0810.4135/ms.tex
1: \documentclass[12pt,preprint]{aastex}
2: 
3: \begin{document}
4: 
5: \title{Polarization Patterns in Pulsar Radio Emission}
6: \author{Mark M. McKinnon}
7: \affil{National Radio Astronomy Observatory,\altaffilmark{1}\altaffiltext{1}
8: {The National Radio Astronomy Observatory is a facility of the National 
9: Science Foundation operated under cooperative agreement by Associated 
10: Universities, Inc.} Socorro, NM \ \ 87801\ \ USA}
11: 
12: \begin{abstract}
13: 
14: A variety of intriguing polarization patterns are created when polarization 
15: observations of the single pulses from radio pulsars are displayed in a 
16: two-dimensional projection of the Poincar\'e sphere. In many pulsars, the 
17: projections produce two clusters of data points that reside at antipodal 
18: points on the sphere. The clusters are formed by fluctuations in polarization 
19: amplitude that are parallel to the unit vectors representing the polarization 
20: states of the wave propagation modes in the pulsar magnetosphere. In other 
21: pulsars, however, the patterns are more complex, resembling annuli and bow 
22: ties or bars. The formation of these complex patterns is not understood and 
23: largely unexplored. An empirical model of pulsar polarization is used to show 
24: that these patterns arise from polarization fluctuations that are perpendicular 
25: to the mode vectors. The model also shows that the modulation index of the 
26: polarization amplitude is an indicator of polarization pattern complexity.
27: A stochastic version of generalized Faraday rotation can cause the orientation 
28: of the polarization vectors to fluctuate and is a possible candidate for the 
29: perpendicular fluctuations incorporated in the model. Alternative models 
30: indicate that one mode experiences perpendicular fluctuations and the other 
31: does not, suggesting that the fluctuations could also be due to a mode-selective 
32: random process, such as scattering in the magnetosphere. A polarization stability 
33: analysis of the patterns implies that processes intrinsic to the emission are 
34: more effective in depolarizing the emission than fluctuations in the orientation 
35: of its polarization vector.
36: 
37: \end{abstract}
38: 
39: \keywords{methods: analytical -- polarization -- plasmas -- pulsars: general
40:           -- pulsars:individual (PSR B0823+26, PSR B1929+10, PSR B2020+28)}
41: 
42: \section{INTRODUCTION}
43: 
44: Pulsar radio emission is renowned for its complicated polarization behavior.
45: The polarization is generally elliptical, yet predominantly linear, and highly 
46: variable, often switching between two, orthogonally polarized states (e.g. 
47: Stinebring et al. 1984). A recent trend in the display and interpretation of 
48: the results from polarization observations of individual pulses has been to 
49: plot the measured values of the polarization vector's orientation angles at 
50: a given rotational phase of the pulse in a two-dimensional projection of the 
51: Poincar\'e sphere. The projections reveal a wide variety of organized 
52: polarization patterns. For example, towards the center of the pulse in PSR 
53: B2020+28, the angles reside in a single cluster in one hemisphere of the 
54: Poincar\'e sphere (McKinnon 2004). The orientation angles form a diffuse 
55: structure resembling a bow tie near the pulse center of PSR B0818--13 
56: (Edwards 2004). At the peak of PSR B1133+16, a Hammer-Aitoff projection of 
57: the angles produces two data clusters, each in a separate hemisphere of the 
58: Poincar\'e sphere (Karastergiou et al. 2003). The Lambert equal-area 
59: projections (LEAPs) of the orientation angles in PSR B0329+54 reveal two 
60: extremes in polarization behavior within the pulsar's pulse (Edwards \& 
61: Stappers 2004). In the cone emission on the edges of the pulse, the angles 
62: reside in two, circularly-shaped, bipolar clusters, similar to the angles in 
63: PSR B1133+16. But within the pulsar's core emission at the pulse center, one 
64: of the two clusters stretches into an ellipse or bar, while the other spreads 
65: into an intriguing partial annulus. An accurate description of these patterns 
66: is needed to determine the physical processes that create them.
67: 
68: The origin of the simple polarization patterns is generally understood 
69: within the context of the statistical model of McKinnon \& Stinebring (1998, 
70: 2000), aided by the analyses of McKinnon (2004) and Edwards \& Stappers
71: (2004). A single, compact cluster is produced when the fluctuations in the 
72: Stokes parameters Q, U, and V are comparable to one another, but less than 
73: the mean value of the polarization vector's amplitude, as might be expected 
74: for the measurement of a polarization vector with fixed amplitude and 
75: orientation accompanied by instrumental noise. Orthogonal polarization modes 
76: (OPMs) produce the projections showing two clusters of orientation angles. 
77: The unit vectors representing the mode polarizations are antiparallel to each 
78: other and form a diagonal in the Poincar\'e sphere. Since the mode vectors are 
79: antiparallel, the amplitude of the resultant polarization is the difference 
80: between the mode polarization amplitudes, and the tip of the resultant 
81: polarization vector will reside in one hemisphere of the Poincar\'e sphere 
82: or the other depending upon which mode is the stronger of the two. The 
83: instantaneous orientation of the polarization vector alternates randomly between 
84: hemispheres due to temporal fluctuations in the polarized intensity of each 
85: mode that could be caused by the radio emission mechanism or propagation 
86: effects in the pulsar magnetosphere, such as the birefringence of the modes 
87: (e.g. Allen \& Melrose 1982; Barnard \& Arons 1986). The resulting polarization 
88: fluctuations caused by the OPMs are therefore parallel to the mode diagonal.
89: 
90: The clues to the origin of the more complicated polarization patterns are
91: provided in the analyses of McKinnon (2004) and Edwards \& Stappers (2004), 
92: the polarization observations of Edwards (2004), and the numerical simulations 
93: of Melrose et al. (2006). The shape of the Q-U-V data point clusters created 
94: by the polarization fluctuations is generally an ellipsoid. McKinnon and 
95: Edwards \& Stappers independently developed a technique to quantify the 
96: polarization fluctuations by measuring the ellipsoid's dimensions. The 
97: technique calculates the covariance matrix of the Stokes parameters and
98: subsequently determines the matrix eigenvalues. The three dimensions of the 
99: polarization ellipsoid are related to the eigenvalues, and the eigenvectors 
100: are the three orthogonal axes of the ellipsoid. For some objects, such as 
101: PSR B0809+74 (Edwards 2004), PSR B1929+10, and PSR B2020+28 (McKinnon 2004), 
102: two of the eigenvalues are roughly equal but less than the third, which is 
103: what one would expect for a prolate ellipsoid fashioned by OPMs. But Edwards
104: (2004) also found examples (e.g. PSR B0320+39 and PSR B0818--13) where all 
105: three eigenvalues are different, proving that the polarization fluctuations 
106: possess a component that is {\it perpendicular} to the mode diagonal, in 
107: addition to the {\it parallel} component produced by the OPMs. Melrose et
108: al. (2006) had to incorporate fluctuations in the orientation of the mode 
109: diagonal to replicate the polarization patterns observed by Edwards \& 
110: Stappers in PSR B0329+54. Randomly varying orientation angles equate to 
111: both parallel and perpendicular components in the polarization fluctuations. 
112: Karastergiou et al. (2003) also suggested that random variations in the 
113: orientation of the mode diagonal might be needed to explain their 
114: observations of PSR B1133+16. 
115: 
116: McKinnon \& Stinebring (1998; 2000) proposed that the polarization of the 
117: radio emission is determined by the incoherent superposition of two, highly 
118: polarized, orthogonal modes. The assumption of highly polarized modes has 
119: strong theoretical support on the grounds that any plasma has two natural 
120: modes of wave propagation that are completely polarized (Petrova 2001). The 
121: assumption of incoherent modes means the modes propagate independently, 
122: which has two important consequences. First, when the radiation components
123: are independent, the intensity of the combined radiation is the sum of the 
124: intensities of the individual radiation components (Chandrasekhar 1960; 
125: Ishimaru 1978). This realization simplifies the modeling of the radiation 
126: and its polarization. Second, the independence of the modes requires the 
127: difference in mode phases to be greater than unity. The mode phase difference, 
128: $\Delta\chi=\Delta kL$, is the product of the difference between the mode 
129: wave numbers, $\Delta k$, in the plasma and the distance, $L$, the modes 
130: propagate through the plasma. Generalized Faraday rotation (GFR) has been
131: defined as the physical process responsible for creating the difference in 
132: mode phases (Melrose 1979). If this definition is correct, then GFR must be 
133: operative in the pulsar magnetosphere. Stochastic GFR can also alter the 
134: orientation of the polarization vector and, therefore, is a tantalizing 
135: prospect for the origin of some polarization patterns we observe (Edwards 
136: \& Stappers 2004).
137: 
138: Melrose et al. (2006) made great strides in simulating the polarization 
139: patterns with their numerical model and called for further use of 
140: empirical models to constrain the mechanism that leads to the separation of 
141: the modes. The objective of this paper is to incorporate the perpendicular 
142: fluctuations in an analytical model of pulsar polarization in an attempt 
143: to replicate the observed polarization patterns and identify potential 
144: processes responsible for the fluctuations. The perpendicular fluctuations 
145: are incorporated in the model in \S\ref{sec:patterns}. The polarization 
146: patterns produced by different fluctuation geometries are also determined. 
147: The fluctuations can depolarize the emission, and, in \S\ref{sec:depol}, the 
148: depolarization is quantified with a polarization stability factor for most 
149: of the fluctuation geometries. In \S\ref{sec:perp}, the perpendicular 
150: fluctuations are interpreted in the context of stochastic GFR or scattering 
151: in the pulsar magnetosphere. The results and implications of the analysis 
152: are discussed in \S\ref{sec:discuss}. Conclusions are summarized in 
153: \S\ref{sec:conclude}. The Appendix lists the joint probability densities 
154: of the polarization vector's orientation angles for each of the fluctuation 
155: geometries. 
156: 
157: \section{ANALYTICAL MODEL OF POLARIZATION PATTERNS}
158: \label{sec:patterns}
159: 
160: The polarization patterns are two-dimensional representations of the joint 
161: probability density of the polarization vector's longitude, $\phi$, and 
162: colatitude, $\theta$. The functional form of the joint probability density 
163: can be estimated from a statistical model of pulsar polarization. An 
164: analytical, empirical model of pulsar polarization has been summarized in 
165: McKinnon (2003). Here, the model is generalized to accommodate polarization 
166: fluctuations perpendicular to the mode diagonal.
167: 
168: The Stokes parameters Q, U, and V completely describe the polarization of 
169: the radiation. The measured values of Q, U, and V are the linear sums of 
170: the pulsar-intrinsic polarization fluctuations and the instrumental noise. 
171: An analytical description of many types of polarization patterns can be 
172: determined by assuming that all fluctuations are independent, normal random 
173: variables (RVs). Since the sum of independent, normal RVs is also a normal 
174: RV, this assumption allows us to interpret the polarization as being 
175: composed of fixed and fluctuating parts, where the fluctuations follow a 
176: normal distribution. The fixed part is the mean value, $\mu$,  of the 
177: polarization vector amplitude. The analysis can be simplied by defining a 
178: new Cartesian coordinate system, q, u, and v, within the Poincar\'e sphere 
179: where the new v-axis is aligned with the mean orientation of the polarization 
180: vector. The simple equations that specify the model are then
181: %
182: \begin{equation}
183: {\rm q} = x_{\rm q},
184: \label{eqn:q}
185: \end{equation}
186: %
187: \begin{equation}
188: {\rm u} = x_{\rm u},
189: \label{eqn:u}
190: \end{equation}
191: %
192: \begin{equation}
193: {\rm v} = \mu + x_{\rm v}.
194: \label{eqn:v}
195: \end{equation}
196: %
197: The normal RVs $x$ account for the polarization fluctuations in each of q, u, 
198: and v. They each have a zero mean, and their standard deviations are denoted 
199: by $\sigma_q, \sigma_u$, and $\sigma_v$ in what follows. The q, u, v coordinate
200: system forms the eigenbasis of the polarization ellipsoid, and $\sigma_q, 
201: \sigma_u$, and $\sigma_v$ are the square roots of its eigenvalues (McKinnon 
202: 2004; Edwards \& Stappers 2004). Equations~\ref{eqn:q}-\ref{eqn:v} give the 
203: Cartesian coordinates that define the instantaneous orientation and amplitude 
204: of the polarization vector. The joint probability density of the vector's 
205: orientation angles can be computed by a transformation from Cartesian to 
206: spherical coordinates. 
207: 
208: Five cases with different fluctuation geometries, from simple to complex, are 
209: explored in the subsections that follow. Four cases investigate possible 
210: geometries where at least two of the eigenvalues are equal. The fifth case is 
211: more general, and evaluates a scenario where all the eigenvalues are different.
212: Granted, other cases could be considered, but they generally produce patterns 
213: that are simple rotations of the patterns produced here. The cases show that 
214: the polarization pattern is determined by the relative magnitudes
215: of $\sigma_q, \sigma_u$, and $\sigma_v$ and the value of $\mu$. 
216: 
217: Despite the simplicity of Equations~\ref{eqn:q}-\ref{eqn:v}, the joint density 
218: produced from them has a complicated mathematical form. Interestingly, but 
219: quite understandably, the mathematical form of the joint density is basically 
220: the same in all cases, differing only in a parameterization determined by the 
221: geometry of the polarization fluctuations. The form of the joint density and 
222: its geometrical parameterization for all cases considered is given in the 
223: Appendix. The functional form of the joint density can be captured in a much 
224: simpler conditional density, which is the probability density of the orientation 
225: angles at a fixed value of the polarization amplitude, $r_o$. The detailed 
226: procedure for calculating both the joint and conditional densities is 
227: described in McKinnon (2003, 2006) and is not reproduced here. The conditional 
228: probability densities of the orientation angles are derived for each case 
229: below. 
230: 
231: Table 1 summarizes the results obtained for each case by listing the 
232: resulting shape of the q-u-v ellipsoid and the formal name of the orientation 
233: angles' conditional density. The axis orientation in the table describes the 
234: orientation of the ellipsoid's axis of symmetry with respect to the v axis. 
235: 
236: Figures~\ref{fig:pattern1} and~\ref{fig:pattern2} show LEAP examples for each 
237: case. A LEAP is simply a polar plot where a data point's azimuth is $\phi$ 
238: and its radius is $2\sin(\theta/2)$. The most attractive feature of a LEAP 
239: is it preserves the density of data points on the sphere when projecting 
240: them in two dimensions (Fisher et al. 1987). In the figures, the left side 
241: of the LEAP is the projection of the top hemisphere of the q-u-v sphere as 
242: viewed down the v axis, where $\theta=0$. The right side of the LEAP is the 
243: projection of the bottom hemisphere at $\theta=\pi$. The circular edge of 
244: the projections is the equator of the q-u-v sphere, where $\theta=\pi/2$. 
245: 
246: \begin{figure}
247: \plotone{f1.eps}
248: \caption{Examples of possible polarization patterns in pulsar radio emission.
249: The patterns are shown as contour plots of Lambert equal area projections of 
250: the Poincar\'e sphere. {\it Top panel:} Fisher distribution with $\kappa = 4$ 
251: (Case 1). {\it Middle panel:} Bingham-Mardia bipolar distribution with 
252: $\kappa = 3, \gamma = 0.015$ created by fluctuations parallel to the 
253: polarization vector (Case 2). {\it Bottom panel:} Bingham-Mardia girdle 
254: distribution with $\kappa=4,\gamma=0.6$ created by fluctuations perpendicular 
255: to the polarization vector (Case 3). Contour levels are at 0.2, 0.4, 0.6,
256: and 0.8 of the peak value in each projection.}
257: %
258: \label{fig:pattern1}
259: \end{figure}
260: 
261: \subsection{Case 1: $\sigma_q = \sigma_u = \sigma_v$}
262: 
263: The first case to consider is the simplest of the five, and evaluates the 
264: conditional density when the standard deviations of the polarization 
265: fluctuations are identical (i.e. $\sigma_q = \sigma_u = \sigma_v = \sigma$). 
266: The shape of the q-u-v cluster in this case is a spheroid because the standard 
267: deviations are equal. Since the cluster is a spheroid, it does not have a
268: unique symmetry axis. This case was evaluated by McKinnon (2003), and 
269: represents the statistics of a fixed polarization vector accompanied by 
270: instrumental noise. The conditional density of the polarization vector's 
271: orientation angles is a Fisher distribution
272: %
273: \begin{equation}
274: f_1(\theta,\phi |r_o) = {\sin\theta\over{4\pi}}
275:                       {\kappa^2\exp(\kappa^2\cos\theta)\over{\sinh(\kappa^2)}},
276: \label{eqn:cond1}
277: \end{equation}
278: %
279: where 
280: %
281: \begin{equation}
282: \kappa^2={\mu r_o\over{\sigma^2}}.
283: \label{eqn:kappa1}
284: \end{equation} 
285: %
286: The parameter $\kappa$\footnote{The equivalent of Equation~\ref{eqn:kappa1} 
287: in McKinnon (2003) is used to define $\kappa$, instead of $\kappa^2$. The 
288: definition used here is preferred because it implies a connection with 
289: variance.} is inversely related to the conditional density's standard 
290: deviation. Since $\sin\theta$ is proportional to the derivative of 
291: $\cos\theta$, the Fisher distribution is an exponential in $\cos\theta$. 
292: The density is not a function of $\phi$ because the distribution of data 
293: points is azimuthally symmetric about the mean orientation of the polarization 
294: vector (i.e. $\phi$ is uniformly distributed over $2\pi$ and is independent 
295: of $\theta$). Therefore, the contour shapes in a LEAP for this case are 
296: always circles. 
297: 
298: The conditional density in Equation~\ref{eqn:cond1} is a probability density 
299: function (PDF), but the equation to plot in a LEAP is the probability density 
300: element (PDE). For spherical data, the PDE is the PDF without the leading
301: $\sin\theta$ term (Fisher et al. 1987; see also Edwards \& Stappers 2004). 
302: When $\kappa>1$, the polarization pattern derived from the PDE is a single 
303: set of concentric, circular contours with a peak at the center of the left 
304: LEAP hemisphere, where $\theta = 0$ (top panel of Figure~\ref{fig:pattern1}). 
305: When $\kappa<1$, the conditional density becomes isotropic, and the LEAP 
306: will show data points uniformly scattered over both hemispheres. The joint 
307: probability density of the orientation angles (Equation~\ref{eqn:joint}) is 
308: parameterized solely by the polarization signal-to-noise ratio, $s=\mu/\sigma$. 
309: The functional form of the joint density is similar to that of the conditional 
310: density. 
311: 
312: An observational example of this case can be found towards the center of the 
313: pulse in PSR B2020+28, as shown in the top panel of Figure~\ref{fig:PSRpattern}.
314: The polarization pattern is a set of circularly-shaped contours in a single 
315: hemisphere of the LEAP. The cluster of polarization data points at this pulse 
316: location is clearly a spheroid because its three dimensions are nearly identical 
317: (see Figures 3 and 5 of McKinnon 2004). 
318: 
319: \subsection{Case 2: $\sigma_q = \sigma_u < \sigma_v$}
320: 
321: The second case investigates fluctuations in q and u that are equal, 
322: $\sigma_q = \sigma_u =\sigma$, but less than those in v by a factor of 
323: $(1+\rho^2)^{1/2}$, where $\rho\ge 0$. This case was evaluated in McKinnon 
324: (2006), and considers a system dominated by fluctuations along the 
325: polarization vector, as is caused by OPMs. The shape of the q-u-v data point 
326: cluster created by these polarization fluctuations is a prolate ellipsoid. 
327: Its major axis is parallel to v. The conditional density of the vector's 
328: orientation angles is a Bingham-Mardia (BM) bipolar distribution
329: %
330: \begin{equation}
331: f_2(\theta,\phi |r_o) = {\sin\theta\over{2\pi}}
332:               {\exp[\kappa^2(\cos\theta + \gamma)^2]\over{w_2(\kappa,\gamma)}}.
333: \label{eqn:cond2}
334: \end{equation}
335: %
336: The distribution is parameterized by the constants $\kappa$ and $\gamma$.
337: %
338: \begin{equation} 
339: \kappa^2 = {r_o^2\rho^2\over{2\sigma^2(1+\rho^2)}} 
340: \label{eqn:kappa2}
341: \end{equation} 
342: %
343: \begin{equation}
344: \gamma = {\mu\over{r_o\rho^2}}
345: \label{eqn:gamma2}
346: \end{equation}
347: %
348: The constant $w_2(\kappa,\gamma)$ normalizes the density and is found by 
349: integrating the numerator of Equation~\ref{eqn:cond2}. The conditional 
350: density becomes the Watson bipolar distribution when $\mu=0$ (McKinnon 2006). 
351: 
352: The conditional density can be unimodal or bimodal depending upon the value 
353: of $\gamma$. When $\gamma > 0$, the PDE always peaks at the center of the 
354: left LEAP hemisphere where $\theta=0$. When $|\gamma|\ll 1$ the polarization 
355: pattern is bimodal (i.e. a secondary peak appears at the center of the right 
356: LEAP hemisphere where $\theta=\pi$). The patterns are unimodal for larger 
357: values of $\gamma$. As with Case 1, the shapes of the density's contours are 
358: always circles because of the azimuthal symmetry of the problem. The middle 
359: panel of Figure~\ref{fig:pattern1} shows an example of the PDE from 
360: Equation~\ref{eqn:cond2}. 
361: 
362: As shown in McKinnon (2006), the functional form of the joint probability 
363: density (Equation~\ref{eqn:joint}) is very similar to that of the conditional 
364: density. The polarization modulation index, $\beta$, determines whether the 
365: joint density is bimodal or unimodal. In this case, the modulation index is 
366: defined by
367: %
368: \begin{equation}
369: \beta = {(\sigma_v^2-\sigma^2)^{1/2}\over{\mu}} = {\rho\over{s}},
370: \end{equation}
371: %
372: where $s=\mu/\sigma$ is the ratio of the polarization amplitude to the 
373: effective noise of the system. The fluctuations along v that are over 
374: and above the effective noise are represented by $\rho\sigma$. When 
375: $\beta > 1$, the fluctuations along v exceed the mean polarization, and 
376: the joint density is bimodal with data points residing in both LEAP 
377: hemispheres.  When $\beta < 1$, the mean exceeds the fluctuations, and the 
378: joint density is unimodal.
379: 
380: An observational example of this case occurs on the trailing edge of the 
381: pulse of PSR B1929+10 (middle panel of Figure~\ref{fig:PSRpattern}). There,
382: the polarization pattern consists of a set of circularly-shaped contours in 
383: each hemisphere of the LEAP. The contours in the right hemisphere are not 
384: centered in the projection because the modes are not precisely orthogonal 
385: (McKinnon 2004). The cluster of polarization data points at this location 
386: in the pulse has the form of a prolate ellipsoid because two of its dimensions 
387: are equal but smaller than the third (Figures 4 and 5 of McKinnon 2004). Another 
388: example of this case occurs in the cone emission of PSR B0329+54. Again, the 
389: pattern consists of a set of circularly-shaped contours in each hemisphere of
390: the LEAP (Figure 2 of Edwards \& Stappers 2004), and the relative dimensions 
391: of the data point clusters are consistent with those expected for a prolate 
392: ellipsoid (Figure 3 of Edwards \& Stappers).
393: 
394: \begin{figure}
395: \plotone{f2.eps}
396: \caption{Additional examples of possible polarization patterns in pulsar radio 
397: emission. {\it Top panel:} A hybrid of the Bingham-Mardia distribution where 
398: the polarization fluctuations are primarily along q (Case 4). This particular 
399: pattern was formed with $\kappa = 2$ and $\gamma=0.65$. {\it Middle panel:} 
400: Same as the top panel but with $\gamma = 1$. {\it Bottom panel:} Another 
401: hybrid of the Bingham-Mardia distribution where the fluctuations in all three 
402: dimensions are different, with $\kappa_o = -4, \kappa = 3,$ and 
403: $\gamma=-0.04$ (Case 5). Contour levels are at 0.2, 0.4, 0.6, and 0.8 of the 
404: peak value in each projection.}
405: \label{fig:pattern2}
406: \end{figure}
407: 
408: \subsection{Case 3: $\sigma_q = \sigma_u > \sigma_v$}
409: 
410: To model fluctuations perpendicular to the polarization vector, the 
411: fluctuations in q and u are taken to be equal, 
412: $\sigma_q = \sigma_u =\sigma(1+\eta^2)^{1/2}$, but greater than those in 
413: v, where $\sigma_v =\sigma$. The constant $\eta$ denotes the magnitude 
414: of the fluctuations perpendicular to the polarization vector, while $\rho$ 
415: is reserved to signify the parallel fluctuations. The polarization fluctuations 
416: in this case create a q-u-v data point cluster in the shape of an oblate 
417: ellipsoid. The minor axis of the ellipsoid is parallel to v. This case does 
418: not preclude the possibility of OPMs; it simply stipulates that fluctuations
419: perpendicular to the polarization vector exceed the parallel fluctuations. 
420: The conditional density is a BM girdle distribution
421: %
422: \begin{equation}
423: f_3(\theta,\phi |r_o) = {\sin\theta\over{2\pi}}
424:          {\exp[-\kappa^2(\cos\theta - \gamma)^2]\over{w_3(\kappa,\gamma)}},
425: \label{eqn:cond3}
426: \end{equation}
427: %
428: where
429: %
430: \begin{equation}
431: \gamma = {\mu(1+\eta^2)\over{r_o\eta^2}}
432: \label{eqn:gamma3}
433: \end{equation}
434: %
435: and $\kappa$ has the same definition as in Case 2, Equation~\ref{eqn:kappa2}, 
436: but with $\rho$ replaced by $\eta$. The constant $w_3(\kappa,\gamma)$ is
437: a normalization factor. The conditional density is normal in $\cos\theta$ with 
438: mean $\gamma$ and a standard deviation proportional to $1/\kappa$. Since the 
439: density is a distribution of $\cos\theta$, which must lie in the range 
440: $0\le|\cos\theta|\le 1$, a strict mathematical interpretation of 
441: Equation~\ref{eqn:cond3} requires $\gamma$ to lie in the range 
442: $0\le|\gamma|\le 1$. A variant of Equation~\ref{eqn:cond3} has been used to 
443: model the motion of volcanic hotspots on the surface of the Earth (Bingham 
444: \& Mardia 1978; Mardia \& Gadsden 1977). 
445: 
446: As shown in the bottom panel of Figure~\ref{fig:pattern1}, the polarization 
447: pattern formed from the conditional density PDE is an annulus in a single 
448: LEAP hemisphere. The PDE peaks at $\cos\theta=\gamma$ for all longitudes, 
449: again because of the azimuthal symmetry of the problem. The outer edge of the 
450: annulus has a steep slope and its inner edge has a more gradual slope. The 
451: sign of $\gamma$ determines the hemisphere where the annulus resides. When 
452: $\gamma=0$, the PDE peaks at $\theta=\pi/2$, and the pattern contours are 
453: concentrated around the equator in each LEAP hemisphere. As $\gamma$ approaches 
454: unity, the annulus collapses into a single cluster with a peak at $\theta=0$. 
455: Unlike the density in Case 2 that can be bimodal, the density in this case 
456: is always unimodal. 
457: 
458: The shape of the polarization pattern derived from the joint probability 
459: density (Equation~\ref{eqn:joint}) is again determined by the polarization 
460: modulation index, here defined as $\beta=\eta/s$. The polarization pattern 
461: is an annulus when the polarization fluctuations are comparable to the mean 
462: polarization ($\beta\simeq 1$). The pattern is a single peak at $\theta=0$
463: when the polarization fluctuations are less than the mean ($\beta<1$).
464: 
465: The polarization annulus was first recognized by Melrose et al. (2006) in 
466: their attempt to model the polarization pattern observed by Edwards \& Stappers
467: (2004) in the core emission of PSR B0329+54. Their numerical model showed 
468: that the annulus formed when the polarization amplitude was weak. The 
469: analytical model developed here reproduces a similar pattern precisely when 
470: $\mu$ (i.e. $\gamma$) is small or when $\beta\simeq 1$.
471: 
472: \subsection{Case 4: $\sigma_q > \sigma_u = \sigma_v$}
473: 
474: The fluctuation geometries described in the preceding cases have all been 
475: symmetric in azimuth. The symmetry is broken in this case by setting the 
476: fluctuations in u and v equal to one another, $\sigma_v = \sigma_u =\sigma$, 
477: but less than those in q, $\sigma_q = \sigma(1+\eta^2)^{1/2}$. Now the shape 
478: of the q-u-v data point cluster is a prolate ellipsoid with its major axis 
479: oriented along q and thus perpendicular to v. The conditional density is a 
480: hybrid of the BM distribution
481: %
482: \begin{equation}
483: f_4(\theta,\phi |r_o) = {\sin\theta\over{2\pi}}
484:            {\exp[-\kappa^2(\cos\theta  - \gamma)^2]
485:            \exp(-\kappa^2\sin^2\phi\sin^2\theta)\over{w_4(\kappa,\gamma)}},
486: \label{eqn:cond4}
487: \end{equation}
488: %
489: where $\kappa$ and $\gamma$ have the same definitions as in Case 3, and
490: $w_4(\kappa,\gamma)$ is a constant that normalizes the distribution. 
491: Equation~\ref{eqn:cond4} is the conditional density of Case 3 (first 
492: exponential term in the equation) that is further shaped by a 
493: longitude-dependent term (the second exponential term). The PDE resides in 
494: one hemisphere of the LEAP and generally peaks at two locations where 
495: $\cos\theta=\gamma$ and $\phi = 0,\pi$. As shown in the top and middle 
496: panels of Figure~\ref{fig:pattern2}, the pattern has a bar shape when 
497: $\gamma=1$, and a bow tie shape when $\gamma<1$. For even smaller values 
498: of $\gamma$, the bow tie separates into two distinct peaks.
499: 
500: Again, the polarization modulation index, $\beta=\eta/s$, determines the 
501: pattern shape derived from the joint probability density. When $\beta>1$,
502: the pattern resembles a bow tie. The pattern is a bar when $\beta<1$.
503: 
504: An observational example of this case appears near the peak of PSR B0823+26
505: (bottom panel of Figure~\ref{fig:PSRpattern}). The polarization pattern 
506: consists of a set of highly elongated contours in a single hemisphere of the
507: LEAP. Another observational example of this case may occur in the precursor 
508: to the core component of PSR B0329+54, as shown in Figure 2 of Edwards \& 
509: Stappers (2004). There, an ellipitical bar appears superimposed upon a noise 
510: background in the left LEAP hemisphere. Only the noise background of uniformly 
511: distributed data points appears in the right hemisphere. As shown in their 
512: Figure 3, two of the dimensions of the Q-U-V cluster are equal but smaller 
513: than the third; therefore, the shape of the cluster is a prolate ellipsoid. 
514: For the polarization pattern to be confined primarily to one LEAP hemisphere, 
515: the ellipsoid's major axis must be perpendicular to the mode diagonal, as 
516: required in this case.
517: 
518: \subsection{Case 5: $\sigma_u < \sigma_q = \sigma_v$}
519: 
520: The final and most complex, yet general, geometry to consider is when 
521: the fluctuations in q and v are greater than those in u, but not 
522: necessarily equal to one another. Here, the fluctuations in u are defined
523: as $\sigma_u =\sigma$, and the fluctuations in q and v are
524: $\sigma_q = \sigma(1+\eta^2)^{1/2}$ and $\sigma_v = \sigma(1+\rho^2)^{1/2}$,
525: respectively. Generally, the shape of the q-u-v ellipsoid in this case is 
526: an irregular ellipsoid that does not have an axis of symmetry. When 
527: $\eta=\rho$, the ellipsoid is oblate with its minor axis perpendicular to v,
528: and the conditional density is
529: %
530: \begin{equation}
531: f_5(\theta,\phi |r_o) = {\sin\theta\over{2\pi}}
532:            {\exp[\kappa^2(\cos\theta + \gamma)^2]
533:            \exp(\kappa^2\sin^2\phi\sin^2\theta)
534:            \over{w_5(\kappa,\gamma)}}
535: \end{equation}
536: %
537: where $\kappa$ and $\gamma$ are now defined by Equations~\ref{eqn:kappa2} 
538: and~\ref{eqn:gamma2}, respectively, under Case 2, and $w_5(\kappa,\gamma)$ 
539: is a normalization factor. 
540: 
541: When $\eta\neq\rho$, the conditional density is 
542: %
543: \begin{equation}
544: f_5(\theta,\phi |r_o) = {\sin\theta\over{2\pi}}
545:            {\exp[-\kappa_o(\cos\theta - \gamma)^2]
546:            \exp(-\kappa^2\sin^2\phi\sin^2\theta)
547:            \over{w_5(\kappa,\kappa_o,\gamma)}}
548: \label{eqn:cond5}
549: \end{equation}
550: %
551: where $\kappa$ has the same definition as in Cases 3 and 4, and $\kappa_o$ 
552: and $\gamma$ are given by
553: 
554: \begin{equation}
555: \kappa_o = {r_o^2(\eta^2-\rho^2)\over{2\sigma^2(1+\rho^2)(1+\eta^2)}}
556: \end{equation}
557: 
558: \begin{equation}
559: \gamma = {\mu(1+\eta^2)\over{r_o(\eta^2-\rho^2)}}
560: \label{eqn:gamma5}
561: \end{equation}
562: %
563: Notice that $\kappa_o$ and $\gamma$ can be negative depending upon the 
564: values of $\eta$ and $\rho$. Equation~\ref{eqn:cond5} is very general in that 
565: it reproduces the conditional density for Case 4 (Equation~\ref{eqn:cond4}) 
566: when $\rho=0$. Similarly, it becomes the conditional density for Case 2 
567: (Equation~\ref{eqn:cond2}) when $\eta=0$. When $\rho\ge\eta$ such that 
568: $|\gamma|<1$, the polarization pattern derived from Equation~\ref{eqn:cond5} 
569: is bimodal and very similar to that of Case 2, with the exception that the 
570: PDE contours now have elliptical, instead of circular, cross sections because 
571: of the asymmetry introduced by the fluctuations in q (see the bottom panel 
572: of Figure~\ref{fig:pattern2}). When $\eta > \rho$, the polarization pattern 
573: can be a bar or bow tie in a single hemisphere, as in Case 4, depending upon 
574: the value of $\gamma$.
575: 
576: An observational example of this case can be found in PSR B0818--13, where
577: the LEAP at the pulse peak is a diffuse structure that resembles a bow tie 
578: (Figure 5 of Edwards 2004). The cluster dimensions are not equal to one 
579: another, but two of the dimensions are noticeably larger than the third 
580: (Edwards, Figure 3), suggesting that the cluster shape is an irregular 
581: ellipsoid, as this case requires. To get the bow tie shape in its LEAP, the 
582: fluctuations along q would have to be slightly larger than those along v.
583: 
584: \begin{figure}
585: \plotone{f3.eps}
586: \caption{Lambert equal area projections of the observed orientation angles 
587: of the polarization vector in three pulsars. {\it Top panel:} Polarization 
588: pattern measured towards the center of PSR B2020+28. Contour levels are -12, 
589: -9, -6, and -3 decibels (dB) referenced to 0 dB at the peak of the projection. 
590: {\it Middle panel:} Pattern measured on the trailing edge of PSR B1929+10.
591: Contours are -20, -15, -10, and -5 dB. {\it Bottom panel:} Pattern measured 
592: near the peak of PSR B0823+26. Contours are -8, -6, -4, and -2 dB. The data 
593: were recorded by Stinebring et al. (1984) with the Arecibo radio telescope at
594: 1404 MHz.}
595: \label{fig:PSRpattern}
596: \end{figure}
597: 
598: \section{DEPOLARIZATION BY THE FLUCTUATIONS}
599: \label{sec:depol}
600: 
601: The polarization patterns observed in pulsars are indicators of how the 
602: orientation angles of a polarization vector fluctuate on the Poincar\'e 
603: sphere. The fluctuations are polarization instabilities that can depolarize 
604: the emission. The conditional densities derived in \S\ref{sec:patterns} 
605: replicate many of the observed polarization patterns, and can be used to 
606: quantify the degree of depolarization caused by the fluctuations.
607:  
608: Manchester et al. (1973) first observed that the percentage linear 
609: polarization of pulsar radio emission decreased with increasing radio 
610: frequency, and Manchester et al. (1975) suggested that the depolarization may 
611: arise from an increase in the randomization of polarization position angle. 
612: Manchester et al. (1975) and Cordes \& Hankins (1977) quantified the stability 
613: of the linear polarization with a two-dimensional polarization stability 
614: factor, here defined as
615: %
616: \begin{equation}
617: \sigma_p = \Biggl({\langle Q\rangle^2 +\langle U\rangle^2\over
618:       {\langle Q^2\rangle + \langle U^2\rangle}}\Biggr)^{1/2}.
619: \end{equation}
620: %
621: Assuming that the position angle, $\psi$, is a normal RV with a zero mean 
622: and a standard deviation, $\sigma_\psi$, Cordes \& Hankins (1977) calculated 
623: the moments of Q and U to derive a polarization stability factor given by
624: %
625: \begin{equation}
626: \sigma_p=\exp(-\sigma_\psi^2/2).
627: \end{equation}
628: %
629: An identical expression is used to describe the depolarization caused by
630: stochastic Faraday rotation in the interstellar medium (e.g. Spangler 1982;
631: Melrose \& Macquart 1998). 
632: 
633: Using the same methodology described in \S\ref{sec:patterns}, McKinnon (2003) 
634: showed that the conditional density of the position angle follows a von Mises 
635: distribution
636: %
637: \begin{equation}
638: f(\psi|r_o)={\exp(\kappa^2\cos2\psi)\over
639:                    {\pi\rm{I_0}(\kappa^2)}},
640: \end{equation}
641: %
642: where $\rm{I_0}(x)$ is the modified Bessel function of order zero and 
643: $\kappa$ is inversely related to the position angle dispersion. The 
644: polarization stability factor derived from the von Mises distribution is 
645: %
646: \begin{equation}
647: \sigma_p = {\rm{I_1}(\kappa^2)\over{\rm{I_0}(\kappa^2)}},
648: \end{equation}
649: %
650: where $\rm{I_1}(x)$ is the modified Bessel function of first order. When 
651: the fluctuations in $\psi$ are small ($\kappa\gg 1$), the von Mises 
652: distribution is almost indistinguishable from a normal distribution 
653: with a standard deviation of $\sigma_\psi = (2\kappa)^{-1}$. Of the two 
654: distributions, the von Mises distribution is the more accurate representation 
655: of position angle fluctuations because $\psi$ is distributed on a semi-circle, 
656: not a line, and lies in the range $0\le\psi<\pi$, instead of 
657: $-\infty\le \psi < \infty$ (Fisher et al. 1987). The stability factors 
658: derived from the two distributions are compared in the top left panel of 
659: Figure~\ref{fig:psf}.
660: 
661: The observations of Edwards (2004) and Edwards \& Stappers (2004), however,
662: have clearly shown that the fluctuations occur in all three Stokes parameters,
663: and not just in Q and U. Consequently, polarization fluctuations in the 
664: three dimensional case cause the vector colatitude to vary in addition to 
665: its longitude (position angle). The depolarization can be quantified by 
666: extending the definition of the polarization stability factor to 
667: %
668: \begin{equation}
669: \sigma_p=\Biggl({\langle Q\rangle^2 + \langle U\rangle^2 + \langle V\rangle^2
670:         \over{\langle Q^2\rangle + \langle U^2\rangle + \langle V^2\rangle}}
671:         \Biggr)^{1/2}.
672: \end{equation}
673: %
674: For any joint probability density of $\theta$ and $\phi$ that is properly
675: normalized, the sum of the second moments of the Stokes parameters is 
676: constant because of the definitions of the Stokes parameters and 
677: trigonometric identities. The first moments of Q and U are equal to zero 
678: for all the distributions derived in \S\ref{sec:patterns}. Therefore, the
679: only term that contributes to the stability factor is the first moment of
680: V.
681: 
682: The polarization stability factor derived from the Fisher distribution 
683: (Case 1) is
684: %
685: \begin{equation}
686: \sigma_p = \coth(\kappa^2)-{1\over{\kappa^2}},
687: \end{equation}
688: %
689: and is shown in the top right panel of Figure~\ref{fig:psf}. Very little
690: depolarization occurs when the fluctuations are small ($\kappa\gg 1$),
691: but the depolarization can be significant when the fluctuations are large
692: ($\kappa\simeq 1$).
693: 
694: The polarization stability factor derived from the BM bipolar distribution 
695: (Case 2) is 
696: %
697: \begin{equation}
698: \sigma_p = {\exp[\kappa^2(1+\gamma^2)]\sinh(2\kappa^2\gamma)
699:            \over{\kappa^2 w_2(\kappa,\gamma)}} - \gamma,
700: \end{equation}
701: %
702: where the constant $w_2(\kappa,\gamma)$ is given by
703: %
704: \begin{equation}
705: w_2(\kappa,\gamma)=\int_{\gamma-1}^{\gamma+1}\exp(\kappa^2x^2)dx.
706: \end{equation}
707: %
708: The stability factor for Case 2 is plotted against $\kappa$ for different 
709: values of $\gamma$ in the bottom left panel of Figure~\ref{fig:psf}. The 
710: fluctuations can depolarize the radiation, but the figure shows that the 
711: depolarization is more pronounced for small values of $\gamma$.
712: 
713: The polarization stability factor derived from the BM girdle distribution 
714: (Case 3) is 
715: %
716: \begin{equation}
717: \sigma_p = \gamma - {\exp[-\kappa^2(1+\gamma^2)]\sinh(2\kappa^2\gamma)
718:            \over{\kappa^2 w_3(\kappa,\gamma)}}.
719: \end{equation}
720: %
721: The constant $w_3(\kappa,\gamma)$ is
722: %
723: \begin{equation}
724: w_3(\kappa,\gamma) = {\sqrt{\pi}\over{2\kappa}}\{{\rm erf}[\kappa(1-\gamma)]
725:                    + {\rm erf}[\kappa(1+\gamma)]\}.
726: \end{equation}
727: %
728: The stability factor for Case 3 is shown in the bottom right panel of 
729: Figure~\ref{fig:psf}. As with Case 2, the depolarization is more severe
730: for small values of $\gamma$. The stability factor approaches 
731: $\sigma_p = \gamma$ when $\kappa\gg 1$.
732: 
733: The effect of $\kappa$ and $\gamma$ on the frequency-dependent polarization 
734: can be summarized as follows. The term $\kappa$ represents the magnitude of 
735: the fluctuations and sets the overall size of the polarization pattern; 
736: larger patterns are formed from smaller values of $\kappa$. The frequency 
737: dependence of $\kappa$ may be set by the process responsible for the 
738: polarization fluctuations. The term $\gamma$ is proportional to $\mu$, and 
739: thus is set by the OPMs or a process intrinsic to the emission mechanism. 
740: Its frequency dependence may be determined by the spectral index of the 
741: individual modes (e.g. Karastergiou et al. 2005; Smits et al. 2006; Johnston 
742: et al. 2008) or that of the emission's intrinsic polarization. As can be seen 
743: from Figure~\ref{fig:psf}, the depolarization caused by pure fluctuations alone
744: (the Fisher distribution from Case 1, top right panel) forms a rough, upper 
745: envelope to the depolarization caused by the other fluctuation geometries. 
746: The depolarization becomes more substantial in Cases 2 and 3 (bottom panels 
747: of the Figure) as the value of $\gamma$ decreases. In other words, the 
748: depolarization is influenced more by the intrinsic effects represented by 
749: $\gamma$ than by the polarization fluctuations represented by $\kappa$. 
750: Therefore, the main conclusion to draw from this polarization stability 
751: analysis is pulsar-intrinsic effects are more effective at depolarizing 
752: the emission than random fluctuations in the orientation of the polarization 
753: vector.
754: 
755: \begin{figure}
756: \plotone{f4.eps}
757: \caption{The dependence of the polarization stability factor upon
758: different types of fluctuations in the position angle or colatitude of
759: the polarization vector. The top left panel shows the stability factor 
760: when the fluctuations in position angle follow the two-dimensional von 
761: Mises and normal distributions. The top right panel shows the stability 
762: factor when the colatitude of the polarization vector fluctuates 
763: according to the three-dimensional Fisher distribution (Case 1). The 
764: stability factors determined from the Bingham-Mardia bipolar distribution 
765: (Case 2) and the Bingham-Mardia girdle distribution (Case 3) are shown 
766: in the bottom left and right panels, respectively, for different values 
767: of $\gamma$. The factors are plotted versus $\kappa$, which is related 
768: inversely to the dispersion in colatitude or position angle.}
769: \label{fig:psf}
770: \end{figure}
771: 
772: \section{POSSIBLE ORIGIN OF PERPENDICULAR FLUCTUATIONS}
773: \label{sec:perp}
774: 
775: A single, physical process is not likely to be responsible for all the
776: polarization patterns we observe. As mentioned in the Introduction, the
777: origin of the simple polarization patterns is qualitatively understood, 
778: particularly in the case of OPMs where the patterns are formed by 
779: fluctuations parallel to the mode polarization vectors. The origin of 
780: the more complicated patterns created by perpendicular polarization 
781: fluctuations is not understood. But as discussed below, they may arise 
782: from a stochastic version of GFR or scattering in the pulsar magnetosphere.
783: 
784: \subsection{Stochastic Version of Generalized Faraday Rotation}
785: 
786: In general terms, Faraday rotation is the physical process that alters the 
787: difference between the phases of the modes as they propagate through a plasma 
788: (Melrose 1979), regardless of the particle energy in the plasma, the strength 
789: of the magnetic field threading the plasma, or the coherence of the modes. 
790: The modes are incoherent when the difference in their phases at a given 
791: wavelength is large ($\Delta\chi\gg 1$) and are coherent (coupled) as long 
792: as the phase difference is small ($\Delta\chi < 1$). The modes retain their 
793: individual polarization identity in an observation when they are incoherent, 
794: but effectively lose their identity when they are coherent. Faraday rotation 
795: can become stochastic when the fluctuations in phase difference are large 
796: ($\sigma_\chi \gg 1$), in which case $\Delta\chi$ can be treated as a random 
797: variable (Lee \& Jokipii 1975; Simonetti et al. 1984; Melrose \& Macquart 1998). 
798: 
799: GFR alters the component of the radiation's polarization vector that is 
800: perpendicular to the polarization vectors of the plasma's wave propagation 
801: modes. For any plasma, the unit vectors representing the polarization states 
802: of the two modes are anti-parallel on a diagonal through the Poincar\'e 
803: sphere. For the cold, weakly-magnetized plasma that is the interstellar 
804: medium (ISM), the propagation modes are circularly polarized, and the mode 
805: diagonal defined by their polarization vectors connects the poles of the 
806: Poincar\'e sphere. Faraday rotation in the ISM causes the orientation of 
807: the radiation's polarization to vary in a plane perpendicular to the mode 
808: diagonal, either on the Poincar\'e sphere's equator or on a small circle 
809: parallel to it, depending upon the polarization state of the plasma-incident 
810: radiation. For the relativistic plasma in the strong magnetic field of a 
811: pulsar's magnetosphere, the modes are thought to be linearly polarized 
812: (Melrose 1979; Allen \& Melrose 1982; Barnard \& Arons 1986) so that the 
813: mode diagonal lies in the equator of the Poincar\'e sphere. The modes in 
814: a relativistic, magnetized plasma of a synchrotron radio source or pulsar 
815: wind are, in general, elliptically polarized (Kennett \& Melrose 1998; 
816: Pacholczyk \& Swihart 1970; Sincell \& Krolik 1992). In these latter cases, 
817: GFR causes the polarization vector to rotate on a small circle in the 
818: Poincar\'e sphere that is perpendicular to and centered on the mode 
819: diagonal (e.g.  Figure 3 of Kennett \& Melrose 1998). If the modes are 
820: coherent, GFR causes the amplitude of the linear and circular polarization 
821: to vary periodically (Pacholczyk \& Swihart 1970; Cocke \& Pacholczyk 1976). 
822: 
823: For GFR to occur, the polarization of the radiation incident on the plasma 
824: must be different from the polarization of the plasma's wave propagation 
825: modes (e.g. Pacholczyk \& Swihart 1970). How this might occur in a pulsar's 
826: magnetosphere can be understood if we view the magnetosphere as being 
827: composed of layers, where the properties of the wave propagation modes are 
828: constant within a given layer, but different between layers. The mode 
829: properties, such as their indices of refraction and polarization states, 
830: are likely to vary quickly with distance, $r$, from the center of the pulsar, 
831: because both the particle density and magnetic field strength are thought to 
832: decrease as $r^{-3}$ (Goldreich \& Julian 1969). For simplicity, let us 
833: assume the radiation is generated in the lowest magnetospheric layer and its 
834: polarization state is identical to that of the wave modes in the layer to 
835: ensure efficient coupling between the generation and propagation of the 
836: radiation. The radiation's polarization is unaffected in the lowest layer, 
837: but gets slightly altered by GFR in subsequent layers, because the 
838: polarization of the radiation incident on each layer is different from the 
839: polarization of that layer's propagation modes. The radiation's polarization 
840: in each layer then contains a component that is perpendicular to the mode 
841: polarization because of GFR. Ultimately, a polarization limiting region 
842: (PLR) is reached in the upper layers of the magnetosphere where GFR is 
843: no longer effective in altering the radiation's polarization (Melrose 1979; 
844: Kennett \& Melrose 1998). The PLR occurs where, or when, $\Delta\chi = 1$. 
845: At or just beyond the PLR, the radiation must couple to the ISM wave modes 
846: for propagation through the ISM (Stinebring 1982). With this simple cartoon, 
847: we see how GFR may be capable of producing a polarization component that is 
848: perpendicular to the polarization of the wave propagation modes.
849: 
850: The relatively long wavelength ($\lambda > 10$ cm) observations of individual 
851: pulse polarization conducted to date suggest that the wave propagation 
852: modes in pulsar radio emission are incoherent (MS1; MS2). Consequently, the 
853: differences in mode phases must be large, implying GFR is operative in the 
854: pulsar magnetosphere. Furthermore, the emission's intensity and polarization 
855: are highly variable, most likely due to rapid changes, both spatially and 
856: temporally, in the flowrate and physical properties of the magnetospheric 
857: plasma. This in turn suggests that the fluctuations in the mode phase 
858: difference, $\sigma_\psi$, are also very large. We are then led to the 
859: prospect that GFR in the pulsar magnetosphere is likely to be stochastic. 
860: This interpretation is consistent with the analysis in \S\ref{sec:patterns} 
861: where the orientation angles of the polarization vectors are treated as 
862: random variables.  
863: 
864: The theory of stochastic Faraday rotation in the ISM is well-developed
865: (e.g. Spangler 1982; Simonetti et al. 1984; Melrose \& Macquart 1998).
866: The theory probes a two dimensional problem, in the Stokes parameters
867: Q and U, where the fluctuations in the position angle of the linear
868: polarization vector map directly to the fluctuations in the mode phase
869: difference. The extension of the theory to the more general, three 
870: dimensional case (Q, U, and V) of stochastic GFR has not been made and 
871: is beyond the scope of this paper. In the absence of the theory, knowing 
872: that the conditional densities derived in \S\ref{sec:patterns} are 
873: reasonable approximations to the polarization patterns we observe, and 
874: realizing that stochastic GFR may occur in the magnetosphere, we are faced 
875: with the possibility that some subset of these conditional densities, or 
876: some other joint probability density, may describe how GFR causes the 
877: orientation of a pulsar's polarization vector to fluctuate. Interestingly, 
878: the polarization stability analysis in \S\ref{sec:depol} produces the same 
879: conclusions developed by Melrose \& Macquart (1998) in their assessment of 
880: depolarization by stochastic Faraday rotation in the ISM. Specifically, 
881: when the conditional densities are used to calculate the moments of the 
882: Stokes parameters, the first moments decay and the sum of their second 
883: moments remains constant. In fact, the radiative transfer equation for 
884: the Stokes parameters (e.g.  Equation 1 of Kennett \& Melrose 1998) requires 
885: the second moment of the polarization, ${\rm p^2 = q^2 + u^2 + v^2}$, to 
886: remain constant as the radiation propagates through the plasma. All of 
887: the conditional and joint probability densities listed in 
888: \S\ref{sec:patterns} and the Appendix comply with this requirement.
889: 
890: \subsection{Scattering in the Magnetosphere}
891: 
892: Edwards \& Stappers (2004) simulated the polarization patterns in the core 
893: component of PSR B0329+54. They suggested that the modes are not precisely 
894: orthogonal, and retained the assumption of superposed modes. More importantly,
895: and in contrast to the models discussed in \S\ref{sec:patterns}, they also
896: proposed that the orientation angles for the polarization vector of one 
897: mode are highly dispersed while the angles of the other mode are not. They 
898: attributed the orientation angle fluctuations to GFR. More specifically, 
899: however, their model is consistent with any mechanism that selectively alters 
900: the orientation angles of only one mode. Melrose et al. (2006) also modeled 
901: the polarization patterns in the core component of PSR B0329+54 using 
902: assumptions similar to those of Edwards \& Stappers. They invoked the 
903: non-orthogonality of the modes and required the fluctuations in the 
904: orientation angles of one mode to be different from those in the other.
905: But unlike Edwards \& Stappers, they suggested that the modes are disjoint 
906: (i.e. they occur separately, not simultaneously) part of the time, and did 
907: not specifically advocate a physical mechanism for the cause of the orientation 
908: angle fluctuations.
909: 
910: One mechanism that may alter the orientation angles of one polarization mode 
911: and not the other is scattering in the pulsar magnetosphere (Blandford \& 
912: Scharlemann 1976; Sincell \& Krolik 1992; Lyubarskii \& Petrova 1996; 
913: Petrova 2008). Induced scattering may occur in the magnetosphere because 
914: large photon occupation numbers are implied by the high brightness 
915: temperatures observed in the radio emission. When the radiation frequency, 
916: $\omega$, is much less than the electron gyrofrequency, $\omega_B$, as might 
917: be expected near the stellar surface, the only wave propagation mode that has 
918: a non-zero scattering cross section is the ordinary mode (O-mode), which is 
919: polarized in the plane defined by the wave vector, k, and the ambient magnetic 
920: field, B (Blandford \& Scharlemann 1976; Sincell \& Krolik 1992). This occurs 
921: for two reasons. First, the extreme strength of the field ($B\simeq 10^{12}$ G) 
922: causes the electrons in the plasma to occupy their lowest Landau level, 
923: such that their motion is constrained along the field, like a bead on a 
924: wire. Second, in an overly simplistic description, only an incident O-mode 
925: wave can accelerate an electron along the field, thus causing it to radiate,
926: because it is the only mode having a component to its electric field that 
927: is parallel to the ambient magnetic field. The extra-ordinary mode (X-mode) 
928: cannot accelerate an electron along the field because its electric field is 
929: always perpendicular to the magnetic field. The scattering cross section 
930: varies as $\sin^2\theta$ (Blandford \$ Scharlemann 1976), where $\theta$ is 
931: now the angle between k and B, and no scattering will occur if k and B are 
932: parallel ($\theta=0$). The observed polarization of the scattered radiation 
933: depends upon the geometry of the magnetic field in the scattering region, 
934: as projected on the plane of the sky, and the temporally and spatially varying 
935: distribution of charged particles along the magnetic field lines.
936: 
937: Petrova (2008) has suggested that induced scattering at different altitudes
938: within the magnetosphere may lead to a depolarization of the radiation. The 
939: scattered radiation from different altitudes may have different position
940: angles, and it is the superposition of these waves with different position 
941: angles that leads to the depolarization. This scenario is qualitatively
942: consistent with what is described in \S\ref{sec:patterns}, Cases 3 and 4.
943: 
944: In summary, the empirical models appear to place three requirements on the 
945: physical processes that create the more complex polarization patterns we 
946: observe. First, the polarization must have a large modulation index, through a 
947: combination of large polarization fluctuations and a small mean polarization.
948: Second, the processes must create fluctuations in polarization that are
949: both parallel and perpendicular to the mean polarization vector. Third,
950: in some cases, the process may create fluctuations in the orientation of 
951: the polarization vector of one orthogonal mode that exceed those in the
952: other mode.
953: 
954: 
955: \section{DISCUSSION}
956: \label{sec:discuss}
957: 
958: A fundamental question remains for the observation of PSR B0329+54 by Edwards 
959: \& Stappers (2004). Why does the polarization annulus appear in the core 
960: component of the pulsar but not in its conal outriders? The annulus must be 
961: intrinsic to the pulsar because a propagation effect in the pulsar wind or 
962: the ISM presumably would affect the core and cone emission in a similar 
963: fashion. Explanations for the difference may reside in the distinction 
964: between core and cone emission made by Rankin (1983; 1990). Cone emission 
965: has a moderate spectral index and can be highly linearly polarized. It is 
966: thought to originate high above the pulsar polar cap. Core emission, on the 
967: other hand, has a steep spectral index, and its polarization signature is 
968: often a change in the handedness of the circular polarization near the core 
969: peak. It is thought to originate at or near the polar cap. The core and cone 
970: radiation from PSR B0329+54 appear to follow this general characterization 
971: of the emission components (Karastergiou et al. 2001). We can surmise that 
972: GFR occurs throughout the pulse of PSR B0329+54 because observations (Edwards 
973: \& Stappers 2004) indicate that independent OPMs occur at most locations within 
974: its pulse. Of course, the core emission would be more susceptible to GFR 
975: because its propagation path length is presumably much larger than that for 
976: the cone emission. Stochastic GFR may occur predominantly near the stellar 
977: surface where a turbulent plasma outflow causes fluctuations in the mode 
978: phase difference. Stochastic GFR may get suppressed in the propagation region 
979: for the cone emission because the cross section of the plasma flux tube 
980: increases with distance from the star and, thus, the plasma outflow becomes 
981: more laminar. Alternatively, a popular model for pulsar radio emission (e.g. 
982: Ruderman \& Sutherland 1975) calls for an intense photon beam to be created 
983: by primary charged particles that are accelerated in a voltage potential gap 
984: near the stellar surface. These photons produce secondary pairs of charged 
985: particles that go on to radiate via coherent curvature radiation, for example. 
986: In this scenario, perhaps the polarization annulus appears only in the core 
987: emission because only the photon beam can provide the large photon occupation 
988: numbers that are conducive to induced scattering. Scattering would be most 
989: evident where the wave vector of the O-mode radiation is perpendicular to 
990: the ambient magnetic field, as might be expected in the multi-polar structure 
991: of the field near the stellar surface where the core emission is thought to 
992: orginate. Furthermore, the specific scattering process described in 
993: \S\ref{sec:perp} will not occur higher in the magnetosphere where the 
994: electrons can occupy higher Landau levels and the cyclotron frequency can 
995: approach the radiation frequency. Finally, the presence of both the 
996: polarization annulus and the sign-changing circular polarization in the core 
997: component of PSR B0329+54 begs a secondary question of whether the two 
998: phenomena are related or entirely coincidental. Single pulse observations 
999: of other pulsars dominated by core emission, such as PSR B1933+16, could 
1000: test whether the phenomena are associated with one another.
1001: 
1002: The analytical model of pulsar polarization presented in this paper can 
1003: describe most, but not all, of the polarization patterns observed in PSR 
1004: B0329+54. In the cone emission of the pulsar, the LEAPs show a data cluster 
1005: with circular cross-section in each hemisphere, consistent with the pattern 
1006: described in Case 2. The cluster in the precursor to the core component has 
1007: the shape of an elliptical bar, which is the pattern produced by Case 4. 
1008: The LEAP observed at the core component shows an elliptical bar in one 
1009: hemisphere and a partial annulus in the other. None of the cases considered 
1010: in \S\ref{sec:patterns} can reproduce this pattern. Additional assumptions, 
1011: such as those described by Melrose et al. (2006) and Edwards \& Stappers
1012: (2004), must be invoked to model the polarization at this pulse location.
1013:  
1014: The oft-stated objective of polarization observations of individual pulses 
1015: (e.g. Manchester et al. 1975; Stinebring et al. 1984) is to understand the 
1016: radio emission mechanism of pulsars. Progress has been made on this front, 
1017: particularly with total intensity measurements that reveal a carousel of 
1018: subbeams circulating about the star's magnetic pole (e.g. Deshpande \& 
1019: Rankin 1999). But in most cases, propagation effects in the pulsar 
1020: magnetosphere, not radio emission mechanisms, are invoked to interpret the 
1021: results of these observations. Afterall, a multitude of propagation effects 
1022: in the ISM complicates our reception of the pulsar signal, so we should not 
1023: be surprised that propagation effects in the pulsar magnetosphere complicate 
1024: our view of what happens there. For example, the occurrence of OPMs has been 
1025: attributed to the birefringence of the magnetospheric plasma above the pulsar 
1026: polar cap (Allen \& Melrose 1982; Barnard \& Arons 1986; Petrova 2001). 
1027: Cyclotron absorption high in the magnetosphere (Luo \& Melrose 2001; 
1028: Melrose 2003) has been proposed as a possible origin of circular polarization 
1029: in the emission. In this paper, GFR and scattering are suggested as possible 
1030: candidates for the origin of some polarization patterns. Individually and 
1031: collectively, these propagation effects can obscure the polarization of the 
1032: underlying emission mechanism. If these interpretations are correct, they 
1033: imply that single pulse polarization observations are best suited for probing 
1034: propagation effects in pulsar magnetospheres. The emission mechanism may 
1035: reveal its secrets through total intensity observations, similar to the high 
1036: time resolution measurements by Hankins \& Eilek (2007) who recently found 
1037: multiple, narrow, radiation bands in the emission from the Crab pulsar.
1038: 
1039: \section{CONCLUSIONS}
1040: \label{sec:conclude}
1041: 
1042: An empirical, analytical model of pulsar polarization has been generalized
1043: to accommodate a wide variety of polarization fluctuation geometries. The 
1044: model is based upon the proposition that the observed polarization of 
1045: pulsar radio emission is due to the incoherent superposition of highly 
1046: polarized orthogonal modes. For the modes to propagate independently,
1047: generalized Faraday rotation may be operative in the pulsar magnetosphere 
1048: for the modes to get significantly out of phase. The model replicates the 
1049: polarization patterns observed in many objects and reproduces the numerical 
1050: results from other work. When the fluctuations are parallel to the 
1051: polarization vectors of the wave propagation modes, the patterns consist 
1052: of two tight clusters, each in a separate hemisphere of the Poincar\'e 
1053: sphere. The patterns assume shapes of bars, bow ties, and annuli when the 
1054: fluctuations are perpendicular to the vectors. The more interesting patterns 
1055: occur when the polarization modulation index exceeds unity. The diverse 
1056: polarization patterns are not likely to originate from the same physical 
1057: process. The parallel polarization fluctuations are caused by fluctuations 
1058: in the polarized intensities of the orthogonal modes. The perpendicular 
1059: fluctuations may be caused by a stochastic version of generalized Faraday 
1060: rotation, which would require large fluctuations in the difference between 
1061: mode phases. An expansion of the two dimensional theory of stochastic 
1062: Faraday rotation in the ISM to the three dimensional case for pulsar 
1063: magnetospheres may aid the interpretation of the observed polarization 
1064: patterns. An alternative model suggests that one mode may experience 
1065: fluctuations perpendicular to its polarization vector while the other 
1066: does not, implying the presence of a mode-selective, random process, such 
1067: as scattering in the pulsar magnetosphere. The polarization patterns reflect 
1068: polarization instabilities that can depolarize the emission in a way that 
1069: is similar to stochastic Faraday rotation in the ISM. The depolarization 
1070: has been quantified with a polarization stability factor for the simpler 
1071: fluctuation geometries. The stability factors imply that pulsar-intrinsic 
1072: effects are more effective in depolarizing the emission than fluctuations 
1073: in the orientation of its polarization vector. For all geometries evaluated 
1074: with the model, the joint probability density of the polarization vector's 
1075: orientation angles follows the same functional form apart from parameters 
1076: determined by the geometry of the polarization fluctuations. The conditional 
1077: density of the orientation angles in all cases follows the Fisher and 
1078: Bingham-Mardia family of distributions. 
1079: 
1080: \acknowledgements
1081: 
1082: I thank Dan Stinebring for providing the data used in the analysis. 
1083: 
1084: \clearpage
1085: 
1086: \begin{center}
1087: \bigskip
1088: {\bf APPENDIX\\
1089: \medskip
1090: Joint Probability Density of Colatitude and Longitude}
1091: \end{center}
1092: 
1093: The basic functional form of the joint probability density of a polarization 
1094: vector's colatitude, $\theta$, and longitude, $\phi$, for all cases considered 
1095: in \S\ref{sec:patterns} is given by
1096: 
1097: \begin{equation}
1098: g(\theta,\phi) = z{\rm C}{\sin\theta\over{4\pi}}
1099:  \Biggl\{\exp{\Biggl({y^2\over{2}}\Biggr)} \Biggl[1+{\rm erf}
1100:  \Biggl({y\over{\sqrt{2}}}\Biggr)\Biggr](1 + y^2)+y\sqrt{{2\over{\pi}}}\Biggr\},
1101: \label{eqn:joint}
1102: \end{equation}
1103: 
1104: \noindent where $\rm{erf}(x)$ is the error function, C is a constant, and the
1105: parameters $y$ and $z$ are functions of $\theta$ and $\phi$ determined by the 
1106: geometry of the polarization fluctuations. The analytical expressions for C, 
1107: $y$, and $z$ for each case are listed below. 
1108: 
1109: {\bf Case 1:}
1110: 
1111: \begin{equation}
1112: {\rm C} = \exp\Biggl(-{s^2\over{2}}\Biggr)
1113: \end{equation}
1114: 
1115: \begin{equation}
1116: y(\theta,\phi) = s\cos\theta
1117: \end{equation}
1118:  
1119: \begin{equation}
1120: z(\theta,\phi) = 1
1121: \end{equation}
1122: 
1123: {\bf Case 2:}
1124:  
1125: \begin{equation}
1126:  {\rm C}= \exp{\Biggl[-{s^2\over{2(1+\rho^2)}}\Biggr]}
1127: \end{equation}
1128: 
1129: \begin{equation}
1130: y(\theta,\phi) = {s\cos\theta\over{[(1 + \rho^2\sin^2\theta)
1131:                  (1 + \rho^2)]^{1/2}}}
1132: \end{equation}
1133:  
1134: \begin{equation}
1135: z(\theta,\phi) = {(1+\rho^2)\over{(1 +\rho^2\sin^2\theta)^{3/2}}}
1136: \end{equation}
1137: 
1138: {\bf Case 3:}
1139:  
1140: \begin{equation}
1141: {\rm C} = \exp\Biggl(-{s^2\over{2}}\Biggr)
1142: \end{equation}
1143: 
1144: \begin{equation}
1145: y(\theta,\phi) = s\cos\theta\Biggr({1+\eta^2
1146:                  \over{1 + \eta^2\cos^2\theta}}\Biggl)^{1/2}
1147: \end{equation}
1148: 
1149: \begin{equation}
1150:  z(\theta,\phi) = {(1+\eta^2)^{1/2}\over{(1 +\eta^2\cos^2\theta)^{3/2}}}
1151: \end{equation}
1152: 
1153: {\bf Case 4:}
1154:  
1155: \begin{equation}
1156: {\rm C} = \exp\Biggl({-s^2\over{2}}\Biggr) 
1157: \end{equation}
1158: 
1159: \begin{equation}
1160: y(\theta,\phi) = s\cos\theta\Biggr[{1 + \eta^2\over{1 + \eta^2
1161:    (\cos^2\theta +\sin^2\theta\sin^2\phi)}}\Biggl]^{1/2}
1162: \end{equation}
1163:  
1164: \begin{equation}
1165: z(\theta,\phi) = {(1+\eta^2)
1166:                  \over{[1+\eta^2(\cos^2\theta+\sin^2\theta\sin^2\phi)]^{3/2}}}
1167: \end{equation}
1168: 
1169: {\bf Case 5:}
1170:  
1171: \begin{equation}
1172: {\rm C} = \exp\Biggl[{-s^2\over{2(1+\rho^2)}}\Biggr] 
1173: \end{equation}
1174: 
1175: \begin{equation}
1176: y(\theta,\phi) = s\cos\theta
1177:                   \Biggr({1+\eta^2\over{1+\rho^2}}\Biggr)^{1/2}
1178:                   \Biggr\{{1\over{(1+\eta^2) + \sin^2\theta
1179:                  [(\rho^2-\eta^2)+\eta^2(1+\rho^2)\sin^2\phi]}}\Biggl\}^{1/2}
1180: \end{equation}
1181: 
1182: \begin{equation}
1183: z(\theta,\phi) = {(1+\eta^2)(1+\rho^2)\over{\{(1+\eta^2)
1184:        + \sin^2\theta[(\rho^2-\eta^2)+\eta^2(1+\rho^2)\sin^2\phi]\}^{3/2}}}
1185: \end{equation}
1186: 
1187: 
1188: \noindent Case 5 provides a general joint probability density for $\theta$ 
1189: and $\phi$. It becomes the joint density for Case 1 when $\eta=\rho=0$, for 
1190: Case 2 when $\eta=0$, and for Case 4 when $\rho=0$. When $s=0$, the joint 
1191: probability density becomes $g(\theta,\phi)=z(\theta,\phi)\sin\theta/4\pi$ 
1192: for all cases since both C and the bracketed term in Equation~\ref{eqn:joint} 
1193: become equal to one.
1194: 
1195: \clearpage
1196: 
1197: \begin{references}
1198: 
1199: \reference{} Allen, M. C. \& Melrose, D. B. 1982, Proc. Astron. Soc. 
1200:              Aust., 4, 365
1201: 
1202: \reference{} Barnard, J. J. \& Arons, J. 1986, \apj, 302, 138
1203: 
1204: \reference{} Bingham, C. \& Mardia, K. V. 1978, Biometrika, 65, 379
1205: 
1206: \reference{} Blandford, R. D. \& Scharlemann, E. T. 1976, \mnras, 174, 59
1207: 
1208: \reference{} Chandrasekhar, S. 1960, Radiative Transfer, (New York: Dover)
1209: 
1210: \reference{} Cocke, W. J. \& Pacholczyk, A. G. 1976, \apj, 204, L13 
1211: 
1212: \reference{} Cordes, J. M. \& Hankins, T. H. 1977, \apj, 218, 484
1213: 
1214: \reference{} Deshpande, A. A. \& Rankin, J. M. 1999, \apj, 524, 1008
1215: 
1216: \reference{} Edwards, R. T. 2004, \aap, 426, 677
1217: 
1218: \reference{} Edwards, R. T. \& Stappers, B. W. 2004, \aap, 421, 681
1219: 
1220: \reference{} Fisher, N. I, Lewis, T., \& Embleton, B. J. J. 1987,
1221:              Statistical Analysis of Spherical Data, (Cambridge: Cambridge)
1222: 
1223: \reference{} Goldreich, P. \& Julian, 1969, \apj, 157, 869
1224: 
1225: \reference{} Hankins, T. H. \& Eilek, J. A. 2007, \apj, 670, 693
1226: 
1227: \reference{} Ishimaru, A. 1978, Wave Propagation and Scattering in Random
1228:              Media, (New York: Academic)
1229: 
1230: \reference{} Johnston, S., Karastergiou, A., Mitra, D., \& Gupta, Y.
1231:              2008, \mnras, 388, 261
1232: 
1233: \reference{} Karastergiou, A., von Hoensbroech, A., Kramer, M., Lorimer, D. 
1234:              R., Lyne, A. G., Doroshenko, O., Jessner, A., Jordan, C., \& 
1235:              Wielebinbski, R. 2001, \aap, 379, 270
1236: 
1237: \reference{} Karastergiou, A., Johnston, S., \& Kramer, M. 2003, \aap,
1238:              404, 325
1239: 
1240: \reference{} Karastergiou, A., Johnston, S., \& Manchester, R. N. 2005, 
1241:              \mnras, 359, 481
1242: 
1243: \reference{} Kennett, M. \& Melrose, D. 1998, PASA, 15, 211
1244: 
1245: \reference{} Lee, L. C. \& Jokipii, J. R. 1975, \apj, 196, 695
1246: 
1247: \reference{} Luo, Q. \& Melrose, D. B. 2001, \mnras, 325, 187
1248: 
1249: \reference{} Lyubarskii, Y. E. \& Petrova, S. A. 1996, Astron. Let, 22, 399
1250: 
1251: \reference{} Manchester, R. N., Taylor, J. H., \& Huguenin, G.R. 1973, \apj,
1252:              179, L7
1253: 
1254: \reference{} Manchester, R. N., Taylor, J. H., \& Huguenin, G. R.  1975, 
1255:              \apj, 196, 83
1256: 
1257: \reference{} Mardia, K.V. \& Gadsden, R. J. 1977, Applied Statistics,
1258:              26, 238
1259: 
1260: \reference{} Melrose, D. B. 1979, Aust. J. Phys., 32, 61
1261: 
1262: \reference{} Melrose, D. B. 2003, in ASP Conf. Ser. 302,  Radio Pulsars,
1263:              ed. M. Bailes, D. J. Nice \& S. Thorsett (San Francisco: ASP),
1264:              179
1265: 
1266: \reference{} Melrose, D. B. \& Macquart, J.-P. 1998, \apj, 505, 921
1267: 
1268: \reference{} Melrose, D., Miller, A., Karastergiou, A., Luo, Q. 2006,
1269:              \mnras, 365, 638
1270: 
1271: \reference{} McKinnon, M. M. 2003, \apjs, 148, 519
1272: 
1273: \reference{} McKinnon, M. M. 2004, \apj, 606, 1154
1274: 
1275: \reference{} McKinnon, M. M. 2006, \apj, 645, 551
1276: 
1277: \reference{} McKinnon, M. M. \& Stinebring, D. R. 1998, \apj, 502, 883
1278: 
1279: \reference{} McKinnon, M. M. \& Stinebring, D. R. 2000, \apj, 529, 435
1280: 
1281: \reference{} Pacholczyk, A. G. \& Swihart, T. L. 1970, \apj, 161, 415
1282: 
1283: \reference{} Petrova, S. A. 2001, \aap, 378, 883
1284: 
1285: \reference{} Petrova, S. A. 2008, \mnras, 385, 2143
1286: 
1287: \reference{} Rankin, J. M. 1983, \apj, 274, 333
1288: 
1289: \reference{} Rankin, J. M. 1990, \apj, 352, 247
1290: 
1291: \reference{} Ruderman, M. A. \& Sutherland, P. G. 1975, \apj, 196, 51
1292: 
1293: \reference{} Simonetti, J. H., Cordes, J. M., \& Spangler, S. R. 1984,
1294:              \apj, 284, 126
1295: 
1296: \reference{} Sincell, M. W. \& Krolik, J. H. 1992, \apj, 395, 553
1297: 
1298: \reference{} Smits, J. M., Stappers, B. W., Edwards, R. T., Kuipers, J., and
1299:              Ramachandran, R. 2006, \aap, 448, 1139.
1300: 
1301: \reference{} Spangler, S. R. 1982, \apj, 261, 310
1302: 
1303: \reference{} Stinebring, D. R. 1982, PhD dissertation, Cornell University
1304: 
1305: \reference{} Stinebring, D. R., Cordes, J. M., Rankin, J. M., Weisberg, 
1306:              J. M., \& Boriakoff, V. 1984, \apjs, 55, 247
1307: 
1308: \reference{} von Hoensbroech, A., Kijak, J., \& Krawczyk, A. 1998, \aap,
1309:              334, 571
1310: 
1311: \end{references}
1312: 
1313: \begin{deluxetable}{ccccc}
1314: \tablenum{1}
1315: \tablewidth{450pt}
1316: \tablecaption{Summary of Fluctuation Geometry Cases}
1317: \tablehead{
1318:  \colhead{Case} & \colhead{Fluctuations} & \colhead{Cluster Shape} & 
1319:   \colhead{Axis Orientation} & \colhead{Conditional Density}}
1320: \startdata
1321:  1 & $\sigma_q=\sigma_u=\sigma_v$ & Spheroid & NA & Fisher  \\
1322:  2 & $\sigma_q=\sigma_u<\sigma_v$ & Prolate Ellipsoid& Parallel & BM bipolar \\
1323:  3 & $\sigma_q=\sigma_u>\sigma_v$ & Oblate Ellipsoid& Parallel & BM girdle \\
1324:  4 & $\sigma_q>\sigma_u=\sigma_v$ & Prolate Ellipsoid& Perpendicular & 
1325:       BM hybrid\\
1326:  5 & $\sigma_u<\sigma_q,\sigma_v$ & Irregular Ellipsoid & $\sigma$-dependent
1327:    & BM hybrid\\
1328: \enddata
1329: \end{deluxetable}
1330: 
1331: \end{document}
1332: