You are on page 1of 36

21 227

MARINE

HYDRODYNAMICS

©
DR P G SAYER
CONTENTS

Introduction 1

Approach to mathematical modelling 3

Continuity equation for incompressible flow 4

Fundamental modelling tools - stream function 8

Velocity potential and irrotational flow 10

Circulation and vorticity 12

Sources and sinks 14

Doublets/dipoles and geometrical transformations 17


P G Sayer
Vortices and propeller flow 19

Lift 22

Boundary layers
NAME NAME 24

Boundary layer velocity profiles 26


P G Sayer
Displacement thickness, momentum thickness and skin friction 26

NAME
Boundary layer thickness
NAME 28

Skin friction resistance in naval architecture 29

Bibliography 31

Tutorial exercises 32
MARINE HYDRODYNAMICS

Introduction

This is a course involving applied mathematics!!....... Mathematics applied to


marine hydrodynamics for naval architects. By marine hydrodynamics I mean
the dynamic properties and characteristics of water, either on the large scale
where we might wish to examine the flow of water along a ship hull and past the
propeller, or on a smaller scale if we are looking at the flow of water (or some
other fluid) through pipework or a piece of machinery. So there will be
relevance to both naval architects and marine engineers.

Naturally you may ask whether there is a need for yet another mathematics
course. After all, surely naval architects and marine engineers simply need
‘tools to do the job’. The job in this case might be the prediction of how a ship
performs in a given condition, how safe it is, how much power is needed, and
similar factors. In a sense, maybe we just need to consider safety and economics.
Whether you accept this view or not, some means is needed for quantifying the
relevant criteria and data.

Think for a moment about what most engineers take for granted when using
Newton’s laws of motion, or some similar equations or techniques fundamental
to mechanics. Inherent in these equations are the principles of vectors (force and
momentum), calculus (differentiation for acceleration and integration for
positions) and all the underlying mathematical properties. Although we might
perfectly happily treat the solutions of these equations via a ‘black box’ , e.g.
some software package, what happens if something goes wrong? - Invalid data,
lack of convergence of some algorithm, etc. We need to be able to identify not
only when something goes wrong, but also when something is likely to go
wrong. This can only be achieved if we understand the fundamental
mechanisms which we are using to try to solve the particular problem. At an
earlier stage still, the greatest difficulty facing many people is how to formulate
the physical problem in mathematical terms.

In the case of ships or other marine vehicles, the laws of motion are complicated
by the details of the excitation forces - the forces arising from the presence of
the water (the hydrostatic and hydrodynamic effects) are complicated to
compute because of the double difficulty of determining the pressure due to the
waves and currents and the effect of the water’s viscosity. Sometimes we can
safely neglect the viscosity; sometimes we cannot, e.g. in the case of ship
resistance. An engineer needs to know when various factors must be included
and when they can be neglected; otherwise the governing equations become

© 2007 P.G. Sayer


Page 1
unmanageable in practice. It may be difficult to compute the solution and also,
importantly, to enable the naval architect to ‘get a feel’ for what is happening
physically in the real situation.

This course therefore introduces the fundamental characteristics of the flow of


water and how forces are then exerted on a marine vehicle, by expressing the
relevant physical properties in appropriate mathematical terms. Only then will it
be possible to build on this knowledge and understanding to model ship motions.
At each stage, we shall try to identify and explain any approximations and
simplifications which can be made, or indeed which might be necessary.

It is important to make clear at an early stage that engineers and scientists know
the general equations that govern the flow of water: the so-called Navier-Stokes
equations developed in the nineteenth century tell us about the way a ‘lump’ of
fluid (water) behaves when it is subjected to pressure forces (e.g. due to waves
or current on a ship hull; or the action of a propeller or pump). The main
practical difficulty, however, is that it has only become possible recently, with
the advent of extremely powerful computers, to solve the equations with
acceptable accuracy for many realistic situations in naval architecture and
marine engineering. But such computational facilities are very expensive, not
widely available and, importantly, we often do not need the level of detail or
accuracy produced by this sophistication; it is a luxury that we do not need!
Instead, we wish to gain more insight into the importance of various parameters
and their effects (trends).

Lots of factors complicate the ‘real’ world, and often we have to decide what
assumptions and simplifications are needed, and whether these are acceptable in
a particular context. For example, most of us would perhaps accept that water is
a non-viscous, incompressible fluid; in many situations this would be perfectly
acceptable.......but not if we wish to study the resistance of a ship hull, for which
a large contributing factor is the skin friction; this depends on the thin region of
water adjacent to the hull where viscosity is important, and which gives rise to a
boundary layer where our fluid ‘lump’ is subjected to large shear forces caused
by rapid changes in flow velocity across the boundary layer. Fortunately, the
incompressibility of water is a very accurate assumption except in the very
deepest parts of the ocean, where at a depth of 10,000m the ambient pressure
will be around 1,000 atmospheres (108 Pa). [Interestingly, in aerodynamics, the
compressibility of air usually only has to be allowed for at speeds greater than
Mach 0.7].

Many of the fundamental concepts and properties of fluid flow are common to
all branches of engineering and science. Bernoulli’s equation, for example, has
very widespread use. But it has several forms, depending on the nature of the

© 2007 P.G. Sayer


Page 2
pressure forces giving rise to the flow of the fluid. We shall always attempt to
emphasise how such principles relate to the marine context, and in doing so give
equations and results in forms which tend to be most useful for naval architects.

After these fundamentals of fluid flow, we proceed to introduce the theory of


water waves. It turns out that the presence of a free surface (the so-called air/sea
interface) is the source of many theoretical complications! To make a start,
however, we shall adopt a widely-used first approximation and consider linear
theory, in which the actual free surface is effectively replaced by the mean
undisturbed water level, in other words by a horizontal surface. Nevertheless,
this simplification turns out not only to be important in a systematic
development of the theory but also leads to (surprisingly!) accurate results for
many real-life situations. Only in the most extreme seas, and/or where large-
amplitude motions of a ship may be expected, does the linear theory break
down...but that, too, is beyond the scope of an introductory set of notes such as
these.

Approach to mathematical modelling

We have already mentioned that water can be safely regarded as being


incompressible. Now, how can this property be expressed mathematically?
Quite often, this is the time when most students start to become alarmed!! As a
general rule, we should try to clear in our minds the physical properties of the
flow; in other words, either what is happening to our lump of fluid as it moves,
or how the fluid moves with respect to a fixed point (or region) upon which we
focus our attention. [The former gives rise to what is known as the Lagrangean
description and Lagrange’s equations, while the latter leads to the Eulerian
description and Euler’s equations, in honour of these two eminent scientists].
The streamlines are the paths traced by the fluid particles; in other words, the
fluid velocity vectors are aligned with the streamlines

If we look at a region of incompressible fluid, then obviously however much


fluid flows into this region must be matched by an equal amount flowing out.
Otherwise, there would be a net build-up of fluid, which would contradict our
assumption of the physics, namely the incompressibility criterion. To fix ideas,
much of this course will concentrate on flow in two dimensions although an
extension to the three-dimensional case is normally straightforward. [I say
‘normally’ because, although the governing equations may become quite
complicated, and correspondingly more difficult to solve, the underlying
physics, and hopefully our understanding of the physics, is not affected. In
three dimensions, it is often easier if we re-formulate the equations in terms of
vectors and matrices]. We will need to set up our equations and formulae

© 2007 P.G. Sayer


Page 3
according to two sets of co-ordinates: Cartesian co-ordinates for the most part,
but sometimes, especially for those flow descriptions which inherently involve
circular-type motion (flow near a propeller, for example), it will be more
convenient to adopt polar co-ordinates; once again, we choose the ‘tools’ (in this
case, the co-ordinate system) best suited to the job. Of course, co-ordinates
other than Cartesians and polars could be used.

Cartesian co-ordinates, however, are probably the best to start with. We shall
use x and y, although in order to conform with the symbols and conventions
widely used by naval architects, the two-dimensional study of water waves tends
to use x and z (for horizontal and vertical directions respectively). You must not
be put off by this!! Concentrate on understanding the approach to the modelling
process; this should always take priority over the detailed mathematics. Forgive
the digression, but this is an important point ... in a practical (work) situation, it
will usually be possible to call on the services of someone with the appropriate
expertise (a mathematician, for example) to solve some equations, but often a far
greater and more important challenge is to determine which approach, model or
set of equations is needed to formulate the real physical problem in
mathematical terms.

Continuity equation for incompressible flow

Returning to our incompressible flow, we shall use u and v to denote the


velocity components of the flow in the x and y directions, respectively. In
mathematical terminology, we are saying that u is a function of x and y, namely
u=u(x,y). Likewise for v. In Cartesian co-ordinates, it is easiest if we consider
‘changes’ only along the co-ordinate directions, which in turn suggests that we
should consider a rectangular region of fluid, to which we need to impose the
equality of inflow and outflow from the physical statement of incompressibility.

Note that we are at liberty to choose an arbitrary region of fluid to which we


apply physical principles; the laws of physics must apply to all shapes, and so it
is sensible to choose something that helps keep the mathematics as simple as
possible.

In standard mathematical style, the x-component of velocity at a horizontal co-


ordinate x+δx will be denoted by u+δu; similarly the y-component at a vertical
location y+δy will be written as v+δv.

© 2007 P.G. Sayer


Page 4
The inflow across the sides of the rectangular fluid region, of length δx and δy,
is easily seen to be u δy + v δx. The outflow is (u + δu) δy + (v + δv) δx. Now
"u "v
from calculus, !u = !x , !v = !y , which leads to the result
"x "y

!u !v
"x "y + "y "x = 0
!x !y

The area of the rectangle is δx δy and this is non-zero. We conclude that

!u !v
+ =0
!x !y

This is referred to as the continuity equation. In practical terms, we shall usually


employ Cartesian co-ordinates, certainly if we wish to consider the flow along a
ship’s hull. Although you may correctly observe that such a flow is evidently
three-dimensional in nature, we could easily extend our two-dimensional model
to successive water-lines, and thereby account for the vertical (third) dimension.

However, when looking at flow through or near a propeller, or near a bilge keel
as a ship rolls, the resulting vorticity (see later) shows itself as a definite rotation
of flow, for which polar co-ordinates are sometimes easier to work with. For r
and θ, therefore, where x = r cos θ and y = r sin θ, the corresponding equations
for the velocity components ur and uθ are:

ur =ucos! +vsin! , u! = " usin! +vcos!

© 2007 P.G. Sayer


Page 5
By similar reasoning as used for Cartesian co-ordinates, the incompressibility
condition leads to the equation

! !u
!r
(rur ) + " = 0
!"

The speed (i.e. the magnitude of the velocity) is sometimes denoted by q;


therefore q2 = u2 + v2 = ur2 + uθ2. The local angle that the streamline makes with
the x-axis is tan-1(v/u) = θ + tan-1(uθ/ur).

So, we now know a mathematical constraint that the velocity components must
satisfy for an incompressible fluid. We progress by considering what causes the
fluid to move in the first place. Our lump of fluid will move only as a result of
some force applied to it, either externally or internally. For the present, we
consider the simple case of a fluid moving purely by the difference in pressure
from one point in the fluid to another. We can apply Newton’s second law
directly: let us consider the case of motion along the x-axis. Denote by p the

© 2007 P.G. Sayer


Page 6
pressure acting on the side of the rectangle having horizontal co-ordinate x;
following the same terminology as above, p+δp denotes the pressure at x+δx.
Then, since pressure acts equally in all directions (it is a scalar quantity), the
fluid lump experiences a (vector) force +pδy from left to right over the left-hand
edge [since force equals (pressure x area), which in two-dimensions is
effectively pressure x length], together with a force -(p+δp)δy over the right-
hand edge. Now, pressure is a function of both x and y; but since the only
difference in co-ordinates between the left-hand and right-hand edges of our
fluid region occurs in the horizontal (i.e. x) direction, we make use of calculus to
!p
express δp as δx. Since the acceleration of the fluid in the horizontal
!x
• !u
direction (positive in the sense of positive x, i.e. left to right) is u or then by
!t
equating the net force in the positive x-direction to the (mass x acceleration) of
the fluid region we have:

#u
p( x )!y $ p( x +!x )!y = ( "!x!y )
#t

This equation can be integrated to give one form (the simplest!) of Bernoulli’s
equation for fluid flow in the absence of external forces such as gravity
(equivalently, flow in a horizontal plane). The integration is best done by noting
that the acceleration can be re-written as u ∂u/∂x, so that we can integrate both
sides with respect to x. The above equation becomes

-∂p/∂x = ρ u ∂u/∂x

© 2007 P.G. Sayer


Page 7
which integrates to give the Energy Equation, or Bernoulli’s Equation (more
precisely, the simplest form of Bernoulli’s Equation):

1 2
p+ !u = constant
2

The constant is the same for all fluid particles following the same streamline,
and so we can generalise the above to give

1 2
p+ !q = constant
2

Fundamental modelling tools - stream function

Let us pause and consider the information that we now have: the velocity
components u and v satisfy the continuity equation; and the pressure and the two
velocity components satisfy Bernoulli’s equation; so we have two equations for
three unknowns, which does not enable us to solve for p, u and v separately. At
this stage it is appropriate to introduce two new concepts: stream function
(which is a physical measure of fluid flow, somewhat like a meter which
measures the flow-rate of the water) and velocity potential (which is a rather
abstract mathematical quantity, but has some direct similarities with potential
energy in mechanics). Both turn out to be very important in the sense that they
provide fundamental information about the flow, from which the physically
meaningful quantities such as pressure and velocity can be obtained (and, of
course, once the pressure is known, it can be integrated/summed around the
surface of a body to get the total force acting on the body).

Which of these two functions we use again depends on which one is better suited
to the constraints of the real situation we are trying to model. For example,
although stream function relates more closely with the physics of the problem,
and so may well be the first choice for most people, it can only be defined in two
dimensions. This causes no difficulties just now (nor indeed in the two-
dimensional study of water waves later) but when we wish to extend to a full
three-dimensional model then other ways have to be sought. On the other hand,
velocity potential can be defined for both two and three dimensions, but (and
this is a crucial point) it is defined only for so-called potential flow (hence its
name!) which means that there is no rotation of the flow. Rotation, and the
related circulation will be precisely defined later; for now, it is sufficient to
know that these terms relate to the ‘twisting’ or ‘spinning’ of the flow particles.
Returning to the physics of what is happening, such rotation stems from the
action of viscosity within the fluid...so, although boundary layers in water are

© 2007 P.G. Sayer


Page 8
thin, they are nonetheless responsible for generating circulation within the flow,
which in turn modifies the pressure and velocity distribution compared to what
would exist in the absence of boundary layers.

If all this seems too abstract and perhaps unnecessary, let us not lose track of the
purpose being modelling the hydrodynamic flow characteristics. We ultimately
wish to calculate the forces acting on marine vehicles and structures. In a
laboratory experiment, we could mount the model on a sensitive balance and
measure the forces or moments directly. Alternatively, particularly if we also
wish to learn more of the characteristics of the flow as it passes the model, we
could measure the fluid pressure at a number of points around the surface of the
model. In practice, either at full-scale or in a laboratory experiment, we could
measure fluid pressure using a Pitot tube (or pressure transducer) and the fluid
velocity using a current meter for water or a hot-wire anemometer for air speed
in a wind tunnel (or, in a more sophisticated way, by means of laser
anemometry). So, for both theoretical model and laboratory experiment, we
need to measure flow pressure and velocity. Regard this new concept of stream
function, therefore, as a device to measure flow velocity in our mathematical
model.

The S.I. units of the stream function are m2s-1 (equivalent to volume flow per
metre width of fluid). Consideration of the flow in the diagram below enables
us to define the velocity components u and v in terms of x- or y-derivatives of
the stream function. Based purely on a dimensional argument, we can obtain the
units of velocity (ms-1) by dividing those of the stream function by length; which
is equivalent to a spatial derivative with respect to x or y. The symbol ψ is
usually used to denote stream function; whether it is positive or negative
depends solely on how we define the positive direction of flow; so the definition
is arbitrary but once we have chosen a particular sign convention then we must
stick to it! (There is no standard convention in the literature and there is no need
to worry about this ambiguity in sign). If we can determine ψ then we can
calculate u and v by differentiation, then p by Bernoulli, and so the
hydrodynamic force.

If we consider the flux of flow δψ between two neighbouring stramlines, then

© 2007 P.G. Sayer


Page 9
continuity (i.e. equating inflow and outflow) gives

!" = u!y # v!x

Noting from differential calculus that

"# "#
!# = !x + !y
"x "y

we can equate the coefficients of the δx and δy terms to obtain

"# "#
u= and v=!
"y "x

By considering the flow in a ‘wedge’ of fluid, comprising arcs of a circle, we


can derive the following corresponding results in polar co-ordinates:

!" = rur !# % u# !r

1 $" $"
ur = and u# = %
r $# $r

As a simple example of a stream function, we consider ψ for a uniform stream


flowing at speed U in the negative x-direction. So u = -U and so trivially it
follows that ψ = -Uy. Why is there no integration constant? Think about this;
we shall return to it in a moment.

Velocity potential and irrotational flow

As mentioned above, the concept of velocity potential is somewhat more


difficult to grasp. Its applicability to three dimensions, however, together with
the fact that there are many situations where we can neglect viscosity, means
that it does play a vital role in hydrodynamic modelling. It is usually denoted by
the symbol φ (and once again whether it is defined as a positive or negative
quantity is quite arbitrary, though having selected a sign convention fixes
everything thereafter). Its units are those of (velocity x distance) and so it is
clear that its spatial derivative will generate a velocity.

An important point to note is that the definition and value of φ at any particular
point in the flow is completely independent of the path taken by the flow in
reaching that point. Path-independence turns out to be a key property in
defining the potential energy of a body subjected to the force of gravity (where

© 2007 P.G. Sayer


Page 10
we implicitly neglect other forces, such as air resistance, in comparison). The
only mathematical difference between potential energy and velocity potential is
that potential energy has the units of (force x distance) whereas velocity
potential has units of (velocity x distance). [Note that for ‘distance’ we could
equally well write ‘length’]. The relationships between u and v and φ are given
in the following diagram.

The change δφ in the velocity potential is seen to be u δx + v δy. We then adopt


the same reasoning as for the stream function by using calculus to express δφ as
(∂φ/∂x) δx + (∂φ/∂y) δy and then equating coefficients:

!" !"
u= and v =
!x !y

Using the same technique but without going through the details, the
corresponding formulae in polar co-ordinates are:

!" 1 !"
ur = and u# =
!r r !#

Before we continue to see how ψ and φ are used in practice, we need to


remember that a body in general motion (and this includes the motion of a fluid
particle) can be split into translational components as well as rotational
components - you can perhaps relate these to a ball hit through the air during its
flight, the path followed by its centre of gravity represents the translational
motion, whilst any spin about its centre of gravity corresponds to rotation, in
every sense of the word! In the absence of external forces, we have already
noted that only the fluid viscosity can generate shear forces, which in turn cause
the fluid to spin. So for a non-viscous, or inviscid, or frictionless flow, there is
no rotation and so a velocity potential can be defined. In this case we use the
term irrotational flow. How we actually calculate φ may be somewhat

© 2007 P.G. Sayer


Page 11
complicated!

Nevertheless, many textbooks are available which give φ or ψ for particular


situations. We shall see that wherever both φ and ψ are valid, then one can be
easily derived from the other. So, if both exist, it makes sense to know which is
the easier to derive and work with. Naturally some worked examples will be
given in class but, more importantly in some ways, I want you to remember how
we can derive φ or ψ for simple case and then build up more complicated fluid
flow models/pictures by using the simpler examples in a kind of building-block
approach. No-one should expect to be able to spot the form of φ or ψ for
complicated flows; moreover, it is not really helpful just to be able to memorise
a formula from a book. The real advances are possible only when the student
understands how to tackle the formulation of a problem.

Circulation and vorticity

If a velocity potential can be defined (we sometimes say ‘exists’) then the
characteristics of the flow are path-independent. Therefore, if we consider two
points A and B in the fluid, then the value of φ at B is not only independent of
the path taken by the flow in moving from A, but also φ will return to the same
value at A no matter how we travel from B back to A. [This agrees with the
definition of potential energy: in raising a body between two levels, the change
in potential energy is independent of the path taken between the two levels;
furthermore, there is net zero change in potential energy if the body returns to its
original level, regardless of the path taken to get there].

But, for a general type of fluid motion, there may be a difference in the velocity
and pressure which do depend on the paths taken. We therefore define the
circulation Γ to be the integral of velocity around any closed path (there is
another criterion which is important mathematically, namely that the path must
not enclose a flow singularity, i.e. a singular point in the mathematical model of
the flow).

© 2007 P.G. Sayer


Page 12
If we divide the expression for the circulation by the area enclosed by the path
taken, the δ terms disappear and we are left with what is termed the vorticity ω
of the flow. [In lay terms, most of you will be aware of flow down a bath plug-
hole, or the flow in a tornado, as being like a vortex, or as having vorticity; so
this vorticity measure the spin of the flow and clearly will be important when we
try to model the flow past a propeller, impeller pump or similar device].
"v "u
!= #
"x "y

So, if the flow is irrotational, then the vorticity is zero. You may be tempted to
assume, therefore, that if the vorticity is zero then the flow must also be
irrotational. This turns out to be true in three dimensions, but not for the two-
dimensional problem. The explanation lies in the mathematical result, known as
Stokes’s theorem, which relates the strength of the vorticity (or vortex ‘tubes’)
to the cross-sectional area of the region within the closed path used to define
circulation. Path-independence turns out to require that the irrotationality result
must hold for any closed path, regardless of its size or shape; in particular, as the
length of the path and the area within it shrink to zero. This can be done for all
paths in 3-d, but not in 2-d (e.g. consider a path surrounding a 2-d circular

© 2007 P.G. Sayer


Page 13
cylinder). In practice, though, we are attempting to model a fully 3-d situation,
so the 2-d ‘special case’ tends to be mainly of theoretical interest.

Do not, however, incorrectly suppose that movement of fluid along a circular-


type path must imply vorticity is present (regardless of whether we are looking
at a 2-d or 3-d model). Vorticity is a local property of the motion of a lump of
fluid and measures how the lump of fluid spins about its centre of gravity. The
centre of gravity of the fluid can still follow a circular path while not spinning -
this is irrotational motion (and so can be described by a velocity potential).

A practical example of this will be illustrated later by the flow past (or through)
a propeller. However, before we demonstrate how to bring together the relevant
mathematical ideas and concepts, it is instructive, and also allows us to follow a
more systematic route, if we introduce some other modelling tools.

Sources and sinks

A source is a mathematical device which causes flow (again mathematically) to


be generated at a singularity (where the source is located) and projected
outwards uniformly in the radial directions. A sink has simply the opposite
effect; it is a negative source. The fact that such singularities do not exist in
reality is not important. Provided the mathematical model generates the correct
physical behaviour of the flow, in the region of the flow where it is needed, then
we need not worry about these mathematically singular points. We have seen
that a singularity exists at the centre of our mathematical model of a vortex; the
fact that a singularity (where something tends to infinity) does not occur at the
centre of a real vortex (viscosity prevents infinite values, even though large
finite values can still occur) means that we must limit the range of validity of our
model. Therefore, we have to realise that the mathematical model must break
down before we reach the origin of the singularity (or the ‘eye’ of the vortex).
We cannot expect to model everything using relatively simple tools!

© 2007 P.G. Sayer


Page 14
The stream function for a source can be derived by considering the uniform
radial flow that it generates outwards from the origin. Denoting its strength by
m means that ur = m/(2πr) whereas uθ = 0. Using the formulae for velocity
components in terms of derivatives of ψ leads to

ψ = mθ/(2π)

Should there be a constant of integration, or some other function? The answer is


that we can set constants of integration arbitrarily to zero, without loss of
generality, because the physically relevant quantities, namely velocity,
acceleration, pressure, etc. depend only on the derivatives of ψ.

The next question, then, is probably: “What use is a mathematical source?” The
answer lies in the velocity that it imparts to the fluid. Since velocity is a vector,
it has both magnitude and direction. A flow model can be thought of as a means
of determining the velocity components at every point in the fluid. So in any
locality, we might compute the fluid velocity components and represent them as
arrows, the length of the arrow corresponding to the magnitude of the velocity,
with the arrow aligned with the local direction of the flow. In such a vector
diagram, we can make use of standard vector algebra: two or more ‘arrows’ can
be combined to find the resultant vector. In this way, it is possible to imagine
the vectors cancelling each other, or reinforcing each other. In the former case,
the resultant velocity can become exactly zero; this occurs at a stagnation point.
From Bernoulli’s equation, if the velocity is zero, then the pressure is a (local)
maximum.

Extending these ideas, a source (or line of sources) can ‘deflect’ an incident flow
to such an extent that we could envisage the flow lines tracing out the lines of an
actual ship hull. If this can be done, then our model lets us take a next step
along the route of modelling the velocity (and hence pressure) distribution along
a hull, and so to the net force acting on the hull. [Remember, though, that this
non-viscous model can only partly account for the force acting on the hull; more
precisely, this model leads to what is known as the form drag; there are other
important components, for example skin frictions which depend intrinsically on
the viscosity of the boundary-layer flow and the hull roughness]. Again,
examples in class will be calculated to show how these ideas are implemented in
practice.

One of the simplest examples is that of a source place in a uniform incident


flow. This results in the incident flow being deflected laterally, before
continuing downstream parallel to its original direction. We can extend this in a
simple way by considering a source and a sink of equal strengths, giving flow
around an oval-shaped body. The lateral deflection of the flow, or equivalently

© 2007 P.G. Sayer


Page 15
the maximum breadth (beam) of the body, can be calculated directly as a
function of the source strength and incident flow speed. Conversely, given the
breadth of the body, we can calculate the required source strength to model the
correct flow behaviour.

In the case of an actual ship’s waterline, its shape is rather more complicated
than the example just given! A large number of sources are needed, their
strengths being determined from the condition that the deflected flow must
follow the waterline itself. This leads to a set of simultaneous equations for the
source strengths. So, for practical purposes, the ship’s beam and the other
waterline offsets form the input, while the solution of the simultaneous equations
generate the source strengths as the output.

A major weakness of the non-viscous model is that it would appear that a body
that has fore and aft symmetry will have zero net force acting on it. The
situation is even worse!! This model predicts zero force on a body of any shape
when subjected to a steady flow: this is known as d’Alembert’s Paradox. Its
resolution lies in realising that the model neglects viscosity, which means that
no separation of flow occurs after the flow encounters the body and therefore
that no wake is generated downstream of the body. In reality the separation of
flow results in an imbalance between the fluid pressures fore and aft, and hence
to a non-zero net force in the direction of the flow (put simply, this is called
drag!). [We have also neglected any wave generation that may occur - this also
leads to a component of drag, independent of viscous effects]. So, once more,
we need to appreciate the limitations of our models, while at the same time
building on what we can already model accurately. We must do things
gradually. We must walk before we run!

© 2007 P.G. Sayer


Page 16
Doublets/dipoles and geometrical transformations

Intuitively, the effects of a source and a sink should cancel out. But think more
carefully about the vector arrows, and what happens if the source and sink are
brought close to each other in such a special way that the product of their
strength and the separation distance remains constant, i.e. the strengths increase
in inverse proportion to the distance between their centres. The result is a
doublet or dipole. What do the streamlines look like?

The expressions for a doublet or dipole are as follows:

µ
! = sin#
2"r

µ
ur = cos#
2"r 2

µ
u# = sin#
2"r 2

If we consider the streamlines for a source and a sink, then there is much in
common with the lines of force of a bar magnet. But as the source and sink
coalesce, without annihilating each other, then so the streamlines bend more and
more to form circles. If we take things one step further by putting the doublet in
a uniform stream then the ‘zero’ streamline (across which flow cannot pass; and
so physically can be viewed as an impermeable barrier) is a circle.

We simply add the stream function for a uniform stream to the stream function
for a doublet:

ψ = -U (1 - a2 / r2 ) r sin θ, where we have written a2 = µ/(2πU)

© 2007 P.G. Sayer


Page 17
For offshore engineers, the flow past circular sections is commonly encountered,
e.g. platform legs. But mathematical modelling turns out to have far greater
scope. By using conformal mapping techniques (a special kind of geometrical
transformation) many flows around complicated shapes can be reduced to flows
around much simpler shapes, including circles.

In the field of ship hull generation, the 1930’s saw the introduction of Lewis
transformation techniques, in which several typical hull cross-sections could be
mapped into semi-circles, and hence for which flow characteristics could be
determined analytically (just as well, in those pre-computer days!!). Earlier still,
in the 1900’s, Joukowski showed how flow around foil sections could also be
reduced to flow around circles or ellipses. This led to pioneering developments
in the design of aircraft wings for several decades afterwards. Much more
recently (but still before the arrival of personal computers) the same ideas were
extended to hydrofoils and then to sails, the latter being somewhat complicated
examples of thin, cambered and flexible foils. Little wonder, then, why much
advance in the design of yachts has had to be done empirically and via extensive
physical model testing. Although I have mentioned previously that the
governing equations of fluid motion, including the effects of viscosity, have
been known for over a century, a detailed study of viscous flow past a single
circular cylinder was only feasible computationally in the 1980’s. To put
matters into perspective, it is still impossible to predict mathematically the time
history of droplets of water as they drip from a tap! [Nor, incidentally, can we
predict the eventual configuration of, say, a piece of string which is held
vertically by one end and then dropped to the ground......and there are many
other examples in many other fields].

© 2007 P.G. Sayer


Page 18
Vortices and propeller flow

The streamlines of a vortex are circular. To derive the stream function (or the
velocity potential) we use elementary mechanics and consider the motion of our
fluid ‘lump’ following a circle path. Polar co-ordinates are appropriate and we
consider the resultant of the pressure acting over the four edges of the ‘lump’
and the acceleration towards the centre of rotation uθ2 / r.

Newton’s second law leads to a fairly simple equation:

dp !u" 2
=
dr r

which also shows that pressure increases as we travel outwards across the
vortex.

In irrotational flow, the energy is constant along a streamline, and so we can


make use of Bernoulli’s Equation in conjunction with the above vortex pressure
equation to obtain

c p # p1 c2 $1 1
u! = and 2 = 2 # &
r " 2 r1 r2 2 %

Such a motion is what we term a free vortex.

© 2007 P.G. Sayer


Page 19
In contrast, for a forced vortex the fluid rotates rather like a solid body; so for
angular velocity Ω we have uθ = rΩ and, of course, ur = 0. The pressure satisfies

p2 ! p1 # 2 2
"
=
2
(r2 ! r12 )

Exercise: Show that the stream function for a free vortex is given by

$ r
! =# ln
2" a

where Γ is the circulation and ψ is defined to be zero on the circle r = a. This

© 2007 P.G. Sayer


Page 20
last condition is equivalent to saying that the fluid is bounded internally by a
circle of radius a; no fluid can cross the surface of the circle, so in practice we
could use this to model the flow around a circular cylinder of radius a.

A more obvious marine application, however, is a ship propeller. The flow


between the propeller shaft and the blade tips is acted upon by the blades, and so
an external force (provided, of course, by the engine) alters the energy of that
region of fluid. The resultant flow is rotational (and so only the stream function,
not the velocity potential, can be used). In contrast, just beyond the blade tips,
the fluid flow is irrotational – so, although the centre of gravity of the fluid
‘lump’ traces out a circular path, the fluid lump does not spin about its centre of
gravity. The fluid as a whole still of course follows as circular-type path, and
continues to do so for quite some distance away from the blade tips (the flow
clearly cannot just ‘switch off’ beyond the blade tips), but the vorticity in this
region ‘exterior’ to the propeller itself has no circulation; its total energy remains
constant, and this means that Bernoulli’s equation can be used (in contrast to the
‘inner’ region encompassed by the blades where Bernoulli’s equation is not
valid).

These physical differences between the flow interior and exterior to the propeller
blades can be modelled by forced and free vortices, respectively. Since we
cannot use Bernoulli’s equation in the forced vortex case we need to employ
other results of mechanics (analogous to rigid-body rotation) to make further
progress. Beyond the blade tips, however, the free-vortex motion can be
determined using the Bernoulli relation between pressure and velocity. Without
going into the mathematical details, we can still make some initial observations
which turn out to serve as boundary conditions; these enable us to determine, for
example, constants of integration when we solve the equations, as well as
checking that the pressure and velocity changes tend to zero as we move away
from the propeller. It is always sensible to carry out such checks. There is no
point at all in striving to solve complicated equations if the equations themselves
do not satisfy fundamental conditions. As a specific example, the equations for
the pressure and velocity in the inner and outer regions of a propeller must
generate equal values of the pressure and equal values of the velocity at the
boundary between the inner and outer regions; i.e. the two models must agree at
the circle described by the tips of the propeller blades.

Lift

As a final illustration of our combined ‘Lego-brick’ modelling we examine the


flow past a spinning circular cylinder placed at right angles to a uniform stream
of speed U. Mathematically, this is achieved by adding together the flow due to
a uniform incident stream, a doublet and a vortex.

© 2007 P.G. Sayer


Page 21
% a2 ( r
! = $ U 1 $ 2 ' r sin" $ ln
r & 2# a

1 ƒ! % a2
ur = = $ U 1 $ 2 ' cos"
r ƒ" r &

ƒ! % a2 )a 2
u" = $ = U 1 + 2 ' sin" +
ƒr r & r

On the surface of the cylinder r = a, we have ur = 0 and uθ = 2U sin θ + Ωa,


giving a stagnation point at

# &a
! = !stag = sin "1 " %
2U $
for which the pressure coefficient is
2
p ! p0 " 'a
cp … = 1 ! 4 sin& + $
1 2 2U #
%U
2

The following diagram shows the variation of cp with θ when Ω = 0. What can
we say when Ω ≠ 0?

© 2007 P.G. Sayer


Page 22
From a practical viewpoint, the resulting transverse (lift) force which is
generated (known as the Magnus Effect) gives a means of controlling the depth
and orientation of an underwater vehicle or towed body, more efficiently than
using flaps and thrusters, as well as increasing the turning effect of a rudder (in
the latter case, however, there are more efficient ways of achieving the same end
via flaps on the rudder). Whilst the expression for ψ may appear complicated,
you should recall that it comprises simply a summation of simple flow models.
After going through the mathematical details, we can integrate the pressure
distribution around the circle to obtain the net force. Of course, we must
remember that force is a vector and so take into account its direction when doing
the pressure integration; hence the sin θ and cos θ factors in the integrals. The
final results for the in-line (drag) and transverse (lift) forces are interesting: the
model still predicts zero drag (because viscosity and hence separation are not
modelled) whereas the lift force is computed very accurately: its magnitude is
ρUΓ, quite a remarkably simple expression for a complex piece of flow
modelling. In fact, the result is the same for any lifting surface: different shapes
of bodies are accounted for by different values of Γ.

Symmetrical foil sections are widely used in marine applications. Examples


include the NACA 4-digit series for rudders and the NACA 60 series for yacht
keels. Some important characteristics of lift and drag are given in the following
figure.

© 2007 P.G. Sayer


Page 23
Boundary Layers

We have already remarked that all fluids have some viscosity. Although the
viscosity of water is often neglected, we cannot do this in those narrow regions
of fluid immediately next to a body, i.e. in the boundary layer. The resultant
shear stresses give rise to a drag force exerted on the body.

There are two basic types of flow, namely laminar and turbulent. Especially
within the boundary layer it is important to determine the flow characteristics
accurately; otherwise, the drag predictions can be seriously in error.

For laminar flows we can make use of a simple relationship, known as Newton’s
Law of Viscosity. This relates the shear stress τ (i.e. force per unit area, or force
per unit length in two dimensions) in the boundary layer to the viscosity µ and
the rate of change of (tangential) velocity ∂u/∂y as we cross the boundary layer
at right angles to the surface of the body:

τ = µ ∂u/∂y

The extent of the boundary layer occupies that region of fluid adjacent to the
body in which the tangential velocity component decreases from its mainstream
(undisturbed) value at the edge of the boundary layer, to zero (relative velocity)
at the surface. This implies that the fluid satisfies a no-slip condition at the body
surface, again an essential criterion if basic laws of physics are considered. The
thickness of the boundary layer is usually denoted by δ. For convenience two
more non-dimensional quantities are also introduced: the ratio of the boundary
layer velocity component u to the mainstream velocity U, i.e. u/U, and a
normalised boundary layer thickness parameter η = y/δ. Therefore both
parameters u/U and y/δ vary from 0 at the surface of the body to 1 at the edge of
the boundary layer. Note that in practice we need to place a definite, and
sensible, limit on the edge of the boundary layer (which mathematically extends
indefinitely); it is usual to define the edge of the boundary layer where
u/U=0.99.

© 2007 P.G. Sayer


Page 24
The study of flow over a flat plate gives a very good insight into general
boundary layer characteristics as well as having direct practical applications in
naval architecture; for example, one important component of ship resistance,
namely skin friction, is computed from extensive tests and empirical formulae
based on flat plates. The boundary layer characteristics obviously depend on
whether the incident flow itself is laminar or turbulent, but regardless of this,
there will be a (short) region extending from the leading edge of the plate where
the boundary layer flow must be laminar. Here, the effects of viscosity are so
pronounced that the flow is ‘forced’ to exhibit this laminar behaviour. In
general, flows for which viscous effects are strong, or which have sufficiently
low speeds (in fact, for sufficiently low values of the Reynolds number Rn ) have
streamlines corresponding to ‘layers’ of fluid slipping past each other, in a fairly
well-structured manner. In contrast, turbulence characterises itself by flow
instabilities and a subsequent ‘mixing’ of the layers of fluid; energy and
momentum are interchanged between fluid layers, across the whole boundary
layer, and it is no longer possible to specify, or determine computationally, the
exact, deterministic properties of the flow, such as velocity. We have to be
content to use mean values, to which are added random perturbations.

Further downstream, along the plate, inherent perturbations in the flow will
‘trip’ the flow, changing it from laminar to turbulent. This occurs in the so-
called transition zone. Even further downstream, where the boundary layer can
be regarded as turbulent, there will nevertheless still be a thin laminar sub-layer
next to the body; as before, the effects of viscosity very close to the body are
sufficient to retain the laminar characteristics, and of course at the body surface
itself there will be the no-slip condition. If x measures distances downstream
from the leading edge of the plate, then the transition zone is located where the
value of the local Reynolds number Rx = Ux/ν lies between about 3x105 and 106.
[For flow inside a pipe of circular cross-section, of diameter D, the
corresponding transition value of Reynolds number is given by RD = UD/ν ≅
2300].

It seems reasonable that streamlining a body will help prevent separation of the

© 2007 P.G. Sayer


Page 25
flow and therefore this might be expected to reduce the drag or resistance. Of
course, there are several components of drag, and excessive streamlining will
actually tend to increase skin friction, at the expense of reducing form drag or
separation drag. So, in practice, a compromise has to be sought. One of the
most startling results for drag from classical fluid mechanics relates to the drag
on a two-dimensional foil section compared to that for a circular cylinder. The
separation drag for a circular cylinder of diameter D is at least an order of
magnitude (a factor of 10) larger than that for the foil.

Boundary layer velocity profiles (b.l.v.p.)

Given our non-dimensional parameters defined above, we often make use of


‘standard’ expressions for the velocity profile of the flow within the boundary
layer; these are based on practical observations/measurements, although a few
originate from theoretical arguments. The most frequently used expressions for
laminar boundary layers are:
u/U = 2η - η2

u/U = (3η - η3)/2

u/U = 2η - 2η3 + η4

u/U = sin (πη/2)

For a turbulent boundary layer, the most widely used expression is

u/U = η1/7

The thickness of the boundary layer, not surprisingly, affects the magnitude of
the drag of a body. However, before we can derive expressions for this and
other important results, we need to introduce two additional concepts, namely
displacement thickness δ* and momentum thickness Θ.

Displacement thickness, momentum thickness and skin friction

If we track a fluid particle as it enters the boundary layer near the leading edge
of the plate, we observe that the streamline is deflected away from the body.

Physically, the reason is quite simple: the fluid near to the plate is retarded by
viscosity and so, to maintain continuity, all streamlines have to be shifted
laterally outwards. For flow incident along a thin flat plate, we can view the

© 2007 P.G. Sayer


Page 26
displacement thickness as the difference between the original streamline spacing
and the boundary layer thickness where the streamline penetrates the boundary
layer.

By continuity, the volume flow (i.e. flux of fluid) between any pair of
streamlines remains constant. So
!
U y = u dy
0

Then, noting δ* = δ - y and η = y/δ, we have

1
!*
!
= (1 # u / U ) d"
0

So, for example, the laminar velocity profile u/U = 2η - η2 leads to δ*/δ = 1/3,
whereas the turbulent profile u/U = η1/7 gives δ*/δ = 1/8.

To calculate the drag arising from skin friction (in principle this is the only
component of drag for a thin flat plate) we use Newton’s second law and
consider the initial momentum per second of a thin strip of fluid before it enters
the boundary layer. Its value is (ρ u dy) U. In contrast, within the boundary
layer, the corresponding velocity component is u rather than U, and so the final
momentum per second of the strip is (ρ u dy) u, from which the total rate of
change of momentum for the whole boundary layer is obtained by integrating
the difference between these two expressions across the boundary layer:
"

( ! u dy ) ( u # U )
0

By Newton’s third law, the skin friction R is minus this value.

Alternatively, we may write R = ρ U2 Θ where the momentum thickness is

© 2007 P.G. Sayer


Page 27
given by
&
u # u
!= 1 " % dy
0
U U$

Equivalently, in non-dimensional form,

R = ρ U2 α δ

where
1
u $ u
!= 1 # & d"
0
U U%

But we still do not have an explicit expression for the actual boundary layer
thickness δ ! This we now obtain.

Boundary layer thickness

For laminar flow we can simply equate the shear stress to the gradient (spatial
derivative) of the skin friction, via Newton’s law of viscosity.

ƒu d$ dR
!=µ = " U2# =
ƒy dx dx

!
When we use u/U = 2η - η2 then integration gives
1
= 5.48 Rn " 2
x

The important point here is that the laminar boundary layer thickness is
proportional to Rn-1/2. A more general form of the above result is
1
! $2" 2
= &
x #Rn %

d #u
where != %
d" U $"=0

Turbulent flow is more difficult, partly because we cannot use Newton’s law of
viscosity, since the actual value of, say, the velocity component u comprises
both a mean value and a turbulent variation (perturbation). [It is beyond the
present scope of work to introduce further concepts such as eddy viscosity].

© 2007 P.G. Sayer


Page 28
Instead we take an empirical result, based on the turbulent flow of a fluid
through a circular cylindrical pipe and use this as a means of ‘extracting’ an
expression for δ. It turns out that the resultant expression agrees remarkably
well with actual measurements (and so that may be regarded as our justification
for using it!). We assume that a sufficiently accurate expression for the shear
stress is given by
1
4
2%#
! = 0.0225 " U '
U$ &

d# dR
The next step is to equate this to ! U 2 " = and then integrate. [In other
dx dx
words we assume that the empirical expression gives a result for an ‘equivalent’
flat plate]. For the usual turbulent profile u/U = η1/7 this gives

! 1
= 0.37 Rn " 5
x

So a fundamental difference is that the boundary layer thickness is proportional


to Rn-1/5 for turbulent flow, but proportional to Rn-1/2 in laminar flow. This has
important practical consequences. For example, when Rn = 106 the skin
friction in turbulent flow is more than three times its value in laminar flow;
so there are obvious benefits in trying to retain laminar flow over as long a
length of the plate as possible. [Without wishing to complicate matters, the
reader should be aware that turbulence can be put to beneficial use, for example
in controlling the width of the wake, or in retaining flow attachment to the body;
but these effects are properties of the shape of the body, and are not relevant to a
simple thin, flat plate].

Skin friction resistance in naval architecture

Over the first half of the twentieth century, several famous theoreticians and
experimentalists in fluid mechanics gradually refined the boundary-layer
formulae, according to greater accuracy in experiments and advances in
theoretical knowledge.

Not surprisingly, a mean skin friction coefficient cf is often used to compare


various expressions and approaches, where cf = 2 α δL / L , and δL is the
maximum boundary layer thickness for a flat plate of length L. Our present
approach gives the result

cf = 0.072 [Rn(L)]-1/5

© 2007 P.G. Sayer


Page 29
for turbulent flow, whereas experimental values now suggest a slightly different
value

cf = 0.074 [Rn(L)]-1/5

In fact, as Reynolds number increases, the turbulent velocity profile u/U = η1/7
becomes less accurate. For Rn up to about 108 we can use u/U = η1/N , increasing
the value of N. However, a logarithmic profile is better, for example

u !y %
= 2.5 ln #+ C
U $ &"

where the constant C lies typically between 6 and 8.5.

Schlichting obtained the empirical result

cf = 0.455 [log10 Rn(L)]-2.58

using a logarithmic velocity profile. Prandtl then made a correction to allow for
the laminar region near the plate’s leading edge: thus we have the Prandtl-
Schlichting skin friction formula

cf = 0.455 [log10 Rn(L)]-2.58 - A [Rn(L)]-1

where A = Rcrit ( cfT - cfl )

For Rcrit ≅ 5 x 105 we find A ≅ 1650.

Using a different approach, Schoenherr proposed the formula

[
c f = 0.242 / log 10 c f Rn ( L) ]

Finally, in the 1950’s the I.T.T.C. put forward a formula which is still in
widespread use today:

0.075
cf = 2 + 0.0004
(log 10 Rn ! 2 )

The factor 0.0004 makes an allowance for average effects of hull roughness.

© 2007 P.G. Sayer


Page 30
Courses on resistance and propulsion of ships will expand on these formulae
and how they are used in practice.

Bibliography

Fundamentals of fluid mechanics - B R Munson, D F Young and T H Okiishi, Wiley, 1998.

Fluid physics for oceanographers and physicists - J Williams and S A Elder, Pergamon Press,
1989.

Mechanics of fluids – B.S. Massey and J Ward-Smith, Chapman and Hall, 1998.

Introduction to fluid mechanics - J A Fay, MIT Press, 1994.

Mechanics of fluids - I H Shames, McGraw-Hill, 1982.

Solving problems in fluid mechanics - J F Douglas, Longman Scientific and Technical, 1986.

Fluid mechanics with engineering applications - R L Daugherty, J B Franzini and E J


Finnemore, McGraw-Hill, 1985.

Fluid flow - a first course in fluid mechanics -R H Sabersky, A J Acosta and E G Hauptmann,
Macmillan, 1971.

Fluid mechanics - VL Streeter and E B Wylie, McGraw-Hill, 1985.

The fluid mechanics and dynamics problem solver - The Research and Education Association,
New York, 1987.

Introduction to Naval Architecture - Eric Tupper, Butterworth Heinemann, 1996.

© 2007 P.G. Sayer


Page 31
Tutorial Exercises 1

1.
(i) Verify that ψ=2xy represents a stream function that is valid for irrotational,
incompressible flow. (Hint: consider continuity and vorticity).
(ii) Sketch the streamlines for x>0, y>0.
(iii) Calculate the magnitude and direction of the fluid velocity at the points (0.5,1.0), (2.0,0.25) and
(4.0,0.125), indicating these vectors on your sketch.
(iv) What type of real flow could ψ represent?
(v) Determine the corresponding velocity potential φ
(vi) Sketch the equipotentials for x>0, y>0. What is the angle between corresponding streamlines
and equipotentials?
(vii) Does ψ=xyα represent a valid incompressible flow field? (α=real number).

2.
(i) What type of flow can be represented by ψ=2y(1+x) ?
Sketch the streamlines.
Derive the corresponding velocity potential.

(ii) Repeat this for ψ=2y(1+x+y-1) and comment on your answer.

3. Determine which of the following velocity fields satisfy continuity for an


incompressible flow:
(i) v=6xi-6yj-7tk (t=time)

(ii) v=10i+(x2+y2)j-2xyk

(iii) v=10i+(x2+y2)j-2tk

(iv) v=(ur,uθ), where ur= -U cos θ + a2 r -2 U cos θ, uθ=U sin θ + a2 r -2 U sin θ,


(U=constant).

4.

(i) Verify that ψ=ax2-y2, where a=constant, is a valid stream function.


(ii) Calculate the velocity at a general point (x,y).
(iii) Determine the asymptotes.
(iv) Sketch the streamlines when a=tan θ, θ=30o, indicating the direction of the fluid flow.
(v) What general type of flow could be modelled by ψ ?
(vi) Determine whether this flow is irrotational.

© 2007 P.G. Sayer


Page 32
Tutorial Exercises 2

1. For an irrotational, incompressible fluid show that both φ and ψ satisfy


Laplace's equation

! 2" ! 2" ! 2# ! 2#
+ = + =0
!x2 !y2 !x2 !y2

2. Sketch the streamlines for a doublet, indicating the direction of flow.


Calculate the velocity vector at (x,y)=(1,1) for a doublet of unit strength.

3. A source of unit strength is placed in a uniform stream of velocity 1 ms-1

(i) Sketch the streamlines.


(ii) Calculate the position of the stagnation point.
(iii) Calculate the transverse width of the zero streamline far downstream.

4. Show that the potential flow generated by a doublet in a uniform stream has velocity
components

ur = - U (1-a2 r - 2 ) cos θ
uθ = U (1+a2 r - 2 ) sin θ
in the usual notation.

The pressure coefficient cp is defined by the relationship

cp = (p-p∞)/(0.5ρU2)

Show that cp=1-4 sin2θ for the combined doublet and stream flow.
Hence determine where the fluid pressure is least.

5. A simple pump comprises an impeller of diameter 1.9 m inside a casing of diameter 2 m.


Sketch the pressure distribution for an impeller speed of 1000 rpm. and
hence calculate the maximum difference in radial pressure across the pump.

6. A circular cylinder is spinning steadily about its longitudinal axis, normal to a uniform incident
stream.

Determine the maximum variation in pressure on the surface of the cylinder


in terms of the cylinder and flow parameters. Calculate the magnitude and direction of the
resultant force.

© 2007 P.G. Sayer


Page 33
Tutorial Exercises 3

1. You are given the following velocity profiles:

Laminar u/U=2η-2η3+η4

Turbulent u/U=η1/7

Calculate the ratio of

(a) displacement thickness


(b) momentum thickness

between laminar and turbulent boundary layers when R n=105

2. For the above laminar boundary layer velocity profile, calculate the boundary layer thickness
at the trailing edge of a flat plate of length 0.6m immersed in water having a uniform velocity of
2ms-1, ν=1.2x10-6 m2s-1, ρ=1025 kg m-3.

Calculate also the power required to overcome the resistance of both sides of the plate which
is 1m wide.

3. For the velocity profile u/U=2η-η2 show that the mean skin friction coefficient is approximately
equal to 1.46 Rn-1/2.

A thin flat plate of length 10m is placed in a uniform flow of velocity 1ms-1.
Compare the maximum thickness of the boundary layer with that given by the velocity profile
u/U = sin (πη/2). What flow speed in air would be needed to generate the same boundary layer
characteristics?

4. For the two velocity profiles in Question 3, calculate the ratio of the corresponding skin-friction
coefficients cf and comment on this value.

Show that when Rn=106 the plate has a maximum boundary layer thickness in turbulent flow
more than four times that generated by laminar flow. (You may assume u/U=η1/7 for turbulent
flow).

How important is the skin friction component when considering the overall drag of a marine
vehicle?

© 2007 P.G. Sayer


Page 34

You might also like