ORBITAL ELEMENTS OF THE SYMBIOTIC STAR Z ANDROMEDAE FROM OPTICAL LINEAR POLARIZATION DURING THE QUIESCENT PHASE

, , , , and

Published 2010 June 11 Copyright is not claimed for this article. All rights reserved.
, , Citation M. Isogai et al 2010 AJ 140 235 DOI 10.1088/0004-6256/140/1/235

1538-3881/140/1/235

ABSTRACT

We present low-resolution spectropolarimetry for the symbiotic star Z Andromedae at four different epochs during the quiescent phase. The linear polarization of the continuum showed a temporal variation; the difference between the maximum and the minimum is 0.3%–0.6% in Stokes q and is larger with shorter wavelengths. Applying scattering models to this variation, we found the variation in the continuum may be correlated with the orbital motion of the binary and estimated the orbital inclination angle ic = 73° ± 14° and the orientation angle Ωc = 80° ± 5°. We also confirmed that the intrinsic linear polarization of the Raman line λ683 varies with the orbital phase; from this modulation, the orbital elements were derived as ir = 41° ± 8° and Ωr = 82° ± 2°. The inclination derived from the continuum has a large error, and the value is larger by twice the error than the inclination angle value derived from the Raman line. The derived orientation, in contrast, is comparable with that derived from the Raman line. The possible inconsistency in the inclination may be due to the simplicity of our adopted model, or it may be caused by a bias effect due to the low quality of the observed continuum polarization data. An accurate estimation of the inclination from the continuum polarization could settle the question, but that estimation requires more frequent observations that cover at least more than a few orbital cycles during the quiescent phase when the observations are not interrupted by the activity of the hot component.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Symbiotic stars are long-period interacting binary systems consisting of a late-type star (usually a giant) and a hot component (usually a white dwarf) that accretes and ionizes a portion of stellar wind from the late-type companion (e.g., Kenyon 1986; Corradi et al. 2003). For these binary systems, the orbital inclination angle, i, is one of the most fundamental parameters because it is a key to estimating the mass of the hot component from the mass function derived from the radial velocity measurement of the cool component,

Equation (1)

where Mc and Mh are the mass of the cool and hot components, respectively. The inclination i can be inferred by photometry, spectroscopy, or polarimetry. Photometry yields a range of inclinations by analyzing the optical light curve for an eclipse feature of the hot component by the cool companion. Since emission lines and most of the bluer optical continuum emanate not from the hot component, but from the more extended ionized nebula, the eclipse feature is not clearly seen on the optical light curve except when the object is in the "active phase," during which the hot component becomes the dominant source of the optical continuum as the temperature of this component decreases and its radius expands (e.g., Mikolajewska & Kenyon 1992; Munari et al. 1995; Nussbaumer & Vogel 1996; Brandi et al. 2005). Spectroscopy constrains i by examining the eclipse feature in the light curves of not only the continuum but also the emission lines of various species with different ionization and excitation potentials in the optical or ultraviolet (UV) region. Since the hot component (Teff ∼ 105 K) in the quiescent phase emits most of its energy in the far-UV region, the eclipse feature is most clearly seen in the light curve of the far-UV continuum flux (e.g., Vogel 1991; Pereira et al. 1995; Kenyon & Mikolajewska 1995; Dumm et al. 1998).

Polarimetry is applicable even to non-eclipsing objects if the polarization is correlated with orbital motion (e.g., Rudy & Kemp 1978; Brown et al. 1978; Manset & Bastien 2000, 2001). For symbiotic binary systems, Raman scattered lines λλ683 and 708, originating from the scattering of O vi resonance lines λλ103 and 104 by neutral hydrogen atoms (Schmid 1989), are used for estimating i (Harries & Howarth 2000; Schmid & Schild 1997, and references therein). Raman scattered lines are suitable because (1) these lines are strongly polarized, (2) the type of scattering particle is known, (3) and the scattering geometry (between the source O vi emitting region, the scattering H i region, and the observer) is relatively simple. Unfortunately, these lines are only detectable in roughly half the symbiotic stars (Allen 1980). Hence, symbiotic binaries without Raman lines require another polarimetric tool to estimate i.

Recently, Brandi et al. (2000) carried out frequent time-series, multicolor polarimetry for 10 symbiotic systems and found that eight of 10 objects showed temporal variations in the continuum polarization. This suggests that the continuum polarization would be useful for estimating i independent of Raman lines. However, no study has examined the variability of the continuum polarization in order to estimate i. Therefore, we have started a program of multi-epoch, low-resolution spectropolarimetry for bright symbiotic stars. Low-resolution spectropolarimetry provides a wide range of optical spectra involving Raman lines and other emission lines, and can extract only the continuum from the prominent emission lines. In this paper, we present the first results of the program for the symbiotic star Z Andromedae (Z And).

Z And is a prototypical symbiotic star and has a circular orbit with a well-established orbital period of 759 days (Formiggini & Leibowitz 1994; Mikolajewska & Kenyon 1996; Fekel et al. 2000). Since the orbital inclination angle i = 47° ± 12° and the orientation angle Ω = 72° ± 6° are also derived from spectropolarimetry of Raman lines (Schmid & Schild 1997), this object is useful as a test of our ability to estimate i from the continuum polarization. Until now, few reports of multicolor polarimetry for this object have appeared (Piirola 1983; Schulte-Ladbeck 1985). Because of the shortage of the observation periods (about one month's worth), these studies did not detect any sign indicating the presence of intrinsic polarization and concluded that the observed polarization is mainly interstellar in origin based on the wavelength dependence in both polarization degree and position angle (P.A.).

In contrast, we carried out long-period (roughly two years) monitoring for Z And. Owing to multiple observations at multiple orbital phases and to our unique capabilities, we succeeded in detecting variability in the continuum. This variability indicates the presence of an intrinsic component. Moreover, we found that this variation may be correlated with orbital motion; we estimated orbital elements by fitting the variation with a scattering model. In Section 2, we explain our observations and data reductions. In Section 3, we describe the observational results during the quiescent phase and discuss the results. In Section 4, we summarize this study.

2. OBSERVATIONS AND REDUCTIONS

2.1. Observations and Reductions

We observed Z And between 1998 December and 2000 August, during which this object was quiescent. The observations were carried out using a low-resolution spectropolarimeter, HBS (Kawabata et al. 1999) at the Dodaira Observatory (DO) and the Okayama Astrophysical Observatory (OAO) of the National Astronomical Observatory of Japan. At the DO, HBS was attached to the Cassegrain focus of the 0.91 m telescope (f/18; 12farcs4 mm−1 at the focal plane). At the OAO, HBS was attached to the Cassegrain focus of the 1.88 m telescope (f/18; 6farcs1 mm−1 at the focal plane). HBS has a superachromatic half-wave plate and a quartz Wollaston prism; the orthogonally polarized spectra are simultaneously recorded on either a front-illuminated TI CCD (1024 × 1024 pixels, 12 μm square pixel−1, used before 1999 August) or a back-illuminated SITe CCD (512 × 512 pixels, 24 μm square pixel−1, after 1999 August). In 1999 August, with the introduction of an SITe CCD, all of the lenses in HBS were replaced with specially designed UV-achromatic ones. The log of the observations is given in Table 1.

Table 1. Log of Observations

UT Date JD Orbital Telescope Total Aperturec Seeingd NSSe
  (2,450,000) Phasea   Exposure Timeb (s) [''] [''] UP/SP/GT
1998 Dec 26 1174.0 1.201 DO 300 × 40 17.4 3.0 ± 0.1 24/7/6
1999 Jan 16 1194.9 1.229 DO 300 × 19 17.4 2.6 ± 0.1 24/7/6
1999 Dec 7 1520.0 1.657 DO 300 × 32 17.4 3.3 ± 0.1 36/18/9
2000 Aug 27 1784.2 2.005 OAO 200 × 32  8.5 1.8 ± 0.1 3/3/4

Notes. aWe adopted the spectroscopic ephemeris of Fekel et al. (2000). bThe total exposure time is expressed as the integrated time per single frame multiplied by the number of frames. cAngular diameter of the D1 diaphragm used in all observations—circular aperture of a 1.4 mm diameter. dFWHM derived from a Gaussian fit for count profiles of both orthogonally polarized spectra along the slit-length direction around 600 nm. This FWHM value includes the effect of the tracking error of the telescope (typically 0farcs5–1farcs0). eThe number of the standard stars observed in each run. See the text (Section 2.1) about the meanings of UP, SP, and GT.

Download table as:  ASCIITypeset image

We used the 1.4 mm diameter circular hole (D1) as the focal diaphragm. This diaphragm has two holes of the same dimension; we put a target star in one hole and the nearby sky in the other. The spectral resolution depends on the seeing conditions on each night (typically 10 nm) because the stellar images were sufficiently smaller than the aperture of this diaphragm as shown in Table 1. All of observations were carried out in three shared-use runs: from 1998 December 21 to 1999 January 31, from 1999 November 6 to 1999 December 21, and from 2000 August 25 to 2000 September 3. The calibration data such as the instrumental polarization, the instrumental depolarization, and the zero point of the P.A. are produced using all of standard stars observed in each run.

The instrumental polarization was derived from unpolarized standard stars (UPs) observed almost every night in each run. The level of instrumental polarization was well expressed by a smooth function of wavelength and vectorially removed from the observed Stokes Q and U spectra. In any run, the instrumental polarization was less than 0.7% over the 380– 900 nm range, and its stability (1σ) was within 0.05%. The factor of the instrumental depolarization, the ratio of the decrease of polarization of the incident radiation through the instrument, was measured by observing UPs through a Glan–Taylor prism (GTs) that practically produces P = 100% linear polarization. This observation also gave us the wavelength-dependent P.A. of the equivalent optical axis of the superachromatic half-wave plate. The zero point of the P.A. on the sky (north on the celestial plane) was determined by observing strongly polarized standard stars (SPs). The observations of SPs and GTs were carried out roughly once every three nights in each run. The UPs and SPs we used are selected from Serkowski (1974a, 1974b), Schmidt et al. (1992), and Wolff et al. (1996); these UPs and SPs are listed in Tables 2 and 3, respectively. The number of standard stars (UPs, SPs, and GTs) observed in each run is also listed in Table 1.

Table 2. Unpolarized Standard Stars

HD No. mv(σ) Spc. P(σ) (%)
21447 5.10(.012) A1 IV 0.051(0.020)
10476 5.24(.012) K1 V 0.016(–)
18803 6.62(.022) G6 V  ... 
20630 4.83(.011) G5 V 0.006(–)
39587 4.40(.010) G0 V 0.013(–)
42807 6.44(.014) G6 V  ... 
65583 7.00(.016) G8 V  ... 
98281 7.29(.011) G8 V  ... 
102870 3.61(.009) F8 V 0.017(–)
103095 6.45(.008) G8 V  ... 
114710 4.26(.014) G0 V 0.018(–)
115617 4.74(.008) G6 V 0.010(–)
142373 4.62(.015) F9 V 0.012(–)
154345 6.76(.012) G8 V  ... 
165908 5.05(.012) F7 V 0.002(–)
185395 4.48(.015) F4 V 0.003(–)

Download table as:  ASCIITypeset image

Table 3. Strongly Polarized Standards

HD No. mv(σ) Spc. P(σ) (%) PA(σ) (deg)
7927 4.99(.016) F0 Ia 3.32(0.04) 92.1(0.2)
19820 7.11(.010) O9 IV 4.82(0.03) 115.1(0.3)
25443 6.74(.027) B0.5 III 5.15(0.03) 135.1(0.2)
204827 7.94(.021) B0 V 5.34(0.02) 58.7(0.4)
25090 7.30(.024) B0.5 III 5.70(–) 135(–)
197770 6.31(.005) B2 III 3.88(–) 131(–)
21291 4.21(.019) B9 Ia 3.5(–) 115(–)
43384 6.25(.009) B3 Ia 3.0(–) 170(–)
198478 4.83(.014) B3 Ia 2.8(–) 3(–)

Download table as:  ASCIITypeset image

A unit of the observing sequence consisted of successive integrations at four P.A.s, 0°, 22fdg5, 45°, and 67fdg5, of the half-wave plate. The obtained images were processed using the reduction package for HBS data, which was outlined in Kawabata et al. (1999). The instrumental polarization, the instrumental depolarization, and the zero point of the P.A. were properly corrected for in the reduction package. Since the uncertainties of the reduced Stokes q and u are the unbiased standard deviation of the average for all observed frames, these uncertainties include all random errors. The systematic error of the calibration data is basically constant in each run. Since these calibration data are obtained by averaging multiple measurements, the systematic error is negligibly small compared to the random error of single measurement. We achieved an overall calibration accuracy of ΔP < 0.1% (see Kawabata et al. 1999, for details of the performance of HBS).

2.2. Orbital Ephemeris

Several solutions have been presented for the orbital ephemeris of Z And; photometrically determined solutions (Formiggini & Leibowitz 1994; Skopal 1998) and spectroscopically determined ones (Mikolajewska & Kenyon 1996; Fekel et al. 2000). Except for the ephemeris given by Skopal (1998), these are consistent with one another to within 0.011 in phase during 1998–2000; the difference is less than the uncertainty in each ephemeris. Owing to the small difference, the choice of the ephemeris among these three ephemerides does not affect the results; we used the most recent ephemeris by Fekel et al. (2000),

Equation (2)

for conjunction with the cool companion in front, where ϕ denotes the orbital phase.

3. RESULTS AND DISCUSSION

The flux and polarization spectra observed in the quiescent phase are plotted in Figure 1; the flux spectra are normalized by the flux at λ = 550 nm. Note that since these flux spectra are not corrected for atmospheric extinction, temporal variation of the flux spectra cannot be discussed. Despite the low spectral resolving power (Δλ∼ 10 nm), we can recognize not only the most prominent line Hα, but also Hβ and He ii λ469 in the flux spectra; for these lines, no clear changes in polarization across the lines were detected. The continuum polarization depends on wavelengths similar to that of interstellar polarization (Serkowski et al. 1975), and the P.A. is roughly constant over the observed wavelength region. In addition to these properties, the continuum polarization appears to be variable. Although the Raman line λ683, indicated by an arrow in the figure, is faint in the flux spectra, significant changes across the line are noticeable in the polarization spectra. In contrast, another Raman line λ708 is too weak to be recognized in our polarization spectra. We describe each component appearing in the spectra individually in the following.

Figure 1.

Figure 1. Observed flux and polarization spectra during the quiescent phase. The top panel shows the unbinned flux uncorrected against atmospheric extinction, normalized by the flux at λ = 550 nm, and shifted by adding the number at the left position; the middle panel displays the polarization degree Pobs; the bottom panel shows the PAobs. Polarization spectra Pobs and PAobs are binned with 20 nm to improve the statistical error (1σ) of each bin. In both polarization spectra, data points with a large error (σP>0.5%) are excluded. An arrow inside the top panel denotes the position of Raman line λ683.

Standard image High-resolution image

3.1. Continuum

3.1.1. Variation of Polarization

As seen in Figure 1, the observed continuum polarization seems to show a temporal variation. To investigate this variation in more detail, we calculated wideband averages of Stokes q and u using Equation (9) in Kawabata et al. (1999). The binning uses boxcar functions for five wavelength ranges of 400–500, 500–600, 600–640, 720–800, and 800–850 nm, excluding Hα and Raman emission lines and covering a wide wavelength region; the binning improved photon noise and the random one while retaining wavelength dependence information. The derived Stokes q, u for each band are listed in Table 4 and plotted as a function of the orbital phase ϕ in Figure 2. This figure indicates that the variation of Stokes parameters is more significant with shorter wavelength bands. This tendency is clearly seen in the wavelength dependence of Δq and Δu, the differences between the maximum and the minimum values among all observations, as listed in Table 5; Δq and Δu increase with shorter wavelength bands.

Figure 2.

Figure 2. Continuum polarizations observed at five wavelength bands and the best-solution curves. The left and right panels show Stokes q and u, respectively. The curves for the bluer three bands (400–500, 500–600, and 600–640 nm) are derived from the simultaneous multiband fit of the BME model.

Standard image High-resolution image

Table 4. Observed Polarizations at Five Wavelength Bands

Phase Band (nm) q (%) u (%) P (%) PA (°)
1.201 400–500 −0.23 ± 0.09 1.41 ± 0.09 1.43 ± 0.09 49.7 ± 1.9
  500–600 −0.17 ± 0.04 1.41 ± 0.04 1.42 ± 0.04 48.5 ± 0.7
  600–640 −0.18 ± 0.04 1.39 ± 0.04 1.40 ± 0.04 48.6 ± 0.8
  720–800 −0.11 ± 0.03 1.22 ± 0.03 1.22 ± 0.03 47.7 ± 0.7
  800–850 −0.10 ± 0.03 1.09 ± 0.03 1.10 ± 0.03 47.5 ± 0.7
1.229 400–500 −0.50 ± 0.18 1.34 ± 0.18 1.44 ± 0.18 55.3 ± 3.6
  500–600 −0.27 ± 0.04 1.36 ± 0.04 1.39 ± 0.04 50.5 ± 0.8
  600–640 −0.30 ± 0.03 1.29 ± 0.03 1.33 ± 0.03 51.5 ± 0.6
  720–800 −0.12 ± 0.02 1.16 ± 0.03 1.17 ± 0.03 48.1 ± 0.5
  800–850 −0.13 ± 0.05 1.14 ± 0.05 1.15 ± 0.05 48.2 ± 1.1
1.657 400–500 −0.15 ± 0.07 1.39 ± 0.07 1.39 ± 0.07 48.0 ± 1.4
  500–600 −0.13 ± 0.06 1.29 ± 0.06 1.29 ± 0.06 47.8 ± 1.2
  600–640 −0.13 ± 0.06 1.29 ± 0.07 1.30 ± 0.07 47.8 ± 1.3
  720–800 −0.09 ± 0.06 1.14 ± 0.07 1.15 ± 0.07 47.2 ± 1.5
  800–850 −0.06 ± 0.07 1.01 ± 0.07 1.02 ± 0.07 46.7 ± 1.9
2.005 400–500  0.10 ± 0.08 1.22 ± 0.07 1.22 ± 0.07 42.7 ± 1.8
  500–600  0.01 ± 0.08 1.27 ± 0.07 1.27 ± 0.07 44.8 ± 1.8
  600–640  0.01 ± 0.09 1.23 ± 0.08 1.23 ± 0.08 44.8 ± 2.2
  720–800  0.03 ± 0.10 1.10 ± 0.09 1.10 ± 0.09 44.3 ± 2.6
  800–850  0.02 ± 0.15 1.02 ± 0.13 1.02 ± 0.13 44.6 ± 4.2

Download table as:  ASCIITypeset image

Using the averaged Stokes q, u in Table 4, we evaluated the variability by χ2 test for q and u, respectively. A reduced chi-square χ2ν(q) was calculated by

Equation (3)

where ν is the degree of freedom, qj and σj are Stokes q and its uncertainty at a certain band in jth observing run, and $\overline{q}$ is a weighted average over all observing runs from j = 1 to 4, calculated by

Equation (4)

Similarly, we calculated χ2ν(u). The results are listed in Table 5. We adopted the criterion for variability as either χ2ν(q) or χ2ν(u) being larger than that at a significance level of 0.01 (= 3.78 for ν = 3). From this criterion, we recognized that the continuum polarizations of the bluer three bands (400–500, 500–600, and 600–640 nm) were variable during the quiescent phase. This recognition is consistent with the wavelength dependence of Δq and Δu; the variability is detected in the shorter three wavelength bands that have larger Δq.

Table 5. Results of the Variability Test

Band (nm) $\overline{q}$ (%) Δqa (%) χ2ν(q)b $\overline{u}$ (%) Δua (%) χ2ν(u)b
400–500 −0.11 ± 0.04 0.60 4.63 1.33 ± 0.04 0.20 1.34
500–600 −0.18 ± 0.02 0.28 3.88 1.36 ± 0.02 0.14 1.62
600–640 −0.23 ± 0.02 0.31 5.95 1.32 ± 0.02 0.16 1.81
720–800 −0.11 ± 0.02 0.15 0.82 1.18 ± 0.02 0.12 1.00
800–850 −0.10 ± 0.02 0.14 0.42 1.09 ± 0.02 0.13 0.93
Raman 683 −0.11 ± 0.03 0.20 2.34 1.17 ± 0.04 0.19 0.23

Notes. aΔqqmaxqmin and Δuumaxumin. bThe degree of freedom ν = 3 for both q and u.

Download table as:  ASCIITypeset image

3.1.2. Model Fit

For a binary embedded in circumstellar matter, two possible processes can result in a temporal variation of polarization: (1) a secular change in the density distribution of scattering particles in circumstellar space, (2) a periodic modulation due to the orbital motion of binary components without any change in the circumstellar environment. By investigating the periodicity in the variation, we can determine which process causes the detected variation; the investigation needs observations covering at least a few orbital cycles. However, it is reasonable to suppose that during the quiescent phase the circumstellar environment is stable. Thus, in this study, we examined the latter possibility, that is, the temporal variation being due to orbital motion, by fitting this variation with a scattering model.

As for the scattering model, we assumed that the scattering particle was a free electron, a neutral hydrogen atom, or a dust grain with a size much smaller than the optical wavelength. These scattering particles are abundant in symbiotic binary systems, although circumstellar dust grains are probably not abundant in Z And, which is classified as an S-type symbiotic star (Belczyński et al. 2000). The scattering phase functions of these particles have the same dependence on a scattering angle as that of a free electron (Thomson scattering). Thus, this assumption enabled us to use the formulae of Brown et al. (1978, BME) for symbiotic binaries, describing the polarization originating from a binary embedded in a free electron envelope with arbitrary distribution. Additionally, we adopted the geometrical model of Seaquist et al. (1984) for the distribution of the scattering particles; in this model, the density distributions of both ionized and neutral regions have symmetries not only about the orbital plane but also about the axis joining both binary components (binary axis). Although the geometries of both ionized and neutral regions depend on many variables through the dimensionless parameter X (Seaquist et al. 1984), the density distribution symmetry of scattering particles about the binary axis (simultaneously satisfying the symmetry about the orbital plane) holds for any X. Thus, we can use this model for Z And without regard to the value of X.

With this additional constraint for the distribution, the BME formulae can be rewritten as

Equation (5)

Equation (6)

where qco(λ) and uco(λ) denote the constant part of the polarization independent of orbital motion, composed of interstellar and/or intrinsic components, and Pv(λ, ϕ, i) and Θv(Ω, ϕ, i) are defined as

Equation (7)

Equation (8)

with an orbital phase ϕ and the maximum polarization Pmax(λ) at ϕ = 0.25 (and ϕ = 0.75). The polarization of this model has a periodicity of half of the orbital period, and the loci drawn by orbital motion in the Stokes qu plane are an ellipse.

The modulation due to orbital motion is expected to be independent of the wavelength, although the amplitude and the constant component depend on the wavelength. Hence, we applied the formulae to the bluer three bands simultaneously. This multiband fit had 11 free parameters, qco(λ), uco(λ), and Pmax(λ) for each band, and i and Ω, which were common among the three bands. The multiband solution is listed in Table 6 and shown in Figure 2. As shown in this figure, the model accurately reproduces the variation in all three bands despite the simplicity of the model. This indicates that observed polarimetric variation in the continuum may originate from orbital motion.

Table 6. Fit Solutions for Continuum and Raman Line Polarizations

Model Band i (deg) Ω (deg) χ2ν ν
BME 3 bandsa 72.9 ± 13.7 79.7 ± 4.5 1.01 13
Raman 683 40.5 ± 8.3 81.6 ± 1.6 0.21 2

Note. aMultiband fit for the bluer three bands (400–500, 500–600, and 600–640 nm).

Download table as:  ASCIITypeset image

The inclination derived from the continuum ic differs by about twice the error (2σi) from that found from the Raman lines by Schmid & Schild (1997), although the orientation Ωc is comparable with that of Raman lines. As described in the opening paragraph in Section 3, significant changes across the Raman line λ683 are detected in our observed polarization spectra. Thus, to confirm the possible inconsistency in the inclination, we extracted the intrinsic polarization of Raman line λ683 and estimated the orbital elements from our data in Section 3.2.

3.2. Raman Line λ683

To estimate orbital elements from the Raman line, we must extract the intrinsic polarization of the line itself. The intrinsic P.A. of the line P.A.L is

Equation (9)

where qil and uil are intrinsic Stokes parameters of the line region, which consist of both line and continuum components and are derived by subtracting interstellar polarization from the observed one; qiC and uiC are those of the continuum interpolated by the nearby continuum; R is the flux ratio of the line itself to the continuum. For estimating orbital elements, we used only the P.A. of the line because (1) accurate measurement of R is difficult in our low-resolution spectra and (2) the P.A. is less subject to R than the polarization degree when R < 0.1, as shown below.

To extract P.A.L by Equation (9), we needed to determine the interstellar polarization around the Raman line. The observed continuum polarization around the Raman line does not vary significantly with the orbital phase. This is supported by the negative detection of the variability in χ2 test using Equation (3) in Section 3.1.1, as listed in Table 5. Moreover, this continuum polarization, calculated from the weighted average of q and u, $\overline{P} = 1.18\% \pm 0.04$% and $\overline{\mbox{PA}} = 47\mbox{$.\!\!^\circ $}7 \pm 0\mbox{$.\!\!^\circ $}7$, is in good agreement with the estimate of Schmid & Schild (1997), P = 1.20% ± 0.04% and PA = 46fdg9 ± 1fdg2 derived from the spectral range 673–719 nm. Therefore, we supposed that the continuum polarization at this wavelength is mainly interstellar in origin and accordingly adopted the weighted average of the continuum polarizations as the interstellar polarization.

Using this interstellar polarization, the intrinsic polarizations of the line region and the continuum were extracted (Table 7). We confirmed that the intrinsic polarization of the line region is considerably larger than that of the continuum. In addition, the measured flux ratio R, as listed in Table 7, is small (R < 0.1) for all observing runs. Under both conditions, Equation (9) is approximated as

Equation (10)

The extracted P.A. is also listed in Table 7.

Table 7. Intrinsic Polarizations of Raman Line λ683

Phase PAL (°) R qil (%) uil (%) qiC (%) uiC (%)
1.201 −20.8 ± 2.6 0.04 0.76 ± 0.07 −0.49 ± 0.08 0.02 ± 0.05 0.17 ± 0.06
1.229 −14.8 ± 2.7 0.09 0.94 ± 0.09 −0.52 ± 0.10 −0.12 ± 0.06 0.09 ± 0.06
1.657 −36.6 ± 4.2 0.05 0.23 ± 0.06 −0.63 ± 0.08 0.03 ± 0.08 0.05 ± 0.11
2.005 82.6 ± 8.2 0.09 −0.38 ± 0.12 0.10 ± 0.12 0.08 ± 0.06 −0.02 ± 0.06

Download table as:  ASCIITypeset image

The orbital elements are derived by the same formula used by Schmid & Schild (1997),

Equation (11)

The best solution, ir = 40fdg5 ± 8fdg3 and Ωr = 81fdg6 ± 1fdg6, is listed in Table 6 and shown in Figure 3. The curve in this figure is in excellent agreement with the modulation in the extracted P.A.; as shown in the latter panel in the figure, the scatter of the residuals between the model curve and the measurements is small compared to the uncertainty (1σ). Derived elements agree with those of Schmid & Schild (1997); i = 47° ± 12° and Ω = 72° ± 6°.

Figure 3.

Figure 3. Upper panel: intrinsic polarization P.A. for Raman line λ683 and the best-fit curve as a function of the orbital phase. Lower panel: residual of the P.A. between measurements and the model.

Standard image High-resolution image

3.3. Comparison of Orbital Elements between the Continuum and the Raman Lines

In the previous sections, we estimated the orbital elements i and Ω from the continuum and the Raman line polarizations, respectively. The orientation Ω from the continuum is in good agreement with that from the Raman line, but the inclination i from the continuum is inconsistent with that from the Raman line. We consider that the orbital elements estimated from the Raman line are more reliable than those estimated from the continuum for the following three reasons.

  • 1.  
    Our understanding of the scattering particle types and geometry for the Raman line is better than our understanding of the same for the continuum. As described in Section 1, the scattering particle for the Raman line is already known and the line formation scattering geometry is relatively simple. In contrast, the scattering particle for the continuum polarization is not clearly identified, although circumstellar dust grains are likely the major candidate, if present (Schulte-Ladbeck et al. 1990). Therefore, the scattering geometry producing the continuum polarization is also unclear. Moreover, if the continuum polarization is produced through Mie scattering by the (spherical) circumstellar dust grains with a circumference comparable to the optical wavelength, the modulation is expected to be more complicated than that of our scattering model (Simmons 1983; Manset & Bastien 2001); this complication would affect the accuracy of estimating i. The functional form of the modulation also depends on the scattering geometry. Although our model assumes a symmetry for the distribution of the scattering particles about the binary axis, this symmetry would not be realized in the actual circumstances (e.g., Gawryszczak et al. 2003; Mitsumoto et al. 2005, and references therein). Whether our scattering model is too simple for the continuum polarization can be verified by examining the functional form of the observed modulation; this examination requires more data points, that is, more frequent observations during the quiescent phase.
  • 2.  
    The estimation by Schmid & Schild (1997) is based on more abundant observations (six epochs) than ours (four epochs). In addition, the amplitude of the modulation for the Raman line is much larger than that for the continuum polarization. Therefore, the inclination from the continuum may be influenced more strongly by biasing due to low data quality, as is reported by Wolinski & Dolan (1994).
  • 3.  
    The orbital elements from our Raman line data agree with those of Schmid & Schild (1997).

On the basis of the above points, we consider that the inconsistency of i may be due to the simplicity of our adopted model or may be caused by a bias effect as reported by Wolinski & Dolan (1994). To investigate the origin of this inconsistency, more frequent observations during the quiescent phase are urgently required. In addition, to confirm that the variability in the continuum polarization originated from the orbital motion, monitoring observations covering at least more than a few orbital cycles are needed. Moreover, the amplitude of the modulation in the continuum polarization depends strongly on the scattering geometry, which would be changed by the activity of the hot component. Therefore, the monitoring should be carried out during the quiescent phase, uninterrupted by the activity of the hot component.

4. SUMMARY

Our studies of multi-epoch, low-resolution spectropolarimetry for Z And during the quiescent phase are summarized as follows.

  • 1.  
    We detected a temporal variation in the continuum polarization during the quiescent phase; the variation is strong evidence for an intrinsically polarized continuum.
  • 2.  
    Applying the BME scattering model and the geometrical model from Seaquist et al. (1984) to the variation, we found that the variation in the continuum polarization may be correlated with orbital motion and derived ic = 73° ± 14° and Ωc = 80° ± 5°.
  • 3.  
    We confirmed that the intrinsic polarization of the Raman line λ683 showed a modulation correlated with orbital motion; from this modulation, the orbital elements were derived as ir = 41° ± 8° and Ωr = 82° ± 2°. These values agree with those of Schmid & Schild (1997) but are inconsistent in i with the continuum-derived values.

We found more than one possible explanation for the inconsistency in i. To determine the correct explanation, more frequent observations that cover at least more than a few orbital cycles during the quiescent phase (uninterrupted by the activity of the hot component) are strongly required.

We thank T. Cho, K. Ito, S. Nakayama, S. Hamasaka, and K. Homma for their help with observations. We also acknowledge the staff members of DO and OAO for their excellent support. We are grateful to Dr. H. Kawakita and Dr. A. Yonehara for their valuable support and suggestions on this study. We also thank the anonymous referee for his/her helpful comments and suggestions on the manuscript.

Please wait… references are loading.
10.1088/0004-6256/140/1/235