1: % Beisbart, Barboni, Wagner and F. da Costa
2: \documentclass[epj]{svjour}
3: %\documentclass[epj,referee]{svjour}
4: \usepackage{graphics}
5: \usepackage{graphicx}
6: \usepackage{amsmath}
7: \usepackage{hyperref}
8: \usepackage{amssymb}
9: \usepackage{subfigure}
10: \usepackage{epic}
11: \usepackage{eepic}
12: \newcounter{subequation}
13: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}\ifnum\thesubequation>0{\alph{subequation}}\fi}
14: \newcommand{\subnumbers}{\setcounter{subequation}{1}\addtocounter{equation}{1}}
15: \newcommand{\nosubnumbers}{\setcounter{subequation}{0}}
16: \newcommand{\stepsubnumber}{\addtocounter{subequation}{1}\addtocounter{equation}{-1}}
17: \def\c{{\mathbf{c}}}
18: \def\dis{{\rm dis}}
19: \def\anis{{\rm anis}}
20: \def\e{{\mathbf{e}}}
21: \def\p{{\mathbf{p}}}
22: \def\R{{\mathbb{R}}}
23: \def\v{{\mathbf{v}}}
24: \def\x{{\mathbf{x}}}
25: \def\V{{\mathbf{V}}}
26: \def\n{{\mathbf{n}}}
27: \def\d{{\rm{d}}}
28: \def\tr{{\mathbf{Tr}}}
29:
30: \begin{document}
31:
32: \title{Extended morphometric analysis of neuronal cells with
33: Minkowski valuations}
34:
35: \author{Claus Beisbart\inst{1,2}, Marconi S. Barbosa\inst{3}, Herbert
36: Wagner\inst{2} \and Luciano da F. Costa\inst{3} }
37:
38: \institute{
39: P.P.M. Research Group\\
40: Center for Junior Research Fellows, University of Konstanz, P.O. Box M 682\\ D--78457 Konstanz, Germany
41: \and
42: Arnold Sommerfeld Center for Theoretical Physics, Ludwig-Maximilians-Universit\"at \\
43: Theresienstra\ss e 37 \\
44: D--80333 M\"unchen,
45: Germany
46: \and
47: Cybernetic Vision Research Group, DFI-IFSC\\
48: Universidade de S\~ ao Paulo, S\~{a}o Carlos, SP\\
49: Caixa Postal 369, 13560-970, Brasil
50: }
51:
52:
53:
54: \date{Received: date / Revised version: date}
55:
56: \abstract{Minkowski valuations provide a systematic framework for
57: quantifying different aspects of morphology. In this paper we apply
58: vector- and tensor-valued Minkowski valuations to neuronal cells
59: from the cat's retina in order to describe their morphological
60: structure in a comprehensive way. We introduce the framework of
61: Minkowski valuations, discuss their implementation for neuronal
62: cells and show how they can discriminate between cells of different
63: types. \PACS{ {07.05.Kf}{Data analysis: algorithms and
64: implementation; data management} \and {87.19.La}{Neuroscience} \and
65: {02.40.Ft}{Convex sets and geometric inequalities} } % end of PACS
66: codes }
67:
68: \authorrunning{Beisbart, Barbosa, Wagner \& da F. Costa}
69: \titlerunning{Minkowski valuations for neuronal cells} \maketitle
70:
71: \section{Introduction}
72:
73: Natural phenomena can be understood as causes and consequences of a
74: continuing interplay between geometry and dynamics, or form and
75: function~\cite{Douady&Coulder:1992,Percolation:2003}. Spatial
76: adjacencies, the specific geometrical features of natural entities, as
77: well as the dimensionality of the space where they are embedded,
78: largely constrain their dynamics and function. For instance, the
79: proper operation of a mammal's heart depends on a suitable diffusion
80: of potentials and waves across the heart surface. It is at the
81: central nervous system, however, that the interplay between form and
82: function reaches its greatest complexity. To begin with, the velocity
83: of signal transmission in neuronal fibers (i.e. dendrites and axons) depends on the width and
84: length of the fibers. The importance of
85: geometry for proper neuronal function is further underlined by the
86: fact that neurons are cells specialized to establish selective
87: connections. Given the spatial constraints imposed by
88: three-dimensional space, these cells have to resort to the most
89: diverse geometries in order to implement the required interconnections
90: -- as they do in a dynamical way during the whole lifetime of an
91: individual.
92: %
93: \\
94: %
95: As a consequence it is to be expected that morphological analyses of
96: neuronal cells provide clues for understanding neural dynamics
97: and function. Although a large number of investigations have
98: been directed at the neural anatomy and geometry
99: (e.g.~\cite{wassle:1981,fuckuda:1984,wassle:1981,pelt:2002}), only
100: the taxonomic organization of neuronal cells or the consideration
101: of shape abnormalities as subsidies for diagnosis have been
102: concentrated on so far. The study of the shape of neuronal cells with
103: objective and mathematically well characterized morphometric
104: descriptors is just at the beginning
105: (e.g.~\cite{velte:1999,Ascoli_Krichmar:2000,Coelho_Costa:2001,Potts_PNAS:2003}).
106: %
107: \\
108: %
109: In order to be useful tools, such morphological descriptors should fulfill the
110: following criteria: First, the extracted quantitative features should
111: obey simple transformation rules, when the neuronal cell under investigation is subject to
112: elementary geometrical transformations such as affine or conformal
113: transformations ({\em in-} and {\em covariance}). Second, the obtained
114: measurements should discriminate between different classes of
115: neuronal cells. Finally, it is important that the estimated
116: features allow for intuitive interpretations from the neuroscience
117: point of view.
118: %
119: \\
120: %
121: Because of their long tradition in modeling and analysis, mathematics,
122: physics and engineering provide a large number of concepts and
123: measures that may be considered for studies in neuroscience and
124: neuromorphometry. A good example is entropy, which has been used in
125: neuroscience because of its close association with the concept of
126: information~\cite{Rieke:1999}. Other such measures include the
127: fractal dimension~\cite{Morigiwa:1989,Coelho:1996,Manoel:2002},
128: lacunarity~\cite{Smith:1996,Rodrigues:2005}, percolation critical
129: density~\cite{Percolation:2003} and curvature~\cite{Cesar:1998}.
130: Recently, concepts from Integral Geometry and in particular the
131: (scalar) Minkowski shape functionals were applied in order to
132: characterize the geometry of ganglion cells from the cat's
133: re\-tina~\cite{Barbosa:2003a,Barbosa:2003b}. Minkowski shape functionals
134: are particularly interesting because they meet the criteria mentioned
135: above: They are invariant under rigid body transformations, seem to
136: have good discriminative power~\cite{Barbosa:2003a}, and can be
137: squared with basic concepts from neuroscience. Moreover, they can
138: easily be implemented: Usually, the original data are filtered with methods known from MIA (Morphological Image
139: Analysis). This preprocessing introduces a free parameter, which can later
140: be varied in order to probe the morphology at different scales. In
141: previous works, the singular points (branching and terminating
142: points)~\cite{Barbosa:2003b} or the whole cell
143: outline~\cite{Barbosa:2003a} were dilated, where the dilation radius
144: entered as a parameter. For each dilation radius, the
145: preprocessed neuron image was decomposed into components (or basic
146: building blocks). The Minkowski functionals can then be calculated by
147: counting certain multiplicities of the basic building blocks. This
148: approach makes use of mathematical results from Integral
149: Geometry~\cite{Raedt:2001}.
150: %
151: \\
152: %
153: In this paper, we use extensions of the Minkowski shape functionals,
154: viz. the {\em Min\-kowski valuations}, in order to further improve
155: neuromorphometric characterization and ana\-lysis. These extensions
156: were only very recently investigated by
157: mathematicians~\cite{alesker:tensor,Schneider:2000,Schneider:2002} and
158: include vector- and ten\-sor-valued measures. They are therefore
159: sensitive to directional information and also allow for valuable
160: graphical visualizations. Minkowski valuations have been successfully
161: applied to describe the morphology of galaxies~\cite{Claus:2002} and
162: galaxy clusters~\cite{Claus:2001}.
163: %
164: \\
165: %
166: In the following, we will illustrate the potential of the Minkowski valuations
167: for neuromorphometry by analyzing a set of ten ganglion cells from the
168: cat's retina. We consider two-dimensional projections of the
169: cells. The set used~\cite{Masland:2001} includes cells with diverse
170: shapes, corresponding to a recently revised classification of those
171: types of cells~\cite{Berson:2002}. In addition, ganglion cells from
172: the retina exhibit branching patterns which are predominantly planar,
173: and therefore compatible with the two-dimensional Minkowski valuations
174: considered in the present work.
175: %
176: \\
177: %
178: The article starts by presenting the higher-order Min\-kows\-ki
179: functionals and proceeds by illustrating their application to the
180: characterization of neuronal cells.
181:
182: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
183: \section{Minkowski valuations}
184: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
185: Morphometry deals with measures for the content, shape and
186: connectivity of spatial patterns (``bodies''). Consider a body $P$ in
187: $2$--dimensional space such as constituted by the pixels of a neuron
188: image (see Figures~\ref{fig:toy} and \ref{fig:smoothing} for
189: examples). A straightforward way to measure its ``content'' is to
190: calculate its area $V_0(P)$ or -- equivalently -- to count its pixels.
191: The area clearly meets the requirement of motion invariance.
192: Furthermore, it is additive; that is, the area of the set union $P$ of
193: two bodies $P_1$, $P_2$ can be decomposed as
194: $V_0(P) = V_0(P_1)+ V_0(P_2) - V_0(P_1\cup P_2)$. Thus, the area can
195: always be calculated by summing up over local contributions from basic
196: building blocks (pixels, e.g.).
197: % more generally, one has to integrate
198: % over the parts of the body: $V_0 = \int_P \d^2 A$.
199: Finally, the area
200: of a convex body can be continuously approximated by the areas of a
201: sequence of convex polygons (conditional continuity of $V_0$).
202: %
203: \\
204: %
205: There are other geometric descriptors that share these properties with
206: the area. The perimeter is a case in point. However, the class
207: of motion-invariant, additive and conditionally continuous descriptors
208: is not unbounded. Let us point this out in full generality for $d$
209: dimensions. Consider an arbitrary pattern $P$ that can be decomposed
210: into a set union of finitely many convex bodies. According to
211: {\em Hadwiger's characterization
212: theorem}~\cite{hadwiger:1957,weil:stereology} there are only $(d+1)$
213: linearly independent measures $V_0(P)$, ..., $V_{d+1}(P)$ that obey
214: motion-invariance, additivity and conditional continuity. They are
215: called {\em (scalar) Minkowski functionals}. Thus, in our case of
216: $d=2$, the area $V_0$, the perimeter $4V_1$ and the Euler
217: characteristic $V_2$ constitute a {\em complete} family of scalar
218: morphological measures. Note, that the Euler characteristic is a
219: topological invariant and equals the number of connected components
220: minus the number of holes for patterns in $\R^2$. The
221: Minkowski functionals were applied to neuronal cell classification in
222: \cite{Barbosa:2003a,Barbosa:2003b}.
223: %
224: \\
225: %
226: Like the area $V_0$, the perimeter $V_1$ and the Euler
227: characteristic $V_2$ can be decomposed into local contributions. This
228: time they arise from the boundary $\partial P$ of the body $P$. For
229: smooth boundaries one has
230: \begin{equation}
231: V_1 = \frac{1}{4} \int_{\partial P} \d^1 S,
232: \quad V_2 = \frac{1}{2\pi} \int_{\partial P} c\,\d^1 S ,
233: \end{equation}
234: where $c$ denotes the curvature of $\partial P$ and varies as one
235: moves along $\partial P$. The factor $\frac{1}{4}$ is a pure matter
236: of convention. For pixel sets, which do not have a smooth boundary,
237: $V_1$ and $V_2$ can be calculated by summing up contributions from the
238: bonds that confine the pixels, and the corners, see~\cite{Raedt:2001}.
239: %
240: \\
241: %
242: A natural way of generalizing the concept of the Minkowski
243: functionals is to replace the requirement of
244: motion {\em in}vari\-ance by motion {\em co}variance. Motion covariance
245: means that the Minkowski valuations obey simple transformation rules,
246: when the body is moved in space: they transform exactly as vectors or
247: tensors do under transformations of a coordinate system.
248: %
249: \\
250: %
251: The class of motion-covariant, additive and conditionally continuous descriptors can
252: be completely characterized by a generalization of Hadwiger's
253: theorem~\cite{alesker:tensor,alesker:rotation}. It turns out that they
254: can be reconstructed as moments of the Minkowski functionals.
255: %
256: \\
257: %
258: In two dimensions there are three {\em first}-order moments of the
259: Minkowski functionals, the so-called {\em Minkowski vectors}. For
260: bodies with a smooth boundary, they can be represented as follows:
261: \begin{gather}
262: \V_0 = \int_P \x\,\d^2 A ,\quad \V_1 = \frac{1}{4} \int_{\partial P}
263: \x\ \d^1 S\notag\\
264: \quad \V_2 = \frac{1}{2\pi} \int_{\partial P} c \, \x \,\d^1 S ,
265: \end{gather}
266: where $\x$ denotes the position vector of the area (perimeter) element
267: $\d^2 A$ ($\d^1 S$) to be integrated over. Minkowski vectors can also
268: be defined for pixelized images, which lack a smooth boundary.
269: %
270: \\
271: %
272: For the purposes of our analysis, it will be useful to normalize the
273: Minkowski vectors and to consider the {\em centroids}:
274: \begin{equation}
275: \p_i = \V_i/V_i\quad (i=0,1,2 \quad {\rm if}\;\; V_i\neq 0).
276: \end{equation}
277: The centroids specify where some aspect of the geometry (area,
278: perimeter, curvature) is concentrated. Note, that the
279: centroids $\p_i$ may, but need not coincide with each other. It can
280: be shown that all centroids coincide for spherically symmetric
281: bodies.
282: %
283: \\
284: %
285: Moving to {\em second}-order moments yields the {\em second-rank
286: Minkowski tensors}. They are built upon the symmetric tensor product
287: denoted by $\x\otimes \x =: \x \x =: \x^2$. In two dimensions there are more than three
288: tensors, because, for $\partial P$-integrals, instead of calculating
289: moments with respect to the spatial position $\x$, one may also
290: consider the local normal $\n$ of the boundary, which points outwards and is normalized to
291: one.\footnote{First-order moments regarding the normal vectors always
292: vanish, as is shown in \cite{hadwiger:vect2}} Thus, for the
293: integrals $\int_{\partial P}
294: \d^1 S$ and $\int_{\partial P} c
295: \d^1 S$ three types of second-order weights for
296: building moments are available, viz. $\x^r\n^s$, where $(r,s)=$
297: $(2,0)$, $(1,1)$ and $(0,2)$ (since we only consider symmetric moments,
298: $\n \x$ and $\x\n$ are identical). Thus, altogether the following
299: seven tensors can be formed:
300: \begin{alignat}{1}
301: V_0^{2,0}&= \int_P \x \x\,\d^2 A ,\\
302: V_1^{r,s} &= \frac{1}{4} \int_{\partial P} \x^r\n^s \,\d^1 S,
303: \\
304: V_2^{r,s} &= \frac{1}{2\pi} \int_{\partial P} c \,\x^r \n^s \,\d^1 S.
305: \end{alignat}
306: In practice, however, we need not consider all of these tensors,
307: because some of them are linearly dependent~\cite{Schneider:2000}. It can be
308: shown that only the following tensors carry independent information:
309: \begin{gather}
310: V_0^{2,0}= \int_K \x \x\,\d^2 A ,\\
311: V_1^{2,0} = \frac{1}{4} \int_{\partial K} \x\x \,\d^1 S,
312: \quad
313: V_1^{0,2} = \frac{1}{4} \int_{\partial K} \n\n\,\d^1 S,\\
314: \quad
315: V_2^{2,0} = \frac{1}{4} \int_{\partial K} c \,\x \x \,\d^1 S.
316: \end{gather}
317: In the following we will concentrate on these tensors. They are
318: listed together with their names in Table~\ref{tab:tens}. The numerics
319: for calculating the Minkowski valuations for pixelized data sets is
320: described in \cite{beisbart:tensor}.
321: \begin{table}
322: \begin{tabular}{|l|l|l|}\hline
323: Symbol & Formula & Name\\\hline\hline
324: $V_0$ & $\int_P \d^2 A$ & area \\\hline
325: $\p_0$ & $\int_P \x \d^2 A /V_0$ & center of mass \\\hline
326: $V_0^{2,0}$ & $\int_P\x\x \d^2 A $ & mass tensor \\\hline
327: $V_1$ & $\int_{\partial P} \d^1 S$ & length of perimeter \\\hline
328: $\p_1$ & $\int_{\partial P}\x \d^1 S /V_1$ & center of perimeter \\\hline
329: $V_1^{2,0}$ & $\int_{\partial P}\x\x \d^1 S $ & perimeter tensor \\\hline
330: $V_1^{0,2}$ & $\int_{\partial P} \n\n\d^1 S $ & $\n$-weighted perimeter tensor\\\hline
331: $V_2$ & $\int_{\partial P}c \d^1 S$ & Euler characteristic \\\hline
332: $\p_2$ & $\int_{\partial P}c \x\d^1 S /V_2$ & curvature centroid \\\hline
333: $V_2^{2,0}$ & $\int_{\partial P} c \x\x \d^1 S $ & curvature tensor \\\hline
334: \end{tabular}
335: \caption{ The Minkowski valuations used in this paper. \label{tab:tens}}
336: \end{table}
337: %
338: \\
339: %
340: Because of motion covariance, the numerical values of the second-rank
341: Minkowski tensors depend on the choice of the coordinate system. But
342: in many applications, there is a natural choice for the origin of the
343: coordinate system. For our neuronal cells we will simply take the
344: position of the soma as the origin (in other cases it might be useful to
345: calculate the second-rank Minkowski tensors $V_i^{r,s}$ with respect
346: to the corresponding centroid $\p_i$ for $i=0,1,2$).
347: %
348: \\
349: %
350: In order to illustrate very briefly how the Minkowski valuations work for pixelized
351: \begin{figure}
352: \centering\includegraphics[width=8.7cm]{a1}
353: \centering\includegraphics[width=8.7cm]{a3}
354: \centering\includegraphics[width=8.7cm]{a4}
355: \caption{Three toy examples to be discussed as an illustration. For
356: the centroids the following point styles are used: red (medium grey)
357: filled square: $\p_0$; blue (dark grey) open square: $\p_1$;
358: green (light grey) x:
359: $\p_2$. The ellipses carry information about the Minkowski tensors;
360: for more information about the construction of the ellipses see the
361: main text. Red (medium grey) ellipse: $V_0^{2,0}$; blue (dark grey) ellipse:
362: $V_1^{2,0}$; green (light grey) ellipse: $V_2^{2,0}$. Ellipse at the left
363: hand side: $V_1^{0,2}$.
364: \label{fig:toy}}
365: \end{figure}
366: data sets, let us consider three simple toy examples (some more
367: examples can be found in \cite{beisbart:tensor}). They are shown in
368: Figure~\ref{fig:toy}. The red (medium grey) filled square, the blue
369: (dark grey)
370: open square and the green (light grey) x denote the
371: centroids $\p_0$, $\p_1$ and $\p_2$, respectively. The
372: tensors are calculated around the center of the black square in the
373: middle of the pixel sets as origin. The red (medium grey), blue (dark grey) and
374: green (light grey) ellipses
375: within the neurons visualize the tensors $V_0^{2,0}$, $V_1^{2,0}$ and
376: $V_2^{2,0}$, respectively. The ellipse for the tensor $V_1^{0,2}$
377: is shown at the left-hand side. The equation defining
378: the ellipses is always: $\x = \c + a\left(\frac{\tau_>}{\tau_<} \cos (\phi)
379: \e_>+\sin(\phi)\e_<\right)$, where $\phi$ runs from 0 to $2\pi$,
380: $\e_{>}$ ($\e_{<}$) is the eigenvector corresponding to the larger
381: (smaller) eigenvalue $\tau_>$ ($\tau_<$) of the tensor and $\c$ is the
382: center of the soma (except for $V_1^{0,2}$; its
383: ellipse is shifted to the edge of the panels). So the axis ratios
384: of the ellipses are the ratios of the eigenvalues, and the ellipses
385: point into the direction of the eigenvector with the larger
386: eigenvalue. The size of the ellipses does not carry specific
387: information because of the free scale factor $a>0$.
388: %
389: \\
390: %
391: In the top panel of Figure~\ref{fig:toy} the pixel set displays an axial symmetry and is
392: almost point symmetric. Accordingly, the centroids are very close to
393: each other; they fan out along the symmetry axis. The tensors
394: $V_i^{2,0}$ align perpendicular to the symmetry axis, because the
395: whole pixel set is more elongated along the horizontal axis. The
396: tensor ellipses for the mass tensor $V_0^{2,0}$ and the perimeter tensor $V_1^{2,0}$ almost coincide,
397: whereas the ellipse corresponding to $V_2^{2,0}$ is a bit more
398: elongated. The reason is that the corners, which play an important
399: role for the curvature tensor
400: $V_2^{2,0}$ are further away from the middle black square, which only
401: contributes to $V_0^{2,0}$ and $V_1^{2,0}$.
402: %
403: \\
404: %
405: For the middle panel, the figure has been slightly modified: in order
406: to destroy the symmetry, we rearranged one of the ``arms''. As a
407: consequence, the average pixel is lower down than in the first panel,
408: so all centroids move downwards. The effect is most prominent for the
409: curvature centroid $\p_2$, because it depends on corners, some of which
410: disappear for the rearranged dendrite. Note, furthermore, that the
411: centroids span a non-degenerate triangle, a fact that can be
412: taken as indicating asymmetry. The lack of symmetry is also reflected
413: by the tensor ellipses, which are not parallel any more. Note,
414: furthermore, that the ratios between the bigger and the smaller
415: eigenvalues are larger for the second pixel set. The reason is that --
416: due to the ``movement'' of the upper right arm -- the vertical
417: extension of the pixel set shrinks on average, such that the whole
418: body is more elongated.
419: %
420: \\
421: %
422: The bottom panel shows a variation of the body in the middle panel,
423: where two holes have been added. This results in an Euler
424: characteristic of $-1$. There is no major effect for both $\p_0$ and
425: $\p_1$ and the related tensors. But for $\p_2$ a big jump can be
426: observed, and the ellipse for the curvature tensor $V_2^{2,0}$ is
427: twisted and more elongated. The position of $\p_2$ can be explained
428: as follows: The hole at the right-hand side makes a big negative
429: contribution to $\V_2$. So, if $\V_2$ is calculated around the
430: center of the black square, it points to the left hand side. But
431: since the Euler characteristic $V_2$ itself is negative, $\p_2$ is
432: bounced back to the
433: right hand side due to its
434: normalization through $V_2$. For the curvature tensor ellipse there is some kind of
435: repulsion from the right hole, because this hole makes a big negative
436: contribution to the tensor; the effect of the other hole is much
437: smaller because it is closer to the soma.
438: %
439: \\
440: %
441: The tensor $V_1^{0,2}$ is shown at the left hand
442: side. It always aligns parallel to the grid axis,
443: the reason for this being that it crucially depends on normals that
444: can only point into four directions for a square lattice.\footnote{
445: %
446: For an elementary proof, you can start with a single pixel and then
447: use additivity.
448: %
449: } The shape of the $V_1^{0,2}$ ellipse can be
450: understood as follows: The eigenvalues of $V_1^{0,2}$ count the number
451: of bonds with horizontal or vertical normals, respectively. For all
452: toy examples there are more vertical normals, so the tensor is
453: anisotropic. By moving from the top to the middle panel, more
454: horizontal than vertical normals are destroyed; in this way the tensor
455: becomes even more anisotropic.
456: %
457: \\
458: %
459: Let us conclude this part by adding two comments. First, note, that
460: by considering the eigenvalues of a tensor with respect to an origin
461: which is given by the body itself, motion-invariance is regained. But
462: does this mean that we have been returning to the scalar Minkowski
463: functionals themselves? The answer is no. Additivity has been lost,
464: because forming eigenvalues is not a linear operation, and, as a
465: consequence, the eigenvalues of a Minkowski tensor cannot be
466: decomposed in the same way as the area is. So we have significantly
467: extended the Minkowski framework without having given up its
468: conceptual foundations.
469: %
470: \\
471: %
472: Second, there is a natural extension of
473: our framework to three-dimensional neuron data.
474: %
475: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
476: \section{The analysis of pixelized neuron data}
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: \paragraph{Data.}
479: We analyze two-dimensional neuron data made a\-vailable by the courtesy
480: of Prof. Berson~\cite{Masland:2001}. We have pixelized maps of ten
481: neurons. They are assigned different types ($\alpha$, $\beta$,
482: $\delta$, $\epsilon$, $\eta$, $\iota$, $\kappa$, $\lambda$, $\theta$,
483: $\zeta$). The neuron maps greatly differ in terms of scale. Each
484: neuron can be thought of as a subset of filled pixels within a
485: square lattice. Not all of the neuron pixel sets are connected;
486: some of them consist of disconnected parts. This is probably due to
487: an artifact of the neuron observations. We will therefore apply a
488: simple smoothing.
489: \paragraph{Method.}
490: For each cell we construct parallel sets with a ball of radius $r_s$
491: on a pixel approximation. The parallel set $P_{r_s}$ of a body $P$
492: comprises all points $\x$ such that the distance between $\x$ and $P$
493: is $r_s$ at most. The smoothing is illustrated in
494: Figure~\ref{fig:smoothing}, where the $\lambda$-neuron is considered.
495: In the sequel, the smoothing length will be varied and used as a diagnostic
496: parameter. It serves to probe structures at different scales.
497: \begin{figure}
498: \centering\includegraphics[width=8cm]{lambda_2}
499: \centering\includegraphics[width=8cm]{lambda_8}
500: \caption{Two smoothed versions of the $\lambda$-neuron. Top panel:
501: smoothing length: 2 pixels. Bottom panel: 8 pixels. The pictures are
502: based on data obtained by~\cite{Masland:2001} (their figure no. 5,
503: copyright permission by
504: \href{http://www.nature.com/neuro}{Nature Neuroscience}). \label{fig:smoothing} }
505: \end{figure}
506: %
507: \\
508: %
509: \noindent For each neuron that has been smoothed with a particular
510: smoothing length, we calculate the scalar Minkowski functionals, the
511: centroids and the second-rank tensors. For the tensors we
512: choose the center of the soma as a natural origin. The soma and its
513: center are identified visually, in an interactive way.
514: %
515: %
516: %
517: \begin{figure}
518: \centering\includegraphics[scale=0.55]{alpha_1}
519: \begin{minipage}[h]{.22\linewidth}
520: \centering\includegraphics[scale=0.55]{beta_1}
521: \end{minipage}
522: \begin{minipage}[h]{.77\linewidth}
523: \centering\includegraphics[scale=0.55]{delta_1}
524: \end{minipage}
525: \caption{Neurons of type $\alpha$ (top panel), $\beta$ (middle panel)
526: and $\delta$ (bottom panel). The smoothing length is one pixel. The
527: meaning of the points and the ellipses is explained in
528: Fig.~\ref{fig:toy}. The small dash in the upper right corner of each
529: panel has a length of 20 pixels. The pictures are based on data
530: obtained by~\cite{Masland:2001} (their figure no. 5, copyright
531: permission by \href{http://www.nature.com/neuro}{Nature
532: Neuroscience}). Note, that in all panels of this figure as well as
533: of Figs.~\ref{fig:n2} and \ref{fig:n3} the tensor ellipses for
534: $V_0^{2,0}$ and $V_1^{2,0}$ almost coincide. }
535: \label{fig:n1}
536: \end{figure}
537: \begin{figure}
538: \centering\includegraphics[scale=0.55]{epsilon_1}
539: \centering\includegraphics[scale=0.55]{eta_1}
540: \centering\includegraphics[scale=0.55]{iota_1}
541: \caption{Neurons of type $\epsilon$ (top panel), $\eta$ (bottom left
542: panel) and $\iota$ (bottom right panel). The
543: smoothing length
544: is one pixel. The pictures are
545: based on data obtained by~\cite{Masland:2001} (their figure no. 5,
546: copyright permission by
547: \href{http://www.nature.com/neuro}{Nature Neuroscience}). }
548: \label{fig:n2}
549: \end{figure}
550: \begin{figure}
551: \begin{minipage}[h]{.66\linewidth}
552: \centering\includegraphics[scale=0.55]{kappa_1}
553: \end{minipage}
554: \begin{minipage}[h]{.33\linewidth}
555: \centering\includegraphics[scale=0.55]{zeta_1}
556: \end{minipage}
557: \centering\includegraphics[scale=0.55]{theta_1}
558: \centering\includegraphics[scale=0.55]{lambda_1}
559: \caption{Neurons of type $\kappa$ (top panel), $\zeta$ (middle left
560: panel), $\theta$ (middle right panel) and $\lambda$ (bottom
561: panel). The smoothing length is one pixel. The pictures are
562: based on data obtained by~\cite{Masland:2001} (their figure no. 5,
563: copyright permission by
564: \href{http://www.nature.com/neuro}{Nature Neuroscience}).}
565: \label{fig:n3}
566: \end{figure}
567: \noindent
568: %
569: \paragraph{Results.}
570: %
571: We show the neurons with some of the results for a smoothing length of one pixel in Figs.~\ref{fig:n1}
572: -- \ref{fig:n3}.\footnote{ In the following, one has to be cautious in
573: interpreting the green (light grey) ellipses, because for our neuronal cells, the
574: tensor $V_2^{2,0}$ sometimes has one or two negative eigenvalues. In
575: this case, the ellipse will become smaller and point into the
576: direction of $\e_<$ instead of $\e_>$, if $|\tau_>|<|\tau_<|$. }
577: %
578: \\
579: %
580: Let us start with some qualitative observations. First, the
581: centroids $\p_0$ through $\p_2$ are typically not within the soma.
582: Recalling that the centroids are
583: morphological centers, we can equivalently say that the soma is quite often eccentric. It would
584: be interesting to know whether the eccentricity of the soma is
585: characteristic for some types of neurons (for this we would have to
586: investigate larger statistical samples of neurons). We suspect
587: that the eccentricities depend on the function and the local
588: environment of the cells. Further investigations are needed to
589: explore this effect.
590: %
591: \\
592: %
593: Second, we observe that typically $\p_0$ and $\p_1$ almost coincide,
594: whereas $\p_2$ may be further away from them. Something similar is true
595: about the tensors: The tensor ellipses of $V_0^{2,0}$ and $V_1^{2,0}$
596: often closely resemble each other, whereas the ellipse for $V_2^{2,0}$ greatly
597: differs. The reason is as follows: As our toy examples have shown,
598: $\p_2$, $V_2^{2,0}$ and the corresponding Minkowski functional
599: (viz. the Euler characteristic) are sensitive to holes. For positive
600: Euler characteristics, every hole that is off-soma pushes $\p_2$ onto
601: the other side of the soma. As a consequence, the location of $\p_2$
602: and the form of $V_2^{2,0}$ very much depend on the holes, their
603: forms and positions. The holes in turn depend on tiny details of the
604: branching structure that are not reflected in $\p_0$ and $\p_1$ and
605: the corresponding tensors $V_0^{2,0}$ and $V_1^{2,0}$. -- Note, that
606: most of the holes are probably due to the projection of the neuron
607: into two dimensions.
608: %
609: \\
610: %
611: We will now turn to a more quantitative analysis. We will show several
612: morphological characteristics that are based upon the Minkowski
613: valuations as a function of smoothing length $r_s$. The point styles
614: designating the different kinds of neurons are explained in the top
615: panel of Fig.~\ref{fig:smooth_sc0}.
616: %
617: \\
618: %
619: We show the first scalar Minkowski functional $V_0$ for a large range
620: of smoothing lengths $r_s$ in the bottom panel of
621: Figure~\ref{fig:smooth_sc0}. For very small $r_s$, $V_0$ grows very
622: quickly, as $r_s$ increases; whereas for larger smoothing lengths, a
623: more moderate growth can be seen. For some neurons it appears to be
624: linear, for other cell types the function $V_0(r_s)$ is clearly
625: convex in this range. Bigger neurons typically grow faster than
626: smaller ones.
627: The explanation is as follows: Let us consider
628: the $\beta$ cell first. Its overall shape is roughly spherical, and
629: its extension $2r_0$ is about $50$ pixels. If the $\beta$ cell is
630: smoothed with a very large $r_s>r_0=25$, all of its substructure is
631: washed out, and we have approximately the same result as if a circle
632: of radius $r_0$ was smoothed by $r_s$. So the volume is about
633: $V_0\approx \pi (r_0+r_s)^2 = \pi r_0^2 + 2 \pi r_0 r_s+\pi r_s^2$,
634: which is parabolic in $r_s$. For $r_s<r_0$, the linear term $ 2 \pi
635: r_0 r_s$ is most significant, so the function $V_0(r_s)$ appears to
636: be linear in a certain range.
637: %
638: \\
639: %
640: More generally, let $CP$ denote the convex hull of a pixelized data set
641: $P$ (or, more precisely, the pixel approximation of its convex hull). For large smoothing lengths, the parallel bodies of $P$ and $CP$,
642: $P_{r_s}$ and $CP_{r_s}$ are very close to each other; consequently
643: the difference $V_0\left(P_{r_s}\right)- V_1\left(CP_{r_s}\right)$ is
644: small compared to $V_0\left(P_{r_s}\right)$. The size of
645: the parallel body $CP_{r_s}$ can be calculated using {\em Steiner's
646: formula} (see \cite{weil:stereology}, p. 367, e.g.):
647: \begin{equation}
648: V_0(CP_{r_s})=V_0(CP)+r_s 4 V_1 (CP) + \pi r_s^2\,.
649: \end{equation}
650: This again
651: defines a parabola, where the Minkowski functionals $V_0$ and $V_1$ of
652: $CP$ arise as coefficients. As a consequence, if $r_s$ is large
653: enough, the volume $V_0(P_{r_s})$ is largely determined by the
654: Minkowski functionals of the convex hull $CP$. For small neuronal
655: cells such as the $\beta$ neuron, this behavior sets in quite
656: early. Bigger neurons will have larger values of $V_0(CP)$ and $V_1(CP)$ such
657: that their area $V_0$ is larger.\footnote{
658: %
659: Similar considerations apply to $V_1$ and $V_2$.
660: }
661: %
662: \\
663: %
664: In order to observe the fine-grained structure of the cells where the
665: neurons significantly differ from their convex hull, we have to
666: concentrate on smaller smoothing lengths $r_s<20$. In
667: Figure~\ref{fig:smooth_sc} the scalar Minkowski functionals are
668: plotted vs. the smoothing length $r_s$. For most neurons, initially, $V_0$ grows comparatively
669: quickly; around $r_s=5$, however, the growth slows down. As a
670: reason, the arms of the neurons that have been blown up, when
671: the parallel set was constructed, start to overlap with each other, such
672: that increasing $r_s$ will not necessarily fill many pixels that have
673: not yet been occupied so far.
674: %
675: \\
676: %
677: For some bigger neurons ($\alpha$, $\delta$, $\kappa$, e.g.) a kind of
678: crossover can be observed around $r_s=5$. For the other types of
679: neuronal cells, the crossover is less pronounced.\footnote{Note, by
680: the way, that there are plateaus at the zero points for the $V_i$
681: vs. $r_s$ curves. More generally, these curves are not continuous,
682: but change stepwise because of our pixelwise smoothing. This can be
683: seen, if the $r_s$ resolution is enhanced. In the following we will
684: neglect discontinuities of this kind; they are a pure artifact of our
685: smoothing and do not carry any physical meaning. }
686: \begin{figure}
687: \centering\includegraphics[width=8.4cm]{helper}
688: \centering\includegraphics[width=8.4cm]{mf_rs_sc1_large}
689: \caption{Top panel: the point styles to be used for the different
690: types of neuronal cells. Bottom panel: the volume $V_0$ as a
691: function of $r_s$ for a large range of
692: $r_s$-values. }\label{fig:smooth_sc0}
693: \end{figure}
694: \begin{figure}
695: \centering\includegraphics[width=8.4cm]{mf_rs_sc1}
696: \centering\includegraphics[width=8.4cm]{mf_rs_sc2}
697: \centering\includegraphics[width=8.4cm]{mf_rs_sc3}
698: \caption{The scalar Minkowski functionals as a function of the
699: smoothing length for all cells. (Note, that in the first panel the
700: curve for the $\theta$ cell is between the curves for the $\eta$ and
701: $\zeta$ types.) }\label{fig:smooth_sc}
702: \end{figure}
703: %
704: \\
705: %
706: \noindent We will now consider $V_1$. For small $r_s$, $V_1$
707: decreases as a function of $r_s$, since
708: $V_1$ is dominated by small scale features that are smoothed
709: away stepwise. $V_1$ reaches a constant value later on. This is not what one would expect for a convex body. The
710: reason, of course, is that the figure is far from being convex: As
711: $r_s$ increases, $V_1$ will gain at the outer parts of the cells, but
712: loose in the inner parts, because holes are being filled. Gains and
713: losses roughly compensate each other. Note, that the curves for the
714: $\alpha$, $\delta$ and $\kappa$ cell type show an inflection point, which
715: very roughly coincides with the position of their
716: crossover in $V_0$.
717: %
718: \\
719: %
720: \noindent The curves for the Euler characteristic display a number of
721: discontinuities, but there is some more general pattern. The negative
722: values indicate that the cells are dominated by holes. For the bigger
723: cell types ($\alpha$, $\delta$, $\epsilon$ and $\kappa$) there is a
724: characteristic dip for small $r_s$. Up to this point, additional holes
725: are formed, as branches of the neuron start to touch each other. The
726: minimum of the dip roughly seems to coincide with the point where
727: $V_0$ shows the crossover for the bigger neurons. Similar results for
728: the Minkowski functionals have been obtained in~\cite{Barbosa:2003a}.
729: %
730: \\
731: %
732: A useful way of combining the information present in the scalar
733: Minkowski functionals is to construct the following dimensionless
734: quantity $Q$:
735: \begin{equation}
736: Q := \frac{4 V_1^2}{\pi V_0} \, .
737: \end{equation}
738: This is a variation of the so-called isoperimetric ratio. For a convex
739: body $P$ we have $Q (P)\ge 1$, where the equality holds for a
740: circle~\cite{fenchel:iso,alexandrov:iso,schmalzing:webI}. $Q$ is
741: considerably larger than one, whenever the body under investigation
742: has an ``excess perimeter'' as compared to its area. We show the
743: logarithm of $Q$ as a function of $r_s$ in Fig.~\ref{fig:iso}.
744: \begin{figure}
745: \centering\includegraphics[width=8.4cm]{logiso_20}
746: \caption{The logarithm of $Q$ (a variety of the isoperimetric ratio)
747: as a function of $r_s$.}\label{fig:iso}
748: \end{figure}
749: The interpretation is as follows: For small smoothing lengths $r_s$,
750: most of the dendrites are still present; they produce huge excess
751: areas for which reason $Q$ starts with very high $Q$-values. As the
752: smoothing length increases, $Q$ goes down. The $\alpha$, $\epsilon$
753: and $\kappa$ cells have the largest $Q$-values, whereas the
754: $\beta$-cell has the lowest $Q$-values for a large range of smoothing
755: lengths because of its smallness and its overall spherical shape. For
756: $r_s<6$ the decrease in $\log_{10}(Q)$ seems roughly to be
757: linear, where the slopes vary with the cell type.\footnote{
758: %
759: Note, that ``linearity'' holds only up to discreteness effects due to
760: our pixelwise smoothing.}
761: %
762: \\
763: %
764: In Figs.~\ref{fig:smooth_dis} and \ref{fig:smooth_dis2} we consider
765: \begin{figure}
766: \centering\includegraphics[width=8.4cm]{mf_rs_dis_0}
767: \centering\includegraphics[width=8.4cm]{mf_rs_dis_1}
768: \caption{The distance soma -- $\p_0$ (top panel) and soma -- $\p_1$
769: (bottom panel) as a function of the smoothing length.}
770: \label{fig:smooth_dis}
771: \end{figure}
772: \begin{figure}
773: \centering\includegraphics[width=8.4cm]{mf_rs_dis_2_1}
774: \centering\includegraphics[width=8.4cm]{mf_rs_dis_2_2}
775: \centering\includegraphics[width=8.4cm]{mf_rs_dis_2_3}
776: \centering\includegraphics[width=8.4cm]{mf_rs_dis_2_4}
777: \caption{The distances soma -- $\p_2$ as a function of the smoothing
778: length for four cells. If $\p_2$ is not defined for some $r_s$
779: (because of $V_2=0$), no
780: point is shown.} \label{fig:smooth_dis2}
781: \end{figure}
782: the centroid distances $\p_i$-soma, $\dis_i$. For $i=0$ they are
783: relatively stable as a function of $r_s$, whereas for $i=1$ more
784: variation can be observed. How is this to be explained? Look at the
785: $\kappa$ neuron as an example (Fig.~\ref{fig:n3}). In the lower half
786: of the image the distribution of small arms is a bit denser than in
787: the upper half. Consequently, for small $r_s$, there is a significant
788: contribution to the perimeter from this part, and this is also
789: reflected in the position of $\p_1$, which is the center of the
790: perimeter. For larger $r_s\approx 10$, however, the lower, denser part
791: is filled more quickly, whereas in the upper part quite big holes are
792: left, which contribute to the perimeter, such that the position of
793: $\p_1$ moves upwards. In this way the curve for $\dis_1$ contains very
794: detailed information about the morphology of the neuron.
795: %
796: \\
797: %
798: In terms of $\dis_0$ and $\dis_1$ the soma is most eccentric for the
799: $\epsilon$ neuron. This is also reflected in our visual
800: impressions. It might be useful, however, to normalize the $\dis_i$
801: parameters by some estimate of the cell size. If we would do so,
802: smaller cells would have a reasonable chance of having bigger
803: eccentricities.
804: %
805: \\
806: %
807: For $i=2$ (Fig.~\ref{fig:smooth_dis2}) we observe even larger
808: variations of the centroid distances. Plateaus alternate with jumps
809: that can ultimately be traced back to discontinuities of the Euler
810: characteristic. For small neurons, such as the $\beta$ type, however,
811: there is not much variation, because the cell is very small and gets
812: completely filled pretty soon. For the $\alpha$, $\delta$ and
813: $\epsilon$-type, there is a common pattern: As the smoothing length
814: increases, the jumps become larger. The reason is probably, that for
815: larger smoothing lengths only a few holes will appear far off
816: centered. When one of these outer holes disappears, $\p_2$ jumps
817: considerably.
818: %
819: \\
820: %
821: \noindent
822: In Figures~\ref{fig:smooth_anis} through \ref{fig:smooth_anis5} we
823: consider the anisotropy of the cells. In order to quantify anisotropy
824: we take the eigenvalues of the tensors $V_i^{j,k}$, $\tau_>$ and
825: $\tau_<$ and calculate the quantity $ \anis:=2(
826: \tau_>-\tau_<)/(|\tau_>|+|\tau_<|)\leq 2$. The anisotropy parameters
827: derived from different tensors focus on different kinds of anisotropy
828: (the area elements belonging to a body might be distributed
829: differently from those of its perimeter elements, for instance). As
830: can be seen from Fig.~\ref{fig:smooth_anis}, the anisotropies in
831: $V_0^{2,0}$ and $V_1^{2,0}$ are quite stable; most often they decrease
832: slowly, as
833: \begin{figure}
834: \centering\includegraphics[width=8.4cm]{mf_rs_anis1}
835: \centering\includegraphics[width=8.4cm]{mf_rs_anis2}
836: \caption{The anisotropy parameters derived from the mass tensor
837: $V_0^{2,0}$ (top panel) and the perimeter tensor $V_1^{2,0}$ (bottom
838: panel) as a function of the smoothing
839: length.}\label{fig:smooth_anis}
840: \end{figure}
841: %
842: \begin{figure}
843: \centering\includegraphics[width=8.4cm]{mf_rs_anis_5_1}
844: \centering\includegraphics[width=8.4cm]{mf_rs_anis_5_2}
845: \caption{Another anisotropy parameter (derived from $V_2^{2,0}$) for
846: two cells as a function of the smoothing
847: length.}\label{fig:smooth_anis4}
848: \end{figure}
849: the smoothing length increases. This indicates that the cells display
850: large-scale anisotropies that are not destroyed by smoothing the
851: cell. For some cells ($\eta$, $\kappa$, $\zeta$) the anisotropies
852: are considerable. For each cell type the anisotropies of area and
853: perimeter do not differ greatly. The $V_1^{2,0}$ tensor is a bit more
854: sensitive to small-scale variations of the morphology, however; so the
855: $\anis(V_1^{2,0})$-$r_s$ curves appear less smooth than the
856: $\anis(V_0^{2,0})$-$r_s$ curves. On the other hand, across the range
857: of cell types, the variation is quite high. Thus anisotropies seem to
858: have a significant discriminative power.
859: %
860: \\
861: %
862: It is different with the tensor $V_2^{2,0}$, which is considered in
863: Figure~\ref{fig:smooth_anis4}. The anisotropy derived from this
864: tensor jumps back and forth and sometimes reaches values that exceed
865: those derived from the other tensors. This performance should not
866: come as a surprise, since we have already seen that other
867: characteristics that are related to the Euler characteristic
868: such as $V_2^{2,0}$
869: typically show discontinuities. At some point, however, when the
870: smoothing has produced one connected pattern without holes (visible
871: for the $\beta$ cell, e.g., where this point is reached quite
872: early), the anisotropy stabilizes at a constant value. Apart from
873: this, the dependence on $r_s$ looks rather chaotic; so far, we are
874: not able to extract information that might help to discriminate
875: between the different cell types.
876: %
877: \\
878: %
879: As mentioned before, on the square lattice, the last tensor to be
880: considered, $V_1^{0,2}$, has a simple interpretation. It checks
881: whether the majority of normals are parallel to the horizontal or to
882: the vertical grid axis. If $\partial P$ is dominated by vertical or
883: horizontal normals, $V_1^{0,2}$ will display a corresponding
884: anisotropy; if not, $V_1^{0,2}$ will roughly be isotropic. In
885: Figure~\ref{fig:smooth_anis5} we show some results for single neurons.
886: One can learn from this that the anisotropies arising from $V_1^{0,2}$
887: are quite small. The anisotropy is comparatively large for the $\eta$
888: type cell, because this cell is clearly elongated. For small values of
889: the smoothing length, $\anis(V_1^{0,2})$ is not so much influenced by
890: the overall shape of the neuron, but rather by the directions of the
891: single arms. Interestingly, the graphs shown are qualitatively
892: different for the different types of cells: One cell (viz. the
893: $\alpha$ cell) starts with zero anisotropy, whereas others begin with
894: a non-zero anisotropy. Moreover, there are significant peak
895: structures. But because of its relation to normals, $\n$, the value
896: of $V_1^{0,2}$ depends to a large extent on the orientation of the
897: cell with respect to the grid. For this reason $V_1^{0,2}$ is only of
898: limited use.
899: %
900: \\
901: %
902: In Fig.~\ref{fig:smooth_tr}, the {\em traces} of the tensors
903: $V_i^{2,0}$ are considered (the trace of the fourth tensor,
904: $V_1^{0,2}$ need not to be taken into account at this point, because
905: it equals $V_1$). Qualitatively, the viewgraphs for $V_i^{2,0}$
906: resemble the curves
907: \begin{figure}
908: \centering\includegraphics[width=8.4cm]{mf_rs_anis_4_1}
909: \centering\includegraphics[width=8.4cm]{mf_rs_anis_4_2}
910: \centering\includegraphics[width=8.4cm]{mf_rs_anis_4_3}
911: \centering\includegraphics[width=8.4cm]{mf_rs_anis_4_5}
912: \caption{The anisotropy parameter derived from $V_1^{0,2}$ for four
913: cells as a function of the smoothing
914: length.}\label{fig:smooth_anis5}
915: \end{figure}
916: of their scalar counterparts, $V_i$ for $i=0,..,2$. In order to
917: extract more specific information, it is thus useful to divide $\tr
918: \left(V_i^{2,0}\right)$ by $V_i$, respectively, for $i=0,..,2$. The result is a measure of
919: how concentrated a cell is in terms of area, perimeter or curvature: $\tr
920: (V_0^{2,0})/V_0$, for instance will be the bigger, the further the soma
921: and those parts of the cell that bear most of its volume lie apart.
922: \begin{figure}
923: \centering\includegraphics[width=8.4cm]{mf_rs_tr1}
924: \centering\includegraphics[width=8.4cm]{mf_rs_tr2}
925: \centering\includegraphics[width=8.4cm]{mf_rs_tr5}
926: \caption{The traces of the tensors $V_i^{2,0}$ for $i=0,..,2$ as a
927: function of the smoothing length.}\label{fig:smooth_tr}
928: \end{figure}
929: Results can be seen from Figure~\ref{fig:smooth_trn}. $\tr ((V_0^{2,0})/V_0$ increases continuously, as $r_s$
930: is enhanced. The reason is that more and more pixels are added at the
931: outer parts of the neuron, so the neuron becomes less and less
932: concentrated. In $\tr ((V_1^{2,0})/V_1$ there is a kink at
933: least for some neurons ($\alpha$, $\delta$, $\epsilon$, $\kappa$). It
934: indicates an additional growth effect. Very probably the explanation
935: is that for small $r_s$, the small branches within the cell
936: significantly contribute to $V_1^{2,0}$, so the neuron appears to be
937: very concentrated; for larger values of $r_s$ the arms merge and do
938: not contribute to the perimeter any more, so most of the neuron's perimeter
939: is found at its outer parts. Note, that the kinks roughly set in at
940: the $r_s$-locations of the crossover point in $V_0$ and the
941: inflection point in $V_1$ for the $\alpha$, $\delta$ and $\kappa$
942: cells.
943: \begin{figure}
944: \centering\includegraphics[width=8.4cm]{mf_rs_trn1}
945: \centering\includegraphics[width=8.4cm]{mf_rs_trn2}
946: \centering\includegraphics[width=8.4cm]{mf_rs_trn5}
947: \caption{The traces of $V_0^{2,0}$, $V_1^{2,0}$, and $V_2^{2,0}$, now
948: normalized by the corresponding scalar $V_0$, $V_1$, or $V_2$,
949: respectively. If $V_2=0$ for some $r_s$, no data point is shown at
950: all. }\label{fig:smooth_trn}
951: \end{figure}
952:
953: \paragraph{Constructing global measures}
954: A multiscale analysis like that presented in this paper, leads
955: to rich and detailed information on the geometrical aspects of an
956: object. Nevertheless, once such a description of the data has been
957: obtained, it is often useful to derive a compact set of global
958: measures that summarize the most important morphological aspects. In this
959: paper, we consider several ways of condensing multiscale information,
960: i.e. a function of some scale, into
961: simple parameters: The \emph{monotonicity
962: index}~\cite{Barbosa:2003a,Barbosa:2003b} is defined as
963: \begin{equation}
964: i_s=\frac{s}{s+d+p}\,,
965: \end{equation}
966: where $s$, $d$ and $p$ count each time the function increases,
967: decreases and remains unchanged, respectively. Thus $i_s$ quantifies
968: the fraction of the interval where the function is monotonically
969: increasing. The \emph{mean value} is the average value of the function
970: over the interval. The \emph{half scale} is the scale at which the
971: area below a curve reaches half of its total value. A different way of
972: constructing global parameters is to consider the \emph{slope} of some
973: characteristic in some particular range of $r_s$-values.
974: %
975: \\
976: %
977: In Figure~\ref{fig:slopes} we visualize the average slopes of $V_0$ in the range
978: \begin{figure}
979: \centering\includegraphics[width=8.4cm]{lines_sl}
980: \caption{The slopes of $V_0$ (top panel) and $\log_{10}(Q)$ for the
981: different cells. The point styles are as above. }\label{fig:slopes}
982: \end{figure}
983: $r_s\in [10,60]$ and of $\log_{10}(Q)$ in the range $r_s\in[1.5,4.5]$
984: for the different cells. In both cases we choose a range of
985: $r_s$-values for which the functionals under investigation look
986: roughly linear for most cell types. Results are shown in
987: Figure~\ref{fig:slopes}. One can immediately see that the slopes
988: discriminate amongst the different cell types.
989: %
990: \\
991: %
992: In order to further illustrate our approach, we selected two two-dimensional
993: feature spaces, which are spanned by size-independent morphological
994: characteristics. In order to calculate them, we considered the
995: interval $r_s \in \left[0,20\right]$ and a spacing of $0.2$.
996: %
997: \\
998: %
999: Our first feature space is spanned by the mean of the anisotropy parameter
1000: derived from $V_0^{2,0}$, $\overline{\anis}\left(V_0^{2,0}\right)$, and the
1001: mean of the anisotropy parameter corresponding to $V_0^{2,0}$, $\overline{
1002: \anis}\left(V_1^{2,0}\right)$. It is shown in Figure~\ref{fig:scatter}(a).
1003: There appears to be some systematic correlation between both characteristics:
1004: cells with higher $\overline{ \anis}\left(V_0^{2,0}\right)$ tend to have
1005: higher $\overline{ \anis}\left(V_1^{2,0}\right)$ as well. Given the meaning of
1006: these characteristics, this should not come as a surprise, although it is in
1007: principle possible to have high anisotropy in $V_0^{2,0}$ and low anisotropy in
1008: $V_1^{2,0}$. Thus, for discriminating between different cells, one dimension
1009: of this feature space is essentially redundant. But the presence of some
1010: correlation might be used to describe some common trait shared by all cells.
1011: %
1012: \\
1013: %
1014: A different situation can be observed for our second feature space. It is spanned
1015: by the monotonicity index $i_s\left(\dis_0\right)$ and by the half scale
1016: $h\left(\tr\left(V_1^{2,0}\right)\right)$. As can be seen from
1017: Figure~\ref{fig:scatter}(b), the scatter is larger, and the cells form kind of
1018: groups. Note, in particular, that neuronal cells that look similar at least in
1019: some respect tend to appear close to each other in this scatter plot. For
1020: instance, cells $\beta$, $\eta$, $\theta$ and $\zeta$ are close to each other
1021: in the bottom panel of Figure~\ref{fig:scatter}, especially regarding the
1022: position of their center of mass $\\p_0$ relatively to their soma.
1023: Figure~\ref{fig:tree} presents a dendrogram obtained by a simple hierarchical
1024: agglomerative clustering~\cite{Cluster:book} of the scatter plot distribution
1025: shown in Figure~\ref{fig:scatter}(b). Such a structure suggests a possible
1026: \emph{taxonomy} for the ten types of cells. As expected, the cells $\beta$,
1027: $\eta$, $\theta$ and $\zeta$ are similar, inhabiting the same branch at the
1028: lower part of the dendrogram. For the remaining subset, the cells
1029: $\alpha$ and $\kappa$ end up markedly distinct from the group of cells formed
1030: by $\delta$, $\epsilon$, $\iota$ and $\lambda$.
1031: %
1032: \\
1033: %
1034: Although the proposed methodology may have a bearing on
1035: the classification of cat ganglion cells, it is difficult to make
1036: more
1037: definitive conclusions at this point, because the original
1038: classification~\cite{Berson:2002} takes into account not only the
1039: neuronal morphology, but also the cell stratification and the size of
1040: the soma. Moreover, except for the more common $\alpha$ and $\beta$
1041: types, only a small number of examples of the cell types have been
1042: analyzed in the related literature~\cite{Berson:2002}. A more
1043: detailed examination of which feature spaces are most useful has
1044: to wait for further data.
1045: \begin{figure}
1046: \begin{center}
1047: \subfigure[]{\includegraphics[width=7cm]{scatter_1}}
1048: \subfigure[]{\includegraphics[width=7cm]{scatter_0}}
1049: \caption{Scatter plots from selected features of an extended Minkowski
1050: analysis showing the population of the feature space with the
1051: neuronal cells.~\label{fig:scatter}}
1052: \end{center}
1053: \end{figure}
1054: %
1055: \begin{figure}
1056: \centering\includegraphics[width=11.2cm, angle=270]{tree_scater}
1057: \caption{
1058: The classification pattern according to an agglomerative hierarchical
1059: clustering analysis considering the two features selected for the
1060: scatter plot in Figure~\ref{fig:scatter}(b).~\label{fig:tree}}
1061: \end{figure}
1062: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1063: \section{Conclusions}
1064: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1065:
1066: We have analyzed two-dimensional projections of neuronal cells using
1067: higher-order Minkowski valuations. Our measures detect different kinds
1068: of substructures, providing a natural extension of previous works that
1069: deal with the more traditional shape
1070: functionals~\cite{Barbosa:2003a,Barbosa:2003b}. An extensive
1071: discussion of the results obtained for a set of ten neuronal cells was
1072: included that illustrates the interpretation of the suggested measures
1073: and implications for neuromorphometric studies. As far as our limited
1074: set of samples is concerned, significant similarities and differences
1075: between the cell types have been found, leading to a putative
1076: taxonomy. It is a pending question whether the differences found will
1077: still be characteristic of the types in a statistical sense.\\[2cm]
1078: {\small This work was financially supported by FAPESP (processes
1079: 02/02504-01 and 99/12765-2) and CNPq (process 308231/03-1). It was
1080: also supported by the "Sonderforschungsbereich 375-95 für
1081: Astro-Teilchenphysik" der Deutschen Forschungsgemeinschaft. C.B.
1082: thanks the Alexander von Humboldt Foundation, the German Federal
1083: Ministry of Education and Research and the Program for the
1084: Investment in the Future (ZIP) of the German Government for
1085: supporting this research. He also thanks Jens Schmalzing for
1086: providing software on which parts of the codes for this paper are
1087: built upon.}
1088:
1089:
1090:
1091:
1092: \bibliographystyle{vunsrt}
1093: \bibliography{n}
1094:
1095:
1096: \end{document}
1097:
1098:
1099: