You are on page 1of 29

Ocean Engineering 27 (2000) 625653

www.elsevier.com/locate/oceaneng

Designing against parametric instability in


following seas
K.J. Spyrou
Ship Design Laboratory, Department of Naval Architecture and Marine Engineering, National
Technical University of Athens 9, Iroon Polytechniou, Zografou, Athens 157 73, Greece

Received 14 October 1998; accepted 6 January 1999

Abstract

We investigate the characteristics of parametric instability when very large variations of


restoring between the wave trough and the wave crest are taking place, creating a restoring
that is alternating between negative (or nearly negative) and strongly positive values. The
possible ways to consider the nonlinearities in damping and in restoring are discussed in detail.
The boundary separating parametric instability from pure-loss is identified. In depth studies
are carried out to ascertain the practical relevance of the parametric mechanism. Instability
regions are identified in terms of transient motions, rather than in terms of the customary
asymptotic stability chart associated with Mathieus equation. A basis for comparing para-
metric roll behavior for different representations of restoring is established. Asymmetric vari-
ation laws and bi-chromatic waves are considered. 1999 Elsevier Science Ltd. All rights
reserved.

Keywords: Parametric; Roll; Capsize; Ship; Instability; Mathieu

Nomenclature
a frequency ratio ( 420/2e )
B1, B2 linear and quadratic roll damping coefficients
c1 nondimensional linear damping, B1/[(I I)0]
c2 nondimensional quadratic damping, B2v/(I I)
g acceleration of gravity
(GM) Metacentric height
h amplitude of cyclic variation in restoring
I, I roll moment of inertia and added roll moment of inertia, respectively

0029-8018/00/$ - see front matter 1999 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 2 9 - 8 0 1 8 ( 9 9 ) 0 0 0 1 9 - 0
626 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

k equivalent linear damping factor


M ship mass
p(t) the time-periodic function representing the variation of metacentric
height
R(,t) The time-varying restoring function
t time
z scaled roll angle (z /v)
Greek symbols

nondimensionalized time ( et/2)


roll angle
v angle of vanishing stability
angle of heading in respect to the direction of wave propagation
e frequency of encounter
0 natural frequency of undamped system, 0 (Mg(GM))/(I I)

1. Introduction

It has been known for a long time that in a following seaway the roll-righting-
arm of a ship can become seriously dependent on position relatively to the waves,
Kempf (1938). The simplest possible scenario is when the forward velocity is nearly
constant and the ship can be assumed in quasi-static equilibrium in the vertical direc-
tion. Then the effect of the wave is primarily geometric arising from a consecutive
loss or gain of waterline, combined with vertical shifts of the centre of buoyancy,
Paulling (1961). Some other effects may also be influential, albeit normally to a lesser
extent, such as the motion of water particles and the forward speed, Helas (1982).
The effect of the wave is commonly accounted in the roll equation by a time-
varying restoring term:1
(I I) B1 B2 R(,t) 0 (1)
As was recognized since earlier days, the linearized version of Eq. (1) is Hills-
type:
2k p(t)20 0 (2)
The problem with linearization however, is that it is quite dubious how it should be
carried out: At first instance, since the instability is incurred on the upright equilib-
rium position, it is reasonable to base it on the range of small roll angles around
this equilibrium. This leads to taking only the linear component of damping,
2k B1/(I I), and a linear restoring based on the metacentric height. But from
a capsize perspective, as one has to deal with the build-up of this unstable motion,
the nonlinear components of damping and restoring must be taken into account. Thus

1
All symbols are explained in the nomenclature.
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 627

for preliminary assessment an equivalent linearization method has to be devised in


the hope that a better overall assessment of the ships propensity for capsize could
be obtained. Such a linearization might however be misleading as far as the encounter
frequency and the critical amplitude of (GZ)s fluctuation where the instability is
inflicted are concerned, especially if the slopes of the two alternative linear restorings
are significantly different. Although this effect is of second order, this offers only
little comfort; because for a container recently tested in Japan, the two slopes were
different by a factor higher than 3.0 (Hamamoto et al., 1995). Extra complications
can arise from the nonlinearity of damping.
Another problem of Eq. (2) is that the time-dependent term p(t) is very rarely
exactly cyclic even if, for the sake of simplicity, a simple monochromatic wave had
been considered. As a matter of fact, in practical applications the first few sinusoidal
terms in a Fourier series expansion of p(t) are required. This however creates con-
siderable uncertainty because different hulls may have widely differing laws of
restoring variation; and of course this law cannot be known at an early stage of ship
design. What is well known however is that consideration of only the first harmonic
leads to Mathieus equation whose stability chart as regards the trivial solution
0 has been discussed several times in applied mechanics (McLachlan, 1947;
Bolotin, 1964) as well as in naval architecture (Grim, 1952; Kerwin, 1955; Paulling
and Rosenberg, 1959):
2k 20[1 h cos(et)] 0 (3)
After the substitutions et 2, a 420/2e and z (/), we can obtain an alter-
native form of Eq. (3) to which we shall refer extensively later:
d 2z 2ka dz
a(1 h cos2)z 0 (3)
d2 0 d

If ship rolling is Mathieu-type then there is a possibility of parametric resonance


when a is in the region of n2, with n an integer. It will be manifested by an oscillatory
build-up of roll, despite the absence of direct excitation, Fig. 1. The first two reson-
ances (the principal at e 20 and the fundamental at e 0)2 are usually
thought as the most relevant although in general the importance of the parametric
mechanism for ship rolling has been a matter of controversy. This stems from the
fact that, whilst the domain of the first and to a lesser extent of the second resonance
may, for low damping, extend to relatively low h-levels, this picture corresponds in
fact to asymptotic roll behavior as the considered number of cycles goes to infinity.
Practically however, what is important is whether the instability becomes noticeable
within a small number of encounter wave cycles. When the number of cycles is
small, the build-up of substantial roll requires very intensive parametric variation of
restoring which may be unrealistic. Moreover, whilst at low encounter frequencies

2
Often in the literature the principal is called parametric and the fundamental harmonic. Here however
we shall use the term parametric for all the resonances of the system.
628 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Fig. 1. The development of parametric roll can be paralleled with the movement of the ball in the above
potential well. If the speed of the ball is right, a small initial perturbation off the centre-line of the well
can build-up resulting eventually in escape.

the build-up of roll may require less parametric forcing, for that region even a small
number of cycles normally would demand excessively long time. Lastly, it is not
completely understood how these responses are affected by the nonlinearities of
damping and restoring which, naturally, will become more influential for larger roll.
It would be very interesting to know for example how an initially hardening or
softening (GZ) can influence the arrangement of the capsize regions.
Motivated by the above observations, we have tried to develop a better understand-
ing of the transient roll behavior in regular following seas with realistic restoring
curves and for h near, or even higher, than 1.0. Very little is currently known about
this range of h, beyond of course the obvious fact that capsize is very likely. If
restoring is negative for a part of the wave-cycle, then both pure-loss and parametric
instability could arise depending on the frequency of encounter and the intensity of
restorings variation. In reality pure-loss and parametric resonance, although usually
treated separately, they represent alternative manifestations of instability exhibited
by a Mathieu-type roll model.
The paper addresses the following aspects: After a short review, we present the
most appropriate analytical predictions concerning the first few resonances from the
extensive literature of analytical mechanics. We interface these with some recently
obtained experimental data. But as for large h only little could be expected from
application of analytical methodologies, a numerical approach is also established
where pure-loss and parametric instability are treated in a unified manner. Concepts
such as the equivalent linearized restoring are set under scrutiny. Then a systematic
approach for the consideration of higher-order, and thus more accurate, restoring
curves is put forward while making sure that a basis for meaningful comparisons
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 629

exists. Equivalent cubic, quintic and 7th order restoring polynomials are con-
sidered. Then we investigate the effect of the nonlinearity of damping and we discuss
also from dampings perspective the effectiveness of equivalent linearization. Finally,
two more specialized matters are briefly addressed: the possibility of vertically asym-
metric modulation of (GM) which is taken into account through consideration of a
higher harmonic; and the build-up of roll due to bi-chromatic following waves which
create secondary spikes on each resonance region.

2. Brief background

Whilst the first studies of parametric roll were based on Mathieu equation, in
later years several modeling improvements were introduced: The nonlinearity in
restoring and/or in damping were considered by a number of researchers, including
Blocki (1980); Sanchez and Nayfeh (1990); Kan (1992). It has been shown that
if restoring and/or damping are nonlinear, then stable oscillations may exist. The
nonlinearity may in fact be beneficial, impeding capsize by arresting the growing
oscillation incurred at an instability point. Remarkably, this can happen although the
nonlinear system may possess less restoring at similar angles because the change in
restoring causes de-tuning from resonance.
In some studies simultaneous parametric and direct excitation has been considered
(for example, Kan, 1992), which is akin to a quartering-sea configuration. If direct
roll excitation coexists with the parametric then, as pointed out in Paulling and
Rosenberg (1959) on the basis of earlier work of Rosenberg (1954), this does not
affect the number and position of the resonances. It adds however extra gravity to
the fundamental resonance. As the angle between the ship and the direction of
wave propagation increases, e should increase too since e (2/)(cUcos).
The amplitude of parametric forcing will tend to reduce but direct forcing will
increase with sin . A rather generic roll equation in terms of scaled angle z
(/) for the study of parametric resonance could obtain the following form:
(Ak)2 I
z c1z c2zz R(,z,t) sin cos(et ) (4)
(I I)
Equations such as Eq. (4) are frequently considered in Analytical Mechanics. Minor-
sky (1962) applied the stroboscopic method in order to identify the systems stable
and unstable oscillations analytically, with cubic-type nonlinearity in the time-depen-
dent restoring term (with no direct forcing). A similar problem was solved by Skalak
and Yarymovych (1960); Struble (1963) through, respectively, energy balance and
asymptotic expansion. Solutions for time-independent nonlinearity are presented in
McLachlan (1956); Zavodney et al. (1990). A study including quadratic damping,
and thus quite relevant to the ship problem, was presented by Hsu (1975). Combined
parametric and direct excitation was considered by Troger and Hsu (1977).
A hydrodynamically more rigorous model for the oblique sea, including coupled
motions, was proposed by Boroday (1990), an approach followed also recently by
Nabergoj et al. (1997). Several other studies dealing with practical aspects have also
630 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

been published, see for example Arndt and Roden (1958); Lindeman and Skomedal
(1983); Hua (1992); Neves et al. (1997); Rakchmanin and Vilensky (1997).
The randomness of the wave environment has also been under consideration, with
the parametric forcing seen as a stochastic random process that enters into the
roll equation. By building upon the earlier works of Kozin, Arnold, Ariaratnam,
Stratonovitch and others, a number of authors including Vinje (1976); Haddara
(1980); Muhuri (1980); Roberts (1982); Dunwoody (1989) have sought random-wave
stability criteria for parametric rolling. Random waves have been considered also
from some other perspectives: Price (1975) obtained expectations of the solutions
of the equation of motion; Helas (1982), applied the effective-regular-wave concept
of Grim. The probability of encountering a fairly regular wave group of given charac-
teristics was considered by Blocki (1980). Some comparisons of experiments and
numerical simulations in random seas for a containership were presented by Hamam-
oto and Panjaitan (1997).
Parametric resonance of roll may also arise in head-seas where couplings of hydro-
dynamic nature are more pronounced, offering new interesting possibilities for reson-
ance. It is well-known that the directly excited pitch can create growth of roll, which
will influence back pitch and so forth. A critical consequence can be saturation of
the pitch response and transfer of all the energy that enters the pitch mode into roll,
see for example Nayfeh et al. (1973). Notably, Paulling and Rosenberg (1959) had
discussed earlier a more elementary configuration without waves, with roll under
autoparametric excitation from a prescribed heave. Sethna and Bajaj (1978) had pre-
dicted amplitude modulated motions due to Hopf bifurcations in the averaged equa-
tions of coupled roll and pitch. Nayfeh (1988) has considered the coupled pitch and
roll for 0(pitch)/0 2 and demonstrated the saturation of roll when e 0(pitch)
and the existence of Hopf bifurcations due to nonlinear coupling between these two
modes. Several bifurcations (Hopf, period-doublings, pitchforks and folds) as well
as chaotic motions were identified by Liaw et al. (1992) in their numerical study of
coupled heave-roll of a box-like vessel.

3. Analytical predictions and insights from some recent experiments

3.1. Exact resonances

As is well known, for the undamped system the vertices of the stability transition
curves lie at 20/2e n2/4, n 1, 2, 3, .... For an overtaking following sea the
frequency of encounter is positive and we can write 0/e n/2. The expression of
the vertices in terms of Froude number is the following (see also Fig. 2 for a plot
in terms of the length ratio):

Fn Fnc
L
0
or, Fn
L


0
L
(5)
n 2 n
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 631

Fig. 2. The required combinations of wave length to ship length ratio versus Froude number for exact
(that is, with no damping) resonance. The loci of the first three resonances are shown for three different
natural frequencies (in the brackets the first number is the nondimensional natural roll frequency and the
second the order of the resonance).

where Fnc is the Froude number of the wave celerity and 0 0L/g.
During recent model experiments in Japan two ships were investigated for para-
metric resonance: a container with LBP 150.0 m, 0 0.144 s-1 (0 0.566)
and a purse-seiner with LBP 34.5 m, 0 0.840 s-1 (0 1.577), Umeda et al.
(1995). The (theoretical) Froude numbers of the vertices of the first four resonances
of these vessels appear in Table 1. Seemingly, for the purse-seiner parametric insta-
bility could arise only in connection with the second or possibly a higher resonance.
But, unlike with the container, during the experiments parametric resonance for the
purse-seiner was not observed at all.

3.2. Equivalent linear restoring

The nonlinearity of (GZ) may, at first instance, be taken into account by consider-
ing an averaged linear (GZ) which fulfills the condition that the potential energy at

Table 1
Froude numbers at exact resonance

/L 1.0 /L 2.0

container purse-seiner container purse-seiner

n 1 0.219 0 n 1 0.203 0
n 2 0.309 0.148 n 2 0.387 0.062
n 3 0.339 0.232 n 3 0.444 0.229
n 4 0.359 0.274 n 4 0.479 0.313
632 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

the angle of vanishing stability is the same both for the real and the linearized (GZ).
From Hamamoto and Panjaitan (1996) we can take for the container an equivalent
linear (GM) 0.525 m which leads to 0 1.059. The corresponding resonance
Froude numbers are shown in Table 2 and, as is realized, they are quite different
from those of Table 1. It will be possible to ascertain which of these predictions are
nearer to the truth, through the numerical study of the next section.

3.3. Linearization of damping

The predictions shown in Tables 1 and 2 do not take into account the existence
of damping. Commonly, the linear damping coefficient is derived from the roll
extinction curves by assuming logarithmic decrement of the roll amplitude. As is
well known, this leads to the relationship 2k/0 ( 2/)ln(1 ) or k
(2)/(T0)ln(1 ), where represents the average peak-to-peak roll angle decrement
during a roll-decay test (see for example Kerwin, 1955) and T0 is the natural roll
period. By measuring the effective extinction coefficients ln(1 ) of the two
models (based on averaged values at small angles), Hamamoto et al. (1995) derived
the following damping coefficients: 2k 0.117 for the purse-seiner, (corresponding
to a damping ratio k/0 0.069 and 2k 0.025 for the container. As the
maximum roll velocity from small angles cannot be very high, essentially this corre-
sponds to a measurement of a perturbation of the nondimensional version of B1 of
Eq. (1). These values of damping ratio may be used in combination with the natural
frequency 0 in order to identify critical encounter frequencies where the upright
equilibrium could tend to become unstable. But, it is not exactly appropriate to use
this damping coefficient for a capsize investigation utilizing the equivalent linear
roll equation; because the higher velocity damping component is not taken into
account. A simple way to overcome this inconsistency is the following:
Consider the scaled roll equation based on the equivalent linear restoring of the
previous sub-section, and with quadratic nonlinearity in damping:
z c1z c2zz z 0 (6)
Then, as it is well known, the equivalent linear damping ratio (assuming harmonic
response) is given approximately by:
1 4A
c c (7)
2 1 3 2

Table 2
Froude numbers at exact resonance for the container based on equivalent linear restoring

/L 1.0 /L 2.0

n 1 0.061 0
n 2 0.230 0.227
n 3 0.287 0.339
n 4 0.315 0.395
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 633

where A is the scaled amplitude of the motion. When parametric resonance is real-
ized, the build-up of roll amplitude is gradual from some small initial value. We
may reasonably assume therefore that, on the way to capsize the effect of damping
may be considered on average at some constant amplitude A. What is an appropriate
value for A will be discussed in Section 4.

3.4. The quantitative effects of damping

Consider now Eq. (3) and apply the transformation z w.e( 2ka/0). Then
we obtain:

d 2w
d 2
a 1 k2
20 1
1
1
k2
h cos 2 z 0 (8)
20
The natural frequency has shifted from 0(roll) to 20 k2 and the amplitude of
parametric forcing h is increased to h1 h/(1 k2/20) > h. Thus when damping is
taken into account, the minimum amplitude of parametric forcing that can cause
instability is higher. Within the instability regions the growth rate of the roll angle
will be reduced as the dissipation of energy produces a stabilizing effect. For rela-
tively low forcing and damping we can use directly the well-known low-order
approximations of the transition curves based on the perturbation method, see for
example Nayfeh and Mook (1979). From their expressions (1.23) we can derive,
after some algebra, the following approximations of critical h:

(a 1)2

4k2
principal: hp 2 (9)
a2 a20
5a4 4 (a 4)a2 2 24k2a
f undamental: h hf (a 4)2 0 (10)
8 2 f
2 3 2
2 3 20
The above yields for hf:

4a 4 (a 4)
9 2

225k2a
20

2.3
hf 2 2
(10)
5 a
In Fig. 3 is plotted h versus a for the container on the basis of the above expressions.
It is shown also how the analytical predictions compare with numerical derived with
the method of Jordan and Smith (1977). It is remarkable that the analytical predic-
tions are very good in the left vicinity of each exact resonance.
From Eqs. (9) and (10) it is possible to also work out the approximate forcing
required for instability at 0/e n/2, n 1,2,$
634 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Fig. 3. Various approximations for the domain of instability: next to each curve is indicated the number
of the expression in the text on which it is based (2k 0.025, 0 0.144). Predictions of the principal
and the fundamental resonance are shown also for k 0 according to, respectively, Eqs. (9) and (10).
The dashed lines are the true transition lines, identified numerically. The dots indicate the predictions
based on Eq. (18) which seem most successful.

4k
principal: hpn 1 (11)
0


26 k
f undamental: hfn 2 (12)
4
5 0

Table 3 shows the analytically predicted minima at principal and at fundamental


resonance. However, these do not represent global minima. Because if we allow
k to be a second-order quantity, the minimum of the principal resonance region
shifts at:
4kk2 apmin20 2k2
hpmin where apmin 1 2 (13)
2k 0
2 2
0
A similar expression, concerning the fundamental, can also be derived from Eq. (10).
Note that although these expressions are based on the assumption of low forcing,
the required forcings in Table 3 turned out to be high.

Table 3
Required h for damped motions

Container Purse-seiner

hpn 1 0.347
hfn 2 0.967 hfn 2 0.864
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 635

Approximations of the lower bound of forcing strength h for instability, concerning


a wide range of a, can be derived by several different methods: For example, Taylor
and Narendra (1969) by applying three different stability theorems could arrive
respectively at the following predictions of critical h (given in increasing degree of
sophistication and converted to fit our own analysis):


2k k2
h 1 (14)
0 20
2k
h 1 a (15)
0
k
h a (16)
0

Later, Gunderson et al. (1974) were able to derive a new expression for the stab-
ility boundary [essentially an improvement on Eq. (16)] which could be applied also
for large damping (a 1):

h 1 k2
20
tanh
k
0
a (17)

Eqs. (14)(17) tend to give conservative predictions compared with Eqs. (9) and
(10). Recently, Turyn (1993) has proposed (for k small and a 1), the following
lower bounds for the resonances at a n2 which, on the evidence of Fig. 3, are
almost coincident with the real values of these bounds:


1/n
2
8 k(n!)
h
n2 20

at: a 1 for n 1 and

at a n2
8 k(n!)2
n2 1 20 2/n

k2n2
20
for n 2 (18)

3.5. Nonlinear behavior

Whilst the above are useful in providing information about the stability of the
upright equilibrium position of the ship, they cannot give insights about purely non-
linear aspects which can be encountered when the ship is away from that equilibrium,
such as the stable oscillatory behavior mentioned earlier. In fact it is known that for
a parametric system with nonlinear restoring the curves of stability transition that
we have been discussing so far, represent bifurcation boundaries leading to the oscil-
latory behavior through a sub-critical (left part of each resonance curve) or supercriti-
636 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

cal (right part) pattern (Skalak and Yarymovych, 1960; Soliman and Thompson,
1992). These oscillations, which notably extend even outside the (a, h) domain of
instability of the linear system, at some critical combinations of h and a start losing
their basin of attraction due to interwiggling of their manifolds. Expressions for the
critical (a, h) combinations are available for low damping by applying the so-called
Melnikov analysis. This method provides a criterion of stability concerning the finite-
amplitude parametric rolling and is therefore less stringent than a criterion concerning
the upright equilibrium. Such an analysis has been carried out, for example, by Kan
(1992); Esparza and Falzarano (1993). According to the Melnikov criterion, for a
softening, time-independent cubic nonlinearity in restoring the critical parametric
forcing is:

h
2 k
3 0
a sinh 2 a
2
(19)

4. Numerical studies

4.1. General procedure

The development of significant roll often requires a very large number of wave
cycles when e is high, or unrealistically long time when e is low. The transition
curves are relevant for the system under the assumption that the two critical para-
meters, namely number of wave cycles and available time, are infinite. However in
reality such an environment could hardly be representative for a ship. Moreover, it
is obvious that h cannot be increased indefinitely because it is constrained by the
maximum slope of the true water waves and beyond that limit ship behavior would
not be amenable to the present analysis. So given a hull, there is an associated
limiting h. Still however, from h alone it is not possible to deduce the relevance of
the higher-order resonances. It is essential to examine also the required number of
cycles as well as the time requirement for reaching capsize levels and then to consider
whether these can be feasible.
In order to investigate these matters in detail, a numerical scheme was
implemented based on direct integration of the roll equation. More specifically,
given:
a certain initial roll angle and velocity at t 0 (z0, z0),
a phase angle which determines the initial position on the wave (the smaller
the number of wave cycles, the more influential becomes); 0 is at the
wave crest,
a pair (a, h) obtained consecutively from a dense grid of the plane defined by
these two parameters,
we examine whether the nondimensional roll angle exceeds a critical nondimensional
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 637

roll angle zlim either within m wave cycles, or within a certain amount of time, t.
We shall begin our investigation with a very simplistic linear equation of roll and
we shall introduce subsequently a number of considerations which will tend to make
the analysis progressively more realistic.

4.2. Number of wave cycles and time

In Fig. 4 we have identified the amplitude of forcing h required for a 100-fold


increase of the roll angle from its initial value z0 0.01 (with z0 0, 0), in
respect to the first four resonances, as the number of wave cycles in increased and
2k/0 0.025/0.144. Thus we have set3 zlim 1.0. The principal resonance appears
practically irrelevant for the 4-cycle scenario because the parametric forcing h
required is too high. But the level of h becomes more feasible as m is increased.
The effect of a larger initial heel, z0 0.1 (thus with capsize being now equivalent
with a 10-fold increase), is shown in Fig. 5. With four cycles allowed, the vertices
of the first two resonances are now at approximately the same h-level (h 1.45)
An interesting feature is the existence of disconnected domains, most obvious in the
third resonant region. Simulations in and around the survival wedge are shown in
the middle graph. Such disconnected domains exist when the number of wave cycles
m is small to moderate and the simulations are started from a point of the wave
where the ship has positive (GM). The process of creation of disconnected domains
in the second resonance region is shown in the lower graph. It is notable that a line

Fig. 4. The relationship between number of wave cycles and parametric forcing at the vertices of reson-
ances 1, 2, 3 and 4.

3
For a linear Mathieu equation capsize does not exist as such and one has to set a range of admissible
roll angles. On the other hand if the equation is nonlinear, the vanishing angle is not the edge of capsize
since this depends also on the initial velocity. However it is still sensible to define the range of stability
on the basis of this angle. Of course in a more practical context the largest acceptable heel can be much
below the vanishing angle.
638 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Fig. 5. Capsize regions in respect of the first six resonances: The upper graph shows these for 4 and
32 wave cycles. Simulations were started from the trough with z0 0.1; 2k/0 0.025/0.144. In the
middle graph are shown simulations around the survival pocket of the third resonance region: h
1.64, a 12, ...... h 1.64, a 11.5, --- h 1.64, a 11. In the lower graph is depicted the
process of creation of disconnected domains as the number of wave cycles becomes lower. The curves
shown correspond to m 24, 16, 12 and 8 (the curves for m 4 and 32 appear in the upper graph).
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 639

seems to exist which depends on but not on m, and which runs across the reson-
ances on the plane (a, h) dividing them into subdomains. For small m, the regions
of capsize which correspond to starts from the trough and starts from the crest are
significantly different, with the second containing the first. Although these domains
tend to come nearer as m is increased, this happens however very slowly, see Table 4.
We should note here that, unlike the standard stability chart of Mathieus equation
(Strutt diagram), we prefer to plot a 420/2e versus h rather than a versus 2q
ah. Superficial reading of the latter could be misleading because, as is well known,
for the higher order resonances 2q should be very large for instability. Also, for
a constant 2q the corresponding unstable domains become progressively thinner.
However, as e approaches zero, a 420/2e becomes very large and thus contrib-
utes substantially in 2q. From Figs. 4 and 5 it is deduced that the minimal h required
at the higher-order resonances may not be unrealistically high. Should the value of
equivalent damping had been lower, the resonances could have their vertices below
the h 1 level. One could infer further that, the curve that connects these mimima
should tend to 1.0 as e 0 and thus 0/e. This is because the condition
e 0 represents the static case where the ship runs with the waves phase velocity
and the marginal value of the parametric amplitude h that is required for instability
is apparently 1.0 in order to achieve exactly zero restoring at the crest.4 In other
words, the minimal h for pure-loss cannot be different from 1.0.
A study based solely on the number of wave cycles is bound to be of little practical
value because at low frequency of encounter it will take very considerable time in
order to advance, say, from trough to crest; and during this time, it is unlikely that
the characteristics of the wave could in practice be preserved. It is vital therefore to
examine whether the prescribed capsizal angle is reached within a reasonable amount
of time t. Several sub-domains of the capsize domain that corresponds to m 8 are

Table 4
Effect of number of cycles on the required minimal h for the first two regions of resonance

Number
of cycles: m 4 m8 m 12 m 16

Order of
resonance: 1st 2nd 1st 2nd 1st 2nd 1st 2nd

Initially
on crest: 1.28 1.27 0.81 1.06 0.645 0.985 0.565 0.95
Initially
on trough 1.445 1.46 0.915 1.14 0.73 1.05 0.64 0.99

4
This argument has to be in fact more complicated since for e 0 it depends on the position on the
wave whether capsize will occur for a certain h 1.0.
640 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Fig. 6. The time required in order to reach capsize levels is an important parameter which should be
considered along with the number of wave cycles and the amplitude of restorings variation. This is
investigated here considering up to 8 cycles (z0 0.1). 1, 25 t 30; 2, 30 t 35; 3, 35 t
40; 4, 40 t 45; 5, 45 t 50; 6, 50 t 200; 7, 200 t 400; 8, t > 400.

indicated in Fig. 6. Of course, when more time is allowed the relevant boundary
extends towards lower a and lower h. The boundary of the high-e region where
the time requirement is minimal should be regarded however as the most practically
important. For the above investigation the simulations were started from the point
of the up-slope of the wave where restoring is zero, as the ship enters into a negative-
restoring region. The phase that corresponds to this point can be found from the
relation 1 h cos(2T ) 0 which yields arccos(1/h).

4.3. More realistic representations of restoring

Consider once more the container of the previous Section. Several points of its
(GZ) curve are shown in Fig. 7. We shall attempt to develop polynomial approxi-

Fig. 7. Equivalent (GZ) curves of various orders. The dots correspond to the true (GZ) of the container.
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 641

mations of this (GZ) which are as near to it as possible; and the area
0

R()d,
where R() is the true (dimensional) restoring, is the same. For this we shall consider
a family of odd-power polynomials of z which have unit slope at the origin and
become zero at z 0, 1:
Q1(z) z Q3(z) z z3
Q5(z) z z3 (1 )z5 Q7(z) z 1z3 2z5 ( 1 1 2)z7
As evidenced from Fig. 7, the seventh-order polynomial with 1 27.5 and 2
49.5 is a very good approximation. With the above 1, 2 the normalised area under
the curve, F7, is:


1
z2 z4 z6 z8 1
F7 Q7(z)dz 1 2 ( 1 1 2)
2 4 6 8 0
0

9 31 2
1.75.
24

It can easily be derived that the dimensional value of this area, which is basically
the potential energy at , is given by Mg(GM)72F7. Equating these potential ener-
gies in respect to a 7th-order and to a quintic polynomial, while utilising the same
(GM) 0.15 m for both (this is of course one of the possible options), we obtain:
(1 31 2)/2 17. The quintic curve, shown in Fig. 7, was drawn with
this .
For the equivalent cubic and linear (GZ) (indicated by subscripts 3 and 1
respectively) we obtain by following a similar procedure:

4.3.1. Cubic
F7
Mg(GM)32F3 Mg(GM)72F7 or, (GM)3 (GM)7
F3
Normalized (GZ) area:


1
z2 z4 1
F3 Q3(z)dz 0.25.
2 4 0
0

Therefore
1.75
(GM)3 0.15 1.05 m and 03 0.381 s1.
0.25
642 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

4.3.2. Linear
Normalized area:


1
z2 1
F1 Q1(z)dz 0.5
2 0
0

Therefore, (GM)1 0.15 (1.75)/(0.5) 0.525 and 01 0.269 s-1. These are
basically the values (GM) and 0 that were used in Section 3.

4.4. Interpretation of critical h

There are many possible ways that a ships restoring can be changing due to the
waves and unfortunately it is not possible to find a generic method which may be
applicable to all ships. The simplest option however is to consider a cyclic variation
of its linear term while treating the higher-order terms as time-independent. Thus
R(z,) (1 h cos 2)z o(z) where o(z) represents the higher-order terms of the
normalized still-water restoring function. If restoring is of linear or cubic-type, then
at h 1 it will turn negative for all positive z. Obviously, this will not be the case
if the restoring curve is initially hardening. Take for example the quintic curve: the
critical h to have negative restoring for all positive z can be found from the condition:
(1 h)z z3 (1 )z5 0, z 0 or, (1 )z4 z2 (1 h) 0
This is satisfied when the discriminant is negative: 2 4(1 h)(1 ) 0. Thus
the critical h is 1 2/[4(1 )]. For 17 this yields h 5.014. It can also
be shown that the critical h for the 7th-order curve is h 5.5, Fig. 8.
Another important, and perhaps more relevant, condition is when the signed (GZ)
area up to becomes zero, in reference to a ship that is sitting on a crest. We
shall find how this condition reflects in terms of h for the different types of (GZ):

4.4.1. Linear
It is easily seen that h 1.

4.4.2. Cubic


1
1 1
[(1 h)z z3]dz 0(1 h) 0h 0.5.
2 4
0

4.4.3. Quintic


1
1h 1 4
[(1 h)z z3 (1 )z5]dz 0 0h
2 4 6 6
0
17 yields h 3.5
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 643

4.4.4. 7th order


1
1 h 1
[(1 h)z 1z3 2z5 ( 1 1 2)z7]dz 0
2 4
0

2 ( 1 1 2) 9 31 2
0h
6 8 12
For 1 27.5 and 2 49.5 we obtain again h 3.5.

4.5. Linear versus equivalent linear restoring

Capsize domains based on linear [(GM) 0.15 m, dashed line] and equivalent
linear [(GM) 0.525 m] restoring, are compared in Fig. 9 for the same damping,

Fig. 8. Effect of magnitude of h on 7th order (GZ) from trough to crest: - - - - calm water (GZ);
h 3.0; --- h 5.5.

Fig. 9. Comparison of capsize domains for linear and linearized restorings (z0 0.01): 0 0.144
s1 (dashed); 0 0.269 s1.
644 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

2k 0.025 s-1. In addition to the expected shift of the resonances towards lower
values of a, we note also a seemingly counterintuitive result: the minimal h tends
to become lower and thus capsize easier, in spite of the stronger restoring.

4.6. Cubic restoring

Consider the nonlinear restoring of cubic type with (GM)3 1.05 m which was
shown to yield 03 0.381 s-1. The corresponding 8-cycle capsize regions, with
initial heel z0 0.01, can be seen in Fig. 10. Using the true (GM) of the container
(0.15 m) but with z0 0.1, results in Fig. 11. This figure shows also, in a quite
unique combined view, the domain of pure-loss capsize (white upper-right area).
But from a design perspective, Fig. 12 is perhaps the most interesting since it summa-
rizes the relationship between number of wave cycles, damping and minimal required
forcing, in respect to the principal (upper) and the fundamental (lower) resonance.
Covering a wide range of damping ratios and based on a rather generic nonlinear
restoring, these figures can provide useful guidance for the early design stage.

4.7. Quintic and 7th order

These two, and especially the 7th order, are the most realistic representations of
the containers restoring. Fig. 13 shows the 8-cycle capsize domain based on the
7th-order. A first observation is that the resonance at a 1, predicted on the basis
of the true (GM) 0.15 m, is still the one representing the greatest threat to cause
capsize as it requires the minimal h. We note also the irregularly-looking boundaries
for higher a whose nature was found to be quite insensitive to the number of wave
cycles. The capsize regions show little resemblance to those obtained with the linear
or cubic restoring.

Fig. 10. Capsize domains for equivalent cubic restoring (m 8; z0 0.01). For comparison, when
z0 0.1 the required h for the first three resonances is respectively (approximately): 0.8, 0.8 and 0.97.
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 645

Fig. 11. As Fig. 10 but with 2k/0 0.025/0.144 and z0 0.1. Capsize in the less dark regions required
t > 300 s. The white upper-right area represents capsize events that occurred very quickly and correspond
to the pure-loss type.

As the above phenomena must relate with the initially hardening nature of restor-
ing, we have tried to find how the boundaries are transformed into fractal-like by
varying the intensity of hardening. For simplicity, we based this investigation on the
-parameterized family of quintic restoring curves which display, on this aspect,
very similar behavior with the 7th-order family. Starting from 1 (where the
quintic polynomial degenerates into cubic) we set gradually higher and we observe
the change of the capsize boundary. For each considered we make sure that the
area under (GZ) is kept constant. From this condition we can identify the correspond-
ing (GM)5 and then the appropriate value of natural frequency 05. The process of
transformation of the boundary is portrayed through the sequence of graphs of Fig.
14 (include also Fig. 10 which shows the degenerate case, when 1): For each
of the resonances it is the right-side of the corresponding boundary which becomes
brittle first. Gradually, with increasing (i.e. more hardening initial restoring) these
regions are enlarged while the original spikes degenerate and are lifted towards
higher h. Eventually the whole capsize boundary becomes highly irregular, although
some picks, still existing at low a, remind the initial resonance mechanism.

4.8. Nonlinearity of damping

We shall compare now the numerical predictions of the resonances based on equiv-
alent linear and on fully nonlinear damping in order to assess the workability of the
idea of linearization. To make the comparisons more straightforward, we have used
in both cases the equivalent linear restoring where 0 0.269 s1. Based on Eq. (7),
the equivalent linear damping will be found from the relationship
k k 4A 0c2/3, where k represents now only the linear damping term. It is
646 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Fig. 12. The h-level at the vertex of the principal (upper graph) and the fundamental (lower) capsize
region for nonlinear restoring and various levels of damping, the parameter of the curves is k/0; z0 0.1.

evident that, in order to calculate k an appropriate amplitude of roll oscillation A


must be specified first. To identify what is an appropriate value for this amplitude
in a rigorous manner is not easy. By trying different values we have found that the
amplitude A 1/4 results in good agreement between the predictions of the two
methods (Table 5). For such a value of A, the pattern is that linearization of damping
results in slightly lower prediction of h for the principal resonance but in slightly
higher h for the others.
We have repeated the calculations associated with Fig. 14 for relatively high quad-
ratic damping, c2 0.25. The main conclusion is that, in addition to the expected
shift of the capsize region upwards, in terms of h, the qualitative character of the
capsize boundary remains essentially unchanged.
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 647

Fig. 13. The initially hardening (GZ) created a highly irregular capsize boundary (m 8; z 0.01;
(2k)/(0) (0.025)/(0.144); 7th order polynomial).

4.9. Effect of higher-order harmonics

As mentioned in the Introduction, the assumption of cyclically varying (GM) is


highly idealized and, in general, the first few harmonics of the Fourier expansion of
the periodic function p(t) should normally be present in the restoring function. Whilst
the type of modulation can vary widely depending on the hull geometry, a feature
that is often encountered in practice is an asymmetric pattern, where the loss of
(GM) nearer to the crest is less than the gain nearer to the trough (Chou et al., 1974;
Hamamoto and Panjaitan, 1996). This means that the average (GM) would be higher
than the calm-sea value. Therefore a natural frequency based on the average (GM)
will be higher than the still-water natural frequency and the resonances will be shifted
towards lower values of a. Assuming for simplicity that the average (GM) appears
at the nodes of the wave, the modulation of (GM) will be now a[1 - h cos(2)],
where is the parameter accounting for the higher average (GM). This may also be
written as: a[1 h cos(2)] with a (1 )a and h (h)/(1 ). Therefore,
not only the resonances will shift, but also, the forcing will be reduced.
Another interesting feature, linked with the existence of asymmetry, is that the
troughs in (GM)s modulation are often wider than the crests. To take this into
account in a simple way, let us consider linear restoring and a second harmonic:
a[1 h cos(2) fh cos(4)] where f is the coefficient representing the strength
of the second harmonic. The effect is shown in Fig. 15 (compare with Fig. 5). Whilst
the first resonance is not seriously affected (but we note that its vertex tends to be
slightly lower), the higher resonances move towards considerably higher h-values.
This effect seems to make even more unlikely the occurrence in practice of a reson-
ance above the principal.
648 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Fig. 14. The process of creation of irregular capsize boundaries for the quintic restoring as is
increased.

Table 5
Required h for the first three resonances, 2k 0.025

Equivalent linear damping Linear quadratic

c2 2k 1st 2nd 3rd 1st 2nd 3rd

0.1 0.0307 1.07 1.125 1.145 1.07 1.12 1.135


0.15 0.03356 1.09 1.145 1.165 1.095 1.13 1.145
0.25 0.03927 1.13 1.185 1.205 1.145 1.16 1.17
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 649

Fig. 15. Asymmetric variation of restoring with consideration of the second harmonic whose strength
is assumed f 0.25. In the upper graph is shown the degree of difference in the variation of (GM) from
crest to trough between the sinusoidal symmetric (dashed line) and the considered asymmetric. In the
lower graph is shown the corresponding capsize diagram which notably remains qualitatively unchanged.
A very large number of wave cycles was considered (m 32). Also, z0 0.1 and other parameters
unchanged.

4.10. Bi-chromatic waves

A matter that had been touched upon by Chou et al. (1974) but not addressed in
detail so far, is the behavior of a ship under the effect of a wave group containing
at least two independent frequencies. In very simplistic terms, this could bring about
a quasiperiodically changing restoring. Such an effect may easily be accounted by
our method and two examples are shown in Fig. 16. In Fig. 16a the two frequencies
were selected to be near to each other. In Fig. 16b on the other hand, they were far
apart. A common characteristic is that a number of new spikes have grown on
each primary resonant spike and the number of these secondary spikes tends to
increase the further the second frequency departs from the first. Without any doubt,
more research will be required in order to understand completely these phenomena.
650 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Fig. 16. (a) Effect of quasi-periodic parametric forcing which may arise if a bi-chromatic wave group
is encountered (m 32, z0 0.1). In the upper graph is shown the pattern of variation of the time-
dependent term for h 1.5. (b) Whilst in Fig. 16a the two frequencies were considered near to each
other, here is shown the effect when the frequencies are far apart.

We should note that some aspects of the asymptotic stability of the undamped
Mathieu equation under quasi-periodic forcing were studied recently by Rand et
al. (1997).

5. Concluding remarks

We have analyzed parametric instability for various types of linear and nonlinear
restoring and we have discussed how to take into account the nonlinearity of damp-
ing. We have argued that parametric instability should be examined in a transient
sense; and practically there is no reason to be seen independently from the pure-
loss mechanism. A successful design objective against capsize should be addressing
the combination of at least the following four main factors: The required amplitude
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 651

of parametric forcing for capsize, which has an upper limit determined, for a certain
design wave, from the hull shape; the available damping which can be significantly
increased by installing extra fins or bilge-keels; and also, the considered number of
wave cycles together with the physical time that is allowed for the capsize, both
aspects being characteristic of the ocean environment rather than of the ship.
In the current literature on parametric resonance, there is a tendency to rely on
simple, Mathieu-type, harmonic modulation of restoring or alternatively to consider
it as a certain random process. It is very doubted whether meaningful conclusions
can be drawn from such investigations since the actual behavior of the ship can vary
widely depending on the nonlinear restoring characteristics and the way they vary
at different positions of a wave.
Fundamental studies based on generic representations of restoring and damping
can provide useful insights about the origins and the character of the instability. For
example, we have now a clearer picture about the effect of nonlinearity and we have
taken note of the considerable difference in behavior between an initially hardening
and a softening restoring curve. However, although these studies bring us nearer to
the truth, they still suffer from the luck of an effective link with the characteristics
of the hull. This means that at this stage the designer, knowing only the link between
restoring/damping characteristics and parametric instability, he faces the task of hav-
ing to find his own interpretations about appropriate hull shapes which minimize the
tendency for capsize. The task of associating families of hulls with capsize bound-
aries is of course a much more ambitious one. But with todays means it seems
realizable and it is hoped that some research will appear in this direction in the future.

References
Arndt, B., Roden, S., 1958 Stabilitat bei vor- und achterlichen Seegang, Schiffstechnik, Bd. 5, Heft 29,
192199.
Blocki, W., 1980. Ship safety in connection with parametric resonance of the roll. International Shipbuild-
ing Progress 27, 3653.
Bolotin, V.V., 1964. The Dynamic Stability of Elastic Systems. Holden-Day Inc., San Francisco.
Boroday, I.K., 1990. Ship stability in waves: On the problem of righting moment estimations for ships
in oblique waves, Proceedings of the 4th International Conference on Stability of Ships and Ocean
Vehicles, STAB 90, September 2428, Naples, Vol II, pp. 441451.
Chou, S.J., Oakley, O.H., Paulling, J.R., Slyke, R..V., Wood, P.D., Zink, P.F., 1974. Ship motions and
capsizing in astern seas. Final report, Department of Naval Architecture, University of California,
Berkeley. Prepared for the Department of Transportation, U.S. Coast Guard, Washington, D.C., Con-
tract DOT-CG-84, 549-A.
Dunwoody, A.B., 1989. Roll of a ship in astern seas-response to GM fluctuations. Journal of Ship Research
33 (4), 284290.
Esparza, I., Falzarano, J.M., 1993. Nonlinear rolling motion of a statically biased ship under the effect
of external and parametric excitations. Proceedings of the Symposium on Dynamics and Vibration of
Time-Varying Systems, OE, 56, ASME, 111122.
Grim, O., 1952. Rollschwingungen, Stabilitat und Sicherheit im Seegang. Schiffstechnik 1 (1), 1021.
Gunderson, H., Rigas, H., Van Vleck, F.S., 1974. A technique for determining stability regions for the
damped Mathieu equation. SIAM J. Appl. Math. 26 (2), 345349.
Haddara, M.R., 1980. On the parametric excitation of nonlinear rolling motion in random seas. Inter-
national Shipbuilding Progress 27 (315), 290293.
652 K.J. Spyrou / Ocean Engineering 27 (2000) 625653

Hamamoto, M., Umeda, N., Matsuda, A., Sera, W., 1995. Analyses of low cycle resonance of a ship in
astern seas. Journal of the Society of Naval Architects of Japan 177, 197206.
Hamamoto, M., Panjaitan, J.P., 1996. Analysis on parametric resonance of ships in astern seas, Proceed-
ings of the 2nd Workshop on Stability and Operational Safety of Ships, Osaka, November 1819, pp.
3646.
Hamamoto, M., Panjaitan, J.P., 1997. A probabilistic approach to parametric resonance of a ship in random
astern seas, Proceedings of the 3rd International Workshop on Theoretical Advances in Ship Stability
and Practical Impact, Crete, October 2829.
Helas, G., 1982. Intact stability of ships in following waves, Proceedings of the Second International
Conference on Stability of Ships and Ocean Vehicles, Tokyo, pp. 689700.
Hsu, C.S., 1975. Limit cycle oscillations of parametrically excited second-order nonlinear system. Journal
of Applied Mechanics 42, 176182.
Hua, J., 1992. A study of the parametrically excited roll motion of a RO-RO ship in following and heading
waves. International Shipbuilding Progress 39 (420), 345366.
Jordan, D.W., Smith, P., 1977. Nonlinear Ordinary Differential Equations. Clarendon Press, Oxford.
Kan, M., 1992. Chaotic capsizing, Proceedings of the 20th ITTC Seakeeping Committee, September 10
11, Osaka, pp. 155180.
Kempf, G., 1938. Die Stabilitatbeanspruchung der Schiffe durch Wellen und Schwinungen, Werftreederei-
Hafen, 19, p. 202.
Kerwin, J.E., 1955. Notes on rolling in longitudinal waves. International Shipbuilding Progress 2 (16),
597614.
Liaw, C.Y., Bishop, S.R., Thompson, J.M.T., 1992. Heave-excited rolling motion of a rectangular vessel
in head seas, Proceedings of the 2nd International Offshore and Polar Engineering Conference, San
Francisco, June, Vol. III, pp. 615622.
Lindeman, K., Skomedal, N., 1983. Modern hullforms and parametric excitation of the roll motion.
Norwegian Maritime Research 13 (2), 220.
Minorsky, N., 1962. Nonlinear Oscillations. D. Van Nostrand Company, Inc., Princeton, New Jersey.
McLachlan, N.W., 1947. Theory and Application of Mathieu Functions. Clarendon Press, Oxford.
McLachlan, N.W., 1956. Ordinary Nonlinear Differential Equations in Engineering and Physical Sciences.
Clarendon Press, Oxford.
Muhuri, P.K., 1980. A study of the stability of the rolling motion of a ship in an irregular seaway.
International Shipbuilding Progress 27, 139142.
Nabergoj, R., Obreja, D.C., Trincas, G., Crudu, L., Stoicescu, L., 1997. Dynamic transverse stability
in longitudinal waves: Theoretical and experimental researches, Proceedings of the 6th International
Conference on Stability of Ships and Ocean Vehicles, STAB 97, September 2227, Varna, Vol. I,
pp. 2943.
Nayfeh, A.H., Mook, D.T., Marshall, L.R., 1973. Nonlinear coupling of pitch and roll modes in ship
motions. Journal of Hydronautics 7 (4), 145152.
Nayfeh, A.H., Mook, D.T., 1979. Nonlinear Oscillations. John Wiley and Sons, New York.
Nayfeh, A.H., 1988. Undesirable roll characteristics of ships in regular waves. Journal of Ship Research
32 (2), 92100.
Neves, M.A.S., Salas, M., Valerio, L., 1997. An investigation on the influence of stern hull shape on the
roll motion and stability of small fishing vessels, Proceedings of the 6th International Conference on
Stability of Ships and Ocean Vehicles, STAB97, Varna, Vol. I, pp. 259269.
Paulling, J.R., Rosenberg, R.M., 1959. On unstable ship motions resulting from nonlinear coupling. Jour-
nal of Ship Research 2, 3646.
Paulling, J.R., 1961. The transverse stability of a ship in a longitudinal seaway. Journal of Ship Research
4, 3749.
Price, W.G., 1975. A stability analysis of the roll motion of a ship in an irregular seaway. International
Shipbuilding Progress 22 (247), 103112.
Rakchmanin, N., Vilensky, G., 1997. Ship crankiness and stability regulation, Proceedings of the 3rd
International Workshop on Theoretical Advances in Intact Stability and Practical Impact, Crete,
October 2829.
Rand, R.A., Zounes, R., Hastings, R., 1997. Dynamics of a quasi-periodically forced Mathieu oscillator,
K.J. Spyrou / Ocean Engineering 27 (2000) 625653 653

Proceedings of the IUTAM CHAOS 97, Applications of Nonlinear and Chaotic Dynamics in Mech-
anics, July, Cornell University, IthacaPublished by Kluwer Acedemic Publishers, Holland (ed.
F.C. Moon).
Roberts, J.B., 1982. Effect of parametric excitation on ship rolling motion in random waves. Journal of
Ship Research 26 (4), 246253.
Rosenberg, R.M., 1954. On the stability of non-linear non-autonomous systems, Proceedings of the 2nd
U.S. National Congress of Applied Mechanics. Ann Arbor, Michigan.
Sanchez, N.E., Nayfeh, A.H., 1990. Nonlinear rolling motions of ships in longitudinal waves. International
Shipbuilding Progress 37 (441), 247272.
Sethna, P.R., Bajaj, A.K., 1978. Bifurcations in dynamical systems with internal resonance. ASME Journal
of Applied Mechanics 45 (4), 895902.
Skalak, R., Yarymovych, M.I., 1960. Subharmonic oscillations of a pendulum. Journal of Applied Mech-
anics 27, 159164.
Soliman, M.S., Thompson, J.M.T., 1992. Indeterminate sub-critical bifurcations in parametric resonance.
Proceedings of the Royal Society of London, A 438, 511518.
Struble, R.A., 1963. On the subharmonic oscillations of a pendulum. Journal of Applied Mechanics 30,
301303.
Taylor, J.H., Narenbra, K.S., 1969. Stability regions for the damped Mathieu equation. SIAM J. Appli.
Math. 17 (2), 343352.
Troger, H., Hsu, C.S., 1977. Response of a nonlinear system under combined parametric and forcing
excitations. Journal of Applied Mechanics 44, 179181.
Turyn, L., 1993. The damped Mathieu equation. Quarterly of Applied Mathematics LI (2), 389398.
Umeda, N., Hamamoto, M., Takaishi, Y., Chiba, Y., Matsuda, A., Sera, W., Suzuki, S., Spyrou, K.,
Watanabe, K., 1995. Model experiments of ship capsize in astern seas. Journal of the Society of Naval
Architects of Japan 177, 207217.
Vinje, T., 1976. On stability of ships in irregular following sea. Norwegian Maritime Research 4 (2),
1519.
Zavodney, L.D., Nayfeh, A.H., Sanchez, N.E., 1990. Bifurcations and chaos in parametrically excited
single-degree-of-freedom systems. Nonlinear Dynamics 1 (1), 121.

You might also like