Vous êtes sur la page 1sur 16

ICCBT2008

Parametric Study of Environmental Load Impacts on a Jack-up


Structure
S. Saiedi, Universiti Teknologi PETRONAS, MALAYSIA
S. Aghdamy, Universiti Teknologi PETRONAS, MALAYSIA

ABSTRACT
The impacts of environmental factors on three major reactions of a jack-up offshore platform
are analysed using computer simulations. The reactions are the total base shear, mud line
overturning moment and platform maximum deflection. A simple in-house computer program
facilitating the manual calculations of the total base shear is developed to ensure a sound
application of a commercial software (SACS) employing finite elements analysis method.
Three major structural reactions are chosen to represent the responses of the jack-up. The
structure is a typical operating jack-up offshore platform in Central North Sea with a water
depth of 97 m, a maximum wave height of 31 m with a period of 18 s, a current speed of 0.67
m/s, and 1-min mean wind speed of 40 m/s at 10 m above the sea level. The loads analysed are
due to wind, current and wave. The first two environmental actions are computed by drag
force equation while the third is expressed by Morrisons equation that accounts for both drag
and inertia forces. Changes in all three reactions under a wide range of major design factors
are monitored and careful conclusions are drawn. The factors are wave height (H), wave
period (T), wind and current speeds, size (D) of the structural members (chords and braces)
and the storm surge. The dominant role of wave forces, as compared to those of wind and
current, is emphasised. It is shown that the shear force is more sensitive to H rather than to
other wave parameters or to D. However, the dominant parameter in the overturning moment
could be H or T depending on the range of the relative change of the respective factors. The
overall behaviour of the platform deflection is similar to that of the overturning moment. The
relative increase of D brings about a low relative change in the overturning moment but a high
change in the base shear. The findings provide for guidance in the conceptual design of jackup offshore platforms.
Keywords: Jack-up offshore structure, parametric study, environmental loads, SACS

*Correspondence: Associate Professor Dr. Saied Saiedi, Civil Eng. Dept., UTP, 31750 Tronoh, Perak, Malaysia.
Tel: +60 5 368 7284, Fax: +60 5 36 56716. E-mail: saiedsaiedi@petronas.com.m y
** Graduating Student (July 2008), Civil Eng. Dept., UTP, E-mail: sanam_aghdamy@yahoo.com

ICCBT 2008 - D - (43) 471-486

Parametric Study of Environmental Load Impacts on a Jack-up Structure

1.

INTRODUCTION

Great majority of the existing offshore platforms are of bottom-supported type as opposed to
floating structures. The floating structures include tension leg platforms (TLP), semisubmersibles, Spars and ship shaped vessels. Bottom-supported structures can be divided into
fixed (jacket, gravity base, jack-up) and compliant structures. While the latter are occasionally
placed in waters as deep as 850 m, the former are usually constructed in shallower waters
(<400 m). Offshore structures are constructed to withstand various types of loads. These are
permanent (dead), operational (live), environmental, construction-installation, and accidental
(ship impact, fire, blast, etc.) loads. The environmental factors are wave, wind, current,
earthquake, temperature, ice, marine growth, and seabed movement. From these, the first three
usually dominate the design.
Offshore disaster rate involving jack-ups has exceeded that of other offshore installations ([1,
2, 3, 4]). The number of failures is increasing as the demand for jack-up operation in deeper
waters and harsher environments has increased. Furthermore, in some regions and for specific
purposes, the Jack-up structures must be designed to operate at one site for long periods which
also increase the failure risks.
Nowadays, the analysis and design of offshore structures call for enormous resources at least
for three reasons: complicated geometry of offshore structures, numerous loading
combinations, and large number of structural checks (static in-place analysis, pile-soil-structure
interaction, plastic collapse analysis, etc.). Commercial computer programs such as SACS [5,
6] and ANSYS-ASAS OFFSHORE [7] have become necessary design tools for major design
firms. Some civil engineering software for the analysis and design of routine structures such as
buildings or bridges have recently moved to incorporate some elements of offshore structures.
STAAD PRO [8] is a recent example. The application of sophisticated software may be
associated with two potential setbacks. The first is that most of these are expensive and hardly
accessible to most practicing engineers, students, and academic educators. In the absence of
such tools, many who need to gain or teach the basic physical understandings of the structural
behavior of offshore structures might be systematically kept away. The second stems from the
facts that the ease of usage, black-box nature of the software, and glamorous
tabular/graphical outputs often sideline the necessity of clear understanding of physical
concepts, real-life behavior of the structure under varying conditions, and intelligent
application of the software for each situation.
The present paper aims at two objectives: to provide with a simple computational tool for
preliminary analysis of a jack-up and to give insight into the reaction of the structure to varying
environmental factors. A simple in-house computer program has been developed to obtain two
major structural reactions, shear and moment, at the base of jack-up structures under three
major environmental loads: wave, wind and current. The program is called BASFJOSEL
(Base Shear Force of a Jack-up Offshore Structure under Environmental Loads) here for ease
of reference, can be equally used for any bottom supported offshore platform. In addition to
the application of BASFJOSEL the commercial software SACS is also employed to provide for
the comparison as well as computation of the jack-up deflection. This paper is an extension of
a previous study on the sensitivity of the base shear of a monopod offshore platform to the
environmental loads [9].
472

ICCBT 2008 - D -(43) 471-486

S. Saiedi & S. Aghdamy

2.

ENVIRONMENTAL LOADS AND BASIC ASSUMPTIONS

A recent study [10] on a large database containing the overall causation data for 71,470
accidents recorded from 1991 to 2001, identified the following as the failure sources: human
(mainly operational) errors 46%, engineering (mainly construction) failures 41%, extreme
environmental (weather related) conditions 11%, and hazardous materials (fire, blast, etc) 2%.
The total number of jack-up accidents in 1979-1988 was 226 with an average overall annual
accident risk of 6% [11]. Figure 1 shows the distribution these accidents among which the
direct contribution of storm is about 7%.

30

% Accidents

25
20
15
10
5

ng
Ja
ck
i

Fa
t ig
ue

O
th
re

St
or
m

De
sig
n

io
n
lis
Co
l

ire

ow
-o
ut
/f

To
w

Bl

So
il

Figure 1. Distribution of 226 jack-up accidents in 1979-1988

BASFJOSEL requires the following structural-environmental date as input.


Type of jack-up (hull type, leg type)
Geometry of the structure : Members (length, orientation, position and diameter of all
members), topside (overall dimensions), air gap (height)
Water depth and storm surge height
Design wind (velocity at a referenced height-10 m- and choice of empirical parameters in
the formula for velocity profile
Design current velocity at sea level
Design wave height and period
The program estimates marine growth thickness according to NORSOK [12]. It also computes
empirical parameters for Morisons equation application such as drag coefficient, inertia
coefficient and shape coefficient according to API Guidelines [13] and Coastal Engineering
Manual [15]. Basic wave mechanics relations for small amplitude waves (Airy theory) are used
to obtain wavelength, maximum horizontal and vertical components of the particle velocity and
acceleration along the water depth [14, 15].
The wind velocity profile is automatically generated using power law to get the design
velocities at specified levels. Wind drag force on hull and dray parts of legs of the structure is
then computed incorporating velocity changes with depth.
ICCBT 2008 - D -(43) 471-486

473

Parametric Study of Environmental Load Impacts on a Jack-up Structure

The current velocity profile is automatically computed with power law to obtain the current
velocity at any required depth. The current drag force on the submerged part of lags is then
calculated by the usual drag force formula incorporating velocity changes with depth.
The wave forces on the structure are computed by the application of Morrisons equation to
each exposed member. For members extending from the seabed to water surface, the total
integrated formulas for forces and moments are employed. For a brace, the force for unit
length of the structure at the member centre is computed. The total brace force is then obtained
by multiplying this unit force by the member length. Knowing that the individual maximums
for inertia and drag forces do not occur simultaneously, a conservative design approach has
been adopted in which the individual maximum inertia and drag components during a wave
cycle are added to get the total maximum force and moments (see pp. 253-4 of VI-5, [15]). The
lift force on submerged structural members has insignificant effect on the total shear force and moment
at the base of the structures. Therefore, its incorporation in BASFJOSEL is not reported here.

The software allows for user-defined values or automatic calculation based on built-in
standards or guides. The equations for these three forces are as follows.
1
airVwind 2 C S A
2

(1)

1
2
= waterVcurrent C D A
2

(2)

Fwind =
Fcurrent

Fwave = Fi + FD
Fi = C M water g

(3)

D
4

HK i

(4)

1
FD = C D water gDH 2 K D
2
2d
1
K i = tanh(
)
2
L

(5)
(6)

1
KD = n
4
Cg 1
4d / L
n=
= (1 +
)
C
2
sinh[ 4d / L]

(7)
(8)

CD=1.2
CD decreases from 1.2 to 0.6-0.7
CD =0.6-0.7

Re < 105
105 < Re < 4 105
Re > 4 105

(9)

CM =2.0
CM =2.5 Re/5 105
CM=1.5

Re < 2.5105
2.5 105 < Re < 5 105
Re > 5 105

(10)

Cs= 1.5
Cs=1.5
Cs=0.5
Cs=1.0

beams
sides of buildings
cylindrical sections
total projected area of platform

474

(11)

ICCBT 2008 - D -(43) 471-486

S. Saiedi & S. Aghdamy

In Equations (1) to (11), Fwave is the total wave force consisting of drag FD and inertia Fi forces,
is mass density, g is gravitational acceleration, D is member diameter, H is wave height, T
is wave period, L is wave length, C is wave celerity, Cg is wave group velocity, n is C/Cg, KD
and Ki are drag- and inertia- related coefficients in Morrisons equation (3), respectively, CM
and CD are drag and inertia coefficients, respectively, d is water depth, Cs is wind shape factor,
A is frontal area against wind and Re is Reynoldss number as defined by the ratio of (water
velocity D)/ (kinematic viscosity of water). The total forces by wind, current and wave are
translated to the base of the structure in terms of shear force providing.
3.

THE CASE STUDY

A typical jack-up structure in Central North Sea (called MSL here for ease of reference) is
selected as the case study to monitor changes in the base shear force, overturning moment and
platform deflection with varying environmental factors. The platform is adopted from a
comprehensive investigation [16], performed by MSL Engineering Limited for the Health and
Safety Executive (HSE) of United Kingdom to determine the effects of air gap and subsequent
hull inundation when wave slams into the deck of a jack-up. The dimensions of major
components and key environmental parameters, applied omni-directionally, are contained in
Table 1. Figures 2 and 3 show the configuration of the case study structure. MSL is a threelegged jack-up structure with a triangular hull founded on them. The legs are triangular,
consisting of three tubular chords braced with a K-bracing arrangement. At the bottom of each
leg is a spudcan. The connection between the leg and hull consists of a set of pinions and rigid
horizontal guides at the bottom of the hull and at the top of the yoke frame.
Table 1. Characteristics of the case study jack-up (MSL)
MSL Model Given Dimensions
Hull
80m 72m 16m
Overall leg length
146m
Leg spacing
55m
Vertical spacing between horizontal braces
6.96m
Length of horizontal bracing
12.2m
MSL Model Assumed Dimensions
Diameter of chords
0.85m
Diameter of braces
0.6m
Leg penetration
23m
Spudcan diameter
14.3 m
Environmental Parameters of Central North Sea
Water depth including storm surge
96.6m
Maximum wave height
31.2 m
Period for the 100-year wave
17.7 sec
Associated current speed at the surface
0.67 m/sec
Maximum wind speed at 10m above SWL
40.1 m/sec
Soil type
Sand

ICCBT 2008 - D -(43) 471-486

475

Parametric Study of Environmental Load Impacts on a Jack-up Structure

D chord
(850mm)
80 m

Hull Height
(16m)

Overall
height
(146 m)

D brace
(600mm)
6.96 m

72 m

12.2 m

a. 3d View

b. Plan View

Figure 2. Case study jack-up (MSL)

Figure 3. Environmental Loads on MSL


476

ICCBT 2008 - D -(43) 471-486

S. Saiedi & S. Aghdamy

4.

RESULTS AND DISCUSSIONS

The reactions of the structure to the environmental forces were analysed by the application of
BASFJOSEL as well as SACS software. A full account of the analysis and an experimental
setup to verify the theoretical findings are given in [17]. A summary of the shear force,
overturning moment and platform deflection is contained in Table 2. Values of some primary
factors are as follows:
L = 433.2 m; C = 24.5 m/s; n = 0.67
Max. water particle velocity at SWL
= 6.25 m/s
Max. water particle acceleration at SWL= 1.97 m/s2
Ki = 0.44 ; KD = 0.17
Cs (Hull) = 1.5 ; Cs (dry part) = 0.5
Wind Frontal Area = 1603 m2
Table 2. Base shear, overturning moment and platform deflection for MSL
Base Shear from BASFJOSEL
Reaction
MN
% of Ftotal % of Ftotal
26.7
Hull: 2.289
0.3
Vertical Members: 0.03
Fwind on whole structure:
Fwind
28.4
0.5
Horizontal Members: 0.039 2.433
0.9
Diagonal Members: 0.075
Vertical Members: 0.100
1.2
Fcurrent on whole
4.0
Fcurrent
Horizontal Members: 0.049
0.6
structure: 0.339
2.2
Diagonal Members: 0.190
Vertical Members: 0.641
7.5
Fwave on whole structure:
Fwave
67.7
Horizontal Members: 2.150 5.804
25.1
35.1
Diagonal Members: 3.013
Ftotal
8.576
100
100
From SACS
Overturning
Moment
Platform Maximum
Reaction
Base Shear
(MY)
Deflection (Dx)
(FX)
% of
MN
% of Ftotal
MN.m
Mtotal
Wind
2.5
26.9
557
44
340 mm
Wave + Current
6.7
73.1
699
56
Total
9.2
100
1256
100

4.1 Breakdown of the total base shear force


Figure 4 shows the breakdown of the total base shear force with wave loads accounting for
68%, wind loads forming 28% and current loads making only 4%. With respect to major
structural components, diagonal braces, hull, horizontal braces, and vertical braces attract 38%,
27%, 26% and 9% respectively, of the total load by wave, wind and current. It should be
pointed out that although diagonal members as a whole attract more environmental loads than
the vertical members (chords), but the role of each diagonal member in taking environmental
loads is less than that of vertical ones. This has been analytically substantiated elsewhere [17].
ICCBT 2008 - D -(43) 471-486

477

Parametric Study of Environmental Load Impacts on a Jack-up Structure

Figure 4. General breakdown of the base shear

4.2 The ratio of drag force to inertia force


Figure 5 gives the breakdown of the total wave force, Fwave on all submerged members
(vertical, diagonal and horizontal braces) into the drag and inertia forces, FD and Fi,
respectively. These proportions are related to the Keulegan-Carpenter KC number and are
especially sensitive to member diameters and wave height.

Inertia force
11%

Drag force
89%

Figure 5. Breakdown of wave force as in Morisons equation for braces

As waves bring greater contribution to the total load, it is interesting to look at the wave force
in terms of the two components in Morrisons equation. The ratio of FD to Fi could be studied
in terms of variation of D and H. It should be noted that for small Keulegan-Carpenter values
(KC<5) associated with large member sizes and small mean particle displacements (i.e., small
waves), the inertia force Fi, rather than the drag force FD, is dominant. Table 3 contains a
useful guide for the validity range of the theories and domination of the drag and inertia forces.

478

ICCBT 2008 - D -(43) 471-486

S. Saiedi & S. Aghdamy

The ratio when D varies: While FD is directly proportional to D, Fi is proportional to D2.


Therefore, increase of D has a greater impact on Fi than on FD. Noting that increase of D
reduces the Keulegan-Carpenter number, this finding is also consistent with the established fact
that drag becomes more dominant when KC increases.
H
gT
The ratio when D varies: The peak particle velocity is Umax= ( /2) ( /2). Increase of the H
leads to an increase of Umax, which in turn increases KC. Thus, the ratio of drag to inertia
forces will also increase.
KC
KC>25
5<KC<25

KC<5

Table 3. Guide for wave load calculation; [13, 15]


D/L<0.2
D/L>0.2
Drag dominated
Diffraction
theory
Morison equation with CM and CD
Drag and inertia dominated range
should be used for
Morison equation applicable, but CM and CD values
computing
wave
show large scatter.
forces.
Inertia dominated range
Morison equation or Diffraction theory is used
Effect of drag is negligible

4.3 Sensitivity of structural reactions


In the preliminary design of offshore platforms, it frequently becomes necessary to know
reactions of the structure to a range of values for each of the parameters involved. An insight
into the sensitivity of the design to these factors, whether environmental or structural, provides
for quick decisions in the conceptual design stage. Several runs of BASFJOSEL and SACS
software were performed to investigate the changes in the reactions (base shear, overturning
moment and platform deflection) with the following parameters: wave height, member
diameters, wind speed, current speed, wave period, water depth.
4.3.1 Base Shear and Moment: The results are depicted in Figures 6, 7, 8 and 9 in nondimensional form. While the origin of the axes denotes the existing situation (i.e., no change),
any relative change in the factors is associated with a relative change in the total base shear
force or overturning moment. In both non-dimensional axes, + means increase and
means decrease. As examples of the application of these curves, the following are noted.
o 50% increase in H (i.e., a wave height 1.5 times that of the original height) will increase the
total shear force by 85% (i.e., a force 1.85 times that of the original force) from BASFJOSEL
(see Fig. 6).
o 50% decrease in T (i.e., a wave period half that of the original period) will reduce the total
shear force by only 15% from SACS (see Fig. 7).
o 25% increase in D (i.e., member sizes 1.25 times those of the original diameters) will
increase the total overturning moment by only 12% from SACS (see Fig. 9).
As for the base shear and moment, the following general observations can be made:
(a) For most real-life jack-ups, the wave has the largest impact on the total stability of the
structure. However, other forces may be dominant locally like the dynamic forces of the wind
for the hull and its various components (cranes, flare booms, etc.).
ICCBT 2008 - D -(43) 471-486

479

Parametric Study of Environmental Load Impacts on a Jack-up Structure

(b) Base shear is the most sensitive to the wave height, H compared to T (and L).
(c) The impact of the wave height to the total base shear is greater than that of the member
diameters. A close look at equations (3) to (5) can physically support this observation. From
equation (4), it is seen that Fi is proportional to D2 and H. Considering the exponents, Fi is
then more sensitive to D than to H. With the same look at equation (5), it is seen that FD is
more sensitive to H than to D. The discussion in (b) reveals that for most jack-up structures,
FD is more significant than Fi. Therefore, the total base shear force is more sensitive to H (that
is the core parameter in FD) than to D (that is the core parameter in Fi).
(d) The variation of overturning moment (OTM) with the parameters does not show the same
pattern in all ranges of relative change of the parameters. When increasing the parameters, the
highest sensitivity of OTM is to H. However, in decreasing the parameters, the highest
sensitivity of OTM is to T.
4.3.2. Leg Deflection: Figure 10 shows the variation of maximum deflection of the platform
(at the top corner of the hull) with various environmental or size factors. The interpretation of
Figure 10, follows the same as performed in (d) above for OTM. As the maximum deflection
is directly proportional to the moment, this similarity of Figure 10 to Figure 9 is well justified.
Figure 11 shows the maximum deflection of a jack-up leg at various elevations, from the mud
line (with zero deflection) to the top (with 34 cm deflection). The deflection of the jack-up is
due to moments from three force sources:
(i)
(ii)
(iii)

Wind force that is fairly concentrated at the hull (top side) with the largest arm. This
force is 28% of the total force (see Table 2).
Current force that is fairly distributed with larger values near water surface and a
resultant force in the top half-water depth. This force is only 4% of the total force (see
Table 2).
Wave force that is fairly distributed on a segment of legs near water surface with
values diminishing rapidly with water depth. This force is 68% of the total force (see
Table 2).

The largest moment is produced by wave, followed by the wind and current. For the sake of a
conceptual discussion, it would be helpful if the whole jack-up structure is likened to a
cantilever beam as shown in Appendix I. The Appendix contains the deflection characteristics
of the beam (one end fixed, the other free) under various loads. Three cases (i) to (iii),
resemble beam type 1 with x=beam length L, beam type 4, and beam type 1 with L/2<x <L,
respectively. Comparing the resulting shape of the deflected leg with those expected from the
respective beams in Appendix I, strong resemblance can be noted. An exercise in application
of curve fitting to the plot in Figure 11 produced a 3-rd degree polynomial with coefficients
close to those from superposition of three loads.
4.4 Comparison between SACS and in-house software
Table 4 compares the base shear forces obtained from BASFJOSEL with those from SACS.
Despite the large approximations involved in the assumptions of both programs, results are
sufficiently close.

480

ICCBT 2008 - D -(43) 471-486

S. Saiedi & S. Aghdamy

An example of the approximations relates to SACS treatment of empirical values of drag and
inertia coefficients, CD and CM, in Morrisons equation. While experimental observations show
that these coefficients depend generally on Re (taking into account both water velocity and
member size) as shown in equations (9) and (10), SACS determines these coefficients as
functions of D only. Table 5 contains the tabular from of these functions. The resulting
deference in the magnitude of CD and CM can partly explain the difference in the predicted base
shear forces.

250

% Changes in Base Shear

200

150

100

50

0
-50

-25

25

50

75

100

-50

-100
% Changes in Parameter
Wave Height

Member Diameter

Wind Speed

Current Speed

Wave Period

Lo/L

Linear (Water Depth)

Figure 6. Sensitivity of base shear to various parameters based on in-house software output

% Changes in Base Shear

150

100

50

0
-50

-25

25

50

-50

-100

% Changes in Parameter
Wave Height

Member Diameter

Wind Speed

Wave Period

Figure 7. Sensitivity of base shear to various parameters based on SACS software output

ICCBT 2008 - D -(43) 471-486

481

Parametric Study of Environmental Load Impacts on a Jack-up Structure

150

% Change in OTM

100

50

0
-50

-40

-30

-20

-10

10

20

30

40

50

-50

-100
% Change in Parameter
Wave height

Member Diameter

Wind Speed

Wave Period

Figure 9. Sensitivity of overturning moment to various parameters based on SACS software output

150

% Changes in Deflection

100
50

0
-50

-40

-30

-20

-10

10

20

30

40

50

-50
-100
-150

% Changes in Parameter
Wave Height

Member Diameter

Wind Speed

Wave Period

Figure 10. Sensitivity of platform deflection to various parameters based on SACS software output

Table 4. Comparison of results between SACS and In-house software


% Difference
BASFJOSEL
SACS
Force
(col. 1)
(col. 2) (col. 1-col. 2)/(col. 1)
Wind (MN)
2.43
2.46
1.14
Wave + Current (MN)
6.14
6.70
9.07
Total Base Shear (MN)
8.58
9.16
6.82

482

ICCBT 2008 - D -(43) 471-486

S. Saiedi & S. Aghdamy

Table 5. Assumption by SACS for CD and CM in Morrisons Equation (from SACS output)
Member size (D, cm)
30.48
60.96
121.92
182.88
243.84
304.80

Drag Coefficient , CD(-)


Normal flow
Axial Flow
0.6100
1.3900
0.6650
1.4000
0.7200
1.4500
0.7560
1.6000
0.7810
1.6700
0.7990
1.7100

Inertia Coefficient , CM (-)


Normal flow
Axial Flow
0.6100
1.3900
0.6650
1.4000
0.7200
1.4500
0.7560
1.6000
0.7810
1.6700
0.7990
1.7100

160

140

120

Elevation (m)

100

80

60

40

Mud-line
20

0
0

10

20

30

40

Deflection (cm)

Figure 11. Deflection vs. elevation

5.

CONCLUSIONS

An in-house computer program and a commercial software (SACS) software have been applied
to a typical jack-up structure to get insight into the sensitivity of these offshore platforms to
environmental parameters and size of the structural members. A typical jack-up as designed
for the Central North Sea was taken as the case study.

ICCBT 2008 - D -(43) 471-486

483

Parametric Study of Environmental Load Impacts on a Jack-up Structure

The breakdown of the base shear force and overturning moment in terms of contributions from
three environmental loads pertaining to wave, current and wind revealed the dominant role of
the wave forces for real-life design conditions. The order of relative magnitude of these forces
was also introduced. Examination of the values and ratios of wave action to the members,
wind action to the hull and to the dry members in the air gap, and current action to various
members showed that among the wave parameters (height H, period T), H causes greater
impact on the base shear whilst the overturning moment and platform deflection show
maximum sensitivity to H only when increase of the parameters is considered. With the
decrease of the parameters, the overturning moment and platform deflection are the most
sensitive to T rather than to H, D and wind speed. It also revealed that the significance of the
wave height is even greater than the role of the member diameters in the total value of the base
shear force.
Referring to deflection of cantilever beams under various loads, a simple interpretation of the
deflection of the jack-up leg is introduced. Numerous applications of in-house computer
program have proved that in the absence of specialised expensive simulators such as SACS and
ANSYS- ASAS OFFSHORE, the program is a handy tool in preliminary checks during the
conceptual design of jack-up platform as well as for educational purposes requiring quick
calculations of the total forces.
REFERENCES
[1].
[2].
[3].
[4].
[5].
[6].
[7].
[8].
[9].
[10].
[11].
[12].
[13].
[14].

484

Boon, B., Vrouwenvelder, A., Hengst, S., Boonstra, H., and Daghigh, M. (1997). System
reliability analysis of jack-up structures: possibilities and frustrations. Proc. of 6th Int. Conf. JackUp Platform Design, Construction and Operation, City University, London
Leijten, S.F. and Efthymiou, M. (1989). A philosophy for the integrity assessment of jack-up
Sharples, B.P.M., Bennett Jr, W.T. and Trickey, J.C. (1989). Risk analysis of jack-up rigs. Proc.
of 2nd Int. Conf. on the Jack-Up Drilling Platform, City University, London, pp. 101-123.
Young, A.G., Remmes, B.D. and Meyer, B.J. (1984). Foundation performance of offshore jackup drilling rigs. J. Geotech Engng Div., ASCE, Vol. 110, No 7, pp. 841-859, Paper No. 18996.
Engineering Dynamics Incorporation (EDI), USA, Structural Analysis Computer System;
http://www.sacs-edi.com/ProductInfo.shtml
SACS Release 5 User Manual. (2005). Engineering Dynamics, Inc.
ANSYS
Incorporation,
USA,
ANSYS ASAS-OFFSHORE
version
14.04;
http://www.ansys.com/default.asp
STAAD. offshore release 2005 manual, Research Engineers International, Bentley Solution
Centre
Saiedi M.R., Saiedi S., Shafiqi N., (2007), Sensitivity of Bottom-Supported Offshore Platforms
to Environmental Forces; CUTSE07, Engineering Conference, Curtin University of Technology
Sarawak, Malaysia, 26-27 November
Baker C.C., McCafferty D.B., 2005, Accident database review of human-element concerns: what
do the results mean for classification?, ABS Technical Papers, also presented at Human Factors
in Ship Design, Safety and Operation held in London, February 23-24
Le Triant. P and Prol. C. (1993). Stability and Operation of Jack-up, Technip, Paris
Standards Norway. (2004) , N-003, NORSOK STANDARDS Actions and action effects, Draft 2
for Revision 2, February
American Petroleum Institute (API). (1993), Recommended Practice for Planning, Designing
and Construction Fixed Offshore Structures - Working Stress Design, RP 2A - WSD, Twentieth
Edition
Chakrabarti S.K. (2005), Handbook of Offshore Engineering, Vol. I, Offshore Structure Analysis,
Inc., Plainfield, Illionois, USA
ICCBT 2008 - D -(43) 471-486

S. Saiedi & S. Aghdamy


[15]. US Army Corps of Engrs. (2006), Coastal Engineering Manual, EM 1110-2-1100,
[16]. MSL Engineering Ltd. (2003), Research Repot 019, Sensitivity of Jack-up Reliability to waveIn-Deck Inundation
[17]. Aghdamy S. (2008), Impacts of Environmental Loads on a Jack-up offshore Structure, Final
Year Project (Thesis), Civil Engineering Department, Universiti Teknologi PETRONAS,
Malaysia
[18]. Pilkey W.D. (1994), Formulas for Stress, Strain, and Structural Matrices, John Wiley and Sons,
Inc., 1458pp.

ICCBT 2008 - D -(43) 471-486

485

Parametric Study of Environmental Load Impacts on a Jack-up Structure

Appendix I. Deflection of Beams (Pilkey 1994)

486

ICCBT 2008 - D -(43) 471-486

Vous aimerez peut-être aussi