The Navier-Stokes Equations

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Chapter 4

The Navier-Stokes equations

In many engineering problems, approximate solutions concerning the overall properties of a fluid
system can be obtained by application of the conservation equations of mass, momentum and en-
ergy written in integral form, given above in (3.10), (3.35) and (3.46), for a conveniently selected
control volume. This approach necessitates in general introduction of simplifying assumptions,
regarding in particular the spatial distributions of the different variables (e.g., uniform velocity
at the open boundaries) and the neglect of terms that are anticipated to give a relatively small
contribution to the overall balances. This integral approach is however not suitable when one is
interested in computing local properties of the flow (e.g., distributions of velocity v̄(x̄, t), density
ρ(x̄, t), pressure p(x̄, t), etc). For that purpose, the conservation equations in integral form
"Z #
d
ρdV = 0, (4.1)
dt Vf (t)
"Z # Z Z Z
d
ρv̄dV =− pn̄dσ + τ̄¯′ · n̄dσ + ρf¯m dV, (4.2)
dt Vf (t) Σf (t) Σf (t) Vf (t)
"Z # Z Z
d 2
ρ(e + |v̄| /2)dV =− pv̄ · n̄dσ + v̄ · τ̄¯′ · n̄dσ
dt Vf (t) Σf (t) Σf (t)

(4.3)
Z Z Z
+ ρf¯m · v̄dV − q̄ · n̄dσ + (Qc + Qr )dV,
Vf (t) Σf (t) Vf (t)

need to be transformed with use made of the Gauss formula to express the rate balances occuring
locally, providing a set of partial differential equations to be integrated with boundary and initial
conditions.

The continuity equation


Using Gauss formula (2.9) it is straightforward to rewrite (4.1) in the form
Z  
∂ρ
+ ∇ · (ρv̄) dV = 0. (4.4)
Vf ∂t
The Navier-Stokes equations

This equation is to be satisfied regardless of the choice of Vf if and only if the identity

∂ρ
+ ∇ · (ρv̄) = 0 (4.5)
∂t
holds at each point of space. This is the continuity (or mass conservation) equation, stating that
the sum of the rate of local density variation and the rate of mass loss by convective outflow
equals zero. An alternative expression is
1 Dρ
= −∇ · v̄, (4.6)
ρ Dt
indicating that the variation of density following the fluid particle is exclusively due to the rate
of volume variation. Note that for an incompressible (constant-density) fluid, the continuity
equation reduces to
∇ · v̄ = 0, (4.7)
a result anticipated previously, whereas for steady gas flow one obtains

∇ · (ρv̄) = 0, (4.8)

that is, the convective rate of mass loss per unit volume is zero.

The momentum equation


The derivation of the momentum equation in differential form follows the procedure used before
to obtain (4.5), that is, Gauss formula (2.9) is used to rewrite all of the surface integrals appearing
in (4.2) as volume integrals and the resulting integrand is set equal to zero to yield


(ρv̄) + ∇ · (ρv̄v̄) = −∇p + ∇ · τ̄¯′ + ρf¯m . (4.9)
∂t
Using now (4.5) and (2.26) provides
 
Dv̄ ∂v̄
ρ =ρ + ∇(|v̄| /2) − v̄ ∧ (∇ ∧ v̄) = −∇p + ∇ · τ̄¯′ + ρf¯m ,
2
(4.10)
Dt ∂t

which is simply Newton’s second law expressed per unit volume of fluid, a result anticipated
in (3.25).
For an incompressible fluid with constant viscosity ∇ · τ̄¯′ = −µ∇ ∧ (∇ ∧ v̄), so that (4.10) reduces
to  
Dv̄ p
= −∇ − ν∇ ∧ (∇ ∧ v̄) + f¯m . (4.11)
Dt ρ
The viscous force per unit volume can be alternatively expressed as1 −µ∇ ∧ (∇ ∧ v̄) = µ∇2 v̄,
yielding  
Dv̄ p
= −∇ + ν∇2 v̄ + f¯m , (4.12)
Dt ρ
1
It is to be noted that in cartesian coordinates each component of the vector ∇2 v̄ is given simply by the
Laplacian of the corresponding velocity component, whereas additional terms appear in cylindrical and spherical
coordinates, thereby complicating the computation.

38
The Navier-Stokes equations

Note that (4.7) and (4.11) suffice to determine the velocity and pressure fields for an incompress-
ible flow with constant viscosity. For such flows, which include those involving water, these two
equations are therefore decoupled from the energy equation, which could be used a posteriori to
determine the temperature field. Also of interest is that, when the mass force is conservative,
introduction of f¯m = −∇U enables (4.11) to be written in the form
 
Dv̄ p + ρU
= −∇ − ν∇ ∧ (∇ ∧ v̄). (4.13)
Dt ρ
This last expression suggests that, for the motion of perfect liquids, introduction of the reduced
pressure P = p + ρU (e.g., P = p + ρgz when f¯m = ḡ) may simplify the description significantly.
For instance, when the pressure does not enter in the boundary conditions, the solution described
with use made of P becomes entirely independent of the mass forces.
A well-known result arises for steady incompressible flows with conservative mass forces when the
flow conditions are such that the effect of viscosity is negligible, so that the momentum equation
reduces to  
p |v̄|2
∇ +U + − v̄ ∧ (∇ ∧ v̄) = 0 (4.14)
ρ 2
as can be seen by using (4.13) with ν = 0 together with
Dv̄
= ∇(|v̄|2 /2) − v̄ ∧ (∇ ∧ v̄), (4.15)
Dt
the corresponding steady form of (2.26). The vector equation (4.14) can be projected along
streamlines by multiplying at each point in space by the unit vector v̄/|v̄|. Since the contribution
of the last term is identically zero, the projection reduces to
 
∂ p |v̄|2
+U + = 0, (4.16)
∂l ρ 2

where ∂/∂l = |v̄|−1 (v̄ · ∇) denotes the derivative along a stream line. It then follows that for
frictionless steady flow of an incompressible fluid the quantity

p + ρU + ρ|v̄|2 /2 = Cl (4.17)

remains constant along any given stream line, with a value Cl that in general is different for
different stream lines. Equation (4.17) is the so-called Bernoulli’s theorem, that provides a
useful relationship between the variations of kinetic energy, potential energy and pressure along
streamlines. Note that, if the flow is irrotational, the constant Cl is the same for all streamlines,
because ∇(p/ρ + U + |v̄|2 /2) = 0, as follows from (4.14).

The energy equation


Internal energy and kinetic energy conservation equations
The same transformation procedure used above in deriving (4.5) and (4.10) can be employed to
derive from (4.3)

[ρ(e + |v̄|2 /2)] + ∇ · [ρ(e + |v̄|2 /2)v̄] = −∇ · (pv̄) + ∇ · (v̄ · τ̄¯′ ) + ρf¯m · v̄ − ∇ · q̄ + Qc + Qr , (4.18)
∂t

39
The Navier-Stokes equations

which can be rewritten by virtue of (4.5) in the simplified form

D
ρ (e + |v̄|2 /2) = −∇ · (pv̄) + ∇ · (v̄ · τ̄¯′ ) + ρf¯m · v̄ − ∇ · q̄ + Qc + Qr . (4.19)
Dt

If the mass force derives from a steady potential such that f¯m = −∇U with ∂U/∂t = 0, then
the above equation admits the alternative writing

D
ρ (e + |v̄|2 /2 + U ) = −∇ · (pv̄) + ∇ · (v̄ · τ̄¯′ ) − ∇ · q̄ + Qc + Qr . (4.20)
Dt

A separate equation for the mechanical (kinetic) energy |v̄|2 /2 can be derived by taking the dot
product of (4.10) and v̄. Since v̄ · [v̄ ∧ (∇ ∧ v̄)] = 0, the product provides the scalar equation

D
ρ (|v̄|2 /2) = −v̄ · ∇p + v̄ · (∇ · τ̄¯′ ) + ρv̄ · f¯m , (4.21)
Dt

which is a purely mechanical energy law. As can be seen, the rate of variation of the kinetic
energy following the fluid particle equals the work rate of the mass force ρv̄ · f¯m plus the work
rate of the surface forces associated with the translational motion of the fluid particle v̄ · (∇ · τ̄¯) =
v̄ · (−∇p + τ̄¯′ ). Subtracting now (4.21) from (4.19) yields

De
ρ = −p∇ · v̄ + τ̄¯′ : ∇v̄ − ∇ · q̄ + Qc + Qr , (4.22)
Dt

clearly indicating that the rate of variation of the internal energy is associated with heat addition
and with the deformation work rate of the surface forces. The term −p∇ · v̄ corresponds to the
compression work, which is a reversible contribution, in that when the fluid volume is reduced
(∇ · v̄ < 0) the compression work produces an increase in the internal energy; conversely, the
internal energy decreases for positive expansion rates ∇ · v̄ > 0. The term φv = τ̄¯′ : ∇v̄ is the
deformation work due to viscous forces. It is non-negative (φv ≥ 0)2 , and corresponds to the
rate of mechanical energy dissipation per unit volume and unit time. Note that both −p∇ · v̄
and τ̄¯′ : ∇v̄ are independent of the orientation, position and motion of the reference frame.

Integral balance equations for mechanical and internal energy


The integral conservation equation for the energy (4.3) is written for the combined contribution
(e + |v̄|2 /2). It is interesting to write separate integral equations for the mechanical and internal
energy. To that end, one may integrate (4.21) in the fluid volume Vf (t), bearing in mind the
equation
D ∂
ρ (|v̄|2 /2) = (ρ|v̄|2 /2) + ∇ · (ρv̄|v̄|2 /2) (4.23)
Dt ∂t
2
From the definition of contraction of two tensors Ā ¯ = P P A B it can be shown that
¯ : B̄
i j ij ij

¯ + (µB − 2 µ)(∇ · v̄)2 = 2µ [(γ11 − γ22 )2 + (γ11 − γ33 )2 + (γ22 − γ33 )2 + 6(γ 2 + γ 2 + γ 2 )] + µB (∇ · v̄)2 ,
¯ : T̄
φv = 2µT̄d d 12 13 23
3 3
thereby demonstrating that φv ≥ 0 provided µ > 0 and µB > 0.

40
The Navier-Stokes equations

and also (3.6), to give


"Z # Z Z Z
d 2
ρ|v̄| /2dV = − pv̄ · n̄dσ + ′
v̄ · τ̄¯ · n̄dσ + ρf¯m · v̄dV
dt Vf (t) Σf (t) Σf (t) Vf (t)

(4.24)
"Z Z #
− −p∇ · v̄dV + φv dV .
Vf (t) Vf (t)

Similarly, one may use (4.22) to derive


"Z # Z Z
d
ρedV = − q̄ · n̄dσ + (Qc + Qr )dV
dt Vf (t) Σf (t) Vf (t)
(4.25)
" Z Z #
+ − p∇ · v̄dV + φv dV .
Vf (t) Vf (t)

The way we have written (4.24) and (4.25) is intended to give a clear picture of the balances of
work and energy in the flow field. Thus, the work done by mass forces in the interior of the fluid
volume as well as that done by surface forces on its bounding surface (the right-hand-side terms
in the first line of (4.24)) are devoted to increase the net amount of mechanical energy contained
in the fluid volume. On the other hand, the heat added by conduction, radiation and chemical
reaction (the right-hand-side terms in the first line of (4.25)) is employed directly to increase
the total internal energy. Viscous dissipation as well as compression work, written for clarity
in separate lines in (4.24) and (4.25), are the mechanisms transforming mechanical energy into
internal energy, with only the latter being reversible.

Enthalpy conservation equation


Equations. (4.5) and (4.10) together with (4.19) are the so-called Navier-Stokes equations gov-
erning fluid motion. The energy equation admits alternative forms, that may be more convenient
than (4.19) for the analysis of specific problems. In particular, often in the analysis of gas flows
it is of interest to use the enthalpy h = e + p/ρ as a replacement for the internal energy e. In
the derivation, one must first use the continuity equation (4.5) to give
D ∂p
∇ · (pv̄) = ρ (p/ρ) − . (4.26)
Dt ∂t
Combining now this result with (4.19) yields
D ∂p
ρ (h + |v̄|2 /2) = + ∇ · (v̄ · τ̄¯′ ) + ρf¯m · v̄ − ∇ · q̄ + Qc + Qr , (4.27)
Dt ∂t
which can also be written with use made of (4.21) as a separate equation for the enthalpy
Dh Dp
ρ = + φv − ∇ · q̄ + Qc + Qr . (4.28)
Dt Dt
Note that, when mass forces are conservative and steady, Eq. (4.27) reduces to
D ∂p
ρ (h + |v̄|2 /2 + U ) = + ∇ · (v̄ · τ̄¯′ ) − ∇ · q̄ + Qc + Qr , (4.29)
Dt ∂t

41
The Navier-Stokes equations

indicating that, when heat addition, temporal pressure variations, and viscous forces are all
simultaneously negligible, the combination h + |v̄|2 /2 + U remains constant in the evolution of
each fluid particle.

Entropy conservation equation


It is also of interest to express the energy equation in terms of the entropy s. Since T ds =
dh − dp/ρ, it is straightforward to use (4.28) to write
Ds
ρT = φv − ∇ · q̄ + Qc + Qr , (4.30)
Dt
indicating that the entropy of a fluid particle increases due to viscous dissipation and heat
addition, while the rate of compression work −p∇·v̄ is reversible and therefore does not change the
entropy of the fluid particle. Equation (4.30) can be seen as an application of the second principle
of thermodynamics, enabling changes of entropy to be quantified for fluids out of equilibrium.
One may divide (4.30) by T and integrate the resulting equation in Vf (t) to give
"Z # Z Z
d
ρsdV = (φv /T )dV − [(q̄ · ∇T )/T 2 ]dV
dt Vf (t) Vf (t) Vf (t)

(4.31)
Z Z
− (q̄/T ) · n̄dσ + (Qc + Qr )/T dV,
Σf (t) Vf (t)

for the variation of the total entropy contained in a fluid volume. The term involving (Qc +
Qr )/T represents a volumetric source of entropy associated with chemical reaction and radiation.
On the Rother hand, q̄ · n̄/T is the entropy flux across the surface due to heat conduction, so
that − Σf (t) (q̄/T ) · n̄dσ is the corresponding rate of entropy gain. Viscous dissipation gives
R R
a nonnegative contribution Vf (t) (φv /T )dV . Finally, the term − Vf (t) [(q̄ · ∇T )/T 2 ]dV is an
additional source of entropy that appears due to conduction inside the control volume. This
contribution is also nonnegative, as can be seen with use of (3.44).

Mathematical description of fluid flows


It is a good time to summarize now what we have done so far to derive a rigorous mathematical
framework for the description of fluid-flow problems. The derivation given above began by
introducing the continuum hypothesis, which was instrumental in defining the concepts of density
ρ, velocity v̄, and internal energy e as continuous functions of the position x̄ and time t. Next,
by introducing the hypothesis of local thermodynamic equilibrium, the local values of all of the
remaining thermodynamic variables (temperature, pressure, enthalpy, etc) were automatically
defined in terms of ρ and e for the non-equilibrium states found in fluid mechanics problems.
This hypothesis enables, in particular, ρ and e to be replaced by p and T as fundamental
thermodynamic variables for the flow-field description. We have also assumed that the viscous
stress tensor τ̄¯′ and the heat-flux vector q̄ are linear isotropic functions of the rate of strain and the
temperature gradient, respectively. In the constitutive equations arising from this assumption
(Navier-Poisson and Fourier laws) there appear three transport coefficients µ, µB and k that
are assumed to be functions of the local thermodynamic state of the fluid. Although some

42
The Navier-Stokes equations

of the hypotheses introduced can be expected to fail under extreme conditions (rarefied flows,
nanofluidics applications, etc), the anticipated range of validity of the mathematical formulation
derived covers most applications of interest in engineering.

Summary of conservation equations, equations of state and constitutive equa-


tions
To solve a given flow problem, the conservation equations of continuity
∂ρ
+ ∇ · (ρv̄) = 0, (4.32)
∂t
momentum
Dv̄
ρ = −∇p + ∇ · τ̄¯′ + ρf¯m , (4.33)
Dt
and energy
De
ρ = −p∇ · v̄ + τ̄¯′ : ∇v̄ − ∇ · q̄ + Qc + Qr , (4.34)
Dt
must be supplemented with the equations of state

ρ = ρ(p, T ) and e = e(p, T ), (4.35)

together with the constitutive equations


2
τ̄¯′ = µ(∇v̄ + ∇v̄ T ) + (µB − µ)(∇ · v̄)¯Ī and q̄ = −k∇T (4.36)
3
and associated state functions

µ = µ(T ), µB = µB (T ) and k = k(T ). (4.37)

In particular, the state equations (4.35) reduce to

ρ = ρo and e = eo + cT (4.38)

for a perfect liquid and to


p
ρ= and e = eo + cv T (4.39)
Rg T
for a perfect gas.
The vector expressions presented earlier in Chapter 2 enable (4.32), (4.33) and (4.34) to be
expressed in any system of orthogonal curvilinear coordinates. For a perfect liquid with constant
viscosity, Eqs. (4.32), (4.33) and (4.34) admit the simplified form

∇ · v̄ = 0 (4.40)

Dv̄
ρ = −∇p − µ∇ ∧ (∇ ∧ v̄) + ρf¯m (4.41)
Dt
DT ¯ : T̄
¯ + k∇2 T + Q + Q ,
ρc = 2µT̄ d d c r (4.42)
Dt
which are written out below for cartesian, cylindrical and spherical coordinates.

43
The Navier-Stokes equations

Initial and boundary conditions

Equations (4.32), (4.33) and (4.34) describe all fluid motions (within the range of validity of the
various assumptions introduced in their derivation). The difference between two flow problems
lies in their initial and boundary conditions, which need to be specified to enable the integration
of the Navier-Stokes equations to be performed.
For unsteady fluid motion, at the initial instant we must define the velocity field v̄ = v̄o (x̄)
together with the thermodynamic state, given for instance in terms of p = po (x̄) and T = To (x̄)
(or any other pair of independent thermodynamic variables). For an incompressible fluid, the
initial velocity field must necessarily satisfy ∇ · v̄o = 0. Because of the presence of the term
∂ρ/∂t in (4.32), in the case of gas flow there is no such restriction. Clearly, no initial conditions
are required for the study of steady or periodic fluid flows.
The Navier-Stokes equations determine the flow evolution within the flow field, and must be
therefore complemented with appropriate boundary conditions on the boundaries of the flow
field. In typical configurations, the fluid may be confined by solid walls, as occurs in wall-
bounded flows. The fluid velocity in contact with the wall will be assumed to be equal to the
wall velocity, i.e., at x̄ = x̄p we impose v̄ = v̄p , which reduces to v̄ = 0 when the wall is at rest
relative to the reference frame used in the description. This so-called non-slip condition comes
from assuming that the interaction of the molecules near the wall with the wall is equivalent
to that occurring between neighboring particles. For the velocity field to be continuous, no
velocity discontinuity is permitted between the wall and the adjacent fluid. There exists ample
experimental evidence confirming the validity of the non-slip condition, at least for flows in which
the mean free path λ is much shorter than the macroscopic length of the flow field L. Note that,
when this is not the case, the non-slip condition must be relaxed, along with the condition of
local thermodynamic equilibrium, thereby complicating the associated analysis.
We shall also assume that there exists local thermodynamic equilibrium at the wall, so that the
fluid temperature in contact with the wall is that of the wall (T = Tp at x̄ = x̄p ). There must also
exist an equilibrium of the heat fluxes at the wall surface, providing an additional equation that
reduces to q̄ · n̄ = ∂T /∂n = 0 at x̄ = x̄p when the wall is thermally insulated (adiabatic wall). In
principle, in integrating the conservation equations for the fluid, one could use alternatively the
temperature distribution Tp or the heat flux ∂T /∂n as boundary condition for the temperature
at the wall x̄ = x̄p . In many problems, however, neither quantity is known a priori, and they
have to be determined as part of the solution, which involves the integration of the Navier-Stokes
equations for the fluid together with the heat conduction equation in the solid, coupled with the
conditions of thermodynamic equilibrium and heat flux balance at the wall surface x̄ = x̄p .
The boundary conditions needed when the fluid extends to infinity (i.e., to distances much larger
than the characteristic macroscopic size of the flow field L), as occurs in external aerodynamic
problems, include specification of the boundary velocity field and its accompanying thermody-
namic state. These external boundary conditions can be written in the form v̄ = v̄∞ (x̄, t),
p = p∞ (x̄, t), and T = T∞ (x̄, t), for |x̄| → ∞. For instance, to study the motion of a uniform
stream of velocity U∞ over a solid body of characteristic size L, we shall impose v̄ = U∞ ēx ,
p = p∞ , and T = T∞ at |x̄|/L ≫ 1.
As an example, consider the motion a liquid stream that moves with velocity U∞ (t)ēx over a
spherical body of radius R. If the density and transport properties of the fluid are constant
and the effect of radiation and chemical reaction are negligible, the problem reduces to that of

44
The Navier-Stokes equations

integrating

∇ · v̄ = 0 (4.43)
∂v̄
ρ + ρv̄ · ∇v̄ = −∇p + µ∇2 v̄ − ρgēz (4.44)
∂t
∂T ¯ : T̄
¯ + k∇2 T.
ρc + ρcv̄ · ∇T = 2µT̄ d d (4.45)
∂t
Boundary conditions must be imposed on the surface of the body

|x̄| = R : v̄ = T − Tw = 0, (4.46)

where Tw is the temperature on the surface of the body, different from the value T∞ of the free
stream. Far from the body the velocity should approach that of the free stream whereas the
pressure equilibrates the gravity and inertial forces according to
dU∞
|x̄| ≫ R : v̄ = U∞ (t)ēx , T = T∞ , p + ρgz + ρ x = constant. (4.47)
dt
If the fluid is initially at rest (i.e., U = 0 for t ≤ 0), appropriate initial conditions are

Tw − T∞
t = 0 : v̄ = 0, T = T∞ + , (4.48)
|x̄|/R

where the temperature field is that obtained for a fluid at rest through solution of the corre-
sponding reduced energy equation ∇2 T = 0 with account taken of the spherical symmetry.

U z
8

y
x
T Tw
8

45
The Navier-Stokes equations

Stress-tensor components for a newtonian fluid


The components of the viscous stress tensor can be determined from (3.31) to give
2
τij′ = 2µγij + (µB − µ)∇ · v̄δij ,
3
¯ and δ represents
where γij = [(∇v̄)ij + (∇v̄ T )ij ]/2 is the ij component of the rate-of-strain tensor T̄ d ij
the Kronecker delta (δii = 1, δij = 0 if i 6= j). From the expressions derived above in Chapter 2, it can
be shown that
   
1 ∂vi X vk ∂hi hj ∂ vj hi ∂ vi
γii = + and γij = + (if i 6= j).
hi ∂xi hi hk ∂xk 2hi ∂xi hj 2hj ∂xj hi
k6=i

Using this last equation together with


 
1 ∂ ∂ ∂
∇ · v̄ = (h2 h3 v1 ) + (h1 h3 v2 ) + (h1 h2 v3 ) ,
h1 h2 h3 ∂x1 ∂x2 ∂x3
one may write expressions for τij′ = τji

in different orthogonal coordinate systems.

∂vx ∂vy ∂vz


Cartesian coordinates (∇ · v̄ = ∂x + ∂y + ∂z )

 
′ ∂vx 2 ′ ∂vx ∂vy
τxx = 2µ + (µB − µ)(∇ · v̄) τxy =µ +
∂x 3 ∂y ∂x
 
′ ∂vy 2 ′ ∂vy ∂vz
τyy = 2µ + (µB − µ)(∇ · v̄) τyz =µ +
∂y 3 ∂z ∂y
 
′ ∂vz 2 ′ ∂vx ∂vz
τzz = 2µ + (µB − µ)(∇ · v̄) τzx =µ +
∂z 3 ∂z ∂x

1 ∂rvr 1 ∂vθ ∂vz


Cylindrical coordinates (∇ · v̄ = r ∂r + r ∂θ + ∂z )

 
′ ∂vr 2 ′ ∂  vθ  1 ∂vr
τrr = 2µ + (µB − µ)(∇ · v̄) τrθ =µ r +
∂r 3 ∂r r r ∂θ
   
′ 1 ∂vθ vr 2 ′ ∂vθ 1 ∂vz
τθθ = 2µ + + (µB − µ)(∇ · v̄) τθz =µ +
r ∂θ r 3 ∂z r ∂θ
 
′ ∂vz 2 ′ ∂vr ∂vz
τzz = 2µ + (µB − µ)(∇ · v̄) τzr =µ +
∂z 3 ∂z ∂r

∂vφ
Spherical coordinates (∇ · v̄ = 1 ∂ 2
r 2 ∂r (r vr ) + 1 ∂
r sin θ ∂θ (vθ sin θ) + 1
r sin θ ∂φ )

 
′ ∂vr 2 ′ ∂  vθ  1 ∂vr
τrr = 2µ + (µB − µ)(∇ · v̄) τrθ =µ r +
∂r 3 ∂r r r ∂θ
    v  
′ 1 ∂vθ vr 2 ′ sin θ ∂ φ 1 ∂vθ
τθθ = 2µ + + (µB − µ)(∇ · v̄) τθφ =µ +
r ∂θ r 3 r ∂θ sin θ r sin θ ∂φ
    v 
′ 1 ∂vφ vr vθ cotθ 2 ′ 1 ∂vr ∂ φ
τφφ = 2µ + + + (µB − µ)(∇ · v̄) τφr =µ +r
r sin θ ∂φ r r 3 r sin θ ∂φ ∂r r

46
The Navier-Stokes equations

Navier-Stokes equations for a perfect liquid (µ and k constants)

CARTESIAN COORDINATES (x,y,z)

Continuity equation
∂vx ∂vy ∂vz
+ + =0
∂x ∂y ∂z
Momentum equation
   2 
∂vx ∂vx ∂vx ∂vx ∂p ∂ vx ∂ 2 vx ∂ 2 vx
ρ + vx + vy + vz =− +µ + + + ρfmx
∂t ∂x ∂y ∂z ∂x ∂x2 ∂y 2 ∂z 2
   2 
∂vy ∂vy ∂vy ∂vy ∂p ∂ vy ∂ 2 vy ∂ 2 vy
ρ + vx + vy + vz =− +µ + + + ρfmy
∂t ∂x ∂y ∂z ∂y ∂x2 ∂y 2 ∂z 2
   
∂vz ∂vz ∂vz ∂vz ∂p ∂ 2 vz ∂ 2 vz ∂ 2 vz
ρ + vx + vy + vz =− +µ + + + ρfmz
∂t ∂x ∂y ∂z ∂z ∂x2 ∂y 2 ∂z 2

Energy equation
   2 
∂T ∂T ∂T ∂T ∂ T ∂2T ∂2T
ρc + vx + vy + vz = φv + k + + + Qc + Qr
∂t ∂x ∂y ∂z ∂x2 ∂y 2 ∂z 2

where
"             #
∂vx 2 ∂vy 2 ∂vz 2 ∂vx ∂vy 2 ∂vx ∂vz 2 ∂vy ∂vz 2
φv = µ 2 +2 +2 + + + + + +
∂x ∂y ∂z ∂y ∂x ∂z ∂x ∂z ∂y

47
The Navier-Stokes equations

Navier-Stokes equations for a perfect liquid (µ and k constants)

CYLINDRICAL COORDINATES (r,θ,z)

Continuity equation
1 ∂rvr 1 ∂vθ ∂vz
+ + =0
r ∂r r ∂θ ∂z
Momentum equation
 
∂vr ∂vr vθ ∂vr v2 ∂vr
ρ + vr + − θ + vz =
∂t ∂r r ∂θ r ∂z
   
∂p ∂ 1 ∂ 1 ∂ 2 vr ∂ 2 vr 2 ∂vθ
− +µ (rvr ) + 2 + − 2 + ρfmr
∂r ∂r r ∂r r ∂θ 2 ∂z 2 r ∂θ

 
∂vθ ∂vθ vθ ∂vθ vr vθ ∂vθ
ρ + vr + + + vz =
∂t ∂r r ∂θ r ∂z
   
1 ∂p ∂ 1 ∂ 1 ∂ 2 vθ ∂ 2 vθ 2 ∂vr
− +µ (rvθ ) + 2 + + 2 + ρfmθ
r ∂θ ∂r r ∂r r ∂θ 2 ∂z 2 r ∂θ

     
∂vz ∂vz vθ ∂vz ∂vz ∂p 1 ∂ ∂vz 1 ∂ 2 vz ∂ 2 vz
ρ + vr + + vz =− +µ r + 2 + + ρfmz
∂t ∂r r ∂θ ∂z ∂z r ∂r ∂r r ∂θ 2 ∂z 2

Energy equation
     
∂T ∂T vθ ∂T ∂T 1 ∂ ∂T 1 ∂2T ∂2T
ρc + vr + + vz = φv + k r + 2 2 + + Qc + Qr ,
∂t ∂r r ∂θ ∂z r ∂r ∂r r ∂θ ∂z 2

where
( "      #
∂vr 2 1 ∂vθ vr 2 ∂vz 2
φv = µ 2 + + + +
∂r r ∂θ r ∂z
      )
1 ∂vz ∂vθ 2 ∂  vθ  1 ∂vr 2 ∂vr ∂vz 2
+ + r + + +
r ∂θ ∂z ∂r r r ∂θ ∂z ∂r

48
The Navier-Stokes equations

Navier-Stokes equations for a perfect liquid (µ and k constants)

SPHERICAL COORDINATES (r,θ,φ)

Continuity equation

1 ∂ 2 1 ∂ 1 ∂vφ
2
(r vr ) + (vθ sin θ) + =0
r ∂r r sin θ ∂θ r sin θ ∂φ
Momentum equation
!
vθ2 + vφ2
  
∂vr ∂vr vθ ∂vr vφ ∂vr ∂p ∂ 1 ∂ 2
ρ + vr + + − =− +µ (r vr ) +
∂t ∂r r ∂θ r sin θ ∂φ r ∂r ∂r r 2 ∂r
  
1 ∂ ∂vr 1 ∂ 2 vr 2 ∂ 2 ∂vφ
sin θ + 2 2 − 2 (vθ sin θ) − 2 + ρfmr
r 2 sin θ ∂θ ∂θ r sin θ ∂φ2 r sin θ ∂θ r sin θ ∂φ

!   
∂vθ ∂vθ vθ ∂vθ vφ ∂vθ vr vθ vφ2 cotθ 1 ∂p 1 ∂ 2 ∂vθ
ρ + vr + + + − =− +µ 2 r +
∂t ∂r r ∂θ r sin θ ∂φ r r r ∂θ r ∂r ∂r
  
1 ∂ 1 ∂ 1 ∂ 2 vθ 2 ∂vr 2 cotθ ∂vφ
(vθ sin θ) + 2 2 + 2 − 2 + ρfmθ
r 2 ∂θ sin θ ∂θ r sin θ ∂φ2 r ∂θ r sin θ ∂φ

    
∂vφ ∂vφ vθ ∂vφ vφ ∂vφ vr vφ vθ vφ 1 ∂p 1 ∂ 2 ∂vφ
ρ + vr + + + + cotθ = − +µ 2 r +
∂t ∂r r ∂θ r sin θ ∂φ r r r sin θ ∂φ r ∂r ∂r
  
1 ∂ 1 ∂ 1 ∂ 2 vφ 2 ∂vr 2 cotθ ∂vθ
(vφ sin θ) + 2 2 + 2 + 2 + ρfmφ
r 2 ∂θ sin θ ∂θ r sin θ ∂φ2 r sin θ ∂φ r sin θ ∂φ

Energy equation
 
∂T ∂T vθ ∂T vφ ∂T
ρc + vr + + =
∂t ∂r r ∂θ r sin θ ∂φ
     
1 ∂ 2 ∂T 1 ∂ ∂T 1 ∂2T
φv + k 2 r + 2 sin θ + 2 2 + Qc + Qr ,
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂φ2
where
( "      #
∂vr 2 1 ∂vθ vr 2 1 ∂vφ vr vθ cotθ 2
φv = µ 2 + + + + + +
∂r r ∂θ r r sin θ ∂φ r r
      )
sin θ ∂  vφ  1 ∂vθ 2 ∂  vθ  1 ∂vr 2 1 ∂vr ∂  vφ  2
+ + r + + +r
r ∂θ sin θ r sin θ ∂φ ∂r r r ∂θ r sin θ ∂φ ∂r r

49
The Navier-Stokes equations

50

You might also like