gr-qc0305083/cvs.tex
1: \chapter{Cataclysmic Variables}
2:   The method which we have used up to now for setting strong upper limits on gravity-induced 
3:   birefringence is very limited. As mentioned before, this is mainly due to the lack of a reliable 
4:   atmospheric model for magnetic white dwarf stars which could provide realistic estimates for 
5:   the emitted degree of polarized radiation as a function of field strengths and opacities.
6:   
7:   In this chapter we circumvent this restriction by focussing our attention on magnetic 
8:   white dwarfs which are members in close interacting binary systems. These are systems where 
9:   two stars move around the common center-of-mass in a close orbit so that matter can fall along a 
10:   ballistic stream from one star to the other. An important subclass of these cataclysmic variables are
11:   the AM-Her type binary systems where the white dwarf posses a strong magnetic field so that matter
12:   is prevented from forming an accretion disc and, instead, impacts directly on the stellar surface near 
13:   the magnetic poles. During this process highly polarized cyclotron radiation is emitted whose rotational 
14:   modulation provides a unique opportunity not only for setting upper limits on gravity-induced 
15:   birefringence but perhaps also for the very first direct observation of this effect.
16:   
17:   We therefore begin this chapter with a brief introduction into the basic physics of close interacting
18:   binary systems. The characteristics of the polarized radiation are determined by the various parameters
19:   like the temperature structure and height of the shock front, formed when the accreting matter hits the 
20:   stellar surface, as well as by the electron number density of the heated plasma and the angular extend 
21:   of the emission region. A realistic estimate of the emitted degree of polarization therefore depends 
22:   heavily on the quality of the shock front model. With this underlying physical structure, the theory of 
23:   radiative transfer of polarized radiation in the case of cyclotron emission is reviewed. Although the 
24:   theory is well known for decades now, we add a new feature by modifying the resulting Stokes vector 
25:   when we allow for gravity-induced birefringence. 
26:      
27:   Using this theoretical framework we present phase dependent light and polarization curves for the AM-Her 
28:   system VV Puppis which are compared with polarimetric data taken by Cropper \& Warner in 1986 \cite{cw86}. 
29:   Although the observed asymmetries are in agreement with our numerical curves, the flat topped character
30:   seen in observations cannot be achieved numerically. However, by allowing for sufficiently strong 
31:   birefringence we obtain a good conformity with the observations. 
32:  \newpage
33:  
34:   \section{Basic model of interacting binaries}
35:     Approximately $50 - 60$\% of all stars are assumed to be members of binary systems. By involving
36:     nearly the whole spectrum of stars from red giants to black holes as possible binary components, 
37:     these systems often provide examples of astrophysical processes which are among the most exciting 
38:     currently known. From a more practical point of view the most precise mass determinations in
39:     astronomy are available from binary systems where the period of revolution is precisely measurable 
40:     due to mutual occultations.
41:     
42:     The physics of binary systems depends of course on the masses of the involved stars and, more 
43:     importantly, on their spatial separation. If the distance between both components is big enough 
44:     so that they can interact at every evolutionary stage only gravitationally without any related 
45:     mass transfer, then each member of the binary will pass through its individual development without being 
46:     disturbed by its companion. An example for this is Sirius with the white dwarf Sirius B as the 
47:     second star. However, the situation becomes more interesting if the stars move so close to each
48:     other that tidal effects are no longer negligible. In this chapter we are interested in such systems
49:     which consist of a white dwarf as the primary component and an, evolutionary younger and more 
50:     massive, secondary component which can be represented by a main sequence or a red giant star.
51:     If both stars are not too close, their initial spherical shapes becomes teardrop-shaped with the tip
52:     pointing towards the companion. Basically a mass transfer stream between the components can now 
53:     be initialized in two ways, i.e. either by an evolutionary expansion of the secondary component 
54:     or by a continuous decrease of the distance between the stars. In both cases, the secondary component
55:     begins to fill its Roche lobe (the volume surrounding an object within matter is gravitationally 
56:     bound to it) and matter is transfered to the primary component via the inner Lagrange point L1.    
57:     The mechanism of how exactly the distance between the stars could be further decreased is not fully 
58:     understood at the moment. One possibility is that they already originated in close proximity or that 
59:     due to friction of the surrounding material from both components the rotational energy of the system 
60:     was partially converted into internal energy.  
61:     
62:     Once the mass transfer has started, the angular momentum of the system prevents
63:     the stream from falling directly onto the white dwarf surface. Instead the material forms, in most 
64:     cases, an accretion disc where the gas slowly spirals inwards while, again by friction, the 
65:     gravitational energy is converted into thermal energy. Finally, the matter falls onto the white dwarf 
66:     from the inner edge of the disc. However, this is not a continuous process since the material is
67:     mostly accumulated in the accretion disc until the disc becomes unstable and, then, is suddenly 
68:     deflated accompanied by a huge brightening of the disc due to conversion of gravitational to thermal 
69:     energy. Such an event is also known as a dwarf nova. Such close binary systems which interact by
70:     means of a mass transfer stream are called {\em cataclysmic variables}. 
71:     
72:     For our purpose, the most interesting among them are those where the white dwarf posseses a 
73:     ultrastrong magnetic field of the of the order $10 \lsim B \lsim 60$ megagauss (MG), which prevents
74:     the formation of a normal accretion disc and leads, instead, to a well-defined accretion flow
75:     from the secondary component to near the magnetic pole(s) of the white dwarf. These objects are
76:     often called AM-Her systems, after the prototype AM-Herculi discovered in 1923. 
77:      
78:     \newpage  
79:   
80:     \begin{figure}[t]
81:       \centerline{\psfig{figure=art-amher.ps,width=4.5in,height=5.5in}}
82:       \caption{Artist impression of an Am-Her type system.}
83:     \end{figure}   
84:     
85:     \vspace*{1cm} 
86:     \noindent The name ''Polar'' for AM-Hers and related systems was introduced by Krzeminski \& 
87:     Serkowski in 1977 \cite{ks77} because of the strong and variable circular $(\sim 10\% - 30\% \,
88:     \cite{cw86,tap77})$ and linear polarization which is typical for these objects. Certainly 
89:     this attribute makes the polars very appealing for utilizing them for possible measurements 
90:     of gravity-induced birefringence. Therefore, the next sections will give some insight into 
91:     the basic physical processes responsible for the emission of polarized radiation. This, in 
92:     turn provides the basis for a critical analysis of the observed polarization properties with respect to 
93:     gravity-induced birefringence.     
94:    
95:     \clearpage
96:     
97:     \begin{figure}[t]
98:       \centerline{\psfig{figure=am-her.eps,width=5.5in,height=3.5in}}
99:       \caption{Schematic representation of the accretion pattern in AM Hers (adopted from
100:                Wickramasinghe \& Ferrario \cite{wf00}).}
101:       \label{amher-scheme}       
102:     \end{figure}
103:     
104:   \section{Mass transfer and shock models}   
105:     The basic model which has been developed for AM-Her type systems is briefly sketched in 
106:     Fig.\ref{amher-scheme}. It is mainly based on X-ray observations \cite{beu98}, modelling 
107:     of the continuum and emission-line radiation from magnetically confined accretion funnels 
108:     \cite{fw99}, and polarized emission from shocks \cite{fwi90}. 
109:     
110:     For an ordinary, nonmagnetic primary component the matter streams from the secondary star and
111:     accumulates in a disc before impacting on the white dwarf surface. On the other hand the detailed 
112:     accretion mechanisms are much more complicated and even controversial (e.g. \cite{liw97}) if 
113:     the primary component posseses a strong magnetic field. However, it is generally accepted that 
114:     after the material has left the companion star near the inner Lagrangian point, a stream is 
115:     formed which falls freely towards the white dwarf on a ballistic trajectory. If the magnetic field
116:     of the accreting object is strong enough $(\gsim 1$MG) the increasing magnetic pressure will
117:     begin to dominate the material flow at some distance from the white dwarf in a region called the
118:     coupling region. Here, the radial velocity component is reduced and the flow aquires first 
119:     toroidal and then poloidal velocity components as the stream couples onto the field lines.
120:     From that point the material falls again freely towards the accreting object in magnetically confined 
121:     funnels that connect the coupling region with the white dwarf surface. Finally a static shock
122:     above one or both magnetic poles slows and heats the infalling material before it settles on the 
123:     surface with a temperature of $kT \approx 10$keV.
124:     \subsection{The shock region}
125:       The shock region can be regarded as the main source of emission in the optical and the X-ray wave
126:       band. Calculations of the circular polarization light curve presented in this chapter are therefore 
127:       based on a sophisticated model of the emission region, first invented by Wickramasinghe \& Megitt
128:       in 1985 \cite{wme85}.
129:       
130:       The first attempts to model the shock fronts of AM-Her type systems, given by Meggitt \&
131:       Wickramasinghe in 1982 \cite{mew82}, were based on the assumptions of constant temperature
132:       accretion columns. Although successful in explaining some of the gross properties
133:       like the continuum energy distribution, this model, based on cyclotron opacity, 
134:       failed to predict the correct degree of polarization (the predicted values were a factor of 
135:       about 2-3 higher than observed). The model which is discussed here can be regarded as a further 
136:       development of this early model since it includes a temperature structure (shock front) as 
137:       well as free-free opacity as a pure absorption process.  
138:      
139:       Assuming that the ionized material falls freely towards the magnetic poles, it moves a radial
140:       distance $r$ with the velocity       
141:       \begin{equation}
142:         V(r) = \left(\frac{2GM}{r}\right)^{1/2} = 5.2 \times 10^8\left(\frac{M}{M_{\odot}}\right)^{1/2}
143: 	\left(\frac{10^9 \mbox{\footnotesize{cm}}}{R}\right)\left(\frac{R}{r}\right)^{1/2}
144: 	\mbox{\footnotesize{cm}}\,s^{-1} \quad ,
145:       \end{equation}
146:       where $M$ and $R$ are the mass and the radius in (cm) of the white dwarf respectively. Before
147:       the material settles down on the surface, the density is increased in a strong shock by a factor
148:       4 and the velocity decreases by the same factor across the shock front. In the case that this 
149:       shock is formed near the white dwarf surface, the shock temperature is given by 
150:       \begin{equation}
151:         T_S = \frac{3\mu m_H GM}{8 kR} = 6\times 10^8 \mu\left(\frac{M}{M_{\odot}}\right)
152: 	\left(\frac{10^9 \mbox{\footnotesize{cm}}}{R}\right)\mbox{K} \quad ,
153:       \end{equation} 
154:       where $\mu$ denotes the mean molecular weight. The postshock electron density is      
155:       \begin{equation}
156:         N_e = 3.8 \times 10^{17}\left(\frac{F_s}{10^2\mbox{\footnotesize{g cm}}^{-2}s^{-1}}\right)
157: 	\left(\frac{M}{M_{\odot}}\right)^{-1/2}\left(\frac{R}{R_{\odot}}\right)^{1/2}
158: 	\mbox{\footnotesize{cm}}^{-3} \quad ,
159:       \end{equation}  
160:       where $F_s$ is the specific accretion rate. In this context it is important to note that the scope
161:       of this model only allows accretion within a small fraction of the stellar surface namely on
162:       a point like structure. This restriction will later be removed by a more general approach 
163:       that also allows for more extended shock regions.
164:       
165:       The bulk of the translational energy is carried by the ions which achieve a Maxwellian velocity
166:       distribution by ion-ion collisions at a distance $h_{\mbox{\tiny{ion}}}\sim V_{\mbox{\tiny{ion}}}
167:       t_{\mbox{\tiny{ion}}}$, where $t_{\mbox{\tiny{ion}}}$ denotes the ion-ion collision time scale
168:       \cite{mew84}, while the electrons are mainly heated by Coulomb collisions with the ions. The
169:       electron temperature usually differs from the ion temperature and both vary with height in the
170:       shock \cite{ima87,wuc90}. As the gas then settles on the stellar surface it is cooled by
171:       bremsstrahlung and cyclotron radiation. In the case that cyclotron radiation dominates, the
172:       electrons cool faster than they can be heated by collisions with ions, with the consequence that 
173:       ions and electrons have different temperatures, so that the gas in the ''two-fluid'' regime. 
174: %      Since the cyclotron opacity depends on the magnetic field $B_p$ and the electron number density 
175: %      $N_e$ it is useful to summarize this dependence in the scaling factor
176: %      \begin{equation}
177: %        s_0 = c\omega_c/\omega_p^2=1.657\times 10^8 B_p/N_e \mbox{\footnotesize{cm}}
178: %%      \end{equation}
179: %      where $\omega_c$ and $\omega_p$ are the cyclotron and the plasma frequency, respectively.
180: %      Therefore, the optical depth parameter $\Lambda$ is given by
181: %      \begin{equation}
182: %        \Lambda(T,\omega/\omega_c,\theta) = 1/\kappa s_0 \quad . 
183: %      \end{equation}   
184: %      Here, $\kappa$ denotes the cyclotron opacity and $\theta$ the inclination angle between the 
185: %      line-of-sight and the magnetic field. 
186: %      
187:       However, it is important to note that the accretion flow in AM-Her systems is not continuous.
188:       Occasionally these systems drop to low states of reduced brightness which could be explained
189:       by interruptions or reductions in the material stream. The reason for this accretion modulation
190:       is not fully understood at the moment, but it may be related to solar-type magnetic activity
191:       in connection with starspots of the secondary component \cite{hgm00}.
192:     \subsection{Extended emission regions}    
193:       \begin{figure}[t]
194:         \centerline{\psfig{figure=axes.eps,width=4.8in,height=3.5in}}
195:         \caption{Orientation of the dipole axis $d$ and the spin axis $l_s$ relative to the symmetry axis
196: 	         $l_p$ of the emission region and the line-of-sight.}
197:         \label{axes}
198:       \end{figure}
199:       The idea that the emission regions are pointlike sources located at the diametrically opposite magnetic 
200:       poles is insufficient and often contradicted by observations. For example the asymmetries seen in 
201:       polarization light curves of most systems or the wavelength dependence of intensity and polarization in 
202:       AM-Her itself are hardly reproduced by these models. Therefore, Wickramasinghe \& Ferrario \cite{wf87,wf88} 
203:       introduced emission regions which extend across and above the stellar surface. Consequently, this 
204:       new concept allowed for field spread, density and temperature structure, and for displacement from 
205:       the magnetic poles. 
206:       
207:       The field geometry is that of a centered dipole. Using a cartesian coordinate system, the spin axis 
208:       $l_s$ is inclined with respect to the line-of-sight by an angle $i$ in the $y-z$-plane, while the 
209:       symmetry axis of the emission region $l_p$ is identified with the positive $z$-axis making an angle 
210:       $\Delta$ with respect to the spin axis. The emission region itself extends over a spherical cap which 
211:       subtends an angle $2\alpha$ at the center of the star. In this coordinate system, the dipole axis has a
212:       polar angle $\Theta_B$ with $l_p$ and an azimuthal angle $\Phi_B$ with respect to a line perpendicular
213:       to $l_p$ in the same plane as $l_p$ and $l_s$. We can therefore identify this line with the (positive)
214:       $y$-axis of the cartesian system in the direction $l_p\times(l_p \times l_s)$. The relative orientations
215:       of the various axes are shown in Fig.\ref{axes}. 
216:       
217:       For our purpose, the best agreement with observations was achieved by assuming uniform electron 
218:       temperature $T_e$ and fixed electron number density $N_e$. The emission region has a constant radial 
219:       thickness $H$ so that the characteristics of the emission region can be summarized by the optical depth 
220:       parameter 
221:       \begin{equation}
222:         \Lambda_H = 2.01\times 10^8(H/10^8\mbox{\footnotesize{cm}})(N_e/10^{16}\mbox{\footnotesize{cm}}^{-3})
223: 	(3\times 10^7 G/B)
224:       \end{equation}
225:       in the radial direction \cite{wf88}. Of course $\Lambda_H$ varies over the emission region due to changes
226:       in $B$. The magnetic field strength $B$ in the above expression is evaluated at the magnetic pole.
227:   \section{Cyclotron radiation and birefringence}
228:       Since the cyclotron emission is highly polarized, the AM-Her systems provide an exellent opportunity
229:       to test for gravity-induced birefringence. For this purpose we have used a FORTRAN program,
230:       written by D.T. Wickramasinghe which calculates the Stokes vector of cyclotron emission in the 
231:       'point source' approximation of Wickramasinghe and Meggitt \cite{wme85} using a fixed electron 
232:       temperature $T_e$ and a fixed optical depth parameter $\Lambda$. This section will present the 
233:       underlying mathematical formalism with the additional new feature of a gravity-induced alteration
234:       of the Stokes vector. Extended emission regions, introduced in the previous section, are easily
235:       obtained from this concept as a homogeneous composite of pointlike sources.
236:   
237:       Consider an electron in helical motion in a uniform magnetic field directed along
238:       the z-axis in a cartesian coordinate system. The energy emitted per unit 
239:       solid angle $\Omega$ making an angle $\theta$ with the magnetic field and per frequency 
240:       interval $d\omega$ is given by
241:       \begin{equation}\label{cyc1}
242:         \epsilon_0d\omega\,d\Omega=\frac{e^2\omega^2}{2\pi c}\sum_{n=1}^{\infty}F_n\delta_n(y)
243:       \end{equation}
244:       where $\epsilon_0$ is also the coefficient of spontaneous emission, with
245:       \begin{equation}
246:         F_n=(\cot\theta-\beta_{\|}\mbox{cosec}\,\theta)^2 J^2_n(n\xi)+\beta_{\perp}^2 J^{'2}_n(n\xi)
247:       \end{equation}
248:       where $J_n$ is the Bessel function of order $n$ and $J_n^{'}(n\xi)\equiv dJ_n(n\xi)/d(n\xi)$
249:       \cite{bekefi}.
250:       Further, $\beta_{\|}$ and $\beta_{\perp}$ are the parallel and perpendicular components of
251:       the dimensionless velocity $\beta=v/c$, so that we define $\xi=\beta_{\perp}\sin\theta/(1-
252:       \beta_{\|}\cos\theta)$. The argument of the $\delta$ function is
253:       \begin{equation}
254:         y = n\omega_c/\gamma-\omega(1-\beta_{\|}\cos\theta)
255:       \end{equation}
256:       where $\omega_c=eB/mc$ is the electron cyclotron frequency and $\gamma=(1-\beta^2)^{-1/2}$ the
257:       usual Lorentz factor. From this one can easily see, that the radiation spectrum consists of
258:       spectral lines occuring at frequencies
259:       \begin{equation}
260:         \omega = \frac{n\omega_c}{\gamma(1-\beta_{\|}\cos\theta)} \quad ,
261:       \end{equation}
262:       which also gives a relation between $\gamma$ and harmonic number $n$.
263:       
264:       In order to describe a uniform plasma with temperature $T$ and electron number density $N$,
265:       one has to introduce a relativistic Maxwellian distribution in (\ref{cyc1}) and integrate over
266:       $\beta_{\|}$  
267:       \begin{equation}\label{cfi}
268:         \epsilon_I\,d\omega\,d\Omega=\frac{e^2\omega^2}{4\pi c}\frac{N\mu}{K_2(\mu)}\sum_{n=1}^{\infty}
269: 	\int_{-1}^1F_I\exp(-\mu\gamma)(\gamma^4/n\omega_c)d\beta_{\|}\,d\omega\,d\Omega
270:       \end{equation}
271:       where the dimensionless electron temperature $\mu=mc^2/kT$ serves as the argument of the modified 
272:       Bessel function of the second kind, $K_2(\mu)$.
273:       
274:       To obtain the emission coefficients $\epsilon_Q$ and $\epsilon_V$ for Stokes $Q$ and Stokes $V$,
275:       one defines a coordinate system where the z-axis serves as the line-of-sight and the uniform
276:       magnetic field lies in the x-z plane, making an angle $\theta$ with the z-axis. Then, the complex
277:       Jones vector for the electric field of the radiation of a single electron is proportional to 
278:       (\cite{bekefi})
279:       \begin{equation}
280:         \left(
281: 	  \begin{array}{c}
282: 	     (\cot\theta-\beta_{\|}\mbox{cosec}\,\theta)J_n(n\xi)\\
283: 	     -i\beta_{\perp} J^{'}_n(n\xi)\\
284: 	     0
285: 	  \end{array}
286: 	\right)\quad .
287:       \end{equation}
288:        The emission coefficients $\epsilon_Q$ and $\epsilon_V$ for the Stokes parameters $Q$ and $V$ are 
289:        obtained by replacing $F_I$ in (\ref{cfi}) with
290:       \begin{equation}
291:         F_Q = (\cot\theta-\beta_{\|}\mbox{cosec}\,\theta)^2 J^2_n(n\xi)-\beta_{\perp}^2 J^{'2}_n(n\xi)
292:       \end{equation}
293:       and
294:       \begin{equation}
295:         F_V = -2(\cot\theta-\beta_{\|}\mbox{cosec}\,\theta)J_n(n\xi)\beta_{\perp} J^{'}_n(n\xi)
296:       \end{equation}      
297:       respectively, while $\epsilon_U=0$. Since $0\leq \xi \leq 1$ the Bessel function $J_n$ and its
298:       derivative can be replaced with the Wild-Hill approximations \cite{whill} which are more easy to
299:       implement in a computer program
300:       \begin{eqnarray}
301:         J_n(n\xi) &\approx& \frac{\xi^n\exp(nw)}{\sqrt{2\pi n}(1+w)^n}[w^3+0.5033/n]^{-1/6} \\
302: 	J^{'}_n(n\xi)&\approx& \frac{\xi^{n-1}\exp(nw)}{\sqrt{2\pi n}(1+w)^n}[w^3+1.193/n]^{1/6}
303: 	[1-1/5n^{2/3}]
304:       \end{eqnarray}
305:       with $w=\sqrt{1-\xi^2}$. The cyclotron absorption coefficients are given by Kirchhoff's law,
306:       since the electrons are in thermal equilibrium, i.e.
307:       \begin{equation}
308:         \epsilon_I = \kappa_{\mbox{\tiny{cyc}}} B_{\omega}, \quad \epsilon_Q = q_{\mbox{\tiny{cyc}}}
309: 	 B_{\omega}, \quad \epsilon_V = v_{\mbox{\tiny{cyc}}} B_{\omega},
310:       \end{equation}
311:       with the Rayleigh-Jeans law
312:       \begin{equation}
313:         B_{\omega} = \omega^2 k T/4\pi^3 c^2 \quad .
314:       \end{equation}
315:       In contrast to the first model, given by Meggitt \& Wickramasinghe in 1982 \cite{mew82} the new
316:       scheme also includes free-free opacity as a pure absorption process. The corresponding opacities
317:       are
318:       \begin{eqnarray}
319:         \kappa_{\mbox{\tiny{ff}}} &=& \frac{\omega_p^2(2\omega^4+2\omega^2\omega_c^2-
320: 	                              3\omega^2\omega_c^2\sin^2\theta+\omega_c^4\sin^2\theta)}
321: 	                              {2c\omega^2(\omega^2-\omega_c^2)^2}\,\nu_c \\
322:         q_{\mbox{\tiny{ff}}}      &=& \frac{\omega_p^2\omega_c^2\sin^2\theta(\omega_c^2-3\omega^2)}
323: 	                              {2c\omega^2(\omega^2-\omega_c^2)^2}\,\nu_c \\
324:         u_{\mbox{\tiny{ff}}}      &=& 0 \\
325: 	v_{\mbox{\tiny{ff}}}      &=& \frac{2\omega_p^2\omega\omega_c\cos\theta}
326: 	                              {c\omega^2(\omega^2-\omega_c^2)^2}\,\nu_c 
327:       \end{eqnarray}
328:       with the plasma frequency $\omega_p=(4\pi N_e e^2/m)^{1/2}$ and the approximately collision 
329:       frequency $\nu_c=3.63 N_e T^{-3/2}\ln(2.95\times 10^{11}T/\omega)$. It is assumed that the 
330:       opacity can be treated as a sum of cyclotron and free-free components, hence
331:       \begin{eqnarray}
332:         \kappa = \kappa_{\mbox{\tiny{cyc}}}+\kappa_{\mbox{\tiny{ff}}}, \quad
333: 	     q = q_{\mbox{\tiny{cyc}}} + q_{\mbox{\tiny{ff}}}, \quad
334: 	     v = v_{\mbox{\tiny{cyc}}} + v_{\mbox{\tiny{ff}}} \quad .
335:       \end{eqnarray}
336:       The transfer equation now becomes
337:       \begin{equation}\label{tfr}
338:         \frac{d}{ds}\left(\begin{array}{c}I\\Q\\U\\V\end{array}\right)=
339: 	\left(\begin{array}{c}\epsilon_I\\\epsilon_Q\\0\\\epsilon_V\end{array}\right)+
340: 	\left(\begin{array}{cccc}-\kappa&-q&0&-v\\-q&-\kappa&-f&0\\0&f&-\kappa&-h\\
341: 	-v&0&h&-\kappa\end{array}\right)
342: 	\left(\begin{array}{c}I\\Q\\U\\V\end{array}\right)
343:       \end{equation}
344:       with the Faraday mixing coefficients
345:       \begin{eqnarray}
346:         f &=& (\omega_p^2/c\omega_c)\cos\theta/(\omega^2/\omega^2_c-1) \\
347:         h &=& (\omega_p^2/c\omega_c)\sin^2\theta/2(\omega^3/\omega^3_c-\omega/\omega_c) \quad .
348:       \end{eqnarray}
349:       For uniform conditions these coefficients are constant along a ray. In order to solve the 
350:       transfer equation (\ref{tfr}) set
351:       \begin{eqnarray}
352:         m &=& 0.5(f^2+h^2), \, n = 0.5(q^2+v^2), \, p = qf-vh, \, r=qh+vf, \\
353: 	R &=& [(m+n)^2-p^2]^{1/2}, \, a_1 = (m+n+R)/p, \, a_3 = 1/a_1, \\
354: 	\lambda &=& (n-m+R)^{1/2}, \, \mu = (m-n+R)^{1/2}, \\
355: 	b_1 &=& (f-qa_1)/\lambda, \, b_3 = (f-qa_3)/\mu \\
356: 	c_1 &=& -(h+va_1)/\lambda, \, c_3 = -(h+va_3)/\mu \quad .
357:       \end{eqnarray}
358:       Setting all intensities equal to zero at $s=0$, the Stokes parameters of the radiation after
359:       passing through a distance $s$ are (see \cite{mew82})
360:       \begin{eqnarray}
361:         I/B_{\omega} &=& 1-(p/2R)(a_1\cosh \lambda s - a_3\cos\mu s)\exp(-\kappa s) , \label{icyc}\\
362: 	Q/B_{\omega} &=& -(p/2R)(b_1\sinh\lambda s - b_3\sin\mu s )\exp(-\kappa s) , \\
363: 	U/B_{\omega} &=& (p/2R)(\cosh \lambda s - \cos\mu s)\exp(-\kappa s) , \label{ucyc}\\
364: 	V/B_{\omega} &=& -(p/2R)(c_1\sinh\lambda s - c_3\sin\mu s)\exp(-\kappa s) \label{vcyc}\quad .
365:       \end{eqnarray}
366:       The question of how gravitational birefringence could be implemented in this scheme depends
367:       heavily on the shock height above the stellar surface where the polarized radiation is emitted. 
368:       Since the phase shift $\Delta\Phi$ in (\ref{mag-form}) is proportional to $1/R$ where $R$ denotes 
369:       the radial distance from the stellar center, an emission region which is extended of approximately
370:       one stellar radius above the surface would require an integration process along the shock height
371:       where a gravitationally modified transfer equation has to be solved for each plane parallel slab.
372:       An upper limit on the shock height is given in the 'single-fluid' approximation when bremsstrahlung
373:       is the dominating cooling mechanism. In this case the height $h_{\mbox{\tiny{br}}}$ is simply 
374:       proportional to the product between the free-fall velocity $V_{\mbox{\tiny{ff}}}$ (the index 'ff', 
375:       of course, must not be confused with the index for free-free absorption) and the specific cooling 
376:       time $t_{\mbox{\tiny{ff}}}$ for electrons due to bremsstrahlung. The explicit expression for 
377:       $h_{\mbox{\tiny{br}}}$, given by Wickramasinghe \& Ferrario \cite{wf00} is
378:       \begin{equation}
379:         h_{\mbox{\tiny{br}}}=9.6\times 10^7\left(\frac{10^{16}\mbox{\footnotesize{cm}}^{-3}}{N_e}\right)
380: 	\left(\frac{M}{M_{\odot}}\right)\left(\frac{10^9\mbox{\footnotesize{cm}}}{R}\right)\,
381: 	\mbox{\footnotesize{cm}} \quad .
382:       \end{equation} 
383:       Assuming a white dwarf mass of $M_{\odot}$ with a corresponding radius of $0.014\, R_{\odot}$
384:       this yields a height of $\sim 10^8$cm with $N_e=10^{15}$. Since we focus on a 'two-fluid' model 
385:       when the electrons are cooled by cyclotron radiation, this means that we have for the cooling time
386:       $t_{\mbox{\tiny{cyc}}} \ll t_{\mbox{\tiny{ff}}}$ and, therefore, $h_{\mbox{\tiny{cyc}}} \ll
387:       h_{\mbox{\tiny{br}}}$. For this reason it is justified to view the polarization as emitted from 
388:       near the surface because $h_{\mbox{\tiny{cyc}}} \ll 10^2$km is certainly a sufficient upper limit.
389:       We can therefore assume that birefringence acts directly on (\ref{ucyc}) and (\ref{vcyc})
390:       according to
391:       \begin{equation}
392:         \left(\begin{array}{c}U_{\mbox{\tiny{grav}}}\\V_{\mbox{\tiny{grav}}}\end{array}\right)=
393: 	\left(\begin{array}{cc}\cos\Delta\Phi & 0 \\ 0 & \cos\Delta\Phi\end{array}\right)
394: 	\left(\begin{array}{c}U_{\mbox{\tiny{cyc}}}\\V_{\mbox{\tiny{cyc}}}\end{array}\right) \quad .
395:       \end{equation}      
396:       Our objective is to include birefringence given by these equations in the calculations of synthetic
397:       polarization light curves from AM-Her type systems. In this way we can set sharp upper limits
398:       on this effect with a new method, independent from techniques presented in chapter two.    
399:   \section{Polarimetry of VV Puppis}  
400:     VV Puppis serves as a key system for our purpose of either detecting weak gravitational birefringence or
401:     setting strong upper limits on it by means of gravitationally modified polarization curves. We therefore 
402:     present here some of the most important results that have been achieved for this system within the 
403:     last decades and which provide the basis for the subsequent analysis.
404:      
405:     VV Puppis was discovered in 1931 by van Gent \cite{gent} as a faint $(V = 14.5-18)$ variable with
406:     a period of 100 min. Sinusoidal variations in radial velocity of the H and He II line, correlated
407:     with the 100 min photometric variation led Herbig in 1960 to the conclusion that the observations 
408:     could best be explained by means of a binary system where the emission lines originate on the brighter 
409:     (primary) component \cite{herb}. Herbig also concluded that changes in the shape of the light curve
410:     could have its origin in variations of the emission area on the visible stellar surface which can be
411:     realized by self eclipses of the relevant area due to rotation. This picture was refined in 1977 by 
412:     Tapia's discovery of strong linear and circular polarization with a maximum of $\sim 16$\%\cite{tap77a} 
413:     in each polarization state of the optical light curve, confirming the suggestion by Bond and Wagner 
414:     (IAUC 3049) that this object is similar to the ''original'' AM-Her system. Two years later, 
415:     Visvanathan and Wickramasinghe identified a series of absorption features in the spectrum with 
416:     the 6th, 7th and 8th harmonics of cyclotron absorption in a nearly uniform magnetic field of 
417:     $\sim 3\cdot 10^7$G. This discovery renewed the general interest in astrophysical cyclotron emission 
418:     models, since it was the first detection of resolvable cyclotron harmonics in the optical spectrum
419:     of a stellar object. However, these results had an uncertainty of one harmonic in the identification 
420:     since the observed features did not have well defined cores. As a consequence the corresponding 
421:     uncertainty in $B$ was $\sim 20$\%. Concerning the visibility of cyclotron harmonics, numerical 
422:     computations, based on the calculations presented in section 4.3 revealed that very special conditions 
423:     are required to produce consistency with the observed spectrum \cite{wf87}. 
424:     \begin{figure}[t]
425:       \centerline{\psfig{figure=vv-pup.ps,width=5.5in,height=5.in}}
426:       \caption{Intensity and polarization light curves of VV Puppis. From: Copper \& Warner (1986) \cite{cw86}.}
427:       \label{pol-curv}
428:     \end{figure}            
429:     For the shock front model related to VV Puppis one can therefore conclude a viewing angle 
430:     $i \sim 70^{\circ} - 90^{\circ}$ with respect to the magnetic field, low optical depths 
431:     $\Lambda \sim 10^5$ and high temperatures of $T \sim 10$ keV in a magnetic field of $\sim 3\cdot 10^7$G. 
432:     The fact that VV Puppis is the only stellar object to show resolvable cyclotron harmonics in its 
433:     optical spectrum is difficult to explain, although it might be possible that magnetic field broadening 
434:     often renders the features undetectable \cite{wf87}.                          
435:     \begin{figure}[t]
436:       \centerline{\psfig{figure=vvpup-phases.ps,width=7in,height=3.3in}}
437:       \centerline{\psfig{figure=cbar.ps,width=4in,height=0.5in,angle=90}}
438:       \caption{Visualization of the longitudinal magnetic field components of VV-Puppis during successive 
439:                rotational phases. The black spot which is visible from $\phi\approx 0.75$ to $\phi\approx 1.25$
440: 	       markes the position of an emission region with size $\alpha = 10^{\circ}$ and displacements
441: 	       $\Phi_B=90^{\circ}$, $\Theta_B=30^{\circ}$ from the dipolar axis.}
442:       \label{vphases}
443:     \end{figure}    
444:    
445:     Fig.\ref{pol-curv} shows phase dependend, wavelength averaged polarization curves of VV Puppis from 
446:     Cropper \& Warner (1986) \cite{cw86}. As one can see, the light curves are strongly modulated at the 
447:     orbital period and the accreting pole is visible only for $\sim 40$\% of the period. A common feature is 
448:     the slow rise after eclipse egress and a more rapid decline before eclipse ingress. Although linear
449:     polarization is present during the entire bright phase, with a second, lower peak at phase 0.7, the
450:     strong linear polarization peak of $\sim 9$\% is expected to be caused by a magnetic field
451:     configuration that can be seen only briefly at a viewing angle of $\sim 90^{\circ}$ with respect to 
452:     the line-of-sight. In this case linear polarization due to cyclotron radiation is strongest and would 
453:     therefore explain the observed feature. The circular polarization reaches a maximum value of $\sim 10$\% 
454:     during the bright phase after a more gradual rise. The steeper descent of the circular polarization is 
455:     followed by a sign reversal and a short dip, lasting 0.05 of a cycle. One of the most important results 
456:     is that VV Puppis shows steady negative circular polarization throughout the entire eclipse phase, which 
457:     could be explained by accretion onto a second region of the white dwarf \cite{ls79}, that is visible at 
458:     all rotation phases \cite{wfb89}. However, the negative polarization dip at eclipse ingress is probably 
459:     related to the shape of the primary accretion region. This conjection is supported by synthetic light 
460:     curves of only one accretion region as we will later show.
461:     
462:     On the basis of the light curves in Fig.\ref{pol-curv}, Cropper \& Warner reported for the system 
463:     geometry an inclination of $i = 77^{\circ}\pm 7^{\circ}$ and a magnetic colatitude of 
464:     $\Delta = 155^{\circ}\pm 6^{\circ}$. In 1987 Wickramasinghe \& Ferrario used $i=75^{\circ}$ and 
465:     $\delta=150^{\circ}$ for the calculation of phase dependent light curves of a displaced accreting pole.
466:     Fig.\ref{vphases} shows a visualization of the longitudinal magnetic field component during a rotation
467:     cycle of VV Puppis. The emission region with size $\alpha = 10^{\circ}$ has an azimuthal and polar 
468:     angle of $\Phi_B=90^{\circ}$ and $\Theta_B=30^{\circ}$, respectively, marked by the black spot. 
469:     One can see that in this configuration the emission region merely grazes the limb during the 
470:     bright phase, which is extraordinarily useful for our purpose, since gravitational birefringence 
471:     is most pronounced for sources located at the stellar limb. Further, the asymmetry in the 
472:     linear pulse and the sign reversal in circular polarization, prior to eclipse ingress can now 
473:     be explained by means of the system geometry. Calculating the angles $\theta$ between the field 
474:     directions in the spot and the line-of-sight one finds, that the angles are mostly greater than 
475:     $90^{\circ}$ at eclipse ingress and mostly less than $90^{\circ}$ at eclipse egress. The observed 
476:     reversal in the sign of circular polarization occurs just shortly before eclipse ingress when 
477:     $\theta > 90^{\circ}$ while the stronger linear pulse as well as the maximum intensity occurs at 
478:     $\theta \approx 90^{\circ}$ as expected from the beaming properties of cyclotron radiation. 
479:     
480:    \subsection{Gravitationally modified lightcurves}
481:      The programm which we have developed for calculations of gravitationally modified polarization curves
482:      is based on a FORTRAN routine written by D.T. Wickramasinghe which calculates cyclotron emission in 
483:      the ''point source'' approximation of Wickramasinghe \& Meggitt \cite{wme85} for fixed optical depth 
484:      parameter $\Lambda$ and fixed electron temperature. Since the point source approximation is insufficient
485:      to account for most of the observed polarization curve characteristics as explained in section 4.2.2,
486:      our code describes an extended emission region as a composition of $\sim 3\times 10^2$ point sources,
487:      depending on the size of the emission region. This emission region is placed on a rotating sphere
488:      according to the coordinate system described in sect. 4.2.2, so that the polarization values (circular
489:      and linear) taken at each rotational phase yields the final polarization curve. 
490:      
491:      We first present light curves without gravitational modification in order to have a comparison to the 
492:      observed curves of Cropper \& Warner \cite{cw86}.     
493:      Fig.\ref{vvp-circ1} shows the phase dependent polarization and light curves for two different models. 
494:      While both models have the same inclination $i=75^{\circ}$ and magnetic colatitude $\Delta=150^{\circ}$ 
495:      for the position of the spot, the left column shows the curves for polar cap size $\alpha=5^{\circ}$ 
496:      and a dipole axis orientation with $\theta_B=10^{\circ}$ and $\Phi_B=90^{\circ}$. The curves in the 
497:      right column have $\alpha=10^{\circ}$, $\theta_B=30^{\circ}$ and $\Phi_B=90^{\circ}$. We take note that
498:      these curves are in good agreement with the curves presented by Wickramasinghe \& Ferrario in 1988 
499:      \cite{wf88} (hereafter WF88) for extended emission regions using similar parameters. At this point it 
500:      is important to note, that these curves (including those in WF88 \cite{dayal}) require an additional constant 
501:      background of unpolarised light in order to reproduce observations at least approximately. This constant, 
502:      unpolarized continuum probably comes from a component of the binary system which is always visible, like 
503:      the accretion stream, white dwarf photosphere or something similar - up to now the genuine origin is 
504:      unknown as well as how to correctly estimate it. Currently the only restriction is placed by the observed 
505:      polarization levels and, so, the continuum has to be introduced by hand to get the observed levels right.
506:       
507:      Since our code only allows for emission from one pole, the circular polarization curve does not show 
508:      a constant negative polarization throughout the hole period which is seen in the observations.       
509:      Basic characteristics like the asymmetry in the linear pulse and the negative dip in circular 
510:      polarization prior to eclipse ingress stand comparison to what is observed by Cropper \& Warner. 
511:      Nevertheless we note that the observations shows a more gradual rise in intensity and polarization 
512:      whereas the rapid declines are in agreement with our models.    
513:        
514:      These ''pure'' curves (without any birefringence influence) could possibly better fit\-ted to observations 
515:      by a different geometrical shape of the emission region. While the present curves are based on a cylindrical 
516:      symmetric source, a more elongated, arc-like structure could perhaps account for the observed gradual rise. 
517:      Observational hints for arc-like emission regions has been presented by Beuermann et al. in 1987 \cite{beu87} 
518:      and also by Cropper \cite{crp87}. We will come to this point again after the discussion of the gravitationally 
519:      modified polarization curves. In accordance with WF88 we emphasize that the characteristics of the 
520:      polarization curves depend sensitively on $\Phi_B$ for a given $\alpha$ and $\theta_B$. For instance, for 
521:      $\Phi_B=0^{\circ}$ we obtain no polarization reversals while for $\Phi_B=180^{\circ}$ we observed two 
522:      symmetrical reversals at eclipse ingress and eclipse egress.      
523:      \clearpage       
524:      \begin{figure}[h]
525:        \centerline{\psfig{figure=vvp-int-e5.eps,width=3.08in,height=2.64in}
526:                    \psfig{figure=vvp-int-e10.eps,width=3.08in,height=2.64in}}
527:        \centerline{\psfig{figure=vvp-lin-e5.eps,width=3.08in,height=2.64in}
528:                    \psfig{figure=vvp-lin-e10.eps,width=3.08in,height=2.64in}}
529:        \centerline{\psfig{figure=vvp-circ-e5.eps,width=3.08in,height=2.64in}
530:                    \psfig{figure=vvp-circ-e10.eps,width=3.08in,height=2.64in}}
531:        \caption{Phase dependent circular polarization curve of VV Puppis for an emission region centered at
532:                 latitude $\Delta=150^{\circ}$. The orbital inclination is $i=75^{\circ}$ and 
533: 		$\omega/\omega_c = 6.2$. The curves in the left column correspond to $\alpha=5^{\circ}$ and
534: 		$\theta_B=10^{\circ},\,\Phi_B=90^{\circ}$, while the right column corresponds to 
535: 		$\alpha=10^{\circ}$ and $\theta_B=30^{\circ},\,\Phi_B=90^{\circ}$.}
536:        \label{vvp-circ1}
537:      \end{figure}
538:      \clearpage  
539:      One of the main problems that these curves and also the curves presented by WF88 have is the 
540:      sinusoidal form of the circular polarization and intensity, in contrast to what is observed.
541:      For this reason it is certainly interesting to see how gravity-induced birefringence modifies
542:      the circular polarization curve. Fig.\ref{gcurve-circ} on the next page shows the phase dependent 
543:      circular polarization using the same system parameters as in Fig.\ref{vvp-circ1} and 
544:      $M_{\star}=0.4\,M_{\odot}$, $R_{\star}=0.014\,R_{\odot}$ but this time additionally with 
545:      increasing values of the metric-affine coupling constant $k$. As expected, for $k^2=0$ the curves 
546:      are identical with the unmodified polarization regime while for a sufficiently big $k^2$, the light
547:      is almost completely depolarized. In between these two ''boundary'' states, one can adjust $k^2$ to a 
548:      value $(0.13\,\mbox{km})^2 \leq k^2 \leq (0.14\,\mbox{km})^2$ so that the circular polarization curve 
549:      is nearly flatten as seen in observations. At the same time, the negative dip which arises
550:      prior to eclipse ingress is only marginally influenced by birefringence, although it but also decreases 
551:      somewhat with increasing $k^2$. 
552:      
553:      In order to see if this range for $k^2$ is in agreement with upper limits, derived with the methods 
554:      of the chapter 3, we have made two consistency checks. By using the oblique dipolar rotator 
555:      technique we got an upper limit of $k^2_{VV\,Pup} \leq (0.35\, \mbox{km})$ for VV Puppis without limb darkening 
556:      and $k^2_{VV\,Pup} \leq (0.42\, \mbox{km})$ with maximum limb darkening and an observed polarization level of 10\%. 
557:      These limits are similar to those of 40 Eridani B from the previous chapter, i.e. $k^2_{40\, Eri\,B} \leq (0.35\, 
558:      \mbox{km})$ and $k^2_{40\, Eri\,B} \leq (0.44\, \mbox{km})$, respectively with $M_{40\, Eri\,B}=0.5 M_{\odot}$ and 
559:      $R_{40\, Eri\,B}=0.0136\,R_{\odot}$ comparable to $M_{VV\,Pup}=0.4M_{\odot}$ and $R_{VV\,Pup}=0.014\,R_{\odot}$.
560:      The other white dwarfs of chapter 3 have higher masses and/or smaller radii so that their upper values 
561:      on $k^2$ are correspondingly lower. The same is true for the $k^2$ value of the sun.
562:      However, one has to note that this dipolar rotator technique assumes that the emitted degree of 
563:      polarization is direct proportional to the longitudinal magnetic field strength, whereas we have seen 
564:      that this relation becomes much more complicate in the case of cyclotron radiation. It is therefore 
565:      more reliable to use the ''plain'' model proposed by Solanki, Haugan and Mann in 1999 \cite{shm99} which 
566:      assumes that the polarized radiation is emitted 
567:      homogeneously over the visible stellar surface. Since this is a rather conservative restriction 
568:      the resulting upper limit is congruously $k^2 \leq (0.4\, \mbox{km})^2$. Nevertheless 
569:      it is important that the range of $k^2$-values for VV Puppis is consistent with the limits from 
570:      the dipolar rotator model as well as with the plain model.
571:   
572:      Since the amount of Stokes $U$ in the linear polarization signal is always $\lsim 5$\% the linear 
573:      polarization curve is hardly affected by gravity-induced birefringence which only acts on Stokes $V$ and
574:      Stokes $U$ as one can see in Fig.\ref{gcurve-lin}. Nevertheless, the first smaller peak becomes sharper 
575:      and more pronounced for $\alpha=5^{\circ}$ while for $\alpha=10^{\circ}$ the peak is washed out.
576:      \clearpage
577: 
578:      \begin{figure}[t]
579:        \centerline{\psfig{figure=gcurve-c05.eps,width=4.in,height=3.5in}}
580:        \centerline{\psfig{figure=gcurve-c10.eps,width=4.in,height=3.5in}}
581:        \caption{Gravitationally modified, phase dependent cicular polarization curves. The physical parameters 
582:        are the same as in Fig.\ref{vvp-circ1}. Upper curves: $\alpha=5^{\circ}$, $\theta_B=10^{\circ},\,
583:        \Phi_B=90^{\circ}$. Lower curves: $\alpha=10^{\circ}$, $\theta_B=30^{\circ},\,\Phi_B=90^{\circ}$.}
584:        \label{gcurve-circ}
585:      \end{figure}
586:      
587:      \clearpage
588:      
589:      \begin{figure}[t]
590:        \centerline{\psfig{figure=vvp-gcurve5-lin.eps,width=4.in,height=3.5in}}
591:        \centerline{\psfig{figure=vvp-gcurve10-lin.eps,width=4.in,height=3.5in}}
592:        \caption{Gravitationally modified, phase dependent linear polarization curves. The physical parameters 
593:        are the same as in Fig.\ref{vvp-circ1}. For comparison the same $k$-values as for circular polarization
594:        have been plotted. Upper curves: $\alpha=5^{\circ}$, $\theta_B=10^{\circ},\,\Phi_B=90^{\circ}$. 
595:        Lower curves: $\alpha=10^{\circ}$, $\theta_B=30^{\circ},\,\Phi_B=90^{\circ}$.}
596:        \label{gcurve-lin}
597:      \end{figure}
598:      
599:      \clearpage
600:  \section{Comparison of the results}
601:    Finally, we want to check how the limits on $k^2$ for VV-Puppis, obtained by lightcurve fitting and
602:    the polarization modelling technique from chapter 3 fit into the empirical scheme, given by (\ref{k-phi}).
603:    For this purpose, Fig. \ref{poutII} shows again the curve from Fig. \ref{pout} additionally with the 
604:    $k^2_{VV\, Pup}$ values, obtained in this chapter.
605:    \begin{figure}[h]
606:      \centerline{\psfig{figure=poutII.eps,width=4.5in,height=4.5in}}
607:      \caption{$k^2(\Phi)$ curve of chapter 3 together with the VV Puppis limits. The upper star $(0.6|0.385)$ 
608:      marks the limit obtained by the polarization modelling technique. The lower star $(0.6|0.15)$ marks the
609:      limit from lightcurve fitting. A better fit to the VV Puppis value is provided by the dashed curve.}
610:      \label{poutII}	     
611:    \end{figure}
612:    The limit which is obtained by the polarization modelling technique deviates slightly from the curve. 
613:    Nevertheless the fitting can be improved with a minimally different function 
614:    \begin{equation}\label{k-phiII}
615:      k^2(\Phi) = 1,026262\cdot (\exp(-\Phi^{1.1})+0.1\,\Phi^{-0.18}) \quad ,
616:    \end{equation}  
617:    which yields the dashed line in the above figure. The bigger deviation of the lower point $(0.6|0.15)$ from
618:    the curve is not surprising, since its intention is not to serve as an upper but as a lower limit on $k^2$.
619:    Certainly more datapoints measured at different celestial bodies are needed to see which curve can best
620:    approximate the upper limits on $k^2$.   
621:      
622:  \clearpage
623:  \section{Conclusions}
624:     The main difference between the results of this chapter and those of chapter 3 is obvious. While for
625:     ''ordinary'' magnetic white dwarfs we merely set upper limits on $k^2$ we now additionally have to
626:     assume a lower bound on $k^2$ in order to link the presented theory of cyclotron radiation in AM-Her type 
627:     systems with observations. However one has to be very careful when making an interpretation of these 
628:     results since the underlying model of interacting binary systems is afflicted by many uncertainties:
629:     \begin{enumerate} 
630:       \item The fact that a constant unpolarized continuum background has to be introduced by hand to 
631:             reproduce observations is unsatisfactory. A certain aspect of the physics of binary systems is
632:             therefore probably not understood and although it is rather unlikely that this aspect could 
633: 	    lead to a modification of the present lightcurves similar to the gravitationally modified 
634: 	    curves, it cannot be completely ruled out.
635:       \item More importantly, Wickramasinghe \& Ferrario \cite{wf90} have shown that an arc-shaped emission 
636:             region is also able to replace the sinusoidal shape seen in the circular polarization curve by a 
637: 	    flat top.
638:     \end{enumerate}
639:     At a first glance, the second point seems to discard any claim for a lower bound on $k^2$. Indeed, 
640:     we are here confronted with two alternative models which try to explain the same observed polarization 
641:     properties with different approaches. Currently there is no reliable and convincing possibility to judge 
642:     in favour of one or the other approach. Although the agreement of the conventional approach with observations 
643:     seems to be very appealing, the new question which now arises is, if the assumption of an arc-shaped emission 
644:     region for VV-Puppis is justified any longer, taking into account the new results from gravitational 
645:     birefringence. One of the main arguments supporting the arc-shaped emission regions is the flat top character 
646:     of the calculated circular polarization curve as well as the correctly predicted degree of polarization - 
647:     both also provided by gravity-induced depolarization. Concerning future projects it is therefore highly 
648:     desirable to look for specific predictions of the conventional model which do not involve polarized radiation 
649:     and are, as a result, not affected by birefringence. 
650:     
651:     However, as an important conclusion of this chapter we can say that, as long as gravitational birefringence
652:     cannot be excluded, the interpretation of polarized radiation from compact astrophysical objects is always
653:     afflicted with many uncertainties since the real nongravitational source properties may be hidden behind a
654:     birefringent curtain.    
655:     
656:     
657:      
658: 
659: %    So, basically we are confronted with two possibilities: In the first place, the underlying model of 
660: %    cyclotron emission which produces the curves in Fig.\ref{vvp-circ1} may be wrong or, at least, uncomplete 
661: %    so that the subsequent birefringence analysis is meaningless. For instance one has to check if the curves 
662: %    could be better fitted by dropping the basic assumption of a cylindrical emission region in favour of a 
663: %    more arc-like structure. Furthermore, the recent evidence for a second accreting pole \cite{wfb89} has 
664: %    introduced a further complication which also must be considered in interpreting the polarization data. 
665: %    However, the fact that the polar cap model is able to reproduce the observed asymmetries at least 
666: %    qualitatively suggests that it is likely not too far away from reality. Nonetheless, much work on this 
667: %    subject still has to be done.  
668: %    
669: %    Secondly it is of course possible that the basic tenets of the polar cap model are indeed correct and 
670: %    depolarization plays an important role in understanding the polarization properties of AM-Her type systems. 
671: %    In the light of the consequences that a confirmation of a lower limit on $k^2$ would have it is therefore 
672: %    important to look for possible conventional depolarization mechanisms that could lead to results similar to 
673: %    Fig.\ref{gcurve}. Such a mechanism must be somehow related to the source properties itself, since the amount
674: %    of depolarisation is strongly correlated with rotational phase as one can see in the figures. The currently
675: %    only known candiate for this is Faraday depolarization although this mechanism is already taken into account
676: %    in the radiative tranfer equation of sec. 4.3. A critical reanalysis of this point is therefore needed.
677:       
678:   
679:     
680: 
681:   
682:   
683:    
684: