You are on page 1of 24

10

2 THEORY
2.1 GENERAL THEORY

The drag of a submarine at a given speed is the fluid force acting on the hull in such a way as to
oppose its motion (Harvald, 1991). For a submarine deeply submerged the drag is due primarily to
viscous effects, which can be split into two categories.

Skin Friction: The component of resistance obtained by integrating the tangential stresses
over the wetted surface of the ship in the direction of motion. The skin friction is related to
the surface area of the hull, therefore it is desirable to reduce the surface area for a given
volume (Burcher & Rydill, 1994); and
Form Drag: The pressure resistance is the component of resistance obtained by integrating the
normal stresses over the surface of a body in the direction of motion. Due to viscosity, there
is a reduction of the fluid momentum resulting in a pressure difference between the bow and
the stern and thus a net drag in the direction of motion. The form drag can be minimized by
having very slow varying sections across the length of the submarine resulting in a high
surface to volume ratio (Burcher & Rydill, 1994).

It should be noted that induced drag due to lift not only on the appendages, but also on the hull itself
also exists when the submarine is travelling at an angle of attack to the flow. It can be seen that the
two forms have opposing requirements. The variation of the two kinds of drag and their summation
when plotted as a function of the length to diameter ratio produce a region at which the minimum drag
can be achieved as seen in Figure 2.1. It should also be noted that appendage drag is not considered.


Figure 2.1: Drag Components for Constant Volume (Joubert, 2004)



11

It can be seen that when determining the hydrodynamic characteristics of a submarine the length to
diameter ratio is critical. In 1948 the American designed Albacore set the bench mark when it opted
for optimal submerged performance at the expense of surface operation. The result was a length to
diameter ratio of 7.723; a figure most submarines have since strived for (Joubert, 2004). However,
submarines typically depart from this shape, mainly due to their diameter being set at a fixed number
of decks. That is, the diameter of a submarine is designed to accommodate a set amount of decks at a
set height, thus small changes are not feasible. Additionally larger, appendages relative to the size of
the boat are typically required for boats with lower L/D ratios; adding further to the appendage drag.

2.1.1 DRAG CLOSE TO A FREE SURFACE

Although the modern, conventional powered submarine (SSK) is designed with emphasis on
submerged performance, many occasions call for surface or near surface operation such as entering or
leaving harbour, lengthy transits to a dive area, reconnaissance and snorting (Burcher & Rydill, 1994).
When a submarine is operating close to, or on, the free surface, wave making resistance has an effect
on the total drag on the hull. Wave making resistance is the component of resistance associated with
the energy expended generating gravity waves (Lewis, 1988).

As water passes around a submarine it speeds up and slows down at various points and, according to
Bernoullis equation, the pressure will also change 9i.e. increased velocities result in a drop in
pressure). The forces acting on the submarine are of a considerable magnitude, however in an ideal
fluid they all act to cancel each other out (Rawson & Tupper, 2001).

As a submarine approaches a free surface, these pressure variations can manifest as elevations or
depressions of the water surface, that is, waves. This process results in an imbalance in the
distribution of pressure across the body and a drag force is created. The magnitude of the drag force
is related to the energy of the wave system (Rawson & Tupper, 2001).

There are two wave systems produced by a body moving close to or on the free surface divergent; and
transverse, as seen in Figure 2.2. As a submarines speed changes, so too does the wave pattern.
Therefore there will be a succession of speeds when the crests of the two wave systems reinforce one
another and vice versa, this process is known as interference (Lewis, 1988).



12


Figure 2.2: Kelvin Wave Pattern Patterns of a Point Disturbance


In the case of a ship, the Froude Number (a dimensionless parameter used to relate inertia and
gravitational forces) is typically defined as:

(2.1)

As submarines are relatively small vessels, obtaining any reasonable speed requires them to operate at
high Froude Numbers (F
r
). Wave making resistance becomes the dominate force at these Froude
Numbers, and the form idealised for underwater performance is not suited to these operational
conditions on the surface and results in the submarine operating at a significantly reduce speed
(Burcher & Rydill, 1994). For a given Froude Number, the effect of wavemaking resistance will
reduce as the submarine moves further below the free surface until it is considered negligible, which
typically occurs at half the length of the body (Rawson & Tupper, 2001). It is clear that a surfaced
submarine will exhibit the highest wavemaking resistance for a given Froude Number; Figure 2.3
shows a typical resistance curve of a surfaced submarine, where the large proportion of wavemaking
resistance at higher Froude Numbers can clearly be seen.

Figure 2.3: Typical Surfaced Submarine Resistance Coefficient Curve (Burcher & Rydill, 1994)



13

2.2 COMPUTATIONAL FLUID DYNAMICS
2.2.1 FUNDAMENTALS

CFD is essentially the integration of fluid mechanics, mathematics and computer science. The key
advantage of using CFD over more conventional model testing is being able to ascertain the impact of
design modifications before a large investment in time or money has been made in model testing.

CFD is fundamentally based on the governing equations of fluid dynamics. They represent
mathematical statements of the conservation laws of physics. According to Tu (2008), the three
equations that govern all computations in CFD are:

Continuity Equation: mass is conserved for the fluid:

. 0 (2.2)

Newtons Second Law: the rate of change of momentum equals the sum of forces acting on
the fluid:

X component:

(2.3)
Y component:

(2.4)
Z component:

(2.5)

First Law of Thermodynamics: the rate of change of energy equals the sum of rate of heat
addition to, and the rate of work done on, the fluid:

(2.6)







14

The above five equations known as the Navier Stokes Equations, represent seven unknowns. They
are completed by adding two algebraic equations; one relating density to temperature and pressure
(equation 2.7) and the other relating static enthalpy to temperature and pressure (equation 2.8):

, (2.7)

, (2.8)

2.2.2 NUMERICAL TECHNIQUES
2.2.2.1 Discretization

Discretization is essentially converting the partial differential equations and auxiliary conditions into a
system of discrete algebraic equations (Tu, 2008). The discretization process does however mean that
in all cases the solutions are approximate (MARNET-CFD, 2002).

In principle the finite-difference method can be applied to any type of grid system, however in
practice it is only applied to a structured mesh. The grid spacing does not need to be identical but
there are limits to the amount of distortion that the grid can undergo. It is important to note that the
numerical calculations do not need to be performed in the physical space but, rather they are
preformed in a transformed computational space. There are essentially two forms of these; known as
the forward and backward difference method, in reference to their particular bias in a given direction
(Tu, 2008).

In contrast, the finite-volume method discretizes the integral of the conservation equation directly in
the physical space. The computational domain is divided into finite number contiguous volumes. As
the finite-volume method works with the control volumes and not at grid intersection points, it has the
capacity to accommodate any type of grid. Thus, the finite-volume method is capable of computing
both structured and unstructured meshes. The finite-volume method is implemented in the majority of
calculations today (Tu, 2008).





15

2.2.3 TURBULENCE MODELS

Whilst theoretically the above equations can sufficiently describe all incompressible flows they are
inherently non-linear and as a result are subject to instabilities, which physically grow to form a
mechanism to describe turbulence (MARNET-CFD, 2002). All flows can be described as either
turbulent or laminar, see Figure 2.4. Laminar flow has a smooth velocity gradient. It moves along in
streamlines, or sheets. The velocity at the wall is zero and there is a slow increase to the free stream
velocity. The velocity of each sheet is affected by the shear stress or friction between the molecular
momentums (Gerhart, Gross, & Hochstein, 1992). Most flows in engineering are turbulent;
turbulence can originate from the free stream or be induced by surface roughness. Turbulence is
associated with random fluctuations in the fluid, making computations based on equations that
describe fluid motion exceptionally difficult. As a result it is assumed that turbulence essentially
causes a variance in velocity with respect to time; this variance is expressed as a mean velocity with
a fluctuating component of

. Accordingly, time averaged calculations, commonly known as the


Reynolds-Averaged Navier-Stokes (RANS) equation, are preformed. (Tu, 2008).


Figure 2.4: (a) Laminar Flow; (b) Transition Flow; (c) Turbulent Flow (White, 2007)





16


Figure 2.5: Laminar and Turbulent Flows Within a Pipe (White, 2007)

After a century of intensive theoretical and experimental research, it is accepted that no single
turbulence model can be considered adequate for all flow situations (Tu, 2008). A large number of
models exist today, and the selection of the best one to use for a particular application must not only
take into account questions of accuracy but also computational requirements. It is acknowledged that
more advanced techniques such as Large Eddy Simulation (LES) and Detached Eddy Simulation
(DES) exist however they will not be considered for the purposes of this study due to their excessively
large computational requirements. The turbulence models mentioned, k-, k- and the Shear Stress
Transport (SST), all rely on developing a relationship between the Reynolds Stresses and the mean
velocity.

2.2.3.1 Equation for Kinetic Energy of the Turbulent Fluctuation

As the complete Navier-Stokes system of equations is not closed, one such equation to solve the
Navier-Stokes equations is the kinetic energy equation for turbulent flow. This equation is obtained
through the balance of kinetic energy of the turbulent fluctuations. The kinetic equation of the
turbulent fluctuations is given by Versteeg & Malalasekera (2007) as:


(2.9)


The k equation describes a balance between the four contributions to the energy of turbulent
fluctuations. These energies are convection, diffusion, production and dissipation. Therefore, the
change in turbulent energy due to convection is compensated by an energy source (production), an
energy sink (dissipation) and energy transportation (diffusion). However, the Navier-Stokes equations


17

cannot be closed without further assumptions being made between the Reynolds stresses and
quantities of mean motion (Tu, 2008). In using this kinetic equation as a base, and assuming the
Reynolds stress or the energy fluctuation, the Navier-Stokes equations can be solved.

2.2.3.2 k- Turbulence Model

The two-equation k- model is the most widely used and validated turbulence model (Tu, 2008). k is
as described in section 2.2.3.1 and is the rate at which turbulent energy is dissipated by the action of
viscosity on the smallest eddies (Launder & Spaulding, 1974). The one dimensional k- model is
expressed as follows (Anderson, 1995):


(2.10)

(2.11)
where; i,j = 1,2,3

The k- model offers a good compromise between versatility and accuracy for most general purpose
calculations (ANSYS CFX, 2008). The k- model is typically inadequate in adverse pressure
gradients such as those found in the boundary layer of a surface vessel (MARNET-CFD, 2002).
Other applications where the k- model is not suitable according to ANSYS CFX (2008) are flows:

with boundary layer separation;
with sudden changes in the mean strain rate;
in rotating fluids; and
over curved surfaces.
The k- model is particularly suited for the following situations:
modelling the free stream around a hull form; and
modelling the turbulence aft of the hull form.






18

2.2.3.3 k-

The second most widely used type of two equation model is the k- model, where is the frequency
of the large eddies (Wilcox, 1986). The k- model performs very well close to walls in boundary
layer flows, particularly under strong adverse pressure gradients. The main problem with the Wilcox
model is its strong sensitivity to free stream conditions (Menter, 1994). Depending on the value
specified for at the inlet, a significant variation in the results of the model can be obtained. (ANSYS
CFX, 2008).

2.2.3.4 Shear Stress Transport

The Shear Stress Transport (SST) model, developed by Menter (1994), is one of the most effective
turbulence models commercially available. To overcome the shortcomings of both of the previous
methods Menter proposed a model that combined the accuracy of the k- model near the wall and the
accuracy of the k- in the free stream by the use of blending functions and limiters. Essentially a
hybrid of both models is used; k- at the wall which is then blended into the k- as the flow
approaches the free stream. Blending functions are introduced to achieve a smooth transition between
the two models (Versteeg & Malalasekera, 2007). To improve the flow in the region of adverse
pressure gradients, viscosity limiters are used. Turbulent kinetic energy is limited to prevent the build
up of turbulence in stagnation regions (Versteeg & Malalasekera, 2007). In a NASA Technical
Memorandum, the SST was rated the most accurate model for aerodynamic applications (Sulficker &
Murali, 2006). ANSYS CFX (2008) recommends the SST model for high accuracy boundary layer
simulations. To benefit from this model, a resolution of the boundary layer of more than 10 points is
required. The SST model is significantly more computationally demanding than the k- and k-
turbulence models, as twice as many equations are being solved in every iteration.

2.2.4 NEAR WALL MODELLING

The modelling of the flow near the wall is important as most engineering flows contain solid
boundaries. Due to the presence of the solid boundary, the flow behaviour and turbulence structure
are considerably different from free turbulent flows and these difference must be accounted for when
using CFD (Malalasekera & Versteeg 2007). To understand how to model this correctly an
understanding of the flow associated within this region is paramount.



19

In the thin wall layer, the flow is dominated by viscous stresses and is not yet influenced by the free
stream. Therefore the mean flow velocity depends on the distance y from the wall, the fluid density ,
viscosity , and the wall shear stress
w
(Malalasekera & Versteeg 2007). Through dimensional
analysis it is shown that,

(2.12)
This equation is called the law of the wall and introduces two important non-dimensional terms, u
+

and y
+
, where u
t
is called the frictional velocity and defined as:

(2.13)
Karman in 1933 found that u in the outer layer is independent of viscosity, but must depend on other
properties. He deduced that,

(2.14)
by dimensional anlysis

(2.15)
This equation is called the velocity defect law.

As seen from Figure 2.6, there are 3 regions in the flow near the wall;

1. wall layer thin layer dominated by viscous stresses;
2. overlap layer viscous and turbulent stresses; and
3. outer Layer dominated by turbulent stresses.




20


Figure 2.6: Velocity Distribution Near a Solid Boundary (White, 2007)

The law of the wall and the velocity defect law are used to model the flow in each layer

2.2.4.1 Thin Viscous Sub-layer

The viscous sub-layer is in practice extremley thin (y
+
<5) and it may be assumed that the shear stress
remains approximatelyy constant and equal to the wall shear stress,
w
, throughout the layer (Versteeg
& Malalasekera, 2007); Therfore:

(2.16)

Upon integration with boundary conditions and then making the equations non-dimensional it is seen
that

(2.17)

This shows that there is a linear relation between the velocity u and the distance from the wall y, and
hence is also known as the linear sub layer.






21

2.2.4.2 Overlap Layer

In this region the viscous and turbulent forces are both considered. Versteeg and Malalasekera (200)
explain that the shear stress varies slowly with the distance from the wall. And using the mixing
length equations, it is possible to derive an equation relating y
+
and u
+
.

ln

(2.18)

where: k=0.4
B=5.5

This equation is called the Log Law and for this reason the overlap layer is often called the log law
layer (Versteeg & Malalasekera, 2007).

2.2.4.3 Outer Layer

For larger values of y (y/>0.2) the velocity defect law is applicable. However this equation needs
modification so that at the overlap region the velocity defect law and the log law are equal, Tennekes
and Lumley (1972) show that a matched overlap is obtained by assuming the following logarithmic
form:

(2.19)













22

Figure 2.7 shows a boundary layer profile with the visous or linear sub-layer transitioning into the
overlap region and the outer turbulent region.

Figure 2.7: Boundary Layer Profile (White, 2007)

A major difference that separates the turbulence models discussed in the previous sections is how they
resolve the boundary layer. k- uses a wall function that completely resolves the viscous sub layer of
the boundary layer while the k- and SST turbulence models do not use wall functions and solve the
boundary layer completely. This means that the k- model needs the first node to be on the edge of
the boundary layer while the k- and SST models need to have the first node very close to the wall to
fully resolve the boundary layer. Turbulence models require the use of a non dimensional parameter
y
+
to determine the height of the first node. y
+
is essentially a function of Reynolds number and
allows the CFD user the ability to determine the required height of the first node based on what
turbulence model is being used, see Figure 2.8. The recommended y
+
values for three of the most
common turbulence models employed in commercial CFD codes can be seen in Table 2.1.



23


Figure 2.8: First Layer Height (ANSYS CFX, 2008)
Table 2.1: Required y
+
for Various Turbulence Models (Ranmuthugala, 2008)
Turbulence Model Required y
+

k- 30-100
k- <1-2
SST <1-2


The y
+
value can be found from;

(2.20)

For simplicity, ANSYS CFX (2008) recommends using the following to calculate initial y
+
values:


(2.21)

where,
y = First prism height
L = Reference length
Re = Reynolds number





24

2.2.5 GRIDDING TECHNIQUES

Grid generation presents an important consideration in computing numerical solutions to the
governing partial differential equations of the CFD problem. The quality of the grid can have a
significant influence on the overall accuracy of the solution. As a general rule the accuracy of the
simulation increases with increasing number of cells, i.e. with decreasing cell size. However, with
increased mesh size the computational demands also increase resulting in longer run times
(MARNET-CFD, 2002). Currently, body fitted grids are used almost universally. The above
publication describes the various forms of mesh topology commonly used as follows:

Structured grid
The points of a block are addressed by a triple of indices. The connectivity is straight-forward because
cells adjacent to a given face are identified by the indices. Cell edges form continuous mesh lines
which start and end on opposite block faces. Cells have the shape of hexahedral, but a small number
of prisms, pyramids and tetrahedral with degenerated faces and edges are sometimes accepted.

Block structured grid
For the sake of flexibility the mesh is assembled from a number of structured blocks attached to each
other. Attachments may be regular, i.e. cell faces of adjacent blocks match, or arbitrary (general
attachment without matching cell faces). Figure 2.9 demonstrates a simple multi-block structured grid


Figure 2.9: Block Structured Grid (Wyman, 2001)








25

Chimera grid
Structured mesh blocks are placed freely in the domain to fit the geometrical boundaries and to satisfy
resolution requirements. Blocks may overlap, and instead of attachments at block boundaries
information between different blocks is transferred in the overlapping region. Figure 2.10
demonstrates the typical way a chimera grid is developed.


Figure 2.10: Chimera Grid (Wyman, 2001)

Unstructured grid
Meshes are allowed to be assembled cell by cell freely without considering continuity of mesh lines.
Hence, the connectivity information for each cell face needs to be stored in a table. The most typical
cell shape is the tetrahedron, but any other form including hexahedral cells is possible. Figure
2.11demonstrates a typical unstructured grid.

Figure 2.11: Unstructured Grid (Wyman, 2001)







26

Hybrid grid
This grid combines structured with unstructured meshes. In Figure 2.12 a hybrid mesh can be seen.


Figure 2.12: Hybrid Grid -Structured Left, Unstructured Right (Wyman, 2001)

2.2.5.1 Grid Topology

Typically in structured multi block meshes there are several topologies adopted. A topology is
essentially a way of arranging the blocks to ensure optimum grid characteristics. Each topology has its
own advantages and the choice of which to use is very much dependent on the situation. It is indeed
not uncommon to employ multiple topologies in the one grid.

A H type mesh, as seen in Figure 2.13, is the standard meshing method used in ANSY-ICEM CFD. A
H mesh can achieve good results for a simple geometry, however to maintain accuracy for complex
shapes the blocking becomes complex.


Figure 2.13: 2D H Grid around a Cylinder (Widjaja 2009)




Unstructured
Grid
Structured
Grid


27

An O type mesh, as seen in Figure 2.14 is ideally suited for circular or curved surfaces; Figure 2.15
shows that when an H mesh is used on a circular geometry highly skewed elements exist at angles of
45 around the geometry; an O type mesh removes this skewness. O type meshing is not well suited
to wake flows. Figure 2.14 shows that as the O expands to outer edges of the geometry the elements
become quite large, and would not accurately capture the wake region of the flow.


Figure 2.14: 2D O Grid around a Cylinder (Widjaja, 2009)


Figure 2.15: O-Grid inside a Cylinder (Widjaja, 2009)

A C mesh, as seen in Figure 2.16, is a combination of an H and C grid, it has the benefit of the O grid
where it accurately models a curved surface, but also allows for refinement of the mesh in the leeward
edge of the geometry. C type meshing is ideally suited for flows where a wake needs to be captured
and anything that has a bluff leading edge and small finite to infinite trailing edge such as foils and
wings as the mesh reduces to H mesh at these sections allowing for mesh edges to fully capture the
geometry of these critical regions.





28


Figure 2.16: 2D C Grid around a Cylinder (Widjaja, 2009)

2.2.6 NUMERICAL SOLUTION OF A FREE SURFACE

A free surface is an interface between a gas and a liquid where the difference in densities is
significant. The gass inertia is ignored; its only effect is to apply a static pressure on the liquid
surface. Two commonly adopted approaches for computing a free surface in CFD are Eularian-
Lagrangian and Eulerian-Eulerian methods, however Eulerian-Elurian methods allow for more
complex free surfaces to be tracked. (Senocak & Iaccarino, 2001)

When simulating a flow using the RANS equations there are essentially two methods that can be
adopted; interface tracking and interface capturing. Interface tracking involves modelling only the
fluid domain. One of the domain boundaries is then assigned to be the free surface and the grid is
adapted to the position of the free surface. This method is not suitable for conditions of steep or
breaking waves as the entrapment of air cannot be accounted for. Interface capturing involves solving
the governing equations for both air and water, which makes this approach typically more versatile
(Senocak & Iaccarino, 2001).

2.2.6.1 Volume-of-Fluid

If the amount of fluid in each cell is known then the surface can be tracked. The Volume-of-Fluid
method (VOF), first introduced by Hirt and Nichols (1981), involves tracking the volume fraction of
each cell. Initially each cell is prescribed a volume faction of either one or zero. The free surface is
then defined as an area of rapid change in the volume fraction. The volume fraction is solved for one
of the phases by means of an extra transport equation, having generally the same mass. One common
algorithm used for this solution is the semi-implicit method for pressure linked equations (SIMPLE).
Interface sharpening algorithms can be used to refine the cells with a value of between one and zero.



29

The main disadvantage of VOF, and indeed all surface capturing techniques, is that the mesh is
required to be more refined at the free surface interface than surface tracking methods, resulting in a
larger grid (Senocak & Iaccarino, 2001). However, VOF remains the most viable option for accurate
three dimensional free surface calculations.

2.2.6.2 Level Set

The level set technique involves enforcing a step-wise variation of the fluid properties. Initially it is
set equal distance from the free surface, positive in one direction and negative in the other. At every
later instance, the function is computed from the condition that its total (material) derivative with
respect to time is zero (MARNET-CFD, 2002); thereby ensuring that the function is constant with
time on all particles. The main issue with the level set technique is the need to arbitrarily induce some
finite thickness across the interface region to promote smooth but rapid change in properties (Sethian,
1996).

2.2.7 SOLUTION PROCEDURE

The most common solution procedure adopted in ANSYS CFX when solving free surface flows is the
SIMPLE algorithm. Essentially this is a pressure correction equation. The equation is obtained from
the discretized mass and momentum conservation equations. The momentum equations are solved
first for an estimate of velocity components. These are then corrected upon solving the pressure-
correction equation. The equation for the volume fraction is then solved, followed by the equations
for turbulent kinetic energy and its dissipation rate. The eddy viscosity is then computed and the
whole process is repeated until all non-linear equations are satisfied. The time is then advanced to the
next time step (Azcueta et al, 1999).










30

2.2.8 TIME STEPS

According to ANSYS 2008 simulations of a physical process can either be transient or steady-state. In
a transient simulation, the behaviour of a physical system as a function of time is investigated. In a
steady-state simulation, it is assumed that the physical system is moving towards an equilibrium or
steady-state solution

Time stepping refers to solving the governing equations for a different value of time (t). A time step
is the change in time per iteration. For steady-state problems, the ANSYS CFX-Solver applies a false
time step as a means of under relaxing the equations as they iterate towards the final solution (ANSYS
CFX, 2008). There are two ways that a time dependent flow can be solved, either implicitly or
explicitly. Essentially explicit methods compute the next time step solution directly based on the
known current time information, whilst implicit methods evaluate the solution at two or more time
spaces, one of which is the current time step. For stability in an explicit equation the time step is
governed by:

(2.22)

This is a stringent rule and if not followed the solution will become unstable and result in divergence.
This represents a serious limitation for the explicit scheme as it becomes computationally expensive to
increase spatial accuracy (Versteeg & Malalasekera, 2007).

In theory the fully implicit scheme will always be stable, i.e. it will not accumulate errors or become
un-bounded, or diverge, (which is why most commercial codes employ this method). However, the
majority of CFD calculations are preformed in single precision or first order to avoid excessive
computational demands (Tu, 2008). As a result, if too large a time step is taken the round-off errors
will increase leading to an inaccurate answer. It is also possible to use a time step that is too small
leading to an increased discretization error. Versteeg and Malalasekera (2007) recommend the
implicit method due to its robustness and unconditional stability. ANSYS CFX (2008) notes that free
surface and multiphase flows typically require very small time scales to ensure an accurate solution.
When simulating a transient or time dependent run, the Courant-Friedrichs-Lewy (CFL) condition is
typically used as a form of monitoring the stability of the equations. For convection-type simulations
the Courant number is written as (Tu, 2008);




31


(2.23)

It is seen that the parameter relates the spatial and time steps to the velocity. For an accurate solution
between each time step the Courant number is required to be below one, however as in the explicit
method this can result in very small time scale for refined grids. As a result it is common to use a
larger Courant number and take time averaged results to account for the fluctuations.

3 APPROACH
3.1 MODEL DESIGN

In order to decide on a suitable geometry for the project, an extensive study of many of the worlds
current operational submarines was conducted. It should be noted that no Ballistic Submarines
(SSBNs) were considered. Of particular interest was the length to diameter ratio of vessels, as it has
been highlighted by Vine (1991) that notable discrepancies between nuclear powered submarines
(SSNs) and SSKs exist. As can be seen in Figure 3.1 and Figure 3.2, no discernable difference
between SSKs and SSNs can be noted in regard to the L/D ratio and, as a result, the DARPA
SUBOFF geometry was adopted as there is a considerable number of studies, both experimental and
numerical, that have been undertaken on this particular hull form. These include the experimental
studies of Roddy (1990) and Huang et al (1992) and the computer modelling studies of Bull (1996),
Toxopeus (2008) and Widjaja et al (2007).


Figure 3.1: Comparison of the Length to Diameter Ratio
of Existing Submarines (Hazegray, 2009)
4.00
6.00
8.00
10.00
12.00
14.00
45 50 55 60 65 70 75
L
/
D
L/V
1/3
L/D as a function of L/V
1/3
SSN
SSK
Collins
Class





Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
Huang
of the free surface.
when using a model of 1.25m in length and 0.25m diameter, a deep water run
surface effects
original model was adopted
form is given in

Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
Huang
of the free surface.
when using a model of 1.25m in length and 0.25m diameter, a deep water run
surface effects
original model was adopted
form is given in
L
/
D
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
Huang et al
of the free surface.
when using a model of 1.25m in length and 0.25m diameter, a deep water run
surface effects
original model was adopted
form is given in
0.00
2.00
4.00
6.00
8.00
10.00
12.00
14.00
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
et al
of the free surface.
when using a model of 1.25m in length and 0.25m diameter, a deep water run
surface effects
original model was adopted
form is given in
0.00
2.00
4.00
6.00
8.00
10.00
12.00
14.00
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
(1992)
of the free surface.
when using a model of 1.25m in length and 0.25m diameter, a deep water run
surface effects,
original model was adopted
form is given in
0.00
2.00
4.00
6.00
8.00
10.00
12.00
14.00
20
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
(1992)
of the free surface.
when using a model of 1.25m in length and 0.25m diameter, a deep water run
was not attainable
original model was adopted
form is given in Figure
20
Figure
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
(1992) - was
of the free surface.
when using a model of 1.25m in length and 0.25m diameter, a deep water run
was not attainable
original model was adopted
Figure
Figure
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
was
of the free surface. Vine
when using a model of 1.25m in length and 0.25m diameter, a deep water run
was not attainable
original model was adopted
Figure 3
Figure
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
was considered too l
Vine
when using a model of 1.25m in length and 0.25m diameter, a deep water run
was not attainable
original model was adopted
3.3 and
Figure 3.
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
considered too l
(1991)
when using a model of 1.25m in length and 0.25m diameter, a deep water run
was not attainable
original model was adopted for both the EFD and CFD
and
40
L/D as a Function of Length
.2: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
considered too l
1991)
when using a model of 1.25m in length and 0.25m diameter, a deep water run
was not attainable
for both the EFD and CFD
and the principal particulars are given in
40
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
considered too l
1991) demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
was not attainable at Froude Numbers upwards of
for both the EFD and CFD
the principal particulars are given in
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
considered too l
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
for both the EFD and CFD
the principal particulars are given in
Figure
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
considered too large to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
for both the EFD and CFD
the principal particulars are given in
Figure
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
for both the EFD and CFD
the principal particulars are given in
Figure 3
60
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
for both the EFD and CFD
the principal particulars are given in
3.3: SUBOFF
Length (m)
L/D as a Function of Length
32
: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
for both the EFD and CFD
the principal particulars are given in
: SUBOFF
Length (m)
L/D as a Function of Length
32
: Comparison of the Length to Diameter Ratio
of Existing Submarines
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
for both the EFD and CFD
the principal particulars are given in
: SUBOFF
Length (m)
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
of Existing Submarines (Hazegray, 2009)
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
for both the EFD and CFD work. A diagram of the
the principal particulars are given in
: SUBOFF
80
Length (m)
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
(Hazegray, 2009)
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
work. A diagram of the
the principal particulars are given in
: SUBOFF (mm)
80
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
(Hazegray, 2009)
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of
work. A diagram of the
the principal particulars are given in
(mm)
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
(Hazegray, 2009)
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
at Froude Numbers upwards of 0.6
work. A diagram of the
the principal particulars are given in
(mm)
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
(Hazegray, 2009)
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
0.6.
work. A diagram of the
the principal particulars are given in Table
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
(Hazegray, 2009)
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
. As a result
work. A diagram of the
Table
100
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio

Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
As a result
work. A diagram of the
Table 3.
100
L/D as a Function of Length
: Comparison of the Length to Diameter Ratio
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run
As a result
work. A diagram of the
.1.
: Comparison of the Length to Diameter Ratio
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
when using a model of 1.25m in length and 0.25m diameter, a deep water run,
As a result,
work. A diagram of the


Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
or run with no free
, a 2.8:1 scale of the
work. A diagram of the axi
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
or run with no free
a 2.8:1 scale of the
axi-symmetric
120
Due to the available depth of the towing tank (1.5m), the standard size model of 4.356m-
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
or run with no free
a 2.8:1 scale of the
symmetric
120
as used by
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
or run with no free
a 2.8:1 scale of the
symmetric

SSK
SSN
Collins
as used by
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
or run with no free
a 2.8:1 scale of the
symmetric
SSK
SSN
Collins
as used by
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
or run with no free
a 2.8:1 scale of the
symmetric hull
Collins

as used by
arge to facilitate the measurement of drag without the effects
demonstrated in experiments conducted in the same facility that,
or run with no free
a 2.8:1 scale of the
hull


33

Table 3.1: Parameters of Scaled SUBOFF Model
Parameter Value
Length 1.55575 m
Diameter 0.181 m
Surface Area 0.3819 m
2

The depths at which the model was tested both experimentally and computationally were expressed
non-dimensionally as D*, where;

(3.1)

3.2 TESTING RANGE

The depths tested both experimentally and computationally were; 1.1D
*
, 1.3 D
*
, 2.2 D
*
, 3.3 D
*
, 4.4
D
*
, 5.5 D
*
(see Figure 3.4). In order to ensure the current project would be relevant to future
submarine design, a full scale length corresponding to Australias existing Collins Class Submarines
(77.5m) was assumed. The speeds tested ranged from 0.5 to 2.5m/s corresponding to a Froude
Number range of 0.128 to 0.640 or a full scale speed range of 6.8 to 34knots.

Figure 3.4: Test Depths

You might also like