Categories: Fluid dynamics, Fluid mechanics, Physics.

Navier-Stokes equations

While the Euler equations govern ideal “dry” fluids, the Navier-Stokes equations govern nonideal “wet” fluids, i.e. fluids with nonzero viscosity.

Incompressible fluid

First of all, we can reuse the incompressibility condition for ideal fluids, without modifications:

v=0\begin{aligned} \boxed{ \nabla \cdot \va{v} = 0 } \end{aligned}

Furthermore, from the derivation of the Euler equations, we know that Newton’s second law can be written as follows, for an infinitesimal particle of the fluid:

ρDvDt=f\begin{aligned} \rho \frac{\mathrm{D} \va{v}}{\mathrm{D} t} = \va{f^*} \end{aligned}

D/Dt\mathrm{D}/\mathrm{D}t is the material derivative, ρ\rho is the density, and f\va{f^*} is the effective force density, expressed in terms of an external body force f\va{f} (e.g. gravity) and the Cauchy stress tensor σ^\hat{\sigma}:

f=f+σ^\begin{aligned} \va{f^*} = \va{f} + \nabla \cdot \hat{\sigma}^\top \end{aligned}

From the definition of viscosity, the stress tensor’s elements are like so for a Newtonian fluid:

σij=pδij+η(ivj+jvi)\begin{aligned} \sigma_{ij} = - p \delta_{ij} + \eta (\nabla_i v_j + \nabla_j v_i) \end{aligned}

Where η\eta is the dynamic viscosity. Inserting this, we calculate σ^\nabla \cdot \hat{\sigma}^\top in index notation:

(σ^)i=jjσij=j( ⁣ ⁣δijjp+η(ijvj+j2vi))=ip+ηijjvj+ηjj2vi\begin{aligned} \big( \nabla \cdot \hat{\sigma}^\top \big)_i = \sum_{j} \nabla_j \sigma_{ij} &= \sum_{j} \Big( \!-\! \delta_{ij} \nabla_j p + \eta (\nabla_i \nabla_j v_j + \nabla_j^2 v_i) \Big) \\ &= - \nabla_i p + \eta \nabla_i \sum_{j} \nabla_j v_j + \eta \sum_{j} \nabla_j^2 v_i \end{aligned}

Thanks to incompressibility v=0\nabla \cdot \va{v} = 0, the middle term vanishes, leaving us with:

f=fp+η2v\begin{aligned} \va{f^*} = \va{f} - \nabla p + \eta \nabla^2 \va{v} \end{aligned}

We assume that the only body force is gravity f=ρg\va{f} = \rho \va{g}. Newton’s second law then becomes:

ρDvDt=ρgp+η2v\begin{aligned} \rho \frac{\mathrm{D} \va{v}}{\mathrm{D} t} = \rho \va{g} - \nabla p + \eta \nabla^2 \va{v} \end{aligned}

Dividing by ρ\rho, and replacing η\eta with the kinematic viscosity ν=η/ρ\nu = \eta/\rho, yields the main equation:

DvDt=gpρ+ν2v\begin{aligned} \boxed{ \frac{\mathrm{D} \va{v}}{\mathrm{D} t} = \va{g} - \frac{\nabla p}{\rho} + \nu \nabla^2 \va{v} } \end{aligned}

Finally, we can optionally allow incompressible fluids with an inhomogeneous “lumpy” density ρ\rho, by demanding conservation of mass, just like for the Euler equations:

DρDt=0\begin{aligned} \boxed{ \frac{\mathrm{D} \rho}{\mathrm{D} t} = 0 } \end{aligned}

Putting it all together, the Navier-Stokes equations for an incompressible fluid are given by:

DvDt=gpρ+ν2vv=0DρDt=0\begin{aligned} \boxed{ \frac{\mathrm{D} \va{v}}{\mathrm{D} t} = \va{g} - \frac{\nabla p}{\rho} + \nu \nabla^2 \va{v} \qquad \nabla \cdot \va{v} = 0 \qquad \frac{\mathrm{D} \rho}{\mathrm{D} t} = 0 } \end{aligned}

Due to the definition of viscosity ν\nu as the molecular “stickiness”, we have boundary conditions for the velocity field v\va{v}: at any interface, v\va{v} must be continuous. Likewise, Newton’s third law demands that the normal component of stress σ^n^\hat{\sigma} \cdot \vu{n} is continuous there.

References

  1. B. Lautrup, Physics of continuous matter: exotic and everyday phenomena in the macroscopic world, 2nd edition, CRC Press.