Waves, Motions and Manoeuvring Lecture Notes
Waves, Motions and Manoeuvring Lecture Notes
LECTURE NOTES
for
SJO 745 Wave loads and Seakeeping
based on
1 INTRODUCTION 1
1.1 Ship Motions in Waves . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Calculation Chain . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 On Wave-Induced Forces . . . . . . . . . . . . . . . . . . . . . . 5
3 REGULAR WAVES 15
3.1 The Velocity Potential . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Linear Airy Wave Theory . . . . . . . . . . . . . . . . . . . . . . 16
3.3.1 Solution for a progressive wave . . . . . . . . . . . . . . 17
3.4 The Application of the Velocity Potential . . . . . . . . . . . . . 20
3.4.1 Real-valued expressions . . . . . . . . . . . . . . . . . . 21
3.5 Velocity Potential in Deep Water . . . . . . . . . . . . . . . . . . 21
3.5.1 Real-valued expressions . . . . . . . . . . . . . . . . . . 22
3.5.2 Particle paths . . . . . . . . . . . . . . . . . . . . . . . . 23
3.6 Wave Properties in Shallow Water . . . . . . . . . . . . . . . . . 23
3.6.1 Real-valued expressions . . . . . . . . . . . . . . . . . . 24
3.6.2 Summary of Airy waves . . . . . . . . . . . . . . . . . . 24
3.7 “Oblique” Waves . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.8 Wave Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
iii
CONTENTS
4 WIND WAVES 45
4.1 Characteristics of Wind Waves . . . . . . . . . . . . . . . . . . . 46
4.2 Fourier Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.1 Fourier series . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.2 Parsevals equation, orthogonality . . . . . . . . . . . . . 50
4.3 Amplitude Spectra and Phase Spectra . . . . . . . . . . . . . . . 51
4.3.1 Wave record . . . . . . . . . . . . . . . . . . . . . . . . 51
4.3.2 Amplitude, variance and energy spectra . . . . . . . . . . 51
4.3.3 Standard spectra . . . . . . . . . . . . . . . . . . . . . . 53
4.3.4 Moments and wave characteristics . . . . . . . . . . . . . 58
4.4 Synthesised Waves . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.4.1 Directional sea . . . . . . . . . . . . . . . . . . . . . . . 64
iv
CONTENTS
v
CONTENTS
8 HEADING AND
ENCOUNTER FREQUENCY 123
8.1 HEADING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.2 ENCOUNTER FREQUENCY . . . . . . . . . . . . . . . . . . . 123
10 SOLUTION METHODS -
STRIP THEORY 133
10.1 Solution methods in general . . . . . . . . . . . . . . . . . . . . 133
10.2 Strip theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
10.3 Lewis forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
10.4 Experimental methods . . . . . . . . . . . . . . . . . . . . . . . 137
vi
11.7.3 Green water . . . . . . . . . . . . . . . . . . . . . . . . . 177
11.7.4 Propeller emergence and risk for slamming . . . . . . . . 179
11.7.5 Slamming . . . . . . . . . . . . . . . . . . . . . . . . . . 180
11.8 Coupled Linear Pitch and Heave Motion at Forward Speed . . . . 183
11.8.1 The coupling between heave and pitch . . . . . . . . . . . 184
11.9 Coupled Pitch and Heave Motion at Zero Speed Including Non-
Linear Viscous Damping . . . . . . . . . . . . . . . . . . . . . . 185
11.9.1 The wave . . . . . . . . . . . . . . . . . . . . . . . . . . 185
11.9.2 Heave . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
11.9.3 Pitch . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
11.9.4 Equations of motion . . . . . . . . . . . . . . . . . . . . 186
11.10Equivalent Linearised Drag Damping . . . . . . . . . . . . . . . 188
REFERENCES 213
vii
1
INTRODUCTION
Traditionally the sea-keeping and manoeuvring properties of ships have been con-
sidered of second importance to the fuel consumption for cargo ships and forward
speed for faster passenger ships. Knowledge of vertical motions and accelera-
tions are however important for estimating loads on cargo and equipment as well
as the possibility for the crew to work safely. The accelerations may also cause
seasickness to crew and passengers. The relative motion between the ship and the
water surface is important, especially in severe sea states, to investigate the risk of
“green water” on deck, propeller emergence and slamming. In the offshore indus-
try, on the other hand, the sea-keeping properties are the most important proper-
ties, as these properties are decisive for the usefulness of floating platforms. The
“downtime” when one cannot drill, have oil-exporting lines (risers) connected to
the sea floor, moor service or living quarters platforms close to fixed or other
floating production platforms, exchange cargo with supply ships nor even stay on
station must be as short as possible. Therefore, over the past decades, considerable
advances have been achieved in the theoretical prediction of motions of floating
platforms and ships, and in computer algorithms. The manoeuvring capabilities
is of great importance for safe operation of a ship in harbors, restricted waters and
at open sea. The manoeuvring theory is the basis both for tools used to predict a
ships manoeuvrability during the design phase, and for ship simulators used for
education and training of bridge officers. The development of computer process-
ing capacity has made methods for simulation of sea-keeping and manoeuvring
available for practical use..
A ship is moving i 6 degrees of freedom (6 DOF), surge, sway, heave, roll,
pitch and yaw. It is common to treat heave, roll and pitch together and call this the
sea-keeping problem. The surge, sway and yaw are all assumed to take place on
the undisturbed free-surface level. This is often called the manoeuvring problem.
This split is motivated by the small coupling between the degrees of freedom in
the two problems.
1
1. INTRODUCTION
This compendium will describe both the sea-keeping and the manoeuvring
problem. First the environment and the equations of motion will be described
followed by the sea-keeping problem and last the manoeuvring problem.
The sea-keeping part starts from the general equations of motion. Assump-
tions and approximations that can be justified for the sea-keeping problem are
introduced to the equations. A description of wave loads and the ships response
due to the loads are explained. Here the strip theory is introduced as a tool to
solve the sea-keeping problem. Full scale sea-keeping trials and model testing are
briefly explained. Other topics covered are derived responses such as slamming,
green water on deck, freeboard exceedance and propeller emergence. Added re-
sistance in waves and involuntary speed loss are discussed as well as voluntary
speed reduction. The effect on passengers and crew due to ship motions is also
discussed and finally design aspects with respect to sea-keeping is covered.
Also the manoeuvring part starts from the general equations of motion. A
final set of linear equations for the manoeuvring problem will be derived based
on assumptions and approximations relevant for the maneouvring problem. The
experimental setup to determine the stability derivatives in the final equations are
explained. Stability theory will be introduced and ways to analyze and improve
the course stability are discussed. Full scale tests used to determine the ships
manoeuvring properties are described together with the stop manoeuvre. The
influence due to restricted water is described and finally some practical aspects
are covered together with a description of rudders and other steering devices.
The ship motion problem including both sea-keeping and manoeuvring is a
three-dimensional time dependent non-linear problem with 6 DOF. It is neces-
sary to introduce assumptions and approximations in order to solve the problem
for practical purposes. In this compendium the focus will be on frequency do-
main methods based on two-dimensional strip theory for the sea-keeping prob-
lem. This approach was initially put forward by Korvin-Kroukovsky and Jacobs
[27] and was later developed by other authors, is still used extensively and has
been proven by both model tests and full scale trials to predict ship motions in a
seaway with acceptable accuracy. It is especially used for conceptual design and
risk assessment procedures [23].
A natural seaway is always irregular, and an irregular wave system can ap-
proximately be thought of as a superposed sum of an infinite number of regular
sinusoidal waves, each of which is characterized by frequency, amplitude, direc-
tion of propagation and a random phase angle. Under the assumption that the
motion response to the waves is linear, the superposition method can be utilized
also for the sea loads and the ship motions. This assumption is valid for moder-
ately steep waves. Thus once the response to regular waves has been assessed,
the response of the ship in an irregular seaway can be determined. The irregular
seaway or sea state is often represented by a spectrum and by multiplication of
this, for each frequency, with e.g. the linear response ratio of motion in that fre-
quency a response spectrum of the motion can be produced. Thereafter statistical
methods can be utilized to assess characteristics of responses in each sea state or
2
1.1. Ship Motions in Waves
A ship with steady forward speed in irregular short-crested sea will oscillate in six
degrees of freedom. In the simplified case of steady speed in meeting or following
regular waves the ship will heave (vertical motion), pitch (tilting motion) and
surge (bow-aft motion) around its mean forward advancing position. In very long
waves its motion will just follow the sea surface motion but for shorter waves
near the vertical heave and pitch resonances of respective motion the motion will
be strongly amplified and out of phase with sea surface motion. For somewhat
shorter waves the motions will be opposed to the wave motion but less amplified,
so when the crest of the wave passes the ship the ship will be at its lowest position,
and when the wave slopes forward the ship will slope backwards with obvious
consequences for risk of green water and propeller emergence. See Fig 1.1.
The chain of calculations to assess the sea keeping properties of a ship is outlined
in Figure 1.2.
1. Gathering of wave data for the route where the ship will operate. Weather
data may be taken from archived observations, satellite observations or
3
1. INTRODUCTION
4
1.3. On Wave-Induced Forces
6. And finally derive the load effects i.e. sectional forces and moments, ten-
sions, risk for propeller emergence, slamming and green water. For moored
ships and structures also the mooring-line tensions are derived.
One may say that there are two fundamentally different ways to calculate wave-
induced forces on structures in the sea. In one method one considers the structure
as a whole and assesses the total wave force from empirical or computed co-
efficients applied on water velocities and accelerations in the undistorted wave
5
1. INTRODUCTION
motion. In the other method the pressure distribution around the surface of the
structure is computed with due consideration to the water motion distorted by the
structure itself, and subsequently integrated around the structure.
In both cases some mathematical model for describing the wave properties is
necessary. For instance, by making the simplified assumption that the wave mo-
tion can be regarded as potential flow, velocities, accelerations and water motion
can be computed in any point under a gravity surface wave by a scalar quantity,
the velocity potential. In Chapter 2 some basics of potential flow theory will be
given and in Chapter 3 it will be used to derive kinematics of linear waves.
6
2
POTENTIAL FLOW FOR WAVES
Many flow problems are elegantly solved by help of potential flow theory. Free-
surface waves is one of them. It is then assumed that the fluid is incompressible
and the flow irrotational. Irrotational flow is a flow where any selected fluid
packed does not rotate around its centre. It may, however be strongly deformed.
Look at an infinitesimal control volume of fluid with the density, ρ, in a flow with
a co-coordinate system (x , y , z) and corresponding velocity components (u , v , w).
From Fig. 2.1 it is evident that to first order the resulting inflow of mass can be
written
∂ (ρu) ∂ (ρv) ∂ (ρw)
− + + ∆x ∆y ∆z . (2.1)
∂x ∂y ∂z
! "
∂(ρ u)
ρ u ∆y∆z ρu + ∆x ∆y∆z
∂x
∆z
z ∆y
y ∆x
x
7
2. POTENTIAL FLOW FOR WAVES
But this mass inflow must equal the increase of mass in the infinitesimal control
volume, ∆x ∆y ∆z, i.e.
∂ (ρu) ∂ (ρv) ∂ (ρw) ∂ρ
− + + ∆x ∆y ∆z = ∆x ∆y ∆z , (2.2)
∂x ∂y ∂z ∂t
which also can be expanded to
∂ρ ∂ρ ∂ρ ∂ρ ∂u ∂v ∂w
+u +v +w +ρ + + = 0, (2.3)
∂t ∂x ∂y ∂z ∂x ∂y ∂z
or written
dρ ∂u ∂v ∂w
+ρ + + = 0. (2.4)
dt ∂x ∂y ∂z
For a homogeneous incompressible fluid each of the derivatives of ρ is zero,
and if the fluid is incompressible it is easily understood that the total derivative,
dρ/dt = 0, i.e. that the density of a chosen control volume of fluid is not changed
during its motion. Observe, on the other hand, that various parts of the fluid may
have different density due to e.g. varying salinity or temperature. The fact that
dρ/dt = 0 leads to the continuity conditions for both three-dimensional and two-
dimensional flow:
∂u ∂v ∂w
3-D + + = 0, (2.5a)
∂x ∂y ∂z
∂u ∂v
2-D + =0 (2.5b)
∂x ∂y
The rotation of a fluid element around its centre can be expressed by the spatial
gradients of the local fluid velocities in the x-, y- and z-directions. For the two-
dimensional case sketched in Fig. 2.2 we find that the gradients generally would
deform the element, because the section a – a rotates counter clockwise and the
section b - b rotates clockwise with the angular velocities
v + ∆x ∂∂ xv − v ∂ v
= , (2.6)
∆x ∂x
u + ∆y ∂∂ uy − u ∂u
− =− . (2.7)
∆y ∂y
The rotation of the element per unit of time, ωz , around the z-axis is defined as the
mean of the two angular velocities above, i.e.
1 ∂v ∂u
ωz = − . (2.8)
2 ∂x ∂y
8
2.2. Irrotational Flow
1 ∂w ∂v
ωx = − , (2.9)
2 ∂y ∂z
1 ∂u ∂w
ωy = − . (2.10)
2 ∂z ∂x
ωx = ωy = ωz = 0 , (2.11)
∂w ∂v
= , (2.12)
∂y ∂z
∂u ∂w
= , (2.13)
∂z ∂x
∂v ∂u
= . (2.14)
∂x ∂y
Note that even in irrotational flow the fluid element can be strongly deformed,
e.g. be deformed from a cube to a diamond. For examples of irrotational and
rotational flows see Figs. 2.3 and 2.4.
9
2. POTENTIAL FLOW FOR WAVES
(a) (b)
(a) (b)
Under which conditions does a continuous and differentiable function, φ (x, z,t),
exist such that
∂φ ∂φ ∂φ
u= , v= , w= . (2.15)
∂x ∂y ∂z
Equation (2.15) means that the spatial gradient of φ in each point shall give the
velocity vector of the flow, U = ∇φ = grad φ . The function, φ , is therefore named
the velocity potential. Observe that the sign in Eq. (2.15) as well could be set
to minus as it is a pure definition, and so is also the convention in many civil
engineering textbooks and some other scientific literature, which one has to note
when studying different sources. A rational motive for choosing a negative sign is
that a minus sign corresponds to the analogy that the gravity force acts downhill
when going uphill i.e. in opposite direction to the slope. Here we will follow
the less intuitive convention according to (2.15) as this is most common in naval
architecture.
If the flow is irrotational then in e.g. in the x − z plane according to Eq. (2.13)
∂ u/∂ z = ∂ w/∂ x and thus that
∂u ∂ 2φ ∂w ∂ 2φ
= = = , (2.16)
∂z ∂z∂x ∂x ∂x∂z
i.e. that the mixed second derivatives are equal, which in turn proves that φ is
continuous and differentiable with respect to x and z. The same is true in the
10
2.4. Bernoullis Equation for Potential Flow
other two orthogonal planes. The potential function φ exists thus if the flow is
irrotational and irrotational flow is therefore also called potential flow.
Further for an incompressible fluid the condition of continuity Eq. (2.5) is true
which with (2.15) substituted gives Laplace differential equation:
∂ 2φ ∂ 2φ ∂ 2φ
3D + + 2 = 0, (2.17a)
∂ x2 ∂ y2 ∂z
2
∂ φ ∂ φ2
2D + = 0, (2.17b)
∂ x2 ∂ y2
or symbolically ∆φ = 0 or ∇2 φ = 0.
Irrotational flow for an incompressible fluid can thus be described by the dif-
ferential equations (2.17), which are linear differential equations of second order.
A surface water wave motion with “small” amplitude in relation to its wavelength
and the water depth can with good precision be described as potential flow. The
deviation from the true physical wave motion for higher waves depends on ap-
proximations of the boundary condition and on viscous and rotational effects.
Due to the linearity of the differential equations (2.17) a wave motion decomposed
into many harmonic wave components with different amplitudes, frequencies and
phase lags, first the flow for each component can be solved and then linearly be
added to a total solution which gives all velocities and accelerations anywhere in
the fluid.
Velocities and accelerations in potential flow can thus be obtained directly from
the velocity potential, φ , by taking the derivatives with respect to space and time.
To calculate pressures and wave elevations it is, however, necessary to use an ad-
ditional condition, namely the Bernoulli Equation for incompressible, irrotational
flow. Most of us have met it before in the context of one-dimensional pipe flow,
but the version here is somewhat different and valid in three dimensions. In the
next paragraph we will derive it for two dimensions from Navier-Stokes’ Equa-
tions for an incompressible fluid.
11
2. POTENTIAL FLOW FOR WAVES
where t is time, x and z are the space coordinates, u and w denote the velocities in
the x- and z-directions, respectively. The earth acceleration is denoted by g, while
h is height, a coordinate in negative g-direction i.e. h is positive upwards. Finally,
ρ is the density of the fluid and µ the dynamic viscosity of the fluid.
12
2.4. Bernoullis Equation for Potential Flow
and the result is one single equation, the Bernoulli Equation, for an incompressible
fluid in irrotational two-dimensional flow:
∂ φ u2 w2 p
+ + + + gh = f (t) . (2.24)
∂t 2 2 ρ
∂ φ u2 w2 p
+ + + + gh = C , (2.25)
∂t 2 2 ρ
where C is a constant.
For three-dimensional flow the Bernoulli Equation is written:
∂φ 1 2 p
+ u + v2 + w2 + + gh = C (2.26)
∂t 2 ρ
For stationary flow ∂ φ /∂t = 0, and then we recognise the Bernoulli Equation
familiar from stationary one-dimensional pipe flow.
13
3
REGULAR WAVES
If the motion in a water mass due to a surface gravity wave can be approximated by
potential flow, we can derive the properties of the wave motion from the Laplace
Differential Equation as stated in Chapter 2,
∆φ = 0 , (3.1)
where φ is the velocity potential. From φ we can then derive i.e. velocities,
accelerations and pressures everywhere in the water mass.
The Laplace Differential Equation is widely applicable for field problems i.e.
treating heat, sound, electromagnetism and structural mechanics. In civil engi-
neering we meet it for description of ground water flow and for diffusion of heat
and chemical matters in structures. Note, however, that potential flow is free from
losses caused by viscosity, which ground water flow is not.
In this chapter we will treat “small-amplitude” wave theory – also called Airy
wave theory, first order wave theory or linear wave theory – which well describes
waves with the wave amplitude much smaller than the wavelength and the water
depth. For steep waves or finite amplitude waves in shallow water, higher-order
wave theories and non-linear wave theories for shallow water must be used. See
e.g. [53, 10, 29, 45].
The small-amplitude wave theory is, in spite of the underlying simplified as-
sumptions, very useful for many applications. It functions well for wave steep-
ness1 up to H/λ = 0.03 and furthermore, as it is linear, one can superpose solu-
tions for different frequencies and with varying direction of propagation and thus
calculate motion in irregular sea states.
1 Note the difference between the wave steepness H/λ = 2a/λ and the wave slope, which latter
is the slope of the water surface ∂ ζ /∂ x and sometimes the maximum slope which is ak = 2πa/λ
for a sinusoidal wave.
15
3. REGULAR WAVES
The velocity potential, φ , of the wave, Eq. (3.2) must satisfy the Laplace
Equation (3.1) and the boundary conditions (3.3) to (3.5). The solution of this
problem is not easy because the free-surface boundary conditions are non-linear
following the moving free surface.
The simplest, but still very useful, wave theory is the Airy [1], linear or small
amplitude wave theory, which is based on the assumption that the wave amplitude,
a, is small compared to the wavelength, λ .
16
3.3. Linear Airy Wave Theory
The linear wave theory is also called the first-order theory because one can ne-
glect terms that are above first order when expanding the solution in a perturbation
series. The solution of φ and the wave profile are then assumed to be expanded in
power series of a non-dimensional perturbation parameter, ε, in terms of the wave
slope at the zero down crossing of the wave:
2πa
ε= = ka , (3.6)
λ
in which λ is the wavelength, a the wave amplitude and k the wave number. Then
we can write for the potential
∞
φ= ∑ ε n φn , (3.7)
n=1
17
3. REGULAR WAVES
We can from Eq. (3.13) see that the waveform, which is a function of both
time and space, can be separated into a product of two functions, each a function
of only one independent variable. Assuming the solution can be written as a
product of three single-variable functions, the solution could be written
∂ 2 φ ∂ 2 φ 00 00
+ = X (x) Z(z) + X(x) Z (z) T (t) = 0 , (3.16)
∂ x2 ∂ z2
which can be separated into two equations for T (t) not being identically zero:
(
X 00 Z 00 X 00 + k12 X = 0 ,
= − = −k12 ⇒ (3.17)
X Z Z 00 − k12 Z = 0 .
The solution to these two second-order differential equations in x and z have the
forms (
X(x) = Aeı k1 x + Be−ı k1 x ,
(3.18)
Z(z) = Cek1 z + De−k1 z ,
where the sign of Eq. (3.17) has been chosen to get a harmonic solution in the
x-direction. Thus
φ = Aeı k1 x + Be−ı k1 x Cek1 z + De−k1 z T (t) , (3.19)
18
3.3. Linear Airy Wave Theory
Two constants remain to be determined from the bottom and free surface
boundary conditions. The bottom boundary condition gives
∂φ
−k d kd
= k Ce − De X(x) T (t) = 0 . (3.21)
∂z z=−d
As X(x) and T (t) are not identically zero this can be solved for instance for C:
which gives
Z(z) = D ek d ek(z+d) + e−k(z+d) = D1 ek d cosh(k(z + d)) (3.24)
∂ 2φ ∂φ
2
+g = 0,
∂t ∂z
will finally yield a functional relation between k and ω. Starting with
dividing by X(x), and substituting Z(z) (3.24) and T (t) = e−ı ω t gives
D1 ek d cosh(k(z + d)) −ω 2 e−ı ω t + g k D1 ek d sinh(k(z + d)) e−ı ω t = 0 . (3.29)
Finally
ω2
= tanh(k d) (3.30)
gk
19
3. REGULAR WAVES
which implicitly gives the wave celerity, ω/k = f (k), and is called the dispersion
relation, because it shows that the wave celerity depends on the wave length,
which fact in turn makes groups of waves disperse.
The final complex velocity potential for plane progressive waves in finite water
depth is then after having substituted the constants and the dispersion relation:
a ω cosh(k(z + d)) ı (k x−ω t) a g cosh(k(z + d)) ı (k x−ω t)
φc = −ı e = −ı e (3.31)
k sinh(k d) ω cosh(k d)
and its corresponding real valued expression is
a g cosh(k(z + d)) ı (k x−ω t)
φ = Re −ı e
ω cosh(k d)
a g cosh(k(z + d))
= Re −ı (cos(k x − ω t) + ı sin(k x − ω t))
ω cosh(k d)
a g cosh(k(z + d))
= sin(k x − ω t) (3.32)
ω cosh(k d)
The velocity potentials (3.31) or (3.32) for a progressive plane harmonic wave are
very useful, because from one single scalar equation we can derive all particle
velocities and pressures in the water mass beneath the still-water level.
According to the definition (2.15) the complex particle velocities will be
∂ φc a g k cosh(k(z + d)) ı (k x−ω t)
uc = = e , (3.33a)
∂x ω cosh(k d)
∂ φc a g k sinh(k(z + d)) ı (k x−ω t)
wc = = −ı e , (3.33b)
∂z ω cosh(k d)
and the complex particle accelerations becomes, after taking the time derivatives,
∂ 2 φc cosh(k(z + d)) ı (k x−ω t)
u̇c = = −ı a g k e , (3.34a)
∂t ∂ x cosh(k d)
∂ 2 φc sinh(k(z + d)) ı (k x−ω t)
ẇc = = −a g k e , (3.34b)
∂t ∂ z cosh(k d)
The pressure above the atmospheric reference pressure can be calculated using
the Bernoulli equation ( 2.26)
∂ φc 1 2 2
pc = − ρ + ρ uc + wc + ρ g z . (3.35)
∂t 2
Here the first and second term are the dynamic fluctuating pressure due to the
waves and the third term the hydrostatic pressure. To first order the velocity
squared term can be neglected. Thus
∂ φc cosh(k(z + d)) ı (k x−ω t)
pc = −ρ +gz = ρ ag e −ρ gz. (3.36)
∂t cosh(k d)
20
3.5. Velocity Potential in Deep Water
The wave form can be retained from the linearised free-surface boundary con-
dition (3.10):
1 ∂ φc
ζc = − = a eı(k x−ω t) . (3.37)
g ∂t
The real valued pressure above the atmospheric reference pressure to first or-
der is
cosh(k(z + d))
p = ρg a cos(k x − ω t) − ρg z . (3.40)
cosh(kd)
Furthermore the particle paths in the wave can be calculated from the Eq. (3.38)
by integration with respect to time. The resulting form is elliptic with the largest
axis horizontal. Close to the bottom the vertical axis is very short and the water
is oscillating horizontally only. In deep water the particle paths are circular, see
Fig. 3.2.
21
3. REGULAR WAVES
∂ φc
uc = = a ω ekz eı (k x−ω t) , (3.42a)
∂x
∂ φc
wc = = −ı a ω ekz eı (k x−ω t) , (3.42b)
∂z
and the complex particle accelerations in deep water,
∂ 2 φc
u̇c = = −ı a ω 2 ekz eı (k x−ω t) , (3.43a)
∂t ∂ x
∂ 2 φc
ẇc = = −a ω 2 ekz eı (k x−ω t) . (3.43b)
∂t ∂ z
22
3.6. Wave Properties in Shallow Water
where r(z) is recognized as the radius of the particle path. Thus in deep water the
particles move in circular orbits with radii that decrease exponentially with depth.
Exercise 3.1
How large is the particle radius at the level z = −λ /2?
In shallow water the general expression for the velocity potential must be used,
but the derived velocities, accelerations and pressure can be simplified using the
fact that when d → 0 so tanh(kd) → kd.
The complex velocities in shallow water are thus
a ω ı(k x−ω t)
uc = e , (3.52a)
kd
z + d ı(k x−ω t)
wc = −ı aω e , (3.52b)
d
23
3. REGULAR WAVES
For the discussion of motions of ships in obliquely approaching waves, the two-
dimensional linear plane wave above can be considered to be a three-dimensional
wave train with straight, infinitely long wave crests. If the considered wave is
travelling at the angle, µ, anticlockwise to the x-axis it can be described by the
equation:
ζ (x, y,t) = a cos (k(x cos(µ) + y sin(µ)) − ω t) .
In fact, all arguments (k x−ω t) can be exchanged by(k(x cos(µ) + y sin(µ)) − ω t)
in all equations. In some literature the notations kx = k x cos(µ) and ky = k y sin(µ)
are used giving
ζ (x, y,t) = a cos(kx x + ky y − ω t) .
24
3.7. “Oblique” Waves
Table 3.1: Summary of linear (Airy) wave theory. Wave characteristics. Here
L = λ . From [? ]. !"#$$$%&'&$$%%#()*+,#--.
$#/01#%2#(34*156#'.
7850+6#--&$&9:###;0<<*+=#>?#@816*+#(A8+=.#B*C6#,46>+=#&#B*C6#D4*+*D,6+8E,8DE
$33C;*%&-2(;$>(L"(RC"3%&-2$L#"9((<(%+&0P(P&;"23&-2#"33(*$0$;"%"0S(.+&=+(;$>(L"(C3"P(%-(0"*#$="("&%+"0(%+"
.$'"(3%""*2"33(-0(0"#$%&'"(.$%"0(P"*%+S(;$>(L"(P":&2"P($3(%+"(0$%&-(-:(.$'"(3%""*2"33(%-(0"#$%&'"(.$%"0(P"*%+9
1+C3S
$D" $
! ETTU5UNNF
!D" !
!"#$%&!"'$&($)*"+,)- 25
../0/10
3. REGULAR WAVES
Figure 3.3: The mass of the lifted water between the two dashed verticals is
ρgζ dx and this mass is lifted the height ζ /2.
The local mean energy in a regular progressive wave can be estimated by integrat-
ing the potential and kinetic energy in the wave over a wavelength.
The potential energy of the wave at a time instant, e.g. t = 0, can be calculated
as the deformation work needed to give the form of the wave, see Fig. 3.3 . Thus
the mean energy over one wavelength is
Z λ Z λ
1 ζ ρg 1
Ep = ρ gζ dx = ζ 2 dx = ρ g a2 . (3.58)
λ 0 2 2λ 0 4
The kinetic energy is the total kinetic energy contained in all the water from the
free water surface to the bottom of the sea. To first order we can only integrate
from the mean water surface, as the theory is not valid above the mean water
surface, and would in fact give large errors for finite waves. For the wave in deep
water it is especially easy to make the integration as the water particle moves with
constant speed V along their circular paths with the radius decreasing with the
level of submergence, z,
2π
r = r0 ek z = r0 e λ z , (3.59)
where r0 = a is the radius at the mean water level, z = 0, equal to the wave ampli-
tude. The velocity is thus
2πr(z) p 2
V (z) = = u + w2 , (3.60)
T
and the kinetic energy per unit volume is
1
ρ (V (z))2 . (3.61)
2
It is interesting to note that the kinetic energy is constant over the horizontal planes
at each level, while the potential energy varies with time.
26
3.9. Wave Power
The energy transport or wave power per unit width of a plane wave can be calcu-
lated by estimating the work done in the propagation direction on the water mass
to the left.
In deep water the work done at the level z in the vertical x = 0 m during the
time dt is
dW = p dx dz , (3.64)
and per time unit of time
dW dx
= p dz = pu dz . (3.65)
dt dt
Integrated along the vertical the wave power is
Z 0
P(t) = ρg aekz cos(−ω t) − z a ω ekz cos(−ω t)dz
−∞
Z 0 Z 0
= ρga2 ωe2kz cos2 (ω t)dz + ρg(−z)a ωekz cos(ω t)dz
−∞ −∞
Z 0 Z 0
= ρg a2 ω cos2 (ω t) e2kz dz − ρg a ω cos(ω t) zekz dz
−∞ −∞
Z 0 Z 0
1
= ρg a2 ω (1 + cos(2ω t)) e2kz
dz − ρg a ω cos(ω t) zekz dz .
2 −∞ −∞
(3.66)
27
ρ (V ( z ) ) . …(5.61)
2
2 2k 4
The last equality comes from the dispersion relation in deep water
gk = !2. The total wave energy follows as
1
E = E p + Ek = ρga 2 ,
2
which has the unit J/m2, shows that the energy content averaged
over a horizontal area in a small-amplitude, harmonic wave is
proportional to the wave amplitude squared.
dW = p dx dz …(5.62)
The first term on the right hand side in the last line in Eq. (3.66) thus oscillates
with twice the frequency of the wave, but is always larger than zero, while the
second term has the same frequency as the wave but gives no net transport as its
time average is zero.
Z 0 Z
1 1 T
P(t) = ρ g a2 ω (1 + cos(2ω t)) dt e2k z dz
T 2 0 −∞
1 1 1 1ω C
= ρ g a2 ω = ρ g a2 = E = Cg E , (3.67)
2 2k 2 2k 2
from which it is seen that in deep water the wave energy is transported at half the
wave celerity C. This transport velocity, Cg , is also called the group velocity due
to the fact that a group of waves must propagate with this velocity. Otherwise it
would leave its energy behind and disappear. See also the next paragraph.
28
3.10. Wave Celerity, Group Velocity and Particle Velocities
Figure 3.5: The water level as a function of time for a composed wave train
at three points along its route of propagation. The uppermost trace
is “upwind” and the lower ones are increasingly farther downwind.
The arrows indicate how the shorter waves arrive later (solid arrow)
downwind than the longer waves (dashed arrow). This wavetrain
was produced by a moving ship.
λ ω
C= = , (3.68)
T k
where ω and k are interdependent through an implicit dispersion relation, Equa-
tion (3.30), that was derived from the velocity potential
ω 2 = g k tanh(k d) , (3.69)
29
3. REGULAR WAVES
Table 3.2: Some important relations for linear waves at various relative depths.
g g √
Celerity C= C= tanh(k d) C= gd
ω ω
gT 2 gT 2 p
Wavelength λ= λ= tanh(k d) λ =T gd
2π 2π
C g C 2k d √
Group velocity Cg = = Cg = 1+ Cg = C = gd
2 2ω 2 sinh(2k d)
Amplitude of
vertical particle aω aω aω
velocity at the
water surface
Amplitude of p
horizontal particle aω a ω cosh(k d) a g/d
velocity at the
water surface
30
Wave-Induced Loads and Ship Motions
02 December
3.10. 2008 Group Velocity and Particle Velocities
Wave Celerity,
Page 51
Water level
Distance
An alternative to the calculation of the group velocity from the energy transport is to
superpose two progressive waves with almost the same frequencies ω and ω’ and
when it leaves
consequently thethe
almost group
sameinwave
the front.
numbers k and k’. See Figure 5.7.
The speed of propagation of the energy and the wave group is called the group
ζ ( xvelocity
, t ) = a(cos(and ωt )half
kx − is + cos(k ′x −
the ω ′t )) =celerity in deep water and equal to the wave celerity
wave
,2 very
in & 1 shallow # water. & 1 Equations#/ 2 for& the1 group# velocity & 1 are given #/ ) in Table 3.2,
= a +0cos2 $ (kx − ωt ) ! − sin 2 $ (kx − ωt ) !- + 0cos2 $ (k ′x − ω ′t ) ! − sin 2 $ (k ′x − ω ′t ) !- ( =
*1
and %
the 2 group "velocity % 2 in deep ". 1water % was
2 derived " from% 2 the energy". ' ...(5.68)
transport in
Eq. &(3.67).
1 1 # &1 1 #
= 2a cos$ (k ′x − ω ′t ) + (kx − ωt ) ! cos$ (k ′x − ω ′t ) − (kx − ωt ) ! =
% 2 alternative
An 2 to the"calculation
%2 of the 2 group" velocity from the energy trans-
port& is to superpose two# progressive
1 & 1
= 2a cos$ (kx − ωt ) + (!kx − !"t ) ! cos$ (!kx − !"t ) !
#
waves with almost the same frequencies ω
and %ω ∗ and consequently
2 " almost
%2 the same " wave numbers k and k∗ . See Fig. 3.7.
whereζ (x,t) k′ −
!k = = k and x − ω t) + cos(k∗ x − ω ∗t))
a (cos(k
!" = ω ′ −ω .
12 2 1
= a cos (k x − ω t) − sin (k x − ω t)
As δk and δω are assumed 2 to be small we can write
2
2 1 ∗ ∗ 2 1 ∗ ∗
+ cos (k x − ω t) − sin (k x − ω t)
&1 #
(kx − ωt ) cos$ (!kx − !"t ) !
2
ζ ( x, t ) = 2a cos 2 …(5.69)
∗% "
1 ∗ 2 1 1 ∗ ∗ 1
= 2a cos (k x − ω t) + (k x − ω t) cos (k x − ω t) − (k x − ω t)
2 2 2 2
The “carrier wave” a cos (kx − ωt ) will thus be modulated
by the function
1 1
= 2a cos (k x − ω t) + (δ k x − δ ω t) cos (δ k x − δ ω t) , (3.71)
2 2
&1 #
cos$ (!kx − !"t ) !
where δ%k = k − k and
2 ∗ " δ ω = ω ∗ − ω. As δ k and δ ω are assumed to be small we
31
3. REGULAR WAVES
Figure 3.7: Snapshot of a wave composed of two sinusoidal waves of almost the
same wavelength. The longer-period wave is the modulated wave
and the shorter-period wave is the “carrier” wave.
can write
1
ζ (x,t) = 2a cos(k x − ω t) cos (δ k x − δ ω t) . (3.72)
2
The “carrier wave” a cos(k x − ω t) will thus be modulated by the function
cos(0.5(δ k x − δ ω t) which has the phase velocity δ ω/δ k.
Each group of waves that is contained between the zero-crossings of the mod-
ulating function is thus moving with this velocity, and therefore it is denoted the
group velocity.
In the limit when δ k → 0 the group velocity will be given by the partial deriva-
tive of k in Equation (5.30) with respect to ω
δk ∂k
lim = = Cg . (3.73)
δ k→0 δ ω ∂ω
Recall that the linear wave theory assumes that the wave height is small compared
to the wavelength and water depth. In natural wind waves the steepness H/λ sel-
32
3.11. Finite Amplitude Waves and Higher-Order Waves
dom exceeds 0.05 to 0.08 in deep water, so the small amplitude theory is often rea-
sonably valid. In some applications it is, however, necessary to use non-linear or
finite-amplitude wave theory. Physically the difference between linear and finite-
amplitude theories is that finite amplitude-theories consider the influence of the
wave itself on its properties. Therefore the phase speed, wavelength, water sur-
face elevation and other properties are functions of the actual wave height.
There are a number of different finite-amplitude wave theories. For deep to
intermediate-depth water (d/λ < 1/8) the most commonly adopted is the theory
by Stokes (1847, see e.g. [53]). For shallower water Cnoidal Wave Theory (Ko-
rteweg de Vries, 1895, see e.g. [53]) or Stream Function Theory (see e.g. [10]) is
more applicable. Williams [54] has produced tables of progressive gravity waves
covering the full range of wavelength from solitary to infinite-depth waves, and
up to the waves of limiting heights with sharp crests.
33
3. REGULAR WAVES
Figure 3.8: An example of a second order Stokes wave as a function of the hori-
zontal co-ordinate x. The crest and trough elevations are also given.
wave profile of the Airy theory. In the figure the parameter, d/gT 2 , is a shallow-
water parameter (∼ water-depth to deep-water wavelength) and the parameter,
H/Hb , is a kind of steepness parameter (ratio between actual wave height and the
breaking wave height for the considered water depth and wavelength). In the left
low corner the profile of a moderately steep wave in intermediate water is shown,
and it can be seen that this profile is reasonably sinusoidal. As a contrast, in the
right high corner a maximally steep wave in very shallow water shows a profile
far from sinusoidal.
In Fig. 3.10 a graph over areas of best fit for wave theories according to
Le Méhauté, as published by [49], is shown. The horizontal axis is a water-depth
to deep-water wavelength parameter, d/gT 2 , and the vertical axis a wave height
to deep-water wavelength parameter, H/gT 2 .
34
3.11. Finite Amplitude Waves and Higher-Order Waves
Figure 3.9: Dimensionless wave profiles for 40 cases of steep periodic waves in
shallow to intermediate water depth (the numbers on each plot rep-
resent the value of H/gT 2 for each case). Here η = ζ λ . From [49].
35
3. REGULAR WAVES
Figure 3.10: Regions of validity for various wave theories. Here L = and Lo is
the deep water wave length. From [49].
36
3.12. Propagation and Transformation
Figure 3.11: Comparison of measured and theoretical wave profiles . From [53].
Weather systems usually move at a much slower speed than the wind velocity
within them. As the celerity of the waves in a well developed sea state are ap-
proximately the same as the wind velocity, and the group velocity is less than that
- in deep water only half the celerity - the result will be that the waves run out of
the windy area where they are generated. After leaving the generation area they
are no longer acted upon by the wind, and rather soon internal friction (viscosity),
parasitic capillary waves and air resistance will dissipate the sharp wave crests
and the shortest components of the spectrum. Such a free wave system is referred
to as swell. Also, due to dispersion the longer faster waves with longer periods
will arrive at distant points long before the shorter waves. The total effect is that
at distant points the swell can become almost monochromatic with a wave period
slowly decreasing with time, as decreasingly shorter waves continue to arrive. In
the low atoll islands of Polynesia, traditionally, the first arriving swell served as
an alarming forewarning of approaching hurricanes [53, 26].
37
3. REGULAR WAVES
Figure 3.12: Variation of the shoaling coefficient, Ks = H/H0 , with the non-
dimensional water depth, d/L0 . L0 is the deep-water wavelength.
3.12.2 Shoaling
P̄ = Cg0 E0 = Cg E , (3.75)
where the index 0 denote ‘deep water’, Cg is the group velocity and E is the wave
energy.
With the energy at any depth proportional to amplitude squared – in deep
water E0 = 0.5ρ g a20 and in finite-depth water E = 0.5ρ g a2 – the ratio between
the wave amplitude or wave height in finite depth water to that in deep water can
be solved. This ratio is called the shoaling coefficient,
s
a H Cg0
Ks = = = . (3.76)
a0 H0 Cg
As shown in Fig. 3.12, going from deep water into shallower water, the shoal-
ing coefficient first decreases slightly below one before increasing rapidly. At the
same time the wavelength becomes shorter due to decreasing celerity, and finally
the waves may break. See the paragraph on wave breaking below.
38
3.12. Propagation and Transformation
Figure 3.13: Oblique waves traversing a uniformly sloped shelf. From [44].
3.12.3 Refraction
As pointed out above, the wave celerity decreases with water depth. Then, if a
long-crested wave approaches a uniformly sloped shelf at an oblique angle the
wave slows down and the wave crest will bend to become more parallel to the
depth contours, see Fig. 3.13. This phenomenon is equivalent to the refraction of
light in optics.
Going into more detail to see how the refraction affects the wave height one
can follow wave rays or wave rays normal to the wave crests and parallel to the
wave celerity in each point. Being parallel in deep water they will either spread
or approach each other approaching land, depending on the bottom topography.
Further, assuming that no energy will be transported across the wave rays but be
contained between two adjacent wave rays, the effect will be decreasing wave
height if they spread and increasing wave height if they approach each other.
Letting them be evenly distributed in deep water at a distance b0 , the distance
will become b somewhere closer to land. The result will be a change pin wave
height by a factor additional to the shoaling effect. The factor, Kr = b0 /b, is
called the refraction coefficient. The total effect of shoaling and refraction will be
s r
a H Cg0 b0
= = = Ks Kr . (3.77)
a0 H0 Cg b
39
a H go bo
= = = Ks Kr . …(5.74)
ao H o Cg b
On a3.straight, shoaling coast the wave rays will spread and thus the refraction
REGULAR WAVES
coefficient will become lower than one: Kr < 1. See Figure 5.14.
Figure 3.14: Orthogonal spacing over a uniformly sloped beach. From [44].
Figure 5.14 Orthogonal spacing over a uniformly sloped beach.
(From Silvesterxxvi, 1974)
On On a straight,
a coastline withshoaling coast and
headlands the wave
bays rays will spread
the result and thusand
of refraction the shoaling
refraction will be
coefficient
increased wavewillaction
become
onlower r < decreased
than one: Kand
the headlands 1, see Fig. 3.14.
action in the bays. See Figure
5.15. On a coastline with headlands and bays the result of refraction and shoaling
will be increased wave action on the headlands and decreased action in the bays.
See Fig. 3.15.
In water deeper than half the wavelength there is no refraction, but the waves will
spread around steep rock peninsulas, piers and structures with steep walls due to
diffraction, when energy spreads along the wave crests into “shadow” areas.
While waves approaching a gradually shoaling coastline will be absorbed by
bottom friction and wave breaking, waves hitting steep rocks, piers or other struc-
tures will be reflected. In the examples shown in Figure 3.16 e.g. waves will be
reflected on the “wave ward” side of the structures, so that a complicated wave
pattern composed of reflected, incident and diffracted reflected waves will set up
on the “wave ward” side. An example is shown in Fig. 3.17. If thus the bottom is
steep or structures dominate an area, where wave propagation shall be modelled,
models must be used that can take diffraction and reflection into account, while
simpler models may be used for mildly sloping bottom topographies.
40
3.12. Propagation and Transformation
with γ = 0.142 [33]. In Fig 3.10 this limit is denoted H0 /L0 = 0.14. In practice
however γ = 0.12. In deep water this means that the steepness will normally not
exceed 0.12. See, e.g., Silvester [44] for further details. Although steeper waves
than given by Eq. (3.78) have been registered now and then, it is not until lately
it has been recognised that these rogue or freak waves may be more frequent than
anticipated, and thus are responsible for many losses of ships and some unfore-
seen damage to offshore drilling platforms. Such waves are three-dimensional in
character and just now (2004) subject to much research2 . In an irregular sea the
steepness breaking will transfer some energy to longer wave components [44] but
largely involves dissipation of energy of shorter wave components. See Young
[55] for further details.
If the wave does not break before it has entered the sloping beach the slope
itself has an influence on the breaking process. One criterion by Collins, as cited
by [44], takes this into account:
Hb
= 0.72 + 5.6S , (3.79)
d
2 http://www.esa.int/esaCP/SEMOKQL26WD index 0.html
41
3. REGULAR WAVES
42
3.12. Propagation and Transformation
Figure 3.17: Sketch of wave crests of regular waves incident, reflected and
diffracted against a semi-infinite breakwater. From [44].
where S is the bottom slope. The breaker height is thus increased at breaking.
Often however a simplified expression is used only containing the water depth:
Hb
= 0.78 . (3.80)
d
In Fig. 3.10 this limit is denoted H/d = 0.78. Different types of breakers are
shown in Fig. 3.18. The breaker type depends on deep-water steepness, beach
slope and wave period. See e.g. [49, 44].
For irregular waves there are some different approaches to depth limited break-
ing. Battjes and Jansen (as cited by [55]) e.g. look at individual waves assuming
them to be Rayleigh distributed, and let all waves with heights above the limiting
criterion be dissipated. Young [55] limits the total energy of the spectrum by a
criterion containing the average wave celerity.
43
3. REGULAR WAVES
e.g. for tidal waves, the “Bernoulli wave”, or shoaling secondary ship generated
waves. Silvester [44] gives an account for regular waves, and Young [55] for ir-
regular waves. The degree of dissipation is governed by the bottom roughness,
which depends on grain diameter of the bottom material, the geometry of ripples
or dunes etc. Permeable and soft bottoms increase the dissipation.
The lost wave energy is partly dissipated into heat and partly used for erosion,
ripple and dune formation, and net transport of bottom materials.
44
4
WIND WAVES
In this chapter some properties of real wind waves are described, it is shown how
they can be looked upon as a linear combination, superposition, of regular waves,
and how realistic wind waves can be synthesised or simulated. Last some basic
wave statistics are given. For broader information, see e.g. the classical book by
Kinsman [26] or a more recent book by Dean and Dalrymple [10].
Waves at sea show a constantly changing, never repeated pattern. They grow
under the action of the wind, and during the growth phase the wave-height, wave
period and wavelength are due to the wind force (wind speed), the duration of the
wind and the length, fetch, of which the wind can act on the waves. On the high
seas the possible wave height is thus limited by the strength, diameter and motion
of the low-pressures, in lakes and landlocked seas of the wind speed and distance
to the upwind shore. At Fully Arisen Sea (FAS) the celerity of dominating waves
approach the wind speed and as a result the wind cannot transfer more energy
to the waves. The exact mechanisms for the generation of waves from a smooth
wave-surface are still not completely explained, but there are hypothetical models
describing the energy transfer from the wind to the waves, e.g. Jeffreys’ and
Phillips-Miles’ models (See Massel [32]). Based on these, empirical functions
were developed during the first half of the 20th century notably by Sverdrup,
Munk och Bredtschneider (See SPM 1983). Later more sophisticated models
have been developed, that in some countries are run on a daily basis to give wave
prognoses. An orientation over recent models is given by Young [55].
Under the progression towards coasts and beaches the waves will be affected
by the bottom so that their height, wavelength and direction of propagation are
changed due to variation in depth, bottom friction and currents still under the
influence of the wind. These effects, refraction, reflection, diffraction and wave-
breaking are equivalent to those of regular waves and were explained in Chapter
3.
If the waves progress out of the low-pressure where they are generated by
45
Wave-Induced Loads and Ship Motions
4. WIND WAVES 02 December 2008
Page 114
0.5
+
H a Still water
Level (m)
0 level
a-
0.5
T trough
1
60 80 100 120 140
Time (s)
Figure 7.1 Some fundamental definitions of wave properties.
Figure 4.1: Some fundamental definitions of wave properties.
In space, the wavelength, λ, corresponding to the wave period is defined as the
horizontal distance between two consecutive up-crossings in the direction of wave
advance,
the strongandwind
the wave heightofand
into areas amplitude
little withinthe
or low wind, a wavelength are defined
shortest waves in
will gradually
analogy
be dissipated and the longer waves will due to their higher celerity overtake and
to those in time. For an irregular wave as in Figure 7.1 the wave heights the
amplitudes in space are normally not the same as those in time, due to the dispersive
shorter waves, so as a result, at a place far away from the low-pressure, they will
properties of the wave.
be almost regular but with a frequency gradually increasing with time. Such waves
Inare called
this swell. we will not delve further on the generation of the wind waves,
compendium
An example
but direct our interest of atowards
time trace of a waveofisdifferent
the description shown in seaFig. 4.1.
states andInstatistics
the figure
for
also some basic, fundamental definitions are illustrated. The
calculating of design waves. Wave generation is described thoroughly by e.g. wave height, H, is
the difference
Kinsman (1965). in level between a wave crest and the following wave trough, the
positive amplitude, a+ , is the crest height above the still-water level, the negative
7.1
amplitude, a− , is the trough
Characteristics of Wind
depthWaves
below the still-water level, and the wave period,
T , is the time lapse between the up-crossings of the still-water level (in some
treatises the down-crossings are used instead of the up-crossings). Note that an
up-crossing in time corresponds to a down-crossing in space.
In space, the wavelength, λ , corresponding to the wave period is defined as
the horizontal distance between two consecutive up-crossings in the direction of
wave advance, and the wave height and amplitude within a wavelength are defined
in analogy to those in time. For an irregular wave as in Fig. 4.1 the wave heights
and amplitudes in space are normally not the same as those in time, due to the
dispersive properties of the wave.
In this compendium we will not delve further on the generation of the wind
waves, but direct our interest towards the description of different sea states and
Figure
statistics 7.2 Point record
for calculating of awaves.
of design wave elevation. The wave
Wave generation is periods defined
described as zero
thoroughly
by e.g. Kinsman [26]. up-crossing periods.
In Figure 7.2 an example of a time trace of a wave elevation in a point is shown. The
wave
4.1 periods are there definedOF
C HARACTERISTICS as W
theIND
times between the zero up-crossings. Before the
WAVES
advance of computers such traces were in the form of paper graphs and were
evaluated byan
In Fig. 4.2 hand. The result
example of such
of a time traceanofevaluation was a list
a wave elevation inof wave is
a point periods, i, and
shown.TThe
connected waveare
wave periods heights, Hi. From
there defined asthat
the atimes
seriesbetween
of characteristics of the wave can
the zero up-crossings. be
Before
defined and is still used, although nowadays the methods of evaluation
the advance of computers such traces were in the form of paper graphs and were are different,
which will be described later. One used to say that the analysed wave record should
evaluated by hand. The result of such an evaluation was a list of wave periods, Ti ,
contain at least 200 waves for the analysis to be meaningful.
and connected wave heights, Hi . From that a series of characteristics of the wave
Mean zero up-crossing period, often called only zero up-crossing period:
46
4.1. Characteristics of Wind Waves
Figure 4.2: Point record of a wave elevation. The wave periods defined as zero
up-crossing periods.
can be defined and is still used, although nowadays the methods of evaluation are
different, which will be described later. One used to say that the analysed wave
record should contain at least 200 waves for the analysis to be meaningful.
The mean zero up-crossing period, often called only zero up-crossing period,
read:
1 Nz τ
Tz = ∑ Tzi = , (4.1)
Nz i=1 Nz
where Nz is the number of waves in the record, Tzi the individual zero-crossing
periods of Fig. 4.2 and τ the length of the record. This definition leaves shorter
small waves riding on the long waves uncounted.
For certain purposes we also need to define periods between local time max-
ima or crests, the crest periods, (see Fig. 4.3) from which the mean crest period is
defined as
1 Nc τ
Tc = ∑ Tci = , (4.2)
Nc i=1 Nc
where Nc is the number of waves in the record and Tci the individual crest periods
of Fig. 4.3.
In the example case above, Tz = τ/5, Tc = τ/7 and generally Nc ≥ Nz and
Tc ≤ Tz .
The wave heights are normally referred to the zero crossing definition, and the
zero-crossing wave heights and number of waves are used from here and onwards
Figure 4.3: The same point record of a wave elevation as in Fig. 4.2. The wave
periods defined as wave crest periods.
47
4. WIND WAVES
without index. There are many possibilities to characterise the wave height. The
most common are defined below.
2 1 N 2
Hrms = ∑ Hi . (4.4)
N i=1
• Significant wave height the mean of the highest 1/3 of the N wave heights:
H1/3 = Hs . (4.5)
Assessment of wave characteristics directly from time traces is not very repro-
ducible, mostly because the number of waves and consequently Tz are very much
depending on how small undulations of the time trace that is taken into account.
Due to modern computer processing and algorithms developed for Fast Fourier
Transformation (FFT) this latter Fourier technology is used in stead, which makes
possible objective, reproducible filtering.
48
4.2. Fourier Analysis
From analysis we know that every piecewise continuous function can on a finite
interval be approximated by a sum of sine and cosine functions. A point registra-
tion of the wave elevation, ζ (t) = f (t), can thus be written on (0 < t < T ):
N
1 2π 2π
f (t) ≈ a0 + ∑ ai cos i t + bi sin i t . (4.8)
2 i=1 T T
It is then implicitly assumed that the point registration is repeated for the registra-
tion period, T . See Fig. 4.21 for a simulated point registration of a wave elevation.
The coefficients ai and bi can be calculated by
Z
2 T +t0 2π
ai = f (t) cos i t dt , i = 0, 1, . . . , N , (4.9)
T t0 T
Z T +t0
2 2π
bi = f (t) sin i t dt , i = 1, 2, . . . , N . (4.10)
T t0 T
49
4. WIND WAVES
N
1 2π
f (t) ≈ a0 + ∑ ci cos i t + εi , (4.11)
2 i=1 T
where
q
ci = a2i + b2i , (4.12)
and
bi
εi = arctan − . (4.13)
ai
The Fourier series (4.8) or (4.11) approximates f (t) well if the number of com-
ponents in the series is sufficient. Observe also that the longest wave that can be
detected by a record with the length, T , is a wave with the period, T . The mea-
surement must thus be long enough in relation to the wave periods contained in
the seastate to give any relevant information. Usually 100 to 200 components are
sufficient to approximate a smooth function such as a train of non-breaking waves.
In Chapter 3 the energy in a harmonic water gravity wave was shown to be pro-
portional to the square of its amplitude, 0.5ρga2 . For a Fourier series with zero
mean, (a0 = 0), the sum of the square of the component amplitudes equals the
variance of the function itself
1 Z T +t0
1 2 1 N 2 1 2 1 N 2
a0 + ∑ ci = a0 + ∑ ai + b2i = ( f (t))2 dt = Var ( f (t)) .
4 2 i=1 4 2 i=1 T t0
(4.14)
This equation is a form of Parseval’s equation.
The consequence for water waves, where we only take the elevation in relation
to the mean water level into account, is then that the waves are not only geomet-
rically additive but that also the sum of the energy of the components equals the
energy of the composed sea state, at least over one period of analysis, T . Also, all
wave-induced forces and resulting motions are orthogonal and can be calculated
for each component frequency independently of the motions at other frequencies.
Then the resultant irregular motions are given by superposition of the component
motions with due respect to the random phases of the wave components.
The derivation of Fourier series and Parseval’s equation is based on the or-
50
4.3. Amplitude Spectra and Phase Spectra
where i and j are positive integers. For a more exhaustive treatise on Fourier
analysis see e.g. Hildebrand [18], or other textbooks in calculus.
The Fourier series, resulting from the analysis of the wave records can be illus-
trated graphically as amplitude and phase spectra. To illustrate this we will in this
section analyse a sample wave record.
51
4. WIND WAVES
is 214 = 16 384 which gives the useful time-length to TR = 2.5744 h. The maxi-
mum resolution of the discrete variance spectrum is then ∆ f = 1/TR , where TR is
the length of the record, and the resulting amplitude, variance or energy spectra
will in this case contain 8 192 discrete amplitudes. In Fig. 4.7 the resulting com-
ponent amplitudes as a function of angular frequency or the amplitude spectrum
is shown and in Fig. 4.8 the corresponding component phases. The graphs reveal
that the amplitudes, although having stochastic magnitudes, show some kind of
pattern with the dominant amplitudes around 0.7 rad/s, period around 8 s, but that
the phases seem to be completely random.
If the original time record should be reconstructed both the amplitudes and
phases must be saved. However, mostly only the component amplitude spectra
are saved, as any particular wave record is seen as one realisation of many pos-
sible. Even a few wavelengths downwind or upwind the shape of the elevation
graph would have been different due to the different celerity, dispersion, of the
component waves.
Traditionally the information is saved as a variance spectrum or “wave-energy
spectrum”. Recall that the variance of each component is 0.5a2i and the energy
0.5ρg a2i . Such a discrete variance spectrum of the sample wave record is shown
in Fig. 4.9. To make the further discussion more clear the variance spectrum is
52
4.3. Amplitude Spectra and Phase Spectra
a2i
Si = . (4.19)
2∆ω
The result of this action is shown in Fig. 4.12, and we now have a spectrum
that resembles the standard, empirical spectra used in the ship industry. Such stan-
dard spectra are the result of assembling and taking the mean of many measured
spectra. We will try to fit some standard spectrum to the smoothed measured
spectrum. But to do that we will describe the standard spectra, and how wave
characteristics are derived from spectra by using spectral moments.
53
4. WIND WAVES
Figure 4.10: The discrete variance spectrum of Fig. 4.9, blown up between 0
and 2 rad/s, as a function of angular frequency.
Figure 4.11: The smoothed discrete variance spectrum, blown up between 0 and
2 rad/s, as a function of angular frequency.
54
4.3. Amplitude Spectra and Phase Spectra
Figure 4.12: The smoothed density variance spectrum or “wave energy spec-
trum” of the sample record.
For this spectrum T01 = 1.086T02 and Tm = 1.408T02 = 1.14T0 . The periods
T01 and T02 are estimates of the zero-upcrossing period, Tz , and will be defined
in the next section. Tm is the modal period or the period for the spectral peak.
T0 = 2π/ω0 .
A variant of the one-parameter PM-spectrum is the International Towing Tank
Conference (ITTC) spectrum:
− 3.11 −4
2 ω
SIT TC (ω) = α g2 ω −5 e Hs , (4.21)
In this spectrum the significant wave height, Hs , is used instead of the wind speed
or mean period.
Mostly the sea state is, however, not fully developed as the wind speed and di-
rection change, the fetch is too short, or the duration is not long enough, especially
for strong winds and high waves. Then two-parameter spectra for developing seas
can be used, e.g. some where the wave height and frequency are the parame-
ters. This was originally proposed by Bredtschneider and offers more flexibility,
because the energy of the spectra can be placed at arbitrary locations on the fre-
quency axis:
1.25 2 ωm4 −1.25( ωm )4
SB (ω) = H e ω , (4.22)
4 s ω5
55
4. WIND WAVES
56
4.3. Amplitude Spectra and Phase Spectra
Spectral moments and the periods T01 and T02 will be defined in the next section.
2
−4
T01
Hs −0.44
SISSCa (ω) = 0.11 2 ω −5 e
ω
2π
, (4.23a)
T01
2π
2
−4
T02
1 Hs −5 −( π1 ) 2π ω
SISSCb (ω) = 2 ω e . (4.23b)
4π T02
2π
Recommended values are, when the fetch, F, and the wind speed, U10 , at
10 m height is used: γ = 3.30 , σa = 0.07 , σb = 0.09 , α = 0.076F0−0.22 , ωm =
7π(g/U10 )F0−0.33 . Here F0 = gF/U10 2.
The expression (4.25) gives identical results to (4.24) provided T01 and Hs are the
same.
1 Notations are different but numerical difference is in the order of 1 %.
2 Using T1 instead of ωm it looks different.
57
4. WIND WAVES
58
4.3. Amplitude Spectra and Phase Spectra
Figure 4.14: A JONSWAP spectrum and PM spectrum with the same modal
frequency, ωm = 0.878ω0 = 0.431 rad/s.
Figure 4.15: A JONSWAP spectrum and PM spectrum with the same significant
wave height Hs .
59
4. WIND WAVES
Figure 4.16: Fit of ISSCa, ISSCb, ITTC and JONSWAP spectra to the measured
smooth spectrum (S j) shown in Fig. 4.12 using calculated T01 , T 02
and Hs respectively for the first three and a manual fit of the five
parameters of the JONSWAP spectrum.
60
4.4. Synthesised Waves
In Fig. 4.17 the same wave record as in Fig. 4.5 is shown but now lines for plus
and minus half the calculated significant wave height have been added in order to
give a feeling for the connexion between the significant wave height and the wave
record. In Fig. 4.18 T01 , T02 and T24 to show that these are of the expected order
of magnitude as the individual zero crossing periods, Tzi , and wave crest periods
Tci . However, as would be expected it is almost impossible to find an individual
period exactly equal to the corresponding mean.
The observation of a wavy lake or ocean surface reveals that the surface normally
is pretty confused with significantly varying wave heights, periods and wave-
lengths, as well as waves progressing in various directions. By using a the spectral
Fourier model described in the previous section, i.e. assuming that the wavy sur-
face is built up by an addition (superposition) of linear waves one can describe
records of such wavy surfaces and reproduce or synthesise them.
The water-surface elevation in a point x is then approximated by a sum
where ai , ωi , ki and εi are the amplitude, circular frequency, wave number and
phase of the ith component, respectively. A fundamental prerequisite for the
waves to be modelled in such a way is that they should be part of a stationary
process whose statistical properties may not change abruptly with time.
Further, the sum above is usually considered to be a Fourier series with or-
thogonal components on a finite time interval, e.g. TR = 20 minutes. This has
the consequence that the elevation is repeated exactly every time interval. The
repeating time, TR , implies the frequency division, ∆ f = 1/TR . Now, choosing
amplitudes of the components from some established standard spectrum and as-
signing random phases distributed evenly between 0 and 2π radians (Fig. 4.20), a
synthetic irregular wave can be formed. In Figs. 4.19 to 4.21 the steps leading to
such a wave is illustrated. We have used a PM spectrum.
The amplitudes are normally chosen as
p
ai = 2S(ωi )∆ω , (4.37)
61
4. WIND WAVES
Figure 4.17: The sample wave elevation with +/- half the significant wave Hs ≈
Hm0 height for comparison.
Figure 4.18: The sample wave elevation with characteristic periods estimated
from its spectrum.
62
Wave-Induced Loads and Ship Motions 4.4. Synthesised Waves
02 December 2008
Page 132
ai
0 5 10 15 20 25 30 35
i
Figure Fig.
4.19:7.19 Thirty-two
Thirty-two component
component amplitudes
amplitudes chosen
chosen by help
by help of aofPM
a PM
spec-
trum.spectrum
εi
0
0 10 20 30 40
i
The assigned amplitudes and phase angles inserted into equation (7.35) will then
produce the wave “record” shown in Figure 7.21, where the randomness seem to be
OK within each time interval, TR. The simulated synthetic wave is shown to repeat
itself, which is a consequence of using components evenly distributed over the
frequency range with the division ∆f. In fact the repetition time TR = 1/∆f, and vice
Figureif 4.20:
versa TR is the length ofrandom
Thirty-two a record or observation
component phasethe highest
angles frequency
between 0 andresolution
2π ra-
would be ∆f =dians.
1/TR. (See Bendat and Piersol, 1986)
t 64 t 128
which will produce a discrete spectrum of the same shape as the original PM
Elevation
spectrum, Fig. 4.19. The choice can be criticised because it does not produce
a realistic amplitude spectrum as the one in Fig. 4.7, and consequently the ran-
domness of the resulting synthesised record will be too small. Instead random
amplitudes from a Gaussian distribution withTime mean, ai , and standard deviation
2 1/2
(0.5ai ) should be used according to Tucker et. al [48].
Fig. 7.21
The assigned Simulatedand
amplitudes water surface
phase elevation.
angles Note
inserted theEq.
into repeated pattern.
(4.36) will then
produce the wave “record” shown in Fig. 4.21, where the randomness seem to
be OK within each time interval, TR . The simulated synthetic wave is shown to
repeat itself, which is a consequence of using components evenly distributed over
the frequency range with the division ∆ω. In fact the repetition time TR = 1/∆ f ,
63
The assigned amplitudes and phase angles inserted into equation (7.35) will then
produce the wave “record” shown in Figure 7.21, where the randomness seem to be
OK within each time interval, TR. The simulated synthetic wave is shown to repeat
itself, which is a consequence of using components evenly distributed over the
4. WIND range with the division ∆f. In fact the repetition time TR = 1/∆f, and vice
WAVES
frequency
versa if TR is the length of a record or observation the highest frequency resolution
would be ∆f = 1/TR. (See Bendat and Piersol, 1986)
t 64 t 128
Elevation
Time
Fig. 7.21 Simulated water surface elevation. Note the repeated pattern.
Figure 4.21: Simulated water surface elevation. Note the repeated pattern.
and vice versa if TR is the length of a record or observation the highest frequency
resolution would be ∆ω = 2π∆ f = 2π/TR . (See Bendat and Piersol, 1986)
with
K
ai = ∑ ai j . (4.39)
j=1
where θ j is the angle between the x-axis and the direction of propagation of the
component.
The variance of the directional waves are distributed according to some spread-
ing function, D(θ ), fulfilling the demand that the integral around the horizon
Z 2π
D(θ )dθ = 1 , (4.40)
0
Usually the spread is restricted to +/- 90◦ around the main direction of propaga-
tion.
The simplest spreading functions are independent of the wave frequency, and
a commonly used one is the cosine square distribution:
(
2 2 π
D(θ ) = π cos (θ ) , if |θ | < 2 , (4.42)
0, otherwise .
64
4.4. Synthesised Waves
65
5
WAVE STATISTICS AND PROBABILITY
Up till now we have only discussed how to describe sea states and the statistical
properties of the water surface elevation in these. This is often called short-term
statistics. However, to be able to design structures and ships, we need know the
probability of the appearance of sea states of different wave heights and periods
or the probability of extreme individual waves during the expected life time of
the structures. This is often called long-term statistics. Founded on the long-term
statistics, the probability of extreme loads on fixed stiff structures, and extreme re-
sponses of flexible structures or floating bodies. For a flexible or floating structure
it is not certain that the “worst sea state” or the “worst wave” will give the largest
stress or motion. The response probabilities are also of interest for estimation of
fatigue, downtime of floating offshore production or drilling platforms, frequency
of green water, slamming, accelerations etc. for ships.
A basic term used in design codes and guidelines is the return period or recur-
rence interval. It is an estimate of the mean interval of time between events like
an earthquake, flood or river discharge flow of a certain intensity or size. It is a
statistical measurement denoting the average recurrence interval over an extended
period of time, and is usually required for risk analysis (i.e. whether a project
should be allowed to go forward in a zone of a certain risk) and also to dimension
structures so that they are capable of withstanding an event of a certain return
period (with its associated intensity of a design quantity).
In the guidelines concerning loads and load effects on load-bearing structures
in the petroleum activity in Norwegian waters [39] it is stated that one should
use a design sea state given by a JONSWAP spectrum with one-hundred-year-
100 yr 100 yr
return significant wave height, Hm0 , combined with the peak period, Tp ,
or other spectra with larger probability and other shape if such spectra give larger
load effects. There is also an option to use a one-hundred-year design wave,
100 yr 100 yr 100 yr 100 yr
Hmax = 1.9Hm0 combined with the period Tmax = 0.92Tp . For float-
ing structures, especially, the wave period should be varied in order to investigate
67
5. WAVE STATISTICS AND PROBABILITY
the responses for shorter waves with maximum steepness. The factor 1.9 above is
founded on the assumption that the duration of the sea state is taken as 3 h. One
can note that the 100 year maximum wave height derived in this way is a little
smaller than if the maximum probable wave height is taken as the most probable
largest of all individual waves. Compare results in Table 5.10 and Table 5.12 for
the North Atlantic.
In Section 5.1 we discuss the concept of risk. Section 5.2 is a short account of
design with safety factors versus with load and material coefficients. In Section
5.3follows short-term statistics of waves and in Section 5.5 long-term statistics.
Before we look at the limit states for design we will acquaint ourselves with the
concept of risk. A hazard is a source of danger but does not contain any likelihood
or actual impact it will have on people, environment or economics. Risk combines
both likelihood and impact, and a risk analysis tries to answer the questions:
and consequently the probability, E, that it would be exceeded at least once during
the L years:
L 1 L
E = P(at least once) = 1 − (1 − Q) = 1 − 1 − . (5.2)
TR
68
5.1. The Concept of Risk
Table 5.1: The probability, E, of an event with the return period TR (years) to be
exceeded at least once during the lifetime L (years).
TR LR
1 5 10 50 100 500 1000 5000 10000
1 1 1 1 1 1 1 1 1 1
5 0.2 0.672 0.893 1 1 1 1 1 1
10 0.1 0.41 0.651 0.995 1 1 1 1 1
50 0.02 0.096 0.183 0.636 0.867 1 1 1 1
100 0.01 0.049 0.096 0.395 0.634 0.993 1 1 1
500 0.002 0.01 0.02 0.095 0.181 0.632 0.865 1 1
1000 0.001 0.005 0.01 0.049 0.095 0.394 0.632 0.993 1
5000 2E-4 0.001 0.002 0.01 0.02 0.095 0.181 0.632 0.865
10000 1E-4 5E-4 0.001 0.005 0.01 0.049 0.095 0.393 0.632
Example 5.1
In many rules it is stated that the weighted probability per year of a
combination of design events should be 10−4 . This means a return period
of 10 000 years. One can then ask: How large is the probability, E, that it
would occur during a lifetime of 50 years?
One can invert the problem and ask: If you accept a probability, E, during a
lifetime, which return period should be used? This can be solved from Eq. (5.2):
1
TR = . (5.3)
1 − (1 − E)1/L
Example 5.2
A probability, E, of 1 % is accepted for a design event to be exceeded in
the lifetime 50 years. Which return period should be used?
For E = 0.01 = 10−2 and L = 50 Table 5.1 gives TR = 5000 years. Alter-
natively, Eq. (5.3) yields 4975 years.
69
5. WAVE STATISTICS AND PROBABILITY
Observe that the real probability of damage certainly is different, as one uses
safety factors or load and material coefficients at the design to safeguard the limi-
tation in knowledge of statistical distributions of events, loads, load effects, mate-
rial properties as well as the quality of construction work. Finally it is actually so
that damage often appears due to unforeseen types of events and human mistakes.
There are typically two approaches for designs. One is to use safety factors i.e. to
calculate the load effect, e.g. a stress, for given design loads and compare the load
effect with a given criteria, e.g. the breaking strength, divided by a safety factor
S > 1 depending of construction loads etc. The other is to use a safety format
with load and material coefficients and assign coefficients γ f > 1 to each load,
and material coefficients γm < 1 to each component. Then check that the total
load effect calculated with the load coefficients is smaller than the design capacity
calculated with the material coefficients.
Sd ≤ Rd . (5.4)
Sd = ∑ S(γ f Fk ) . (5.5)
The characteristic loads shall be established for each load category, Table 5.3, and
limit state, Table 5.4.
70
5.2. Design Approaches
Table 5.2: Safety factors According to DnV’s mooring rules for mobile offshore
units (POSMOOR [11]).
71
5. WAVE STATISTICS AND PROBABILITY
Table 5.5: Load coefficients, γ f , and their combinations in the ultimate limit
state, ULS, According to DnV’s tentative rules for fish farms.
Limit state Steel struct. Concrete struct. Chain Wire rope Synthetic rope
Concrete Reinforced
ULS 1.15 1.40 1.25 1.50 1.30 3.00 (2.50)∗
FLS 1.00 1.20 1.10
ALS 1.00 1.20 1.10 1.30 1.15 2.50 (1.30)∗
SLS 1.00 1.00 1.00
∗ Coefficients within brackets apply if there is satisfying documentation of the rope.
• Motion;
• Acceleration;
• Stress;
• Deformation.
• In the accidental limit state, ALS, all load coefficients are set to 1.0.
• The environmental loads for ULS and ALS shall have a yearly probability
of 0.01 or less.
• The fatigue limit state, FLS, shall be investigated. All load coefficients are
set to 1.0.
72
5.3. Short Term Statistics
which, however, only is defined on the interval [0 , π) and therefore p(εi ) has to
be doubled on this interval. Finally substituting Eqs. (5.10) and (5.12) into (5.9)
the frequency distribution of ζi is
1
p(ζi ) = q , −ai ≤ ζi ≤ ai . (5.13)
π a2i − ζi2
This frequency function is shown in Fig. 5.1 and its distribution function in Fig.
5.2. In the same figures also the corresponding functions of a Gaussian process
73
5. WAVE STATISTICS AND PROBABILITY
(
dnorm ζi , 0⋅ m , σ ) 0.5
1
( )
p ζi
0
(
dnorm ζi , 0⋅ m , σ ) 0.52 1 0
ζi
1 2
ai
µ i = ! ζ i p(ζ i )dζ i = 0
and its variance …(7.50)
− ai
ai
1 2
i = ! ζ i p (ζ i ) dζ i =
and itsσvariance
2 2
ai …(7.51)
− ai
2
i a
1 2
!−a ζarei pthe
74
which, ofσcourse
2
= 2
(ζ i )same
dζ i =as for
ai the time function itself. …(7.51)
i
i
2
The distribution of the water level in an irregular wave
which,
The wateroflevel
course are the same
variation around asthe
for mean
the time
can,function itself.
as stated before, be written as a sum of
cosine functions with varying frequencies, ωi, and random phase angles, εi.
The distribution of the water level in an irregular wave
5.3. Short Term Statistics
with the same mean and standard deviation are shown. One can note that the
frequency function is far from similar to the Gaussian frequency function, while
the distribution functions are more similar.
The mean of the frequency-distribution function is
Z ai
µi = ζi p(ζi )dζi = 0 , (5.14)
−ai
The water level variation around the mean can, as stated before, be written as a
sum of cosine functions with varying frequencies, ωi , and random phase angles,
εi :
ζ (t) = ∑ ζi (t) = ∑ ai cos (ωi t + εi ) . (5.16)
i i
Each component is a stochastic variable with mean, µi , and variance, σi2 . Then,
according to the central limit theorem the sum of many components, ∑i ζi (t), ap-
proaches a Gaussian process with the mean
µ = ∑ µi = 0 , (5.17)
i
where m0 is the area under the wave spectrum or 0th moment and σ the standard
deviation.
Thus the water level in an irregular wave is Gaussian distributed with the fre-
quency function
2 2
1 −ζ 1 −ζ
p(ζ ) = √ e 2σ 2 = √ e 2m0 . (5.19)
σ 2π 2πm0
75
5. WAVE STATISTICS AND PROBABILITY
Example 5.3
How large part of the time will the surface level be above 1 m around a
platform leg in a sea state with the significant wave height, Hs = 3 m?
Assuming the duration of the sea state is T and the sum of all times the
water level is above ζ = 1 m is t, the fraction of time is
Z ∞ 2 Z ∞ ζ2
t 1 −ζ2 1 − 2m
= P(ζ > 1 m ) = √ e 2σ dζ = √ e0 dζ .
T 1 σ 2π 2πm0 1
(5.20)
√
As Hs = 4 m0 and σ = Hs /4, Eq. (5.20) will give 0.0912 e.g. by help of
the standard normal distribution with (1 m)/σ = 1.33 as found in table in
Standard Mathematical Tables:
For a reasonably narrow-banded Gaussian sea state the probability density for the
level of wave crests or local maxima, ζmax , is approximately Rayleigh distributed
with the frequency function
2
ζmax − ζ2m
max
p (ζmax ) = e 0 (5.21)
m0
which has the variance of the sea surface elevation or 0th moment of the spectrum
as its sole parameter. This expression can be used for maxima of all other derived
processes too.
The probability density for the wave heights, H, is also approximately Rayleigh
distributed with the frequency function
H2
H − 8m
p (H) = e 0. (5.22)
4m0
Its distribution function or probability that the wave height H < Hq is then
Z Hq 2 H2
H H
− 8m − 8mq
P(H < Hq ) = e 0 dH = 1 − e 0 . (5.23)
0 4m0
The probability Q that the wave height should be higher than Hq is then
H2
− 8mq
Q(H > Hq ) = 1 − P(H < Hq ) = e 0 . (5.24)
76
5.4. Extremes in a Sea State
From this follows that the wave Hq that is exceeded with the probability Q is
s
1
Hq = 8m0 ln . (5.25)
Q
Example 5.4
How many waves will be greater than 5 m during 6 hours in a sea state
with the significant wave height, Hs = 3 m and mean wave period 10 s?
nq = QN = 8.35 ≈ 8
Of course one cannot expect that the number of waves larger than 5 m
should be just 8 in all such sea states, but the number should be seen as
an expected number in such sea states. Many sea states should give the
arithmetic mean around 8.4.
77
5. WAVE STATISTICS AND PROBABILITY
The wave height that is exceeded by 10, 1, 0.1 and 0.01 % of the number of
waves in a sea state can also be assessed by Eq. (5.25) and are sometimes given
in literature as references. See Table 5.7.
The maximum wave, Hmax , in a sea state is often given as the most probable
maximum wave height of a thousand waves, that is, corresponding to 0.1 % above.
In a sea state with the mean period 10 s this corresponds to a duration of 10 000 s
or 2.8 hours. In a certain sea state the largest wave can of course be higher or
lower but the mean of many measurements or simulations should be 1.86Hs .
The same methods for estimating periods, variances and extremes, that have
been used for the waves in this chapter can be applied to all derived motions,
accelerations, loads, etc in Chapter 13.
4H −2( HH )2
p(H) = e s . (5.30)
Hs2
The significant wave height varies in its turn between the recorded sea states rep-
resented by point measurements of water elevations. Often the measurements are
performed at regular intervals and recorded and analysed for a set period of time,
during which it is assumed that the sea state is stationary. Often measurements
are done for 20 minutes every three hours, but as computer capacity increases it
may be done more often, although the period of analysis should not contain large
78
5.5. Long Term Wave Statistics
changes in significant wave height or mean period invalidating the stationary as-
sumption.
79
Table 5.8: Scatter diagram from the North Atlantic, Area 16, All directions: adapted from Hogben et al. [19]. The sum column and row
column denoted Hogben are taken directly from Hogbens table, while the other sum column and row column are results of
summing the figures in the table.
Hs (m) Tz (s)
<4 4–5 5 –6 6–7 7–8 8–9 9–10 10–11 11-12 12–13 > 13 Hogben Row sum Cumulated sum
<1 0 2 13 22 14 5 1 0 0 0 0 57 57 57
1–2 0 0 11 53 78 51 18 4 1 0 0 218 216 273
2–3 0 0 4 31 77 80 44 15 4 1 0 255 256 529
3–4 0 0 1 13 45 64 47 21 6 2 0 197 199 728
4–5 0 0 0 4 21 37 34 18 7 2 0 124 123 851
5–6 0 0 0 1 9 19 20 13 6 2 0 70 70 921
5. WAVE STATISTICS AND PROBABILITY
6–7 0 0 0 1 3 9 11 8 4 1 0 37 37 958
7–8 0 0 0 0 1 4 6 5 3 1 0 20 20 978
8–9 0 0 0 0 1 2 3 3 2 1 0 11 12 990
9–10 0 0 0 0 0 1 2 1 1 0 0 6 5 955
10–11 0 0 0 0 0 0 1 1 1 0 0 3 3 998
11–12 0 0 0 0 0 0 0 1 0 0 0 2 1 999
12–13 0 0 0 0 0 0 0 1∗ 0 0 0 1 1 1000
13–14 0 0 0 0 0 0 0 0 0 0 0 1 0 1000
> 14 0 0 0 0 0 0 0 0 0 0 0 1 0 1000
Total 1003 1000
Hogben 0 3 28 124 249 271 186 90 34 10 3 998
Column sum 0 2 29 125 249 272 187 91 35 10 1000
80
∗ 1 ‰ observations are added for 12–13 m at 10–11 s to make the table sum 1000 ‰.
5.5. Long Term Wave Statistics
Wave-Induced Loads and Ship Motions
02 December 2008
Page 149
4
( )
ln Hc
ln ( − ln ( 1 −P i) )
0
( ( ( )))
ln − ln P Hs
i
4
0 0.5 1 1.5 2 2.5 3
( )
ln Hs
i
Figure 8.1 Sample probability distribution Pi and fitted Weibull distribution.
Figure 5.3: Sample probability distribution Pi and fitted Weibull distribution.
γ
Table 8.8 Sample probabilities of P( H s < x ) = 1 − P( H s > x ) = 1 − e − ( x / H ) c
<1
Significant Class No i 1Permille of 57 Permille of 57 Probability 0.057
1-2 height
wave 2observations 216observations of Hs being 0.273
273
2-3 (m)
classes 3in each class 256
below upper 529smaller than up- 0.529
3-4 4 199 728 0.728
limit of each per class limit
4-5 5 123 851 0.851
6
class
70
P(H < i∆Hs ) 0.921
921 s
5-6
<6-7
1 1 757 57
37 0.057
958 0.958
7-8
1–2 2 8216 20
273 978
0.273 0.978
8-9 9 12 990 0.990
2–3
9-10 3 10256 529
5 0.529
995 0.995
10-11
3–4 4 11199 3
728 998
0.728 0.998
11-12 12 1 999 0.999
4–5 5 123 851 0.851
12-13 13 1 1000 1
5–6
Column sum 6 70 921
1000 0.921
6–7 7 37 958 0.958
Now the fitted distribution can be used directly to assess the probability of significant
7–8
wave heights. If8the probability is20P ( H > x ) that978 0.978 wave height
an observed significant
s
8–9
is larger than or9equal to x then the
12 inverse is called
990the return period
0.990
9–10 10 5 995 0.995
1
10–11 R p = 11 , 3 998 0.998
P( H s > x )
11–12 12 1 999 0.999
12–13 13 1 1000 1
81
5. WAVE STATISTICS AND PROBABILITY
Table 5.10: Significant wave heights and maximum wave heights in the North
Atlantic for various return periods.
R(years) HsR R
Hmax
10 16.84 31.84
20 17.37 32.84
50 18.06 34.14
100 18.57 35.10
500 19.71 37.25
1000 20.18 38.25
where P(H < x) is the long term sample probability that a wave height does not
exceed x. See Jensen (2001) or Faltinsen (1990). The method described below
82
5.5. Long Term Wave Statistics
where τ is the average wave period during the return period, R, or equivalently
the number of waves are R/τ. The mean wave period derived from Table 5.8 is
Tz = 8.42 s which can be set to τ. Then the following maximum expected indi-
vidual wave height for some return periods will be given as in Table 5.12. These
maximum wave heights are a little higher than those derived from the significant
R-year return wave heights as would be expected.
Jensen (2001) derives the 20 year maximum wave to 36 m from the same data
as here. He uses a one sided normal distribution for the distribution of significant
wave heights which then gives a Gumbel distribution of individual wave heights.
• wind speed and sea state corresponding to a 100-year return period com-
bined with a 10 year-return-period current;
• current velocity and sea state with 100-year return period combined with a
wind speed with a 10-year-return period.
83
5. WAVE STATISTICS AND PROBABILITY
Table 5.11: The parameters C and D of the long term distribution of individual
wave heights Equation (5.29).
γd C D
d = 4/3 d=1
∞ 1.189 0.707 2.000
10.00 1.056 0.628 1.780
8.00 1.029 0.612 1.712
6.00 0.992 0.590 1.614
4.00 0.930 0.553 1.444
84
5.5. Long Term Wave Statistics
Table 5.12: Maximum wave heights in the North Atlantic for various return pe-
riods.
R(years) R
Hmax
10 31.89
20 33.14
50 34.80
100 36.06
500 38.97
1000 40.23
A recent treatise of the joint probability of wind and waves in the Northern
North Sea is given by Johannessen et al. (2002). In this the marginal distribu-
tion of the wind speed is taken as a two-parameter Weibull distribution. In each
wind-speed class the significant wave heights are found to follow a two-parameter
Weibull distribuition and the peak periods a log-normal distribution.
85
6
EQUATIONS OF MOTION FOR A SHIP
Three coordinate systems are normally used when the equations of motion for a
ship are defined. Unfortunately there is no universally accepted coordinate sys-
tem convention. For the seakeeping problem most authors and organizations use
x in the forward direction, y to port and z upwards, while in the manoeuvring
problem x is forward, y is to starboard and z is downwards. There are however
exceptions from this ”rule”. These Lecture Notes deal with both the seakeeping
problem and the manoeuvring problem and it is not obvious how to define the
coordinate systems. It was decided to follow the ”rule” described above and use
separate coordinate system definitions for the two problems. This is probably a
combination the students will meet if they deal with seakeeping and manoeuvring
in their professional life as Naval Architects. The three coordinate systems shown
in Fig. 6.1 follow the convention for the manouevring problem and they are here
used for the derivation of the equations of motion for a ship.
The three coordinate systems are:
1. an earth fixed coordinate system (x0 , y0 , z0 ) with the origin at any desirable
location. The z0 axis is positive downwards and the x0 , y0 plane is normally
located at the calm water level. The x0 axis is pointing in the forward di-
rection of the ship. This coordinate system is used as the reference system
when deriving the equations of motion for a ship. It is also used to define
the incident wave system.
87
6. EQUATIONS OF MOTION FOR A SHIP
Figure 6.1: Definition of the earth-fixed, the inertial and the body-fixed coordi-
nate systems
3. a third coordinate system(x̄, ȳ, z̄) is body fixed and the motions of the ship is
determined by the orientation of the body fixed coordinate system relative
to the inertial coordinate system.
The transformation from the earth fixed to the inertial coordinate system is
given by:
x0 = x +U · t (6.1)
y0 = y (6.2)
z0 = z (6.3)
Three translations
• η1 - surge is a translation in the x̄-direction, positive forward
• η4 - roll is a rotation about the x̄-axis, positive when starboard side goes in
to the water
• η5 - pitch is a rotation about the ȳ-axis, positive when the bow goes out of
the water
88
6.1. COORDINATE SYSTEMS AND DEFINITIONS
• η6 - yaw is a rotation about the z̄-axis, positive when the bow turns to star-
board
are used to define the motions of the ship, see Figs. 6.2 , 6.3 and 6.4. It must
be noted that the rotations about the x̄, ȳ and z̄-axes are not commutative. The final
89
6. EQUATIONS OF MOTION FOR A SHIP
position depends on the order in which the rotations are carried out. However, if
the rotations are infinitesimal, it can be shown that they do satisfy the commutative
law of addition. Thus the angular velocity vector can be expressed as the time rate
of change of infinitesimal rotations.
A convention is therefore used for the order of the rotations in order to obtain
the angular position of the ship with respect to the earth fixed coordinate system.
The convention is as follows:
The translation velocities are denoted η̇1 in the x̄-direction, η̇2 in the ȳ-direction
and η̇3 in the z̄-direction. The angular velocities are denoted η̇4 about the x̄-axis,
η̇5 about the ȳ-axis and η̇6 about the z̄-axis.
In an earth fixed coordinate system the equations of motion for the center of grav-
ity (xG0 , yG0 , zG0 ) of a rigid body are
90
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
The ship motion relative to the earth fixed coordinate system is normally large
which means that the forces, moments, moments of inertia and the products of
inertia vary with time. They are therefore difficult to evaluate. A body fixed
coordinate system where the moments of inertia and the products of inertia are
constant is therefore introduced to avoid these difficulties.
The price we have to pay is that the equations of motions must also be trans-
ferred to the body fixed coordinate system which make them more complicated.
After solving the equations of motion in the body fixed coordinate system we
91
6. EQUATIONS OF MOTION FOR A SHIP
get information about the motion of the origin of this system and about the rota-
tion about the axes of the body fixed coordinate system. The origin of the body
fixed coordinate system is normally located in the center plane at Lpp/2 at the
undisturbed free surface level. The position of the center of gravity is denoted
(xG , yG , zG ).
6.3.1 Translation
First assume that the ship performs translations in surge, sway and heave only
while the rotations are all zero. A small mass element dm at (x̄, ȳ, z̄) will then
have the same velocities η̇1 , η̇2 , η̇3 in the x̄, ȳ, z̄ directions as the origin of the body
fixed coordinate system, see Fig. 6.6. The angle between the axes of the (x̄, ȳ, z̄)
and the (x, y, z) systems do not change with time.
• rη̇6
is added to the velocity of the small mass element dm. The added velocity is
perpendicular to the position vector of the mass element dm. The components of
92
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
the velocity due to the angular velocity are then added to the translation velocities
η̇1 , η̇2 , η̇3 in the body fixed coordinate system
• rη̇5
The added velocity is perpendicular to the position vector of the mass element
dm. The components of the velocity due to the angular velocity η̇5 are then added
to the translation velocities and the velocity contributions from yaw in the body
fixed coordinate system.
93
6. EQUATIONS OF MOTION FOR A SHIP
94
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
• rη̇4
The added velocity is perpendicular to the position vector of the mass element
dm. The components of the velocity due to the angular velocity η̇4 are then added
to the translation velocities and to the velocity contributions from yaw and pitch in
the body fixed coordinate system. We can now write the total velocity components
for motion in six degrees of freedom in the x̄, ȳ and z̄ directions respectively:
(η̇1T OT +d η̇1T OT ) cos dη6 − η̇1T OT −(η̇2T OT +d η̇2T OT ) sin dη6 = d η̇1T OT − η̇2T OT dη6
(6.13)
95
6. EQUATIONS OF MOTION FOR A SHIP
96
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
(η̇2T OT +d η̇2T OT ) cos dη6 − η̇2T OT +(η̇1T OT +d η̇1T OT ) sin dη6 = d η̇2T OT + η̇1T OT dη6
(6.15)
97
6. EQUATIONS OF MOTION FOR A SHIP
(η̇1T OT +d η̇1T OT ) cos dη5 − η̇1T OT +(η̇3T OT +d η̇3T OT ) sin dη6 = d η̇1T OT + η̇3T OT dη5
(6.17)
(η̇3T OT +d η̇3T OT ) cos dη5 − η̇3T OT −(η̇1T OT +d η̇1T OT ) sin dη5 = d η̇3T OT − η̇1T OT dη5
(6.19)
(η̇2T OT +d η̇2T OT ) cos dη4 − η̇2T OT −(η̇3T OT +d η̇3T OT ) sin dη6 = d η̇2T OT − η̇3T OT dη4
(6.21)
98
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
(η̇3T OT +d η̇3T OT ) cos dη4 − η̇3T OT +(η̇2T OT +d η̇2T OT ) sin dη4 = d η̇3T OT + η̇2T OT dη4
(6.23)
99
6. EQUATIONS OF MOTION FOR A SHIP
Z
F1 = η̈1 dm =
Z Z LZ
Z
F2 = η̈2 dm =
Z Z LZ
Z
F3 = η̈3 dm =
Z Z LZ
100
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
Z
z̄dm = mz̄G (6.37)
L
The contributions to the moments from a small mass element dm about the z̄,
ȳ and x̄-axes are shown in Fig. 6.13, Fig. 6.14 and Fig. 6.15 respectively. The
contributions are
101
6. EQUATIONS OF MOTION FOR A SHIP
102
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
Z
F4 = (ȳη̈3 − z̄η̈2 )dm =
Z Z L Z
η̈3 ȳdm − η̈5 x̄ȳdm + η̈4 ȳ2 dm −
Z L ZL ZL
2
η̇5 η̇1 ȳdm + η̇5 η̇6 ȳ dm − η̇52 ȳz̄dm +
ZL ZL ZL
η̇4 η̇2 ȳdm + η̇4 η̇6 x̄ȳdm − η̇42 ȳz̄dm −
L Z ZL ZL
η̈2 z̄dm − η̈6 x̄z̄dm + η̈4 z̄2 dm −
Z L Z L ZL
η̇6 η̇1 z̄dm + η̇62 ȳz̄dm − η̇6 η̇5 z̄2 dm +
LZ L Z LZ
103
6. EQUATIONS OF MOTION FOR A SHIP
Z
F5 = (z̄η̈1 − x̄η̈3 )dm =
Z Z L Z
η̈1 z̄dm − η̈6 ȳz̄dm + η̈5 z̄2 dm −
Z L Z L ZL
η̇6 η̇2 z̄dm − η̇62 x̄z̄dm + η̇6 η̇4 z̄2 dm +
ZL ZL ZL
η̇5 η̇3 z̄dm − η̇52 x̄z̄dm + η̇5 η̇4 ȳz̄dm −
LZ LZ ZL
η̈3 x̄dm + η̈5 x̄2 dm − η̈4 x̄ȳdm +
Z L ZL ZL
η̇5 η̇1 x̄dm − η̇5 η̇6 x̄ȳdm + η̇52 x̄z̄dm −
LZ LZ LZ
Z
F6 = (x̄η̈2 − ȳη̈1 )dm =
Z Z L Z
2
η̈2 x̄dm + η̈6 x̄ dm − η̈4 x̄z̄dm +
Z L Z L ZL
η̇6 η̇1 x̄dm − η̇62 x̄ȳdm + η̇6 η̇4 x̄z̄dm −
ZL LZ ZL
η̇4 η̇3 x̄dm + η̇5 η̇4 x̄2 dm − η̇42 x̄ȳdm −
LZ ZL ZL
η̈1 ȳdm + η̈6 ȳ2 dm − η̈5 ȳz̄dm +
Z L Z L ZL
η̇6 η̇2 ȳdm + η̇62 x̄ȳdm − η̇6 η̇4 ȳz̄dm −
LZ LZ LZ
104
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
Z Z Z Z Z Z
ȳ2 dm − z̄2 dm = ȳ2 dm + x̄2 dm − z̄2 dm − x̄2 dm =
L L L L L L
(6.51)
Z Z
(x̄2 + ȳ2 )dm − (x̄2 + z̄2 )dm = Iz̄ − Iȳ (6.52)
L L
F4 = Ix̄ η̈4 + (Iz̄ − Iȳ )η̇6 η̇5 − Ix̄ȳ (η̈5 − η̇6 η̇4 ) − Iȳz̄ (η̇52 − η̇62 ) − Ix̄z̄ (η̈6 + η̇4 η̇5 ) +
(6.53)
m[ȳG (η̈3 − η̇5 η̇1 + η̇4 η̇2 ) − z̄G (η̈2 + η̇6 η̇1 − η̇4 η̇3 )]
(6.54)
F5 = Iȳ η̈5 + (Ix̄ − Iz̄ )η̇6 η̇4 − Iȳz̄ (η̈6 − η̇5 η̇4 ) − Ix̄z̄ (η̇62 − η̇42 ) − Ix̄ȳ (η̈4 + η̇5 η̇6 ) +
(6.55)
m[z̄G (η̈1 − η̇6 η̇2 + η̇5 η̇3 ) − x̄G (η̈3 − η̇5 η̇1 + η̇4 η̇2 )]
(6.56)
F6 = Iz̄ η̈6 + (Iȳ − Ix̄ )η̇5 η̇4 − Ix̄z̄ (η̈4 − η̇6 η̇5 ) − Ix̄ȳ (η̇42 − η̇52 ) − Iȳz̄ (η̈5 + η̇6 η̇4 ) +
(6.57)
m[x̄G (η̈2 + η̇5 η̇1 − η̇4 η̇3 ) − ȳG (η̈1 − η̇6 η̇2 + η̇5 η̇3 )]
(6.58)
105
6. EQUATIONS OF MOTION FOR A SHIP
F4 = Ix̄ η̈4 + (Iz̄ − Iȳ )η̇6 η̇5 − Ix̄ȳ (η̈5 − η̇6 η̇4 ) − Iȳz̄ (η̇52 − η̇62 ) − Ix̄z̄ (η̈6 + η̇4 η̇5 ) +
(6.65)
m[ȳG (η̈3 − η̇5 η̇1 + η̇4 η̇2 ) − z̄G (η̈2 + η̇6 η̇1 − η̇4 η̇3 )]
(6.66)
F5 = Iȳ η̈5 + (Ix̄ − Iz̄ )η̇6 η̇4 − Iȳz̄ (η̈6 − η̇5 η̇4 ) − Ix̄z̄ (η̇62 − η̇42 ) − Ix̄ȳ (η̈4 + η̇5 η̇6 ) +
(6.67)
m[z̄G (η̈1 − η̇6 η̇2 + η̇5 η̇3 ) − x̄G (η̈3 − η̇5 η̇1 + η̇4 η̇2 )]
(6.68)
F6 = Iz̄ η̈6 + (Iȳ − Ix̄ )η̇5 η̇4 − Ix̄z̄ (η̈4 − η̇6 η̇5 ) − Ix̄ȳ (η̇42 − η̇52 ) − Iȳz̄ (η̈5 + η̇6 η̇4 ) +
(6.69)
m[x̄G (η̈2 + η̇6 η̇1 − η̇4 η̇3 ) − ȳG (η̈1 − η̇6 η̇2 + η̇5 η̇3 )]
(6.70)
6.3.12 Equations of motion when the origin of the coordinate system is at center
of gravity
Assume that the origin of the coordinate system is located at the center of gravity
of the ship. All terms containing x̄G , ȳG and z̄G are then zero.
F4 = Ix̄ η̈4 + (Iz̄ − Iȳ )η̇6 η̇5 − Ix̄ȳ (η̈5 − η̇6 η̇4 ) − Iȳz̄ (η̇52 − η̇62 ) − Ix̄z̄ (η̈6 + η̇4 η̇5 )]
(6.74)
106
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
F5 = Iȳ η̈5 + (Ix̄ − Iz̄ )η̇6 η̇4 − Iȳz̄ (η̈6 − η̇5 η̇4 ) − Ix̄z̄ (η̇62 − η̇42 ) − Ix̄ȳ (η̈4 + η̇5 η̇6 )]
(6.75)
F6 = Iz̄ η̈6 + (Iȳ − Ix̄ )η̇5 η̇4 − Ix̄z̄ (η̈4 − η̇6 η̇5 ) − Ix̄ȳ (η̇42 − η̇52 ) − Iȳz̄ (η̈5 + η̇6 η̇4 )]
(6.76)
F1 = m[(η̈1 − η̇6 η̇2 + η̇5 η̇3 ) − x̄G (η̇62 + η̇52 ) + z̄G (η̈5 + η̇6 η̇4 )] (6.77)
F2 = m[(η̈2 + η̇6 η̇1 − η̇4 η̇3 ) + x̄G (η̈6 + η̇4 η̇5 ) − z̄G (η̈4 − η̇6 η̇5 )] (6.78)
F3 = m[(η̈3 − η̇5 η̇1 + η̇4 η̇2 ) − x̄G (η̈5 − η̇4 η̇6 ) − z̄G (η̇52 + η̇42 )] (6.79)
F4 = Ix̄ η̈4 + (Iz̄ − Iȳ )η̇6 η̇5 − Ix̄z̄ (η̈6 + η̇4 η̇5 ) +
(6.80)
m[−z̄G (η̈2 + η̇6 η̇1 − η̇4 η̇3 )] (6.81)
F6 = Iz̄ η̈6 + (Iȳ − Ix̄ )η̇5 η̇4 − Ix̄z̄ (η̈4 − η̇6 η̇5 ) +
(6.84)
m[x̄G (η̈2 + η̇6 η̇1 − η̇4 η̇3 )] (6.85)
107
6. EQUATIONS OF MOTION FOR A SHIP
F1 = m[(η̈1 − η̇6 η̇2 + η̇5 η̇3 ) − x̄G (η̇62 + η̇52 ) + z̄G (η̈5 + η̇6 η̇4 )] (6.86)
F2 = m[(η̈2 + η̇6 η̇1 − η̇4 η̇3 ) + x̄G (η̈6 + η̇4 η̇5 ) − z̄G (η̈4 − η̇6 η̇5 )] (6.87)
F3 = m[(η̈3 − η̇5 η̇1 + η̇4 η̇2 ) − x̄G (η̈5 − η̇4 η̇6 ) − z̄G (η̇52 + η̇42 )] (6.88)
108
6.3. SHIP MOTION IN A BODY FIXED COORDINATE SYSTEM
F4 = Ix̄ η̈4 + (Iz̄ − Iȳ )η̇6 η̇5 − Ix̄z̄ (η̈6 + η̇4 η̇5 ) (6.98)
F5 = Iȳ η̈5 + (Ix̄ − Iz̄ )η̇6 η̇4 − Ix̄z̄ (η̇62 − η̇42 ) (6.99)
F6 = Iz̄ η̈6 + (Iȳ − Ix̄ )η̇5 η̇4 − Ix̄z̄ (η̈4 − η̇6 η̇5 ) (6.100)
These equations are often used in the analysis of ship motion, but it is then
important to be aware of the assumptions included and check that the assumptions
are not violated.
109
7
MOTION RESPONSE TO LOADING
The loads on a platform or ship can be constant in time, transient – i.e. of short
duration – or harmonic. Irregular or random loads from e.g. sea waves can to a
first, linear, approximation be treated as a superposition of harmonic loads. The
responses are fundamentally different for the three types of loads. To clearly
illustrate this we will in this chapter use a simple, one-degree-of-freedom system
as shown in Figure 7.1. The equation of motion for this system can be written
Figure 7.1: A mechanical system with one degree of freedom, mass, m, added
mass, a, damping coefficient, b, and spring stiffness, c.
111
7. MOTION RESPONSE TO LOADING
(m + a) ẍ + bẋ + cx = 0 , (7.2)
x = Ceκt , (7.3)
κ 2 + 2ξ ωN κ + ωN2 = 0 , (7.4)
p
where ωN = c/(m + a) is the “natural”
p angular frequency, that is, the undamped
angular frequency and ξ = b/(2 c(m + a)) is the damping factor.
The roots of Eq. (7.4) are
q
κ1,2 = −ξ ωN ± ωN ξ 2 − 1 . (7.5)
These roots are complex, zero or real depending on the value of ξ . The damping
factor can thus be used to distinguish between three cases:
• underdamped (0 ≤ ξ < 1)
• critically damped (ξ = 1)
• overdamped (ξ > 1)
See Fig. 7.2 for the motion of a body released from the position x(0) = 1 m at
t = 0 s. The underdamped case displays an attenuating vertical oscillation, while
the other cases display motions monotonously approaching the vertical equilib-
rium position. A floating ship in heave, roll and pitch would normally display
underdamped characteristics with a damping factor of the order of 10−1 in heave
and pitch and of the order of 10−2 in roll. In the horizontal degrees of freedom
there are no stiffnesses so the horizontal motions do not exhibit resonant charac-
teristics.
The damping factor is often called the damping ratio, as it is equal to the ratio
between
p the current damping coefficient b and the critical damping coefficient
2 c(m + a).
112
7.2. Response to Constant Loads
Figure 7.2: Response of a damped SDOF system with various damping ra-
tios. Dashed line: overdamped case (ξ = 1.5), solid line: critically
damped case (ξ = 1) and dotted line: underdamped case (ξ = 0.1).
A constant or static load Fo (see Fig. 7.3) acting on the system in Fig. 7.1 gives
as response a displacement to a static equilibrium position, x = xo , because the
equation of motion, Eq. (7.1), gives cx = Fo ⇒ x = Fo /c = xo if Fo is constant, as
ẍ and ẋ must be identically zero.
In dynamic problems one usually calculates the equilibrium position for con-
stant, static loads first, and thereafter, new co-ordinates are defined from this static
equilibrium position. Static loads can for instance be the weight of a ship in calm
water balanced by the equally constant buoyancy to yield a specific draught, or
wind and current forces acting on a moored platform giving a constant offset bal-
anced by the mooring arrangement.
113
Wave-Induced Loads and Ship Motions
7. MOTION RESPONSE TO LOADING
02 December 2008
Page 14
Kraft, förskjutning
Force F ( t )
x ( t ) 4 2 0 2 4
Displacement
t
Time
Tid
We can solve Equation (3.1) for the given harmonic load, Equation (3.6), simply by
substituting the particular solution Equation (3.7) into it
7.3 R ESPONSE TO H ARMONIC L OADS
(m + a ) !x! + bx! + cx = Fo cos (ωt)
A harmonic load
x = x" cos(ωt − ε )F(t) = Fo cos(ωt) , (7.6)
x! = −ωx" sin(ωt − ε )
where Fo is the load amplitude, ω = 2π/T is the angular frequency and T is the
!!
x = −ω 2
x" cos(ωt − ε )
time period. A harmonic load, as from regular waves for instance, gives a response
of the same harmonic type:
The substitution gives
x(t) = x̂ cos(ωt − ε) , (7.7)
(c − (m + a)ω 2 ) xˆ cos(ωt − ε ) − bωxˆ sin(ωt − ε ) = Fo cos(ωt )
in which x̂ is the amplitude of displacement and ε denotes the phase lag between
Using the trigonometric expressions for sine and cosine of angle differences then
the force and displacement. The motion x(t) is the stationary response to the
yields:
harmonic load and is the particular solution to Eq. (7.1) with the right hand side
F(t) given by Eq. (7.6), see 2Fig. 7.4.
(c − (m + a )ω ) xˆ (cos(ωt ) cos(ε ) + sin(ωt ) sin(ε )) −
We can solve Eq. (7.1) for the given harmonic load, Eq. (7.6), simply by
− bωxˆ (sin(ωt ) cos(ε ) − (cos(ωt ) sin(ε )) = F cos(ωt )
substituting the particular solution, Eq. (7.7), into it o
Identification and(m
separation
+ a)ẍ + bofẋ +
terms
cx =with sin(ωt) ,and cos(ωt) gives
F cos(ωt) o
114
7.3. Response to Harmonic Loads
Using the trigonometric expressions for sine and cosine of angle differences then
yields:
c − (m + a) ω 2 x̂ (cos(ω t) cos(ε) + sin(ω t) sin(ε))
−b ω x̂ (sin(ω t) cos(ε) − cos(ω t) sin(ε)) = Fo cos(ω t) .
Identification and separation of terms with sin(ω t) and cos(ω t) gives
c − (m + a)ω 2 x̂ sin(ε) − b ω x̂ cos(ε) = 0 , (7.8)
and
c − (m + a)ω 2 x̂ cos(ε) + b ω x̂ sin(ε) = Fo . (7.9)
From Eq. (7.8) follows directly
bω
tan(ε) = , (7.10)
c − (m + a)ω 2
i.e.
bω
ε = arctan , (7.11)
c − (m + a)ω 2
where also the correct quadrant must be decided.
Squaring and adding Eqs. (7.8) and (7.9) gives after using the trigonometric
unity
2
2 2
x̂ c − (m + a)ω + b ω = Fo2 ,
2 2
(7.12)
and finally, as the amplitude, x̂, by definition is positive,
Fo
x̂ = q . (7.13)
2 2 2 2
(c − (m + a)ω ) + b ω
115
7. MOTION RESPONSE TO LOADING
This amplification factor is illustrated in Fig. 7.5 together with the corresponding
phase lag ε. In the figure the frequency axis is nondimensionalised by the natural
frequency, N = ωN /(2π), which by definition is the frequency of the undamped
eigenvibration. Thus the abscissa, frequency axis, in Fig. 7.5 is scaled in the
nondimensionalised frequency, which reads
f ω
Ω= = . (7.16)
N ωN
Equation (7.15) can now be written
c
Y (Ω) = q
2
(c − (m + a)(ωN Ω)2 ) + b2 (ωN Ω)2
1
=q , (7.17)
2 2 b2 2
(1 − Ω ) + c2 (ωN Ω)
and finally introducing also the damping ratio defined in section 7.1
1
Y (Ω , ξ ) = p . (7.18)
(1 − Ω2 )2 + 4(ξ Ω)2
because
eı ω t = cos(ω t) + ı sin(ω t) . (7.21)
Observe that when analysing performed work, power etc. the real valued quan-
tity, F(t), must be used. Sticking to complex notation then the sum of, Fc (t), and
116
7.3. Response to Harmonic Loads
(a) (b)
Figure 7.5: (a) Amplification factor, Y (Ω , ξ ), and (b) phase lag, ε(Ω , ξ ), for
a system with one degree of freedom. Ω is the non-dimensional
frequency and ξ is the damping ratio.
its complex conjugate can be used, which sum equals 2F(t). Here we will use
Re(Fc (t)) when necessary.
The complex motion, xc (t), is then set to
117
7. MOTION RESPONSE TO LOADING
which includes both the amplification factor, Y (ω), and the phase lag, ε(ω). The
amplification factor is given by the modulus of Yc (ω), i.e. Y (ω) = |Yc (ω)|, and
the phase lag by its argument, i.e. ε(ω) = arg (Yc (ω)).
Exercise 7.1
Confirm that this gives the same result as Eqs. (7.15) and (7.11).
1
xc (t) = Fc (t) = Tc (ω)Fc (t) , (7.32)
c − (m + a)ω 2 + ı b ω
i.e. the motion is for any regular frequency of excitation given by a complex
multiplication of a transfer function, Tc (ω), and the driving force, Fc (t).
A comparison with Eq. (7.14) reveals that
and thus the amplitude response function equals the modulus of the complex trans-
fer function and the phase lag equals ε(ω) = arg(Tc ), i.e. arctan(Im(Tc )/Re(Tc ))
with additional information of into which quadrant Tc (ω) points.
118
7.3. Response to Harmonic Loads
Figure 7.7: Comparison in the time domain between the transient motion of
Fig. 7.6 and steady-state oscillation for the same load amplitude.
119
7. MOTION RESPONSE TO LOADING
The motion in six degrees of freedom of a ship at zero speed can be described
by the following six coupled equations of motion, which looks the same as Eq.
(9.13):
(M + A)η̈ + Bη̇ + Cη = F . (7.34)
Here M, A, B and C are 6 × 6 matrices, η a 6 × 1 vector with the six motions and
F a 6 × 1 vector with the three forces and the three moments. See Fig. 9.1. The
forces and moments, F, are functions of time which implies that the motions, η,
also are so.
If the coefficients, the elements of the matrices, in this equation all are con-
stants, it can be solved directly in the time domain for arbitrary loads by methods
equivalent to those in Paragraph 3.3. For large motions the elements of damping
matrix, B, is however depending on the quadratic drag forces and the velocity am-
plitudes relative to the water motion. For floating bodies the matrices A and B are
functions of the frequency of the motion and Eq. (7.34) can only be solved for
one harmonic motion at a time.
120
7.4. System with Six Degrees of Freedom
which makes the total response. This method requires that A and B are functions
of only the frequency of the wave and not the wave amplitude.
In the equation of motion the motion, η(t), is simplest written as a complex
position vector, η c (t), with the elements (i = 1 , 2 , . . . , 6):
121
8
HEADING AND
ENCOUNTER FREQUENCY
8.1 HEADING
The heading is defined with respect to the propagation direction of the waves as
shown in Fig. 8.1a. The angle µ between the intended track of the ship and the
direction of wave propagation is denoted the heading angle or heading.
Fig. 8.1b shows some common definitions used when defining a ships heading.
• µ = 0◦ Following waves with the waves and the ship traveling in the same
direction
• µ = 90◦ Beam waves with the waves coming from the starboard side.
• µ = 180◦ Head waves with the waves traveling in the opposite direction of
the ship.
• µ = 270◦ Beam waves with waves coming from the port side.
Encounter frequency is one of the most important parameters for a ships behavior
in waves. Wave disturbances at or close to a ships resonance frequency can cause
large motion amplitudes even at moderate sea states. It is therefore important to
know how the frequency experienced by the ship ωe is related to the ships speed
and heading.
123
8. HEADING AND
ENCOUNTER FREQUENCY
(a) (b)
Figure 8.1: (a) Definition of the heading angle µ and (b) names of common
heading directions.
The component of the ships velocity in the direction of the propagating wave
is
U cos(µ) (8.1)
The waves will overtake the ship with the relative velocity
c −U cos(µ) . (8.2)
A ship progressing through a following wave with the speed, U, at the heading,
µ, in relation to the direction of wave propagation, will be passed by a wave crest
every encounter period
λ
Te = . (8.3)
c −U cos(µ)
The angular frequency of encounter is thus
2π 2π
ωe = = (C −U cos(µ)) , (8.4)
Te λ
2π g
Since T = λc ;ω = T ;ω = c for deep water Eq. (11.73) can be written
ω2U
ωe = ω − cos(µ) . (8.5)
g
124
8.2. ENCOUNTER FREQUENCY
In forward seas (90◦ < µ < 270◦ ) cos(µ) is always negative and the encounter
frequency is greater than the wave frequency. In beam sea (µ = 90◦ or 270◦ )
cos(µ) = 0 and the encounter frequency is equal to the wave frequency and is
not influenced by the by the ship speed.
An important but much more complicated situation occurs for (0◦ < µ < 90◦ )
or (270◦ < µ < 360◦ ). cos(µ) is then always positive and the encounter frequency
now has a maximum value
g
ωe max = (8.6)
4U cos(µ)
U cos(µ) = c (8.9)
since ω = g/c .
The encounter frequency is negative for higher values of ω which means that
the ship is overtaking the waves. Positive encounter frequency means that the
waves are overtaking the ship.
In following and quartering waves a given (absolute) value of encounter fre-
quency may be experienced in three different wave systems if | ωe |< ωe max as
shown in Fig. 8.2. Two of these wave systems will give positive encounter fre-
quencies and the third will give a negative encounter frequency. If | ωe |> ωe max
there will be one negative encounter frequency only.
The physical interpretation of this is that long waves having a high celerity
will give the same encounter frequency as shorter waves with lower celerity when
| ωe |< ωe max . Very short waves having a celerity lower than the ship speed will
be overtaken by the ship resulting in a negative encounter frequency.
In head or bow waves there will be only one encounter frequency for each
wavelength while following and quartering waves may allow for up to three dif-
ferent wave systems to give the same encounter frequency. The relation between
ship speed, heading, wavelength and encounter frequency is shown in Fig. 8.3.
For regular waves a wide range of wavelengths (100-1000m) will give nearly the
same encounter frequency for 0.25 < ωe < 0.15. For this encounter frequency
range a large number of wave components may contribute to excitations of ship
motions at the same frequency.
125
8. HEADING AND
ENCOUNTER FREQUENCY
126
8.2. ENCOUNTER FREQUENCY
Figure 8.3: Relation between encounter frequency, ship speed and wave length.
127
9
THE SEAKEEPING PROBLEM
For hydrodynamic purposes a floating body can mostly be regarded as rigid but
moving. It then exhibits six motional degrees of freedom. See Fig. 9.1, where
also the co-ordinate system used for the seakeeping problem is shown. Such a
space-fixed, right-handed co-ordinate system is usually oriented with respect to
the position of the body in rest. Its origin is either placed in the centre of gravity
of the body or in the still water surface vertically above the centre of buoyancy as
is chosen here. Usually the z-axis is vertical and points upwards, the x-axis points
horizontally forward, as in Fig. 9.1.
The six oscillating motions of a floating body have established names in En-
129
9. THE SEAKEEPING PROBLEM
glish but lack this in some other languages, see Fig. 9.1 and Table 9.1.
Symbolically the equation of motion of a floating body can be written
M η̈ = F , (9.1)
In Chapter 6 the equations of motion for a rigid ship were derived for dif-
ferent levels of approximation regarding symmetry and position of the coordinate
system.
NOTE that the equations in Chapter 6 are derived for the manoeuvring prob-
lem with the z-axis pointing downwards. But the rigid body equations will be the
same also for z upwards.
The most general equations of motion Eq. 6.60 to Eq. 6.70 are nonlinear since
they contain products of the unknown degrees of freedom. It is therefore difficult
to solve the equations. A first step towards a practical method to solve the sea-
keeping problem is to linearize the equations. It is then assumed that the motions
are small and that products of the unknowns can be neglected. The linearized
equations are:
130
Figure 9.2: Superposition of wave excitation (left); added mass, radiation damp-
ing and restoring loads (middle); to the total hydrodynamic problem
(right). From Faltinsen [12].
It can be noted that heave and pitch are coupled through Equations 9.8 and
9.10 while the equation 9.9 for roll is uncoupled.
The seakeeping problem for ship motion in waves can be solved by introduc-
ing:
1. the forces on the body when this is forced to oscillate with an arbitrary
amplitude in calm water,
2. the forces on the body when this is fixed in the incident waves,
131
9. THE SEAKEEPING PROBLEM
The force vector F in Equation (9.1) can be split into the exciting forces, the
reaction forces from the water and from moorings if any. Neglecting, for the time
being, other exciting forces than the wave excited forces a convenient split of the
forces are:
F = F e + F r + F rs , (9.11)
in which F e contains the wave-excited forces on the fixed structure, F r denotes
the hydrodynamic reaction forces from the water on the moving body in the ab-
sence of the waves and F rs are the reaction forces from the mooring system.
Substituting Eq. (9.11) and Eq. (9.12) into Eq. (9.1) the following simple expres-
sion results:
(M + A) η̈ + +Bη̇ − Cη = F e + F rs . (9.13)
NOTE that the direction of the z-axis is now upwards. The direction of the z-
axis will have an influence for the sign of the wave excitation, hydrodynamic and
mooring forces. This sign convention will be used for the seakeeping problem.
Linear reaction forces from e.g. the mooring system can be included in the
coefficient matrices A, B and C above, while non-linear reaction forces must be
included in the right-hand side of the equation of motion. Additionally, for bodies
floating in the water surface or bodies being submerged but close to the water
surface, the hydrodynamic properties A and B are functions of the frequency of
the motion, which is caused by the generation of waves around the body. Then
Eq. (9.13) can be easily solved only for cases when the excitation is a harmonic
function.
132
10
SOLUTION METHODS -
STRIP THEORY
The methods used for practical design work today are based on the potential flow
assumption, but in addition it is assumed that the ship can be represented by a
number of 2D strips, see Fig. 10.1. The 3D effects are thus assumed to be small
and are therefore neglected. This assumption implies that:
133
10. SOLUTION METHODS -
STRIP THEORY
• Deep water
With this assumption the coefficients for added mass, damping and restoring
force can be determined by simple 2D potential flow solutions for each strip. Al-
ternatively the coefficients can be computed based on analytical solutions of each
strip after coordinate transformation to a circular 2D cylinder, see Section 10.3.
The coefficients can also be determined by experiments for 2D cylinder having
the shape of the hull segments. A summation over the strips will then give the
properties for the hull.
Experience shows that the strip methods in many cases can be used for practi-
cal applications even if the assumptions are violated. The strip methods can also
be extended to take the nonlinear effects of the wavy intersection between the hull
and the free surface into account when the pressure forces are computed. These
methods are referred to as ”Nonlinear Strip Methods”.
A widely used method to determine the sectional added mass and damping prop-
erties for heave, sway and roll is via Lewis transforms. This method is based
on analytical solutions of added mass and damping for ship-like sectional forms
134
10.3. Lewis forms
known as Lewis forms. The Lewis forms are controlled by two parameters: the
Beam/Draught ratio H and the sectional area coefficient σ . Typical Lewis forms
are shown in Fig. 10.2 for 0.5 < H < 6 and 0.5 < σ < 0.9.
The analytical expressions for the Lewis forms are:
B 2(1 + a1 + a3 )
H= = (10.1)
D 1 − a1 + a3
A π 1 − a21 − 3a23
σ= = (10.2)
BD 4 1 − a21 + 2a3 + a23
where
H −2
a1 = (1 + a3 ) (10.3)
H +2
√
3 −C + 9 − 2C
a3 = (10.4)
C
135
10. SOLUTION METHODS -
STRIP THEORY
4σ 4σ H −2 2
C = 3+ + 1− (10.5)
π π H +2
In addition there is a relation between the Beam/Draught ratio H and the sec-
tional area coefficient σ that must be satisfied:
π
σ< (H 2 + 20H + 4) (10.6)
64H
Examples of invalid Lewis forms are shown in Fig. 10.3
The sectional 2D added mass and damping properties for heave, sway, and roll
are available through analytical expressions for valid forms. For details see [30].
Comparison between results obtained from Lewis forms and by experiments are
shown in Figs. 10.7, 10.8 and 10.9.
It should be noted that the Lewis form is a two-parameter approximation of
the real sectional shape. A comparison between sections of a real body plan and
the corresponding Lewis forms is shown in Fig. 10.4
136
10.4. Experimental methods
Figure 10.4: Comparison of real sectional shape and Lewis form representation
Sectional properties of damping and added mass can also be determined by ex-
periments. Some of the experiments concerns sectional forms that can be exactly
represented by Lewis forms, while other experiments include details of the shape
that cannot be captured by Lewis forms. Experiments were carried out by Vugts
at Delft Shipbuliding Laboratory [51] for a number of sectional forms. The ex-
perimental set up is indicated in Fig. 10.5 and the forms are shown in Fig. 10.6.
Note that the differences between cylinders A, B and C cannot be captured
by Lewis forms since they have the same Beam/Draught ration and sectional area
coefficient. Results for the cylinders are compared to Lewis form theory in Figs.
10.7, 10.8 and 10.9.
137
10. SOLUTION METHODS -
STRIP THEORY
138
10.4. Experimental methods
139
10. SOLUTION METHODS -
STRIP THEORY
140
10.4. Experimental methods
141
11
SHIP IN REGULAR WAVES
The arising forces are due to the vertical position, η3 , Eq. (11.1), vertical velocity
143
11. SHIP IN REGULAR WAVES
c33 = ρ g B . (11.5)
The dynamic forces due to the acceleration and velocity of the body are associated
with the forced oscillatory motion of the ambient water. For example the accelera-
tion of a floating body is associated with a local, evanescent wave (standing wave,
clapotis) when the water is forced to shift back and forth between the bottom and
the sides. See Fig. 11.1.
As the force a33 η̈3 is in phase1 with the force needed to accelerate the body, the
coefficient a33 is often called the “added mass” and sometimes the hydrodynamic
mass.
For bodies floating in the water surface or positioned close to the water surface
the added mass is a function of the frequency of oscillation due to the evanescent
waves and degree of resonance. Far from the free surface the added mass is con-
stant and depends only on the shape of the body and its vicinity to other bodies
or fixed boundaries. Also for floating bodies, in the limit, as ω → ∞, no waves
are formed, and the added mass becomes independent on the frequency. See Fig.
11.2b for an example of added-mass coefficients for a rectangular cross section in
water shallow compared to the draft.
1 Or sometimes in antiphase for slightly submerged bodies, that is, the added mass a33 is nega-
tive.
144
11.1. The Two-Dimensional Heave Problem
η̂32 1 ω
ω 2 b33 = 2 ρ g a2 , (11.9)
2 4 k
from which b33 can be solved
2 2
a 1 a g2
b33 = ρ g =ρ , (11.10)
η̂3 kω η̂3 ω3
which is valid for all frequencies and shows that the damping is always positive,
which is not necessarily true for the added mass. Remember that according to the
145
11. SHIP IN REGULAR WAVES
(a)
(b)
146
11.1. The Two-Dimensional Heave Problem
(c)
147
33 $ ηˆ ! kω $ ηˆ ! ω 3
% 3" % 3"
which is valid for all frequencies and shows that the damping is always positive,
which is not necessarily true for the added mass. Remember that according to the used
linear assumptions and the potential theory, the given coefficients are only valid for
11. SHIP IN REGULAR WAVES
small-amplitude motion and viscous non-linear damping is not taken into account
either. In reality the total damping is always larger than the radiation damping.
ζ ( x, t )
ζ( x, 0.25⋅ s )
ζ( x, 0.5⋅ s)
148
11.1. The Two-Dimensional Heave Problem
149
11. SHIP IN REGULAR WAVES
150
11.1. The Two-Dimensional Heave Problem
In Fig. 11.4 B and β are indexed n to stress that they are valid for a section No
n along the ship. The angular frequency ωe is the encounter, angular frequency
experienced by the ship. At zero speed the encounter frequency is identical to the
wave frequency.
Similarly the radiation-damping coefficient can be assessed by reading the
coefficient A from the diagrams in Fig. 11.5. Here A is the ratio between the
amplitude of the radiated waves and the amplitude of the driving heave motion.
Compare Equation (6.10).
ρ g2 A2
b33 = . (11.12)
ω3
A is a function of the same parameters as the added mass, namely B/T , β , and
ωe2 B/(2 g).
Jensen et al. [23] simply used for a ship at speed V and heading, θ , the constant
added mass, a33 = ρ g B T , with astonishingly good result, and a somewhat more
complicated closed-form expression for b33 :
ρ g2 A2
b33 = 3 3 ,
ω α
where
ωe2 B −
ωe2 T 1 2
A = 2 sin e g = 2 sin k B α e−k T α .
2
2g 2
ζ = a cos(k y + ω t) , (11.13)
and ask ourselves which forces that will act on the body. See Fig. 11.6. This so-
called wave-diffraction or scattering problem can be solved in a similar manner as
used for solving the added mass and the radiation-damping coefficient. The differ-
ence is now that the boundary condition on the hull states that the water velocities
perpendicular to the hull surface should be nil, that is, the water particle velocities
of the scattered wave must compensate the normal water particle velocities of the
incident wave. This will cause a part of the wave to be reflected and the other part
to be transmitted passed the hull.
To make it simple here we will restrict our treatise to using the small-body
assumption, which demands that the cross-section width, beam, should be less
than a fourth to fifth of a wavelength for producing reasonable results. Then the
151
11. SHIP IN REGULAR WAVES
variation of incident wave properties across the beam can be neglected2 , which is
equivalent of setting y = 0 in Eq. (11.13).
Sticking to the small-body assumption the wave excited force acting on the
cross-section can be expressed by help of the same coefficients a33 , b33 and c33 as
above.
Assume temporarily the cross-section is rectangular with a flat, horizontal bot-
tom and the draught T , and further that the wave is not distorted by the presence
of the ship, which by help of the more complete theory can be shown to be ap-
proximately valid. Then the pressure from the incident wave on the ship’s bottom
is
−k T
p = ρ g T +ae cos(ω t) , (11.14)
where the first term is balanced by the displacement of the ship and can be dis-
missed in this context, and the wave excited force due to relative displacement
will be
as
c33 = B ρ g . (11.16)
That is, the wave-excited force depends on the relative motion between the wave
and the fixed ship, in the same way as the reaction force on the ship in still water
depended on the motion of the ship. This force is calculated using the undisturbed
2 Some authors use the term long-wave approximation, but this notation is in wave contexts
mostly used to denote wave-theory approximations for waves with a longer wavelength than 10 to
20 water depths.
152
11.1. The Two-Dimensional Heave Problem
pressure in the incident wave on the body surface and is often named the Froude-
Krylov force. It can with small errors be used under the small-body assumption,
wile for shorter wave in relation to the body diameter the exciting pressure must
be calculated from the disturbed potential flow and the Bernoulli equation taking
the presence of the body into account. See e.g. discussion in Korvin-Krokowsky
and Jacobs [27] in connexion with the ship motion problem.
When the wave passes the fixed ship the relative velocity will be
Again the easiest way to solve this problem is to use complex notation. Compare
Eq. 11.27.
153
11. SHIP IN REGULAR WAVES
Let, here, the complex wave progress from portside to starboard, i.e. in the
negative y-direction, compare Eq. 11.13.
the vertical displacement, velocity and acceleration of the water at the depth
z = −T are
The dynamic pressure is, again dismissing the statically balanced mean pres-
sure ρ g T in Eq. (3.44)
where now also the amplitude is complex containing the information of the phase
angle between the wave and the heave motion. The vertical velocity and acceler-
ation are then
Substitute this into the equation of motion Eq. (11.20) and solve for the complex
amplitude η̂3c :
c33 − a33 ω 2 + ı ω b33
η̂3c = a e−k T . (11.29)
c33 − (m + a33 )ω 2 + ı ω b33
The motion amplitude is finally given by the modulus, η̂3 = |η̂3c |, the phase angle
between wave motion and heave by the argument, arg (η̂3c ), and the complex
transfer function from wave motion to heave motion by Tc = η̂3c /a. The factor
e−k T is often called the Smith effect and shows that the deeper the draught the less
are the excitation forces. In Fig. 11.7 the response amplitude operator RAO3 =
|η̂3c | /a and the wave-excited force normalised by the displacement force ρ S are
shown as functions of angular frequency. In Fig. 11.8 the phase angles between
wave and heave motion; and between wave and wave-excited force as functions
of angular frequency are shown.
154
11.1. The Two-Dimensional Heave Problem
Figure 11.8: The phase angles between wave and heave motion (continuous
line) and between wave and wave-excited force (dashed line) as
functions of normalised angular frequency, Ω = ω/ωN , for a two-
dimensional ship.
155
11. SHIP IN REGULAR WAVES
In Fig. 11.7 it is seen that for long waves ω → 0 the amplitude of the heave
motion is the same as the amplitude of thepwave, then grows to a maximum just
below the natural angular frequency, ωN = c33 /(m + a33 ), and finally attenuates
to nil for higher frequencies.
In Fig. 11.8 it is seen that for long waves ω → 0 the wave-excited pforce is
in phase with the wave, then as the wave frequency increases to ω = c33 /a33
the force is 90◦ before the wave and for higher frequencies the force becomes in
opposition to the waveform. For all frequencies the maximum of the wave-excited
force appears before the maximum of the wave elevation, 0 < εwF < 180◦ . The
heave motion, on the other hand, appears for all frequencies after both the wave
motion and the wave-excited force, εwm < 0◦ < εwF .
Note, however, that strictly speaking the small-body assumption in this ex-
ample is violated for angular frequencies above approximately 1 rad/s. We have
furthermore assumed that the added mass and the radiation-damping coefficient
are constant. Yet, the features of the response are valid. A more thorough calcu-
lation using correct potential forces and coefficients derived from potential theory
would give the same result in principle, although with somewhat different graphs,
especially for angular frequencies above 1 rad/s.
The sway problem can for a two-dimensional ship be treated similarly but as con-
cerns the roll problem the dynamic equilibrium must be solved for the ship as
an entity as the roll stiffness is due to the mass distribution of the entire ship.
Therefore we will save the roll problem till later.
The simple method used above for the two-dimensional ship can be used also
for real three-dimensional ships, in an approximate variant of the so called strip
theory, where the ship is divided into slices - or strips of the hull surface - and
the two-dimensional flow problem is assumed to hold for each slice or strip (see
Fig. 11.9). The problem is then reduced to integrating the forces or hydrodynamic
characteristics along the ship knowing the added-mass and radiation-damping co-
efficients for each strip. The ship has to be reasonably slender, at least, L/B > 5
to neglect the end effects. The approximation is also better for ships with pointed
ends as the three-dimensional end effects are less pronounced for such ships.
156
11.2. The Uncoupled Three-Dimensional Heave Problem Using Strip Theory
Figure 11.9: In the strip theory the ship is divided into slices or strips of the hull
surface. For each strip the two-dimensional hydrodynamic prob-
lem is then solved.
The vertical reaction force acting on the body from the water can after integrating
the two-dimensional reaction force, Eq. 11.4, along the ship be written
where now the coefficients A33 , B33 and C33 are the integrated or summed quanti-
ties along the ship
Z
A33 = a33 (x)dx = ∑ a33n ∆xn , (11.32)
L n
Z
B33 = b33 (x)dx = ∑ b33n ∆xn , (11.33)
L n
Z
C33 = c33 (x)dx = ∑ c33n ∆xn = ∑ ρ g Bn ∆xn . (11.34)
L n n
157
11. SHIP IN REGULAR WAVES
where T is the draught; p(x, −T ) the undisturbed pressure at the box bottom
z = −T ; ẇ(x, −t) the water acceleration at the bottom z = −T ; w(x, −T ) the water
velocity at the bottom z = −T ; a33 the vertical two-dimensional added mass; b33
the vertical two-dimensional radiation damping coefficient; B the beam (breadth)
of the ship section and, finally, L the length of the ship.
Also in this case we have utilised the “small-body” assumption that the beam
of the ship should be small in relation to the wavelength, i.e. B < λ /4. For
following or meeting waves this is not a restrictive assumption as the apparent
wavelength for theses two conditions is infinitely long.
Again we will, for simplicity use complex notation. Also, because it is more
useful and not very complicated we will directly formulate the wave-excited force
for waves oblique to the ship in the following example. The potential for such
waves in complex notation is
a g k z ı k (x cos(θ )+y sin(θ )) −ı ω t
φc = ı e e e . (11.37)
ω
For the ship heading along the x-axis i.e. y = 0 it reduces to:
a g k z ı k x cos(θ ) −ı ω t
φc = ı e e e .
ω
Further, at y = 0 we also have the water vertical motion at the bottom z = −T :
158
11.2. The Uncoupled Three-Dimensional Heave Problem Using Strip Theory
or
Z L/2
F3e = a e−k T e−ı ω t ρgB − ω 2 a33 − ı ωb33 eı k x cos(θ ) dx . (11.39)
−L/2
Using Eq. 11.31 with the complex motion above the equation of motion finally
gives
C33 − (m + A33 )ω 2 − ı ω B33 η̂3c e−ı ω t =
sin(k cos(θ )L/2)
a e−kT ρgB − ω 2 a33 − ıωb33 e−ı ω t 2 , (11.42)
k cos(θ )
which can be solved for the complex motion to form a nice closed-form expression
159
11. SHIP IN REGULAR WAVES
Figure 11.10: Heave wave-excited force as a function of ship length over wave-
length, L/λ .
Exercise 6.1
Inspecting Eq. (11.43) one can see that the amplitude response function
|η̂3c |/a has one minimum around ρgB = ω 2 a33 , minima in deep water
for sin(k cos(θ )L/2) = sin (ω 2 /g) cos(θ )L/2 , and a maximum around
ρgBL = (ρT BL + a33 L)ω 2 .
The natural angular frequency is obtained by psetting the driving force and
damping to nil in Eq. (11.42), which gives ωN = C33 /(m + A33 ).
In Fig. 11.11 the heave motion as a function of time at zero speed in head
waves are shown, in Fig. 11.12 the amplitude response and in Fig. 11.13 the phase
lag between heave motion and wave are shown as functions of angular frequency.
160
11.3. The Pitch Problem Using Strip Theory
Figure 11.11: Heave motion and water level as a function of time in head waves
with the wave amplitude a = 1 m.
where now the coefficients A55 , B55 and C55 are the integrated quantities along the
ship:
Z L/2
U2
A55 = a33 (x) x2 dx −A33 , (11.46)
−L/2 ωe2
Z L/2
U2
B55 = b33 (x) x2 dx + 2 B33 , (11.47)
−L/2 ωe
Z L/2
C55 = ρgV (zB − zG ) + ρg B(x) x2 dx = ρgV GM L . (11.48)
−L/2
Here the first terms in the added mass and radiation-damping coefficient are caused
by the motion in still water as in Eq. 11.4 or 11.31 taking into account that the
vertical motion due to pitch at each cross section is the pitch motion multiplied
by the lever, x, and that the reaction moment also is the sectional reaction force
multiplied by the lever, x. From this fact comes the x2 in the integrations. The
second terms, the forward-speed terms, are caused by the forward-speed potential
and will not be explained further here, see Salvesen et al. [43]. U is the speed of
the ship and ωe is the angular frequency of encounter.
For the box-like ship, again the coefficients a33 and b33 are functions of the os-
cillation frequency but independent of x, and the beam is B(x) = B. For simplified
estimates a characteristic frequency of the exciting wave can be used and, actu-
ally, for each sea state the resulting sea-keeping properties will be astonishingly
realistic.
161
11. SHIP IN REGULAR WAVES
Figure 11.12: Amplitude response function at the headings: 0, π/4 and π/2 rad.
Figure 11.13: Phase lag between heave motion and wave at the headings: 0, π/4
and π/2 rad.
162
11.3. The Pitch Problem Using Strip Theory
L3
A55 = a33 , (11.49a)
12
L3
B55 = b33 , (11.49b)
12
L3 L3
C55 = ρgV (zB − zG ) + ρgB ≈ ρgB . (11.49c)
12 12
where T is the draught; pc (x, −T ) the undisturbed pressure at the box bottom
z = −T ; ẇc (x, −T ) the water acceleration at the bottom z = −T ; wc (x, −T ) the
water velocity at the bottom z = −T ; a33 the vertical two-dimensional added
mass; b33 the vertical two-dimensional radiation damping coefficient; B the beam
(breadth) of the ship section and L the length of the ship.
Again we will, for simplicity use complex notation. Also, because it is more
useful and not very complicated we will directly formulate the wave-excited mo-
ment for waves oblique to the ship in the following example. The potential and
derived properties for such waves in complex notation were given above in con-
nection with the heave problem Eq. (11.37). Substituting these expressions into
Eq. (11.50) gives
Z L/2
F5e = − ρ g a B − a ω 2 a33 − ı a ω b33 x e−kT eıkx cos(θ ) e−ı ω t dx , (11.51)
−L/2
or
Z L/2
−kT −ı ω t
2
F5e = −a ρ g B − ω a33 − ı ω b33 e e eıkx cos(θ ) x dx . (11.52)
−L/2
which approaches zero in the limit as θ → π/2, and thus confirms that in beam
regular waves the pitch is zero.
163
11. SHIP IN REGULAR WAVES
Figure 11.14: The amplitude of the pitch wave excited forces at zero speed as
a function of ship length divided by wave length in deep water,
L/λ .
Exercise 6.2
Look at Fig. 11.14. Why are the forces zero for meeting waves at 1.5,
2.5, 3.5 etc.?
Now set the pitch motion to η5c = η̂5c e−ı ω t , which gives the pitch angular
velocity and acceleration of the ship to
Using Eq. (11.45) with the complex motion above, the equation of motion
finally gives
C55 − (I5 + A55 )ω 2 − ı ω B55 η̂5c e−ı ω t = −a ρgB − ω 2 a33 − ı ω b33 ×
−k T −ı ω t 2ı L L L
e e sin k cos(θ ) − k cos(θ ) cos k cos(θ ) ,
(k cos(θ ))2 2 2 2
(11.54)
which can be solved for the complex pitch motion amplitude in a closed-form
164
11.4. The Roll Problem
expression
−kT ρgB − ω 2 a33 − ıωb33
η̂5c = − a e ×
C55 − (I5 + A55 )ω 2 − ı ω B55
2ı L L L
sin k cos(θ ) − k cos(θ ) cos k cos(θ ) . (11.55)
(k cos(θ ))2 2 2 2
Exercise 6.3
Inspecting Eq. (11.55) one can see that the amplitude response func-
tion |η̂5c |/a has one minimum around ρgB 2
L L
= ω a33 , minima for
sin(k cos(θ )L/2) = k cos(θ ) 2 cos k cos(θ ) 2 , and a maximum around
C55 = (I5 + A55 )ω 2 .
165
11. SHIP IN REGULAR WAVES
Figure 11.15: Pitch amplitude response functions at the headings: 0, π/4 and
π/2 rad.
Figure 11.16: Phase lags between pitch motion and wave at the headings: 0,
π/4 and π/2 rad.
166
11.4. The Roll Problem
Figure 11.17: Pitch motion and wave slope at the origin of a ship as functions
of time in a swell with the period around 43 s. Observe that the
ship follows the slope of the wave-surface exactly for this swell,
but due to the definitions of slope and pitch angle, the functions
are 180 deg out of phase.
B3
C44 = ρgV (zB − zG ) + ρgL = ρgV GM T . (11.60)
12
The dynamic moments due to the roll angular acceleration and roll angular
velocity of the ship are - as for the other degrees of freedom - associated with
the forced oscillatory motion of the ambient water. The roll acceleration of the
167
11. SHIP IN REGULAR WAVES
ship is thus associated with a local, evanescent wave in which the water is forced
to shift back and forth between the bottom and the sides alternately to starboard
and portside. The effect is an “added moment of inertia”, which can be calculated
by help of two-dimensional potential theory in combination with strip theory or
directly from three-dimensional potential theory for short ships or floating objects
with complicated shapes like offshore drilling platforms. As, for a ship, the order
of magnitude of the added moment of inertia is only 10 to 20 % of the mass
moment of inertia of the ship itself, it can suffice here to use an approximate
value. Thus
A44 = 0.15 I4 . (11.61a)
For bodies floating in the water surface or positioned close to the water surface
this added moment is a function of the frequency of oscillation. Far from the free
surface the added moment of inertia is constant and depends only on the shape
of the body and its vicinity to other bodies or fixed boundaries. Also for floating
bodies, in the limit, as ω → ∞, no waves are formed, and the added moment
becomes independent of the frequency and is half that of the body mirrored in the
water surface and submerged deeply below the free surface.
To have an approximate value of the roll moment of inertia one can use the
fact that usually the radius of inertia in roll i4 is of the order of 0.4B, where B is
the beam of the ship, and thus
168
11.4. The Roll Problem
1. radiated waves
where Ac is the cross-sectional area of the submerged part of the section. For
triangular cross-sections with 3 ≤ B/T ≤ 6 the functions a( · ), b( · ) and d( · )
became:
B B
a = 0.256 − 0.286 , (11.63a)
T T
B B
b = −0.11 − 2.55 , (11.63b)
T T
B B
d = 0.033 − 1.419 . (11.63c)
T T
169
11. SHIP IN REGULAR WAVES
Table 11.1: Added damping ratio ξ for a Panamax container vessel in different
headings. From [23].
For ships with fuller lines, e.g. container ships and tankers, the same procedure
was used for rectangular cross-sections with 1 ≤ B/T ≤ 3 to obtain:
B B
a = −3.94 + 13.69 , (11.64a)
T T
B B
b = −2.12 − 1.89 , (11.64b)
T T
B B
d = 1.16 − 7.97 . (11.64c)
T T
170
11.4. The Roll Problem
This moment is in phase with the cross-beam slope velocity of the wave because
here we have neglected the force in phase with the slope and with the slope accel-
eration, which are less important when the magnitude of the roll motion shall be
assessed.
Integrating along our box-like ship yields the moment
Z L/2 r 2
ρg
F4 = a b44 eı k x cos(θ ) e−ı ω t sin(θ ) dx
−L/2 ω
r
ρ g2 2 sin(θ ) −ı ω t L
= b44 e sin k cos(θ ) . (11.68)
ω k cos(θ ) 2
171
11. SHIP IN REGULAR WAVES
As the wave excited roll moment should be added to the reaction moment from
the water due to the roll of the ship in still water, Eq. (11.59), and be balanced by
the roll moment of inertia the following equation of roll motion is yielded:
(I4 + A44 ) η̈4c + B44 η̇4c +C44 η4c = F4 . (11.70)
Although the wave frequency variation in an irregular sea state influence the ship
motion at zero speed, the effect of the frequency change due to ship moving
through the waves is much more pronounced. A ship progressing through a fol-
lowing wave with the speed, U, at the heading, θ , in relation to the direction of
wave propagation, will be passed by a wave crest every encounter period
λ
Te = . (11.72)
c −U cos(θ )
Negative encounter period means that the ship moves faster than the wave. This
can only happen for waves abaft the beam, −90◦ < θ < 90◦ . In head waves,
meeting (encountered) waves, 90◦ < θ < 270◦ , the encountered period is always
positive and shorter than the wave frequency.
The angular frequency of encounter is thus
2π 2π
ωe = = (c −U cos(θ )) , (11.73)
Te λ
or
λ
= c −U cos(θ ) ,
Te
For deep water Eq. (11.73) can be written
ω2U
ωe = ω − cos(θ ) . (11.74)
g
See Fig. 11.24.
172
11.5. Forward-Speed Effect
Figure 11.21: A blow up of Fig. 11.20 around the roll resonance frequency.
173
11. SHIP IN REGULAR WAVES
(a) (b)
The heave, pitch and roll motions are resonant motions, while the horizontal mo-
tions sway, surge and yaw are non-resonant because there are no stiffnesses in the
latter modes of motion for unmoored ships.
s s
C33 ρgAw
in heave ωN3 = = ,
ρV + A33 ρV + A33
r
C44
in roll ωN4 = ,
I4 + A44
s r
C55 ρgIw5
in pitch ωN5 = ≈ .
I5 + A55 I5 + A55
The last approximation can be used because the hydrostatic stiffness dominates in
pitch. Aw is the water-plane area and Iw5 = BL3 /12 is the area moment in pitch of
the water-plane area.
175
11. SHIP IN REGULAR WAVES
Hitherto we have assessed the global motion of the ship in six degrees of freedom
referred to the origin of the chosen co-ordinate system. For applications we have
to be able to describe the translational motion in three degrees of freedom at any
point of the ship. The aim can be to investigate the freeboard, the risk for pro-
peller emergence out of the water, the risk for slamming and also accelerations
for assessing cargo fastenings, comfort and seasickness. These local motions are
called derived responses.
As we in the strip theory cannot predict the surge motion with any degree of
success the first row of Eq. 11.76 describing the horizontal motion in the surge
direction is of no interest in this chapter. Furthermore as we in this treatise have
176
11.7. Derived Responses
not calculated the yaw motion, although it is fully feasible, we cannot use the
second row for the crossbeam horizontal motion. The most important motion is,
however, the vertical motion described by the third row. Thus
Here we have retained all the vertical motions heave, pitch and roll [η3 , η4 , η5 ]
contributing to the vertical motion of a point. It should be warned that as the
used method for calculating the roll motion is not phase correct, it should not be
included in Eq. (11.77) if it lessens the vertical motion. For statistical estimates
in irregular waves see Chapter 9.
The vertical velocity and acceleration of the same point are likewise
and
s̈3 = η̈3 + yη̈4 − xη̈5 . (11.79)
The vertical motions heave, pitch and roll [η3 , η4 , η5 ] are most conveniently in-
troduced into Eqs. (11.77) to (11.79) in their complex time-domain form. The
real motion is then assessed by taking the real part of the derived motion s3 .
An example of a calculation of the vertical bow motion of the example ship
is shown in Fig. 11.25 with y = 0 m, or no roll motion. The ship is a box 100 m
long, 20 m wide and with 10 m draught. In Figs. 11.26 and 11.27 the response
amplitude operator and phase lag is shown.
11.7.2 Acceleration
It is important to assess the accelerations of a ship because it is difficult to work
if the accelerations are too large, say g/3; accelerations at certain frequencies
also causes seasickness; cargo may get loose or fastenings must be attached and
designed. In the passenger and cruising trade the comfort is important, in the off-
shore industry the focus is to avoid downtime of operations like drilling or pump-
ing oil and gas, while in cargo traffic the safety of the goods is most important but
also speed reductions should be avoided. In Figs. 11.28 to 11.30 the acceleration,
Eq. (11.79), at the bow of the example ship is shown.
sFB (t) = η3 (t) + yη4 (t) − xη5 (t) + zFB − ζ (x cos(θ ) , y sin(θ ) ,t) > 0 , (11.80)
177
11. SHIP IN REGULAR WAVES
Figure 11.25: Vertical bow motion and wave elevation near pitch and heave res-
onance.
Figure 11.26: The response amplitude operator for the vertical bow motion.
178
11.7. Derived Responses
Figure 11.28: Vertical bow acceleration and wave elevation near pitch and heave
resonance.
Figure 11.29: The response amplitude operator for the vertical bow accelera-
tion.
where sFB is the instantaneous freeboard; zFB the static freeboard at station (x , y)
and actual trim and ζ (x cos(θ ) , y sin(θ ) ,t) the instantaneous wave elevation at
station (x , y). See Figs. 11.31 and 11.32.
The risk for the propeller to emerge out of the water and the risk that the bottom
will rise above the water and experience slamming at re-entry must be assessed.
179
11. SHIP IN REGULAR WAVES
This can be tested by a similar expression as was used for assessing the freeboard.
sBE (t) = −η3 (t) − yη4 (t) + xη5 (t) + T + ζ (x cos(θ ) , y sin(θ ) ,t) > 0 , (11.81)
11.7.5 Slamming
Slamming will appear when the ship’s bottom has risen out of the wave and hits
back at re-entry. The slamming pressure depends on the relative velocity squared,
between the ships bottom and the water surface, the angle between them, their
irregularity and content of air bubbles. The slamming is complicated and it is
referred to e.g. Faltinsen [12] for a deeper description. The slamming pressure can
very roughly be approximated by ps = C ρ Urel 2 , where C is a constant. Then the
∂
Urel (t) = (−η3 (t) − yη4 (t) + xη5 (t) + T + ζ (x cos(θ ) , y sin(θ ) ,t))
∂t
= −η̇3 (t) + xη̇5 (t) + ζ˙ (x cos(θ ) , y sin(θ ) ,t) . (11.82)
180
11.7. Derived Responses
Figure 11.31: Elevation of railing aft and wave elevation at the same place.
181
11. SHIP IN REGULAR WAVES
Figure 11.33: The elevation of the propeller centre and the wave surface as a
function of time. In this case the propeller is submerged all the
time.
182
11.8. Coupled Linear Pitch and Heave Motion at Forward Speed
This velocity should be evaluated for all moments when Eq. (11.81) is zero
−η3 (t) − yη4 (t) + xη5 (t) + T + ζ (x cos(θ ) , y sin(θ ) ,t) = 0 , (11.83)
and Urel is positive. See further for statistical simplifictions in irregular waves in
Paragraph 9.6.
Assume that, in calm water, the ship is forced to oscillate and pitch with the am-
plitudes, η̂3 and η̂5 , at the angular frequency, ω. The vertical reaction force and
pitch reaction moment acting on the body from the water can be written, after
integrating the two-dimensional reaction forces along the ship,
F3 = −A33 η̈3 − B33 η̇3 −C33 η3 − A35 η̈5 − B35 η̇5 −C35 η5 , (11.84)
and
F5 = −A53 η̈3 − B53 η̇3 −C53 η3 − A55 η̈5 − B55 η̇5 −C55 η5 , (11.85)
or in matrix form
F r = −A η̈ − B η̇ − C η , (11.86)
and Z
C35 = C53 = ρg x ds . (11.91)
AW P
183
11. SHIP IN REGULAR WAVES
184
11.9. Coupled Pitch and Heave Motion at Zero Speed Including Non-Linear Viscous
Damping
Figure 11.35: Definition of the chosen coordinate system for a ship. The z-axis
is drawn vertically through the centre of buoyancy, CB, and the
centre of gravity, CG. (From Faltinsen).
For the box-like ship used in the examples we will investigate the effect of viscous,
drag damping, which is quadratic and therefore hinder us from solution in the
frequency domain, the use of complex numbers and linear superposition. We will
therefore formulate the problem in the time domain, using the notations we used
before and the coordinate system presented in Fig. 11.35.
11.9.2 Heave
Displacement
η3 (t) = η̂3 cos(−ω t − ε3 ) . (11.99)
Velocity
η̇3 (t) = η̂3 ω sin(−ω t − ε3 ) . (11.100)
Acceleration
η̈3 (t) = −η̂3 ω 2 cos(−ω t − ε3 ) . (11.101)
185
11. SHIP IN REGULAR WAVES
11.9.3 Pitch
Displacement
η5 (t) = η̂5 cos(−ω t − ε5 ) . (11.102)
Angular velocity
η̇5 (t) = η̂5 ω sin(−ω t − ε5 ) . (11.103)
Acceleration
η̈5 (t) = −η̂5 ω 2 cos(−ω t − ε5 ) . (11.104)
Heave
Z L/2 Z L/2
m η̈3 = ¨
a33 ζ − η̈3 + xη̈5 dx + b33 ζ˙ − η̇3 + xη̇5 dx
−L/2 −L/2
| {z } | {z }
added mass force linear damping force
Z L/2
1
+ ρCD B ζ˙ − η̇3 + xη̇5 ζ˙ − η̇3 + xη̇5 dx
−L/2 2
| {z }
viscous drag force
Z L/2 Z L/2
+ ρgBζ dx + ρgB (−η3 + xη5 ) dx . (11.105)
−L/2 −L/2
| {z } | {z }
Froude-Krylov force buoyancy force
Assemble exciting forces on the right hand side and linear reaction terms on the
left hand side yields
Z L/2
(m + A33 ) η̈3 + B33 η̇3 +C33 η3 = ¨ ˙
a33 ζ + b33 ζ + ρgBζ dx
−L/2
Z L/2
1
+ ρCD B ζ˙ − η̇3 + xη̇5 ζ˙ − η̇3 + xη̇5 dx .
−L/2 2
(11.106)
We have then used that
Z L/2
A33 = a33 dx , (11.107a)
−L/2
Z L/2
B33 = b33 dx , (11.107b)
−L/2
Z L/2
C33 = ρgB dx , (11.107c)
−L/2
186
11.9. Coupled Pitch and Heave Motion at Zero Speed Including Non-Linear Viscous
Damping
ζc = a eı(k x−ω t) ,
ζ = Re (ζc ) = a cos(k x − ω t) ,
ζc |ζc | = a2 eı (k x−ω t) = a2 (cos(k x − ω t) + ı sin(k x − ω t)) ,
Re (ζc |ζc |) = a2 cos(k x − ω t) ,
Pitch
Z L/2 Z L/2
I5 η̈5 = − ¨
a33 ζ − η̈3 + xη̈5 x dx − b33 ζ¨ − η̈3 + xη̈5 x dx
−L/2 −L/2
| {z } | {z }
added mass moment linear damping moment
Z L/2
1 ˙ ˙
− ρCD B ζ − η̇3 + xη̇5 ζ − η̇3 + xη̇5 x dx
−L/2 2
| {z }
viscous drag moment
Z L/2 Z L/2
− ρgBζ x dx − ρgB (−η3 + xη5 ) x dx . (11.109)
−L/2 −L/2
| {z } | {z }
Froude-Krylov moment buoyancy moment
Assemble exciting moments on the right hand side and linear reaction terms on
the left hand side.
Z L/2
(I5 + A55 ) η̈5 + B55 η̇5 +C55 η5 = − a33 ζ¨ + b33 ζ˙ + ρgBζ x dx
−L/2
Z L/2
1
− ρCD B ζ˙ − η̇3 + xη̇5 ˙
ζ − η̇3 + xη̇5 x dx .
−L/2 2
(11.110)
187
11. SHIP IN REGULAR WAVES
i.e. again all linear coupling terms vanish and we are left with the last non-linear
coupling term.
The non-linear coupled drag term constitutes a problem, when making assessment
of sea-keeping properties in the frequency domain.
In roll, the linear radiation damping is very small and the viscous non-linear
drag damping dominates and must therefore be assessed to get realistic motion.
Jensen increased the linear damping coefficient, Equation (6.66), by comparing
the calculated motion with seakeeping model tests. In this section we will show
that such an equivalent linear drag-damping coefficient depends on the amplitude
of motion.
Neglecting the coupling between roll and sway we can symbolically write the
drag damping moment in beam regular sea as
˙ ˙
FD4 = K ζy − η̇4 ζy − η̇4 , (11.113)
where K can be set to ρ CD B2 L/2 and ζ˙y = ∂ 2 ζ /∂t∂ y is the angular velocity of
the wave slope in the y direction i.e. in the starboard-portside direction.
When the non-linear roll damping is important usually ζ˙y η̇4 and then
188
11.10. Equivalent Linearised Drag Damping
189
12
SHIP IN IRREGULAR WAVES
In Chapter 6 the heave motion at zero speed in a plane regular wave with the
amplitude, a, and the propagation direction, θ , was deducted. For a box-like ship
191
12. SHIP IN IRREGULAR WAVES
the wave excited heave motion was then represented by a complex heave motion:
ρg B − ω 2 a33 − ı ω b33
η3c = η̂3c eı ω t =a e−k T ×
ρgBL − ω 2 (ρT BL + a33 L) − ı ω b33 L
sin (k(cos(θ )L/2) −ı ω t
2 e . (12.1)
k cos(θ )
Here the modulus |η̂3c | of the complex amplitude represents the heave ampli-
tude, the argument, arg (η̂3c ), the phase lag to the wave motion and the real part,
Re (η3c ), the real heave motion.
By dividing the complex amplitude by the wave amplitude, a, we get a com-
plex transfer function between the wave motion and the heave motion as
Now forming the product of the transfer function (12.2) and the wave function
(12.3) at x = 0 m, we get the complex heave motion in the time domain as
The wave function (12.3) is the same as Equation (5.13), but complemented with
a phase angle, ε.
The transfer function (12.2) transfers both the amplitude, |T3c (ω , θ )|, and the
phase lag, arg(T3c (ω , θ )), to the wave motion. The real heave motion can thus be
written
The amplification factor, |T3c (ω ,t)|, or frequency response function is the heave
amplitude divided by the wave amplitude (m/m) and is often called the heave-
response amplitude operator or RAO1 and has it equivalents in other degrees of
freedom as well as for derived responses.
1 In some literature the square of |T (ω , θ )| is called RAO because the square is used in the
3c
multiplication by the wave spectrum to form the motion spectrum.
192
12.1. The Heave Motion in Irregular Waves
From Eqs. ( 12.4) and (12.6) now follows by superposition that the heave motion
in the time domain can be written
∞
η3c (t) = ∑ ai T3c (ωi , θ ) eı(−ω t+εi ) . (12.7)
i=1
This time-domain heave is shown in Fig. 12.2 in three wave directions for a
box-like ship with dimensions length × beam × draught = L × B × T = 100 ×
20 × 10 m3 . The above simulation was done in the frequency domain starting out
with a wave spectrum, multiplying with a transfer function to produce a response
spectrum and from this response spectrum simulating a time trace of response by
an inverse fast-Fourier transformation (IFFT). By preserving the random phase
angles, εi , of the component waves and “adding” the phase lags of the response
components by the complex transform multiplication the time trace of response
shows the correct time response – to first order that is.
In Fig. 12.3 the heave spectrum for following (or meeting) waves in two dif-
ferent sea states from Table 5.8 are shown using the ISSC wind-sea spectrum Eq.
(4.25). It is seen how the sea state with shorter mean period does not excite the
ship much within the frequency range where the transfer function is appreciable.
In Fig. 12.4 the response spectra for three headings are shown. There it is clearly
seen that beam sea is unfavourable for the heave motion.
193
12. SHIP IN IRREGULAR WAVES
(a)
(b)
(c)
Figure 12.2: Heave motion in 300 seconds of simulated plane waves with Hs = 2
m and T02 = 5 s using an ISSCb spectrum: (a) following waves, (b)
quartering waves and (c) beam waves.
194
12.1. The Heave Motion in Irregular Waves
Figure 12.3: Heave-response spectra of the box-like ship in the same heading
(θ = 0◦ ) at two different sea states.
195
12. SHIP IN IRREGULAR WAVES
Table 12.1: Some characteristics of the two wave spectra and their heave re-
sponses.
The pitch response can be treated exactly as the heave response with the complex
transfer function
η̂5c (ω , θ )
T5c (ω , θ ) = , (12.10)
a
which has the unit rad/m. The complex pitch amplitude η̂5c (ω , θ ) is given by
Equation (11.55). One can also choose to form a wave-slope spectrum from the
wave amplitude spectrum first and use a dimensionless transfer function
η̂5c (ω , θ )
T5c,slope (ω , θ ) = , (12.11)
k(ω)a
196
12.3. The Response of Vertical Motion at Station x
197
12. SHIP IN IRREGULAR WAVES
Figure 12.5: Pitch-response spectra of the box-like ship in the same heading
(θ = 0◦ ) at two different sea states.
Figure 12.6: Pitch-response spectra of the box-like ship in three headings at two
different sea states.
198
12.3. The Response of Vertical Motion at Station x
Figure 12.7: Vertical motion response of the box-like ship in three stations.
Figure 12.8: Vertical velocity response of the box-like ship at bow in following
waves.
Figure 12.9: Vertical acceleration response of the box-like ship at bow in fol-
lowing waves.
199
13
DERIVED RESPONSES
η1
s = η2 + Ω × r , (13.1)
η3
η1 + zη5 − yη6
s = η2 − zη4 + xη6 . (13.2)
η3 + yη4 − xη5
η̇1 η̈1
ṡ = η̇2 + Ω̇ × r , s̈ = η̈2 + Ω̈ × r . (13.3)
η̇3 η̈3
For large angles of rotation these simple expressions are not valid but more
complicated expressions must be made like those in the manoeuvring compendium.
201
13. DERIVED RESPONSES
Exercise 2.1
Assume that a ship is moving in a closed elliptical orbit in the x − z plane
without pitching, i.e. η2 = η4 = η5 = η6 = 0. The orbit is described by
" # " #
η1 η̂1 sin(ωt)
η= = .
η3 ηˆ3 cos(ωt)
The constant buoyancy force is balanced by the weight of the body. The
sum of all varying forces acting on the platform must then be given by the
following simple equation of motion:
" # " #
m 0 d 2 η1
F= .
0 m dt 2 η3
As was told in Chapter 11 , if the wave elevation comes above the instantaneous
position of the railing or above the freeboard at any point of the ship this will result
in water on deck, so called green water. To treat this problem in the frequency
domain we form the transfer function for the relative vertical motion between the
ship and water elevation for any heading and at any station, x, but only along the
longitudinal axis of the ship, y = 0:
Examples of response spectra for the vertical relative motion are shown in Fig. 13.1
202
13.3. Propeller emergence
Figure 13.1: Response spectra of relative motion at bow, amidships and aft.
203
13. DERIVED RESPONSES
Figure 13.2: No of green water occurrences per hour as a function of static free-
board.
complete analogy to what was done for green water. If the draught is T or the
propeller is T below the mean-water surface.
2
T (x)
− 12 σs3rel (x)
PT (x) = P (s3rel (x) > T (x)) = 1 − P (s3rel (x) < T (x)) ≈ e . (13.9)
204
13.4. Slamming, Whipping, Springing
Examples of response spectra for the vertical relative velocity are shown in Fig. 13.3.
As the relative vertical position and the vertical relative velocity can be shown
to be statistically uncorrelated processes [22] the probability of slamming is
2 2
T (x) v0
− 12 σs3rel (x)
+ σs3relvel (x)
Pslam (x) = [P (s3rel (x) > T (x))] [P (s3relvel (x) > v0 )] ≈ e .
(13.14)
The number of slamming occurrences within a time period, t, can be assessed by
t
Nslam = Pslam (x) . (13.15)
T02s3rel
205
13. DERIVED RESPONSES
It is difficult to work if the accelerations onboard are too large. This is impor-
tant for the safety on ships and in the offshore industry, where also downtime of
operations should be avoided. In the passenger and cruising trade the comfort is
important. In freighters cargo may get loose so fastenings must be attached and
designed. In tankers the sloshing in incompletely filled tank compartments may
cause structural problems.
13.5.1 Seasickness
Vertical acceleration is probably the prime reason for seasickness (motion sick-
ness, kinetosis). For instance in Fig. 13.5 a motion-sickness index (MSI) is given
as the share of exposed persons throwing up within a given time at combinations
of mean frequency and standard deviation1 of acceleration.
As an application we can from the acceleration spectrum above calculate the
spectral moments and estimate the mean period of acceleration to T02acc = 9.7 s
and the standard deviation to 0.6 m/s2 . Then assessing seasickness in Fig. 13.5,
we find that between 10 and 20 % of passengers will through up within 2 hours
under such conditions.
1 RMS in Figure 9.9 and Table 9.2 is defined for the acceleration itself. As the mean is zero it
is identical to the standard deviation. The RMS value used in other contexts in this compendium
is defined for the amplitudes or double amplitudes of the quantities e.g. Hrms .
206
13.5. Passenger and Crew
For the performance of the crew the standard deviation 0.1 g for vertical acceler-
ation at the bridge can be adopted, a criterion denoted by “Intellectual work” by
a Nordforsk study [36]. In cabins and restaurants the criterion of “Transit pas-
sengers” is stricter and given as 0.05 g for vertical acceleration. Also criteria for
transversal (horizontal) acceleration are given. See Table 13.1.
207
13. DERIVED RESPONSES
To calculate wave excited first order motions for moored ships in the open sea
or moored offshore platforms is a straightforward matter nowadays - with the
exception of roll motions of ships. To calculate motions for a ship at berth in a
harbour introduces the complications of the wave penetration into and possible
resonance of the harbour to the waves, the non-linear mooring arrangements, but
also the necessity to take the proximity of the quay walls, shore slopes and sea
floor into account, when assessing the hydrodynamic properties of the ship. This
is all automatically “included” when making physical model tests with regular or
pseudo random waves, albeit with model scale deficiencies. Field measurements
are also useful, but cannot be used for testing changes of harbour and mooring
layout. It is however gradually becoming possible to make “complete” numerical
modelling of ships moored in harbours [3].
208
13.6. Motions of moored ships in harbours
209
13. DERIVED RESPONSES
Table 13.3: Criteria for vessel movements for safe mooring conditions at berth.
The movements are peak-peak values. For the berth to be accept-
able, the frequency of these movements should be less than 3 h/year.
From Nordic Council [35].
Table 13.4: Recommended velocity criteria for Safe mooring conditions of var-
ious ships [40].
210
13.6. Motions of moored ships in harbours
Table 13.5: Wave criteria for small craft and pleasure boats. The acceptable
frequency of occurrence is one to a few times per year [40].
211
REFERENCES
[1] B.G. Airy. Encycl. Metrop., chapter Tides and waves. London, 1845.
[2] L. Bergdahl and M. Johansson. Time simulation of the motion of a tension leg
platform. In BOSS ’88, Trondheim, June 1988.
[5] J.P. Comstock, editor. Principles of Naval Architecture. The Society of Naval Ar-
chitecture and Marine Engineers, 1967.
[6] R.R. Craig. Structural Dynamics. John Wiley and Sons, New York, 1981.
[7] W.E. Cummins. The impulse response function and ship motions. Schiffstechnik,
9:101–109, 1962.
[8] J.W. Daily and D.R.F. Harleman. Fluid Dynamics. Addison-Wesley Pub. Co., Read-
ing, Mass., 1966.
[9] R.G. Dean. Evaluation and development of water wave theories for engineering
application. presentation of research results. Special report I, US Army, Corps of
Engineers, Coastal Engineering Research Center, 1974.
[10] R.G. Dean and R.A. Dalrymple. Water Wave Mechanics for Engineers and Scien-
tists. World Scientific, 1991.
[11] DnV. Position mooring (POSMOOR), rules for the classification of mobile offshore
units; part 6, chapter 2. Technical report, Det Norske Veritas, 1989.
[12] O.M. Faltinsen. Sealoads on Ships and Offshore Structures. Cambridge University
Press, 1990.
213
REFERENCES
[13] O.M. Faltinsen and F. Michelsen. Motions of large structures in waves at zero
Froude number. In R.E.D. Bishop and W.G. Price, editors, International Symposium
of Dynamics of Marine Vehicles and Structures in Waves, pages 91–106, London,
1974. Mechanical Engineering Publications Ltd.
[14] L. P. Graham. Safety Analysis of Swedish Dams: Risk Analysis for the Assessment
and Management of Dam Safety. Licentiate Thesis, Division of Hydraulic Engi-
neering, Department of Civil and Environmental Engineering, Royal Institute of
Technology, 1995.
[16] O. Grim. Oscillation of buoyant two dimensional bodies and the calculation
of the hydrodynamic forms. Technical Report 1171, Hamburgische Schiffbau-
Versuchsanstalb, 1959.
[17] K. Hasselmann and et al. Measurements of wind-wave growth and swell decay
during the joint Nort sea wave project (JONSWAP), volume Reihe A(8◦ ) 12 of
Ergänzungsheft zur Deutschen Hydrographischen Zeitschrift. Deutsches Hydro-
graphisches Institut, Hamburg, 1973.
[19] N. Hogben, N.M.C. Dacunha, and G.F. Olliver. Global Wave Statistics. British
Maritime Technology, Unwin Brothers Ltd., UK, 1986.
[21] ITTC. Recommended procedures and guidelines – testing and extrapolation meth-
ods for responses, sea keeping, sea keeping experiments. Technical report, Interna-
tional Towing Tank Conference, 2005.
[22] J.J. Jensen. Load and Global Response of Ships, volume 4 of Ocean Engineering
Book Series. Elsevier, 2001.
[23] J.J. Jensen, A.E. Mansour, and A.S. Olsen. Estimation of ship motions using closed-
form expressions. Ocean Engineering, 31(61–85), 2004.
[24] J.J. Jensen and T.P. Pedersen. Wave-induced bending moments in ships a quadratic
theory. Transctions of The Royal Institute of Naval Architects, 1978.
[26] B. Kinsman. Wind Waves. Prentice Hall, Englewood Cliffs, N.J., USA, 1965.
214
REFERENCES
[27] B.V. Korvin-Kroukovsky and W.R. Jacobs. Pitching and heaving motions of a ship
in regular waves. Transactions SNAME, 1957.
[30] A.R.J.M. Lloyd. SEAKEEPING: Ship Behaviour in Rough Weather. ELLIS HOR-
WOOD LIMITED, Chichester, England, 1989.
[31] M.S. Longuet-Higgins. Statistical properties of wave groups in a random sea state.
Philosophical Transactions of the Royal Society of London Series A, 312:219, 1984.
[32] S.R. Massel. Ocean Surface Waves: Their Physics and Prediction, volume 11 of
Advanced Series on Ocean Engineering. World Scientific, Singapore, 1996.
[34] D. Mollison. The prediction of device performance. In B. Count, editor, Power from
Sea Waves, pages 135–172. Academic Press, London, 1980.
[35] NC. Ship motions in harbours (in Norwegian. Technical report, Official Committee
of the Nordic Council on Transport Questions, 1986.
[36] NCP. Assessment of ship performance in a seaway. Technical report, The Nordic
Cooperative Project, 1987.
[37] NMD. Regulations of 10 January 1994 for mobile units with technical installations
for production on board (in Norwegian). Technical report, Norwegian Maritime
Directorate, 1994.
[40] PIANC. Criteria for movements of moored ships in harbours, a practical guide. Ptc
ii, report of working group no. 24, supplement to bulletin no. 88, PIANC, 1995.
[41] J.B. Roberts and P.D. Spanos. Random Vibration and Statistical Linearization. John
Wiley and Sons, Chichester, 1990.
[42] H. Rouse, editor. Engineering Hydraulics. Wiley, New York, 5th edition, 1950.
[43] N. Salvesen, E.O. Tuck, and Faltinsen O.M. Ship motions and sea loads. Trans.
SNAME, 78:250–287, 1970.
215
REFERENCES
[45] L. Skjelbreia and J. Hendrickson. Fifth order gravity wave theory. In Proc. 7th
Coastal Engr. Conf., volume 1, pages 184–196, 1961.
[46] M. St. Denis and W. J. Pierson. On the motion of ships in confused seas. Transac-
tions SNAME, 61:280–357, 1953.
[47] W.T. Thompson. Theory of Vibration with Applications. Prentice-Hall Inc., Engle-
wood Cliffs, 1972.
[48] M.J. Tucker, P.G. Cahallenor, and D.J.T. Carter. Numerical simulation of a ran-
dom sea: a common error and its effect upon wave group statistics. Applied Ocean
Research, 6(2), 1984.
[49] U.S. Army Coastal Engineering Research Center. Shore Protection Manual, 1984.
[50] G. van Oortmerssen. The motions of a moored ship in waves. Technical Report 510,
Neth. Ship Model Basin, 1976.
[51] J.H. Vugts. The hydrodynamic coefficients for heaving, swaying and rolling cylin-
ders in a free surface. NSRC report 112S, 2002.
[52] G. Wahl. Wave statistics from swedish coastal waters. In R.E.D. Bishop and W.G.
Price, editors, The Dynamics of Marine Vehicles and Structures in Waves, volume 4,
pages 33–40. Office of Naval Research, and the Royal Institution of Naval Archi-
tects, 1974.
[54] J.M. William. Tables of Progressive Gravity Waves. Pitman Publishing Ltd, 1985.
[55] I.R. Young. Wind Generated Ocean Waves, volume 2. Elsevier Ocean Engineering
Series, 1999.
216