You are on page 1of 13

LAGRANGIAN DENSITY EQUATIONS OF SINGLE-FLUID AND TWO-FLUID FLOWS

Alejandro Clausse


CNEA, CONICET and Universidad Nacional del Centro
7000 Tandil, Argentina
Email clausse@exa.unicen.edu.ar



ABSTRACT

Variational principles for problems in fluid dynamics are discussed in the context
of the general problem of finding an equivalent derivation of a system of equations
by means of an optimization procedure. Starting with the extended Hamilton
principle of least action for fields, the Lagrange equation for fields is derived,
equivalent to the Euler-Lagrange equations of particle mechanics. It is shown that
the equation of motion of an invicid fluid can be derived from the Lagrangian
density. In addition, the flow equations of a fluid dispersed in a continuum are
derived from the average Lagrangian density, leading to the classical two-fluid
model of two-phase flow. A possible way to solve the ill-posedness problem of the
two-fluid model, by means of the correct averaging of the kinetic energy density of
the mixture, is presented.

Key Words: Lagrangian, variational principle, two-phase flow, virtual mass.


I. INTRODUCTION

The basic equations of many conservative systems can be formulated applying the Hamilton
variational principle [1]. Such formulation has several appealing features, namely:

1) The use of the principle helps in recognizing convenient physical variables to represent complicated
systems.
2) There is an aesthetic attraction in condensing all the basic equations by extremizing a scalar quantity.
3) Variational principles consist in a set of standard rules that minimizes the possibility of errors during
derivations.
4) These rules provide a systematic bookkeeping of the pertinent variables in large systems.

The general equations of mechanics are derivable from variational principles [2]. This procedure
furnishes the most suitable basis for the systematic study of the conservation theorems. In the same way,
the fluid equations can be derived from the extremization of a scalar. This procedure is not widely used
in the field of fluid mechanics for several reasons. For start, turbulence and other irreversible phenomena
are not simple to treat. The extension of the Halmilton principle from particles described by Lagrangian
coordinates to a continuum described by Eulerian coordinates is not trivial.
Recently, the variational derivation of the fluid equations attracted the interest of researchers
working in the modelling of complex fluids for which the basic force equations is not well known [3-6].
In some of these fluids the calculation of energy components are more direct, and hence the Halmilton
principle can be useful leading to a variety of conservation laws.
One of the problems that the variational treatment proves to be useful for is two-phase flow. The
derivation of the field equations governing bubbly mixtures was traditionally approached by means of

averaging procedures. The formulation starts from the microscopic equations describing the local flow of
separate phases, i.e. the gas motion inside the bubbles and the liquid flow around the bubbles. In order to
arrive at a macroscopic description different kinds of averaging procedures have been proposed, viz. time
averaging, space averaging, ensemble averaging [7]. Unfortunately, the macroscopic equations resulting
from the averaging procedure contains quantities that are very complicated to model in terms of
macroscopic variables. On the other hand, Halmiltons variational principle was applied in the past to
characterize flows of inmiscibles mixtures of superfluids [6]. The scalar magnitude used to characterize
the system is the Lagrangian density, calculated as the difference of kinetic and the internal energy
densities. In this way, the averaging procedure is applied to energies instead of forces, at the beginning of
the analysis.
Despite of the fundamental importance of this derivation procedure, there is no readily available
account of it that is adapted to the needs of engineering and physics students. The goal of this article is to
provide a simplified account of the variational derivation of the fluid mechanics equations, which it is
hoped will be of assistance to the reader in gaining an idea of the concepts underlying this problem.

II. THE HAMILTON PRINCIPLE

The Hamilton principle for classical mechanics states that the motion equations of a particle is
equivalent to the minimization of the action:

=
2 2
1 1
,
,
) , (
t x
t x
dt x x L A (1)

where L is a function which characterize the system called the Lagrangian, and equal to the kinetic
energy minus the potential energy.
From all possible trajectories in the (x,t) plane binding the points (x
1
, t
1
) and (x
2
, t
2
) (Fig. 1), the
particle will follow the one that minimize A, call it x
o
(t). Let us calculate the action for a trajectory x(t) =
x
o
(t) + (t) close to the solution (i.e. small ). To first order this will be:


+ = + + =
2 2
1 1
2 2
1 1
,
,
,
,
) , (
t x
t x
o
t x
t x
o o
dt
x
L
x
L
A dt x x L A
D
D
D
D (2)


















Figure 1. Minimum action principle.

x
t
(x
1
, t
1
)
(x
2
, t
2
)
A
1
A
2
A
3
A
min

Table 1 Nomenclature.
Symbol Parameter
A Action
a Sphere radius
C
o
Virtual mass coefficient, Eq. (43)
c
1
Virtual mass coefficient, Eq. (43)
D Pipe diameter
e Eccentricity
H Slug cell length, H
g
+ H
l
H
g
Bubble length, Fig. 3
H
l
Slug length, Fig. 3
J Magnitude defined in Eq. (59)
K Kinetic energy
k Spring constant, kinetic energy density
K
v
Virtual mass term
L Lagrangian, Lagrangian density
M Mass
p Pressure
R Density average defined in Eq. (41)
R Radial coordinate
T Time
u Fluid velocity
u
r
Relative velocity, u
g
- u
l
v

Polar velocity
v
r
Radial velocity
x Particle position
z Spatial coordinate
Void fraction
Determinant, Eq. (35)
Perturbation
Small parameter
Drift flux, Eq. (31)
Potential field
Chemical potential
Field perturbation
Density

v
Exertia, Eq. (38)

*
Density average defined in Eq. (30)
Polar coordinate
Scalar field

g Gas
l Liquid
. Temporal derivative
Spatial derivative



The second term in the last integral can be worked out as:


2 2
1 1
2 2
1 1
2
1
2 2
1 1
,
,
,
,
,
,
t x
t x
t x
t x
t
t
t x
t x
dt
x
L
dt
d
x
L
d
x
L
dt
x
L

(3)

The first term in Eq. (3) vanishes since is zero at (x
1
, t
1
) and (x
2
, t
2
).
Combining Eqs. (2) and (3):

(

'

+ =
2 2
1 1
,
,
t x
t x
o
dt
x
L
dt
d
x
L
A A

(4)

Since A
o
is a minimum, the Lagrange equation follows:

0 =

x
L
dt
d
x
L

(5)

For instance, the Lagrangian of the harmonic oscillator is:

2 2
2
1
2
1
kx x m L = (6)

Applying Eq. (5):

0 = + kx
dt
x d
m
D
(7)

III. THE HAMILTON PRINCIPLE FOR FIELDS

Consider now a physical system represented by a scalar field (x,t), from which we can calculate
a Lagrangian density

t x
L

, , . The action is written as:

dx dt
t x
L A
t x
t x

=
2 2
1 1
,
,
) , , (

(8)

We expand now the field around the solution
o
:

) , ( ) , ( ) , ( t x t x t x
o
+ = (9)

where (x
1
, t
1
)= (x
2
, t
2
)=0.
The action will be to first order:


dt dx
L
x
L
t
L
A
dt dx
L L L
A A
t x
t x
o
t x
t x
o

'

'

+ =

'

+ =
2 2
1 1
2 2
1 1
,
,
,
,


D
D
D
(10)

where the second and third terms in the integral were solved by parts following the same procedure as in
the particle action. Since A
o
is the minimumaction, Eq.(10) leads to the Lagrange equation for fields:


L
x
L
t
L

(11)


III. THE LAGRANGIAN DENSITY OF AN INVICID FLUID

Consider now the velocity field of an invicid fluid. For simplicity, we shall assume only one-
dimensional flow (a pipeline for example). The Lagrangian density is:


2
1
2
1
2
= u L (12)

Hence, three fields are involved, namely density , velocity u and chemical potential . The
density and the velocity fields should satisfy the continuity equation:

0 =

z
u
t

(13)

To account for the condition given by Eq. (13), and u can be derived from a potential field :

D =

=
=

=
t
u
z
(14)

Combining Eqs. (12) and (14):

=
2
2
1 D
L (15)

Applying the Lagrange equations to Eq. (15):

0 =

z z
u
u
t
u
(16)

Since

dp
d = , the Euler equation follows:


z
p
z
u
u
t
u

(17)


IV. THE LAGRANGIAN DENSITY OF TWO-FLUID FLOW

Consider an adiabatic flow of two fluids, gas and liquid, in a channel, with the following features:

No mass interchange between phases
Neglectable friction and gravity forces
Constant physical properties
The gas is trapped in bubbles flowing in a continuum liquid

The scalar field that characterizes a dispersion of bubbles in a liquid environment under static
conditions is the volume fraction of the gas, called the void fraction. Averaging the energies over a
certain spatial volume, we can write the Lagrangian density for a gas-liquid mixture as:

( ) ( )
l l g g l l g g
u u L + = 1 1
2
1
2
1
2 2
(18)

with the continuity constraints:

0 =

z
u
t
g g g

(19)
( ) ( )
0
1 1
=


z
u
t
l l l

(20)

Following the same procedure as in the derivation of the Euler equation, let us write the continuity
equations in terms of potential fields:

( )
( )



D
D
=
=
=
=
l l
l
g g
g
u
u
1
1
(21)

The Lagrangian of the mixture is then:

+ +

=
l g
L
2 2
2
1
2
1 D D
(22)

Considering pressure equilibrium between phases:

l
l
g
g
d
d
dp

= = , (23)

the equations of the two-fluid model of two-phase flow follows:


z
p
z
u
t
u
g
g g


2
(24)
( ) ( ) ( )
z
p
z
u
t
u
l
l l

1 1 1
2
(25)

Eliminating
z
p

from Eqs. (24) and (25):



( ) ( )
0
1
1
1
1
2
2
=

z
u
z
u
t
u
t
u
l l
g g
l l
g g

(26)

Eq.(26) can be better viewed by changing the frame of reference to a coordinatemoving with the
center of volume of the mixture, that is:

( ) [ ] ( )
( ) [ ]
r l g l l
r l g g g
u u u u u
u u u u u


= +
= +
1
1 1
(27)

where
l g r
u u u = .
The mass and momentumbalance are then written as:

0 =

z t

(28)
0
2
* 2 *
=


d
d
z t
(29)

where:

+ =
1
* l
g
(30)

( )
r
u = 1 , usually called the drift-flux. (31)

The ill-posedness problem. Eqs. (28) and (29) look nice, but they are deadly ill posed. To understand the
meaning of this, let us linearize Eqs. (28) and (29) around steady-state values
o
and
o
:

0
2
0
*
*
*
*
=

+ +

z z d
d
t t d
d
z t
o o
o
o
o


(32)

Combining Eqs. (32):

( ) ( ) ( )
0
1 1
2
1
2
2
3 3
2
2
2 2 2
2
=

+ +

+
z z t t
o
l
o
g
o
o
g
o
l
o
o
l
o
g

(33)


Eq. (33) is a second order partial differential equation whose type depends on the sign of the
determinant:

( )
( )
( )
o
l
o
g
o
g
o
l
o
l
o
g
o

+ =

1
1
1
2
2 2
3 3 2
(34)

Negative leads to hyperbolic equations like the wave equation, positive leads to elliptic
equations like the Laplace equation. In our case:

( )
>

=


0
1
3 3 *
2
o o o
o l g
Always elliptic ! (35)

This means that to determine the solution it is necessary to give conditions not only in the spatial
boundary, but also in the time boundary. The latter clearly violates causality, for the occurrence of future
events would influence the flow in the past. This paradox is known as the two-fluid model ill-posedness,
and it cannot be avoided by including gravity or friction terms.

The virtual mass. The ill-posedness of the two-fluid equations can be avoided by adding a term to the
Lagrangian accounting for the kinetic coupling between phases. The rationale behind that comes from
realizing that the kinetic energy density of the dispersed fluid (the liquid in our case) is not proportional
to the square of the average velocity as we wrote in Eqs. (24) and (25) but proportional to the
average of the square of the velocity. For instance, consider the case of a spherical bubble moving at
velocity u
g
in an invicid stagnant liquid (Fig. 2). The liquid velocity field can be obtained from potential
theory, resulting [8]:
















Figure 2. Sphere moving in a stagnant liquid.

sin
2
cos
3
3
3
3
r
a
u v
r
a
u v
g
g r
=
=
(36)


Integrating over the whole liquid space, the kinetic energy results:

( )
( )
2 3
0
2
4
2 6
0
2 2 2 2
3
1
sin cos 3 1
4
sin
g l
a
g l
a
r g l l
u a
d
r
dr
u a
d v v dr r u K

=
+ =
+ =

(37)

Had we calculated the kinetic energy as the square of the average velocity the result would have
been zero, since the mean liquid velocity is null. The energy expression given by Eq. (37) is the
consequence of the inertial coupling between the sphere and the liquid, and it is often known as added or
virtual mass term. For other bubbles distributions the virtual mass term can be written in general as:

2
) (
2
1

v v
k = (38)

where we have introduced a phenomenological function
v
() , called the exertia [9-10], which depends
on the configuration of the bubbles.
Including the virtual mass term into the Lagrangian of the mixture, we have:

( )
l l g g v
l
g
L

+ = 1 ) (
1 2
1
2
(39)

Applying the Lagrange equations to Eq. (39) leads to:

0
2
2
=


d
dR
z t
R
(40)

where

) (
1
) (

v
l
g
R +

+ = (41)

Thus, the form of the exertia determines the character of the equations, since Eq. (40) is ill posed
if:

0 2
2
2
2
<

d
dR
d
R d
R , (42)

which imposes a condition on the exertia.
A simple constitutive model of exertia which cover a wide range of practical cases is:

( )

+
=
1
1
c c
o
l
v
(43)

In such case well-posedness is assured if:


( )( ) 0 1
1
< + + + c c c
o l o g
(44)

The value of the coefficients c
o
and c
1
depends on the topology of the flow. Therefore,
speculations were sustained that the well-posedness condition of the two-fluid equations can be a method
to determine flow pattern transitions.

A simple case of slug flow. Slug flow is a particular flow pattern in vertical gas-liquid mixtures, where
the bubbles coalesce occupying almost all the pipe cross section. An estimate of the function f
v
() can be
produced considering unit cells of diameter D and length H, containing a cylindrical bubble of length H
g

and diameter D
g
, followed by a liquid slug of length H
s
(Fig. 3). The void fraction is given by:

2
2
HD
D H
g g
= (45)

Let us call u
s
the velocity of the liquid slug and u
f
the velocity of the liquid film surrounding the
bubble. The average liquid velocity is:

( )
( )
2 2 2
2 2 2
D H D D H
u D D H u D H
u
s g g
f g g s s
l
+
+
= (46)

The total kinetic energy density is:

















Figure 3. Slug-flow model.


( )

+ =
+ +
=
2
2
2 2 2
2
2 2 2 2 2 2 2
1
2
1
2
1
2
1
2
1
2
1
2
1
D
D
H
H
u
H
H
u u
HD
D D H u D H u D H u
k
g g
f l
s
s l g g
g g f l s s l g g g g


(47)

u
g
u
s
u
f
u
f
D
g
H
g
H
s

Had we naively calculated k using the square of the average liquid velocity, the result would
have been:

( )
2 2
2
1
1
2
1
g g l l
u u k + = (48)

The virtual mass term is given by the difference of Eqs (47) and (48):

( )
2
2
2
2 2
1
2
1
1
2
1
2
1
l l
g g
f l
s
s l v
u
D
D
H
H
u
H
H
u k

= (49)

Since the volumetric flux should be the same along the tube, the velocities are related by:

( )
f g g g s
u D D u D u D
2 2 2 2
+ = (50)

Combining Eqs. (49) and (50):

( ) ( )
( )

=
2
2
2
2
2
2
2
2
2
2
1
1 2
1
1
1
2
1
D
D
D
D
D
D
D
D
u u k
g
g
l
g
g
l g l v

(51)

Comparing Eqs. (43) and (51), we have for the slug-flow model:

2 2
2
g
g
o
D D
D
c

= (52)
2 2
2
1
g
D D
D
c

= (53)

Models based in potential flow. Potential flow theory can be applied quite easily to produce
expressions for ) (
v
. The kinetic energy of an irrotational fluid flowing around objects can be written
as [8]:

dV K
V
l l

=
2
1
(54)

Since the fluid is incompressible, 0
2
= , and therefore:

( ) = (55)

Hence, applying the divergence theorem:

dS K
S
l l
=


2
1
(56)

where the integral is taken over the envelope of the liquid domain.












Figure 4. Unit cell containing a bubble array.


When the flow is confined to pipes, it is possible to estimate K
l
solving the Laplace equation for
the potential inside unit cells with periodic boundary conditions (Fig. 4). Wallis [10] shows that for an
ensemble of random well-mixed collection of spherical bubbles:

( )

=
1 2
1 ) (
l
v
(57)

Smereka and Milton [11] consider mixtures of ellipsoidal bubbles finding:

( )( )
( )

+
+
+
=
1
) (
3
1 1 2
) (
2
O
J J
J
l
v
(58)

where J is related to the ratio of the major axis to the minor axis, e, according to:

( ) ( )
( ) ( )
2
2
1
2 1 1
1 1
2
1
2
1 cos
cos 1




=
e e e
e e
J (59)


IV. CONCLUSIONS

The purpose of this paper was to discuss the application of variational principles in fluid dynamics. It
was shown that the equation of fluids motion can be derived from the Lagrangian density, opening an
elegant perspective to study complex flows whose basic forces are difficult to formulate in terms of state
variables. The procedure was applied to the flow equations of two-phase mixtures, providing a new way
to tackle the virtual-mass force terms.
Many interesting fluid mechanics problems can be analysed using the variational technique presented
here. I would like to suggest two research lines that deserve further study. The first is the identification of
symmetry properties of fluid equations, which can be searched applying the Noether theorem. In most
cases in particle mechanics, the standard constants of motion combined with other conserved quantities
associated with space-time symmetries of the action lead to algebraic solutions of the particles motion. It
would be interesting to extend these findings to the fluid realm. Another issue that is important from the
practical point of view, is the treatment of dissipative flows. Starting with the derivation of the Navier-
Stokes equations, the variational modeling of non-conservative terms may open new channels to study
more complicated phenomena, as turbulence and radiating plasmas.


A

B

REFERENCES

[1] Goldstein H., Classical Mechanics, Addison-Wesley Pub. Co., Massachusetts, 1950.
[2] Hill E., Halmilton's principle and the conservation theorems of mathematical physics, Rev. Modern
Phys., 23, 253-260, 1951.
[3] Geurst J ., Two-fluid hydrodynamics of bubbly liquid-vapor mixture including phase change, Phillips
J . Res., 40, 352-374, 1985.
[4] Geurst J ., Variational principles and two-fluid hydrodynamics of bubbly liquid-gas mixtures, Physica
135A, 455-486, 1986.
[5] Pauchon C. and Smereka P., Momentum interactions in dispersed flow: an averaging and a
variational approach, Int. J . Multiphase Flow, 18, 65-87, 1992.
[6] Penfield P., Hamilton's principle for fluids, Phys. Fluids, 9, 1184-1204, 1966.
[7] Ishii M., Thermofluid dynamic theory of two-phase flow, Eyroles, Paris, 1975.
[8] Lamb H., Hydrodynamics, Cambridge University Press, 1957.
[9] Wallis G., Inertial coupling in two-phase flow. Multiphase Sci. Technol. 5, 239, 1989.
[10] Wallis G., The averaged Bernoulli equation and macroscopic equations of motion for the
potential flow of a two-phase dispersion, Int. J . Multiphase Flow, 17, 683-695, 1991.
[11] Smereka P. and Milton G., Bubbly flow and its relation to conduction in composites, J . Fluid
Mech., 233, 65-81, 1991.

You might also like