Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

A publishing partnership

Modifying the Standard Disk Model for the Ultraviolet Spectral Analysis of Disk-dominated Cataclysmic Variables. I. The Novalikes MV Lyrae, BZ Camelopardalis, and V592 Cassiopeiae

, , , and

Published 2017 August 31 © 2017. The American Astronomical Society. All rights reserved.
, , Citation Patrick Godon et al 2017 ApJ 846 52 DOI 10.3847/1538-4357/aa7f71

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/846/1/52

Abstract

The standard disk is often inadequate to model disk-dominated cataclysmic variables (CVs) and generates a spectrum that is bluer than the observed UV spectra. X-ray observations of these systems reveal an optically thin boundary layer (BL) expected to appear as an inner hole in the disk. Consequently, we truncate the inner disk. However, instead of removing the inner disk, we impose the no-shear boundary condition at the truncation radius, thereby lowering the disk temperature and generating a spectrum that better fits the UV data. With our modified disk, we analyze the archival UV spectra of three novalikes that cannot be fitted with standard disks. For the VY Scl systems MV Lyr and BZ Cam, we fit a hot inflated white dwarf (WD) with a cold modified disk ($\dot{M}\,\sim $ a few 10−9 M yr−1). For V592 Cas, the slightly modified disk ($\dot{M}\sim 6\times {10}^{-9}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$) completely dominates the UV. These results are consistent with Swift X-ray observations of these systems, revealing BLs merged with ADAF-like flows and/or hot coronae, where the advection of energy is likely launching an outflow and heating the WD, thereby explaining the high WD temperature in VY Scl systems. This is further supported by the fact that the X-ray hardness ratio increases with the shallowness of the UV slope in a small CV sample we examine. Furthermore, for 105 disk-dominated systems, the International Ultraviolet Explorer spectra UV slope decreases in the same order as the ratio of the X-ray flux to optical/UV flux: from SU UMa's, to U Gem's, Z Cam's, UX UMa's, and VY Scl's.

Export citation and abstract BibTeX RIS

1. Introduction: Ultraviolet Disk Spectra and Optically Thin Hard X-Ray in Novalikes

The properties of accretion disks in binary systems are best studied in the brightest types of non-magnetic cataclysmic variables (CVs): the novalikes (NLs) in high state and dwarf novae (DNe) in outburst. These systems are compact binaries with a white dwarf (WD) accreting matter and angular momentum via a disk from a low-mass companion filling its Roche lobe. Dwarf nova systems are found mostly in a state of low mass accretion rate (quiescence) and have periods of high mass accretion rate (outburst) when the disk dominates the UV (Hack & la Dous 1993). NLs are found mostly in a state of high mass accretion in which the UV emission is predominantly originating from the disk (la Dous 1990, 1991). For that reason, the DNe in outburst and NLs in high state are of special interest because they are expected to have an accretion disk radial temperature profile given by the analytical expression Teff(R) of the standard disk model (Pringle 1981).

However, 30 years ago, Wade (1984, 1988) showed, using International Ultraviolet Explorer (IUE) spectra, that some CV systems systematically disagree with the standard disk model when either black body or Kurucz stellar models were used to represent the accretion disk. A better model was developed by Hubeny (1990) explicitly including the calculation of synthetic spectra using the standard disk model (the code TLUSTY; Wade & Hubeny 1998). This modeling was more successful and has been used since in standard procedures to assess the mass accretion rate of systems in a state of high mass accretion rate (e.g., Knigge & Long 2002; Hamilton et al. 2007). With time, however, it appeared that a large fraction of disk systems could not be modeled using the standard disk temperature profile, and instead the temperature of the disk had to be modified, e.g., IX Vel, QU Car, RW Sex; Linnell et al. (2007, 2008, 2010). A statistical study (Puebla et al. 2007) of 33 disk systems indicated that a revision of the temperature profile, at least in the innermost part of the disk, is required: a large fraction of the theoretical spectra were too blue compared to the observed spectra. Some of the UV spectral modelings of disk-dominated systems that were in better agreement with the observations consisted of truncating the hot inner disk (Linnell et al. 2005). By truncating the hottest region in the disk, the slope of the UV continuum became shallower as the spectrum became redder. In her statistical analysis of IUE spectra of CVs, la Dous (1991) already pointed out that dwarf novae in outburst have a continuum flux level increasing toward the shorter wavelength (blue), while the continuum slope is shallower for non-magnetic NLs and it is almost flat for intermediate polars (IPs—systems in which the inner disk is expected to be truncated by the WD magnetic field). la Dous (1991) showed that IUE spectra of IPs were similar in slope/color to theoretical spectra of magnetically truncated disks with a truncation radius ≈5 Rwd, while in contrast, a large number of DNe in outburst have been sucessfully modeled with non-truncated standard disk spectra (Hamilton et al. 2007). This suggests that the disk spectra of non-magnetic NLs, showing a UV continuum slope intermediate between IPs and DNe in outburst, might have a truncated disk with a truncation radius smaller than that in IPs.

In addition to the above problem in the UV spectroscopy of disk-dominated systems, X-ray data of CVs have also been in strong disagreement with theoretical expectations for more than three decades (Ferland et al. 1982). The culprit has been the difficult to study boundary layer (BL) between the accretion disk and the WD star. About half of the disk accretion energy (in the form of kinetic energy) is expected to be dissipated in the BL between the Keplerian disk and slowly rotating stellar surface (Pringle 1981). Because of its small size, the BL was predicted to emit in the X-ray band: at low accretion rates (when the WD is dominant in the UV), the BL was expected to be optically thin and emit hard X-rays (Pringle & Savonije 1979; Tylenda 1981; Narayan & Popham 1993); at large accretion rates, typical of NLs in a high state and DNe in outburst, the BL was expected to be optically thick and emit soft X-rays (Pringle 1977; Narayan & Popham 1993; Popham & Narayan 1995). Systems in the low state indeed reveal optically thin hard X-ray emission (Szkody et al. 2002; Pandel et al. 2005; Mukai et al. 2009). However, systems in a state of high mass accretion often do not show an optically thick soft X-ray component; instead, many exhibit optically thin hard X-ray emission (Patterson & Raymond 1985a, 1985b; Mauche et al. 1995; van Teeseling et al. 1996; Baskill et al. 2005; Balman et al. 2014), with an X-ray luminosity much smaller than expected, i.e., much smaller than the disk luminosity. While optically thin hard X-ray emission from high mass accretion rate systems was unexpected, it is, however, not especially inconsistent with the theoretical work: optically thin BLs can occur in high mass accretion rate systems, since the transition to being optically thin depends not only on the mass accretion rate, but also on the WD mass, the WD rotation rate, and the (unknown) alpha viscosity parameter (Popham & Narayan 1995). Simulations of optically thin BLs (Narayan & Popham 1993; Popham 1999) show that the inner edge of the Keplerian (and optically thick) disk starts at an actual radius R0 = Rwd + δBL, where the size δBL of the BL is of the order of the stellar radius Rwd: δBL ∝ Rwd (the BL is actually geometrically thick). The direct consequence of having an optically thin and geometrically thick BL is that the optically thick Keplerian disk will appear to have an inner hole of size δBL (possibly of the order of the radius of the WD Rwd). Two decades ago, it had already been pointed out that optically thin BL can explain the inner hole observed in circumstellar disks around young stellar objects (T Tauri stars; Godon 1996). In other words, optically thin BLs are consistent not only with the X-ray data, but also with the UV data, as truncated optically thick disks produce a UV continuum with a shallow slope in better agreement with the UV observations than non-truncated disks.

It appears that the high mass accretion rate NL systems seem to systematically exhibit both optically thin hard X-ray emission and a UV continuum with a shallow slope that cannot be fitted with standard disk models, e.g., MV Lyr (Linnell et al. 2005; Balman et al. 2014), IX Vel (Long et al. 1994; Linnell et al. 2007), BZ Cam (Prinja et al. 2000; Balman et al. 2014), V592 Cas (Hoard et al. 2009; Balman et al. 2014), and RW Sex (Córdova et al. 1981; Linnell et al. 2010). Such a behavior indicates that this might be the rule for NL systems as a whole. We suspect that NLs, due to the optically thin BL, have a truncated inner disk with a radius of truncation ≈2 Rwd. Our recent Swift X-ray observations of the three NLs BZ Cam, MV Lyr, and V592 Cas (Balman et al. 2014) show optically thin BLs merged with advection-dominated accretion flows (ADAFs) and/or hot coronae. In such systems, the BL energy is advected inwards as optically thin BLs cannot cool efficiently (Abramowicz et al. 1995). The heating of the stellar surface can further result in an increased/inflated stellar radius.

Motivated by this, we decided to carry out an improved UV spectral analysis of the three NLs BZ Cam, MV Lyr, and V592 Cas by modifying (truncating) the standard disk model. The three NL systems are reviewed in the next section together with their archival FUSE and IUE spectra. The modification of the standard disk model, presented in Section 3, consists mainly in taking the inner radius of the optically thick disk to be at R0 = Rwd + δBL, where Rwd is the non-zero temperature WD radius and δBL is the size of the optically thin BL (which is now also a free parameter of size Rwd or smaller). We caution that generating a standard disk model with an inner disk radius R0 > Rwd is not the same as truncating at R0 a standard disk model generated with an inner disk radius at Rwd, because the inner boundary condition is changed and, therefore, generates a new mathematical solution that has a lower temperature (although similar to the standard disk model, it is different). We briefly present the stellar spectral modeling in Section 4, and show how the disk spectra are affected by our modification. In Section 5, the results show that the modified disk model provides a better fit to the observed UV spectra of these three NLs. In Section 6, we check a possible correlation between the slope of the UV spectrum and the X-ray hardness ratio for a sample of CVs. We also discuss the possibility that the optically thin (truncated) inner region of the disk might be extended in the VY Scl subclass of NLs, and we also quantitatively check the departure of UV spectra from the standard disk model for disk-dominated CV subtypes. We present our conclusion in Section 7.

2. The Three Novalikes MV Lyr, BZ Cam, and V592 Cas

The three systems MV Lyr, BZ Cam, and V592 Cas were the three targets of our Swift X-ray observations (Balman et al. 2014), and while this might seem to be the main reason for choosing them for our UV spectral analysis, these systems have some characteristics that make them especially relevant for the kind of spectral analysis we are performing here. In order to demonstrate that some NL systems are best modeled with truncated disks, we chose systems in which the shallow slope of the UV flux continuum is incontrovertible. Namely, we want to ensure that the shallow slope is not due to an unknown reddening or to the wrong choice of system parameters such as the WD mass, radius, the distance to the system, or its inclination.

With spectral coverage from ∼910 Å to ∼1190 Å, FUSE covers the Lyman series and is therefore more useful in helping determine the hot component (WD temperature and/or mass accretion rate) of CVs than any other UV telescope. However, we wish to use the broadest spectral UV coverage possible, and we try to choose systems that have both FUSE and IUE spectra, thereby extending from the Lyman limit ∼910 Å to the near-ultraviolet ∼3200 Å. There are many hundreds (if not thousands) of known CVs, and almost 300 of them have good IUE spectra from ∼1150 Å to ∼3200 Å. Usually, the IUE spectra are easily used to derive the reddening (E(BV)) toward the system using the 2175 Å feature (with an accuracy of about 25%; Fitzpatrick 1999). Since it is not clear that one can use the same standard reddening law for all CVs in the Galaxy (e.g., Fitzpatrick 1999), it would be preferable to choose systems with zero or negligibly small reddening, E(BV) ≈ 0. In the present work, we use the extinction curve from Fitzpatrick & Massa (2007) with R = 3.1 to deredden the UV spectra.

Another desirable feature is the inclination of the systems. A high inclination system will be more affected by the orbital phase. In addition, the standard disk model is a two-dimensional flat disk and its departure from a realistic three-dimensional disk will be more pronounced at higher inclinations. We therefore prefer to choose systems with a small inclination.

When modeling the accretion disk, it is preferable to know the distance and/or WD mass to derive a more precise mass accretion rate (knowing only one of the two is most often good enough).

These conditions ensure that the failure of the standard disk model is not due to a wrong reddening (or reddening law), an unknown WD mass, or distance. Although it has already been shown statistically that the failure of the standard disk model is due to the model itself, see Puebla et al. (2007).

We list the system parameters in Table 1, from which it appears that MV Lyr is possibly the best choice: it has no reddening, a very small inclination, and the distance, mass (and radius), and temperature of the WD have all been well determined. BZ Cam too is a very good choice, but its WD mass is not known accurately. BZ Cam is known to exhibit an IUE spectrum that has a remarkably flat (see Section 2.2) continuum slope, therefore presenting a challenging test to our modified disk model. As for V592 Cas, it has a large reddening value, but its parameters are otherwise relatively well known; its spectrum presents a UV continuum slope that is rather common for disk-dominated CVs (as discussed in Section 6).

Table 1.  System Parameters

Parameter MV Lyr BZ Cam V592 Cas
${M}_{\mathrm{wd}}\langle {M}_{\odot }\rangle $ 0.73–0.8(1), (2) 0.4–0.7(7), (8) 0.75(13)
${M}_{2\mathrm{nd}}\langle {M}_{\odot }\rangle $ 0.3(1) 0.3–0.4(7) 0.21(13)
$i\langle \deg \rangle $ 10 ± 3(3), (4), (5) ∼15 < 40(9) 28 ± 10(14)
E(BV) 0(6) 0.05(10) 0.22(13)
$P\langle \mathrm{hr}\rangle $ 3.19(4) 3.69(11), (12) 2.76(15)
$d\langle \mathrm{pc}\rangle $ 473 ± 37(1), (2) 830 ± 160(9) 330–364(13), (14)

References: (1) Hoard et al. (2004), (2) Godon et al. (2012), (3) Schneider et al. (1981), (4) Skillman et al. (1995), (5) Linnell et al. (2005), (6) Bruch & Engel (1994), (7) Lu & Hutchings (1985), (8) Greiner et al. (2001), (9) Ringwald & Naylor (1998), (10) Prinja et al. (2000), (11) Patterson et al. (1996), (12) Honeycutt et al. (2013), (13) Hoard et al. (2009), (14) Huber et al. (1998), (15) Taylor et al. (1998).

Download table as:  ASCIITypeset image

In the following subsections, we describe in detail these three systems and their archival UV spectra, providing also their important characteristics.

2.1. MV Lyrae

MV Lyr has an inclination i ∼ 10° (Schneider et al. 1981; Skillman et al. 1995; Linnell et al. 2005) and zero reddening (Bruch & Engel 1994). It is a VY Scl novalike system, i.e., spending most of its time in a high state and occasionally undergoing short-duration drops in brightness, or "low state," when accretion is almost shut off and the WD is revealed in the UV. It has a distance d ∼ 500 pc and a WD mass of Mwd ∼ 0.73–0.80 M (Hoard et al. 2004; Linnell et al. 2005; Godon et al. 2012). Compared to other CVs, the values of these parameters are relatively well known. The orbital period of the binary is P = 3.19 hr and the mass ratio was found to be $q={M}_{2\mathrm{nd}}/{M}_{\mathrm{wd}}=0.4$ (Schneider et al. 1981; Skillman et al. 1995). The system parameters are listed in Table 1.

MV Lyr spends most of its time in the high state, at a visual magnitude of ∼12–13, when the emission is primarily from the accretion disk. From time to time it drops into a low state, reaching a magnitude of ∼17.5, when the emission is possibly entirely from the WD. MV Lyr also has periods during which it is in an intermediate state at a magnitude of ∼14–15. MV Lyr was observed with FUSE and IUE in a low state, with IUE in an intermediate state, and with FUSE and HST/STIS in a high state. The IUE observations of MV Lyr during its low state revealed a hot WD possibly reaching 50,000 K (Szkody & Downes 1982) or higher (Chiappetti et al. 1982). These results were later confirmed with a FUSE spectroscopic analysis in a low state (Hoard et al. 2004), during which the mass accretion rate was estimated to be no more than $\dot{M}\approx 3\times {10}^{-13}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ and the WD temperature was found to be 47,000 K. A follow-up analysis to study the different states of MV Lyr was carried out by Linnell et al. (2005), who found that standard disk models did not fit the observed spectra. Their models improved with the truncation of the inner disk, an isothermal radial temperature profile in the outer disk, or both, and gave a mass accretion rate of the order of 3 × 10−9 M yr−1 for the high state and 1 × 10−9 M yr−1 for the intermediate state.

The departure from the standard disk model is more pronounced in the intermediate state and most evident in the longer wavelengths of IUE. Since the high state was only observed down to ∼1800 Å, in our analysis we only model the system in the low and intermediate states extending above 3000 Å. The modeling of the low state is first carried out (see Section 5.1) to demonstrate the numerical method and to re-derive the system parameters. The observation log is recapitulated in Table 2. The spectra are shown in Figure 1 together with a standard disk model for comparison. The shallow slope of the IUE spectra in the intermediate state appears clearly.

Figure 1.

Figure 1. MV Lyrae has been observed in low, intermediate, and high states; however, only the low and intermediate states have IUE spectral coverage all the way to 3200 Å. Shown here are the FUSE (black) and IUE (SWP+LWP) spectra of MV Lyr obtained in the low state (red), as well as an IUE (SWP+LWP) spectrum taken in an intermediate state (green). The low state reveals the WD, while the intermediate state is expected to be dominated by emission from the disk. However, the slope of the continuum of the spectrum obtained in the intermediate state is significantly shallower than the spectrum of a standard disk model, shown here in blue. The disk model has a central WD mass of 0.8 M, mass accretion rate of 10−8 M yr−1, and inclination of 41°. The disk spectrum has been shifted upwards for clarity as it is displayed only to show the slope of its continuum (there is no attempt here to fit the observed spectra with the theoretical disk spectrum). The observed spectra have not been shifted.

Standard image High-resolution image

Table 2.  Observation Log

System Telescope/ Obs ID Date (UT) Time (UT) Exp. Time State
Name Instrument   yyyy mm dd hh:mm:ss seconds l-i-h
MV Lyr FUSE C0410301 2002 Jul 07 11:56:39 11209 l
  IUE/SWP 07296 1979 Dec 02 20:09:28 5400 i
  IUE/LWP 06288 1979 Dec 02 18:52:24 4200 i
  IUE/SWP 10905 1980 Dec 27 11:29:05 2400 l
  IUE/SWP 10906 1980 Dec 27 13:18:44 9000 l
  IUE/LWP 09589 1980 Dec 27 12:12:51 3600 l
  IUE/LWP 09590 1980 Dec 27 15:50:09 7080 l
BZ Cam FUSE D9050101 2003 Dec 01 17:39:28 13187 h
  IUE/LWP 14485 1988 Nov 19 14:25:08 1500 h
  IUE/LWP 14486 1988 Nov 19 14:45:36 1500 h
  IUE/LWP 14487 1988 Nov 19 17:51:15 1500 h
  IUE/LWP 14491 1988 Nov 20 13:44:17 1500 h
  IUE/LWP 14492 1988 Nov 20 15:38:36 1380 h
  IUE/LWP 14493 1988 Nov 20 17:31:46 1380 h
  IUE/SWP 34775 1988 Nov 19 14:04:30 1800 h
  IUE/SWP 34776 1988 Nov 19 16:23:32 3600 h
  IUE/SWP 34777 1988 Nov 19 18:25:06 1440 h
  IUE/SWP 34784 1988 Nov 20 12:16:20 3600 h
  IUE/SWP 34785 1988 Nov 20 14:21:57 3600 h
  IUE/SWP 34786 1988 Nov 20 16:16:59 3600 h
  IUE/SWP 34787 1988 Nov 20 17:05:03 2520 h
V592 Cas FUSE D1140101 2003 Aug 05 21:22:24 23751 h
  IUE/LWP 12084 1981 Dec 05 19:48:23 1200 h
  IUE/LWP 12085 1981 Dec 05 22:24:02 1680 h
  IUE/SWP 15658 1981 Dec 05 18:48:15 900 h
  IUE/SWP 15659 1981 Dec 05 21:20:02 3600 h

Note. (1) The flux level is h = high, i = intermediate, l = low.

Download table as:  ASCIITypeset image

All of the FUSE spectra in this work were calibrated with the latest and final version of CalFUSE (Dixon et al. 2007), and processed using our suite of FORTRAN programs, UNIX scripts, and IRAF procedures written for this purpose. The details can be found in Godon et al. (2012).

2.2. BZ Camelopardalis

BZ Cam is a VY Scl NL type (Garnavitch & Szkody 1988; Greiner et al. 2001), with a relatively small reddening E(BV) = 0.05 (Prinja et al. 2000), an orbital period of 221 minutes (Patterson et al. 1996), a distance of 830 ± 160 pc (Ringwald & Naylor 1998), and a moderately low inclination ∼12° < 40° (Ringwald & Naylor 1998). Its WD mass is unknown, but in our analysis we assume Mwd ≈ 0.4 M ≤ Mwd ≤ 0.7 M (Greiner et al. 2001). BZ Cam is not a typical CV in that it is surrounded by a faint emission bow-shock nebula (Ellis et al. 1984; Krautter et al. 1987; Greiner et al. 2001), which is proof of an outflow (Patterson et al. 1996). All the system parameters of BZ Cam are listed in Table 1 with their references.

BZ Cam spends most of its time in a high state at a magnitude of ∼12–13. It has been shown to drop occasionally to a magnitude of ∼14. All the UV spectra of BZ Cam were collected in the high state, but show some small variations in the continuum flux level. The first ultraviolet (UV) spectra of BZ Cam were obtained with IUE (Krautter et al. 1987; Woods et al. 1990, 1992; Griffith et al. 1995). We retrieved here 13 IUE spectra from a single epoch (1988 November 19 and 1988 November 20) with the same continuum flux level. They consist of six IUE LWP spectra and seven IUE SWP spectra, which we co-added and then combined (see Table 2). The IUE (SWP and LWP) and optical continuum is consistent with a ∼12,500 K Kurucz LTE model atmosphere with $\mathrm{log}g=2.5$ (Prinja et al. 2000), but it cannot be attributed to the WD, because of its small emitting surface area. The IUE spectra of BZ Cam has a relatively shallow slope when compared to an accretion disk. BZ Cam also has a FUSE spectrum (Froning et al. 2012), which was never previously modeled with realistic disk or WD atmosphere model spectra. The FUSE spectrum is affected by interstellar (ISM) absorption, mainly molecular hydrogen. The FUSE and IUE spectra of BZ Cam are shown (using a coarse binning for clarity) in Figure 2 together with a standard disk model and a 12,500 K stellar atmosphere (with $\mathrm{log}g=8$) spectrum for comparison. The continuum flux level in the FUSE spectrum is slightly lower than the continuum flux level of the combined IUE spectrum in the region where the two spectra overlap. This difference is likely due to a small variation in the flux of BZ Cam, although we cannot exclude the possibility that it could also be due, at least in part, to the fact that the data were obtained with different telescopes, gratings, detectors, and software.

Figure 2.

Figure 2. Dereddened FUSE (in black) and IUE (red) spectra of BZ Cam in its high state drawn on a log–log scale together with the theoretical spectra of a standard disk model (blue) and a 12,500 K stellar atmosphere (orange) for comparison. The accretion disk has a WD mass of 0.8 M, mass accretion rate of 10−8 M yr−1, and inclination i = 18°. The theoretical spectra have been shifted upwards for clarity. We do not attempt here to "fit" the theoretical spectra to the observed ones. The relatively shallow slope of the IUE spectrum does not match the steeper slope of a standard accretion disk, and is even shallower than the 12,500 K isothermal component. On this log-scaled figure, it appears that the IUE spectrum does not especially match the FUSE spectrum both in its continuum flux level and slope. The observed spectra were not obtained at the same time, and since BZ Cam is a variable UV source, the spectra cannot be combined together for modeling. In this figure, the FUSE and IUE spectra have been binned coarsely for clarity.

Standard image High-resolution image

2.3. V592 Cassiopeiae

V592 Cas is a UX UMa novalike subtype and has never been observed in a low brightness state. It has an inclination of i = 28° and a mass ratio of 0.19 (Huber et al. 1998), and its orbital period is Porb = 2.76 hr (Taylor et al. 1998; Witherick et al. 2003). It shows evidence for a bipolar wind outflow (Witherick et al. 2003; Prinja et al. 2004; Kafka et al. 2009) with velocities reaching up to vwind ∼ 5000 km s−1 (Kafka et al. 2009). The reddening toward the system is E(BV) = 0.22 (Cardelli et al. 1989), and its distance is about 240–360 pc (Taylor et al. 1998). The mass accretion rate of the system is believed to be of the order of ≈1 × 10−8 M yr−1 (Taylor et al. 1998; Hoard et al. 2009). In the present work, we assume the WD to have a standard CV WD mass of ∼0.8 M and a temperature of 45,000 K (Hoard et al. 2009).

V592 Cas was observed with IUE on 1981 December 5. We retrieved the two LWR spectra and two of the four SWP spectra that were obtained, and combined and co-added them to generate a spectrum with a spectral range from ∼1150 Å to ∼3200 Å. With a flux of a few 10−13 erg s−1 cm−2 Å−1, the spectrum has a relatively good signal in the SWP segment but it is rather noisy in the LWR segment, as shown in Figure 3 (in red). V592 Cas has three FUSE data sets, obtained on 2003 August 5, 7, and 8. The data sets all have the same quality and about the same exposure time and reveal the same spectrum. We decided to retrieve the first data set (D1140101) and extracted a final spectrum following the same procedure we used for BZ Cam and MV Lyr. The FUSE spectrum of V592 Cas has the same continuum flux level as the IUE spectrum in the region where the two spectra overlap. Like the FUSE spectrum of BZ Cam, the FUSE spectrum of V592 Cas is heavily affected by ISM molecular hydrogen lines, appearing in Figure 3 (in black) as if the spectrum had literally been sliced.

Figure 3.

Figure 3. Dereddened FUSE (solid black) and IUE (red) spectra of V592 Cas in its normal high state plotted on a log–log scale against two standard disk model spectra for comparison. The theoretical spectra have been shifted upward for clarity; there is no attempt here to "fit" the observed spectra with the theoretical ones. The standard disk models have a central WD mass of 0.8 M with a mass accretion rate of $\dot{M}={10}^{-8}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ (green/upper plot) and $\dot{M}={10}^{-9}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ (blue/lower plot), and inclinations of 41° and 18°, respectively. The FUSE flux continuum shape is in better agreement with the 10−8 M yr−1 model at wavelengths shorter than 1000 Å, while the IUE flux continuum slope is even shallower than the 10−9 M yr−1 model. These are clear indications that the spectra do not agree with the standard disk model. The FUSE spectrum is strongly affected by molecular hydrogen absorption lines from the interstellar medium (ISM).

Standard image High-resolution image

In Figure 3 we also plot two standard disk models for comparison. The FUSE continuum is in better agreement with the 10−8 M yr−1 model, while the IUE continuum slope is even shallower than the 10−9 M yr−1 model. These indicate that the combined FUSE + IUE spectrum does not agree very well with the standard disk model. However, the discrepancy is not as strong as for MV Lyr and BZ Cam.

3. The Modified Disk Model

We wish to model the three NL systems in high and intermediate states using a disk model generated by modifying the standard disk model. Therefore, we first review the standard disk model, and then we introduce the modification we make.

3.1. The Standard Disk Model

In the standard disk theory (Pringle 1981), the total accretion luminosity that can be released by accretion onto a star is given by

Equation (1)

where G is the gravitational constant, $\dot{M}$ the mass accretion rate, Mwd the mass of the accreting star (here a WD), and Rwd its radius.

In disk systems, accretion energy is released through the disk. All current disk models are based on the Shakura−Sunyaev "alpha" disk model known as the standard disk model (Shakura & Sunyaev 1973; Lynden-Bell & Pringle 1974). The disk is assumed to be geometrically thin (in the vertical dimension), and the energy dissipated between adjacent rings of matter (Φν, the dissipation function due to the shear) is radiated locally in the vertical direction, $\sigma {T}_{\mathrm{eff}}^{4}={{\rm{\Phi }}}_{\nu }$. The viscosity ν is due to turbulence and has been shown to come from a magneto-rotational instability (Balbus & Hawley 1991). The turbulent viscosity is parametrized with an a priori unknown parameter α (α < 1), which is the reason the standard disk model is also called the alpha disk model. The standard disk model is obtained from the steady-state Navier–Stokes equations in cylindrical coordinates (R, ϕ, z)). The equations reduce to one dimension (R) after integrating in the vertical dimension (z) and assuming axisymmetry (∂ϕ/∂ = 0). Using the conservations of mass (for $\dot{M}$) and angular momentum and imposing the boundary condition that the angular velocity gradient vanishes (∂Ω/∂R = 0) at the inner boundary R = Rwd, one obtains the standard disk model (Pringle 1981). In that approximation, the BL between the star and disk is geometrically thin, and its thickness can be neglected δBL ≪ Rwd.

3.2. The New Boundary Condition

We modify the standard disk model by truncating the inner disk at a radius R0 > Rwd. In order to do this correctly, one cannot just remove the inner disk between Rwd and R0 from a standard disk model starting at Rwd. Instead, one has to generate a standard disk model starting at R = R0. We give below several reasons for having a disk start at R0 rather than Rwd.

The optically thin BL simulations of Narayan & Popham (1993) and Popham (1999) show that the optically thin BL has the size δBL ≈ Rwd and that the angular velocity gradient vanishes at the outer edge of the BL. Namely, the no-shear boundary condition ∂Ω/∂R = 0 has to be imposed at R = R0 = Rwd + δBL > Rwd.

If the inner hole in the disk is due to a weakly magnetized WD truncating the disk (Hameury & Lasota 2002) at R0 (as in IPs), one still has the same new boundary conditions, as R0 is now the corotation radius, which by definition is the radius where ∂Ω/∂R = 0, since Ω = constant there.

If the inner hole in the disk is due to evaporation (Hameury et al. 2000), the same new boundary conditions can also be imposed as the disk only starts at the outer radius (now R0) of the evaporated inner disk.

Also, it is important to note that the radius of the WD is expected to increase as the WD temperature increases. For example, a 10,000 K WD of mass 0.7 M has a radius Rwd = 7.9 × 108 cm, and at 50,000 K, the radius increases by 16% to Rwd = 9.2 × 108 cm (Wood 1995). The increase in radius is even more pronounced for a smaller mass WD and basically reaches a factor of 2 for a 0.4 M WD, with a zero-temperature radius of Rwd = 1.1 × 109 cm versus a 50,000 K radius of Rwd = 2 × 109 cm (Wood 1995). Since the VY Scl systems are notorious for having a temperature much hotter than other CVs at the same period, the change in radius cannot be ignored. So far, most standard disk model spectra were computed assuming a constant (zero-temperature) radius (Wade & Hubeny 1998; Godon et al. 2012), thereby not taking into account the increased WD radius and possibly overestimating the disk temperature. Modeling the disk properly requires that the WD temperature is known, and requires a fine-tuning of the modeling such as the one performed in Linnell et al. (2008).

The modified disk model can therefore be applied to different inner disk structures, including the magnetically truncated disks in IP systems and around hot CV WDs to provide a more realistic approach for the modeling of the disk.

3.3. The Temperature of the Modified Disk

Our modified disk model differs from the standard disk model in that we now impose the no-shear boundary condition at r = R0 > Rwd. Namely,

Equation (2)

and, following Pringle (1981), we assume a Keplerian velocity Ω = ΩK for R ≥ R0.

With these assumptions and boundary conditions, one obtains the disk radial temperature profile

Equation (3)

where the angular velocity in the disk has been substituted with its Keplerian value. For convenience, this expression can be written as

Equation (4)

where x = R/R0. This model gives a maximum temperature (${T}_{\max }=0.488{T}_{0}$) at x = 1.36, and T = 0 K at R = R0 (x = 1).

The expression for the temperature of the modified disk model is the same as that for the standard disk model, except for the star radius Rwd, which has simply been replaced by the inner disk radius R0. For example, for an inner disk radius twice as large as the star radius (R0 = 2 Rwd), the maximum effective surface temperature in the disk (reached at R = 1.36 R0) has dropped to ∼65% of the value it has in the standard disk model (at R = 1.36 Rwd). In Figure 4 we show the temperature profile for a modified disk model for different values of the truncated radius R0. The mass of the WD is 0.8 M, and the mass accretion rate is 1 × 10−9 M yr−1. The radius R0 is given in units of the WD radius Rwd (which is itself obtained from the WD mass–radius relation; e.g., Wood 1995). From Figure 4, it is clearly apparent that truncating a standard disk model by solely removing the region R < 2 Rwd gives a disk model with a much higher temperature in the inner disk than generating a model starting at R0 = 2 Rwd. As the size of the BL is possibly not negligible and as the WD can easily increase its radius with increasing temperature, the standard disk model can overestimate the temperature in the inner disk by a large fraction.

Figure 4.

Figure 4. Temperature profile of the modified disk model shown as the radius R0 (in units of Rwd) of the inner boundary of the disk is increased from one stellar radius (the standard disk model) to two stellar radii. Here, the WD mass is 0.8 M and the mass accretion rate is $\dot{M}=1\times {10}^{-9}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$. As R0 increases, the disk temperature profile flattens and the disk luminosity decreases. For low Mwd and $\dot{M}$, and for large R0, the temperature becomes almost isothermal.

Standard image High-resolution image

Although our modified disk model is actually a standard disk model in which the radius of the star Rwd has been replaced by the inner radius of the disk R0 (where R0 > Rwd), we use in this paper the terms modified disk and standard disk to differentiate between the two.

3.4. The Luminosity of the Modified Disk

The disk luminosity Ldisk is given by integrating Φν (or $\sigma {T}_{\mathrm{eff}}^{4}$) over the surface of the disk

Equation (5)

In the standard disk model, the disk luminosity amounts to half the accretion energy; in the modified disk model, it is smaller than half the accretion energy. For example, for R0 = 2 Rwd, the disk luminosity is only half the disk luminosity expected for a standard non-truncated disk. The important point is that a truncated disk cannot be generated by simply truncating a standard disk model, but it has to be obtained as a standard disk model in which Rwd is replaced by R0, which produces a much lower disk temperature and luminosity. The idea of replacing Rwd by R0 was discussed briefly in the modeling of IX Vel (Long et al. 1994) but was never followed through. Similarly, the modeling of MV Lyr by Linnell et al. (2005) includes a truncated disk, but with the "usual" temperature profile of the standard disk model (Rwd was not replaced by R0; instead, the standard disk was just truncated).

3.5. The Boundary Layer Luminosity

The remaining fraction of the accretion energy is dissipated at the inner edge of the disk, where the disk meets the surface of the star (r = Rwd), in the so-called BL. The fast (Keplerian) rotating (ΩK(R0)) matter adjusts itself to the slowly rotating stellar surface (Ωwd) and dissipates its remaining kinetic energy. The luminosity of the BL, LBL, in the standard disk model, is of the same order of magnitude as Lacc/2, although it is usually smaller as some of the energy goes into spinning up the WD (Kluźniak 1987). This is true for optically thick BLs.

X-ray observations of disk-dominated CVs (e.g., van Teeseling et al. 1996; Balman et al. 2014) show that the BL is optically thin and that the X-ray luminosity is much smaller than the disk luminosity; the BL is underluminous. For the optically thin BL, one can only refer to the work of Narayan & Popham (1993) and Popham (1999), and to observations.

3.6. The Fate of the Remaining Accretion Energy

If the disk emits less than half the accretion energy, and the BL luminosity (in the X-ray) is itself only a fraction of the disk luminosity, one might ask, where does the remaining energy go? When an accretion flow cannot cool efficiently, due to being optically thin, the energy is advected with the flow into the outer layer of the star (Abramowicz et al. 1995). The advected BL energy will then increase the WD temperature and will likely drive an outflow as in ADAFs (Narayan & Yi 1995). This is supported by three observational facts. (i) The WD temperature of NLs is rather elevated when compared to other CVs with the same period. (ii) Many NL systems exhibit hot, fast wind outflows seen in the P Cygni line structure and blueshifted absorption. These winds originate in the inner disk/BL region but their source of energy has not yet been identified. Some CV systems even have visible ejected nebular material (e.g., BZ Cam, Ellis et al. 1984). (iii) Recent X-ray observations of NLs (Balman et al. 2014) suggest that high-state NLs may have optically thin BLs merged with ADAF-like flows and/or hot coronae. Balman et al. (2014) showed that the observed discrepancy between the X-ray/BL and disk luminosity can be accounted for if the BL energy is heating the WD through advection (Twd ∼ 50,000 K) and is also used to drive the wind outflows. The efficiency of the BL in radiating its energy is then greatly reduced (<0.01), consistent with the observed ratio of the X-ray to the disk UV-optical luminosities.

For this reason, we expect NLs to be modeled with the modified disk model and the addition of a hot WD, itself possibly with an inflated radius.

3.7. The Outer Disk Modeling

In CVs, the accreting WD is the more massive component and the Roche lobe radius is always larger than half the binary separation, i.e., >a/2. Due to the impact of the stream (from the first Lagrange point L1) on the rim of the disk and due to tidal interaction, the outer disk is expected to have a temperature larger than dictated by both the standard and modified disk models (Buat-Ménard et al. 2001). We assume here that the outer region of the disk has a temperature of 10,000 K to 12,000 K. We note that such a practice is not uncommon, and some authors (e.g., Linnell et al. 2005) have considered disk models with an elevated temperature (∼10–12,000 K) in the outer region of the disk.

4. UV Spectral Modeling

4.1. TLUSTY and SYNSPEC Codes

We use the FORTRAN suite of codes TLUSTY, SYNSPEC, ROTIN, and DISKSYN (Hubeny 1988; Hubeny et al. 1994; Hubeny & Lanz 1995) to generate synthetic spectra of stellar atmospheres and disks. The codes include the treatment of hydrogen quasi-molecular satellite lines (low temperature) and LTE and NLTE options (high temperature). SYNSPEC generates a continuum with absorption lines. In the present case, we do not generate emission lines. For disk spectra, we use solar abundances, and for stellar spectra, we vary the abundances as needed. An introductory guide, a reference manual, and an operational manual for TLUSTY and SYNSPEC have just been released and are available with the full details of the codes (Hubeny & Lanz 2017a, 2017b, 2017c).

4.2. WD Stellar Atmosphere Models

The code TLUSTY is first run to generate a one-dimensional (vertical) stellar atmosphere structure for a given surface gravity, effective temperature, and surface composition of the star. Computing a single model is an iterative process that needs to converge. In the present case, we treat hydrogen and helium explicitly, and treat nitrogen, carbon, and oxygen implicitly (Hubeny & Lanz 1995).

The code SYNSPEC is then run, using the output stellar atmosphere model from TLUSTY as an input, and it generates a synthetic stellar spectrum over a given (input) wavelength range (it has the capabilities to cover a spectral range from below 900 Å into the optical). The code SYNSPEC derives the detailed radiation and flux distribution of the continuum and lines, and generates the output spectrum (Hubeny & Lanz 1995). SYNSPEC has its own chemical abundances input to generate lines for the chosen species. For temperatures above 35,000 K, the approximate NLTE treatment of lines is turned on in SYNSPEC.

Rotational and instrumental broadening as well as limb darkening are then reproduced using the routine ROTIN. In this manner, we have already generated a grid of WD synthetic spectra covering a wide range of temperatures and gravities with solar composition (see http://synspecat.weebly.com/).

4.3. Disk Models

The disk spectra are generated by dividing the disk into rings, each with a given radius ri and temperature (T(ri)), and the density and effective vertical gravity obtained from the modified disk model for a given stellar mass, inner and outer disk radii, and mass accretion rate.

The code TLUSTY generates a one-dimensional vertical structure for each disk ring. The input parameters are the mass accretion rate, the mass of the accreting star, the radius of the star, and the radius of the ring. The radius of the star in our modified disk model is set to be the inner radius of the disk R0.

SYNSPEC is then used to create a spectrum for each ring, and the resulting ring spectra are integrated into a disk spectrum using the code DISKSYN, which includes the effects of (Keplerian) rotational broadening, inclination, and limb darkening.

For disk rings below 10,000 K, we use Kurucz stellar spectra of appropriate temperature and surface gravity.

In Figure 5, we present the synthetic spectra of two disks generated with the suite of codes: a standard disk model (in black) and a modified disk model (in red). Both models have a 0.55 M WD accreting at a rate of $\dot{M}={10}^{-9}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$, and are displayed for an inclination i = 40°. The standard disk has an inner radius at R0 = Rwd, while the modified disk has an inner radius at R0 = 2 Rwd. The slope of the continuum of the modified disk is almost flat.

Figure 5.

Figure 5. Synthetic spectra of two accretion disks shown in the ultraviolet range. Both disk models have a 0.55 M WD accreting at a rate of 10−9 M yr−1. The upper spectrum (in black) is a standard disk model with the inner disk radius at R0 = Rwd, and the lower spectrum (in red) is that of a modified disk with the inner radius at R0 = 2 Rwd. Because of the change in the disk temperature profile, the lower spectrum is characterized by a continuum slope that appears almost flat on this log–log scale. The continuum exhibits an "elbow" at λ ∼ 2400 Å, which is a real feature observed in the IUE spectrum of BZ Cam.

Standard image High-resolution image

5. Results

5.1. MV Lyrae

The Low State. The modeling of the FUSE spectrum of MV Lyr in the low state has been presented elsewhere (Godon et al. 2012) and gives identical results to the modeling of the combined FUSE + IUE spectrum presented here, which covers a much larger wavelength range. However, it is important to re-derive and confirm the distance to MV Lyr, its WD mass, radius, temperature, and rotational velocity, because these are important parameters serving as input for the modeling of the intermediate state. Namely, these parameters are used as constrains in the modeling of the combined FUSE + IUE spectrum of MV Lyr in the intermediate state.

While the FUSE spectrum presents many absorption lines commonly detected in the atmosphere of accreting WDs (see Godon et al. 2012 for details), the IUE (SWP+LWP) spectrum is noisier and presents only a continuum, in which the only identified absorption feature is the C iv 1550 line. The FUSE and IUE spectra match in their flux level and slope, in spite of the fact that they were obtained with different telescopes more than 20 years apart. This is possibly a sign that MV Lyr had reached a similar (if not identical) low state in which the WD completely dominates the FUV (Hoard et al. 2004).

Our fitting results yield (Godon et al. 2012) a WD temperature Twd = 44–47,000 K, with a projected stellar rotational velocity ${V}_{\mathrm{rot}}\sin i=150\mbox{--}250\,\mathrm{km}\,{{\rm{s}}}^{-1}$, chemical abundances Z = 0.1–1.0 Z, and stellar surface gravity $\mathrm{log}g=8.06\mbox{--}8.28$ (Mwd = 0.7–0.8 M with Rwd = 9010–7450 km). The resulting distance we obtained is ∼464 pc for a 0.8 M WD and ∼560 pc for a 0.7 M WD. In Figure 6 we present a 45,000 K WD fit as one of the solutions we obtained. The remarkable characteristic of this model fit is that, although it was originally generated to fit the FUSE spectrum alone (Godon et al. 2012), as shown in Figure 6, it actually also fits the IUE spectrum all way to the longest wavelengths, ∼3000 Å (see Figure 7; the region beyond 3000 Å is very noisy and has been discarded). From this fit, one obtains relatively accurate values for some of the important system parameters.

Figure 6.

Figure 6. Synthetic WD atmosphere spectrum fit (in black) to the combined FUSE + IUE spectrum of MV Lyr (in red) in the low state (the blue shows the contamination by terrestrial airglow). This figure shows the FUSE range, while the next figure shows the IUE range. The sharp emission lines are artifacts due to either airglow (marked with a plus sign inside a circle at the top of each panel) or indirect reflection of sunlight inside the telescope (such as, e.g., He ii 1168 Å). The WD model has a temperature of Twd = 45,000 K, solar composition, and a projected rotational velocity of ${V}_{\mathrm{rot}}\sin (i)=200\,\mathrm{km}\,{{\rm{s}}}^{-1}$.

Standard image High-resolution image
Figure 7.

Figure 7. Synthetic WD atmosphere spectrum fit (in black) to the combined FUSE + IUE spectrum (in red) of MV Lyr in the low state presented here showing only the IUE range. Note that the vertical scale is different from that in Figure 6 for clarity.

Standard image High-resolution image

These values of the parameters are consistent with the analysis of Hoard et al. (2004), who obtained Twd = 47,000 K with $\mathrm{log}g=8.25$ for a non-rotating WD, and Twd = 44,000 K with $\mathrm{log}g=8.22$ for a WD rotating at 200 km s−1 with elemental abundances of C = 0.5, N = 0.5, and Si = 0.2 (solar), and a derived distance of 505 ± 50 pc.

The Intermediate State. MV Lyr was caught in an intermediate state with IUE, covering that region of the spectrum where the continuum slope departs strongly from a standard disk model all the way to ∼3200 Å. In order to model the intermediate state spectrum, we use the parameters we derived from fitting the low state: we adopt a WD mass of Mwd = 0.73 M, with a temperature of ∼45,000 K, a radius Rwd = 8,500 km, and a distance of ∼500 pc. We use the system inclination i = 10° from the literature (see Table 1).

We first attempt to fit the combined IUE SWP+LWP spectrum with a standard disk model, where the only varying parameter is the mass accretion rate (since the inclination and WD mass are fixed). The inner disk radius is simply Rwd and the outer radius of the disk is placed at a distance a/3, where a is the binary separation. Using the system parameters, we have a/3 ∼ 30 Rwd. We also include the expected contribution of the 45,000 K WD. In order to fit the distance to the system (by scaling the theoretical spectrum to the observed one), we find a mass accretion rate of 8 × 10−10 M yr−1. However, this model does not fit the observed spectrum at all (see Figure 8)—the theoretical spectrum is too "blue" when compared to the observed spectrum, consistent with the statistical analysis of Puebla et al. (2007).

Figure 8.

Figure 8. Combined SWP + LWP IUE spectrum of MV Lyr in an intermediate state (solid red line; strong emission lines in blue) modeled with a hot WD (45,000 K) and standard accretion disk with $\dot{M}=8\times {10}^{-10}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ and i = 10°. The inner radius of the disk is at 1 Rwd and the outer radius is at 30 Rwd. The combined model WD+disk is shown with the solid black line, the dotted line in the lower part of the panel shows the WD contribution, and the dashed line shows the contribution from the disk. The model cannot match the continuum flux level—it is far too "blue."

Standard image High-resolution image

Next, we fit the observed spectrum with modified standard disk models. For each given mass accretion rate, we built disks with an increasingly larger inner radius, R0/Rwd = 1.0, 1.2, 1.5, 1.7, 2.0, 2.5, and 3.0. We kept the outer disk radius at 30 Rwd and set the outer disk at a temperature of 10,000, 11,000, or 12,000 K. The width of this isothermal outer disk region was also varied from 1 Rwd to 10 Rwd (i.e., to a maximum of one-third of the disk radial extent). The increase in the inner disk radius produces a decrease in the flux that is more pronounced in the shorter wavelength region of the spectrum, while the setting of the outer isothermal disk region produces a relative increase in the flux in the longer wavelength region. We find that a best fit is obtained for a mass accretion rate of 2.4 × 10−9 M yr−1 with an inner hole of size 2.0 Rwd and an outer disk, from 20 Rwd to 30 Rwd, set with a temperature of 10,000 K. This model is presented in Figure 9 and also includes the contribution of a 45,000 K WD. The strong emission lines are shown in blue; there are more emission lines on the edges of the Lyα profile which we are not modeling, as they form in an optically thin medium. The fit to the continuum is much improved, when compared to the standard disk model.

Figure 9.

Figure 9. The same IUE spectrum (in red) as in Figure 8 is modeled here with a hot WD (dotted line) and our modified disk model (dashed line); the combined model is the solid black line. The mass accretion rate is $\dot{M}=2.4\,\times {10}^{-9}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ and i = 10°. The inner edge of the disk is placed at R0 = 2.0 Rwd. The outer region (20 Rwd to 30 Rwd) of the disk has been set at a temperature of 10,000 K. The strong emission lines are shown in blue; there are more emission lines on the edges of the Lyα profile which we are not modeling, as they form in an optically thin medium.

Standard image High-resolution image

5.2. BZ Camelopardalis

BZ Cam has a FUSE spectrum and an IUE spectrum that do not have exactly the same flux level; in addition, the IUE spectrum is consistent with a rather cold component, while the FUSE spectrum seems to be consistent with a hotter component. Consequently, we first model these two spectra separately, and next we try to model the two spectra in a self-consistent manner.

Modeling the FUSE Spectrum. The FUSE spectrum shows the presence of a hot component contributing flux all the way to the shortest wavelengths of FUSE. We try a standard disk model, assuming Mwd = 0.8 M, and we find that the mass accretion rate has to be as large as 10−8 M yr−1 to produce the observed flux in the short wavelengths of FUSE. This model, however, gives a distance of more than 2 kpc, i.e., more than twice the accepted distance to the system. As we decrease the WD mass to 0.55 M, the distance decreases slightly to 1.8 kpc, and for a 0.35 M WD, the distance is still too large; however, the flux in the shorter wavelengths of the theoretical spectrum is now too low. For a slightly lower mass accretion rate of 10−8.5 M yr−1, only the larger WD model with Mwd = 0.8 M provides enough flux in the short wavelengths, but here, too, the distance is too large and reaches 1.4 kpc.

These results show that the single-disk model does not provide a satisfactory fit. Since our modified disk model provides colder disks, we do not try to use such a modeling for the FUSE spectrum, as it would not provide enough flux in the shorter wavelengths.

Instead, we try single-WD models. The best-fit WD model has Mwd = 0.4 M, Twd = 45,000 K, solar composition, and ${V}_{\mathrm{rot}}\sin (i)=200\,\mathrm{km}\,{{\rm{s}}}^{-1}$, and the distance obtained from the fit is 791 pc, when the WD radius is set to Rwd = 18,540 km. This model is shown in Figure 10 and includes a basic modeling of the ISM molecular hydrogen absorption lines (see Godon et al. 2007 for details). In the short wavelengths (∼960 Å upper panel), the model has too much flux, while in the longer wavelengths (lower panel), the model has too little flux. A 40,000 K WD gives a distance of 654 pc while a 50,000 K WD gives a distance of 933 pc. We note that at a temperature of about 50,000 K, the radius of the 0.4 M WD model (∼2 × 109 cm) is about twice that of the zero-temperature model (∼1 × 109 cm).

Figure 10.

Figure 10. FUSE spectrum of BZ Cam (in red) fitted with a hot WD to agree with the distance to the system as well as with the flux in the short wavelengths. The WD model (including the ISM absorption model) is shown with the solid black line; it has Mwd = 0.4 M, Twd = 45,000 K, solar composition, and ${V}_{\mathrm{rot}}\sin (i)=200\,\mathrm{km}\,{{\rm{s}}}^{-1}$ (the same WD model without the ISM absorption model is shown with the dotted black line). The distance obtained from the fit is 791 pc. The ISM absorption features as well as the airglow emission lines are marked in blue. We have marked some of the prominent absorption lines that are usually observed in an accreting WD with a high temperature; these lines are not observed here, a sign that the observed spectrum might have some emission lines on top of the absorption lines or that the photosphere has a low metallicity.

Standard image High-resolution image

Next, we increase the WD mass to 0.7 M, and accordingly decrease the radius to ∼0.9 × 109 cm. We find that the scaled distance from the fit becomes too short for the same temperature, and higher temperature models (T > 50,000 K) do not provide the best fit. The fit of the 45,000 K 0.7 M WD is as good as the 0.4 M fit with the same temperature, and indicates that a 0.7 M WD with a radius of ∼2 × 109 cm provides a valid solution, indicating that the radius of the WD might be inflated.

We also try to add a standard disk model to the WD disk model, but we find that the disk model deteriorates the fit and increases the distance well above 1 kpc.

To summarize, the FUSE spectrum of BZ Cam is best fitted with a single WD with a temperature of 45,000 K ± 5000 K, and a mass of 0.4–0.7 M (from the literature) as long as the WD radius is ∼2 × 109 cm (to fit the known distance to the system). Since this radius really corresponds to a 0.4 M WD at T ∼ 50,000 K (see Section 3.2), this is the WD mass we adopt as the best solution.

Modeling the IUE Spectrum. The IUE spectrum of BZ Cam was modeled with a 12,500 K Kurucz model with $\mathrm{log}g=2.5$ by Prinja et al. (2000), indicating that a standard disk model does not fit the spectrum. Therefore, we do not attempt a standard disk model but instead we use a modified disk model.

We assume a WD mass of 0.4 M and a low inclination (i = 10° and i = 20°), and we vary the mass accretion rate from 10−10 M yr−1 to 10−8 M yr−1. We assume a WD radius Rwd = 1.15 × 109 cm, which actually corresponds to a 10,000 K 0.4 M WD (see Section 3.2), and we vary the inner disk radius R0 from 1 Rwd to 3 Rwd. We keep the outer part of the disk at an isothermal temperature of 10,000, 11,000, or 12,000 K.

We find that the best modified disk model fit has a mass accretion rate $\dot{M}={10}^{-8.5}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$, i = 20°, and a disk inner radius R0 = 2 Rwd (=23,000 km), and the outer half of the disk is set to an isothermal temperature T = 10,000 K. This outer half of the disk has a surface area three times larger than the inner half of the disk, making this modified disk model resemble an isothermal disk model. However, in order for this modified disk model to agree with the known distance (830 ± 160 pc), the outer disk radius had to be set at ∼0.5 ± 0.1a, corresponding to Rd ≈ 22 Rwd. Although such a disk radius is about twice the accepted radius (0.3a) in a CV binary, it is still within the limits of the Roche lobe. Choosing an outer disk radius smaller than this value decreases the scaled distances. The fit is shown in Figure 11.

Figure 11.

Figure 11. IUE spectrum of BZ Cam (in red, with masked emission lines in blue) modeled with a modified disk model (in black). The WD mass is 0.4 M, the disk has a mass accretion rate of $\dot{M}={10}^{-8.5}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$, an inclination i = 20°, and the outer half of the disk has been set to an isothermal temperature T = 10,000 K. The inner disk radius is R0 = 2 Rwd (zero temperature WD), and the outer disk radius is set to 0.5 times the binary separation (within the limits of the Roche lobe) in order to match the distance to the system. Note that within the Lyα region, the model seems to be deficient in flux.

Standard image High-resolution image

Since this modified disk model is close to an isothermal disk, we next try an isothermal disk model, with the same inner and outer radii. The isothermal disk model gives very similar results when its temperature is set to 11,000 K. In this model, shown in Figure 12, the deficiency in flux in the shorter wavelength range (the Lyα region) is a little more pronounced than in the modified disk model. In the isothermal disk model, the exact value of the inner disk radius, i.e., whether it is 1 or 2 Rwd, has very little effect on the results as it only changes the surface area of the disk by a tiny fraction. Also, the value of the WD mass is not taken into account. However, when integrating the luminosity over the surface of the isothermal disk, the resulting mass accretion rate corresponds to 3.7 × 10−9 M yr−1 for a 0.4 M mass WD and 2.7 × 10−9 M yr−1 for a 0.7 M mass WD.

Figure 12.

Figure 12. Same as in the previous figure, but now the spectrum is modeled with an isothermal disk only. The disk has a temperature of 11,000 K. Here, too, in order to scale the model to the known distance, the radius of the disk has to be 0.5a, departing the most from a standard disk model. The deficiency in flux in the Lyα region is even more pronounced than in the modified disk model.

Standard image High-resolution image

Self-Consistent Modeling of the FUSE and IUE Spectra. Since the FUSE and IUE spectra were obtained years apart, in slightly different brightness states, and with different telescopes/instruments, we do not expect them to match and we do not attempt to combine them. However, when modeling each spectrum separately (IUE or FUSE), we must take into consideration the results obtained from the modeling of the other spectrum (FUSE or IUE, respectively). Namely, we now model the IUE spectrum with a cold disk and a hot WD model, similar but not identical, to the WD model obtained from the modeling of the FUSE spectrum, and we model the FUSE spectrum with a hot WD and a cold disk model, similar but not identical, to the cold disk model obtained from the modeling of the IUE spectrum.

For the disk model, we chose the isothermal disk model as it is easier to parametrize, and gives results similar to the modified disk model. Namely, the WD mass does not enter the disk model, and the WD radius, taken as the inner radius of the disk, has a minimal effect on the results. For the WD model, we chose a 0.4 M WD with a radius increasing with temperature. As we have mentioned earlier, only the WD radius and temperature affect the results.

For the FUSE spectrum, we find that the 45,000 K WD model, with a radius of 18,540 km, is improved significantly with the addition of a cold isothermal disk with T = 11,000 K. The cold disk contributes mainly in the longer wavelengths of the FUSE spectrum and helps improve the fit in that region (see Figure 13). The disk has an outer radius of 0.5 times the binary separation, and the distance obtained from the fit is 854 pc. The mass accretion rate obtained by integrating the luminosity of the disk is 3.7 × 10−9 M yr−1, and the heated WD luminosity is of the same order.

Figure 13.

Figure 13. FUSE spectrum of BZ Cam is modeled with a 0.4 M 45,000 K WD (dotted line) plus a 11,000 K isothermal disk (dashed line) of radius 0.5a. The combined model (solid black line) includes a basic ISM absorption model. Note that the isothermal disk, contributing only 7% of the flux in the FUSE range, improves the overall fit and more particularly in the longer wavelength. The distance obtained from the fit is 854 pc. The region of C iii (1160–1180) shows a possible P Cygni profile, although it is also a region exhibiting a detector artifact known as the "worm."

Standard image High-resolution image

The fit is improved further by increasing the WD temperature to 50,000 K and the isothermal disk temperature to 12,000 K; however, the distance increases to 1137 pc. This model is shown in Figure 14. The disk-integrated luminosity gives a mass accretion rate $\dot{M}=5\times {10}^{-9}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$, and here, too, the WD luminosity is of the same order.

Figure 14.

Figure 14. FUSE spectrum of BZ Cam is modeled with a 0.4 M 50,000 K WD (dotted line) plus a 12,000 K isothermal disk (dashed line). The combined model (solid black line) includes a basic ISM absorption model. The isothermal disk now contributes 25% to the FUSE range of the flux, and further improves the fit in the longer wavelength. The distance obtained from the fit is 1137 pc.

Standard image High-resolution image

For the IUE spectrum, we find that for the model to fit the spectrum, the temperature of the disk has to be decreased as the temperature of the WD increases. The best fit (Figure 15) is obtained for a disk with a temperature of 10,000 K combined with a 40,000 K WD. This model also gives a distance of the order of 650 pc for a an outer disk radius ∼0.6a. The integrated luminosity of the disk gives a mass accretion rate of 2.5 × 10−9 M yr−1, while the WD luminosity is about 10% less than that of the disk.

Figure 15.

Figure 15. IUE spectrum of BZ Cam is modeled (solid black line) with an isothermal disk plus a hot WD. The disk (dashed black line) has a temperature of 10,000 K and the WD (dotted black line) has a temperature of 40,000 K. Note that the WD contribution greatly improves the fit in the Lyα region.

Standard image High-resolution image

The results of the analysis of the UV spectra of BZ Cam can be summarized as follows: a 0.4 M WD with a temperature of 45,000 ± 5000 K, and a disk with a temperature of 11,000 ± 1000 K, extending to 0.5 ± 0.1 times the binary separation. Taking only the disk luminosity into consideration gives a mass accretion rate of the order of 2.5 × 10−9 M to 5 × 10−9 M yr−1, and twice as large when also taking the WD luminosity contribution (assuming that the WD is heated due to the advection of energy from the inner disk). The WD contributes mainly to the FUSE spectral range while the cold disk contributes mainly to the IUE spectral range. The WD mass could be larger, i.e., 0.7 M, but its radius would have to be of the order of 20,000 km, i.e., twice as large as expected for such a mass (again, see Section 3.2 for details).

5.3. V592 Cas

We assumed a 0.8 M WD mass with a radius Rwd = 7000 km, which is consistent with a temperature of a WD at about 16,000 K. The inner radius of the disk, R0, was placed at 1.0 Rwd, 1.2 Rwd, 1.5 Rwd, and 2.0 Rwd. The outer radius was placed at 36 Rwd ≈ a/3, where a is the binary separation of V592 Cas derived from Table 1. The outer region of the disk was allowed to take different temperatures: either the actual disk temperature, or 10,000 K, 11,000 K, or 12,000 K. The size of that outer region was also varied. Some preliminary modelings were carried out using a grid of standard disk models to obtain an order of estimate of the mass accretion rate, which gave $\dot{M}\sim {10}^{-8}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$. Modified disk models were then computed within the vicinity of this mass accretion rate assuming an inclination angle of 30°.

The models were chosen to provide the best fit and the correct distance (≈350 pc) to the source. The best fit is for a modified disk with a mass accretion rate of $\dot{M}={10}^{-8.2}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$ (∼6.3 × 10−9 M yr−1), an inner disk radius placed at R0 = 1.2 Rwd, an outer disk radius at 30 R0, and an isothermal outer disk between r = 20 R0 and r = 30R0 with a temperature T = 12,000 K. A 45,000 K WD contribution was added to the model for completeness (Hoard et al. 2009) but contributed only a few percent to the flux and did not affect the model. Consequently, the temperature of the WD could not be assessed from our modeling. The fit is presented in Figure 16 in the FUSE spectral range and in Figure 17 in the IUE SWP+LWP spectral range.

Figure 16.

Figure 16. FUSE spectrum of V592 Cas (in red) has been fitted with a combined WD + modified disk model (solid black line). The WD model, dotted line, has a mass of 0.8 M with a temperature of 45,000 K, but contributes very little to the combined model and was added only for completeness. The modified disk model, dashed line, has a mass accretion rate of 6.3 × 10−9 M yr−1 ($\mathrm{log}(\dot{M})=-8.2$), with an inclination of 30°. The inner radius of the disk is set at R0 = 1.2 Rwd (8400 km), and the outer radius is set at 30R0. The outer region of the disk between 20 R0 and 30 R0 was set to a temperature of 12,000 K. This model gives a distance of 342 pc. The regions affected by ISM absorption features and sharp emission lines (due to daylight) are omitted (masked) for the fitting and have been colored in blue. The location of absorption lines observed in the FUSE spectra of some CVs have been marked, but are not especially observed here.

Standard image High-resolution image
Figure 17.

Figure 17. The same disk model shown in Figure 16 is now fitted to the IUE spectrum of V592 Cas. For clarity, the vertical scale is different from that in Figure 16.

Standard image High-resolution image

For a 16,000 K WD with a radius of 7000 km, the modified disk model, with an 8400 km inner radius, starts only at 1.2 Rwd, and for a 45,000 K WD with a radius of 7450 km, the disk starts at 1.13 Rwd. In either case, the size of the BL would be relatively narrow (∼0.1–0.2 Rwd), and this modified disk model is very similar to a standard disk model in its inner region and differs mainly in the assumption of an outer isothermal region. V592 Cas is not as extreme a case as BZ Cam or MV Lyr.

6. Discussion

In the present UV spectral analysis, we find that the data for the UX UMa system V592 Cas are consistent with a slightly modified disk model (very similar to a standard disk) with an inner disk radius at R0 = 1.2 Rwd, $\dot{M}={10}^{-8.2}\,{M}_{\odot }\,{\mathrm{yr}}^{-1}$, and an outer 12,000 K isothermal disk. The small surface of the ∼0.8 M WD is overshone by the disk, as its contribution, even at a temperature Twd = 45,000 K, is negligible. The data for our two VY Scl systems, MV Lyr and BZ Cam, point to a different scenario. For MV Lyr, we find a significantly modified disk model with an inner radius as large as 2 Rwd, and an outer 10,000 K isothermal disk. The contribution of the 45,000 K 0.8 M WD is significant. BZ Cam is an even more extreme case in that it has a slope shallower than MV Lyr and V592 Cas, and its spectrum is consistent with a large isothermal disk with T ∼ 10,000–11,000 K extending to the limit of the Roche lobe (∼0.5a). A modified disk model with an inner radius of 2 Rwd (40,000 km) and an extended outer isothermal region provides the same solution. Its 0.4 M WD with a 20,000 km radius and a temperature of 45,000 K contributes most of the flux in the FUSE range. For both MV Lyr and BZ Cam, the mass accretion rate is about twice as small as for V592 Cas.

The hot WD scenario in the two VY Scl NL systems, MV Lyr and BZ Cam, with an inner disk hole ∼2 Rwd, is consistent with our recent Swift X-ray data analysis, which reveals that the optically thin hard X-ray emitting BL is underluminous, a sign that it cannot radiate its energy to cool efficiently, and, instead, the BL energy is advected onto the outer layer of the white dwarf, heating it, and possibly inflating its radius and launching an outflow. This is the self-consistent global picture we draw from our present UV spectral analysis and our previous Swift X-ray analysis. In this picture, the optically thin BL extends to r ∼ 2 Rwd and heats the WD through the advection of energy. For V592 Cas, the inner hole is much smaller extending only to ∼1.2 Rwd.

To further check this scenario, we select a larger and self-consistent set of X-ray data for CVs to compare to the archival IUE data. If the above scenario is correct, we expect the systems with UV spectra departing from the standard disk model to have a harder X-ray emission than those systems with UV spectra agreeing with the standard disk model. For this purpose, we turn to van Teeseling et al. (1996), who computed the X-ray hardness ratio from ROSAT observations of non-magnetic cataclysmic variables, including five VY Scl NLs and six UX UMa NLs observed in a state of high accretion rate, and a number of DN systems observed in their quiescent state. ROSAT has an energy range (0.1–2.4 keV) that is relatively soft, considering that some IPs can be as hard as 20–300 keV. In spite of the fact that the DNe were observed in quiescence, van Teeseling et al. (1996) found that the NL VY systems had an X-ray hardness ratio larger than the NL UX UMa and quiescent DN SU UMa systems. DN systems in outburst are known to exhibit hard X-rays, too (only a few systems exhibit EUV/soft X-rays). During outburst, however, the X-ray temperature is lower than in quiescence (Balman 2015). Consequently, the hardness ratio of the quiescent DNe computed in van Teeseling et al. (1996) is an upper limit to the hardness ratio of these same DNe in outburst. So, the data in van Teeseling et al. (1996) give us the hardness ratio of disk-dominated UX and VY NLs and an upper limit to the hardness ratio of DN in outburst.

As noted by la Dous (1991), disk-dominated NL systems have a shallower UV continuum slope than the DN in outburst and cannot be modeled with standard disk models. To quantitatively assess the departure from the standard disk model in the UV, we measure the slope of the continuum of the IUE spectra, the flux in $\langle \mathrm{erg}\,{{\rm{s}}}^{-1}\,{\mathrm{cm}}^{-2}\,{\mathring{\rm A} }^{-1}\rangle $ as a function of wavelength in $\langle \mathring{\rm A} \rangle $ on a log–log scale, for the sample of non-magnetic disk systems in van Teeseling et al. (1996) (for which IUE data exist). Using the LWP segment and the longer wavelength region of the SWP spectrum past the 1550 Å carbon line, we obtain a UV spectral coverage from ∼1600 Å to 3200 Å. We deredden all the spectra based on the E(BV) values from the literature (Bruch & Engel 1994) or directly from the 2175 Å absorption feature using the extinction curve from Fitzpatrick & Massa (2007) with R = 3.1. We then measure the slope of the UV continuum directly from the log–log graphs that we generate for these systems (the graphs are similar in appearance to Figures 13). We have here a limited sample of systems as we decided to take only those systems for which the hardness ratio was measured from the same telescope in the same manner, and van Teeseling et al. (1996) provides such a self-consistent sample.

In Table 3, we list the ROSAT X-ray hardness ratio and IUE UV continuum slope for the nine DN systems, six UX NL systems, and four VY NL systems together with the novalike V795 Her and the two old novae V603 Aql and RR Pic. V592 Cas was not included in van Teeseling et al. (1996) and has no hardness ratio listed. MV Lyr is listed twice, once for its intermediate state and once for its high state that exhibits a slightly different UV continuum slopes. At the bottom of the table, we also list the UV continuum slope of some standard disk models computed with TLUSTY and SYNSPEC (e.g., Wade & Hubeny 1998) for comparison. These standard disk models have a slope that flattens as the mass accretion rate decreases.

Table 3.  System Parameters

System Type UV Hardness
Name   Slope Ratio
WZ Sge DN/SU −2.4 0.21
SW UMa DN/SU −2.5 0.00
SU UMa DN/SU −1.9 0.24
Z Cha DN/SU −1.7 0.53
WX Hyi DN/SU −2.0 0.46
T Leo DN/SU −2.3 0.22
VW Hyi DN/SU −2.4 0.05
AB Dra DN/UG −2.1 0.84
WW Cet DN/ZC −1.75 0.50
RW Tri NL/UX −1.8 1.39
AC Cnc NL/UX −1.7 −0.26
UX UMa NL/UX −2.0 0.05
IX Vel NL/UX −2.2 0.10
V3885 Sgr NL/UX −2.1 0.38
V592 Cas NL/UX −2.4 ...
BZ Cam NL/VY −1.15 0.95
KR Aur NL/VY −1.55 0.96
MV Lyr i NL/VY −1.4 0.83
MV Lyr h NL/VY −1.6 0.83
TT Ari NL/VY −1.9 0.83
V795 Her NL/SH −1.9 0.00
V603 Aql CN/SH −2.35 0.83
RR Pic CN/SW −1.95 −0.49
 
System $\dot{M}$ UV hardness
  $\langle {M}_{\odot }\,{\mathrm{yr}}^{-1}\rangle $ slope ratio
Standard Disk 10−9.5 −2.52 ...
Standard Disk 10−9.0 −2.89 ...
Standard Disk 10−8.5 −2.88 ...
Standard Disk 10−8.0 −2.94 ...

Note. The subtypes are as in Ritter & Kolb (2003), except for the novae which are denoted CN (classical nova). The UV continuum slope was measured directly from the dereddened IUE spectra and the ROSAT X-ray hardness ratio was taken from van Teeseling et al. (1996). Some standard disk models (Wade & Hubeny 1998) are also included for comparison.

Download table as:  ASCIITypeset image

In Figure 18, we plot the UV continuum slope of these systems against the ROSAT X-ray hardness ratio reported in van Teeseling et al. (1996). Even though the X-ray hardness ratio for DNe is an upper limit and all the DNe (green triangle) are expected to be located more to the left, had the hardness ratio been measured in outburst, there is a clear separation between the DN and VY Scl systems. Furthermore, for this limited sample, it appears that the VY Scl systems have a shallower UV slope with a harder X-ray spectrum, consistent with the scenario of an optically thin inner disk with a modified disk model. The DNe have a softer spectrum and a steeper UV slope consistent with either a standard disk model or a less modified disk model with a less extended optically thin inner disk (e.g., as we found for V592 Cas). There is an overall tendancy pointing to the fact that the more disk-dominated CVs depart from the standard disk model in the UV (shallower slope), the harder their X-ray emission, respectively, starting with the DN systems and ending with the VY Scl systems.

Figure 18.

Figure 18. UV continuum slope of DNe in outburst and NLs in high state (disk-dominated CVs) plotted against the measured X-ray hardness ratio of quiescent DNe and NLs in high state for the systems listed in Table 3. The subtypes are as indicated in the upper-right panel. V603 Aql is here taken both as a CN and as an NL SH. V592 Cas was not included as it is not listed in van Teeseling et al. (1996). DNe in outburst have softer X-ray than in quiescence (Balman 2015), implying that the DNe (green triangles) would shift to the left if the hardness ratio had been measured during outburst. Even so, there is a clear separation between DNe and VY NLs. Overall, for disk-dominated CVs, the harder the X-ray, the shallower the slope of the UV continuum. The departure from the standard disk model in the UV correlates with the X-ray hardness of the system.

Standard image High-resolution image

Motivated by these findings and to further check the nature of VY Scl systems, we decided to measure the UV continuum slope for all of the available IUE archive of disk-dominated CVs. We found 105 CVs with IUE spectra taken when the disk dominated the system and for which a continuum slope could be measured. Reddening values E(BV) were taken from Bruch & Engel (1994) when available, or were directly assessed using the 2175 Å feature. The UV continuum slope for all of the systems was measured after the spectra were dereddened. For each system in Table 4, we list the resulting UV slope together with the Data ID, reddening, and orbital period. Some systems with only SWP spectra available were included (V794 Aql and CZ Ori); others were not included because the value of the slope could not be assessed (e.g., V380 Oph). Some systems did not reveal any obvious continuum from which a slope value could be derived, as these spectra were affected by noise, emission lines, or possible additional components (e.g., BH Lyn, V348 Pup, DW UMa). For GI Mon, the only two IUE spectra (SWP and LWP) gave different slopes: SWP gave −1.5 and LWP gave −1.1, and the value listed is their average.

Table 4.  UV Continuum Slope of Disk-dominated Cataclysmic Variables

System Type IUE IUE Period UV E(BV) i
Name Subtype Short Long $\langle \mathrm{days}\rangle $ Slope   $\langle \deg \rangle $
WZ Sge DN/SU SP03507 LR03086 0.056694 −2.4 0.00 77
SW UMa DN/SU SP27871 LP07754 0.056815 −2.5 0.00 45
WX Cet DN/SU SP36511 LP15730 0.05827 −2.3 0.00 ...
CC Scl NL/IP SP34427 LP14189 0.0584 −2.0 0.00 81
T Leo DN/SU SP33646 LP13312 0.05882 −2.3 0.00 65
CP Pup CN/SH SP27806 LP07782 0.061264 −2.0 0.20 ...
V1159 Ori DN/SU SP56781 LP31957 0.06218 −2.2 0.00 ...
V436 Cen DN/SU SP54246 LP30319 0.062501 −2.7 0.07 65
BC UMa DN/SU SP50665 LP28027 0.06261 −2.2 0.00 ...
EK TrA DN/SU SP09705 LR08446 0.06288 −2.25 0.03 58
TV Crv DN/SU SP41842 LP20597 0.0629 −2.6 0.10 ...
OY Car DN/SU SP25857 LP05906 0.063121 −2.2 0.05 83
VY Aqr DN/SU SP21719 LP02366 0.06309 −2.2 0.00 ...
ER UMa DN/SU SP40947 LP19846 0.06366 −2.05 0.00 ...
EX Hya NL/IP SP03858 LR03435 0.0682338 −1.6 0.00 78
IR Gem DN/SU SP38524 LP17696 0.0684 −2.15 0.00 ...
VZ Pyx DN/SU SP44128 LP22531 0.07332 −2.35 0.07 ...
AY Lyr DN/SU SP09342 LR08100 0.0737 −2.4 0.00 ...
VW Hyi DN/SU SP48680 LP26405 0.074271 −2.4 0.00 ...
Z Cha DN/SU SP30677 LP10466 0.074499 −1.7 0.00 82
WX Hyi DN/SU SP23952 LP04221 0.074813 −2.0 0.01 40
T Pyx CN SP33034 LP12791 0.076223 −2.6 0.35 ...
SU UMa DN/SU SP34824 LP14532 0.07635 −1.9 0.00 ...
YZ Cnc DN/SU SP03727 LR03308 0.0868 −1.85 0.00 38
V795 Her NL/SH SP22901 LP03269 0.1082648 −1.9 0.00 ...
V592 Cas NL/UX Table 2 Table 2 0.115063 −2.4 0.22 28
TU Men DN/SU SP10665 LR09374 0.1172 −2.6 0.08 ...
V442 Oph NL/VY SP14731 LR11298 0.12433 −2.2 0.20 ...
LQ Peg NL/VY SP17367 LR13618 0.124747 −2.2 0.00 ...
AH Men NL/SW SP43037 LP21666 0.12721 −1.85 0.12 ...
DN Gem CN SP38213 LP17402 0.127844 −1.7 0.10 ...
KQ Mon NL/UX SP15384 LR11915 0.128 −1.8 0.04 ...
MV Lyr NL/VY SP07296 LR06288 0.132335 −1.5 0.00 12
SW Sex NL/SW SP21533 LP02262 0.134938 −1.0 0.00 >75
HL Aqr NL/SW SP23325 LP03647 0.13557 −2.15 0.05 ...
TT Ari NL/VY SP42147 LP20922 0.137550 −1.9 0.03 ...
V603 Aql CN/SH SP05758 LR04994 0.138201 −2.35 0.08 13
WX Ari NL/SW SP55953 LP31494 0.139351 −1.0 0.00 72
V1315 Aql NL/SW SP27096 LP07085 0.13969 −1.1 0.10 82
V1223 Sgr NL/IP SP13365 LR10022 0.140244 −2.15 0.15 24
V2400 Oph NL/IP SP47007 LP24971 0.142 −2.4 0.40 ...
LN UMa NL/SW SP40948 LP19847 0.1444 −1.9 0.15 ...
V751 Cyg NL/VY SP25774 LP05819 0.144464 −1.1 0.20 ...
RR Pic CN/SW SP06625 LR05687 0.145025 −1.95 0.00 65
PX And NL/SW SP39273 LP18414 0.146353 −1. 0.05 ...
V533 Her CN/IP? SP44805 LP23205 0.147 −1.4 0.03 ...
V425 Cas NL/VY SP15267 LR11783 0.1496 −1.5 0.10 25
AO Psc NL/IP SP09706 LR08447 0.1496252 −2.2 0.10 ...
AB Dra DN/UG SP17619 LR13886 0.1520 −2.1 0.10 ...
V794 Aql NL/VY SP50754 ... 0.1533 −2.1 0.10 ...
BP Lyn NL/SW SP32940 LP12690 0.152812 −2.1 0.00 ...
BZ Cam NL/VY Table 2 Table 2 0.15353 −1.15 0.05 ...
LX Ser VY/SW SP08070 LR07037 0.158432 −1.0 0.00 90
CY Lyr DN/UG SP21030 LR16779 0.1591 −2.5 0.15 ...
CM Del NL/UX SP14707 LR11282 0.162 −1.9 0.09 73
KT Per DN/UG SP17712 LR13968 0.16265777 −2.7 0.20 ...
KR Aur NL/VY SP14734 LR11299 0.16280 −1.55 0.05 38
AR And DN/UG SP18877 LR14885 0.16302 −2.4 0.02 ...
CN Ori DN/UG SP32593 LP12364 0.163199 −2.15 0.00 67
X Leo DN/UG SP15951 LR12282 0.1646 −2.4 0.00 ...
VW Vul DN/ZC SP18875 LR14883 0.16870 −2.1 0.15 ...
UZ Ser DN/ZC SP15078 LR11605 0.173 −3.0 0.25 ...
V405 Aur NL/IP SP56765 LP31949 0.17345 −0.5 0.15 <5
LS Peg NL/IP SP39782 LP18948 0.174774 −1.7 0.04 30
WW Cet DN/ZC SP10664 LR09373 0.1758 −1.75 0.03 54
U Gem DN/UG SP10327 LR08987 0.17690619 −2.5 0.04 70
GI Mon CN/IP SP38419 LP17709 0.1802 −1.3 0.10 ...
SS Aur DN/UG SP18610 LR14679 0.1828 −2.4 0.08 38
DQ Her NL/IP SP09201 LR07981 0.19362 −0.75 0.08 90
IX Vel NL/UX SP18578 LR14657 0.193927 −2.2 0.01 57
UX UMa NL/UX SP46816 LP24773 0.196671 −2.0 0.02 70
FO Aqr CN/IP SP15260 LR11774 0.202 −0.5 0.00 ...
V825 Her NL/UX? SP17353 LR13601 0.206 −2.0 0.00 ...
HX Peg DN/ZC SP37458 LP16654 0.2008 −2.0 0.00 ...
V3885 Sgr DN/UX SP25655 LP05709 0.207161 −2.1 0.02 ...
RX And DN/ZC SP17673 LR13942 0.209893 −2.1 0.02 51
PQ Gem NL/IP SP46993 LP24965 0.211636 −0.5 0.00 low
HR Del CN SP02122 LR02606 0.214165 −2.6 0.17 40
CZ Ori DN/UG SP07385 ... 0.214667 −2.35 0.00 ...
TV Col NL/IP SP08672 LR07425 0.2286 −2.2 0.02 70
RW Tri NL/UX SP07915 LR06895 0.231883 −1.8 0.10 71
VY Scl NL/VY SP32594 LP12365 0.2323 −1.9 0.06 30
TX Col NL/IP SP00502 LP11930 0.23825 −1.55 0.05 ...
RW Sex NL/UX SP26313 LP06299 0.24507 −2.0 0.02 34
AH Her DN/ZC SP17671 LR13923 0.258116 −2.2 0.03 46
TZ Per DN/ZC SP17643 LR13907 0.262906 −2.45 0.27 ...
TW Pic NL/VY SP27087 LR17818 0.265 −2.3 0.05 ...
TT Crt DN/UG SP47801 LP25680 0.2683522 −2.3 0.00 ...
SS Cyg DN/UG SP48441 LP26199 0.27513 −2.5 0.07 51
Z Cam DN/ZC SP26717 LP15337 0.2898406 −2.0 0.02 57
EM Cyg DN/ZC SP07297 LR06289 0.290909 −2.1 0.03 67
AC Cnc NL/UX SP18735 LR14787 0.300478 −1.7 0.00 76
V363 Aur NL/UX SP25335 LP05437 0.321242 −1.9 0.15 70
BT Mon CN/SW SP25446 LP05516 0.333814 −2.2 0.24 82
RZ Gru NL/UX SP15385 LR11916 0.360 −1.75 0.03 ...
RU Peg DN/UG SP15079 LR11606 0.3746 −2.1 0.00 43
SY Cnc DN/ZC SP08082 LR07051 0.38 −2.3 0.00 ...
KO Vel NL/IP SP16034 LR12341 0.422 −1.4 0.05 ...
DX And DN/UG SP37687 LP16843 0.440502 −1.9 0.10 45
QU Car NL/UX SP41926 LP20699 0.45 −2.3 0.11 <60
DI Lac CN SP29325 LP09208 0.543773 −2.0 0.26 ...
V841 Oph CN SP07950 LR06925 0.601304 −2.6 0.40 ...
BV Cen DN/UG SP24867 LP05162 0.610108 −1.5 0.10 53
GK Per NL/IP SP20653 LR16563 1.996803 −0.7 0.30 <73

Note. We List here 105 cataclysmic variables for which an IUE spectrum was taken when the system was dominated by emission from the disk. The CV subtypes are according to Ritter & Kolb (2003), except for the novae, which we denote as systems having experienced a classical nova explosion (CN). The reddening values E(BV) were taken from Bruch & Engel (1994), or directly assessed based on the 2175 Å absorption feature. The slope of the UV continuum was assessed for each system after the spectra were dereddened using the relation from Fitzpatrick & Massa (2007) with R = 3.1. Periods and inclinations are also taken from the Ritter & Kolb catalog (Ritter & Kolb 2003).

Download table as:  ASCIITypeset images: 1 2

We plot the UV continuum slope against the orbital period for all of the CVs in Table 4 in two figures: we show in Figure 19 the systems above the period gap, and in Figure 20 we show the systems below the period gap (for clarity, the x-axis has been stretched in Figure 20).

Figure 19.

Figure 19. Slope of the (IUE) UV continuum shown as a function of the orbital period (in days, on a log scale) for the disk-dominated CV systems listed in Table 4. For clarity, only the region above the period gap is shown extending to P ∼ 0.6 days. The subtypes are shown as indicated by the different symbols and colors in the panel on the upper right, where the average value of the slope is also listed for each subtype. Overall, the DNe SU, UG, and ZC subtypes (all in green) all have a similar average slope of about ∼−2.25, significantly steeper than the NLs (shown in blue). The UX UMa NLs have an average slope of about −2, the VY Scl NLs have an average slope of −1.70, and the eclipsing SW Sex NLs seem to cluster around either −1.0 or −2.0 with an average value of −1.63. The IPs have been included (in red) for comparison and have an average slope of only 1.43 (with a very broad scattering), not inconsistent with magnetically truncated inner disks. For the DNe UG and ZC (as a whole), the steepest (most negative) slope value reached at a given orbital period shows an increase (less negative) with orbital period. The SU UMa systems below the gap, shown in detail in Figure 20, exhibit a similar pattern. We included the novae (CN) classified as IP (CN/IP) with the IPs, thereby leaving only seven CN systems.

Standard image High-resolution image
Figure 20.

Figure 20. The UV slope against the orbital period is shown, as in Figure 19, but for the SU UMa systems below the period gap. The horizontal (log) axis has been stretched for clarity. As for the the U Gem and Z Cam systems, the steepest slope value reached for the SU UMa's at a given orbital period shows an increase with orbital period, and overall the UV slope value for the SU UMa systems slightly increases (less shallow) with orbital period.

Standard image High-resolution image

Overall, the DNe present the steepest average slope (−2.20 to −2.27) and, therefore, their disk departs from the standard disk model less than the NLs with an average slope of −1.99 to −1.63. Not surprisingly, the IPs, with their magnetically truncated inner disks, have the shallowest average slope of all, with a value of only −1.43, with a very large scatter ranging from −0.5 to almost −2.5. This is consistent with the suggestion that the slope of the UV continuum becomes shallower as the size of the inner hole in the disk increases. We note, however, that the eclipsing SW Sex stars, with an average slope value of −1.63, seems to cluster around either −2.0 or −1.0.

The small number of DNe that are showing soft X-ray emission in outburst (van Teeseling et al. 1996; Kuulkers et al. 2006), VW Hyi, OY Car, SW UMa, SS Cyg, U Gem, and Z Cam, all exhibit a rather steep UV slope: −2.4, −2.2, −2.5, −2.5, −2.5, and −2.0 respectively. The average value of their UV slope is −2.35, while the average value of the VY Scl with the hardest X-ray emission of all CVs is −1.70, again pointing to the fact that the slope of the UV spectral continuum increases with the softness of the X-ray emission.

We now consider the UV slope by CV subtype, excluding the magnetic IPs and eclipsing SW Sex stars. We obtain the following sequence in order of decreasing steepness of the average UV slope: SU UMa and U Gem systems (∼−2.26), Z Cam systems (−2.20), UX UMa systems (−1.99), and VY Scl systems (−1.70). This sequence is the same for the ratio of the X-ray flux to the optical flux and/or UV flux of CVs in decreasing order: from ∼0.1 for SU UMa systems, to U Gem systems, to ∼0.01 for Z Cam systems, to ∼0.001 for UX UMa systems and VY Scl systems in high state (Kuulkers et al. 2006; Balman 2015), which is due mainly to an increase of the optical/UV flux (or $\dot{M}$). Within the context of our modified disk model, a simple interpretation of this correlation is that the optically thin advective inner disk/BL region increases in size with decreasing mass accretion rate.

From Figures 19 and 20, we also find that for all of the DNe, separately above and below the gap, the slope value is possibly increasing slightly (less negative) with increasing orbital period: the UV slope increases from −2.5 at an orbital period of P ∼ 0.15 days to −2.0 at 0.3 days < P < 0.6 days above the gap (Figure 19); a similar increase is noticeable below the period gap, from −2.5 at P ∼ 0.06 days to −2.0 at P ∼ 0.075 days (Figure 20). The increase of the UV slope value as a function of the orbital period, for the DN systems, is a weak correlation, and it seems to be counterintuitive as one would expect larger mass accretion rate systems, at longer orbital period, to have a disk that is hotter with a steeper UV slope. However, the size of the disk also increases with increasing binary separation at longer orbital periods, thereby increasing the surface area of the colder outer disk, providing a possible explanation for the increase of the UV slope value. This is similar to our modified disk modeling, where the increasing size and the cold temperature (∼10,000–12,000 K) of the outer disk also makes the UV continuum slope shallower.

It has been cautioned that systems with a higher inclination may exhibit a flatter UV continuum slope (la Dous 1991); however, we find that this does not seem to be the case. In fact, there seems to be a selection effect showing the opposite, as most of the VY Scl systems (characterized by a shallow UV continuum slope) in our sample have a low inclination (<40°) while most of the remaining non-magnetic systems (with a steeper UV continuum slope) in our sample have a higher inclination, making it appear as if the low inclination systems have a shallow UV slope. Many systems in Table 4 do not have a known inclination; nevertheless, we draw the UV continuum slope against the inclination in Figure 21. There does not appear to be an obvious correlation between the UV slope and the inclination, except maybe at a very high inclination, where three SW Sex NL systems cluster around −1 (e.g., LX Ser). The IP systems exhibiting the shallowest slope of all (−0.5 to −0.7) do not have an high inclination: FO Aqr has i = 65° and GK Per, i ∼ 46°–73°, but V405 Aur has i < 5°, and PQ Gem has a low (but unknown) inclination. The same is true for the VY Scl systems: V425 Cas with a slope of −1.5 has an inclination of only 25°, BZ Cam with a slope of −1.15 has an inclination less than 40°.

Figure 21.

Figure 21. UV slope against the inclination of the system shown for the CVs in Table 4 for which the inclination is known. Except, possibly, for the few systems close to 90°, there does not appear to be a correlation between the UV slope and the inclination.

Standard image High-resolution image

7. Summary and Conclusion

Our spectral analysis demonstrates that the two VY Scl systems MV Lyr and BZ Cam, departing from the standard disk model in the UV range (with a shallow UV continuum slope of −1.4 and −1.15, respectively), can be modeled with a modified/truncated inner disk (of size ∼Rwd) and a hot WD with an inflated radius. The UX UMa NL system V592 Cas, with a steeper UV continuum slope of −2.4, is modeled with a less modified disk model with a smaller truncated inner disk region (at ∼1.2 Rwd). The picture we present, to justify our modified (truncated) inner disk, is that of a geometrically extended, of size up to ∼Rwd, optically thin BL/inner disk heating the WD through the advection of energy, and which provides an explanation for the observed UV and X-ray data in a self-consistent manner. This scenario is further strengthened by the correlation we found: the non-magnetic disk systems (i.e., excluding quiescent DNe) exhibiting the hardest X-ray (the VY Scl systems in our sample in Table 3) also have the shallowest UV slope, while at the other end, the disk systems exhibiting soft X-ray (five DNe in outburst) have a much steeper UV slope.

We find that the UV continuum slope of the SW Sex systems is possibly affected by their high inclination. As to the magnetically truncated inner disk systems, the IPs, they have the shallowest UV slope of all CVs, consistent with the proposed modified disk models with a truncated inner region providing a shallower UV continuum slope.

In addition, many NL systems exhibit hot, fast wind outflows seen in the P Cygni line structure and blueshifted absorption. These winds originate in the inner disk/BL region, with some systems revealing ejected nebular material (e.g., BZ Cam; Ellis et al. 1984). Also, our X-ray analysis of the NLs MV Lyr, BZ Cam, and V592 Cas (Balman et al. 2014) suggests that these systems have optically thin BLs possibly merged with ADAF-like flows and/or hot coronae.

The X-ray (hard spectra) and UV characteristics (shallow slope) of NL VY systems, including their unexpectedly hot WD (Pala et al. 2017), can be explained with the presence of a geometrically extended optically thin BL/inner disk. When an accretion flow cannot cool efficiently, due to being optically thin, the energy is advected with the flow into the outer layer of the star (Abramowicz et al. 1995), increasing the WD temperature, and likely driving an outflow as in ADAFs (Narayan & Yi 1995). For this reason, we expect the VY Scl systems to be modeled with the modified disk model and the addition of a hot WD, possibly with an inflated radius. The modeling of such systems has to be carried out by creating modified disk models suited to each system individually, taking into account the WD mass, inner and outer disk radius, and the binary separation, itself depending on the secondary mass and orbital period. We note that Balman & Revnivtsev (2012) have shown the existence of a truncated inner disk in quiescent dwarf novae as well, where the flow changes from optically thick into a hot ADAF-like flow. Since quiescent dwarf novae have a very low $\dot{M}$, the optical thickness in the BL must be very small and the truncation radius relatively large.

The UX UMa systems have an average slope somewhat intermediate between that of VY Scl systems and DNe in outburst. Namely, their disk (e.g., V592 Cas) is less modified than for the VY Scl subtype. As to the DN systems, as a whole, below and above the period gap, they have an average UV continuum slope closest to the standard disk and should be modeled with only minimal disk modifications. Their disk has a geometrically small and optically thin inner disk/BL region.

We have proposed here a modified disk model compatible with the X-ray and UV characteristics of non-magnetic CVs, which provides a possible explanation for the discrepancy between the data and theory, simultaneously in the UV (shallow UV slope) and in the X-ray (optically thin and missing BL). We found, by carrying out a quantitative assessment, that the UV and X-ray data self-consistently support this scenario, and suggest that the size of the optically thin inner hole (and advection process) in the disk might be increasing in the following sequence: SU UMa and U Gem systems, followed by Z Cam systems, UX UM systems, and VY Scl systems. In addition, this scenario explains several of the characteristics unique to VY Scl systems, such as their elevated WD temperature and ejected nebular material.

To complete this conclusion, we observe that the optical spectra of disk-dominated CVs also show a continuum slope that is rather shallow when compared to theoretical disk spectra, and the Balmer edge (around ∼3700–4000 Å) is often not detected, or it is seen in absorption but its size is smaller than expected (la Dous 1989). It has been suggested that some of the accretion energy possibly goes into the formation of a disk wind contributing to the UV and optical continuum. Matthews et al. (2017) have shown that the inclusion of such a disk wind significantly improves the spectral fit of the model to the observed UV and optical spectra and fills up the Balmer edge, thereby explaining its non-detection in many optical spectra of disk-dominated CVs and providing a viable explanation for the observed features of UV and optical spectra in CVs. Our present research does not take into account emission from a disk wind as it is limited to the emission from the optically thick disk. In this context, our disk model with a truncated inner region comes as the disk contribution to which a disk wind contribution has to be added to provide a more complete picture. In our modeling (see Section 3.7), we include an isothermal outer disk region with an elevated temperature as the impinging of the L1 stream on the rim of the disk, and the tidal interaction with the secondary is expected to heat the outer region of the disk (Buat-Ménard et al. 2001). Irradiation from the hot inner BL could also contribute to the heating of the outer disk. It is possible that the isothermal outer disk partially compensates for the lack of disk wind emission in our modeling.

We wish to thank The Astrophysical Journal editor Allen Shafter for his promptness in handling the manuscript, and in particular the referee, Ivan Hubeny, for his short, constructive, and encouraging comments and promptness in delivering a report that helped expedite the article. This work was supported by the National Aeronautics and Space Administration (NASA) under grant numbers NNX17AF36G, NNX13AF11G, and NNX13AF12G issued through the Office of Astrophysics Data Analysis Program (ADAP) and grant number NNX13AJ70G issued through the Astrophysics Division Office (Swift Cycle 8 Guest Investigator Program), all at Villanova University. We have used some of the online data from the AAVSO, and are thankful to the AAVSO and its members worldwide for making this data public and their constant monitoring of cataclysmic variables.

Please wait… references are loading.
10.3847/1538-4357/aa7f71