Towards a Truly Unified Model of AGN:
Aspect, Accretion and Evolution

Michael A. Dopita, PASA, 14 (3), 230
The html and gzipped postscript versions of this paper are in preprint form.
To access the final published version, download the pdf file
.

Next Section: Relativistic Jet Model
Title/Abstract Page: Towards a Truly Unified
Previous Section: The Role of Accretion
Contents Page: Volume 14, Number 3

Radiation-Pressure Driven Wind Model

During the period of super-Eddington accretion and possibly beyond, the radiation pressure from the central BH would be sufficient to drive a wind. If this wind is driven by radiation pressure of the central source, then equations similar to those of hot stars apply. For the case of a magnetic wind externally illuminated by the central engine, the detailed theory has already been developed in a series of papers by Murray and his collaborators (Murray and Chiang, 1995; Murray et al. 1995; Chiang and Murray, 1996). However, in our case we would like to consider the possibility that the radiatively driven wind is both dense, so that it is not heated to the Compton temperature, and is optically thick to the escape of X- and tex2html_wrap_inline893- ray photons from the central source so that the continuum observed is produced by a reprocessing photosphere dominated by electron scattering opacity. Such geometry may apply in the high accretion rate limit. In this case, the momentum flux in the wind is given by:


 equation181
where tex2html_wrap_inline1061 is the solid angle covered by the optically-thick radiatively-driven wind, subtended at the BH, and tex2html_wrap_inline1063 is the effective number of scatterings per photon. In stars, this factor is greater than unity; typically tex2html_wrap_inline1065 however, in Wolf-Rayet stars it may rise even higher. In such radiatively-driven winds, the outflow velocity is similar to the escape velocity at the base of the outflow, which in this case will be at the inner edge of the region of super-Eddington flow:


 equation192
where is the mass of the BH in units of 10tex2html_wrap_inline977 and tex2html_wrap_inline1071 is the inner radius in units of 10tex2html_wrap_inline1073cm. In stars, 1tex2html_wrap_inline1075 typically. From 7 the number density in the wind is given by:


 equation212
which corresponds to tex2html_wrap_inline1077 cmtex2html_wrap_inline1079 where tex2html_wrap_inline1081 is the luminosity of the BH in units of 10tex2html_wrap_inline1083 ergs.stex2html_wrap_inline949, and tex2html_wrap_inline1087 is the radius in units of 10tex2html_wrap_inline1089cm and tex2html_wrap_inline1091 is the velocity of the wind in units of 10tex2html_wrap_inline973 km.stex2html_wrap_inline949. The photospheric radius is set by the point where the electron scattering optical depth in the outflowing wind is unity, which for a constant-velocity wind gives:


 equation234
where tex2html_wrap_inline1097 is the electron scattering opacity. Such photospheric radii are consistent with limits produced by variability studies and reverberation mapping analysis of nearby Sy I galaxies. The effective temperature of the photosphere is then determined from Stefan's Law, remembering that electron scattering photospheres have their radiation density diluted by a factor tex2html_wrap_inline109910 with respect to a Black-Body emitter;


 equation247
Thus, when the wind velocity is of order 20,000 km.stex2html_wrap_inline949 as implied by eqn. (8), the effective photospheric temperature is high enough ( tex2html_wrap_inline929 10tex2html_wrap_inline1105 K) to provide a ``big blue bump'' in the continuum spectrum which, thanks to the high outflow velocity and the electron scattering, should provide only weak and broad photospheric lines. In addition, such electron scattering dominated extended atmospheres should show only a weak Lyman Limit discontinuity. It should be noted that the spectral distribution of such an electron scattering photosphere is not characterised by a simple temperature; the temperature given in eqn. (11) is only representative of the hardness of the spectrum. In general, such atmospheres give a power-law below the peak in emergent flux, and roll off sharply above this peak. This is the result of the fact that the peak temperature, given roughly by eqn. (11) is seen in the line of sight to the centre of the structure where the deepest penetration occurs. The photosphere occurs further out in annuli around this point, and so is characterised by lower electron temperatures. The contribution from these annuli produce the power-law extension of the spectrum to lower frequencies. The slope of this power-law depends on the effective curvature of the atmosphere, i.e. on the ratio tex2html_wrap_inline1107 This effect is very familiar to those who work on Wolf-Rayet stellar atmospheres (e.g. Abbott and Conti, 1987). In AGN, the value of this ratio is much higher than in Wolf-Rayet stars, so we expect them to be characterised by little curvature, and a rather flat power-law (tex2html_wrap_inline1109 tex2html_wrap_inline1111 1) at energies below the peak.

Since the photosphere is extended both above, and below the accretion disk, it can illuminate and photoionise the surface layers of the accretion disk. This would give the broad-line region in our model. Since the effective temperature decreases for higher BH luminosities, we would expect that this photoionisation would provide lower ionisation conditions in the more luminous AGN. This is presumably the explanation of the Baldwin Effect in QSOs (Baldwin (1977), Baldwin et al. (1978), where the CIV equivalent width and the ratio of C IV/ Lytex2html_wrap_inline925 is observed to decrease with increasing luminosity. This effect has been most comprehensively studied by Kinney, Rivolo and Koratkar (1990).

Equations (7), (8) and (10) imply together that:


 equation282
consistency therefore requires tex2html_wrap_inline1115 tex2html_wrap_inline929 tex2html_wrap_inline1063 if the BH luminosity is to be of the same order as the Eddington Limit. Equation (7) also implies that the mechanical energy flux in the wind is simply related to the bolometric luminosity of the central object:


 equation292
Finally, we can relate the mass flux in the wind to the mass flux onto the BH, using Eqns (5) and (7):


 equation307
Thus, the mass flux into the radiatively driven wind will dominate over the mass flux into the BH until the accretion disk becomes thin (tex2html_wrap_inline1121

The outflowing radiative wind interacts directly with the accretion flow. The condition that the accretion flow is not seriously dynamically modified by this interaction is that the ram pressure in the wind is less than the ram pressure in the accretion flow. From eqn. (7) this implies that the BH luminosity ( in units of 10tex2html_wrap_inline1123erg.stex2html_wrap_inline949; tex2html_wrap_inline1127) is below tex2html_wrap_inline1129 where tex2html_wrap_inline945 is the rate of mass accretion in units of 200 tex2html_wrap_inline947yrtex2html_wrap_inline1135 tex2html_wrap_inline951 is the velocity of infall in units of 300 km.stex2html_wrap_inline949 and tex2html_wrap_inline955 is the solid angle covered by the accretion flow.


Next Section: Relativistic Jet Model
Title/Abstract Page: Towards a Truly Unified
Previous Section: The Role of Accretion
Contents Page: Volume 14, Number 3

Welcome... About Electronic PASA... Instructions to Authors
ASA Home Page... CSIRO Publishing PASA
Browse Articles HOME Search Articles
© Copyright Astronomical Society of Australia 1997
ASKAP
Public