You are on page 1of 42

Computation of free-surface

ows using interface-


tracking and interface-capturing methods
Samir Muzaferija and Milovan Peric
Fluid Dynamics and Ship Theory Section,
Technical University of Hamburg-Harburg,
Lammersieth 90, D-22 305 Hamburg, Germany
Email: peric@schi bau.uni-hamburg.de

Abstract
The use of popular nite-volume methods for incompressible viscous ows
for solving problems involving free surfaces is described. The methods are
classi ed into two groups: interface-tracking and interface-capturing meth-
ods. The former compute the ow of the liquid only, using a grid which
is adapted to the free surface and which moves and deforms with it. The
latter use a solution domain which extends over both gas and liquid and
solve an additional equation to determine the distribution of the two uids.
Advantages and disadvantages of each class of methods are outlined and
their implementation in nite-volume schemes with arbitrary unstructured
grids is described. Examples of application of methods of both types are
presented to demonstrate their capabilities.

1 Introduction
Flows of two (or more) immiscible uids are often encountered in
both engineering practice and in environment. They can be classi ed
into three groups, based on the structure of the interface between

1
the uids. In the rst class, the two uids are separated by a well-
de ned, sharp interface. Examples are ows in open channels with
non-breaking waves, ows in stirring vessels, and ows in tanks which
shake gently. The second class includes transitional ows in which
parts of the interface break into regions lled by both uids (e.g.
air bubbles trapped in liquid and liquid drops dispersed into air).
Examples are open channel ows, ows in tanks, and ows around
ships with breaking waves. The third class includes dispersed ows
in which the two uids make a suspension without a clearly de ned
interface, e.g. in a vigorously shaken tank.
We shall deal here with free-surface ow problems of the rst and
partly of the second class. We shall further limit the analysis to in-
compressible ows without a signi cant phase change at the free sur-
face. Methods used for predicting such ows with free surfaces can be
grouped into two broad categories: interface-capturing and interface-
tracking methods. In interface-capturing methods, the computation
is performed in a solution domain which extends over regions occupied
by both uids. An additional equation is solved for the concentration
or volume fraction of one uid, say c, which is used to determine the
uid properties at a particular location. This method has a disad-
vantage that errors may accumulate while evolving the eld c. The
most common source of error is the numerical di usion. Also, c has
to stay within bounds 0  c  1, while most second- and higher-
order schemes tend to produce over- and undershoots, thus making
the obtaining of accurate results dicult. The next problem is the
construction of a free surface once the eld of c is calculated since
the interface is not sharply de ned. Interface-tracking methods can
follow the evolution of a simple interface very accurately. However,
they encounter diculties in dealing with changes in interface topol-
ogy. When the ow conditions are such that there are no breaking or
overturning waves, the interface-tracking method is a good approach.
In the following sections we shall discuss the methods of both
kinds and present a representative example of each kind. The anal-
ysis will be limited to nite-volume methods that solve the Euler or
Navier-Stokes equations; methods designed for potential ows (es-
pecially the boundary-element methods) will not be discussed, since
they are covered in other chapters. Results of several typical ows

2
with free surfaces will be presented to demonstrate features of the
solution methods and discuss their advantages and disadvantages.
Finally, an outlook to future developments in this area will be given.

2 Finite-Volume Method
In this section we brie y describe the nite-volume method (FVM)
for solving the Navier-Stokes or Euler equations; more details can
be found in Ferziger and Peric1 or Demirdzic et al.2 . The starting
point are the conservation equations for mass, momentum, and scalar
quantities (e.g. energy or chemical species) in their integral form:
d Z  dV + Z (v v )  n dS = 0 ; (1)
dt b
V S
d u dV + u (v v )  n dS = Z ( i pi )  n dS + Z b dV ;
Z Z
dt i i b ij j i i
V S S V
Z Z Z Z (2)
d  dV + (v v )  n dS = r  n dS + b dV : (3)
dt b 
V S S V
In these equations,  is the uid density, V is the control volume
(CV) bounded by a closed surface S , v is the uid velocity vector
whose Cartesian components are ui , vb is the velocity of the CV
surface, t is time, is the di usion coecient and b the volumetric
source of the conserved scalar quantity , p is the pressure, bi is
the body force in the direction of the Cartesian coordinate xi , n
is the unit vector normal to S and directed outwards, and ij are
the components of the viscous stress tensor de ned (for Newtonian
incompressible uids considered here) as:
!
@ui
ij =  @x + @x ;@u j (4)
j i
with  being the dynamic viscosity of the uid.
When the control volume moves, the so called space conserva-
tion law (SCL), which is expressed by the following relation between

3
the rate of change of CV volume and its surface velocity, has to be
satis ed too:
d Z dV Z v  n dS = 0 : (5)
dt b
V S
The solution domain has to be subdivided into a nite num-
ber of CVs, which can in principle be of any shape and are de-
ned by edges which form a grid. The grids can be structured,
block-structured (non-overlapping or overlapping), or fully unstruc-
tured. Structured grids are limited to solution domains of a rel-
atively simple shape, so that it can be mapped onto a rectangle
(two-dimensional or 2D-problems) or cube (three-dimensional or 3D-
problems). Block-structured grids are more exible and are widely
used; see Lilek et al3 for a description of a method which uses non-
matching, non-overlapping blocks which can slide along each other;
Tu and Fuchs4 describe a method using overlapping blocks (Chimera-
grids). Demirdzic et al.2 describe a method which uses unstructured
grids and CVs of an arbitrary shape. Usually, the unknown vari-
ables are computed at a node centered in the CV, although other
arrangements (vertex-centered schemes) are also possible.
In order to solve the conservation equations using FVM, they are
applied to each CV and discretized in order to obtain one algebraic
equation per CV; each such equation involves the unknown from the
CV-center and from a certain number of neighbor CVs. The equa-
tions also have to be linearized; this means that an iterative solution
method has to be used. Most often the coupled system of equations is
solved in a segregated manner, i.e. for each variable in turn, treating
thereby other variables as known.
In the course of obtaining an algebraic equation for each CV, two
levels of approximation are necessary:
 The integrals over surface, volume, and time need to be evaluated
by a suitable numerical approximation, which uses the value of the
integrand at one or more locations within the integration domain;
 Since the variable values are calculated at CV centers only, values
at other locations { which are required for the evaluation of inte-
grals { have to be obtained by interpolation; also, derivatives of
4
certain quantities are required, which makes numerical di erenti-
ation necessary.
Only the most common approximations will be brie y described here;
more details can be found in textbooks on general computational uid
dynamics (CFD), e.g. Ferziger and Peric1 .
2.1 Approximation of Surface and Volume Integrals
For both surface and volume integrals, it is the most convenient to
use midpoint-rule approximations; they result in simple algebraic ex-
pressions and are second-order accurate, which is usually the best
compromise between simplicity and accuracy. Methods of higher or-
der require evaluation of the integrand at more than one location; see
Lilek and Peric5 for a description of a method in which fourth-order
approximations for all terms were used. Examples of general CVs in
2D and 3D are shown in Fig. 1.
Nk n
n
k Nk
n
k Sk Nk Sk
C n
k
∆V C C
∆V
2D n
3D

Figure 1: An example of a 2D and a 3D control volume.

Since all unknowns and uid properties are calculated and stored
at CV centers, there are no additional approximations involved in the
evaluation of volume integrals when midpoint rule is used: one sim-
ply multiplies the CV-center value of the integrand with CV-volume
V . For the calculation of surface integrals, further approximations
are necessary since the values of the integrand, f = (v vb ) in con-
5
vective and f = r in di usive uxes, see Eq. (3), are not known
at the cell-face center. Therefore, interpolation and numerical di er-
entiation have to be used to express the cell-face values of variables
and their derivatives through the nodal values.
Cell-face values of variables are often approximated using linear
interpolation:
k  Nk k + C (1 k ) ; (6)
where k is the interpolation factor. This is a second-order approxi-
mation at the location k0 on the straight line connecting nodes C and
Nk , cf. Fig. 1. If this line does not pass through the cell-face center
k, the integral approximation
Z
Fk = f  n dS  (f  n)k Sk (7)
Sk
will not be second-order accurate. As long as k0 is close to k relative
to the sell-face size, the rst-order error term is small. The grid
quality can be optimized by minimizing jrk rk0 j, where r denotes
the position vector. The second-order accuracy can be restored by
adding a correction term as follows:
k  k0 + (r)k0  (rk rk0 ) ; (8)
where the gradient at k0 can be obtained by interpolating the cell-
center gradients using Eq. (6).
Other popular approximations for the cell-face center quantities
are: linear extrapolation using two nodes on the upwind side, and
quadratic interpolation using two nodes on the upwind and one on
the downwind side. Polynomials of higher order can also be used;
see Lilek and Peric5 for a description of a fourth-order interpolation
scheme. A rst-order upwind approximation uses the cell-center value
on the upwind side; it is not recommended since the error associated
with it manifests itself as an excessive numerical di usion.
It is advantageous to use the so called deferred correction approach
(introduced by Khosla and Rubin6) when implementing any scheme
of higher order in implicit solution algorithms. In that case the rst-
order upwind approximation (Fku ) contributes to the elements of the
coecient matrix, while the higher-order approximation (Fkh , which
6
may involve more nodes than the nearest neighbors of the cell cen-
ter) is calculated explicitly using the prevailing nodal values of the
unknowns (denoted by the superscript 'old'):
Fk = Fku + (Fkh Fku )old : (9)
The term in brackets is calculated explicitly and treated as a known
source term. It can be multiplied by a factor 0   1, thus blending
the two schemes; this may be necessary in order to suppress oscilla-
tions near discontinuities or pro le peaks (e.g. for the kinetic energy
of turbulence or its dissipation rate near solid walls) when the grid is
too coarse. In the converged solution the new and old upwind approx-
imations are equal and they cancel out, leaving only the higher-order
approximation. This approach allows us to keep the entries in the
coecient matrix limited to the nearest neighbors and enhances its
diagonal dominance, thus saving memory and increasing the stability
of the iterative solution algorithm. Since iterations are necessary any-
way due to the non-linearity and coupling of equations, this deferred
correction does not a ect negatively the convergence of the solution
method (it requires under-relaxation anyway).
In order to calculate the di usive uxes, the gradient vector or
the normal derivative at the cell face is required. Again, there are
many approximations that can be used. When the line connecting
the two cell centers on either side of the cell face is orthogonal to it,
the normal derivative is easily approximated:
 @  N k C :
@n = ( r  ) k  n k  jr r j (10)
k Nk C
If the line connecting cell centers C and Nk is not orthogonal to the
cell face, one can still de ne auxiliary nodes C0 and N0k which lie on
the normal to the cell face, see Fig. 1, and calculate the derivative
from the above approximation; the values of  at these nodes must be
expressed through the cell-center values by means of a suitable inter-
polation, e.g. Eq. (8). Another possibility { which is often employed
when structured grids are used { is to de ne a grid-oriented local
coordinate system at the cell face and express the gradient operator
through derivatives with respect to these (in general non-orthogonal)

7
coordinates, as was done e.g. by Peric7 . Another alternative is to
calculate the components of the gradient vector at cell centers rst,
and then interpolate them to the cell-face center using the same inter-
polation formulae used to compute the variable values there for the
convective ux, see Eqs. (6) and (8). The gradient at the cell center
can also be calculated in many ways; one of the simplest approaches
is based on the Gauss theorem and a midpoint rule:
 @  P i
 k k S k : (11)
@xi C V
Here, Ski = Sni is the ith component of the surface vector S n, and
the summation is over all faces of the control volume (which may be
of an arbitrary shape).
The cell-center derivatives can now be interpolated to the cell-
face centers using the same interpolation technique as for convective
terms. However, oscillatory solutions may develop in this case; see
Ferziger and Peric1 for detailed discussion. Muzaferija8 introduced an
e ective approach which both avoids decoupling problems and acts
as a deferred correction, preserving the simplicity of the coecient
matrix and its diagonal dominance:
!
 N  C old rNk rC
(r)k  nk  jr k
r j (r)k jr r j nk : (12)
Nk C Nk C
The underlined term is calculated using prevailing values of the vari-
ables and treated as another deferred correction, see above. If the
line connecting nodes C and Nk is orthogonal to the cell face, the
underlined term is zero (since the vectors rNk rC and nk are then
co-linear) and the approximation reduces to that of Eq. (10). The ex-
plicitly calculated gradient at the cell face (denoted by over-bar in the
above expression and obtained by interpolation) only accounts for the
cross-derivative. The additional terms in the momentum equations
can be obtained by interpolating cell-center values, as they cause no
problems.
The convective uxes are non-linear terms, so linearization is nec-
essary. The simplest and most widely used approach is the Picard-
iteration scheme, i.e. the mass ux through the cell face is taken from
8
the previous iteration:
Z Z
(v vb )  n dS  k (v vb )  n dS = k m_ k : (13)
Sk Sk
The calculation of mass uxes is described below. Source terms may
also be non-linear; a similar approach can be used to linearize them.
2.2 Approximation of Time Integral
In the case of unsteady ows, the integration needs also to be per-
formed in time. Explicit schemes, which compute the solution at the
new time level at each node using only solutions at preceding time
levels, are very simple but restrictive regarding the allowable time
step size due to stability limitations. This is why implicit schemes
are often preferred, although they require both more memory and
computing e ort per time step: they usually have no step-size limita-
tions. Especially when the unsteadiness of the ow is less important
than its spatial distribution (e.g. in steady or periodic ows), implicit
schemes are more ecient.
The simplest implicit scheme is the rst-order Euler scheme: the
surface and volume integrals are calculated at the new time level, so
the solution from the previous time step appears only in the unsteady
term:
1 Z tn+1 dt  Cn+1 Cn ; (14)
t tn t
where the superscript n denotes the time levels. Here, stands for
the volume integral in the unsteady term of Eqs. (1{3); it involves
only the variable values at CV-center. This scheme is simple and un-
conditionally stable, but the rst-order error term introduces serious
numerical di usion e ects, so the scheme is only useful when time
accuracy is not important (e.g. when marching toward a steady solu-
tion). When the time history of ow is important, at least a second-
order scheme is necessary. One of the simplest implicit second-order
schemes is the three-time-levels scheme. It implies integration over a
time interval t centered around the time level tn+1 . The surface and
volume integrals at the time tn+1 are then centered approximations
with respect to time, so multiplying them by t is a second-order
9
approximation of the time-integral. A second-order approximation
of the time derivative at time tn+1 is achieved by tting a parabola
through solution at three time levels, thus:
1 Z tn+1=2 @ dt  3 Cn+1 4 Cn + Cn 1 : (15)
t tn 1=2 @t 2 t
This scheme is as simple as the implicit Euler scheme (one only needs
to store the solution at one more level; there is no computational over-
head). It is also less prone to oscillations than the Crank-Nicolson
scheme, another popular second-order method. More details about
various time-integration techniques can be found in Ferziger and
Peric1 and other textbooks.
2.3 The E ects of Grid Movement
In interface-tracking methods, the grid is tted to the free surface and
must follow its movement, which requires the use of a moving grid.
The space-conservation law must than be taken into account. When
the mass-conservation equation (1) for a uid of constant density is
written as:
d Z dV Z v  n dS + Z v  n dS = 0 ; (16)
dt b
V S S
we see that the rst two terms represent the SCL and add up to zero,
cf. Eq. (5), leading to:
Z
v  n dS = 0 or div v = 0 : (17)
S
It is therefore important to ensure that the above two terms can-
cel out in the discretized equations as well (i.e. the sum of volume
uxes through CV faces due to their movement must equal the rate of
change of volume); otherwise, arti cial mass sources are introduced
and they may accumulate and spoil the solution, as demonstrated by
Demirdzic and Peric9 .

10
tn+1
swept volume
δVkn
tn
k

tn-1

Figure 2: A control volume at three consecutive time levels; shaded areas


show volumes swept by the face over one time step.

Using the discretization method described above, the SCL equa-


tion can be cast into the following form:
3 V n+1 4 V n + V n 1 X [(v  n) S ]n+1 = 0 : (18)
2 t b k k
k
The di erence between CV volumes at consecutive time levels can be
expressed as the sum of volumes Vk swept by each CV face when
moving from its old to the new position, see Fig. 2, i.e.:
X
V n+1 V n = Vkn : (19)
k
When this expression is introduced in Eq. (18), one nds out that
the SCL is satis ed identically if the volume uxes through cell faces
are de ned as:
n n 1
V_ kn+1 = [(vb  n)k Sk ]n+1  3 Vk 2 V
t
k : (20)
Therefore, the volumes swept by each face over one t, Vk , should
be calculated using grid positions at two consecutive levels and used
to calculate volume uxes V_ kn+1 ; the velocity of the CV-surface, vb ,
need not be explicitly de ned. The same approach can be adopted

11
to any other time integration scheme.
2.4 Algebraic Equation Systems
Upon separation of the terms involving unknown variables and the
explicitly calculated ones, an algebraic equation of the form
X
A C C + A k N k = Q C (21)
k
is obtained for each CV. The coecients Ak contain contributions
from surface integrals over faces common to the cell around node C
and the corresponding neighbors Nk ; AC contains in addition con-
tributions from the unsteady term and possibly from source terms
(volume integrals). QC contains all terms which are treated as known
(source terms, parts of surface integrals treated explicitly as deferred
corrections, part of the unsteady term).
For the solution domain as a whole, the algebraic equation system
can be written as:
A = Q ; (22)
where A is a square M  M coecient matrix,  is the vector of un-
knowns, Q is the vector of right-hand sides, and M is the number of
CVs. The global matrix A is irregular if an unstructured grid is used;
for structured grids, the matrix has a diagonal structure for which
many solvers exist. Fortunately, some of the most ecient solution
methods for linear equation systems can be applied to unstructured
grids as well as to the structured ones, e.g. the conjugate gradient10
and multigrid methods1 .
2.5 Calculation of Pressure
Incompressible ows are special in the sense that the continuity equa-
tion can not be used to compute density (as is the case for compress-
ible ows), but represents an additional constraint on the velocities,
which can be calculated from the corresponding momentum equa-
tions. However, all equations will be satis ed only for a suitable
pressure distribution. Although it is in principle possible to construct
an equation for pressure by taking the divergence of the vector form
12
of momentum equation1 , this is seldom done; it is easier to construct
a pressure or pressure-correction equation using discretized momen-
tum and continuity equations. The so called SIMPLE algorithm11 is
the most widely used approach; Ferziger and Peric1 describe it in de-
tail together with several alternatives, including the fractional-step
method, another popular approach for computing unsteady incom-
pressible ows.
The solution process starts with a guessed pressure eld. Each
time the linearized momentum equations are solved, the mass con-
servation is imposed on the new velocities (to within a certain toler-
ance) by applying a velocity correction, which is proportional to the
gradient of the pressure correction, as dictated by the momentum
equations. Special care is needed in a colocated variable arrange-
ment to avoid pressure-velocity decoupling; see Ferziger and Peric1
for a detailed discussion on this issue.
The velocity eld obtained by solving the linearized momentum
equations is denoted by ui . The normal velocity component at the
cell face k is calculated by interpolating neighbor nodal values and
subtracting a correction term which should detect oscillations and
help smooth them out12 :
 V  2 p 3
(rp)  (rNk rC ) 5 :
(v n)k = (v  n)k A 4 Nk pC
(
C k Nk C r r )  n (rNk rC )  n
(23)
The over-bar denotes interpolation from neighbor nodal values (here
linear; see Lilek and Peric5 for details about using higher-order inter-
polation and integration techniques); double over-bar denotes arith-
metic averaging. The correction term in square brackets vanishes
when the pressure variation is linear or quadratic; it is proportional
to the square of mesh spacing and the third derivative of pressure1 .
Thus, this term is a second-order correction which goes consistently
towards zero as the grid is re ned; it is large only when the pressure
variation is non-smooth.
The normal velocity component is proportional to the normal
derivative of pressure, which can be approximated at the cell-face

13
center by the following central-di erence approximation:
 @p  pN0k pC0

@n k (rNk rC)  n ; (24)

where the auxiliary nodes C0 and N0k are located at the normal pass-
ing through the cell-face center according to the projection of the
vector rNk rC onto that line, see Fig. 1. Here, an approximation is
made in that the pressure values at locations N0k and C0 are replaced
by pressure values at Nk and C. This does not a ect the convergence
of the procedure if the non-orthogonality is not severe (i.e. when the
angle between vectors rNk rC and n is smaller than 45 ). The use
of the correct normal derivative would complicate the resulting alge-
braic equation; the error can be eliminated by applying an iterative
correction1 .
The mass uxes calculated using the above normal velocity do
not { in general { satisfy the mass conservation law. They need
to be corrected by invoking a pressure-correction, in the spirit of
the SIMPLE method11 . From the above equation one can derive
a relation between the corrections of the normal velocity and the
pressure-derivative:
 V  p0 0
0
(v  n)k  A Nk pC : (25)
jr r j
C k Nk C
The mass ux is calculated using the above approximations as follows:
h i
m_ k = k (v  n)k + (v0  n)k V_k Sk : (26)
The requirement that the discretized mass conservation equation is
satis ed:
(3 V n+1 4 V n + V n 1) + X m_ = 0 ; (27)
2 t k
k
and the introduction of the above-de ned expressions for V_ k , (v  n),
and (v0  n), see Eqs. (20){(25), leads to the discrete form of the Pois-
son equation for pressure-correction. It has the same form as Eq.
14
(21); the expressions for the coecients Ak are easily derived from
Eq. (26) and (25). Note that, for ows with constant density, the
unsteady term cancels out with the moving grid contribution to mass
uxes, as outlined above.
2.6 Boundary and Initial Conditions
The initial elds of all variables at t = t0 must be known; they have
to satisfy all the equations. In addition to the initial conditions, at all
future times the boundary conditions must be prescribed. This basi-
cally means that the surface integrals over cell faces lying in the solu-
tion domain boundary must be calculated using prescribed boundary
data: either variable values (Dirichlet conditions) or their gradients
(Neumann conditions). The implementation of these conditions in
FVM is described in detail in many books and will not be repeated
here; see Ferziger and Peric1 for more information. The conditions
at the free surface and their implementation will be described below.
2.7 Solution Algorithm
The solution algorithm follows the well-known SIMPLE-pattern and
can be summarized as follows:
 First, all variables are assigned initial values at t = t0. Time is
then advanced to t1 = t0 +t, the grid is moved to the new position
where and if necessary, and an iterative procedure is started in
order to nd solution of the coupled non-linear equations at the
new time level. Within each time step, the following steps are
repeated:
{ The momentum equations are discretized and linearized, lead-
ing to an algebraic equation systems for each velocity compo-
nent, in which the pressure, the other velocity components, and
all uid properties are treated as known (values from the pre-
vious iteration are used). These linear equation systems are
solved iteratively in turn to obtain an improved estimate of the
velocity at the new time level, ui . The iterations in the linear
equation solver are called inner iterations. There is no need to
15
solve these equation systems very accurately, since the equa-
tions were linearized and de-coupled; experience shows that
it is enough to reduce the residual level by an order of magni-
tude. Here the methods from the conjugate-gradient family are
used: incomplete Cholesky preconditioned conjugate gradient
(ICCG) method for the pressure-correction equation (which has
a symmetric coecient matrix), and the CGSTAB-method10
for the other equations. These iterative solvers are robust (no
parameters which need to be tuned) and can be used on any
grid.
{ The improved velocity eld is used to calculate new mass uxes
through CV faces and to invoke the mass-conservation equa-
tion; the result is the pressure-correction equation. Upon solv-
ing it for p0 , the mass uxes, CV-center velocities, and pressure
are corrected using expressions given above. Only an -fraction
of p0 is added to pressure, where  (0:1 0:8); smaller values
are used for large time steps and highly non-orthogonal grids.
{ If necessary, additional transport equations are solved in the
same manner to obtain better estimates (e.g. for temperature,
turbulence quantities, species concentration, etc).
{ If necessary, uid properties are updated (e.g. the density, vis-
cosity, and Prandtl number may be a function of temperature
or turbulence quantities).
This completes one outer iteration; the above steps are repeated
until residual level before the rst inner iteration in each equation
becomes suciently small. Usually, a reduction by three to four
orders of magnitude corresponds to a convergence error of the
order of 0.1 %.
 When the non-linear, coupled equations are satis ed to a desired
tolerance, the time is advanced by another t, and the above de-
scribed process is repeated. Usually, the solution from the previous
time step serves as the initial guess for the new solution.
Unless time steps are small enough, the momentum equations
(and usually other transport equations too) require an under-relaxation

16
of outer iterations1 . Typical under-relaxation factors range between
0.7 and 1.0, depending on the time step size.

3 Interface-Tracking Methods
Many methods of this type were developed in the past. Most of them
were designed for use in conjunction with structured grids and nite-
di erence or nite-volume methods, with grid nodes being allowed to
move vertically. They were thus limited to ows with a smooth free
surface (due to the limitation of the interface-tracking approach) and
in relatively simple geometries (due to the limitations of structured
grids and a prescribed direction of movement of grid nodes). The
applications ranged from sloshing ows with a moderate amplitude,
over ows in open conduits to ows over or around fully or partly
submerged bodies (hydrofoils under free surface, ows around simple
ship bodies etc.)13 17 .
We shall present here an interface-tracking method developed by
the present authors18 , which can be applied to both structured and
unstructured grids. The di erences and similarities compared to
other methods of the same kind will also be mentioned. Applica-
tions of the method will be presented later, together with the results
obtained using an interface-capturing method.
n
s
n

fσ s

t dS

dl Rt Rs
t
dl
fσ fσ

Figure 3: On the description of boundary conditions at the free surface.

The following boundary conditions have to be satis ed at the free


surface:
17
 The kinematic condition, which implies that the free surface is an
interface between two uids with no ow through it, i.e.:
[(v vb )  n]fs = 0 or m_ fs = 0 ; (28)
where \fs" denotes the free surface.
 The dynamic condition, which implies that the forces acting on
the uid at the free surface are in equilibrium. In addition to the
forces due to pressure p and viscous stresses ij , surface-tension
forces may also be present. The surface tension coecient  may
be a function of temperature or concentration of chemical species.
With reference to Fig. 3, the equilibrium of forces in the normal
and two tangential directions at the free surface can be expressed
as:
(n  T )l  n = (n  T )g  n + K ;
(n  T )l  t = (n  T )g  t + @
@t ; (29)
(n  T )l  s = (n  T )g  s + @
@s :
Here  is the surface tension coecient, n, t and s are the unit
vectors in the local orthogonal coordinate system (n; t; s) at the
free surface (n is normal to the free surface and directed from liquid
to gas), the indices `l' and `g' denote liquid and gas, respectively,
and K is the curvature of the free surface,
K= 1 + 1 ;
Rt Rs (30)
with Rt and Rs being radii of curvature along coordinates t and
s, see Fig. 3. In many cases the surface tension forces can be
neglected. The tangential viscous forces are most often neglected
in the absence of surface tension gradients; the normal viscous
stresses at the free surface are also small and can often be ne-
glected. The dynamic conditions reduce then to the statement
that the pressures on both sides of the free surface must be the
same.
18
Since one is usually interested in liquid/gas interfaces, the ow
of liquid is computed alone and the gas pressure is prescribed along
the free surface. The iterative nature of the segregated solution algo-
rithm is well suited for the implementation of free-surface boundary
conditions. The predictor-corrector scheme inherent to all SIMPLE-
type methods is used to force the satisfaction of both the kinematic
and dynamic conditions at the free-surface boundary.
The dynamic boundary condition is implemented by treating the
free surface as a boundary with a prescribed pressure. This means
that in the discretized continuity equation the velocity of uid at the
free surface boundary is obtained by extrapolation from the interior;
Eq. (23) applies, with interpolation between two nodes being replaced
by a one-sided extrapolation, and k is denoting a cell face located at
the free surface, with Nk being the boundary node. This boundary
velocity may be corrected and the mass ux correction of the same
form as for the interior cell faces results, see Eq. (25). The only
di erence is that the pressure correction p0Nk = 0 for nodes located
at the free surface boundary. The pressure-correction equation thus
has a Dirichlet boundary condition at the free surface.
With this approach, at the end of a SIMPLE-step (one outer
iteration), the dynamic condition is satis ed, but not the kinematic
condition. The cell-face velocities at the free surface, corrected upon
solving the pressure-correction equation in order to enforce local and
global mass conservation, may lead to non-zero mass uxes m_ fs. In
order to satisfy the kinematic condition, the vertices which de ne
the boundary cell face have to be moved so that the correction of
the volume swept by the free surface, V_ fs0 , compensates the mass ux
obtained in the preceding step:
m_ fs + V_ fs0 = 0 : (31)
This is not a trivial task, since vertices which de ne a boundary cell
face cannot be moved without changing shapes and positions of all
neighbor cell faces. Also note that there are more vertices in the free
surface than there are cell faces and Eqs. (31).
Here, the necessity of iteration is used to simplify the calculation
of the free surface position. The rst estimate of the free surface
movement is obtained by assuming that the volume V_ fs0 results from
19
displacing the boundary cell face Sfs by h in the direction de ned
by a unit vector efs . The displacement h obtained from this ap-
proximation is:
_0
h = fs S Vfsn et ; (32)
fs fs
where 0 < fs  1 is an under-relaxation factor and t is the time
step. The direction of efs is arbitrary, but in order to deliver a good
approximation of h, it should be parallel to the direction of the
resultant body force.
efs
efs efs
V3
∆h B3 n fs
V2 B2
B1 V4
efs
Grid before adjustment
∆h
Grid after adjustment B4
Computed volume flux through free surface
Control point before adjusment
Control point after adjusment
Vertices at free surface before adjustment
Vertices at free surface after adjustment

Figure 4: On the movement of control points and grid nodes at the free
surface.

The calculation of h is thus local and inexpensive. However, its


direct implementation results in a discontinuous numerical grid and
a non-smooth free surface. In order to preserve the local and com-
pact nature of the scheme, free-surface control points are introduced
for each boundary cell face which coincides with the free surface, as
shown in Fig. 4. Their position is controlled by the displacements
h calculated from Eq. (32), while they control the movement of the
vertices which de ne the free-surface boundary cell faces, as shown
in Fig. 4. Note that efs need not be the same at each control point.

20
The position vectors of the free-surface control points rBi are ad-
justed according to the expression:
rkB+1
i = rBi + h efs ;
k (33)
where k is here the counter of the outer iterations within one time
step. The situation after applying the correction (33) is shown schemat-
ically in Fig. 4. The next step is to reconstruct the free surface by
modifying the position vectors of vertices rVi that de ne the free-
surface boundary (see Fig. 4), according to the following expression:
" n #
+1 X +1
rVi = rVi efs  rVi
k k k wm rBm efs ;
k (34)
m=1
where n is the number of free-surface boundary cell faces that share
vertex Vi , and wm are interpolation weighting factors; usually, linear
interpolation is used.
Additional corrections of the free surface movement can be su-
perimposed on that controlled by the kinematic boundary condition
described above. The purpose of those corrections is usually to keep
the free surface in contact with solid boundaries and to maintain
good grid quality. However, both these additional corrections and
the ones calculated from Eq. (34) have to become negligible as the
outer iterations converge within each time step.
The volumes swept by free-surface cell faces, V_ fs , are now obtained
using the current vertex coordinates and those from the end of the
preceding time step. In general, the swept volumes computed in this
way, and the new velocity eld at the free surface, do not result in zero
mass uxes (31) through the free surface. However, by repeating the
iterative procedure described above, these mass uxes can be made
arbitrarily small.
The iterative procedure for correcting the free surface and its
incorporation into outer iterations can be summarized as follows:
1. Solve the momentum equations using the geometry de ned by the
current shape of the free surface and the prescribed pressure at it.
2. Enforce local mass conservation in each CV by solving the pressure-
correction equation, using the prescribed pressure boundary con-
dition at the current free surface. Mass is conserved both globally
21
and in each CV, but the non-zero mass uxes through the free
surface may result.
3. Correct the position of the free surface, as described above, so
that the volume de ned by its corrected and previous position
compensates the mass uxes through the free surface obtained in
the preceding step.
4. Return to step 1 and repeat until all equations and boundary
conditions are satis ed (i.e. until all corrections become negligibly
small).
5. Advance to the next time step.
The numerical grid in the interior of the solution domain has to
respond to the movement of the vertices in the free surface, in order
to preserve its good quality. It is dicult to device a simple and
general technique which can be successfully used in all situations and
for all geometries for this purpose. An elliptic grid smoother has been
used by present authors to adjust the inner points to the movement
of nodes lying in the free surface (which need not be done in each
time step if the movements are small relative to CV size). In the case
of structured grids, one can simply re-adjust the location of the inner
nodes to preserve the same relative position between the free surface
and the opposite boundary. However, a problem-dependent solution
may sometimes be necessary if the geometry of the computational
domain is complicated.
The movement of cell vertices which are common to free surface
and other boundaries like inlet, outlet, wall, and symmetry deserves
special attention. As outlined above, the kinematic condition is used
to move the control points, which are placed above cell-face centers.
The cell vertices between these control points are moved to locations
obtained by interpolating coordinates of neighbor control points, but
it remains open how to move the points at the edges of the free
surface. In some cases (e.g. at the edges between free surface and
inlet boundaries), these locations are xed so the edge vertices are
not moved. In other cases (e.g. where the free surface is in contact
with solid walls), one may assume that the edge vertex moves by the
same amount as the control point next to it, as shown in Fig. 4. In
22
the case of outlet boundaries and a subcritical ow regime (Froude
number smaller than unity), the solution is especially sensitive to
the movement of grid points. This is due to the fact that the ow
then depends on the conditions beyond the outlet cross-section. The
uniqueness of the solution can be enforced by specifying the liquid
height at the outlet, which is dicult if waves are present. Some of
the possible approaches will be described when presenting results of
computations for test cases involving outlet boundaries.
Boundary control volumes Computational node
Interior control volume Cell-face center
Free surface
Wall

Figure 5: On the element-based nite-volume method15 .

Another interface-tracking method that { although presented for


two-dimensional problems and a vertical movement of grid nodes {
could be adapted to three-dimensional and/or unstructured grids is
that of Raithby et al.15 . They used an element-based FVM, in which
the grid de nes elements over which the shape functions (or interpo-
lations functions) are prescribed, and the control volumes are con-
structed around each grid node as shown in Fig. 5. The control
volumes are also constructed around boundary nodes, which at the
same time de ne the geometry of the free surface; therefore, there is
no need to introduce control points. The combination of the continu-
ity equation and the SCL enables the derivation of an equation from
which the hight at each free-surface node can be computed, as the
grid velocity can be expressed as the rate of change of height. Since
the pressure at the free surface is prescribed, the continuity equation
for CVs around boundary nodes is not used to compute the pres-

23
sure correction (which is zero at the free surface), but to obtain the
node displacement necessary to keep the kinematic condition satis-
ed. Note that the surface of a control volume consists of more faces
than in the present approach (eight for quadrilaterals in 2D, 24 for
hexaedra in 3D, compared to four and six in the method described
above). This makes the computation of the coecient matrix and
source terms more expensive, which is probably partly compensated
by a higher accuracy than in the present method (for the same grid
and approximations of integrals).
We should also mention the nite-element methods developed and
used for special problems in coating ows (2D, low Reynolds number
ows, strongly a ected by surface tension)19 .

4 Interface-Capturing Methods
One of the earliest interface-capturing methods was the MAC-scheme20 .
In it, massless particles are introduced into the liquid phase near the
free surface initially and their movement is being followed. The com-
putational domain extends beyond the free surface and its location
is determined according to the distribution of marker particles.
The MAC scheme is attractive because it can treat complex phe-
nomena like wave breaking. However, the computing e ort is large,
especially in 3D, since in addition to solving the equations governing
uid ow, one has to follow the motion of a large number of particles.
Another class of interface-capturing methods was initiated by the
introduction of the VOF-scheme21 ; many similar methods have been
presented since then. In addition to the conservation equations for
mass and momentum, one introduces and solves an equation for the
void fraction of one phase, c. The grid extends over both liquid and
gas phase; one sets e.g. c = 1 for CVs lled by liquid and c = 0 in CVs
lled by gas. The change of c is governed by the transport equation:
d Z c dV + Z c(v v )  n dS = 0 : (35)
dt b
V S
Since the grid does not follow the deformation of the free surface,
the grid movement is only necessary if the solution domain changes
24
shape due to the movement of solid walls.
The critical issue in this type of methodS is the discretization of
convective term in Eq. (35). First-order upwind scheme smears the
interface too much and introduces arti cial mixing of the two uids.
Since c must obey the bounds 0  c  1, one has to ensure that the
scheme does not generate overshoots or undershoots. Fortunately,
it is possible to derive schemes which both keep the interface sharp
and produce monotone pro les of c across it; see Ref. 22 for some
examples and Ref. 23 for a scheme speci cally designed for interface-
capturing in free surface ows; another example will be presented
below.
This approach is more ecient than the MAC scheme and can
be applied to complex free surface shapes including breaking waves.
However, the free surface contour is not sharply de ned; it is usually
smeared over one to three cells (like shocks in compressible ows).
Local grid re nement is important for accurate resolution of the free
surface. The re nement criterion is simple: cells with 0 < c < 1 need
to be re ned.
There are several variants of the above approach. In the origi-
nal VOF-method21 , Eq. (35) is solved in the whole domain to nd
the location of the free surface; the mass and momentum conserva-
tion equations are solved for the liquid phase only. The method can
calculate ows with overturning free surfaces, but the motion of gas
enclosed by the liquid phase is not taken into account.
Kawamura and Miyata24 used Eq. (35) to calculate the distribu-
tion of the density function, which is used to nd the location of the
free surface (contour of c = 0:5). The computation of the motion
of the liquid and gas phase is done separately, treating in each sub-
domain the free surface as a boundary at which the kinematic and
dynamic boundary conditions are applied. Cells which become ir-
regular due to being cut by the free surface require special treatment
(variable values are extrapolated to nodal locations lying on the other
side of the interface). The method was used to calculate ows around
ships and submerged bodies.
Alternatively, one can treat both uids as a single e ective uid
whose properties vary in space according to the volume fraction of

25
each phase, i.e.:
 = 1 c + 2 (1 c) ;  = 1 c + 2 (1 c) ; (36)
where subscripts 1 and 2 denote the two uids (e.g. liquid and gas).
If one CV is partially lled with one and partially with the other uid
(i.e. 0  c  1), it is assumed that both uids have the same velocity
and pressure. The free surface does not represent a boundary and no
boundary conditions need to be prescribed at it. If surface tension is
signi cant at the free surface, this can also be taken into account by
transforming the resulting force into a body force25 . These methods
can also deal with merging and fragmentation in multiphase ows26 .
We present below a method of the above kind, which is similar to
that of Ubink23. It can be applied to both structured and unstruc-
tured grids with arbitrarily shaped CVs and can also be extended
to treat ows in which more than two immiscible uids are involved.
However, for the sake of simplicity we shall describe it for two uids
and a 2D-grid made of quadrilaterals.
As noted above, the solution domain extends over both uids, and
all the conservation equations are solved in the whole domain. At the
initial time step, the distribution of c is prescribed, de ning the initial
location of liquid and the shape of the free surface. The discretization
of the transport equation for c, Eq. (35), requires special care. This
is due to the fact that c must be bound by zero and unity, and that
the region in which the cells are only partially lled should be as
small as possible. As in shock-capturing in compressible ows, the
discontinuity in uid properties is smeared and this smearing should
be controlled.
Equation (35) contains only convective uxes and the unsteady
term. The integration in time can be selected according to the re-
quirements on the accuracy of time history of solution: if only a
steady solution is sought or slowly-varying ows are considered, the
fully-implicit Euler method is appropriate. For ows in which time
evolution of the free surface is important, second-order schemes like
Crank-Nicolson, implicit three-time-levels scheme etc. should be used.
The choice of approximations for convective uxes is less obvi-
ous. The only scheme which unconditionally satis es the bounded-
ness criterion is the rst-order upwind scheme; however, it can not
26
be used due to excessive numerical di usion, which smeares the in-
terface so badly that the two uids mix over a wide region. On
the other hand, any of the higher-order methods mentioned above
tends to produce over- and undershoots in the vicinity of disconti-
nuities, so special care is needed. One can resort to a wide range of
special bounded schemes developed for applications in aerodynamics,
like total-variation-diminishing (TVD) and essentially non-oscillating
(ENO) schemes; see Ref. 27 for examples. However, the interface-
capturing in free-surface ows has some specialities which need be
considered. One comes from the fact that the convective ux out of
one CV must not transport more of one uid than is available in the
donor cell. How the cell-face value of c is computed must also take
into account the orientation of the interface and the local Courant
number.
c~j

1 HRIC
DDS

CDS UDS

0
0 1 c~C

Figure 6: The normalized variable diagram (NVD)22 .

The sharpnes of the interface without over- and undershoots can


be achieved by limiting the approximation of the cell-face value (see
Eq. (7)) to lie in the shaded area of the so called normalized variable
diagram22 (NVD) shown in Fig. 6. The local normalized variable c~
in the vicinity of the cell-center C is de ned as:
c~(r) = c(r) cU ;
cD cU (37)
where subscripts `U' and `D' denote nodes upstream and downstream
27
U C j D U C D
j

c c
a) b)

gradc
j θ
C nj c)

Figure 7: On the computation of cell-face volume fraction.

of the cell-center C. Should the cell-center value c~C turn out to be


smaller than zero or larger than unity, this means that the pro le of
c is not monotone and we need numerical di usion to get rid of oscil-
lations; for values of c~C between zero and unity, one can choose any
dependency from the shaded region of the NVD-diagram. The partic-
ular choice chosen here is indicated in Fig. 6 as HRIC (high-resolution
interface capturing) scheme. The reasoning for such a choice can be
seen from Fig. 7: if the cell is almost empty, only the uid present
at the downstream cell will be convected through the cell face. This
use of the downwind scheme keeps the interface sharp. However, this
applies only if the interface is parallel to the cell face and moves in
the direction of the cell-face normal; if the interface is perpendicular
to the cell face, it is likely that the convected uid would be of the
same composition as in the cell center, so upwind scheme is appropri-
ate. Also, downwind approximation must not drain more uid than
is available in the cell, which depends on the local Courant number,
Co = v  nVSj t : (38)
C

28
There is no obvious way how to vary the approximation of cj de-
pending on the Courant number and interface orientation; we use
the following ad-hoc corrections which proved satisfactory in most
applications:
8
>
> c~ if c~ < 0
< 2~CcC if 0  c~CC  0:5
>
c~j = > 1 if 0:5  c~  1 (39)
> C
: c~C if 1  c~C

8
>
< c~j if Co < 0:3
c~j = > 0 : 7 Co
c~C + (~cj c~C ) 0:7 0:3 if 0:3  Co < 0:7 (40)
: c~C if 0:7  Co
p p
c~ 
j = c~j cos  + c~C (1 cos ) : (41)
Here  represents the angle between the normal to the interface (rep-
resented by the gradient vector of c) and the normal to the cell face,
see Fig. 7. Finally, the cell-face value of c is computed according to
Eq. (37) as:
cj = c~
j (cD cU ) + cU : (42)
One of the motivations for the choice of the above approximations
and corrections was the simplicity. Obviously, many other choices
are possible; some are demonstrated by Ubbink23.
The discretized equation for c has the same form as all other
equations and is solved using the same linear equation solver. The
outer iteration loop is extended for solving the equation for c after
the pressure-correction equation is solved; the rest of the sequential
solution algorithm remains the same, as described in Sect. 2.7.

5 Application Examples
Four examples have been chosen to demonstrate the application of
the methods described above. In two cases, both interface-tracking
and interface-capturing methods could be used, while the other two

29
cases involve overturning free surfaces and can be treated only by the
latter approach.
5.1 Critical Flow Over a Semi-Circular Obstacle
A turbulent critical ow over a semi-circular obstacle on the oor
of an open channel was studied theoretically and experimentally by
Forbes28 . We used both interface-tracking and interface-capturing
methods to compute this ow.

Figure 8: Initial (upper) and nal (lower) grid for the ow over a semi-
circular obstacle on the oor (computation using interface-tracking ap-
proach).
Depending on the conditions before and after obstacle, various
ow regimes can be obtained. In the case studied here the ow is
subcritical ahead of obstacle, with
Fr = pv = 0:6 ; (43)
gH
where v is the ow velocity, g is the gravity acceleration and H the
undisturbed water depth. Due to the constriction of the ow cross-
section, uid accelerates above obstacle, while water depth reduces;
the Froude number rises and may become larger than unity. If that
happens (which depends on the downstream conditions, similar to
compressible ows through nozzles), the uid accelerates further and
30
the water level drops to a much lower level behind the obstacle than
before it. The case studied here is the critical one, in which the
subcritical (Fr < 1) ow turns supercritical (Fr > 1). At the out-
let boundary, the pressure was initially prescribed according to the
expected free-surface height; once the velocity became large enough
(Fr > 1), one could switch to the downstream extrapolation of both
pressure and free-surface height, as appropriate for a supercritical
ow regime.

Figure 9: Numerical grid and predicted disctribution of volume fraction


for the ow over a semi-circular obstacle on the oor (computation using
interface-capturing approach, with a local mesh re nement along interface)
Both methods predict the free-surface shape correctly. Figure 8
shows the initial and the nal grid shape obtained with the interface-
tracking method, while Fig. 9 presents the volume fraction distribu-
tion computed using the interface-capturing method. The shape of
the free surface agrees well with experimental data in both cases,
as can be seen from Fig. 10. It also turns out that the free surface
shape is not very sensitive to the grid neness { even the coarsest grid
predicts the shape of the free surface to within few percents correctly.
The ow separates from cylinder surface and recirculates behind
it. The k{ model with wall functions29 was used to model the turu-
lence e ects. Although in the present case the computed free-surface
shape is not sensitive to the turbulence model used, there are cases
in which turbulence modelling errors can signi cantly a ect the pres-

31
2.4
2.2 Experiment
2 Interface-capturing
Interface-tracking
1.8
1.6
y/R

1.4
1.2
1
0.8
0.6
-6 -4 -2 0 2 4 6 8
x/R

Figure 10: Comparison of predicted water surface elevation for the ow over
a semi-circular obstacle on the oor and comparison with experimental data
of Fobres28.

sure distribution and the shape of the free surface.


5.2 Simulation of Water-Entry of a Wedge
Water entry of a solid body is a subject of great interest in marine
engineering. Depending on body shape, entry velocity, and acceler-
ation, very high pressures on body surface may arise, leading to a
material damage (e.g. ship slamming in rough sea). A large num-
ber of both experimental and theoretical studies has been devoted
in the past to studying water entry of simple bodies like a wedge,
cone, sphere etc. Potential ow methods were adapted to compute {
with success { such water-entry ows30 . However, these are usually
not applicable to at bodies or bodies of a complicated shape. We
show here the results of application of FVM for viscous ows using
interface-capturing approach to water entry of a wedge with a nose
angle of 120 . The wedge moves downward with a constant velocity
of 5.5 m/s. A cell-wise locally re ned numerical grid ( ne near body
and coarse further away; very ne near wedge base where separation
is expected) with 17 012 CV was used.
Figure 11 shows free-surface shape and pressure contours at two
instants: one before the wedge base passes the undisturbed free sur-
face, and one thereafter. Before the whole edge has entered water,

32
Figure 11: Free-surface shape (upper) and pressure contours (lower) at two
instants 0.001 seconds apart (left and right column).

pressure maximum is close to the free surface; afterwards, it moves to


the wedge nose, and the drag approaches asymptotically a constant
value, as shown in the plot of pressure force on wedge as a function
of time in Fig. 12.
The ow was computed over 2 000 time steps, with t = 5:0 
10 5 seconds (with the entry velocity of 5.5 m/s and wedge width of
500 mm, the entry process takes only 0.1 second). About 10 outer
iterations per time step were performed, requiring that all normalized
residual norms should be reduced below 10 4 . About 9 MB RAM and
30 hours of computing time were used up on a PC with a Pentium
166 MHz processor, running at about 10 MFlops in double precision.
Obviously, the computing e ort for simulating unsteady viscous

33
20000
18000
16000
14000
12000
Force (N)

10000
8000
6000
4000
2000
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Time (s)

Figure 12: Pressure force acting on the wedge as a function of time (constant
entry velocity).

ows using FVM and interface-capturing approach is much larger


than when potential ow is computed. However, the former can also
be applied to much more complicated ow problems where potential
ow methods are inapplicable, as will be demonstrated below.
5.3 Simulation of a Breaking-Dam Experiment
The spreading of a water column after one side wall is removed {
usually referred to as a breaking dam { is often used as a test case
for computational methods dealing with free surface ows. Here we
consider a problem investigated experimentally by Koshizuka et al.31 ,
in which the collapsing water column rst hits an obstacle on the
oor and then a solid wall. In this case it is important to consider
the ow of both water and air, especially when air becomes trapped
in water; due to a much lower density, trapped air is subjected to a
large buoyancy force and tends to rise up. This is obvious from Fig.
13, which shows the shape of water column at three time instants.
After water jet hits the wall, the air trapped between the wall and
obstacle escapes upwards breaking up the water layer. If the air ow
were not computed in a coupled manner with the water ow, the
results would have not corresponded with experimental observations

34
as well as is the case with present computations, cf. Fig. 13.

Figure 13: Comparison of experimental visualization by Koshizuka et al.31


(left) and present computation (right) of the breaking-dam ow at three
time instants (0.2, 0.4, and 0.5 seconds)
This example demonstrates the large potential of interface-capturing
methods for computing free-surface ows with severe interface defor-
mations. Space does not permit presentation of all applications, but
we mention successful prediction of sloshing in tanks, ows over hy-
drofoils under free surface, lling and emptying of various vessels,
breaking waves in costal region etc. Ubbink30 predicted rising of a
bubble in a pipe, ow under a sluice gate, and hidraulic jump prob-
lems. Flows with dominating e ects of surface tension { which can
also be treated by interface-capturing methods of the kind presented
here { have been studied by Lafaurie et al.26 and Zaleski et al.32 .

35
5.4 Three-Dimensional Flow Around a Ship Hull
Finally, we show the results of computations of three-dimensional vis-
cous ow around ship hull. While the interface-capturing method can
be applied to hulls of any shape, interface-tracking methods have been
used so far only for ship hulls of relatively simple shape which allows
an easy adaptation of grid to the hull surface. Hulls with bulbous
bow and complicated stern form, on the other hand, pose diculties
to interface-tracking methods, since the free surface changes its shape
so much in the bow and stern regions that an adaptation of grid lines
to both free surface and hull shape becomes extremely complicated.

Figure 15: Computed wave pattern for the ow around Wigley ship hull
(interface-tracking method, grid with 192 000 CV; from Azcueta et al.33 ).
We show here application of both methods to a rather simple hull
form known as Wigley-hull. Its shape is given by the equation
"  2 # "  2 #
B
y = 2 1 2Lx 1 Dz ; (44)
where L, B , and D stand for length, width (y-direction), and depth
(z -direction), respectively. Figure 15 shows water surface elevation
36
around one half of the hull (steady mean ow was computed around
one half of the hull only due to its symmetry). This pattern agrees
well with experimental observations. In Fig. 16 free surface elevation
along hull surface, as computed using both interface-tracking and
interface-capturing methods, is compared with experimental data.
For interface-capturing methods, free surface is assumed to be located
where the volume fraction c has the value 0.5, which was found by
linear interpolation between nodal values.

Experiment
Interface-capturing, 192000 CV
0.01 Interface-capturing, 24000 CV
Interface-tracking, 479000 CV

0.005
z/L

-0.005

0 0.2 0.4 0.6 0.8 1


x/L

Figure 16: Comparison of predicted and measured free surface elevation


along Wigley hull surface for the ow at Froude number of 0.267 (interface-
tracking by Lilek34 , interface-capturing by Azcueta et al.33 )
It is worth noting that both approaches predict the shape of the
wetted hull surface well. Interestingly, interface-capturing method
agrees somewhat better with experimental data even though the grid
used was coarser. Especially important for practical applications in
design is the fact that even a very coarse grid with only 24 000 CV
leads to a relatively accurate prediction of water surface elevation.

6 Conclusions
Finite-volume methods for the computation of viscous ows can be

37
extended to yeald the location of the free surface by various ap-
proaches. We have discussed here some general features of two broad
categories of such extensions, namely interface-tracking and interface-
capturing methods. One variant of each method, as developed and
used by the present authors, was described in more detail. The main
conclusions can be summarized as follows:
 Interface-tracking methods resolve the interface very sharply, since
it is treated as a boundary of the solution domain. The approach
is suitable for smooth free surface, when it is important to sharply
resolve the interface, and when the geometry of ow domain is
not too complicated so that grid-adaptation to both free surface
and solid walls can be accomplished automatically during compu-
tation. Computation of both liquid and gas ow is possible, but
it requires a coupled solution in two separated domains with a
common moving boundary (free surface).
 Interface-capturing methods do not resolve the interface sharply;
it is captured at best between one to three CVs. Both liquid and
gas ow are computed. This is important when gas is trapped
in liquid or when gas ows with a high velocity. Since the grid
does not have to be adapted to the shape of the free surface,
problems with grid adaptations mentioned above are avoided. The
free surface may deform in an arbitrary manner. Applications of
the present interface-capturing method have shown that the free-
surface shape is predicted with about the same accuracy as with
the interface-tracking method where both are applicable; however,
the applicability range is much larger for the former. Surface-
tension e ects can also be included.
In spite of generally positive results presented so far, there is much
scope for improvement of both interface-tracking and interface-capturing
methods. One is the marching towards steady state when steady
ows are considered; the fact that the free surface is evolving from
an initial (usually undisturbed) shape to the steady solution through
an unsteady ow computation means that the computing cost { es-
pecially in 3D { is relatively high. One possibility to reduce the
computing e ort is to compute the ow on a very coarse grid rst

38
and use this solution to determine the initial grid for the ner level.
The experience in this eld is also not as large as in ows without
free surface, so that modelling of turbulence in the vicinity of the free
surface { especially for interface-capturing methods { and the e ects
of the choice of various approximations (e.g. pressure terms in mo-
mentum equations, convective terms in the volume fraction equation,
gravity force etc.) need to be investigated in detail while searching
for improvements.
Both methods can be applied to both structured and unstruc-
tured grids. This fact and the possibility of cell-wise local adaptive
grid re nement are further areas in which considerable improvement
is expected in the future.

7 References
1. Ferziger, J.H., Peric, M.: Computational Methods for Fluid Dynamics,
Springer, Berlin, 1996.
2. Demirdzic, I., Muzaferija, S., Peric, M.: Advances in computation of
heat transfer, uid ow, and solid body deformation using nite vol-
ume approach, in W.J. Minkowycz, E.M. Sparrow (eds.), Advances in
Numerical Heat Transfer, chap. 2, pp. 59{96, Taylor and Francis, New
York, 1996.
3. Lilek, Z ., Muzaferija, S., Peric, M., Seidl, V.: An implicit nite-volume
method using non-matching blocks of structured grid, Numer. Heat Trans-
fer, Part B, Vol. 32, pp. 385{401 (1997).
4. Tu, J.Y., Fuchs, L.: Overlapping grids and multigrid methods for three-
dimensional unsteady ow calculation in IC engines, Int. J. Numer.
Methods Fluids, Vol. 15, pp. 693{714 (1992).
5. Lilek, Z ., Peric, M.: A fourth-order nite volume method with colocated
variable arrangement, Computers & Fluids, Vol. 24, pp. 239{252 (1995).
6. Khosla, P.K., Rubin, S.G.: A diagonally dominant second-order accurate
implicit scheme, Computers & Fluids, Vol. 2, pp. 207{209 (1974).
7. Peric, M.: A nite volume method for the prediction of three- dimen-
sional uid ow in complex ducts, PhD Thesis, University of London,
1985.

39
8. Muzaferija, S.: Adaptive nite volume method for ow predictions using
unstructured meshes and multigrid approach, PhD Thesis, University of
London, 1994.
9. Demirdzic, I., Peric, M.: Space conservation law in nite volume calcula-
tions of uid ow, Int. J. Num. Methods in Fluids, Vol. 8, pp. 1037{1050
(1988).
10. Van den Vorst, H.A.: BI-CGSTAB: a fast and smoothly converging vari-
ant of BI-CG for the solution of non-symmetric linear systems, SIAM J.
Sci. Stat. Comput., Vol. 13, pp. 631{644 (1992).
11. Patankar, S.V., Spalding, D.B.: A calculation procedure for heat, mass
and momentum transfer in three-dimensional parabolic ows, Int. J.
Heat Mass Transfer, Vol. 15, pp. 1787{1806 (1972).
12. Rhie, C.M., Chow, W.L.: A numerical study of the turbulent ow past
an isolated airfoil with trailing edge separation, AIAA J., Vol. 21, pp.
1525{1532 (1983).
13. Farmer, J., Martinelli, L., Jameson, A.: Fast multigrid method for solv-
ing incompressible hydrodynamic problems with free surfaces, AIAA J.,
Vol. 32, pp. 1175{1182 (1994).
14. Kawamura, T., Miyata, H.: Simulation of nonlinear ship ows by density-
function method, J. Soc. Naval Architects Japan, Vol. 176, pp. 1{10
(1994).
15. Raithby, G.D., Xu, W.-X., Stubley, G.D.: Prediction of incompressible
free surface ows with an element-based nite volume method, Comput.
Fluid Dynamics J., Vol. 4, pp. 353{371 (1995).
16. Hino, T.: Computation of viscous ows with free surface around an
advancing ship, Proc. 2nd Osaka Int. Colloquium on Viscous Fluid Dy-
namics in Ship and Ocean Technology, Osaka Univ. (1992).
17. Gentaz, L., Alessandrini, B., Delhommeau, G.: Motion simulation of a
cylinder at the free surface of a viscous uid, Ship Technology Research,
Vol. 43, pp. 3{18 (1996).
18. Muzaferija, S., Peric, M.: Computation of free-surface ows using nite
volume method and moving grids, Numer. Heat Transfer, Part B, Vol.
32, pp. 369{384 (1997).
19. Saito, H., Scriven, L.E.: Study of coating ow by the nite element
method, J. Comput. Physics, Vol. 42, pp. 53{76 (1981).

40
20. Harlow, F.H., Welsh, J.E.: Numerical calculation of time dependent
viscous incompressible ow with free surface, Phys. Fluids, Vol. 8, pp.
2182{2189 (1965).
21. Hirt, C.W., Nicholls, B.D.: Volume of uid (VOF) method for dynamics
of free boundaries, J. Comput. Phys., Vol. 39, pp. 201{221 (1981).
22. Leonard, B.P.: Bounded higher-order upwind multidimensional nite-
volume convection-di usion algorithms, in W.J. Minkowycz, E.M. Spar-
row (eds.), Advances in Numerical Heat Transfer, chap. 1, pp. 1{57,
Taylor and Francis, New York, 1997.
23. Ubbink, O.: Numerical prediction of two uid systems with sharp inter-
faces, PhD Thesis, University of London (1997).
24. Kawamura, T., Miyata, H.: Simulation of nonlinear ship ows by density-
function method, J. Soc. Naval Architects Japan, Vol. 176, pp. 1{10
(1994).
25. Brackbill, J.U., Kothe, D.B., Zemaach, C.: A continuum method for
modeling surface tension, J. Comput. Physics, Vol. 100, pp. 335{354
(1992).
26. Lafaurie, B., Nardone, C., Scardovelli, R., Zaleski, S., Zanetti, G.: Mod-
elling merging and fragmentation in multiphase ows with SURFER, J.
Comput. Phys., Vol. 113, pp. 134{147 (1994).
27. Hirsch, C.: Numerical computation of internal and external ows, vol. I
& II, Wiley, New York, 1991.
28. Forbes, L.K.: Critical free-surface ow over a semi-circular obstruction,
J. Eng. Math., Vol. 22, pp. 3{13 (1998).
29. Launder, B.E., Spalding, D.B.: The numerical computation of turbulent
ows, Comput. Meth. Appl. Mech. Eng., Vol. 3, pp. 269{289 (1974).
30. Zhao, R., Faltinsen, O.M., Aarsens, J.: Water entry of arbitrary two-
dimensional sections with and without ow separation, Proc. Twenty-
First Symposium on Naval Hydrodynamics, Trondheim, June 24{28,
1996.
31. Koshizuka, S., Tamako, H., Oka, Y.: A particle method for incompress-
ible viscous ow with uid fragmentation, Comput. Fluid Dynamics J.,
Vol. 4, pp. 29{46 (1995).

41
32. Zaleski, S., Li, J., Succi, S.: Two-dimensional Navier-Stokes simulation
of deformation and breakup of liquid patches, Phys. Rev. Lett., Vol. 75,
pp. 244-277 (1995).
33. Azcueta, R., Muzaferija, S., Peric, M.: Computation of ows around
ships using interface-capturing nite-volume method, Internal Report,
Institut fur Schi bau, University of Hamburg, 1997.
34. Lilek, Z .: Ein Finite-Volumen Verfahren zur Berechnung von inkom-
pressiblen und kompressiblen Stromungen in komplexen Geometrien mit
beweglichen Randern und freien Ober achen, Dissertation, University of
Hamburg, Germany, 1995.

42

You might also like