Chapter 2. Fluid Dynamics

Fluid Approximation

The fluid approximation is valid when the system size $L$ is much larger than the mean free path $l_\text{mfp}$.

  • For ISM (interstellar medium), the typical mean free path is

    lmfp1015cm(n1 cm3)1(d1 A˚)2l_\text{mfp}\sim 10^{15}\text{cm}\left(\frac{n}{1\text{ cm}^{-3}}\right)^{-1}\left(\frac{d}{1\ \AA}\right)^{-2}

Inside each fluid particle, we can define thermal dynamical quantities, $T,\rho,P,S$, as well as a velocity field $\vec v(\vec x,t)$.

For any quantity of the fluid element at $(\vec r_0,t_0)$, say $F(\vec r_0,t_0)$, we have

ΔF=F(r0+v(r0)Δt,t0+Δt)F(r0,t0)=(Ft+vxFx+vyFy+vzFz)Δt+O(Δt2)\begin{align*} \Delta F&=F\left(\vec r_0+\vec v(\vec r_0)\Delta t,t_0+\Delta t\right)-F(\vec r_0,t_0)\\ &=\left(\frac{\partial F}{\partial t}+v_x\frac{\partial F}{\partial x}+v_y\frac{\partial F}{\partial y}+v_z\frac{\partial F}{\partial z}\right)\Delta t+\mathcal O(\Delta t^2) \end{align*}

In fact, we can write the derivative of $F$

limΔt0ΔFΔt\lim_{\Delta t\to 0}\frac{\Delta F}{\Delta t}

in two different ways

  • Lagrange's derivative

    DFDt\frac{\text DF}{\text Dt}

    which follows the fluid element everywhere.

  • Eulerian derivative

    (t+v)F\left(\frac{\partial}{\partial t}+\vec v\cdot \nabla\right) F

    which is more important for numerical calculations, as it considers only fixed fluid cells.

Equation of Continuity - Mass Conservation

  • In Eulerian picture

    For a small, fixed volume $\delta V_0$, in a time span of $\Delta t$, the change in mass inside is

    δm=Δtddt(δV0ρdV0)\delta m=\Delta t\cdot\frac{\text d}{\text dt}\left(\iiint_{\delta V_0}\rho\text dV_0\right)

    Consider the mass flow at the surface

    j=ΔtδS(vn)ρdS=ΔtδV0(ρv)dV0\vec j=\Delta t\oiint_{\delta S}\left(\vec v\cdot\vec n\right)\rho\text dS=\Delta t\iiint_{\delta V_0}\nabla\cdot\left(\rho\vec v\right)\text dV_0

    where we have applied Gauss' divergence theorem. Here $\vec n$ is the normal vector of the surface.

    If there are no source/sink of mass within the fluid element, the mass flow should exactly account for the change in mass

    ddt(δV0ρdV0)+δV0(ρv)dV0=0\frac{\text d}{\text dt}\left(\iiint_{\delta V_0}\rho\text dV_0\right)+\iiint_{\delta V_0}\nabla\cdot\left(\rho\vec v\right)\text dV_0=0

    Since we have fixed $\Delta V$, we can write this formula so that for any $\Delta V_0$

    δV0(ρt+(ρv))dV0=0\iiint_{\delta V_0}\left(\frac{\partial \rho}{\partial t}+\nabla\cdot\left(\rho\vec v\right)\right)\text dV_0=0

    The arbitrariness of the choice of $\Delta V$ directly leads to the famous equation of continuity

    ρt+(ρv)=0\frac{\partial \rho}{\partial t}+\nabla\cdot\left(\rho\vec v\right)=0
  • In Lagrange's picture

    Let us consider a volume element $\delta V$ that moves with fluid particles.

    If no mass creation / annihilation

    0=D(δm)Dt=DDt(ρδV)=DρDtδV+ρDDt(δV)0=\frac{\text D(\delta m)}{\text Dt}=\frac{\text D}{\text Dt}\left(\rho\delta V\right)=\frac{\text D\rho}{\text Dt}\delta V+\rho\frac{\text D}{\text Dt}\left(\delta V\right)
    DρDt=ρδVDDt(δV)=ρi1δxiD(δxi)Dt=ρ(v)\Rightarrow \frac{\text D\rho}{\text Dt}=-\frac{\rho}{\delta V}\frac{\text D}{\text Dt}\left(\delta V\right)=-\rho\sum_{i}\frac{1}{\delta x_i}\frac{\text D\left(\delta x_i\right)}{\text Dt}=-\rho\left(\nabla\cdot \vec v\right)

    Recall that

    DDt=t+v\frac{\text D}{\text Dt}=\frac{\partial}{\partial t}+\vec v\cdot \nabla

    Then

    (t+v+v)ρ=0ρt+(ρv)=0\left(\frac{\partial}{\partial t}+\vec v\cdot \nabla+\nabla\cdot\vec v\right)\rho=0\Rightarrow\frac{\partial\rho}{\partial t}+\nabla\cdot\left(\rho\vec v\right)=0

Two pictures give exactly the same equation!

Now we further explore the relation between two pictures. In the Lagrange's picture,

δV=dxdydz\delta V=\text dx\cdot\text dy\cdot\text dz

while in the Eulerian picture

δV0=dξxdξydξz\delta V_0=\text d\xi_x\cdot\text d\xi_y\cdot\text d\xi_z

$\delta V$ and $\delta V_0$ are related as follows,

δV=(x,y,z)(ξx,ξy,ξz)δV0JδV0\delta V=\left|\frac{\partial\left(x,y,z\right)}{\partial\left(\xi_x,\xi_y,\xi_z\right)}\right|\delta V_0\equiv J\delta V_0

The ratio $J$ is called the expansion of the fluid. Euler's expansion formula claims that

DJDt=(v)J\frac{\text DJ}{\text Dt}=\left(\nabla \cdot \vec v\right)J

Then if we reconsider the conservation of mass in Lagrange's picture,

DDtδVρdV=0\frac{\text D}{\text Dt}\iiint_{\delta V}\rho\text{d}V=0
    0=DDtδVρJdV0=δV[DρDt+ρ(v)]JdV0\iff 0=\frac{\text D}{\text Dt}\iiint_{\delta V}\rho J\text{d}V_0=\iiint_{\delta V}\left[\frac{\text D \rho}{\text Dt}+\rho\left(\nabla\cdot \vec v\right)\right]J\text dV_0

which also gives the continuity formula

DρDt+ρ(u)=0\frac{\text D \rho}{\text Dt}+\rho\left(\nabla\cdot u\right)=0

Before moving to the next section, we can similarly derive several useful identities of the Lagrange's derivative.

For a function $F$,

DDtVFdV=DDtVFJdV0=V[DFDt+(v)F]JdV0=V[DFDt+(v)F]dV\frac{\text D }{\text Dt}\iiint_VF\text dV=\frac{\text D }{\text Dt}\iiint_VFJ\text dV_0=\iiint_V\left[\frac{\text D F}{\text Dt}+\left(\nabla \cdot \vec v\right)F\right]J\text dV_0=\iiint_V\left[\frac{\text D F}{\text Dt}+\left(\nabla \cdot \vec v\right)F\right]\text dV
    DDtVFdV=V[Ft+(Fv)]dV\iff \frac{\text D }{\text Dt}\iiint_VF\text dV=\iiint_V\left[\frac{\partial F}{\partial t}+\nabla\cdot\left(F \vec v\right)\right]\text dV

This is known as the Reynolds transport theorem. Simply let $F=\alpha\rho$, we have

DDtVαρdV=V[DDt(αρ)+(v)αρ]dV=V{α[DρDt+(v)ρ]+ρDαDt}dV=VρDαDtdV\begin{align*} \frac{\text D }{\text Dt}\iiint_V\alpha\rho\text dV&=\iiint_V\left[\frac{\text D }{\text Dt}\left(\alpha\rho\right)+\left(\nabla \cdot \vec v\right)\alpha\rho\right]\text dV\\ &=\iiint_V\left\{\alpha\left[\frac{\text D \rho}{\text Dt}+\left(\nabla \cdot \vec v\right)\rho\right]+\rho\frac{\text D\alpha}{\text Dt}\right\}\text dV\\ &=\iiint_V\rho\frac{\text D\alpha}{\text Dt}\text dV \end{align*}

This identity is extremely useful in the following chapters.

Euler Equation - Momentum Conservation

In the Lagrange's picture, the acceleration onto a mass element is given by

DDtV(t)ρvdV=V(t)ρDvDtdV\frac{\text D}{\text Dt}\iiint_{V(t)}\rho\vec v\text dV=\iiint_{V(t)}\rho\frac{\text D\vec v}{\text Dt}\text dV

With the Reynolds transport theorem, the EoS thus gives

V(t)ρDvDtdV=V(t)fdV+S(t)tdS\iiint_{V(t)}\rho\frac{\text D\vec v}{\text Dt}\text dV=\iiint_{V(t)}\vec f\text dV+\oiint_{S(t)}\vec t\text dS
  • First term - body force (e.g gravity)

  • Second term - surface force (e.g. pressure)

Consider the $i$-th component of the RHS

V(t)fidV+S(t)tidS=V(t)fidV+S(t)TijnjdS=V(t)(fi+jTij)dV\iiint_{V(t)}f^i\text dV+\oiint_{S(t)}t^i\text dS=\iiint_{V(t)}f^i\text dV+\oiint_{S(t)}T^{ij}n_j\text dS=\iiint_{V(t)}\left(f^i+\partial_jT^{ij}\right)\text dV

where $\mathcal T$ is the stress tensor, which satisfies

ti=Tijnjt^i=T^{ij}n_j

An important surface force is pressure.

The force caused by pressure is

Fp=SpndS=VpdV\vec F_p=-\iint_S p\vec n\text dS=-\iiint_V\nabla p\text dV
  • Pressure and Stress Tensor

    If the $i$-th component

    ip=jTij-\partial_i p=\partial_j T^{ij}

    Thus $p$ is strongly correlated with the diagonal components in $\mathcal T$.

Since the choice of $V$ is arbitrary, the EoM is expressed as

ρDvDt=p+f\rho\frac{\text D\vec v}{\text Dt}=-\nabla p+\vec f
    vt+(v)v=1ρp+1ρf\iff \frac{\partial \vec v}{\partial t}+\left(\vec v\cdot \nabla\right)\vec v=-\frac1\rho\nabla p+\frac1\rho\vec f

which is the famous Euler equation for ideal fluid.

Using the equation of continuity and the Euler equation, we can calculate the time derivative of mass flux

t(ρvi)=viρt+ρvit=vij(ρvj)ρvjjvijδijp=j(ρvivj+δijp)jΠij\begin{align*} \frac{\partial}{\partial t}\left(\rho v^i\right)&=v^i\frac{\partial\rho}{\partial t}+\rho\frac{\partial v^i}{\partial t}\\ &=-v^i\partial_j\left(\rho v^j\right)-\rho v^j\partial_jv^i-\partial_j\delta^{ij}p\\ &=-\partial_j\left(\rho v^iv^j+\delta^{ij}p\right)\\ &\equiv -\partial_j\Pi^{ij} \end{align*}

So in the Eulerian picture, the momentum conservation is

t(ρvi)+jΠij=0\frac{\partial}{\partial t}\left(\rho v^i\right)+\partial_j\Pi^{ij}=0

Integrate the above equation over a fixed volume $V$

ddtVρvidV=ipdVj(ρvivj)dV=SpnidSSρvivjnjdS\begin{align*} \frac{\text d}{\text dt}\iiint_V\rho v^i\text dV&=-\iiint\partial^i p\text dV-\iiint\partial_j\left(\rho v^iv^j\right)\text dV\\ &=-\oiint_Spn^i\text dS-\oiint_S\rho v^iv^jn_j\text dS \end{align*}
  • First term - thermal pressure

  • Second term - ram pressure

Equation of State: $f(\rho,p,T)=0$

Two independent quantities are necessary to characterize thermal states.

  • Ideal gas

    p=ρkBTμmpp=\frac{\rho k_BT}{\mu m_p}
  • Adiabatic gas

    p=Kργ,γ=cpcVp=K\rho^\gamma,\quad \gamma=\frac{c_p}{c_V}

    The adiabatic index $\gamma=5/3$ for mono-atomic gas.

Ideal + adiabatic is an example of polytropic EoS

If $p=f(\rho)$, the EoS is known as barotropic.

For inviscid fluid

DsDt=st+(v)s=0\frac{\text Ds}{\text Dt}=\frac{\partial s}{\partial t}+\left(\vec v\cdot \nabla\right)s=0

that is, entropy doesn't change along each fluid stream line.

Furthermore, if $s=const$ everywhere at $r=r_0$, and $s=const$ is preserved, the. motion is known as isentropic motion.

Bernoulli's theorem

In general, we would like to derive the energy conservation with Euler equation with certain assumptions

  • Barotropic EoS: $\rho=f(p)$, thus the specific enthalpy is simply

    h(p)dpρ(p)h(p)\equiv\int\frac{\text dp}{\rho(p)}
  • Potential force: $\vec f=-\rho\nabla \phi$

Then we can rewrite Euler equation

vt+(v)v=1ρpϕ\frac{\partial \vec v}{\partial t}+\left(\vec v\cdot \nabla\right)\vec v=-\frac1\rho\nabla p-\nabla\phi
vt+(12v2+h+ϕ)=v×ω\Rightarrow \frac{\partial \vec v}{\partial t}+\nabla\left(\frac12v^2+h+\phi\right)=\vec v\times\vec\omega

where the vorticity is defined as

ω=×v\vec\omega=\nabla\times\vec v

and we have applied the identity

v×(×v)=(v)v(v)v=12v2(v)v\vec v\times\left(\nabla\times\vec v\right)=\left(\nabla\cdot\vec v\right) \vec v-\left(\vec v\cdot\nabla\right) \vec v=\frac12\nabla v^2-\left(\vec v\cdot\nabla\right) \vec v
  • For steady flow

    ut=0\frac{\partial u}{\partial t}=0

    thus

    v(12v2+h+ϕ)=v(v×ω)=0\vec v\cdot\nabla\left(\frac12v^2+h+\phi\right)=\vec v\cdot\left(\vec v\times\vec\omega\right)=0

    which means $\frac12v^2+h+\phi$ is conserved along each streamline.

  • No vorticity

    ×v=0v=ψ\nabla\times \vec v=0\Rightarrow \vec v=-\nabla\psi

    thus

    ψt+12v2+h+ϕ=F(t)\frac{\partial\psi}{\partial t}+\frac12v^2+h+\phi=F(t)

    and $\nabla F(t)=0$.

  • Steady & no vorticity

    ψt=0\frac{\partial \psi}{\partial t}=0

    Then $\frac12v^2+h+\phi$ remains a constant everywhere.

Energy Conservation

For fluid dynamics, the energy change in a unit time is given by

DDtV(t)ρ(e+12v2)dV=V(t)fvdV+S(t)tvdS\frac{\text D}{\text D t}\iiint_{V(t)}\rho\left(e+\frac12 v^2\right)\text dV=\iiint_{V(t)}\vec f\cdot\vec v\text dV+\oiint_{S(t)}\vec t\cdot\vec v\text dS

where $e$ is the specific internal energy.

If $T^{ij}$ is simply given by $-p\delta^{ij}$, the second term

S(t)tvdS=V(t)(pv)dV\oiint_{S(t)}\vec t\cdot\vec v\text dS=-\iiint_{V(t)}\nabla\cdot\left(p\vec v\right)\text dV

Again with Reynolds transport theorem,

V(t)[ρDDt(e+12v2)+(pv)]dV=V(t)fvdV\iiint_{V(t)}\left[\rho\frac{\text D}{\text Dt}\left(e+\frac12 v^2\right)+\nabla\cdot\left(p\vec v\right)\right]\text dV=\iiint_{V(t)}\vec f\cdot \vec v\text dV
ρDDt(e+12v2)+(pv)=fv\Rightarrow \rho\frac{\text D}{\text Dt}\left(e+\frac12 v^2\right)+\nabla\cdot\left(p\vec v\right)=\vec f\cdot \vec v
    DDt[ρ(e+12v2)](e+12v2)DρDt+(pv)=fv\iff \frac{\text D}{\text Dt}\left[\rho\left( e+\frac12 v^2\right)\right]-\left(e+\frac12 v^2\right)\frac{\text D\rho}{\text Dt}+\nabla\cdot\left(p\vec v\right)=\vec f\cdot \vec v
    t[ρ(e+12v2)]+(v+v)[ρ(e+12v2)]+(pv)=fv\iff \frac{\partial}{\partial t}\left[\rho\left( e+\frac12 v^2\right)\right]+\left(\nabla\cdot \vec v+\vec v\cdot \nabla\right)\left[\rho\left( e+\frac12 v^2\right)\right]+\nabla\cdot\left(p\vec v\right)=\vec f\cdot \vec v
    t[ρ(e+12v2)]+[ρv(e+12v2+pρ)]=fv\iff \frac{\partial}{\partial t}\left[\rho\left( e+\frac12 v^2\right)\right]+\nabla\cdot\left[\rho\vec v\left( e+\frac12 v^2+\frac{p}{\rho}\right)\right]=\vec f\cdot \vec v

which is known as the energy equation.

We can integrate and expand the second term so that

V[ρv(e+12v2+pρ)]dV=ρvn(e+12v2)dS+pvndS\iiint_V \nabla\cdot\left[\rho\vec v\left( e+\frac12 v^2+\frac{p}{\rho}\right)\right]\text dV=\oiint\rho v_{\vec n}\left( e+\frac12 v^2\right)\text dS+\oiint pv_{\vec n}\text dS

These two terms correspond to the energy flux and the work, respectively.

  • Here,

    h=e+pρh=e+\frac p\rho

    is the specific enthalpy.

  • If $\vec f$ is given by a static potential as

    f=ρϕ,ϕt=0\vec f=-\rho\nabla\phi,\quad \frac{\partial \phi}{\partial t}=0

    we can further write the energy equation as

    t[ρ(e+12v2+ϕ)]+[ρv(12v2+h+ϕ)]=0\frac{\partial}{\partial t}\left[\rho\left( e+\frac12 v^2+\phi\right)\right]+\nabla\cdot\left[\rho\vec v\left(\frac12 v^2+h+\phi\right)\right]=0

Kinetic Energy

We derive the equation of the kinetic energy from Euler equation,

v[vt+(v)v+1ρp1ρf]=0\vec v\cdot\left[\frac{\partial \vec v}{\partial t}+\left(\vec v\cdot \nabla\right)\vec v+\frac1\rho\nabla p-\frac1\rho\vec f\right]=0
    ρ(t+v)(12v2)+(v)p=fv\iff \rho\left(\frac{\partial}{\partial t}+\vec v\cdot\nabla\right)\left(\frac12v^2\right)+\left(\vec v\cdot\nabla\right) p=\vec f\cdot\vec v
    ρDDt(12v2)+(v)p=fv\iff \rho\frac{\text D}{\text Dt}\left(\frac12v^2\right)+\left(\vec v\cdot\nabla\right) p=\vec f\cdot\vec v

Thermal Energy (Total - Kinetic)

ρDDt(e+12v2)ρDDt(12v2)+(pv)(v)p=0\rho\frac{\text D}{\text Dt}\left(e+\frac12 v^2\right)-\rho\frac{\text D}{\text Dt}\left(\frac12v^2\right)+\nabla\cdot\left(p\vec v\right)-\left(\vec v\cdot\nabla\right) p=0
    ρDeDt+(v)p=0\iff \rho\frac{\text De}{\text Dt}+\left(\nabla\cdot\vec v\right) p=0

Since

DDt(1ρ)=1ρv\frac{\text D}{\text Dt}\left(\frac1\rho\right)=\frac1\rho\nabla\cdot\vec v

We have

DeDt+pDDt(1ρ)=0\frac{\text De}{\text Dt}+p\frac{\text D}{\text Dt}\left(\frac1\rho\right)=0

Recall that the first law of thermaldynamics gives

DeDt=pDDt(1ρ)+TDsDt\frac{\text De}{\text Dt}=-p\frac{\text D}{\text Dt}\left(\frac1\rho\right)+T\frac{\text Ds}{\text Dt}

thus

DsDt=0\frac{\text Ds}{\text Dt}=0

which suggests adiabatic motion of the fluid. In this way, as long as we assume

Tij=pδijT^{ij}=-p\delta^{ij}

we obtain ideal fluid.

Viscosity & Navier-Stokes Equation

In general, the stress tensor is not diagonal,

Tij=pδij+σijT^{ij}=-p\delta^{ij}+\sigma^{ij}

where $\sigma$ is the viscous tensor.

The $i$-th component of the surface force is thus

FSi=TijnjdS=pnidS+σijnjdSF_S^i=\oiint T^{ij}n_j\text dS=-\oiint pn^i\text dS+\oiint\sigma^{ij}n_j\text dS

and the force vector is

FS=V(σp)dV\vec F_S=\iiint_V\left(\nabla\cdot \sigma-\nabla p\right)\text dV

We claim that $\mathcal T$ is symmetric, that is, $T^{ij}=T^{ji}$. Consider a box with $\delta V=\delta x\delta y\delta z$. In the $z$-direction, the net torque caused leads to the change in angular momentum with respect to the center of $\delta S_{x,y}$

dLzdt=(r×F)z\frac{\text dL_z}{\text dt}=\left(\vec r\times \vec F\right)_z
ρδVδx2+δy26ω˙z(Tx,yTy,x)δV\rho\delta V\cdot\frac{\delta x^2+\delta y^2}{6}\dot\omega_z\sim\left(T_{x,y}-T_{y,x}\right)\delta V

for $\delta x\to0,\delta y\to0$, we have $T{x,y}=T{y,x}$, otherwise the $\dot \omega$ goes to infinite.

Now that we know $\mathcal T$ is symmetric, it has 6 independent components.

Viscous Tensor for Fluid

In a flow with $v_x$ that has a gradient of $\text dv_x/\text dy$, the stress to the $x$-axis is well approximated as

σxy=ηdvxdy\sigma_{xy}=\eta\frac{\text d v_x}{\text dy}

where $\eta$ is the (dynamic) viscosity determined by microscopic processes of particles. Fluid that satisfies such approximation is known as Nowton fluid.

In astrophysics, a more frequently applied viscosity is the kinematic viscosity $\nu$, defined as

ν=ηρ\nu=\frac{\eta}{\rho}

$\nu$ in a microscopic perspective

Consider a flow with coherent bulk velocity $v_x(y)$ along the $x$-axis, we try to count how much momentum is exchanged at the surface of $y=y_0$ per unit area per unit time, that is, the shear stress.

For simplicity, all we need to consider is a thin layer within $[y-l\text{mfp},y+l\text{mfp}]$, since we assume mometum is fully exchanged in the first collision. In this way, statistically, particles outside this layer cannot transport momentum to the surface $y=y_0$.

The number flux of particles hitting $y=y_0$ from one side can be estimated as

N=13ncˉ×12N=\frac13 n\bar c\times\frac12

where $n$ is the number density of particles and $\bar c$ is the average velocity dispersion ($\sim c_s$). We introduce a factor of $1/3$ because particles move in all three dimensions and $1/2$ further distinguish motion upwards and downwards. As a result, the shear stress is given by

σxy=16ncˉm[vx(y0+lmfp)vx(y0lmfp)]=13ρcˉlmfpvx(y0)x\sigma_{xy}=\frac16n\bar cm\left[v_x(y_0+l_\text{mfp})-v_x(y_0-l_\text{mfp})\right]=\frac13\rho\bar cl_\text{mfp}\frac{\partial v_x(y_0)}{\partial x}

Thus

ν=13cˉlmfp\nu=\frac13\bar cl_\text{mfp}

Unfortunately, momentum transition through collisions of particles is not the main cause of viscosity in astrophysics, and this formula is useless... In fact we can estimate the viscous timescale, in which viscosity drives materials to fall into the accretor.

tvis=r2νr2cˉlmfpt_\text{vis}=\frac{r^2}{\nu}\sim\frac{r^2}{\bar cl_\text{mfp}}

For the supermassive black hole (SMBH), this assumption gives a $t\text{vis}$ if 10 Gyr, which is orders of magtitude longer than the typical AGN timescale ($\sim$ Myr). At present, it is widely agreed that the Magneto-Rotational Instability (MRI) causes the viscosity in accretion disks. In this case, the $l\text{mfp}$ is given by the thickness of the disk $H\sim0.01 r$, and the viscous timescale is quite satistying.

Construction of Stress Tensor

  1. $\sigma{ij}$ is symmetric since $T{ij}$ is symmetric, but the original definition.

    σij=ηvixj\sigma_{ij}=\eta\frac{\partial v_i}{\partial x^j}

    is not. In this way, we expand it into symmetric and anti-symmetric parts, as we can do to any second-order tensor

    σij=η2(vixj+vjxi)+η2(vixjvjxi)\sigma_{ij}=\frac{\eta}{2}\left(\frac{\partial v_i}{\partial x^j}+\frac{\partial v_j}{\partial x^i}\right)+\frac{\eta}{2}\left(\frac{\partial v_i}{\partial x^j}-\frac{\partial v_j}{\partial x^i}\right)

    The second (anti-symmetric) term corresponds to the rotation. It has no viscosity, so we will consider a linear relation between $\sigma{ij}$ and $E{ij}$, which is defined as

    Eij12(vixj+vjxi)E_{ij}\equiv\frac{1}{2}\left(\frac{\partial v_i}{\partial x^j}+\frac{\partial v_j}{\partial x^i}\right)
  2. Fluid doesn't have a preferred direction.

    Intuitionally, we write down the (multi-)linear relation as

    σij=αijklEkl\sigma_{ij}=\alpha_{ijkl}E^{kl}

    We don't have any information of $\alpha$, but since fluid is isotropic, we would like to believe the tensor itself is isotropic, which means it has the same components in all rotated coordinate systems. There are only three basic 4-order isotropic tensors, all of which are the combination of the famous Kronecker delta $\delta_{ij}$.

    Aijkl=δijδkj, δikδjl, δilδjkA_{ijkl}=\delta_{ij}\delta_{kj},\ \delta_{ik}\delta_{jl},\ \delta_{il}\delta_{jk}

    So we assume $\alpha_{ijkl}$ is simply their linear combination

    σij=(Bδijδkj+Cδikδjl+Dδilδjk)Ekl=BδijEkk+CEij+DEji=B(v)δij+(C+D)Eij\begin{align*} \sigma_{ij}&=\left(B\delta_{ij}\delta_{kj}+C\delta_{ik}\delta_{jl}+D\delta_{il}\delta_{jk}\right)E^{kl}\\ &=B\delta_{ij}E^k_k+CE_{ij}+DE_{ji}\\ &=B\left(\nabla\cdot \vec v\right)\delta_{ij}+(C+D)E_{ij} \end{align*}
  3. Decompose $\sigma_{ij}$ into trace part and traceless part

    tr(σ)=δijσij=(3B+C+D)v\text{tr}\left(\sigma\right)=\delta^{ij}\sigma_{ij}=(3B+C+D)\nabla\cdot \vec v

    Thus

    σijtr=13(3B+C+D)(v)δij\sigma_{ij}^{\text{tr}}=\frac{1}3(3B+C+D)\left(\nabla\cdot \vec v\right)\delta_{ij}
    σijtrf=C+D3(v)δij+C+D2(vixj+vjxi)\sigma_{ij}^\text{trf}=-\frac{C+D}3\left(\nabla\cdot \vec v\right)\delta_{ij}+\frac{C+D}2\left(\frac{\partial v_i}{\partial x^j}+\frac{\partial v_j}{\partial x^i}\right)
  4. Redefine $\eta$ and $\zeta$ as

    η12(C+D),ζ13(3B+C+D)\eta\equiv \frac12(C+D),\quad\zeta\equiv\frac13(3B+C+D)

    Finally, with only two free parameters, we may obtain

    σij=η[vixj+vjxi23(v)δij]+ζ(v)δij\sigma_{ij}=\eta\left[\frac{\partial v_i}{\partial x^j}+\frac{\partial v_j}{\partial x^i}-\frac23\left(\nabla\cdot \vec v\right)\delta_{ij}\right]+\zeta\left(\nabla\cdot \vec v\right)\delta_{ij}
    • First term - traceless, pure shear viscosity

    • Second term - bulk viscosity

      When Stokes first derived this formula, he believed the bulk viscosity vanished. In fact this term is usually close to 0.

Right now,

Tij=pδij+σijT^{ij}=-p\delta^{ij}+\sigma^{ij}

and we can rewrite the momentum tensor by adding the viscous tensor

Πij=ρvivj+δijpσij\Pi^{ij}=\rho v^iv^j+\delta^{ij}p-\sigma^{ij}

From this correction, the conservation of momentum (Euler equation) gives

DvDt=1ρf+1ρT=1ρp+1ρf+1ρσ\frac{\text D\vec v}{\text Dt}=\frac1\rho\vec f+\frac1\rho\nabla\cdot\mathcal T=-\frac1\rho\nabla p+\frac1\rho\vec f+\frac1\rho\nabla\cdot\sigma

Since

(σ)i=ηj[jvi+ivj23kvkδij]+ζjkvkδij=η[2vi+13i(v)]+ζi(v)\begin{align*} (\nabla\cdot\sigma)_i&=\eta\partial^j\left[\partial_jv_i+\partial_iv_j-\frac23\partial_kv^k\delta_{ij}\right]+\zeta\partial^j\partial_kv^k\delta_{ij}\\ &=\eta\left[\nabla^2v_i+\frac13\partial_i(\nabla\cdot\vec v)\right]+\zeta\partial_i(\nabla\cdot\vec v) \end{align*}
    σ=η2v+(ζ+13η)(v)\iff \nabla\cdot\sigma=\eta\nabla^2\vec v+\left(\zeta+\frac13\eta\right)\nabla\left(\nabla\cdot\vec v\right)

Finally we derive the Navier-Stokes equation

vt+(v)v=1ρp+1ρf+1ρη2v+1ρ(ζ+13η)(v)\frac{\partial \vec v}{\partial t}+\left(\vec v\cdot \nabla\right)\vec v=-\frac1\rho\nabla p+\frac1\rho\vec f+\frac1\rho\eta\nabla^2\vec v+\frac1\rho\left(\zeta+\frac13\eta\right)\nabla\left(\nabla\cdot\vec v\right)

In the end, let us consider the energy conservation in the viscous situation

DDtV(t)ρ(e+12v2)dV=V(t)fvdV+S(t)tvdS\frac{\text D}{\text D t}\iiint_{V(t)}\rho\left(e+\frac12 v^2\right)\text dV=\iiint_{V(t)}\vec f\cdot\vec v\text dV+\oiint_{S(t)}\vec t\cdot\vec v\text dS
    V(t)ρDDt(e+12v2)dV=V(t)fvdV+V(t)(Tv)dV\iff \iiint_{V(t)}\rho\frac{\text D}{\text D t}\left(e+\frac12 v^2\right)\text dV=\iiint_{V(t)}\vec f\cdot\vec v\text dV+\iiint_{V(t)}\nabla\cdot\left(\mathcal T\cdot\vec v\right)\text dV
    ρDDt(e+12v2)=fv+(Tv)\iff\rho\frac{\text D}{\text D t}\left(e+\frac12 v^2\right)=\vec f\cdot\vec v+\nabla\cdot\left(\mathcal T\cdot\vec v\right)

By multiplying N-S equation with $\rho\vec v$, we have

ρDDt(12v2)=fv+v(T)\rho\frac{\text D}{\text Dt}\left(\frac12v^2\right)=\vec f\cdot\vec v+\vec v\cdot\left(\nabla\cdot\mathcal T\right)
ρDeDt=(T)v\Rightarrow \rho\frac{\text De}{\text D t}=\left(\mathcal T\cdot\nabla\right)\cdot \vec v

Now we expand the RHS

(T)v=(ivj)Tij=ivip+(ivj)σij=p(v)+η(ivj)[2Eij23(v)δij]+ζ(ivi)(v)=p(v)+η(ivj)(jvi+ivj)+(ζ23η)(v)2\begin{align*} \left(\mathcal T\cdot\nabla\right)\cdot \vec v&=(\partial_iv_j)T^{ij}\\ &=-\partial_iv^ip+(\partial_iv_j)\sigma^{ij}\\ &=-p(\nabla\cdot\vec v)+\eta(\partial_iv_j)\left[2E^{ij}-\frac23\left(\nabla\cdot \vec v\right)\delta_{ij}\right]+\zeta(\partial_iv^i)\left(\nabla\cdot \vec v\right)\\ &=-p(\nabla\cdot\vec v)+\eta(\partial_iv_j)(\partial^jv^i+\partial^iv^j)+\left(\zeta-\frac23\eta\right)\left(\nabla\cdot \vec v\right)^2 \end{align*}

where

(ivj)(jvi+ivj)=12(ivj+jvi)(jvi+ivj)12(ivjjvi)(jvi+ivj)=2EijEij\begin{align*} (\partial_iv_j)(\partial^jv^i+\partial^iv^j)&=\frac12(\partial_iv_j+\partial_jv_i)(\partial^jv^i+\partial^iv^j)\\ &-\frac12(\partial_iv_j-\partial_jv_i)(\partial^jv^i+\partial^iv^j)\\ &=2E_{ij}E^{ij} \end{align*}

as the second term obviously vanishes. Finally,

(T)v=p(v)+2ηEijEij+(ζ23η)(v)2\left(\mathcal T\cdot\nabla\right)\cdot \vec v=-p(\nabla\cdot\vec v)+2\eta E_{ij}E^{ij}+\left(\zeta-\frac23\eta\right)\left(\nabla\cdot \vec v\right)^2

Define

Φvis=2ηEijEij+(ζ23η)(v)2\Phi_\text{vis}=2\eta E_{ij}E^{ij}+\left(\zeta-\frac23\eta\right)\left(\nabla\cdot \vec v\right)^2

we have

ρ[DeDt+pDDt(1ρ)]=Φvis=ρTDsDt\rho\left[\frac{\text De}{\text D t}+p\frac{\text D}{\text Dt}\left(\frac1\rho\right)\right]=\Phi_\text{vis}=\rho T\frac{\text Ds}{\text Dt}

Thus,

DSDt=DDtVρsdV=VρDsDtdV=VΦvisTdV\frac{\text DS}{\text Dt}=\frac{\text D}{\text Dt}\iiint_{V}\rho s\text dV=\iiint_{V}\rho \frac{\text Ds}{\text Dt}\text dV=\iiint_{V}\frac{\Phi_\text{vis}}{T}\text dV

The second law of thermaldynamics tells us

DSDt0\frac{\text DS}{\text Dt}\ge0

In fact, we can prove

Φvis0\Phi_\text{vis}\ge0

The Reynolds Number

The Reynolds number is defined as

Re=ULν\text{Re}=\frac{UL}{\nu}

where $U$ is the flow speed, $L$ is a characteristic linear dimension, and $\nu$ is the kinematic viscosity. It is naturally connected with the Navier-Stokes equation when we try to derive a dimensionless form of the equation.

By assuming incompressible fluid ($\nabla\cdot \vec v=0$) and neglecting the external force, the Navier-Stokes equation goes like

vt+(v)v=1ρp+ν2v\frac{\partial \vec v}{\partial t}+\left(\vec v\cdot \nabla\right)\vec v=-\frac1\rho\nabla p+\nu\nabla^2\vec v

Define several dimensionless quantities,

r=rL,v=vU,t=tL/u,ρ=ρρ0,p=pρ0v2r'=\frac rL,\quad \vec v'=\frac{\vec v}{U},\quad t'=\frac{t}{L/u},\quad \rho'=\frac{\rho}{\rho_0},\quad p'=\frac{p}{\rho_0v^2}

We have

vt+(v)v=1ρp+νUL2v1ρp+1Re2v\frac{\partial \vec v'}{\partial t'}+\left(\vec v'\cdot \nabla'\right)\vec v'=-\frac1\rho'\nabla' p'+\frac\nu{UL}\nabla'^2\vec v'\equiv -\frac1\rho'\nabla' p'+\frac1{\text{Re}}\nabla'^2\vec v'

The Reynolds number is simply there.

If $\text{Re}\to\infty$, the equation goes back to the Euler equation (ideal fluid).

Small $\text{Re}$ corresponds to laminar flow, a fluid flowing in parallel layers with no disruption between the layers, while large $\text{Re}$ introduces annoying turbulence, a fluid motion characterized by chaotic changes in pressure and flow velocity. In astrophysics, $\text{Re}$ is usually large.

Notes

  1. Dimensionless Navier-Stokes equation does not depend on $\rho_0$ - scale-free. Once a hydrodynamical simulation is done with certain density, it is done with any other density. This beautiful scale-free law breaks when

    • $P-\rho$ relation is not a simple power-law, and has certain breaks.

    • the simulation has to take opacity/cooling/heating into account, none of which is scale-free.

  2. If we introduce gravity as the external force

    fext=GMr2r^\vec f_\text{ext}=-\frac{GM}{r^2}\hat{\vec r}

    the corresponding term in the dimensionless Naiver-Stokes equation is

    1LU2GMr2-\frac{1}{LU^2}\frac{GM}{r'^2}

    It seems to have dependence on $M$, however, if we set

    U=c,L=rgGMc2U=c,\quad L=r_\text{g}\equiv\frac{GM}{c^2}

    where $r_\text g$ is the Schiwarzchild radius. This term is simply

    1r2-\frac1{r'^2}

    Around a black hole, Navier-Stokes equation is again scale-free, this time for the mass of the accretor.

Last updated