Chapter 5. Accretion Disk Theory

General Picture

The inflowing has a non-zero angular momentum and forms a disk. For a specific angular momentum, $j_0=rv$. Let us ignore the pressure for this moment, and compare the gravity and centrifugal force,

GMr2j02r3\frac{GM}{r^2}\sim\frac{j_0^2}{r^3}

At a small radius, to conserve the specific angular momentum, the centrifugal term must dominate. The critical radius is

Rdj02GMR_\text d\equiv\frac{j_0^2}{GM}

It is known as the disk radius, or the centrifugal radius. It corresponds to the radius within which a disk could form due to the net angular momentum from the accreted gas.

  • Estimation of $R_\text d$

    At any radius, the accreted gas has a random motion velocity of the order of the adiabatic sound speed $c_s$. The effective tangential velocity component $v$ is scaled with an order-one factor $f$, $v\sim fc_s$.

    In addition, the accretion starts at Bondi radius,

    rRBGMc2r\lesssim R_B\sim\frac{GM}{c_\infty^2}

    Consequently,

    j0=rvRBfcfGMrRdf2RB=102(f0.1)2RBj_0=rv\sim R_B\cdot fc_\infty\sim f\frac{GM}{r_\infty}\Rightarrow R_\text d\sim f^2 R_B=10^{-2}\left(\frac{f}{0.1}\right)^2R_B

    Usually $f\sim0.1$, so $R_\text d$ is two orders of magnitude smaller than the Bondi radius.

    • Black Hole

      The so-called surface of the object is the event horizon, whose size is approximately Schwarzschild radius

      RSch=2GMc2=2c2c2RBR_{Sch}=\frac{2GM}{c^2}=\frac{2c_\infty^2}{c^2}R_B

      For HI gas, $c_s\sim10$ km/s, thus for $f\sim0.1$

      RSch109RBRdRBR_{Sch}\sim10^{-9}R_B\ll R_\text d\ll R_B
    • Sun-like Star

      Again for $f\sim0.1$,

      RB103R,Rd10R,RRR_B\sim10^{3}R_\odot,\quad R_\text d\sim10 R_\odot,\quad R_*\sim R_\odot

The disk is, obviously, axisymmetric. All our discussion below will be in the cylindrical coordinate $(r,z)$. $z=0$ corresponds to the equator of the disk.

Profile along $z-$Direction

Along $z-$direction, pressure gradient is balanced with gravity.

csρdρdz=GMr2+z2z(r2+z2)1/2\frac{c_s}{\rho}\frac{\text d\rho}{\text dz}=-\frac{GM}{r^2+z^2}\frac{z}{\left(r^2+z^2\right)^{1/2}}

Fix $r$, and consider isothermal case in which $c_s=c_0$ for simplicity, we can solve this ODE simply by integration

c02lnρ=GM(r2+z2)1/2+Cρ(r,z)=f(r)exp[GMc02(r2+z2)1/2]c_0^2\ln\rho=\frac{GM}{\left(r^2+z^2\right)^{1/2}}+C\Rightarrow \rho(r,z)=f(r)\exp\left[\frac{GM}{c_0^2\left(r^2+z^2\right)^{1/2}}\right]

For $z\ll r$, we can expand the exponential index into Taylor series

GMc02(r2+z2)1/2=GMc02r(112z2r2)\frac{GM}{c_0^2\left(r^2+z^2\right)^{1/2}}=\frac{GM}{c_0^2r}\left(1-\frac12\frac{z^2}{r^2}\right)

Define

f~(r)f(r)exp(GMc02r)\tilde f(r)\equiv f(r)\exp\left(\frac{GM}{c_0^2r}\right)

we have

ρ(r,z)=f~(r)exp(12GMc02r3z2)f~(r)exp(12ΩK2c02z2)\rho(r,z)=\tilde f(r)\exp\left(-\frac12\frac{GM}{c_0^2r^3}z^2\right)\equiv\tilde f(r)\exp\left(-\frac12\frac{\Omega_K^2}{c_0^2}z^2\right)

where $K$ denotes for Keplerian. We can further define

Hc0ΩKH\equiv\frac{c_0}{\Omega_K}

and the $\rho(r,z)$ develops a Gaussian profile

ρ(r,z)=f~(r)exp(12z2H2)\rho(r,z)=\tilde f(r)\exp\left(-\frac12\frac{z^2}{H^2}\right)

$H$ is the disk scale height, which corresponds to the disk thickness.

Now we need to think of the ratio of $H$ to $r$. Only if $H/r\ll1$ is our assumption that $z\ll r$ self-consistent.

Hr=csΩKr=csvKM11    M1\frac Hr=\frac{c_s}{\Omega_Kr}=\frac{c_s}{v_K}\sim \mathcal M^{-1}\ll1\iff\mathcal M\gg1

Thus we claim that the thin disk is equivalent to a cold disk, where the thermal velocity is much smaller than the rotational velocity.

Gas Rotation in the Disk

Consider the terms of the Navier-Stokes equation in $r-$direction,

vrt+vrvrrvϕ2r=1ρPrGMr2+[σ]r\frac{\partial v_r}{\partial t}+v_r\frac{\partial v_r}{\partial r}-\frac{v_\phi^2}{r}=-\frac1\rho\frac{\partial P}{\partial r}-\frac{GM}{r^2}+\left[\nabla\cdot \sigma\right]_r

In the steady state, the time derivative vanishes. If there is no viscosity, due to the conservation of angular momentum, there should not be any mass inflow and materials simply rotate, thus $v_r=0$. So

vϕ2r=1ρPr+GMr2=cs2ρρr+GMr2vϕ2=lnρlnrcs2+vK2\frac{v_\phi^2}{r}=\frac1\rho\frac{\partial P}{\partial r}+\frac{GM}{r^2}=\frac{c_s^2}\rho\frac{\partial \rho}{\partial r}+\frac{GM}{r^2}\Rightarrow v_\phi^2=\frac{\partial\ln\rho}{\partial\ln r}c_s^2+v_K^2

For a thin disk, $cs\ll v_K$, and $v\phi$ is simply the Keplerian velocity. But for a relatively thick disk,

vϕ=vKlnρlnrcs2vK2+1=vKlnρlnrH2r2+1v_\phi=v_K\sqrt{\frac{\partial\ln\rho}{\partial\ln r}\frac{c_s^2}{v_K^2}+1}=v_K\sqrt{\frac{\partial\ln\rho}{\partial\ln r}\frac{H^2}{r^2}+1}

Usually the density gradient is negative, so the rotation is sub-Keplerian, where $v_\phi<v_K$. But there could be some regions where the density gradient is positive. The rotation in such region is super-Keplerian.

In fact, the radial velocity is led by the viscosity. Dimensional analysis gives that

vrνrv_r\sim\frac\nu r

In accretion disks, the consensus is that viscosity comes from turbulence excited by magnetorotational instability (MRI). We know that the kinematic viscosity $\nu$ is determined as

ν=cˉlmfp=cˉLtur\nu=\bar cl_\text{mfp}=\bar cL_\text{tur}

where $\bar c$, here the turbulence velocity, is of the same order as $c_s$. On the other hand, the turbulence scale has a upper limit, which is the disk scale height $H$. As a result,

νcsH\nu\sim c_sH

In 1973, Shakura & Sunyaev proposed a trick by introducing a factor $\alpha$ so that

ν=αcsH\nu=\alpha c_sH

throughout the disk. Though they have no idea of the physical origin for $\alpha-$viscosity, recent MHD simulations for MRI proves that although $\alpha$ is a time-dependent, local parameter, the variation is reasonable and the average $\alpha$ is somehow $0.001-0.01$.

With this assumption, the inflow velocity is

vrνrα(Hr)csα(Hr)2vK2v_r\sim\frac\nu r\sim\alpha\left(\frac Hr\right)c_s\sim\alpha\left(\frac Hr\right)^2v_K^2

So $v_r\ll c_s\ll v_K$ for thin disks.

Energy Production

In a axisymmetric accretion system, the accretion rate is

M˙=2πrvrΣ\dot M=2\pi rv_r\Sigma

where

vrΣHHvr(r,z)ρ(r,z)dzv_r\Sigma\equiv\int_{-H}^Hv_r(r,z)\rho(r,z)\text dz

In the thin disk limit, $v_r(r,z)\sim v_r(r)$, thus

ΣHHρ(r,z)dz\Sigma\equiv\int_{-H}^H\rho(r,z)\text dz

$\Sigma$ is the surface density on the disk. Using our estimation of $v_r$, we have

M˙2πrνrΣνΣ\dot M\sim 2\pi r\cdot\frac \nu r\cdot\Sigma\sim\nu\Sigma

Now we can calculate the energy generation rate ([erg/s/cm$^2$], energy flux) due to accretion, during which gravitational energy is transformed into thermal energy via visocity.

Qvis+1πr2GMrM˙νΣΩK2Q^+_\text{vis}\sim\frac1{\pi r^2}\cdot\frac{GM}r\cdot\dot M\sim\nu\Sigma\Omega_K^2

This relation can be derived from Navier-Stokes equation directly. In fact,

Qvis+νΣ(dΩdlnr)294νΣΩK2Q^+_\text{vis}\sim\nu\Sigma\left(\frac{\text d\Omega}{\text d\ln r}\right)^2\sim\frac94\nu\Sigma\Omega_K^2

The energy is carried away mainly via radiation, convection, and advection. The convection can be important, but is really complicated and beyond the scope of this lecture. Here we mainly discuss advection and radiation.

Advection

In advection, the bulk fluid motion carries certain energy away. Again, out of dimensional analysis, we have

Qadv1πr2M˙cs2Q^-_\text{adv}\sim\frac1{\pi r^2}\cdot \dot Mc_s^2

where $\dot M$ describes the bulk motion and the specific internal energy is scaled with $cs^2$. The ratio of $Q^-\text{adv}$ to $Q^+_\text{vis}$ is

QadvQvis+M˙cs2GMrM˙cs2vK21\frac{Q^-_\text{adv}}{Q^+_\text{vis}}\sim\frac{\dot Mc_s^2}{\frac{GM}r\dot M}\sim\frac{c_s^2}{v_K^2}\ll1

which suggests that for thin disk case, advection cannot efficiently takes energy away. For thicker disks, however, advection can play an important role in energy transportation.

Radiation

In a geometrically thin disk, we estimate the optical depth $\tau$.

τ12κΣ12κM˙2πrvr\tau\sim\frac12\kappa\Sigma\sim\frac12\kappa\frac{\dot M}{2\pi rv_r}

We consider the fluid behavior in the vicinity of the black hole, that is, $r\sim R_{Sch}$. Because of the steep potential well predicted in GR, the infalling velocity is approximately the speed of light. In this way,

τ12M˙c24πGMc/κ12M˙c2LEdd12ηM˙M˙Edd\tau\sim\frac12\frac{\dot Mc^2}{4\pi GMc/\kappa}\sim\frac12\frac{\dot Mc^2}{L_{Edd}}\sim\frac1{2\eta}\frac{\dot M}{\dot M_{Edd}}

For a typical efficiency $\eta\sim0.1$, when $\dot M\sim\dot M_{Edd}$, $\tau$ is significantly larger than unity. Note that $c$ is strictly an upper limit of $v_r$, so we have underestimated the optical depth. Absolutely the disk is optically thick. So the energy transporting rate via radiation is

QradσSBT4τQ^-_\text{rad}\sim\frac{\sigma_{SB}T^4}{\tau}

Angular Momentum Transfer and Viscosity

In 1974, Lynden-Bell & Pringle claimed that rotating accretion disks should achieve minimum energy.

Suppose a test particle with specific angular momentum $j$ in a potential $\psi=-\frac{GM}r$. The motion is Keplerian, thus the specific energy is

ε(j)=G2M22j2(1e2)\varepsilon(j)=-\frac{G^2M^2}{2j^2}(1-e^2)

where $e$ is the orbital eccentricity. The minimum energy state achieves with a circular orbit, where

ε(j)=G2M22j2,dεdj(j)=G2M2j3=G2M2ΩK3r6=ΩK\varepsilon(j)=-\frac{G^2M^2}{2j^2},\quad \frac{\text d\varepsilon}{\text dj}(j)=\frac{G^2M^2}{j^3}=\frac{G^2M^2}{\Omega_K^3r^6}=\Omega_K

Now consider two particles orbiting around $M$. The total energy and angular momentum are,

E=m1ε(j1)+m2ε(j2)J=m1j1+m2j2E=m_1\varepsilon(j_1)+m_2\varepsilon(j_2)\\ J=m_1j_1+m_2j_2

To conserve the total mass and angular momentum, we have

dm1+dm2=0d(m1j1)+d(m2j2)=0\text dm_1+\text dm_2=0\\ \text d(m_1j_1)+\text d(m_2j_2)=0

The lowest energy state obviously requires circular orbits. The energy differential is thus

dE=ε(j1)dm1+ε(j2)dm2+dεdj(j1)d(m1j1)dεdj(j2)d(m2j2)=[ε(j1)ε(j2)]dm1+[dεdj(j1)dεdj(j2)]d(m1j1)=[ε(j1)ε(j2)]dm1+(Ω1Ω2)d(m1j1)<0\begin{align*} \text dE&=\varepsilon(j_1)\text dm_1+\varepsilon(j_2)\text dm_2+\frac{\text d\varepsilon}{\text dj}(j_1)\text d(m_1j_1)-\frac{\text d\varepsilon}{\text dj}(j_2)\text d(m_2j_2)\\ &=[\varepsilon(j_1)-\varepsilon(j_2)]\text dm_1+\left[\frac{\text d\varepsilon}{\text dj}(j_1)-\frac{\text d\varepsilon}{\text dj}(j_2)\right]\text d(m_1j_1)\\ &=[\varepsilon(j_1)-\varepsilon(j_2)]\text dm_1+(\Omega_1-\Omega_2)\text d(m_1j_1)\\ &<0 \end{align*}

Without loss of generality, we assume $j_1>j_2$. Since

ΩKj3,ε(j)j2\Omega_K\propto j^{-3},\quad -\varepsilon(j)\propto j^{-2}

and

r=jΩKrrj2\quad r=\frac j{\Omega_K r}\Rightarrow r\propto j^2

we have $r_1>r_2$, $\varepsilon(j_1)>\varepsilon(j_2)$, and $\Omega_1<\Omega_2$. To ensure that $\text dE0$. Thus mass is transported inwards (to particle 2), while angular momentum is transported outwards (to particle 1).

Evolution of Disk Surface Density

Let's now consider a real disk structure. Without viscosity, all materials follow Keplerian orbits, and no accretion actually happens, so viscosity is intrinsic in accretion disks. The most important non-diagonal component of the stress tensor is (especially out of the thin disk geometry) $\sigma_{r\phi}$. In cylindrical coordinate, it writes

σrϕ=ρν(vϕrvϕr)=ρνrΩr\sigma_{r\phi}=\rho\nu\left(\frac{\partial v_\phi}{\partial r}-\frac{v_\phi}{r}\right)=\rho\nu r\frac{\partial \Omega}{\partial r}

The corresponding friction force is

Frϕ=σrϕ2πr2HF_{r\phi}=\sigma_{r\phi}\cdot2\pi r\cdot2H

Thus the net torque onto the belt within $r$ and $r+\text dr$ is

N=(r+dr)Frϕ(r+dr)rFrϕ(r)=r(rFrϕ)dr=drr(2ρH2πr3νΩr)=drr(2πr3ΣνΩr)\begin{align*} N&=(r+\text dr)F_{r\phi}(r+\text dr)-rF_{r\phi}(r)\\ &=\frac{\partial}{\partial r}\left(rF_{r\phi}\right)\text dr\\ &=\text dr\frac{\partial}{\partial r}\left(2\rho H\cdot2\pi r^3\nu\frac{\partial \Omega}{\partial r}\right)\\ &=\text dr\frac{\partial}{\partial r}\left(2\pi r^3\Sigma\nu\frac{\partial \Omega}{\partial r}\right) \end{align*}

while the angular momentum of the belt is

J=2πrdrΣjJ=2\pi r\text dr\Sigma j

Therefore, the angular momentum conservation is

Jt+r(Jvr)=N\frac{\partial J}{\partial t}+\frac{\partial}{\partial r}\left(Jv_r\right)=N
    t(2πrΣj)+r(2πrΣvrj2πr3ΣνΩr)\iff\frac{\partial }{\partial t}(2\pi r\Sigma j)+\frac{\partial}{\partial r}\left(2\pi r\Sigma v_rj-2\pi r^3\Sigma\nu\frac{\partial \Omega}{\partial r}\right)

Simply from the EoC, the accretion rate is

M˙=2πrΣvr\dot M=-2\pi r\Sigma v_r

So we can define the advection flux

FJ,adv=2πrΣvrj=M˙jF_{J,\text{adv}}=2\pi r\Sigma v_rj=-\dot Mj

and the viscosity flux

FJ,vis=2πr3ΣνΩrF_{J,\text{vis}}=-2\pi r^3\Sigma\nu\frac{\partial \Omega}{\partial r}

By simply taking $\Omega=\Omega_K$ in the thin disk approximation,

FJ,vis=3πr2ΣνΩK>0F_{J,\text{vis}}=3\pi r^2\Sigma\nu\Omega_K>0

Again, angular momentum is transferred outwards.

Steady State Solution ($\partial/\partial t=0$)

r(M˙j)=r(3πr2ΣνΩK)=r(3πΣνj)\Rightarrow\frac{\partial}{\partial r}\left(\dot Mj\right)=\frac{\partial}{\partial r}\left(3\pi r^2\Sigma\nu\Omega_K\right)=\frac{\partial}{\partial r}\left(3\pi \Sigma\nu j\right)

If $r_\text{in}$ is the innermost radius of the disk, we have

M˙(jjin)=3πΣνj(3πΣνj)in\dot M(j-j_\text{in})=3\pi \Sigma\nu j-\left(3\pi \Sigma\nu j\right)_\text{in}

Let's impose the torque-free boundary condition at $r=r_\text{in}$, which suggests no viscosity in the inner boundary of the disk.

For black holes, $r\text{in}\simeq r\text{ISCO}\simeq6r_{Sch}$, inside which the GR effect suppresses any possible stable orbital motion, so the torque-free boundary condition holds. However, for accretion onto proton-stars / neutron stars, such condition is not justified.

M˙=3πΣνjjjin=3πΣν(1rinr)1\Rightarrow \dot M=\frac{3\pi \Sigma\nu j}{j-j_\text{in}}=3\pi \Sigma\nu \left(1-\sqrt{\frac{r_\text{in}}{r}}\right)^{-1}

For $r\gg r_\text{in}$, $\dot M\simeq3\pi \Sigma\nu$. When we combine this solution with the EoC, we obtain

vr=32νrv_r=-\frac32\frac\nu r

We used to claim a viscous timescale

tvis=R2νt_\text{vis}=\frac{R^2}{\nu}

out of the dimensional consideration. Now we prove it.

Non-Steady Solution

The EoC now gives

Σt=12πrM˙r,M˙=2πrΣvr\frac{\partial \Sigma}{\partial t}=\frac1{2\pi r}\frac{\partial \dot M}{\partial r},\quad\dot M=-2\pi r\Sigma v_r

And the angular momentum conservation gives

2πrjΣt+r(M˙j2πr3ΣνΩr)=02\pi rj\frac{\partial \Sigma}{\partial t}+\frac{\partial}{\partial r}\left(-\dot Mj-2\pi r^3\Sigma\nu\frac{\partial \Omega}{\partial r}\right)=0
    M˙jr=r(2πr3ΣνΩr)\iff \dot M\frac{\partial j}{\partial r}=-\frac{\partial}{\partial r}\left(2\pi r^3\Sigma\nu\frac{\partial \Omega}{\partial r}\right)
    Σt=12πrM˙r=12rr[r(2r3ΣνΩr)r(Ωr2)]=12rr[r(3r2ΣνΩK)2ΩKr]=3rr[rr(rΣν)]\begin{align*} \iff \frac{\partial \Sigma}{\partial t}&=\frac1{2\pi r}\frac{\partial \dot M}{\partial r}\\ &=\frac1{2r}\frac{\partial }{\partial r}\left[-\frac{\partial}{\partial r}\left(2r^3\Sigma\nu\frac{\partial \Omega}{\partial r}\right)\frac{\partial r}{\partial (\Omega r^2)}\right]\\ &=\frac1{2r}\frac{\partial }{\partial r}\left[\frac{\partial}{\partial r}\left(3r^2\Sigma\nu\Omega_K\right)\frac{2}{\Omega_K r}\right]\\ &=\frac3{r}\frac{\partial }{\partial r}\left[\sqrt r\frac{\partial}{\partial r}\left(\sqrt r\Sigma\nu\right)\right] \end{align*}

If we assume $\nu$ is roughly a constant throughout the disk, we can define

Σ~=rΣ\widetilde\Sigma=\sqrt r\Sigma

so that

Σ~t=3νrr[rΣ~r]=3ν2rΣ~r+3ν2Σ~r2=3νΣ~(r2)+3ν2Σ~r2\begin{align*} \frac{\partial\widetilde\Sigma}{\partial t}&=\frac{3\nu}{\sqrt r}\frac{\partial }{\partial r}\left[\sqrt r\frac{\partial \widetilde\Sigma}{\partial r}\right]=\frac{3\nu}{2 r}\frac{\partial \widetilde\Sigma}{\partial r}+3\nu\frac{\partial^2 \widetilde\Sigma}{\partial r^2}\\ &=3\nu\frac{\partial \widetilde\Sigma}{\partial (r^2)}+3\nu\frac{\partial^2 \widetilde\Sigma}{\partial r^2} \end{align*}

The first term,

Σ~t=3νΣ~(r2)\frac{\partial\widetilde\Sigma}{\partial t}=3\nu\frac{\partial \widetilde\Sigma}{\partial (r^2)}

is a wave equation. The general solution is $\widetilde \Sigma=f(r^2-3\nu t)$, where $f(\cdot)$ is an arbitrary function. As the system evolves with time, the wave propogates while the shape remains the same.

The second term,

Σ~t=3ν2Σ~r2\frac{\partial\widetilde\Sigma}{\partial t}=3\nu\frac{\partial^2 \widetilde\Sigma}{\partial r^2}

is a diffusion equation. The general solution peaks at the same $r$, but the shape is gradually diluted. The overall solution behaves as a mixture of the two.

However, $\nu$ is in general not a constant. For a thin disk,

ν=αcsH=αcs2ΩKT(r)r3/2\nu=\alpha c_sH=\alpha\frac{c_s^2}{\Omega_K}\propto T(r)r^{3/2}

So we have to solve the equation numerically.

Standard Disk Model

Shakura & Sunyaez 1973, 1976

Assumptions

  1. Axisymmetric disk ($\partial/\partial \phi=0$).

  2. Steady disk ($\partial/\partial t=0$). As a result,

    M˙=2πrvrΣ\dot M=2\pi rv_r\Sigma

    is a constant.

    In the last section, we have proved that when this assumption is applied and the torque-free boundary condition is imposed, we have

    νΣ=M˙3π(1rinr)M˙3π\nu\Sigma=\frac{\dot M}{3\pi}\left(1-\sqrt{\frac{r_\text{in}}{r}}\right)\simeq\frac{\dot M}{3\pi}
  3. The gravity is dominated by the central object.

  4. Hydrostatic balance in the vertical direction.

    H=csΩKH=\frac{c_s}{\Omega_K}
    ΩK=GMr3\Omega_K=\sqrt{\frac{GM}{r^3}}
  5. Geometrically thin disk ($H/r=c_s/v_K\ll1$).

  6. Rotation dominated ($|vr|\ll v\phi=v_K$).

  7. Somehow the least reasonable assumption: viscosity is given by the $\alpha$-viscosity model.

    ν=23αcsH\nu=\frac23\alpha c_sH

    where $\alpha=\mathcal O(0.01-1)$.

  8. The disk should be in thermal equilibrium ($T$ is well-defined throughout the disk) and optically thick, $Q^+\text{vis}=Q^-\text{rad}+Q^-_\text{adv}$.

    For the thin disk model, the advection term is negligible, and we have proved that

    Qvis+νΣ(dΩdlnr)294νΣΩK2Qrad2σSBT4τ=8acT43τQ^+_\text{vis}\sim\nu\Sigma\left(\frac{\text d\Omega}{\text d\ln r}\right)^2\sim\frac94\nu\Sigma\Omega_K^2\simeq Q^-_\text{rad}\sim\frac{2\sigma_{SB}T^4}{\tau}=\frac{8ac T^4}{3\tau}

    The disk has two sides, thus we add a factor of two in the radiation term. The optical depth $\tau$ is given by

    τ=κH2\tau=\kappa\cdot\frac H2
  9. The opacity is mainly given by the electron scattering and free-free transition.

    κ=κes+κff=κes+κ0ρT7/2\kappa=\kappa_\text{es}+\kappa_\text{ff}=\kappa_\text{es}+\kappa_0\rho T^{-7/2}

    where the density $\rho$ is given by

    ρ=ΣH\rho=\Sigma H
  10. Equation of state is given by

    P=Pgas+Prad=ρkBTμmp+13aT4P=P_\text{gas}+P_\text{rad}=\frac{\rho k_BT}{\mu m_\text p}+\frac13aT^4

    and the adiabatic sound speed is given by

    cs=γPρPρc_s=\sqrt{\gamma\frac P\rho}\sim\sqrt{\frac P\rho}

For fixed $M$, $\dot M$, and $\alpha$, we still have 11 unkown quantities ($v_r, \Sigma, \Omega_K, H, \rho, \nu, c_s, P, T, \kappa, \tau)$ and exactly 11 equations, so all the quantities can be solved as a function of $r$.

The result gives a broken power law because the formula for $\kappa$ and $P$ include addition. The disk is thus divided into three regions.

  • Region I, $\kappa\text{es}$ and $P\text {rad}$ dominate.

  • Region II, $\kappa\text{es}$ and $P\text {gas}$ dominate.

  • Region III, $\kappa\text{ff}$ and $P\text {gas}$ dominate.

In each region, the profile of each quantity has a individual slope, and is scale-free, by defining dimensionless quatities below

mMM,m˙M˙M˙cr,r^rrSchm\equiv\frac M{M_\odot},\quad \dot m\equiv\frac{\dot M}{\dot M_\text{cr}},\quad \hat r\equiv \frac r{r_{Sch}}

where

M˙crLEddc2,rSch2GMc2\dot M_\text{cr}\equiv\frac{L_{Edd}}{c^2},\quad r_{Sch}\equiv\frac{2GM}{c^2}

and $L_{Edd}$ is the Eddington luminosity,

LEdd4πcGMκesL_{Edd}\equiv \frac{4\pi cGM}{\kappa_\text{es}}

Note that after solving the equation, one need to double check to make sure $|v_r|\ll c_s\ll v_K$ and $H/r\ll 1$ so that our assumpitions shall be valid.

Disk Energetics and Radiation Spectra

The energy equation is

Qvis+=QradQ^+_\text{vis}= Q^-_\text{rad}

So the one-side flux is

F=12Qrad=98νΣΩK2=3GMM˙8πr3(1rinr)F=\frac{1}{2}Q^-_\text{rad}=\frac98\nu\Sigma\Omega_K^2=\frac{3GM\dot M}{8\pi r^3}\left(1-\sqrt{\frac{r_\text{in}}{r}}\right)

In an accretion disk, gravitational energy is released, and viscosity transports it to radiation. If there is significant advection, since $Q^+\text{vis}=Q^-\text{rad}+Q^-_\text{adv}$, the luminosity is reduced.

The total disk luminosity is

L=rinQrad2πrdr=32GMM˙rin1r2(1rinr)dr=GMM˙2rinL=\int_{r_\text{in}}^\infty Q^-_\text{rad}\cdot2\pi r\text dr=\frac32 GM\dot M \int_{r_\text{in}}^\infty\frac1{r^2}\left(1-\sqrt{\frac{r_\text{in}}{r}}\right)\text dr=\frac{GM\dot M}{2r_\text{in}}

If $r\text{in}=r\text{ISCO}=6GM/c^2$, $L=\dot Mc^2/12$.

Last updated