You are on page 1of 12

Derivation of Orr-Sommerfeld equation

Let us consider an incompressible velocity field governed by the Navier-Stokes equation:

u~ v~ w~


Mass:    0 (1)
x y z
u~ ~ u~ ~ u~ ~ u~ 1 ~
p   2 u~  2 u~  2 u~ 
X-Momentum: u v w     2  2  2  (2)
t x x x  x  x y z 

v~ ~ v~ ~ v~ ~ v~ 1 ~


p   2 v~  2 v~  2 v~ 
Y-Momentum: u v w     2  2  2  (3)
t x x x  y  x y z 

~ ~ ~ ~ ~ 2~ 2~ 2~
Z-Momentum:
w
 u~
w
 v~
w ~ w   1 p     w   w   w  (4)
w
t x x x  z  x 2 y 2 z 2 
 

Let us imagine a base or mean velocity field (U(x,y,z),V(x,y,z),W(x,y,z),P(x,y,z)) and


this is perturbed by a disturbance field given by
uˆx, y, z, t , vˆx, y, z, t , wˆ x, y, z, t , pˆ x, y, z, t  . The instantaneous velocity field is then
given by u~  U  uˆ, v~  V  vˆ, w ~  W  wˆ , ~ p  P  pˆ . After substituting the expressions
for the instantaneous field in terms of the mean and the disturbance field in equations (1-
4) and subtracting out the governing equation form for mean flow (which have the same
form as equations 1-4 but with the tilde quantities (such as u~ ) replaced by the large case
quantities (such as U etc., )), one gets the equation governing the disturbance dynamics
as
uˆ vˆ wˆ
  0 (5)
x y z
uˆ U uˆ U uˆ U uˆ uˆ uˆ uˆ
 uˆ U  vˆ V w ˆ W  uˆ  vˆ w ˆ
t x x y y z z x y z
  
non linear terms to be neglected (6)
1 pˆ   uˆ  uˆ  uˆ 
2 2 2
    2  2  2 
 x  x y z 

vˆ V vˆ V vˆ V vˆ vˆ vˆ vˆ


 uˆ U  vˆ V w
ˆ W  uˆ  vˆ w
ˆ
t x x y y z z x y z
  
non linear terms to be neglected (7)
1 pˆ   vˆ  vˆ  vˆ 
2 2 2
   2  2  2 
 y  x y z 
w
ˆ W w
ˆ W w
ˆ W w
ˆ w
ˆ w
ˆ w
ˆ
 uˆ U  vˆ V w
ˆ W  uˆ  vˆ w
ˆ
t x x y y z z x y z
 
non linear terms to be neglected (8)
1 pˆ  w ˆ  w
2
ˆ
ˆ  w 2 2
   2  2  2 
 z  x y z 

We further assume the mean or base flow to be a parallel flow of the form (U(y),0,W(y)).
This along with the linearization assumption, whereby products of the fluctuations terms
are neglected, results in

uˆ vˆ wˆ


  0 (9)
x y z
uˆ uˆ U uˆ 1 pˆ   2 uˆ  2 uˆ  2 uˆ 
U vˆ W       (10)
t x y z  x  x 2 y 2 z 2 
vˆ vˆ vˆ 1 pˆ   2 vˆ  2 vˆ  2 vˆ 
U W     2  2  2  (11)
t x z  y  x y z 
w
ˆ w
ˆ W w
ˆ 1 pˆ  2wˆ 2w ˆ 2w ˆ
U vˆ W     2  2  2  (12)
t x y z  z  x y z 

We assume a travelling wave form for the disturbance field as given by

uˆ x, y, z, t   u  y  expi x cos  z sin   ct 


vˆx, y, z , t   v y  expi x cos  z sin   ct 
wˆ x, y, z , t   w y  expi x cos  z sin   ct 
pˆ x, y, z , t   p y  expi x cos  z sin   ct 

where i   1 ,  is the wave number and c is the phase speed of the wave( both are
complex quantities in general). The temporal frequency is given by   c .Upon
substitution of the travelling waveform into equations (9-12) we get (with primes
denoting differentiation with respect to y)

i u cos  w sin    v'  0 (13)


ip cos
i U cos  W sin   c u  vU '  


 u ' ' 2u  (14)

i U cos  W sin   c v


p'

 v' ' 2v  (15)
.

ip sin 
i U cos  W sin   c w  wU '    w' ' 2 w (16)

We introduce the notations u0  u cos  w sin  ,U 0  U cos  W sin  . By multiplying
equation (14) with cos and equation (16) by sin  and adding them and using the
notation just introduced, the governing equations become

i u 0  v '  0 (17)
i p
i U 0  c u 0  vU 0'  


  u 0''   2 u 0  (18)

i U 0  c v 
p'


  v' ' 2 v  (19)

The essence of this set of equations means that the stability of an incompressible 3D
parallel flow field with respect to an arbitrary 3D disturbance mode in a travelling wave
form. By orienting oneself in the direction of the disturbance wave front and studying the
stability of the base flow component in the wave direction, the stability of a 3D base flow
with respect to a 3D disturbance mode can be equivalently seen as the stability of a 2D
base flow with respect to a 2D wave.

Now we consider our base flow to be two-dimensional (U(y),0,0). Let us first treat the
inviscid stability of this flow. This is done by ignoring the viscous terms in equations (18-
19).

p c
Let us now redefine u  u 0 , p  and c  and. Then equation (17) becomes
cos cos
iu  v' 0 ….. (20)

ip
Equation (18) becomes iU  c u  vU '   (21)

Equation (19) becomes iU  c v  


p'
(22)

Mathematically, the stability of the 2D base flow with respect to a 3D disturbance has
been cast as a problem of a 2D base flow with respect to an equivalent 2D disturbance.
In other words, the set of equations (20-22) are identical to the stability problem of a 2D
base flow with respect to an equivalent 2D disturbance wherein the parameters of the
disturbance waves are indicated by over bars; the equivalent 2D disturbance becomes an
actual 2D disturbance when   0 .
If we consider the stability of equations (20-22), it can be seen that for the equivalent 2D
c
disturbance, the growth rate is given by ci  i . This means the growth rate of an
cos
equivalent 2D wave is larger than the corresponding 3D wave. This means that for every
amplifying 3D (or oblique) wave there is an equivalent 2D wave with a larger growth
rate. The essence of Squire’s theorem (for inviscid stability equation) is that it is enough
to consider the stability of a 2D base flow under the influence of a 2D disturbance as this
gives the ‘worst case scenario’, as it were, to calculate stability.

For the corresponding viscous stability problem now include hitherto neglected viscous
terms. Then under the rescaling to construct the equivalent 2D base flow + 2D

disturbance we additionally get   . For the viscous stability problem, the system
cos
of equations become

iu  v' 0 (20-a)


ip
iU  c u  vU '  


  u ' ' 2 u  (21-a)

iU  c v  
p'


  v' ' 2 v  (22-a)

This means in the viscous counterpart of the 2D equivalent problem, the kinematic
viscosity is increased for the equivalent 2D flow .In non-dimensional terms, we can say
that the Reynolds number for the 2D problem is lower. This means the critical Reynolds
number for the 2D problem is also lower than the corresponding 3D problem and hence
again it is enough to compute the viscous linear stability of a 2D base flow with 2D
disturbances.

So with the understanding that we are considering only 2D disturbances, we can drop all
the over bars in the equation to write the final equations as :

i u 0  v '  0 (23)
i p
i U  c u 0  vU 0'  


  u 0''   2 u 0  (24)

i U  c v 
p'


  v' ' 2 v  (25)

We would like to get rid of the pressure term. For this, differentiate equation (24),
multiply equation (25) by i , and subtract one from the other and make use of equation
(23) to write u in terms of v. The result is the celebrated Orr-Sommerfeld equation in
dimensional form:

(U  c)(v' ' 2 v)  U v 
i
v' ' ' '2 2 v' ' 4 v (26).
where primes denote differentiation with respect to y.

The non-dimensional form of (26) is


(U  c)(v' ' 2 v)  U v 
1
i R
v' ' ' '2 2 v' ' 4 v, (27)
where   y
denotes differentiation with respect to  
, where L is the characteristic
L
dimension used for non-dimensionalisation. In (27) all the velocities have been
UL
normalised with a reference velocity U and the Reynolds number R  . The (non-

dimensional) wave number  in (27) is the dimensional wave number as in (26)
multiplied by the reference length. For a channel, the reference velocity is the centreline
velocity and the reference length is the half-width of the channel; for a boundary layer,
the reference velocity is the free stream velocity and the reference length is the boundary
layer thickness (or the displacement or momentum thickness). And so on.
d
With the notation D  for the differentiation operator, it is sometimes convenient
d
to rewrite (in an easy-to-remember form)
(U  c)( D 2   2 )v  U v 
1
iR
 D2   2 v .
2
(27*)
Equation (27) is a fourth order linear homogeneous (i.e., no forcing term)
ordinary differential equation governing the disturbance dynamics. Even though this is a
linear O.D.E, which means it has nice properties such as linear superposition of solutions
being a solution of the equation, it is still a formidable equation to solve due to singular
behaviour of the solution at certain points in the flow domain. We will see more of it
later.
The boundary conditions are given by the consideration that the vertical velocity
(v()) of the disturbance has to vanish at the boundaries of the flow domain; for the
boundary layer this means the vertical velocity of the disturbance should vanish at the
wall and at the free stream. However, since the O-S equation is a fourth order equation,
we need two more boundary conditions. For these, we could demand that the stream wise
velocity of the disturbance (u()) should also vanish at these boundary locations. If we
remember the form of the continuity equation in the course of derivation of the O-S
equation we have a relation between the u and v velocity components of the disturbance:
v'
iu  v'  0 (for a two-dimensional disturbance). This yields u   , i.e., stream wise
i
velocity component is proportional to the y-derivative of the vertical velocity component.
This means the vertical velocity and its y-derivative have to disappear at the boundaries
of the flow geometry under consideration. To sum up, the boundary conditions are:

Boundary layer : v  0  v'     v  0  v'     0 .


Channel flow: v  1  v'   1  0 .
Free shear flows: v    v'     0 . (28)

For the 2D disturbance (once Squire’s theorem guarantees the validity of using that in our
analysis), we could have written the form of the stream function from which we could
then obtain the velocity components u and v. The stream function is given by
ˆ x, y, z, t     y  expi x  ct , (29)
using the normal mode form for the disturbance wave. Then we could write
ˆ
uˆ x, y, z, t     '  y  expi x  ct   u ( y ) expi x  ct , (30a)
y
and
ˆ
vˆx, y, z, t      i   y  expi x  ct   v( y ) expi x  ct . (30b)
x
These expressions (30a&30b) lead to
u y     y  , (31a)
v y    i   y  , (31b)
respectively leading to the expression iu  v'  0 . In other words, the introduction of a
stream function formulation automatically satisfies the continuity equation for a 2D
disturbance, as it should.
If we substitute expression (31b) in the original dimensional form of the O-S equation
(26), we end up in

(U  c)(    2 )  U  
i
   2 2    4. (32)
Equations (26) and (32) are equivalent descriptions and we will be using them
interchangeably in this course. The non-dimensional form of (32) is
(U  c)(    2 )  U  
1
iR

   2 2    4 , (33)
is likewise equivalent to (27). The boundary conditions for equation (32) are:
Boundary layer:    0   '        0   '     0 .
Channel flow:    1   '   1  0 .
Free shear flows:    1   '   1  0 . (34)

Since the O-S equation (33) is a fourth order equation, we can expect 4 independent
solutions:  1 ,  2 ,  3 and  4 . Hence a general solution to (33) can be written as
  A1  B2  C3  D4 , (35)
where A,B,C and D are arbitrary constants.

Using the boundary conditions for a channel flow, for example, from (34) in the
expression (35), we could write
 1  1  2  1  3  1  4  1 A   0 
 '    
 1  1  2  1  3  1  4  1 B   0 
' ' '

  1  1  . (36)


 3 1  4 1  C   0 
 ' 1 2
   
  1
 1  2' 1  3' 1  4' 1  D   0 

Since A, B, C and D are arbitrary constants


 1  1  2  1  3  1  4  1
 ' 
 1  1  2  1  3  1  4  1
' ' '

Det  0, (37)
1 1  2 1  3 1  4 1 
 ' 
  1  2' 1  3' 1  4' 1 
 1

where Det stands for determinant. Clearly, the determinant elements in (37) are functions
of the disturbance parameters  , c and the Reynolds number R. Therefore (37) leads to
a functional form
F ( , c, R)  0 . (38)
Relation (38) is known as the dispersion relation for the disturbance. Clearly this is an
eigenvalue problem in terms of  , c and R. In general,    r  i i and c  cr  ici are
complex quantities with the corresponding real parts being the physical wave number and
phase speed respectively; the significance of imaginary parts is that exp  i x  and
exp r ci t  are the spatial and temporal amplification rates respectively. Two special
cases arise, when either one of these quantities are taken to be real, for ease of analysis.
They are:
Temporal stability analysis:  real and c complex so that
F  , cr , ci , R  0 (39)
Spatial stability analysis:  complex and c real so that
G r ,  i , cr , R  0 . (40)
It is easier to do a temporal stability analysis and so we will follow that. However, the
results of spatial stability analysis are closer to the spirit of experimental studies of
stability.

Physically, what the expression (39) means is the following. We have assumed a
travelling wave form for the disturbance and plugged them into the governing equations.
In other words, the form of the solutions is assumed but the disturbance parameters  and
c are unknown to start with. When the dispersion relation relation (39) is solved, we are
in effect seeking to obtain the characteristic or the permissible combinations of  and
cr for a given R for decay or growth (i.e., for ci  0 or ci  0 respectively) of the
disturbance. The case ci  0 is clearly the borderline between stability and instability and
hence it is an important case to be considered for determining stability boundaries. ci  0
corresponds to the neutral disturbance wherein the kinetic energy of the disturbance stays
put with time (so that production = dissipation according to the energy equation). The
resulting stability curves (which demarcate the stable and unstable regions and are
typically thumb shaped) are plotted in the form of the so-called neutral stability diagrams
as shown below. The lowest value of Reynolds number for any wave for which instability
is possible is called the critical Reynolds number and is an important parameter for
applications such as transition prediction and management.

Unstable
ci  0

Stable

Rcritical
R

Physical interpretation of the Orr-Sommerfeld equation

Since the Orr-Sommerfeld equation has been derived from the momentum equation and
the LHS contains convective terms and the RHS terms contain viscous diffusion, it is
perhaps possible to make an analogy with the boundary layer equation and interpret O-S
as a convection-diffusion balance. However, the analogy is not obvious. This is because a
question arises for O-S: convection-diffusion balance of which quantity? Momentum? It
cannot be streamwise momentum as this is a governing equation for vertical velocity
component and it also involves fourth derivative of that component (as against N-S where
the corresponding term has a second derivative of the streamwise velocity).
Let us start from the linearised N-S equation for the 2-D disturbance:
uˆ uˆ U 1 p
U  vˆ    2uˆ ………………..(*)
t x y  x
vˆ vˆ 1 p
U    2vˆ ………………………..(**)
t x  y
Differentiating (**) with respect to x and (*) with respect to y and subtracting one from
vˆ uˆ
the other we get the linearised vorticity equation for ˆ   :
x y
ˆ ˆ
U  vˆU    2ˆ (***) (check this for yourself!)
t x
Introducing the normal mode form for the fluctuation vorticity
ˆx, y, t    ( y) expi x  ct  into (***) to obtain the Orr-Sommerfeld equation.
It can now be seen from equation (***) that the O-S equation is a indeed convection-
diffusion balance of the disturbance vorticity with an extra term vˆU  to characterise the
interaction between the mean flow and the fluctuation; this term is an extra element in
the fluctuation dynamics having no such parallel term in the equation for the base flow.
This is so because the base or the mean flow exists by itself whereas the fluctuation has
to depend on the mean flow for its sustenance.

The right hand side of Orr-Sommerfeld equation has a leading order-term of the
1
form D 4 v . For large enough Reynolds number, 1/R goes to zero and it is tempting to
R
neglect the viscous term and solve the resulting Rayleigh equation for inviscid stability
analysis. This is pretty much similar to neglecting the viscous term in the N-S equation to
obtain the Euler equation. In the case of Euler equation it turned out that this is a very
naïve approach as this led to d’Alembert’s paradox. The crux of the issue there, while
considering the N-S equation, is that a small term 1/R multiplied a very large gradient

term  2u (near the wall in boundary layer flows) such that their product cannot be
1
neglected – a singular perturbation problem! Likewise, the term D 4 v in the right hand
R
side of the O-S equation becomes substantial near walls and we have on our hands a
singular perturbation problem. In other words, the solutions of Rayleigh equation are not
necessarily the infinite Reynolds number limit solutions of the O-S equation.

Let us assume we can in fact completely neglect the viscous term and it is justifiable to
do so for flows not near solid walls or not near the so-called critical layer(where U=c)
and look at the consequences. The Rayleigh equation governing the inviscid instability of
parallel flows is then obtained:
(U  c)(v' ' 2v)  U v  0 . (40)
Since the order of the equation has dropped from 4 to 2 as we go from O-S to Rayleigh
equation we cannot keep all the four boundary conditions. So we drop the boundary
condition that the streamwise component of the velocity fluctuation u(y) vanishes at the
boundaries. The boundary conditions are:
v y  y1   v y  y2   0 , (41)
where y1 and y 2 are the boundaries of the flow domain for the parallel flow under
consideration. Note that, since R does not occur as a parameter in the Rayleigh equation,
both the dimensional and non-dimensional forms of the equation are equally convenient
to work with. We will use the dimensional form.

Solution of Rayleigh equation


Consider Rayleigh equation (40) along with the boundary condition (41). Rewrite the
equation in the form
U v
v' ' 2v  . (42)
U c
(It should be noted that if c i = 0 so that c  cr is real, then the right hand side of (42)
could become singular when the denominator goes to zero. We could expect analytical
difficulties in such situations. However, no such difficulty arises if ci  0 so that
c  cr  ici is complex.)

Multiply equation (42) by the complex conjugate of v(y), denoted by v*(y), to get
2
U vv * U ' ' v
v' ' v *  vv* 
2
 U  c  * . (43)
U c U c2
Equation (43) can be integrated between y1 and y 2 , to get:
y2 y2 2
U '' v
 (v' ' v *  vv*)dy 
2
U  cr   ici dy .
 (44)
U c
2
y1 y1

The first term on the LHS in equation (20), can be integrated by parts to yield
y2

 dv   dv 
y2 y2 y y 2
2 2
dv dv * dv
 v' ' v * dy   v * d    v*
 dy    
dy dy
dy   
dy
dy.
y1 y1 dy y1
 y1
 y1

When boundary conditions (41) are applied, the terms in the square bracket in the above
expression become zero. Substituting the resulting expression in (44), one can then write:
y2 2 y2 2
U '' v
  (
dv
  2 v )dy 
2
 U  cr   ici dy . (45)
U c
2
y1 dy y1

By equating the real and imaginary parts in (45) on the LHS and the RHS, we can write

U ' ' v U  cr   dv 2 
y2 2 y2

 dy       2 v dy ,
2
(46)
U c
2
y1  
dy
y1 
y2 2
U '' v
ci  dy  0. (47)
U c
2
y1

Equations (46) and (47) apply equally to damped, neutral or growing disturbances in that
we have not said anything so far about the value of c i . Now in the following we will
consider the growing disturbance.

Growing disturbance or unstable waves case( c i >0)

Let us examine now the conditions needed for the instability to occur. Consider c i >0.

Then for a growing disturbance


y2 2
U '' v

y1 U  c
2
dy  0. (48)

For the integral to be zero, U ' ' has to change sign so that U ' '  0 (an inflexion point)
atleast once in the interval  y1 , y 2  . This is the famous inflexion point theorem of
Lord Rayleigh, which says the necessary condition for instability is that the basic velocity
profile should have an inflection point. Note that the presence of an inflection point in
mean velocity implies the mean vorticity is either a maximum or minimum. So
Rayleigh’s inflexion point theorm does not differentiate between a vorticity maximum or
vorticity minimum. If we try to apply Rayleigh’s theorem to plane Couette flow where
U ' '  0 everywhere in the flow field, it is not clear if there will be instability or not. In
order to resolve this, we invoke the so-called Fjortoft theorem which we prove next.
Fjortoft’s theorem

Consider equation (46) which is true for all values of ci  0 or ci  0 .


U ' ' v U  cr   dv 2 
y2 2 y2

 dy       2 v dy .
2
(46)
U c
2
y1  
dy
y1 

Consider Rayleigh’s result for a growing disturbance ( c i >0) – i.e., equation (48) and
define the quantity U ( y s )  U s where y  y s is the location of the inflection point in the
mean velocity profile i.e., U ' '  y  y s   0 . Now multiply ( (U  U s ) by equation (48)
(which is true for a growing disturbance) so that
y2 2
U '' v
(cr  U s )  dy  0. (49)
y1 U  c
2

Adding equations (46) and (49) , we get the following relation for instability:
U ' ' U  U s  v
y2 2


y1 U c
2
dy  0 , ()

from which the following result follows:

Fjortoft’s theorem: A necessary condition for instability is that U ' ' (U  U s )  0


somewhere in the field of flow.

U ' '  0 everywhere U ' '  0 everywhere.


Stable. Stable

U ' ' ( ys )  0 U ' ' ( ys )  0


U ' ' (U  U s )  0 U ' ' (U  U s )  0
Stable. Possibly unstable.

By examining the four figures above, it is clear that the Fjortoft’s theorem picks out the
profiles with maximum vorticity of the mean flow as the (necessary) condition for
instability. Now if we apply this to Couette’s flow it does not have a vorticity maximum
of the mean flow; in fact the mean flow vorticity is everywhere constant for a Couette
flow. Hence Couette flow is stable according to inviscid analysis.
As has been already mentioned, the presence of an inflection point or a vorticity
maximum mean flow is only a necessary condition. In fact, it was shown by W.Tollmien
by consideration of the velocity profile U  sin y that the presence of an inflection point
or the vorticity maximum in the mean profile are not sufficient conditions for instability.
Tollmien also however, gave a heuristic argument to suggest that the conditions are
sufficient for symmetric profiles in a channel and for monotone profiles of the boundary
layer type. These will not be discussed here and reference is made to Drazin & Reid
“Hydrodynamic Stability”.

You might also like