0805.0101/IP2.tex
1: %&latex
2: %&latex
3: \documentclass[namedreferences]{SolarPhysics}
4: \usepackage[optionalrh]{spr-sola-addons} % For Solar Physics 
5: %\usepackage{epsfig}          % For eps figures, old commands
6: \usepackage{graphicx}        % For eps figures, newer & more powerfull
7: %\usepackage{courier}         % Change the \texttt command to courier style
8: %\usepackage{natbib}         % For citations: redefine \cite commands
9: %\usepackage{amssymb}        % useful mathematical symbols
10: \usepackage{color}           % For color text: \color command
11: \usepackage{url}             % For breaking URLs easily trough lines
12: \def\UrlFont{\sf}            % define the fonts for the URLs
13: 
14: % Definitions for the journal names
15: \newcommand{\adv}{    {\it Adv. Spa. Res.}}
16: \newcommand{\annG}{   {\it Annales Geophysicae}}
17: \newcommand{\aap}{    {\it Astron. Astrophys.}}
18: \newcommand{\aaps}{   {\it Astron. Astrophys. Suppl.}}
19: \newcommand{\aapr}{   {\it Astron. Astrophys. Rev.}}
20: \newcommand{\ag}{     {\it Ann. Geophys.}}
21: \newcommand{\aj}{     {\it Astronom. J.}}
22: \newcommand{\apj}{    {\it Astrophys. J.}}
23: \newcommand{\apjl}{   {\it Astrophys. J. Lett.}}
24: \newcommand{\apss}{   {\it Astrophys. Spa. Sci.}}
25: \newcommand{\cjaa}{   {\it Chinese J. Astron. Astrophys.}}
26: \newcommand{\gafd}{   {\it Geophys. Astrophys. Fluid Dyn.}}
27: \newcommand{\grl}{    {\it Geophys. Res. Lett.}}
28: \newcommand{\ijga}{   {\it Int. J. Geomag. Aeron.}}
29: \newcommand{\jastp}{  {\it J. Atmos. Sol. Terr. Phys.}}
30: \newcommand{\jgr}{    {\it J. Geophys. Res.}}
31: \newcommand{\mnras}{  {\it Mon. Not. Roy. Astron. Soc.}}
32: \newcommand{\nat}{    {\it Nature}}
33: \newcommand{\pasp}{   {\it Pub. Astron. Soc. Pac.}}
34: \newcommand{\pasj}{   {\it Pub. Astron. Soc. Japan}}
35: \newcommand{\pre}{    {\it Phys. Rev. E}}
36: \newcommand{\solphys}{{\it Solar Phys.}}
37: \newcommand{\sovast}{ {\it Sov. Astronom.}}
38: \newcommand{\ssr}{    {\it Space Sci. Rev.}}
39: 
40: \newcommand{\mres}{multiresolution}
41: \newcommand{\Mres}{Multiresolution}
42: \newcommand{\ls}{length-scale}
43: \newcommand{\Ls}{Length-scale}
44: \newcommand{\bgdcl}{$\beta\gamma\delta$}
45: \newcommand{\bgcl}{$\beta\gamma$}
46: \newcommand{\bcl}{$\beta$}
47: \newcommand{\acl}{$\alpha$}
48: \newcommand{\linename}{MOPRS}
49: 
50: \begin{document}
51: \begin{article}
52: \begin{opening}
53: 
54:   \title{\Mres\ analysis of active region magnetic structure and its
55:     correlation with the Mount Wilson classification and flaring
56:     activity}
57: 
58: \author{J.     \surname{Ireland}$^{1}$,
59:         C.A.   \surname{Young}$^{1}$,
60:         R.T.J. \surname{McAteer}$^{2}$,
61:         C.     \surname{Whelan}$^{3}$,
62:         R.J.   \surname{Hewett}$^{4}$,
63:         P.T.   \surname{Gallagher}$^{5}$}
64: 
65: \runningauthor{Ireland et al.}
66: \runningtitle{\Mres\ analysis of active region magnetic structure}
67: 
68: \institute{ $^{1}$ ADNET Systems, Inc., NASA's Goddard Spaceflight
69:   Center, Mail Code 671.1, Greenbelt, MD 20771, USA.\\
70:   $^{2}$ Catholic University of America, NASA Goddard Space Flight
71:   Center, Greenbelt, MD 20771, USA.\\
72:   $^{3}$ School of Physics, University Col lege Dublin, Belfield,
73:   Dublin 4, Ireland.\\
74:   $^{4}$ Computer Science Departement, University of Il linois at
75:   Urbana-Champaign, Urbana, IL 61801, USA.\\
76:   $^{5}$ Astrophysics Research Group, School of Physics, Trinity 
77:   College Dublin, Dublin 2, Ireland.\\
78: }
79: 
80: \begin{abstract}
81:   Two different \mres\ analyses are used to decompose the structure of
82:   active region magnetic flux into concentrations of different size
83:   scales.  Lines separating these opposite polarity regions of flux at
84:   each size scale are found.  These lines are used as a mask on a map
85:   of the magnetic field gradient to sample the local gradient between
86:   opposite polarity regions of given scale sizes.  It is shown that
87:   the maximum, average and standard deviation of the magnetic flux
88:   gradient for $\alpha, \beta, \beta\gamma$ and $\beta\gamma\delta$
89:   active regions increase in the order listed, and that the order is
90:   maintained over all \ls s. Since magnetic flux gradient is strongly
91:   linked to active region activity, such as flares, this study
92:   demonstrates that, on average, the Mt. Wilson classification encodes
93:   the notion of activity over all \ls s in the active region, and not
94:   just those \ls s at which the strongest flux gradients are found.
95:   Further, it is also shown that the average gradients in the field,
96:   and the average length-scale at which they occur, also increase in
97:   the same order.  Finally, there are significant differences in the
98:   gradient distribution, between flaring and non-flaring active
99:   regions, which are maintained over all \ls.  It is also shown that
100:   the average gradient content of active regions that have large
101:   flares (GOES class 'M' and above) is larger than that for active
102:   regions containing flares of all flare sizes; this difference is
103:   also maintained at all \ls.  All the reported results are
104:   independent of the \mres\ transform used.  The implications for the
105:   Mt. Wilson classification of active regions in relation to the
106:   \mres\ gradient content and flaring activity are discussed.
107: \end{abstract}
108: \keywords{Sun: active region, Sun: magnetic field}
109: \end{opening}
110: 
111: \section{Introduction}
112: \label{Introduction} 
113: 
114: Many attempts have been made to assess the complexity of active
115: regions, as a measure of their activity.  The earliest attempt (dating
116: from 1908) - Mount Wilson classification - and still in use today,
117: puts active regions into four broad classes, and is based on the
118: distribution of magnetic flux polarity in the active region.  The
119: magnetic flux distribution, in relation to the location of white
120: light sunspots, in conjunction with some simple classification rules,
121: are used by the observer, to classify the observed region.  The four
122: main classes of active region, and their classification rules are
123: shown in Table \ref{tab:mtwilson}.
124: 
125: The success of the Mt. Wilson classification scheme lies in the fact
126: that it is simple, and has some predictive power when combined with
127: flare frequency rates deduced from long time series (many solar
128: cycles) observations. Other classification schemes exist, notably the
129: McIntosh classification scheme \cite{1990SoPh..125..251M}.  This is
130: a significantly more complex classification scheme because it looks at
131: the magnetic structure of the active region in much greater detail.
132: This too has some predictive ability in divining activity, again based
133: on correlation between flare frequency rates and the the McIntosh
134: class.
135: 
136: Both schemes are implemented manually, and so are subject to the usual
137: human observer biasses.  Research is currently ongoing into automating
138: these classification schemes, with some success
139: (\opencite{2006SoPh..239..519D}; \opencite{2002ApJ...568..396T};
140: \opencite{1998A.AS..131..371B}).  The notable feature about both the
141: Mt. Wilson and McIntosh classification schemes is that they both use
142: the magnetic structure, that is, the relative locations and sizes of
143: concentrations of opposite polarity magnetic flux in the active region
144: in order to determine active region class.  Since opposite polarity
145: flux creates gradients, these classification schemes are in
146: essence classifying the location and size of the magnetic flux
147: gradients in the magnetic data.  Hence these classification schemes
148: act as proxies for the magnetic flux gradient information content of
149: the active region.
150: 
151: The notion of complexity is ill-defined, and that carries over to the
152: study of active regions.  Many studies have attempted to quantify the
153: notion of active region complexity. \opencite{2005ApJ...631..628M}
154: look at around 10,000 active regions and calculate a fractal dimension
155: for the spatial distribution of the absolute magnetic flux.  Separated
156: by Mt. Wilson class, active regions are surprisingly homogenous in
157: fractal dimension, with very little difference between $\alpha$ class
158: active regions and $\beta\gamma\delta$ active regions.  This shows,
159: under the assumption that the flux is a mono-fractal, that all the
160: magnetic flux is basically the same, regardless of polarity and of
161: what class of active region it appeared in.
162: 
163: The mono-fractal restriction can be removed by assuming that the
164: magnetic flux can be described my a multifractal.  This assumes that
165: the flux can be represented by a distribution of fractals, a
166: considerable increase in sophistication.  This is physically
167: motivated, since the multifractals have deep ties to turbulence (
168: \opencite{2005SoPh..228....5G}; \opencite{2005SoPh..228...29A};
169: \opencite{2002ApJ...577..487A} ).  \opencite{conlon} present the
170: results of analyzing the multifractal content of several active
171: regions.  It is shown that the multifractal spectrum changes with the
172: emergence of the flux, indicating that the scale content is changing
173: with emergence.  This study allows for flux cancellation in
174: calculating the multifractal spectrum, and so implicitly takes account
175: of the different flux polarity, a significant difference from the
176: study of \opencite{2005ApJ...631..628M}.
177: 
178: Both these studies are more concerned with the nature of the flux, as
179: opposed to its gross distribution on the Sun's surface, which is what
180: the classification schemes outlined above deal with. In particular,
181: these studies do not make any statements about the gradient between
182: flux elements, which is known to be a significant indicator of
183: activity. This study is complementary, in that it attempts to look at
184: the spatial distribution of flux at different sizes, or \ls s, as
185: opposed to the nature of that flux.  To do this, the study uses four
186: \mres\ analyses to decompose the structure on different \ls s, and
187: then looks at the gradient between flux elements of different size in
188: an attempt to examine how the scale size of the spatial flux
189: distribution is related to the gradient, and to the Mt. Wilson
190: classification, which implicitly encodes gradient information.  It is
191: known that the presence of strong gradients in the magnetix flux
192: distribution is indicative of the active region activity
193: (\opencite{2007ApJ...656.1173L}; \opencite{2006ApJ...644.1258F};
194: \opencite{2003ApJ...595.1277L}(a); \opencite{2003ApJ...595.1296L}(b);
195: \opencite{2002ApJ...569.1016F}; \opencite{1997ApJ...482..519F}).  A
196: recently found indicator of activity is given by
197: \opencite{2007ApJ...655L.117S}, in which it is found that the total
198: unsigned magnetic flux in a 15 arcsecond strip around strong polarity
199: inversion lines is an excellent predictor for the occurence of M and X
200: class flares within 24 hours of observation.  All these studies
201: implicate gradients in the magnetic field as being a crucial component
202: indicating the likelihood of activity (\opencite{cui};
203: \opencite{2002SoPh..209..171G}).
204: 
205: 
206: Section \ref{sec:multi} describes the \mres\ analyses used.  Section
207: \ref{sec:data} describes the data used in this study.  Section
208: \ref{sec:results} describes and discusses the results of the analysis
209: procedure.
210: 
211: \begin{table}
212: \begin{tabular}{cl}
213: class         & feature/classification rule \\
214: $\alpha$      & a single dominant spot often linked with a plage of opposite magnetic polarity\\
215: $\beta$       & a pair of dominant spots of opposite polarity\\
216: $\gamma$      & complex groups with irregular distribution of polarities\\
217: $\beta\gamma$ & bipolar groups which have more than one clear north-south polarity inversion line\\
218: $\delta$      & umbrae of opposite polarity together in a single penumbra\\
219: \end{tabular}
220: \caption{Mt. Wilson classification rules}
221: \label{tab:mtwilson}
222: \end{table}
223: 
224: 
225: \section{Magnetic flux data}
226: \label{sec:data}
227: The magnetic flux data used in this paper is the same as that used by
228: (\opencite{2005ApJ...631..628M}; also \opencite{2005SoPh..228...55M})
229: in their study of active region fractal dimension.  The analyzed
230: dataset is based on Solar and Heliospheric Observatory (SoHO)
231: Michelson Doppler Imager (MDI) images (\opencite{1995SoPh..162....1D};
232: \opencite{1995SoPh..162..129S}, and consists of extracted subfields
233: from full disk MDI magnetograms, centred on active regions present on
234: the disk. Full details of the method of extraction and correction can
235: be found in \opencite{2005ApJ...631..628M}.  The final dataset for use
236: in the present study consists of 19827 $600''\times600''$ FITS
237: (Flexible Image Transport System) files each centred on one or more
238: active regions.  Of these, only those within 60 degrees of disk centre
239: have their magnetic structure decomposed using the algorithm described
240: in Section \ref{sec:algorithm}.  Images more than 60 degrees away from
241: disk centre contain too many artifacts from projection effects and
242: field reversals for the \mres\ analyses to proceed safely.  This
243: leaves 9757 usable active region images.
244: 
245: \section{\Mres\ algorithms}
246: \label{sec:multi}
247: Active regions are complex objects, and to attempt to understand the
248: spatial distribution of their flux sources, any analysis must look at
249: all the \ls s available.  This arises from the observation
250: that active regions emerge and occur in many different shapes, sizes,
251: and configurations, and the fact that activity is often confined to
252: very small localized regions.  This points naturally to a
253: \mres\ approach to understanding the flux distribution.  However,
254: since active regions are complex objects, it is inevitable that a
255: single analysis will not capture all the information of interest, and
256: so, for comparative purposes, four \mres\ algorithms are implemented,
257: enabling cross-checking of results.
258: 
259: Two simple algorithms are chosen, a wavelet transforms (Mexican hat)
260: and one \mres\ morphological transform (\mres\ median transform).
261: These mother wavelets are chosen since they apepar briadly similar to
262: the features we are attempting to isolate.  However, the wavelet-based
263: algorithm does lead to the introduction of spurious features in the
264: transform (which has the possibility of misleading further analysis,
265: see Section \ref{sec:algorithm}) and hence a second, completely
266: different analysis algorithm (which does not create these features)
267: based on the median filter, is also used.
268: 
269: \subsection{Wavelet transforms}
270: \label{sec:mexican}
271: \begin{figure}
272:   \centerline{\hspace*{0.015\textwidth}
273:     \includegraphics[width=0.515\textwidth,clip=]{mexhat_shadesurf.eps}
274:     \hspace*{-0.03\textwidth}
275:     \includegraphics[width=0.515\textwidth,clip=]{mexhat_plot_image.eps}
276:   }
277:   \caption{Mexican hat mother wavelet, as a (a) surface and (b) contour
278:     plot. Interpreting the greyscale as opposite polarity flux, the
279:     single Mexican hat central flux concentration has a ring of opposite
280:     polarity flux.  These features are also evident in the transform at
281:     all \ls s. }
282:   \label{fig:mexhat}
283: \end{figure}
284: 
285: The continuous wavelet transform of a two-dimensional image $S$ in the
286: domain $D\subset R^{2}$ is
287: \begin{equation}
288: W(L,\underline{x}') = \int_{D}
289: \frac{1}{L}\psi
290: \left(
291: \frac{\underline{x}-\underline{x}'}{L}
292: \right)
293: S(\underline{x}) d\underline{x}
294: \end{equation}
295: where $L$ is the wavelet scale and $\underline{x}'$ translates
296: the wavelet across the domain $D$.  The Mexican hat wavelet transform
297: is continuous, and is implemented via the mother wavelet
298: \begin{equation}
299: \psi(\underline{x}) = c(2-|\underline{x}|^2)\exp\left(-|\underline{x}|^{2}/2 \right)
300: \label{eqn:mother}
301: \end{equation}
302: for some normalization constant $c$.  A plot of this mother wavelet is
303: shown in Figure \ref{fig:mexhat} (this wavelet is also used in the
304: analysis of \opencite{hewett}).  The wavelet is isotropic,
305: non-orthogonal, and is a good approximation to the shapes found in
306: active region magnetic fields.
307: 
308: 
309: 
310: 
311: \subsection{\Mres\ median transform}
312: \label{sec:median}
313: As noted above, the wavelet transforms used above are good at
314: identifying point and approximately circular features in images, but
315: have the unfortunate side effect of introducing a ``ringing'' which
316: appears as a false opposite polarity flux in the transform (Figure
317: \ref{fig:mexhat}, Figure \ref{fig:ex:trans}(I:c-h); the same effect
318: is also clearly visible in Hewett et al. 2007).  It is clearly
319: desirable to have a transform in which positive (negative) structure
320: in the image does not create negative (positive) structure in the
321: transform.  One such transform which has this property is the \mres\
322: median transform, based on the median filter.
323: 
324: The median transform for an image $S$ at \ls\ $L$ (denoted by
325: $\mbox{med}(S,L)$) is found by sliding a kernel of dimensions $L\times
326: L$ at all points in the image and calculating the median value\footnote{
327: If the list $X_{1},...,X_{N}$ is ordered such that $X_{j}\le X_{j+1}, \forall j, 
328: 1\le j \le N-1$, then \cite{2003psa..book.....W},
329: \[
330: \mbox{median}(X_{1},...,X_{N}) = \left\{
331: \begin{array}{cc}
332: X_{j} & \mbox{, $j=N/2 + 0.5$ if $N$ is odd}\\
333: (X_{j} + X_{j+1})/2 & \mbox{, $j=N/2$ if $N$ is even}.\\
334: \end{array}
335: \right.
336: \label{eqn:med}
337: \]
338: } of the $L\times L$ subimage extracted from $S$.  The \mres\
339: median transform on $N_{l}$ (specified by the user) \ls s is
340: found by through the following algorithm:
341: \begin{enumerate}
342: \item  Let $c_{j} = S$ with $j = 1$. 
343: \item Determine $c_{j +1} = \mbox{med}( S , 2l + 1 )$. 
344: \item The \mres\ coefficients $w_{j+1}$ are defined as: $w_{j+1} = c_{j}-c_{j +1}$ . 
345: \item Let $j \leftarrow j + 1$; $l\leftarrow 2l$. Return to step 2 if $j < N_{l}$ . 
346: \end{enumerate}
347: The original image $S$ may be reconstructed by $S = c_{N_{l}}
348: +\sum_{j=1}^{N_{l}}w_{j}$ where $c_{N_{l}}$ is the residual image left
349: after exiting the algorithm.  In step 4, the set of resolution levels
350: associated with $S$ lead to a dyadic decomposition.  A useful
351: feature of the median transform for analysis of active regions is that
352: the shapes of structures on the analyzed \ls s are closer to
353: those of the input image than would be case with a wavelet transform.
354: The \mres\ median transform acts like a median filter operating on
355: multiple length-scales.  \cite{gonzaleswoods} state that the median
356: filter forces neighboring pixels to become more like their neighbors.
357: This operation will preserve the structure of the field, whilst
358: ignoring noisy outliers that may skew wavelet based processing.
359: 
360: 
361: \subsection{An algorithm to define a Multi-scale opposite polarity region separator (\linename)}
362: \label{sec:algorithm}
363: \begin{figure}
364: \begin{tabbing}
365: \bf{Algorithm ``\linename''} \\
366: 1 be\=gin   \\
367: 2   \> R\= emove noise in image $S$ thresholding: pixels where $|B|<50$\\
368:     \>  \> are set to zero.   \\
369: 3   \> D\= ecompose $S$ by \mres\ transform onto $n$ \ls s, $L_{i}$ \\
370:     \>  \> to get transforms $W_{i}$, $1\le i \le n$. \\
371: 4   \> for \= i in range 1,n  \\
372: 5   \>          \> Find zero contours in $W_{i}$ \\
373: 6   \>          \> K\= eep pixel $p_{i,a,b}$ on the zero contour if there exists opposite \\
374:     \>          \>  \> polarity fields within a 3x3 box centred on that pixel.\\
375:     \>          \>  \> in the original image $S$. \\
376: 7   \>          \> Report the locations of all the lines at found at this \ls -  $M_{i}$ \\
377: 8   \>          \> Calculate the local transform power around each line in $M_{i}$.    \\
378: 9    \> end \\
379: 10 end \\
380: \end{tabbing}
381: \caption{Algorithm to find Multi-scale Opposite Polarity Region
382:   Separators (\linename).}\protect\label{fig:algorithm}
383: \end{figure}
384: 
385: At a given \ls, the wavelet transforms decompose the active region
386: into objects of the same \ls\ (Figures \ref{fig:ex:trans}(I:a-j).  The
387: \mres\ median transform is slightly different in that the active
388: region field structure is not directly compared to a given shape;
389: rather, the \mres\ coefficients arise from the distribution of
390: magnitudes of active region fluxes in the analysis window (note
391: however that the shape of window used to define the local data at that
392: \ls\ is square.)
393: 
394: The organization of the active region at different scale sizes and
395: different polarities is apparent.  This makes it possible to find
396: lines slicing the active region structure between opposing polarity
397: regions of a given scale size.  These lines are termed multi-scale
398: opposite polarity region separators (\linename), and they delineate
399: the organization of opposite polarity regions in the active region.
400: The algorithm describing the generation of \linename\ is given in
401: Figure \ref{fig:algorithm}.  These lines are geometrical constructions
402: based on the \mres\ analysis used, and since the transforms we are
403: using are different, we can expect the \linename\ defined to vary.
404: 
405: 
406: The \linename\ are used to {\it sample} the gradient of the magnetic flux
407: between regions of opposite polarity flux, at multiple \ls, using the
408: following algorithm.
409: \begin{enumerate}
410: \item Rank all the \linename-lines by local transform power,
411:   regardless of scale.
412: \item Create a mask by overlaying lines with larger local transform
413:   power over lines with smaller local transform power.  This enables
414:   overlapping \linename\ with stronger local wavelet power to be
415:   preferentially represented, since \linename\ with weaker local
416:   wavelet power are replaced.  Each \linename\ carries a label with the
417:   \ls\ it was found at.
418: \item Return gradient of original image $S$ at each point on the remaining
419:   \linename, and assign that gradient to the \linename's \ls\ label.
420: \end{enumerate}
421: Regardless of which \mres\ transform is used, the \linename\ exist to
422: sample the same underlying magnetic field gradients, and so each
423: \mres\ transform is analyzing the same complex, physical object.
424: Different \mres\ transforms will give different samples of the same
425: underlying magnetic field gradient, and the results obtained have to
426: be interpreted with respect to the properties of the transform in
427: mind.  However, this is the case for all decompositions.  For example,
428: one can decompose any one dimensional time series signal using any
429: orthogonal set; using Hermite polynomials is mathematically equivalent
430: to using the more familiar sinusoids, and the scale content of the
431: signal can be assessed.  However, the advantage of sinusoids is that
432: they are also the normal modes of vibration for a number of different
433: systems, and therefore have an additional meaning on top of their
434: convenience in understanding the scale structure of a time series.  It
435: is possible to understand the scale structure of a one dimensional
436: time series with Hermite polynomials, but more difficult.
437: Unfortunately, there is little or no extra guidance in choosing which
438: \mres\ decomposition will be the ``best'' for understanding the
439: complex structure of active regions, other than looking for mother
440: wavelets that represent the object we are looking at.  Therefore, in
441: lieu of a ``best possible choice'', we instead use multiple \mres\
442: transforms.
443: 
444: Figure \ref{fig:ex:trans}(I) implements the Mexican hat transform and
445: \linename\ analysis for some test data.  Figures
446: \ref{fig:ex:trans}(I:a) shows the test data, whilst (I:b) shows the
447: gradient of the image.  Figures \ref{fig:ex:trans}(I) show the Mexican
448: hat wavelet transform of the original image at multiple \ls s.
449: Overplotted in color are the \linename\ found at that \ls, along with the
450: rank of each line.  These lines are then plotted on the original image
451: and the gradient image, Figures \ref{fig:ex:trans}(I:c,d).  Discussion
452: of the transformed images and the concomitant gradient distributions
453: are presented below.
454: 
455: \begin{figure}
456:   \centerline{
457:     \hspace*{-0.015\textwidth}
458:     \includegraphics[width=0.98\textwidth,height=0.25\textheight,clip=]{mex.10.eps}
459:   }    
460:   \vspace*{0.005\textheight}
461:   \centerline{
462:     \hspace*{-0.015\textwidth}
463:     \includegraphics[width=0.98\textwidth,height=0.25\textheight,clip=]{med.10.eps}
464:   }
465:   \vspace{-0.80\textwidth}   % Shift close to the panel top 
466:   \centerline{\large \bf     % Includes the labels (here needs the color package)
467:     \hspace{0.0 \textwidth} \color{white}{(I)}
468:     \hfill}
469:   \vspace{0.375\textwidth}    % Shift back to the panel bottom 
470:   \centerline{\large \bf     % Includes the labels (here needs the color package)
471:     \hspace{0.0 \textwidth} \color{white}{(II)}
472:     \hfill}
473:   \vspace{0.35\textwidth}    % Shift back to the panel bottom 
474:   \caption{ Panels (I): an example \linename\ analysis using the Morlet
475:     wavelet (Section \ref{sec:mexican}). Panel I(a) shows some test
476:     data, I(b) the magnitude of the gradient, and I(i), I(j) show
477:     I(a), I(b), respectively with \linename\ overlaid.  Panels I(c-h) show
478:     the wavelet transform of test data at length-scale $s$, with the
479:     \linename\ found at that length-scale.  Each line is labeled with its
480:     local wavelet power rank.  Panels (ii): the same data is analyzed
481:     by the median transform.  Panels II(a,b,i,j) have the same meaning
482:     as panels I(a,b,i,j). Panels II(c-h) show the median transform of
483:     the test data at length-scale $s$, with the \linename\ found at that
484:     length-scale.  Each line is labeled with its local median
485:     transform power rank.}
486:   \label{fig:ex:trans}
487: \end{figure}
488: 
489: 
490: \section{Results}
491: \label{sec:results}
492: The \mres\ analysis described above generates a large amount of
493: information for each active region.  Gradient information found along
494: the \linename s measures the distribution of gradients in the
495: line-of-sight magnetic field as a function of the \ls\ of opposite
496: polarity regions in active regions.  Summary statistics of the
497: gradient distributions (after grouping these according to Mt. Wilson
498: class, flaring activity and \ls) are discussed below, for the 9757
499: active region observations described in Section \ref{sec:data}.
500: 
501: The discussion begins with examining the behaviors of the transforms
502: themselves.
503: 
504: %
505: % alpha
506: %
507: \begin{figure}
508:   \centerline{
509:     \hspace*{-0.015\textwidth}
510:     \includegraphics[width=0.98\textwidth,height=0.25\textheight,clip=]{cor_mdi_ar7986_19960901_1251cor.fits.mex.ALPHA.all.eps}%{cor_mdi_ar7995_19961114_1251cor.fits.mex.all.eps}
511:   }    
512:   \vspace*{0.005\textheight}
513:   \centerline{
514:     \hspace*{-0.015\textwidth}
515:     \includegraphics[width=0.98\textwidth,height=0.25\textheight,clip=]{cor_mdi_ar7986_19960901_1251cor.fits.med.ALPHA.all.eps}%{cor_mdi_ar7995_19961114_1251cor.fits.med.eps}
516:   }
517:   \vspace{-0.80\textwidth}   % Shift close to the panel top 
518:   \centerline{\large \bf     % Includes the labels (here needs the color package)
519:     \hspace{0.0 \textwidth} \color{white}{(I)}
520:     \hfill}
521:   \vspace{0.375\textwidth}    % Shift back to the panel bottom 
522:   \centerline{\large \bf     % Includes the labels (here needs the color package)
523:     \hspace{0.0 \textwidth} \color{white}{(II)}
524:     \hfill}
525:   \vspace{0.35\textwidth}    % Shift back to the panel bottom 
526:   \caption{Comparison of Mexican hat and median transform \linename\
527:     for an $\alpha$ region (NOAA AR 7995, 1996/11/14 12:51 UT). See
528:     Figure \ref{fig:ex:trans} for a description of the ordering of the
529:     plots.  A blank plot denotes that no information was found at that
530:     scale.  The region is very simple, and only a very few pixels are
531:     retained as \linename\ in the median transform compared to the
532:     Mexican hat results.}
533:   \label{fig:mexmed:a}
534: \end{figure}
535: %
536: % beta-gamma-delta
537: %
538: \begin{figure}
539:   \centerline{
540:     \hspace*{-0.015\textwidth}
541:     \includegraphics[width=0.98\textwidth,height=0.25\textheight,clip=]{cor_mdi_ar551_20040210_1251cor.fits.mex.BETAGAMMADELTA.all.eps}
542:   }    
543:   \vspace*{0.005\textheight}
544:   \centerline{
545:     \hspace*{-0.015\textwidth}
546:     \includegraphics[width=0.98\textwidth,height=0.25\textheight,clip=]{cor_mdi_ar551_20040210_1251cor.fits.med.BETAGAMMADELTA.all.eps}
547:   }
548:   \vspace{-0.80\textwidth}   % Shift close to the panel top 
549:   \centerline{\large \bf     % Includes the labels (here needs the color package)
550:     \hspace{0.0 \textwidth} \color{white}{(I)}
551:     \hfill}
552:   \vspace{0.375\textwidth}    % Shift back to the panel bottom 
553:   \centerline{\large \bf     % Includes the labels (here needs the color package)
554:     \hspace{0.0 \textwidth} \color{white}{(II)}
555:     \hfill}
556:   \vspace{0.35\textwidth}    % Shift back to the panel bottom 
557:   \caption{Comparison of Mexican hat and median transform \linename\
558:     for a $\beta\gamma\delta$ region (NOAA AR 2004/02/10 12:51 UT).
559:     The left two columns are Mexican Hat results, the right two
560:     columns are median transform results.  The individual plots for
561:     each type transform are (a) original magnetogram, (b) original
562:     magnetogram overplotted with \linename\ from all \ls s, (c-h)
563:     \mres\ transforms at \ls s L=1,2,4,8,16 and 32 pixels
564:     respectively, overplotted with \linename\ found at that scale.
565:     Note that the thickness of the \linename\ are weighted by the
566:     local magnetic flux gradient found at that location.  Both
567:     transforms capture the location of opposing polarities at small
568:     scales quite well.  The Mexican hat transform appears to slice the
569:     large scale structure of the region better than the median
570:     transform.}
571:   \label{fig:mexmed:bgd}
572: \end{figure}
573: 
574: 
575: \subsection{Comparative behaviour of the transforms}
576: \label{sec:compare}
577: It is apparent from the example transforms given in Figures
578: \ref{fig:ex:trans}, \ref{fig:mexmed:a} and \ref{fig:mexmed:bgd} that
579: each transform captures different information about the active region.
580: For example, at large scales, Figures \ref{fig:ex:trans}(I:g,h) and
581: \ref{fig:mexmed:a}(I:g,h) divide the plane into approximately two
582: regions of opposite polarity, capturing the large scale distribution
583: of the field (Figure \ref{fig:ex:trans}(I:a),
584: \ref{fig:mexmed:a}(I:a)). The median transform does not return any
585: structure at the largest scale size (Figure \ref{fig:ex:trans}(II:h),
586: \ref{fig:mexmed:a}(II:h)).  The ``ringing'' of the Mexican
587: hat transform is also apparent in Figures \ref{fig:ex:trans}(I:c-h),
588: brought about by the shape of the transform (Section
589: \ref{sec:mexican}).  Both \mres\ analyses report \linename\ where one
590: would expect them, but give them different rankings.  In particular,
591: the highest ranked \linename\ occurs at different (and neighboring \ls
592: s), but at the same location.
593: 
594: 
595: The median transform is much more localized, and so even although it
596: does also find \linename\ at larger \ls s, the pixels retained as
597: having significant gradients are far fewer.  This is due to the
598: interaction of the \linename\ algorithm (Figure \ref{fig:algorithm})
599: with the properties of the transform.  The median transform cleaves
600: more closely to the true shape of the regions
601: \cite{1998ipda.book.....S} at that \ls, whereas the Mexican hat
602: transform imposes a shape on the active region.  As very large scale
603: contiguous features are not very common in active regions, they are
604: not there for the median transform to find, and so \linename\ at these
605: \ls s are deprecated compared to the Mexican hat transform.  This is
606: apparent in Figure \ref{fig:mexmed:bgd}(II:g,h), where the large scale
607: organization of the field is found via the Mexican hat transform
608: because the transform at that scale is smoothing out the original data
609: over a large \ls.  This is analogous to the situation in Morlet
610: wavelet time series analysis, where a spike (delta function) in the
611: time series can leads to power at larger scales, even although there
612: is no real variation at those longer scales in the data
613: \cite{2004SoPh..222..203D}.  The median transform does not detect any
614: information at this \ls\ (Figure \ref{fig:mexmed:bgd}(II:g,h).
615: 
616: Line 5 in the \linename\ algorithm (Figure \ref{fig:algorithm}) looks
617: for zero contours in the transform, and so can be susceptible to the
618: ringing effect of the Mexican hat transform.  This effect is
619: suppressed by line 6 of the \linename\ algorithm, by referring back to
620: the original data to look for pixels on the \linename\ which do have
621: some opposite polarity flux around them.  The overall result is that
622: the Mexican hat transform is quite effective at dividing the active
623: region into intuitively satisfying regions, at a given \ls.  The
624: median transform is more effective at retaining the local shape of
625: opposite polarity regions at a given \ls, and hence more effective at
626: describing where opposite polarity flux is at close proximity. The
627: \linename\ ranking algorithm also reports the most ``important''
628: \linename\ at different \ls\ (Figures \ref{fig:ex:trans},
629: \ref{fig:mexmed:bgd}).  This is not very surprising, as the two \mres\
630: transforms are measuring very different properties of the structure.
631: In general, it is found that the median transform assigns \linename\
632: to smaller \ls\ than the Mexican hat transform, consistent with the
633: behavior noted by \opencite{1998ipda.book.....S}.  The results quoted
634: below from each method are consistent as a function of \ls, as defined
635: by each analysis method.  Since the results are qualitatively the same
636: for each method, this confirms that each method is indeed internally
637: consistent, and both are measuring the same property of the active
638: region field.
639: 
640: \subsection{Gradients along \linename}
641: \label{sec:res:rlm}
642: \begin{figure}
643:   \centerline{\hspace*{0.0\textwidth}
644:     \includegraphics[width=0.5\textwidth,viewport =  0 0 275 360, clip=]{decomp6_sch.rlm.eps}
645:     \hspace*{0.0\textwidth}
646:     \includegraphics[width=0.5\textwidth,viewport =  0 0 275 360, clip=]{decomp6_sch_median.rlm.eps}
647:   }
648:   \caption{Measurements of the gradient statistics as a function of
649:     multi-resolution transform (Mexican hat, a-c; median transform,
650:     d-f), Mt. Wilson classification ($\alpha$, solid line style;
651:     $\beta$, dotted; $\beta\gamma$, dashed; $\beta\gamma\delta$,
652:     dot- dashed) and \linename\ \ls (thick lines with plot
653:     symbols). Plots (a,d) show the average {\it maximum} gradient
654:     found on each \linename; plots (b,e) show the average {\it
655:       average} gradient on each \linename\ and plots (c,f) show the
656:     average {\it standard deviation} of the gradient found on each
657:     \linename.  Also indicated are the total number of images that
658:     have at least one \linename\ at any \ls\ and return a valid
659:     gradient measurement, as a function of Mt. Wilson classification
660:     (thin lines, same line style as the Mt. Wilson classification,
661:     numbers on left hand plot axis).
662:     These can vary per plot since occasionally the analysis will find
663:     a \linename\ that will return a gradient statistic which is
664:     undefined: for example, occasionally the \linename\ consists of a
665:     single point - in such a case the standard deviation of the
666:     gradient underlying this point is undefined and so must be
667:     excluded from further analysis.  See Section \ref{sec:res:rlm}
668:     for more discussion of this result.}
669: \label{fig:res:rlm}
670: \end{figure}
671: Figure \ref{fig:res:rlm}(a-c,d-f) shows the results of
672: measuring the gradients along the \linename\ for the active region
673: data set for the Mexican hat and median transforms respectively.  Four
674: different statistics are returned; (a,d), maximum gradient found along
675: the \linename\, (b,e), average gradient found along the \linename\,
676: (c,f), standard deviation of the gradient along the \linename\.  In
677: all cases, the average value of the quantity is plotted as a function
678: of \ls\ and Mt. Wilson class (gradients across the \linename\ are
679: measured in arbitrary units).
680: 
681: Figure \ref{fig:res:rlm}(a) is indicative of the other plots in this
682: figure, regardless of the transform used.  Over all \ls s, the
683: \bgdcl-class has larger average maximum gradients than the
684: \bgcl-class, which is larger than the \bcl-class, which is larger than
685: the \acl-class (excepting where low numbers of results mean that good
686: averages cannot be obtained).  The same ordering holds for all other
687: plots.  Hence when an active region moves from one class to another,
688: on average, the change in the geometrical structure of the active
689: region is reflected by a change in gradient content at all \ls s, and
690: not on any particular \ls.  All measures of the gradient given here
691: exhibit the same property - gradients in active region classes
692: regarded as being more likely to flare have larger maximum gradients,
693: larger average gradients, and larger standard deviations along the
694: \linename.  Figures \ref{fig:res:rlm}(c,f) suggest a greater
695: variability in the more active classes, and so suggest that more
696: active classes exhibit a wider range of gradient conditions, again, at
697: all \ls s.  This indicates that on average, the gradient content
698: between opposite polarity regions maintains the same Mt. Wilson
699: classification order, regardless of whether the field is understood as
700: being organized on small scales, or large scales.
701: 
702: \subsection{Weighted \ls\ versus weighted gradients}
703: \label{sec:weighted}
704: \begin{figure}
705:   \vspace*{0.05\textheight}
706:   \centerline{
707:     \hspace*{0.015\textwidth}
708:     \includegraphics[width=1.0\textwidth,height=0.40\textheight,clip=]{decomp6_sch.wav.eps}
709:   }    
710:   \vspace*{0.02\textheight}
711:   
712:   \centerline{
713:     \hspace*{0.000\textwidth}
714:     \includegraphics[width=1.0\textwidth,height=0.40\textheight,clip=]{decomp6_sch_median.wav.eps}
715:   }
716:   \vspace{-1.38\textwidth}   % Shift close to the panel top 
717:   \centerline{\large \bf     % Includes the labels (here needs the color package)
718:     \hspace{0.0 \textwidth} \color{black}{(I): Mexican hat}
719:     \hfill}
720:   \vspace{0.655\textwidth}    % Shift back to the panel bottom 
721:   \centerline{\large \bf     % Includes the labels (here needs the color package)
722:     \hspace{0.0 \textwidth} \color{black}{(II): median transform}
723:     \hfill}
724:   \vspace{0.68\textwidth}    % Shift back to the panel bottom 
725: %   \vspace{-0.80\textwidth}   % Shift close to the panel top 
726: %   \centerline{\large \bf     % Includes the labels (here needs the color package)
727: %     \hspace{0.0 \textwidth} \color{white}{(I)}
728: %     \hfill}
729: %   \vspace{0.375\textwidth}    % Shift back to the panel bottom 
730: %   \centerline{\large \bf     % Includes the labels (here needs the color package)
731: %     \hspace{0.0 \textwidth} \color{white}{(II)}
732: %     \hfill}
733: %   \vspace{0.33\textwidth}    % Shift back to the panel bottom 
734:   \caption{Weighted average \ls\ ($\Lambda$) versus weighted average
735:     gradient statistic ($\Psi$(maximum gradient),$\Psi$(average
736:     gradient) and $\Psi$(standard deviation of gradient) ).
737:     Horizontal/vertical lines are annotated with the weighted average
738:     gradient/\ls\ of the distribution.  Also indicated on each plot is
739:     the maximum number of counts in a single histogram bin ('m') and
740:     the total number of counts in the distribution ('\#').  See
741:     Section \ref{sec:weighted} for more details on the calculation of
742:     weighted averages.}
743:   \label{fig:res:wav}
744: \end{figure}
745: The previous results show that active region fields may on average, be
746: ordered by Mt. Wilson classification without regard to any given \ls.
747: The above study also shows that gradients exist between objects of
748: many different \ls\ (active region fields are known to be
749: multi-fractal: \opencite{conlon}; \opencite{2005SoPh..228....5G};
750: \opencite{1996ApJ...465..425L}; \opencite{1993ApJ...417..805L} ).
751: However, a representative \ls s and gradient statistic can be defined
752: for a given active region, which summarizes the \mres\ nature of the
753: active region.  Assume that the \linename\ $q_{i}$ is found at \ls\
754: $L_{q_{i}}$ and has gradient statistic (either a maximum, average or
755: standard deviation of gradient) $G_{q_{i}}$.  Further, the local
756: wavelet power around $q_{i}$ is $w_{q_{i}}$.  If there are $n$
757: \linename s, $1\le i \le n$ for a given active region then the {\it
758:   weighted average \linename\ \ls} is
759: \begin{equation}
760: \label{eqn:Lambda}
761: \Lambda = \sum_{i=1}^{n}w_{q_{i}}{L_{q_{i}}}/\sum_{i=1}^{n}w_{q_{i}}
762: \end{equation}
763: and similarly, the {\it weighted average gradient statistic} is
764: \begin{equation}
765: \label{eqn:Psi}
766: \Psi = \sum_{i=1}^{n}w_{q_{i}}{G_{q_{i}}}/\sum_{i=1}^{n}w_{q_{i}}.
767: \end{equation}
768: Weighting by the local wavelet power (the same local wavelet power of
769: Section \ref{sec:algorithm}) takes into account the ``importance'' of the
770: \linename\ found at that location.
771: 
772: Figure \ref{fig:res:wav} plots frequency distributions of $\Lambda$
773: and $\Psi$ for both \mres\ analyses, for all 9757 magnetograms,
774: binned by Mt. Wilson classification.  Although the distributions are
775: highly scattered, it is clear that in all measures, and for both
776: \mres\ analyses, that average values of $\Lambda$ and $\Psi$
777: increase in the order
778: \acl$\rightarrow$\bcl$\rightarrow$\bgcl$\rightarrow$\bgdcl.  As the
779: Mt. Wilson classification increases in the above order, the
780: organization of the \linename\ moves to progressively longer \ls, and
781: the gradient on these separators also increases.  Hence there are
782: longer separators, with larger gradients, as the Mt. Wilson class
783: increases to those known to have a greater chance of activity.  The
784: order \acl$\rightarrow$\bcl$\rightarrow$\bgcl$\rightarrow$\bgdcl\ also
785: generally connotes an increase in active region size.  However, an
786: increasing size does not necessarily mean that there are longer
787: \linename\ between opposite polarity flux regions on those longer \ls;
788: it depends entirely on the locations and proximity of the opposite
789: polarity field as it emerges.  For a given size, a \bgdcl\ active
790: region has much more opposite polarity structure than other Mt. Wilson
791: classes (indeed, it is implicit in the definitions - see Table
792: \ref{tab:mtwilson}), and so the increase in $\Lambda$ and all three
793: $\Psi$ statistics (maximum gradient on a \linename, average gradient
794: on a \linename, and the standard deviation of the gradient on a
795: \linename), indicate the presence of increasingly diverse field
796: structure.  The qualitative result is the same for both the \mres\
797: results, indicating that both analyses are capturing consistent
798: behavior.
799: 
800: 
801: \subsection{Gradients in flaring and non-flaring active regions}
802: \label{sec:fnf}
803: \begin{figure}
804:   \vspace*{0.05\textheight}
805:   \centerline{
806:     \hspace*{0.015\textwidth}
807:     \includegraphics[width=1.0\textwidth,height=0.40\textheight,clip=]{an_decomp_flare7.flare_and_nonflare.mex.eps}
808:   }    
809:   \vspace*{0.02\textheight}
810:   
811:   \centerline{
812:     \hspace*{0.000\textwidth}
813:     \includegraphics[width=1.0\textwidth,height=0.40\textheight,clip=]{an_decomp_flare7.flare_and_nonflare.med.eps}
814:   }
815:   \vspace{-1.38\textwidth}   % Shift close to the panel top 
816:   \centerline{\large \bf     % Includes the labels (here needs the color package)
817:     \hspace{0.0 \textwidth} \color{black}{(I): Mexican hat}
818:     \hfill}
819:   \vspace{0.655\textwidth}    % Shift back to the panel bottom 
820:   \centerline{\large \bf     % Includes the labels (here needs the color package)
821:     \hspace{0.0 \textwidth} \color{black}{(II): median transform}
822:     \hfill}
823:   \vspace{0.68\textwidth}    % Shift back to the panel bottom 
824: %   \vspace{-0.80\textwidth}   % Shift close to the panel top 
825: %   \centerline{\large \bf     % Includes the labels (here needs the color package)
826: %     \hspace{0.0 \textwidth} \color{white}{(I)}
827: %     \hfill}
828: %   \vspace{0.375\textwidth}    % Shift back to the panel bottom 
829: %   \centerline{\large \bf     % Includes the labels (here needs the color package)
830: %     \hspace{0.0 \textwidth} \color{white}{(II)}
831: %     \hfill}
832: %   \vspace{0.33\textwidth}    % Shift back to the panel bottom 
833:   \caption{Distributions of the maximum gradient found as a function
834:     of multi-resolution transform, analyzing gradient and flare
835:     occurence (only those flares that occur no later than six hours
836:     after the active region gradient measurement are included).  In
837:     all plots, the dotted lines refer to active regions having no
838:     flares occuring in them.  The thin solid line refers to active
839:     regions having at least one flare of class 'A' or above ({\it
840:       all-flare distribution}); the thick solid line refers to active
841:     regions having class 'M' or above ({\it large-flare
842:       distribution}).  Vertical lines without asterisks indicate the
843:     mean value of the gradient distribution; vertical lines with
844:     asterisks indicate the median value of the gradient distribution.
845:     The quantity $\theta$ is the test quantity calculated to apply the
846:     Kolmogorov-Smirnov test.  It is applied to test if the all-flare
847:     and large-flare distributions are different; values above 1.36
848:     confirm the null hypothesis (that the two distributions are drawn
849:     from the same underlying distribution) less than 5\% of the time -
850:     see Section \ref{sec:fnf} for more detail. }
851:   \label{fig:res:fnf}
852: \end{figure}
853: 
854: As suggested in the introduction, field gradients have long been
855: associated with flaring activity.  In this study, active regions are
856: split into three sets in order to quantify the relationship between
857: the presence of gradients between opposite polarity of different size
858: scales and the occurance and size of flares.  The first set of active
859: regions have no flaring activity associated with them at any time.
860: The second set of magnetograms (the {\it all-flare} set) have at least
861: one GOES 'A1.0' class flare, or more energetic, occuring no more than
862: six hours after the active region magnetogram.  The third set of
863: magnetograms (the {\it large flare} set ) have at least a cumulative
864: flare index equivalent to a GOES 'M1.0' class flare, or more
865: energetic, ocurring no more than six hours after the active region
866: magnetogram.  \linename\ for these magnetograms are calculated, and
867: distributions of the {\it maximum} gradients found are shown in Figure
868: \ref{fig:res:fnf}.
869: 
870: It is clear that, on average, the gradient content of a flaring active
871: region field is very different from that of a non-flaring active
872: region field.  Further, this difference is apparent on almost all \ls;
873: it does not matter which size scale one considers the opposite
874: polarity regions to be the one at which this difference is best
875: measured.  The difference between the all-flare and large-flare
876: distributions is less pronounced, but measurable both by looking at
877: the mean and median values of gradients, and also by the
878: Kolmogorov-Smirnov test.  The Kolmogorov-Smirnov test for two samples
879: (\opencite{wasserman}; \opencite{keeping}; \opencite{barlow}) treats
880: the agreement between two cumulative distributions\footnote{For a
881:   probability density function $f(x)$ the cumulative distribution
882:   function is $G(x) = \int_{0}^{x}f(y)dy$} $G_{1}$ and $G_{2}$ against
883: the null hypothesis that the two distributions are samples come from
884: the same underlying distribution function.  The first step is to
885: calculate
886: \begin{equation}
887: \label{eqn:KSA}
888: d = \max\left|G_{1}(x) - G_{2}(x)\right|.
889: \end{equation}
890: If $G_{1,2}$ has $n_{1,2}$ samples then it can be shown that 
891: \begin{equation}
892: \label{eqn:KSB}
893: \lim_{n_{1},n_{2}\rightarrow \infty} \mbox{prob}\left(
894: \theta \ge \lambda \right) = 2\sum_{m=1}^{\infty}(-1)^{m+1}e^{-2m^{2}\lambda^{2}}
895: \end{equation}
896: where $\theta = d\sqrt{N}$ and $N=n_{1}n_{2}/(n_{1} + n_{2})$.  Values
897: of $\theta \ge 1.36$ reject the null hypothesis at greater than or
898: equal to the 5\% level, that is, on average five times out of a
899: hundred the two cumulative distributions come from the same underlying
900: distribution function ($\lambda = 1.63$ is equivalent to 1\%,
901: $\lambda=1.73$, 0.5\% and $\lambda=1.95$, 0.1\%).  It is a
902: nonparametric test and is thus appropriate to use in this case where
903: the true gradient distribution is unknown.  
904: 
905: The Kolmogorov-Smirnov test shows that the Mexican hat derived
906: distributions for all-flare and large-flare maximum gradients, reject
907: the null hypothesis with a high degree of confidence, for all the \ls\
908: studied.  The median transform results also reject the null hypothesis
909: with a high degree of confidence, but only for the first four \ls s -
910: the hull hypothesis cannot be rejected at \ls = 16 and 32 pixels
911: (probably due to insufficient data at these \ls\ - see Sections
912: \ref{sec:median} and \ref{sec:compare}).  Hence the difference between
913: active regions that contain any type of flare, and active regions that
914: contain large flares is measureable in the active region gradient
915: content, without regard to which opposite polarity region \ls\ is
916: considered.
917: 
918: 
919: \section{Conclusions}
920: \label{sec:discuss}
921: 
922: Active region magnetic fields are known to exhibit multi-fractal
923: properties. The study presented here connects the geometrical
924: arrangement of opposite polarity regions (via the \linename\ concept)
925: to the \ls\ properties of the field.  There appears to be no special
926: \ls\ that determines the gradient content of the active region field,
927: which agrees with the notion from previous fractal and multi-fractal
928: studies that there is no preferred \ls\ to the spatial size of the
929: absolute valoe of active region flux elements.
930: 
931: The Mt. Wilson classification does capture some structure information
932: on the field, and through that, gradient content (Figures
933: \ref{fig:res:rlm} and \ref{fig:res:wav}).  The Mt. Wilson
934: classification is based on the very largest \ls\ describing the
935: geometrical arrangement of the field.  Figure \ref{fig:res:rlm} shows
936: that regardless of the \ls\ used, the gradient content of different
937: classes of active region field can be distinguished on average.  This
938: is a result of the multi-scale nature of the active region magnetic
939: field.  Figure \ref{fig:res:wav} shows that even although all
940: Mt. Wilson classes retain their multi-scale structure (i.e., gradients
941: at all \ls\ are present), there is a general increase in the average
942: \ls\ and gradient as a function of Mt. Wilson class, in the order
943: \acl$\rightarrow$\bcl$\rightarrow$\bgcl$\rightarrow$\bgdcl).  Not only
944: is the active region field appearing at larger \ls, the organization
945: of its gradients is also appearing at larger \ls, and those gradients
946: are getting larger.
947: 
948: The equivalence of all \ls s in the active region magnetic field is
949: also shown in Figure \ref{fig:res:fnf}; at all \ls, there are
950: measurable differences in the gradient content for active regions that
951: give rise to large flares, active regions that give rise to all sizes
952: of flares, and active regions that do not have flares.  On average,
953: one can distinguish the difference between these types of active
954: region by considering the geometrical arrangement of the opposite
955: polarity flux at any \ls.  The geometrical arrangement of opposite
956: polarity field preserves the multi-scale nature of the active region
957: at all \ls, through to the gradient content of the field.
958: 
959: When an active region breaks through the photosphere, injection
960: happens at multiple \ls s.  The largest gradients between opposite
961: polarity regions are observed at the smallest \ls\ (see Figure
962: \ref{fig:res:rlm}(a,d)).  Although the largest gradients may be at the
963: smallest \ls s, gradients also exist between opposite polarity regions
964: of longer \ls.  This study suggests that indicators of flare activity
965: and Mt. Wilson classification exist at all \ls s, and that all active
966: regions \ls\ are equivalent, due to the multi-scale properties of the
967: field. 
968: 
969: 
970: \begin{acks}
971:   This work was supported by NASA Living With a Star Targeted Research
972:   and Technology award NNH04CC65C. SOHO is a joint project of
973:   international co-operation by ESA and NASA.
974: \end{acks}
975: 
976: 
977: % format of references provided by the journal (.bst)
978: \bibliographystyle{spr-mp-sola}
979: 
980: % name your Bibtex file containing your references (.bib)
981: \bibliography{ip}  
982: 
983: 
984: 
985: \end{article} 
986: \end{document}
987: 
988: