Sie sind auf Seite 1von 12

Ocean Engineering 76 (2014) 40–51

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

Time domain modeling of a dynamic impact oscillator under


wave excitations
Mingsheng Chen a, Rodney Eatock Taylor a,b, Yoo Sang Choo a,n
a
Centre for Offshore Research & Engineering, Department of Civil & Environmental Engineering, National University of Singapore,
1 Engineering Drive 2, Singapore 117576, Singapore
b
Department of Engineering Science, University of Oxford, Oxford OX1 3PJ, UK

art ic l e i nf o a b s t r a c t

Article history: This paper establishes a methodology for analyzing the dynamics of a wave-induced impact model, with
Received 17 March 2013 emphasis on the modeling of float-over installations. The time domain model described by the Cummins
Accepted 16 October 2013 equation provides an attractive way of analyzing the dynamics of marine structures with nonlinear
Available online 18 December 2013
effects. By replacing the time-consuming convolution terms, the resulting model is very efficient in
Keywords: dealing with nonlinear problems. The established time domain model is applied to investigate Leg
Float-over Mating Unit (LMU) impacts during a float-over operation by considering the heaving motions of the
Wave-induced impact whole system. Both a one-body system (considering that barge and deck move as one rigid body) and a
Time domain model two-body system (barge and deck moving separately) are considered in this paper. The techniques of
Cummins equation
impact maps, Poincaré maps, bifurcation diagrams and phase portraits are used to investigate the motion
State-space model
characteristics of the barge-deck system undergoing vertical impacts with the substructure.
Bifurcation diagram
Impact map & 2013 Published by Elsevier Ltd.
Poincaré map

1. Introduction installing a large deck onto an offshore platform, due to its


relatively lower operational cost and higher installation capacity
Marine operations generally involve nonlinearities such as a (Tahar et al., 2006). By this method, after being constructed at the
nonlinear mooring system, wave-induced impacts during a float- shipyard the deck is loaded onto a transportation barge supported
over installation, and viscous forces. Due to these nonlinearities, on the Deck Support Frame (DSF) using Deck Support Units (DSUs),
the dynamics of marine structures may exhibit sub-harmonic and then towed to the installation site. The deck is then trans-
motion and chaotic motion, under environmental loads. One ferred onto the pre-installed substructure by controlling the
typical example is the dynamics of an articulated tower, that has ballasting of the barge when the prevailing sea state is suitable.
been modeled as a forced bilinear oscillator with a discontinuity in In the load transfer stage, when the system is excited by waves,
the stiffness of the system due to slackening of the mooring cables, the mating cones attached on the deck legs will make intermittent
see, for example, Thompson et al. (1983, 1984) and Gottlieb et al. impact with the receptors at the top of the substructure legs.
(1992). Another example is Virgin and Bishop's (1988) investiga- These impact loads are dissipated by the Leg Mating Units (LMUs)
tion of the complex dynamics of a floating semi-submersible with that are pre-installed within the receptors. In order to ensure an
nonlinear mooring system in the time domain, based on a non- efficient and safe deck installation, it is critical to predict the
linear oscillator model with a constant damping for varying resulting forces and dynamics of the barge-deck system with high
periods. They analyzed the phenomena of period-doubling bifur- confidence.
cations leading to chaos. However, the effects of the frequency The nonlinear dynamics of wave-induced LMU impacts cannot
dependent radiation damping were not well addressed in any of be evaluated by the conventional frequency domain model based
these time domain models, since only a constant damping term on a linear potential flow approach (Chen et al., 2012). Thus, time
was used. domain modeling of the wave-induced LMU impacts is pursued
Another typical nonlinear marine system is the model of wave- here. The Cummins (1962) equation provides an attractive way of
induced impact arising in the installation of an offshore platform analyzing the dynamics of marine structures in the time domain.
by the float-over method. This method is gaining popularity for The equation leads to a linear time invariant framework relating the
motion of the structure to the wave excitations on it. By replacing
the time-consuming convolution term with a state-space model,
n
Corresponding author. Tel.: þ 65 6516 2994; fax: þ 65 6779 1635. one obtains a constant parameter time domain model based on the
E-mail address: ceecys@nus.edu.sg (Y.S. Choo). frequency domain results. Much literature has been devoted to the

0029-8018/$ - see front matter & 2013 Published by Elsevier Ltd.


http://dx.doi.org/10.1016/j.oceaneng.2013.10.004
M. Chen et al. / Ocean Engineering 76 (2014) 40–51 41

replacement of the convolutions by a state-space model—see, for zero forward speed, the equation for a floating rigid body has the
example, Yu and Falnes (1995), Kristiansen et al. (2005) and following form
Taghipour et al. (2008). Compared to the frequency domain model, Z t
½Μ þAð1Þx€ ðtÞ þ hðt  τÞx_ ðτÞdτ þ KxðtÞ ¼ f ðtÞ
exc
a time domain model has the advantage of analyzing nonlinear ð1Þ
0
problems, providing the hydrodynamics remain linear. Any non-
linear effects can be added into the model as additional loads (Wu where, x is the vector of the six degrees of freedom; M is the 6  6
and Moan, 1996). In the authors’ previous paper (Chen et al., 2012), mass matrix; A (1) is a constant positive-definite matrix that is
the time domain model described by the Cummins equation known as the infinite-frequency added mass matrix; the kernel of
focused on investigating the complex dynamics of a floating barge the convolution term hðtÞ, linked to memory effects, is the matrix
with cubic springs in surge. of retardation functions (impulse response functions); K is the
exc
This paper develops a simple wave-induced impact model for hydrostatic stiffness matrix; and f ðtÞ are the wave excitation
float-over operations by considering only the heaving motion of forces and moments.
the barge-deck system, and incorporating additional piecewise Ogilvie (1964) considered Eq. (1) in the frequency domain by
linear terms in the Cummins equation. The simplified single using the Fourier transform and derived the following frequency
degree of freedom (SDOF) wave-induced impact model is similar domain model:
to a bilinear impact oscillator that has been extensively studied   exc
x^ ðjωÞ ω2 ½M þ AðωÞ þ jωBðωÞ þ K ¼ f^ ðjωÞ ð2Þ
since Akashi (1958) analyzed the motion of an electrical bell.
exc
Studies of both hard and soft impacts can be found in the ^ ωÞ and f^ ðjωÞ are Fourier transforms of x(t) and f ðtÞ.
where xðj
exc

literature. Thompson and Ghaffari (1982) investigated the chaotic The hydrodynamic coefficients (added mass A(ω) and radiation
exc
behaviour and sub-harmonic motions of a hard impact oscillator. damping B(ω)) and wave excitation forces f^ ðjωÞ can be readily
Shaw and Holmes (1983) investigated the dynamics of a periodi- obtained from 3d hydrodynamic codes such as the program
cally forced piecewise linear oscillator by both analytical solutions DIFFRACT (Eatock Taylor and Chau, 1992; Zang et al., 2005;
and digital simulations, and provided the fundamentals of a hard Walker et al., 2008; Sun et al., 2012) used in this paper. The
impact oscillator. Lee (2005) investigated the dynamics of a hard relationship between the parameters of Eq. (1) and those of Eq. (2)
impact oscillator based on the Runge–Kutta integration algorithm. were given by Ogilvie (1964)
Andreaus et al. (2010) simulated the dynamics of a bilinear soft Z
1 1
impact oscillator based on the generalized Jacobian Matrix AðωÞ ¼ Að1Þ  hðτÞ sin ðωτÞdτ; ð3aÞ
ω 0
method. In the present paper, a combination of hydrodynamic
Z 1
analysis of the wave-induced dynamics, and dynamic analysis of
an impact oscillator, is applied to investigate the impact behaviour BðωÞ ¼ hðτÞ cos ðωτÞdτ: ð3bÞ
0
of the barge-deck system under wave excitation. Both a one-body
By taking the Inverse Fourier Transform of the above equations,
system (barge and deck moving as one body) and a two-body
the impulse response function can be formulated as follows:
system (barge and deck moving separately) are considered. The Z
numerical scheme described by Lee (2005) and Virgin and Bishop 2 1
hðtÞ ¼  ω½AðωÞ  Að1Þ sin ðωtÞdω; ð4aÞ
(1988) is applied here in deriving the time domain results. π 0
Analysis tools such as Poincaré maps, impact maps, bifurcation or
diagrams and phase trajectories are used to identify the motion Z 1
2
characteristics. hðtÞ ¼ BðωÞ cos ðωtÞdω: ð4bÞ
This paper mainly has two objectives. The first is to demon- π 0

strate the methodology of combining hydrodynamic analysis and Eq. (4b) is the preferred way to calculate hðtÞ since it converges
dynamic analysis to investigate the wave-induced impacts arising faster than Eq. (4a). This then provides the way to evaluate the
in a float-over installation. The second objective is to investigate time domain model based on frequency domain results. In
how the impact behaviour changes as the wave excitations change, practice, due to the limitations of computation time and panel
and perform a parametric study to analyze the sensitivity to sizes, diffraction codes can only produce accurate results up to a
certain control parameters of the overall system. The paper is certain frequency, say s. In order to obtain accurate results at high
arranged as follows. Section 2 briefly describes the theoretical frequencies, however, the size of panels needs to be very fine,
background of the wave-induced impact model based on the resulting in a very large number of computations (Pérez and
Cummins equation. Analysis tools for identifying nonlinear Fossen, 2008a). Therefore, the calculation of the impulse response
dynamics are briefly reviewed in Section 3. Section 4 gives functions hðtÞ is generally undertaken in the following way
validations of the established time domain model. The hydrody- Z Z
2 s 2 1
namic analysis of the float-over system is given in Section 5. hðtÞ ¼ BðωÞ cos ðωtÞdω þ Ba ðωÞ cos ðωtÞdω ð5Þ
π 0 π s
Results and discussions of the wave-induced LMU impacts are
illustrated in Section 6, for both one-body and two-body systems, where, Ba(ω) represents an asymptotic approximation of B(ω) at
and finally some concluding remarks are given in Section 7. high frequencies, which can be determined through polynomial
fitting (Greenhow, 1986). The errors associated with this fit have
been investigated by Chen et al. (2012), and it was shown how an
2. Theoretical background of wave-induced impact oscillator appropriate value of the cut-off frequency s can be determined.

2.1. Parametric time domain model based on the Cummins equation 2.1.2. Parametric model of the convolutions by using a state-space
model (SSM)
2.1.1. Cummins equation and Ogilvie relations The convolution term in the Cummins equation describes a
The forces and responses of floating structures are usually causal linear time-invariant system (Yu and Falnes, 1995). This
obtained by means of a linear analysis in the frequency domain. model is, however, cumbersome for numerical simulation and not
Cummins (1962) derived a linear time domain equation to well suited for the design and analysis of motion control systems
describe the wave excited dynamics of a floating marine structure, (Kristiansen et al., 2005). In addition, it may be very time-
which is now known as the Cummins equation. For the case of consuming to directly evaluate the convolutions depending on
42 M. Chen et al. / Ocean Engineering 76 (2014) 40–51

the simulation time length, time step and degrees of freedom of


the considered problems (Holappa and Falzarano, 1999;
Kashiwagi, 2004). The convolutions can however be approximated
by a state-space model, since they are equivalent in describing a
linear system. The state-space model is convenient for the analysis
of multiple-input and multiple-output (MIMO) systems and well
suited for the design and analysis of motion control tools. In
general, the state-space model has the following form
z_ ðtÞ ¼ Ac zðtÞ þ Bc uðtÞ
yðtÞ ¼ C c zðtÞ þDc uðtÞ ð6Þ
where z(t) is the state vector summarizing all the past information
of the system; and Ac, Bc, Cc and Dc are continuous-time state-
space matrices. The response yðtÞ of the system to any excitation
uðtÞ can be determined by the excitation and the initial value of
the state vector. Fig. 1. Configuration of a typical float-over installation.
The unknown parametric matrices of the SSM can be deter-
mined via system identification based on the frequency domain
data. In this paper, the so-called realization theory method is
adopted to determine the unknown matrices of the SSM, as
proposed by Kristansen and Egeland (2003). The identification
problem is firstly solved in discrete time by applying Kung's (1978)
algorithm based on a single value decomposition (SVD)-factoriza-
tion of the Hankel matrix of the samples of hðtÞ at discrete times.
The matrices of the discrete-time state-space model Ad–Dd are
next derived. The parametric matrices of the continuous-time
state-space model can then be obtained from the conversion of
the discrete-time state-space model. The MATLAB function imp2ss
Fig. 2. Piecewise liner impact oscillator in the unstressed configuration (The spring
in the Robust Control Toolbox implements Kung's algorithm, and is
with rigidity k1 and damping coefficient b1 is always active and the spring with
used in our approach. The state-space model obtained from imp2ss rigidity k2 and damping coefficient b2 is active only if the displacement x is greater
is of high order, and therefore needs model order reduction than the gap δ).
afterwards. The MATLAB function balmr from the Robust Control
Toolbox implements the truncated balanced reduction method,
from which the new lower order state-space matrices Ac–Dc can be hydrodynamic interactions, to focus first on behavior of a single
obtained. The detailed theoretical background of the realization degree of freedom system. Two cases are considered here: a one-
theory method can be found in Kristansen and Egeland (2003) and body system (considering barge and deck as one body) and a two-
Kristiansen et al. (2005). Other identification methods can also be body system (barge and deck moving separately).
found in the literature such as least-square fitting of hðtÞ as
proposed by Yu and Falnes (1995), and regression in the frequency
domain as proposed by Jefferys et al. (1984) and Damaren (2000).
The performances of these different identification techniques have
2.2.1. One-body system
been extensively studied by Taghipour et al. (2008) and Pérez and
For a one-body system, the barge and deck are assumed to
Fossen (2008b), in terms of ease of implementation and quality of
move as one body. This is true during transportation and the
the approximated model. By replacing the convolution term with a
duration before mating. During this time, the wave-induced
state-space model, the resulting model is a constant parametric
impact between the mating cones and the receptors is similar to
model that can be efficiently solved.
a soft impact oscillator that has been extensively studied—see, for
example, Ma et al. (2006) and Andreaus et al. (2010). Fig. 2
2.2. A simplified wave-induced impact model (LMU impacts) illustrates a typical impacting bilinear oscillator, which has inter-
mittent contacts with the spring with rigidity k2 and damping
Fig. 1 illustrates the general configuration for installing a large coefficient b2, under sinusoidal excitation. The governing equation
deck onto a pre-installed jacket by the float-over method. It can be
of the model shown in Fig. 2 can be expressed as follows:
observed that float-over installation involves multi-body dynamics
with nonlinear constraints. The barge and the deck are generally mx€ ðtÞ þbðxÞx_ ðtÞ þ kðxÞ ¼ F sin ðωtÞ ð7Þ
modeled as two bodies interconnected by DSUs. In the load
transfer stage, interactions between the deck and the jacket may where x is the displacement, from the unstressed configuration of
occur due to the wave-induced LMU impacts. In addition, the the spring with rigidity k1 and damping coefficient b1, of the mass
barge may interact with the jacket by impacting with the fenders m at time t; the dot denotes the derivative with respect to time; F
attached on the jacket legs. Due to these nonlinearities, the float- and ω are the external force amplitude and frequency, respec-
over system may exhibit complex dynamics under the environ- tively; k(x) and b(x) represent the total stiffness and damping
mental loads. In this paper, a simple wave-induced impact model coefficients in this model at any displacement. These latter
is implemented by only considering the heaving motion of the parameters are piecewise functions of x depending on the impact
float-over system. This simplification omits certain characteristics condition:
of wave-induced impacts arising in float-over operations, not least
the possibility of horizontal impacts. It is however useful, as a (
b1 ; xoδ
preliminary step in investigating nonlinear features such as sub- bðxÞ ¼ ð8Þ
b1 þ b2 ; xZδ
harmonic motions and bifurcations in systems with complex
M. Chen et al. / Ocean Engineering 76 (2014) 40–51 43

(
k1 x; xoδ 3. Analysis tools for nonlinear systems
kðxÞ ¼ ð9Þ
k1 x þ k2 ðx  δÞ; xZδ
In order to investigate the characteristics of the wave-induced
where δ denotes the position of the impact obstacle (see Fig. 2). impact model, analysis tools such as Poincaré maps, bifurcation
Unlike the conventional soft impact oscillator shown in Fig. 2, diagrams, impacts map and phase trajectories are applied here. For
the damping term for the wave-induced impact oscillator is not a the impact model represented by Eq. (10), one has a three dimensional
constant but involves the convolution term that is replaced by a phase space ðx; x_ ; tÞ. Here, we define the phase ϕ as ϕ ¼ modðt; TÞ
state-space model. The dynamics of the barge-deck system when where T is the excitation period. It is then convenient to use the three
not impacting (x3 4 g, where g is the initial gap between the dimensional phase space ðx; x_ ; ϕÞ where 0r ϕ rT. There are different
mating cones and the receptor as shown in Fig. 1) are governed by ways of presenting the solutions. The phase trajectory plots the
the Cummins equation given in Eq. (1), reduced to the terms displacement x versus the velocity x_ for all the time considered. It is
governing uncoupled heave motions x3. The LMUs are modeled as a closed trajectory for periodic motion. Two different mapping
springs with overall vertical stiffness K LMU and damping BLMU . techniques for the three dimensional space can be found in the
During impacts (x3 r  g), the dynamics of the barge-deck system literature: the Poincaré map and the impact map.
are governed by the following equation: A useful method to identify the periodicity of the motions from
Z t direct numerical simulations is the Poincaré mapping (Pippard,
½M 33 þ A33 ð1Þx€ 3 ðtÞ þ h33 ðt  τÞx_ 3 ðτÞdτ þ K 33 x3 ðtÞ 1985). A Poincaré map can be obtained by studying the values of
0 ðx; x_ Þ at regular intervals of the forcing period (T ¼2π/ω). It projects
¼f3
exc
sin ðωt þ θÞ þ F m ð10Þ a two dimensional space ðx; x_ Þ onto a plane at a particular phase
exc
ϕ ¼ ψ , where ψ is a constant within [0, T]. The Poincaré map is
where, M33 is the sum of barge and deck weights; f 3 is the force actually a set of points determined by the displacement xϕ ¼ ψ and
amplitude and θ is the phase of the incident waves; and Fm velocity x_ ϕ ¼ ψ . It contains one point for periodic motion (period-1
represents the mating forces exerted by the impacts between the motion) and m points for sub-harmonic motion of order m
mating cones and the receptors, which has the following form: (period-m motion). For example, a steady-state sub-harmonic
F m ¼  K LMU ðx3 þ gÞ  BLMU x_ 3 ð11Þ motion of order m is defined by xðtÞ ¼ xðt þ mTÞ and
x_ ðtÞ ¼ x_ ðt þ mTÞ at ϕ ¼ ψ . In the case of chaotic motion, it contains
BLMU is taken to be zero for all the simulations below (This will a complex fractal structure.
highlight the dynamic effects and experimental data from LMUs Similar to a Poincaré map, the impact map also represents a
would be required for a detailed design study). two dimensional phase space, but it is projected onto a particular
section determined by the impact instant and impact velocity, to
enable the identification of the impact behavior (Lee, 2005). For a
2.2.2. Two-body system hard impact problem, the impact map contains n points where n is
In a more complete model of a float-over operation, the deck the number of impacts in one motion period. For the problem
and barge may be allowed to move separately. The connections considered in this paper, which is a soft impact problem, there are
between deck and barge are the DSUs, which are modeled as 2n points in the impact map where n denotes the number of
springs with overall vertical stiffness KDSU and damping coefficient impacts during each motion period, since it plots both the
BDSU. It is assumed that DSUs provide compressions only and no penetrating velocity at which the mating cone starts penetrating
separation between the barge and the deck occurs. The impact with the receptor and the separating velocity at which the mating
condition is then governed by the motion of the deck. Subscript D cone starts separating with the receptor. The point set in the
denotes the deck and subscript 3 is retained for the heave of the impact map is determined by the impact phase ϕ ¼ modðt I ;TÞ and
barge. The dynamics of the deck without impact (xD 4 g) can the impact velocity at xðt I Þ ¼ δ, where t I is the impact instant.
then be described by the following equation: As with the Poincaré map, the impact map shows a complex
M D x€ D þ BDSU ðx_ D  x_ 3 Þ þ K DSU ðxD  x3 Þ ¼ 0 ð12Þ fractal structure for the case of chaotic motions. Plotting the
Poincaré map and the impact map together, the motions can be
During impacts (xD r g), the governing equation of the deck described as period-m and impact-n motions.
has the modified form Bifurcation theory has been widely applied to investigate the
M D x€ D þ BDSU ðx_ D  3 Þ þ K DSU ðxD  x3 Þ ¼  K LMU ðxD þ gÞ  BLMU x_ D ð13Þ stability of system behaviour, using point sets in a Poincaré map or
an impact map, as one control parameter is changed. More detailed
In the analysis, an artificial damping ratio μ is used to derive information about these analysis tools for nonlinear systems can be
the damping coefficient of the DSUs found in Thompson and Stewart (2002). In the following results, two
sffiffiffiffiffiffiffiffiffiffiffi types of bifurcation diagram were considered: plots of impact velo-
K DSU cities in the impact maps versus the changing control parameters
BDSU ¼ 2μM D ð14Þ
MD (wave frequency or wave height); and plots of displacements in
Poincaré maps versus the changing control parameters (wave fre-
The governing equation for the barge has the following form:
quency or wave height).
Z t
½M þ A33 ð1Þx€ 3 þ h33 ðt  τÞx_ 3 ðτÞdτ þ K 33 x3 ¼ f 3 sin ðωt þ θÞ þF D
exc

0
ð15Þ
4. Validation with published results
with
As a preliminary to using the time domain model to analyze the
F D ¼  K DSU ðx3  xD Þ  BDSU ðx_ 3  x_ D Þ ð16Þ
impact problem, two validation cases are presented here. The first is
where M is the barge weight only and FD represents the DSU to validate the accuracy of the time domain model in analyzing the
supporting forces. By combining Eqs. (12) to (16), the motions of dynamics of floating structures. The second case is to validate the
the barge and deck can be solved by using the Rung–Kutta accuracy of the numerical algorithm used in this paper in the analysis
integration algorithm. of a dynamic impact oscillator.
44 M. Chen et al. / Ocean Engineering 76 (2014) 40–51

4.1. Heaving motion of a floating hemisphere

Damaren (2000) established a time domain model based on a


rational approximation of the radiation impedance and diffraction
force. He replaced the time-consuming convolution term by a
transfer function and analyzed the heaving motions of a floating
hemisphere due to initial disturbances. Here, the same hemisphere
model is analyzed and comparisons are made with Damaren's
results.
The hemisphere has a radius of 2 m and floats in infinite water
depth. Frequency domain analysis of this model was conducted by
DIFFRACT from 0.02 rad/s to 5.02 rad/s, with a step size of 0.1 rad/s.
The mesh size was chosen appropriately to produce accurate results
in the frequency range considered here. The calculated hydrody-
namic coefficients are compared with Damaren's results and good
agreement is achieved as shown in Fig. 3. As discussed in Section 2, Fig. 5. Identification results for h33 (t).
the time domain model described by the Cummins equation can be
built based on the frequency domain results, with the impulse Fig. 6 gives the comparisons between the present results and those of
response function hðtÞ being obtained via Eq. (5). Fig. 4 shows the Damaren, which are seen to agree quite well.
values of damping coefficients at very high frequency based on the
asymptotic trend given by Greenhow (1986), with the DIFFRACT 4.2. Simulation of a bilinear soft impact oscillator
results obtained up to a cut-off frequency s ¼5.02 rad/s. Identifica-
tion results for h33(t) are shown in Fig. 5. In the figure, the calculated Andreaus et al. (2010) analyzed the motions of the impacting
h33(t) correspond to the results obtained via Eq. (5) and the results by bilinear oscillator as shown in Fig. 2, based on a generalized
the state-space model are based on Eq. (6) by assuming u(t)¼ δ(t) Jacobian Matrix method. The parameters for the model are specified
with zero initial conditions. It can be observed that a state-space in the caption of Fig. 7, which shows the results. Periodic motion,
model of order 5 can produce accurate results compared to the sub-harmonic motion and chaotic motion were all identified. In the
calculated h33(t). The transient responses of the hemisphere due to present paper, the same dynamic impact model is analyzed by
initial disturbances are analyzed by the resulting time domain model. using the Runge–Kutta integration algorithm to solve Eqs. (7) to (9).
An iterative scheme is used to identify the impact instants at x¼ δ.
For the analysis here, a time step of 0.01 s is used, which shows
good convergence. Comparisons of impact maps and phase trajec-
tories for periodic and sub-harmonic motions are shown in Fig. 7.
It can be seen that the present results agree very well with the results
digitised from the published figures. For the case of chaotic motion
(the 3rd row in the figure), only the present results are shown here,
which also appear to agree quite well with the published results.
The validation cases have shown that the present approach is able
to produce accurate results for both wave-induced dynamics and
impact oscillator problems. By combining these two together, one
can obtain a model for the wave-induced LMU impacts during a
float-over operation. The parametric time domain model established
here can be extended to model wave-induced impacts, as shown in
Eqs. (10) to (16). This is discussed in the following sections.

5. Preliminaries to the wave-induced impact model

Fig. 3. Plot of hydrodynamic coefficients in heave. Float-over operations generally can be divided into different
stages, such as zero clearance, half engagement and 0% load
transfer. Numerical modeling of float-over operations has com-
monly been conducted for the various stages (Chu et al., 1998; Tahar
et al., 2006). In order to maximize the impact effects, the initial gap
between the mating cone and the receptor is fixed as zero, which
corresponds to the 0% load transfer stage. In this paper, a float-over
configuration loosely based on the McDermott Platong Gas II Project
is considered (but as mentioned above, at this stage we are
simplifying the problem by omitting consideration of all motions
except heave). The carrying barge is similar to McDermott's Inter-
mac 650, with overall length of 198 m, a width of 51.2 m at the stern
and 42.06 m at the bow, and a draft of 7.2 m. Frequency domain
analysis of this barge was undertaken by DIFFRACT from
ω ¼0.02 rad/s to 2.47 rad/s with a step of 0.05 rad/s, for the case
of head seas. The mesh size was chosen appropriately to produce
accurate results in the frequency range considered here. The
Fig. 4. Plot of extrapolated data for B33. calculated wave excitation force and response amplitude operator
M. Chen et al. / Ocean Engineering 76 (2014) 40–51 45

Fig. 6. Comparison of transient responses due to initial displacement and velocity.

Fig. 7. Impact map (1st column) and phase portrait (2nd column) (k1 ¼1, k2 ¼103, b1 ¼ 6.2  10  4, b2 ¼ 6.2, F¼1.4881, and δ ¼ m¼1). Three different cases are plotted for
periodic motion (1st row, ω¼ 3.750), sub-harmonic motion (2nd row, ω¼ 2.211) and chaotic motion (3rd row, ω¼ 2.176).

(RAO) for heave are shown in Fig. 8 over a range of representative which demonstrates the accuracy of the established time domain
sea-state periods. As described in the previous section, a parametric hydrodynamic model.
time domain model can be established based on the frequency
domain results. The accuracy of the resulting time domain model
can be validated by comparing with frequency domain results. Fig. 9 6. Modeling of Leg Mating Unit (LMU) impacts
shows the comparisons between time domain results and fre-
quency domain results for regular wave excitation with wave height When wave-induced impacts occur, a number of factors are
h¼1 m, assuming that the barge and deck move as one body. Good critical to identifying the motion characteristics of the float-over
agreement is observed after the initial transient has died down, system, such as wave period, wave height, gap between the mating
46 M. Chen et al. / Ocean Engineering 76 (2014) 40–51

Fig. 8. Plot of wave excitation forces in heave at wave height h ¼1 m and heave RAO for the head seas.

Fig. 9. Comparison of wave-induced dynamics of the FOD system for two wave periods.

Table 1 identified: period-1 and impact-1 motion, period-1 and impact-2


Specifications of the wave-induced impact model.
motion, and period-1 and impact-3 motion. Each is considered in
Items Value detail in Fig. 11, which plots results for three wave periods. In the
impact maps (second row), tI is the time at which impact occurs.
MB (kg) 4.59E þ07 It is seen that the system responses tend to be more nonlinear as
A33(1) (kg) 1.95E þ07 period increases. For a fixed wave height, force amplitude and
MD (tonne) 2.02E þ07
K33 (N/m) 9.37E þ07
motion amplitude increase as period is increased as shown in
KDSU (N/m) 3.97Eþ09 Fig. 8. Therefore, larger penetration into the LMU, longer impact
KLMU (N/m) 6.5E þ08 duration and larger LMU mating force are observed for T¼ 22 s, as
compared with the other results in Fig. 11. In addition, no chaotic
motion is observed in any of these cases, which implies that the
considered wave-induced impact model is out of the chaotic
cone and receptor, and KLMU, and KDSU if considering barge and deck region.
move separately. In this paper, we focus on investigating how the In a similar manner, bifurcations of varying wave heights with
motion changes with varying wave excitations and the effects of the fixed wave period and varying wave periods with fixed force
component stiffnesses such as KLMU and KDSU. Due to the intermittent amplitude may be considered, using the relationships presented in
impacts, the dynamic system is nonlinear. Table 1 summarizes the Fig. 8. In this way, only one control parameter is changed when
parameters of the single degree of freedom wave-induced model, plotting the bifurcation diagrams.
based on those for a realistic float-over project. Both one-body and A common practice is to limit the allowable sea state for a float-
two-body systems are analyzed here. over operation to 1 m wave height. The change in behaviour over a
range of wave heights (1–3 m) at T ¼22 s is illustrated in Fig. 12.
6.1. Analysis of the one-body system Period-1 and impact-3 motions are observed for all the wave
heights considered here. In addition, the increase of impact
Based on the assumption of a rigid connection between the velocity as wave height increases is clear from this figure. Fig. 13
barge and the deck, these units move as one body. As for a shows the bifurcation diagrams for varying wave periods with the
conventional dynamic impact oscillator, bifurcation techniques fixed force amplitude that corresponds to the wave force at h¼1.5
are here applied to investigate how the motion changes as one and T ¼22 s. This choice of parameters highlights some interesting
control parameter, such as wave frequency or wave height, is behaviour, though it may be somewhat unrealistic over part of the
increased. Fig. 10 plots the bifurcation diagrams of the impact range for this configuration. It is clearly seen how the number of
velocity and the displacement at ϕ ¼ 0:75 T for varying wave impacts changes as wave period is increased. At T ¼12.9 s, the
periods in the range of 4–26 s with a step of 1 s, for a fixed motion bifurcates from period-1 and impact-1 to period-1 and
wave height of 1.5 m. Through the plot of displacement versus impact-2 motions. The motion then becomes period-1 and impact-3
wave periods, it is found that the motions are all periodic for the motions after T¼ 19 s and becomes period-1 and impact-4 motion
periods considered here. In the bifurcation diagram of the impact after T¼ 24 s. As T increases further, it can also be observed that the
velocity in Fig. 10, as period is increased three types of motions are motion change back to impact-3 or impact-2 motion. Based on all the
M. Chen et al. / Ocean Engineering 76 (2014) 40–51 47

Fig. 10. Bifurcation diagrams of displacement and impact velocity at ϕ ¼ 0:75 T, h ¼1.5 m.

Fig. 11. Period-1 and impact-1 motion (1st column, T ¼ 5 s), period-1 and impact-2 motion (2nd column, T ¼13 s) and period-1 and impact-3 motion (3rd column, T ¼ 22 s):
1st row: time history; 2nd row: impact map; 3rd row: Poincaré map; 4th row: mating forces, all for h ¼ 1.5 m.
48 M. Chen et al. / Ocean Engineering 76 (2014) 40–51

Fig. 12. Bifurcation diagram of varying wave height and fixed period T ¼22 s.

Fig. 13. Bifurcation diagrams for fixed wave force and varying periods (4–30 s).

Fig. 14. Period-1 and impact-3 motions of two-body system (h ¼ 1.5 m and T ¼22 s).
M. Chen et al. / Ocean Engineering 76 (2014) 40–51 49

bifurcation plots, it is clear that higher wave heights and longer separate from each other, which means that under these condi-
periods may introduce stronger nonlinearity. tions it is reasonable to make the assumption of a rigid connection
between the barge and the deck based on the simplified wave-
6.2. Analysis of the two-body system induced impact model. In addition, it can be seen that the DSU
supporting forces are positive, which in the sign convention used
Although the DSU stiffness shown in Table 1 is very large, it here is consistent with the assumption that the DSUs only provide
would be more realistic to model the barge-deck system as two compressions. The results in Fig. 14 are similar to those for the
bodies. In this section, the DSU is modeled as a constant spring one-body model which are given in the 3rd column of Fig. 11.
with stiffness KDSU and damping BDSU. The damping term BDSU Nevertheless, it is of interest to use the bifurcation techniques
provides the only damping effects on the deck, which is deter- used above to investigate the effects of varying BDSU, KLMU
mined here by Eq. (13) in terms of damping ratio μ. With an and KDSU.
assumption of a damping ratio of μ ¼0.05 and for h ¼1.5 m and Fig. 15 shows the comparison of three scenarios with the same
T ¼22 s, a period-1 and impact-3 motion is obtained, as shown in wave excitation (h¼1.5 m and T ¼22 s): time histories of motions,
Fig. 14. It is observed that the barge and the deck do not obviously impact maps, phase trajectories, and mating forces and DSU forces

Fig. 15. Plots of three motions for different system parameters (1st column: KDSU ¼ 3.97E þ 08 N/m, KLMU ¼ 6.5E þ 08 and μ¼ 0.05; 2nd column: KDSU ¼ 3.97E þ09 N/m,
KLMU ¼ 6.5E þ09 N/m and μ¼ 0.05; 3rd column: KDSU ¼3.97Eþ09 N/m, KLMU ¼6.5E þ 08 N/m, μ¼ 0).
50 M. Chen et al. / Ocean Engineering 76 (2014) 40–51

are all plotted. The first column shows the results with KDSU of (e.g. pitch and roll) will be considered in future work, and the
3.97E þ08 N/m, which is 1/10 of the value shown in Table 1. implications of horizontal impacts.
Significant separation between the barge and the deck is observed
for this much softer DSU. The motion becomes a period-1 and
impact-2 motion, with longer penetration duration. Furthermore, Acknowledgement
the oscillations of the DSU forces become less compared to the
results shown in Fig. 14.
We acknowledge the support of Lloyd's Register Foundation
The second column shows the results with KLMU of 6.5Eþ09 N/m,
towards funding the research and development program in the Centre
which is 10 times the value shown in Table 1. For a stiffer LMU, the
for Offshore Research & Engineering in National University of Singa-
response is found to be more nonlinear and a period-2 and impact-8
pore (NUS). The authors also wish to thank McDermott International
motion is generated. Based on the plots of mating forces and DSU
for providing suggestions for validation. Special thanks are given to
forces, it can be observed that each impact leads to high-frequency
Mr. Trevor Mills for his constant support and valuable suggestions.
oscillations in the DSU forces. The third column shows the results
without any DSU damping (i.e. no damping acts on the deck). It is
found that the deck motion tends to be more oscillatory, as shown in References
the impact map and phase trajectory, and the DSU forces also show
high-frequency oscillations even at the instants when the mating Akashi, H., 1958. The motion of an electric bell. Am. Math. Mon. 65 (4), 255–259.
cones do not impact with the LMUs. Due to the absence of damping, Andreaus, U., Placidi, L., Rega, G., 2010. Numerical simulation of the soft contact
it is difficult to obtain accurate numerical results for a nonlinear dynamics of an impacting bilinear oscillator. Commun. Nonlinear Sci. Numer.
Simul. 15 (9), 2603–2616.
system. It would appear that an appropriate DSU damping should be Chen, M.S., Choo, Y.S., Eatock Taylor, R. (2012). Dynamic simulation of marine
included in the time domain model to avoid these high-frequency structures based on a state-space model. In: Marine Operations Specialty
oscillations in the simulations. Symposium, Singapore.
Chu, N., Cochrane, M., Mobbs, K., Mitchell, D. (1998). Results comparison of
computer simulation, model test and offshore installation for Wandoo inte-
grated deck Float Over installation. In: Offshore Technology Conference,
7. Concluding remarks Houston, USA.
Cummins, W., 1962. The impulse response function and ship motions. Schiffstechnik
9 (1661), 101–109.
This paper briefly introduces the methodology of analyzing wave- Damaren, C.J., 2000. Time-domain floating body dynamics by rational approximation
induced impacts based on the Cummins equation. Before extending of the radiation impedance and diffraction mapping. Ocean Eng. 27 (6), 687.
Eatock Taylor, R., Chau, F., 1992. Wave diffraction theory—some developments in
the time domain model to analyze some impact problems associated
linear and nonlinear theory. J. Offshore Mech. Arct. Eng. 114 (3), 185–196.
with float-over operations, the methodology is validated for both Gottlieb, O., Yim, S.C.S., Hudspeth, R.T., 1992. Analysis of non-linear response of an
hydrodynamic analysis of floating structures and dynamic analysis of articulated tower. Int. J. Offshore Polar Eng. 2, 1.
an impact oscillator. Bifurcation techniques are applied to investigate Greenhow, M., 1986. High- and low-frequency asymptotic consequences of the
kramers-kronig relations. J. Eng. Math. 20 (4), 293–306.
the impact behaviour for varying wave excitations, based on a one- Holappa, K., Falzarano, J., 1999. Application of extended state space to nonlinear
body system assumption for a realistic configuration. In sea states ship rolling. Ocean Eng. 26 (3), 227–240.
appropriate to design operations, no chaotic motion is observed. As Jefferys, E.R., Broome, D.R., Patel, M.H., 1984. A transfer function method of
modeling systems with frequency-dependent coefficients (for flat-bottomed
might be expected, the results tend to be increasingly nonlinear for a cargo barge roll dynamics). J. Guid. Control Dynam. 7, 490–494.
longer wave periods and higher wave heights. For example, in the Kashiwagi, M., 2004. Transient responses of a VLFS during landing and take-off of
example modeled here, in a wave height of 1.5 m and a period an airplane. J. Mar. Sci. Technol. 9 (1), 14–23.
Kristansen, E., Egeland, O. (2003). Frequency dependent added mass in models for
around 13 s, there is a bifurcation such that there become 2 impacts controller design for wave motion ship damping. In: Proceedings of Sixth
per wave cycle, and 3 impacts per cycle above 19 s. The results for a Conference on Maneuvering and Control of Marine Craft, Girona, Spain.
two-body model of the system show that the one-body assumption Kristiansen, E., Hjulstad, A., Egeland, O., 2005. State-space representation of radiation
forces in time-domain vessel models. Ocean Eng. 32 (17–18), 2195–2216.
may be reasonable if the deck support unit (DSU) is quite stiff. It is Kung, S. (1978). A new identification and model reduction algorithm via singular
found that control parameters such as the stiffness and damping of value decompositions. In: 12th Asilomar Conference on Circuits, Systems and
the DSUs and the leg mating units (LMUs) play an important role in Computers, pp.705–714.
Lee, J.Y., 2005. Motion behavior of impact oscillator. J. Mar. Sci. Technol. 13 (2),
the final responses of the float-over system under wave excitation.
89–96.
Using the specified DSU stiffness in Table 1 and appropriate damping, Ma, Y., Agarwal, M., Banerjee, S., 2006. Border collision bifurcations in a soft impact
the one and two body models produced similar results in the wave system. Phys. Lett. A 354 (4), 281–287.
conditions considered. Reducing this stiffness by a factor of 10 Ogilvie, T.F. (1964). Recent progress toward the understanding and prediction of
ship motions. In: Fifth Symposium on Naval Hydrodynamics, Bergen, Norway.
reduced the number of impacts per cycle. The damping of the DSU Pérez, T., Fossen, T.I., 2008a. A derivation of high-frequency asymptotic values of 3d
is found to be very important in mitigating high-frequency vibra- added mass and damping based on properties of the cummins equation.
tions: it is however difficult to determine a realistic value for this Reducing of Maritime Accidents Caused by Human Factors Using Simulators
in Training Process, 5(1): 65–78.
parameter if experimental results are not available. A stiffer LMU will Pérez, T., Fossen, T.I., 2008b. Time- vs. frequency-domain identification of para-
introduce stronger nonlinearity into the system, which may be metric radiation force models for marine structures at zero speed. Model. Ident.
detrimental to the float-over installation: using the specified DSU Control 29 (1), 1–19.
Pippard, A.B., 1985. Response and Stability: an Introduction to the Physical Theory.
stiffness but increasing the LMU stiffness by a factor of 10, led in one Cambridge University Press, Cambridge, MA.
case examined to a complex period-2 and impact-8 motion. Shaw, S.W., Holmes, P., 1983. A periodically forced piecewise linear oscillator. J.
It should be recalled, however, that the numerical simulations Sound Vib. 90 (1), 129–155.
Sun, L., Eatock Taylor, R., Choo, Y.S., 2012. Multi-body dynamic analysis of float-over
here are based on a simple SDOF wave-induced impact model. installations. Ocean Eng. 51, 1–15.
While this does not model all the features of a typical float-over Taghipour, R., Pérez, T., Moan, T., 2008. Hybrid frequency-time domain models for
installation, the results nevertheless are indicative of the complex dynamic response analysis of marine structures. Ocean Eng. 35 (7), 685–705.
Tahar, A., Halkyard, J., Steen, A., Finn, L., 2006. Float over installation method-
impacting behavior which can arise, and illustrate the capability of
comprehensive comparison between numerical and model test results. J.
the methodology which couples frequency domain linear hydro- Offshore Mech. Arct. Eng. 128, 256.
dynamic analysis with a full simulation of the nonlinear dynamics Thompson, J.M.T., Bokaian, A., Ghaffari, R., 1983. Subharmonic resonances and
in the time domain. Further research is needed to extend the chaotic motions of a bilinear oscillator. IMA J. Appl. Math. 31 (3), 207–234.
Thompson, J.M.T., Bokaian, A., Ghaffari, R., 1984. Subharmonic and chaotic motions
impact model, to bring improved understanding to this problem. of compliant offshore structures and articulated mooring towers. J. Energy Res.
The effect of coupled responses in multiple degrees of freedom Technol. 106, 191.
M. Chen et al. / Ocean Engineering 76 (2014) 40–51 51

Thompson, J.M.T., Ghaffari, R., 1982. Chaos after period-doubling bifurcations in the Wu, M., Moan, T., 1996. Linear and nonlinear hydroelastic analysis of high-speed
resonance of an impact oscillator. Phys. Lett. A 91 (1), 5–8. vessels. J. Ship Res. 40 (2), 149–163.
Thompson, J.M.T., Stewart, H.B., 2002. Nonlinear Dynamics and Chaos. John Wiley & Yu, Z., Falnes, J., 1995. State-space modeling of a vertical cylinder in heave. Appl.
Sons Inc. Ocean Res. 17, 265–275.
Virgin, L.N., Bishop, S.R., 1988. Complex dynamics and chaotic responses in the time Zang, J., Taylor, P.H., Eatock Taylor, R. (2005). Non-linear interaction of directionally
domain simulations of a floating structure. Ocean Eng. 15 (1), 71–90. spread waves with FPSO. In: Proceedings of the 20th International Workshop
Walker, D.A.G., Eatock Taylor, R., Taylor, P.H., Zang, J., 2008. Wave diffraction and on Water Waves and Floating Bodies, Spitsbergen, Norway.
near-trapping by a multi-column gravity-based structure. Ocean Eng. 35 (2),
201–229.

Das könnte Ihnen auch gefallen