You are on page 1of 16

Ocean Engineering 28 (2001) 899–914

Statics of a three component mooring line


Russell J. Smith a, Colin J. MacFarlane a, b,*

a
Marine Technology Centre, University of Strathclyde, 100 Montrose Street, Glasgow, G4 0LZ, UK
b
MOSIS Systems Limited, Tilehurst, Farm Lane, Ashstead, Surrey, KT21 ILW, UK

Abstract

Catenary equations are solved for a three component mooring made up of two lines, connec-
ted at a point buoy or sinker where water depth and fairlead tension are given. This configur-
ation is typical of deep water moorings and solutions are required for preliminary design of
mooring systems and measurement of stability in service of moored floating platforms. The
problem is transformed to a single polynomial equation of degree eight and is solved by
Laguerre’s iteration. Where there is no buoy or sinker the equation reduces to a quartic and
can be solved in closed form. Elongation of the lines is shown to be equivalent to small
uncertainties in the weights per unit length. The method is computationally efficient.  2001
Elsevier Science Ltd. All rights reserved.

Keywords: Catenary; Mooring line; Mooring; Statics

1. Introduction

Static analysis of grounded catenaries forms part of the preliminary design of


mooring systems for floating platforms. Placement of several mooring lines around
the platform provides the principal resistance to displacements in the horizontal
plane, induced by environmental loading (Faltinsen, 1990). Displacements in the
vertical plane also induce tension in the mooring lines but the resistance to these
motions is principally provided by buoyancy of the platform. An initial analysis is
likely to consider weight from a single line and end-point tensions as the only sig-
nificant forces (Koburt and Herbert, 1970; Rothwell, 1979; Jain, 1980). The analysis
is refined with the inclusion of multiple components, as shown for example in Fig.

* Corresponding author. Tel: +44(0)1415483304.


E-mail address: c.j.macfarlane.strath.ac.uk (C.J. MacFarlane).

0029-8018/01/$ - see front matter  2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 2 9 - 8 0 1 8 ( 0 0 ) 0 0 0 5 8 - 5
900 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914

1 (Niedzwecki and Casarella, 1976; Oppenheim and Wilson, 1982; Orgill et al.,
1985) elongation of the line (Oppenheim and Wilson, 1982; Ogawa, 1984) distributed
drag from currents (Skop, 1988) quasi-static interaction with the platform (Brown,
1997) and ultimately a complete dynamic simulation (Mavrakos, 1996; Sun and
Leonard, 1998). The methods of lumped masses (Ansari, 1980; Anasri and Khan,
1986) or finite elements (Papoulias and Bernitsas, 1988) are often preferred for com-
plex arrangements of components and loads.
The motivation for this work lies with the measurement of stability in service
(MOSIS) of moored platforms. The MOSIS system provides a statistical statement
of the lightweight and associated centre of gravity of a platform, based on a sequence
of ballast shifts and filtering of rotational motions (Bradley and MacFarlane 1986a,b,
1995; MacFarlane, 1996). It is accepted by marine authorities as equivalent to regular
inclining and is fitted to many drilling and production semi-submersibles world-wide.
Comparison with regulatory KG requires that the measured rotational stiffness is
adjusted to account for the contribution to the displacement and righting moments
of the platform of tensions in the mooring lines. The problem addressed here is to
determine the inclination of the mooring on the seaward side of the lower fairlead,
given the tension recorded at the platform, water depth and immersed weights of a
three component line as shown in Fig. 1 — a typical deep water mooring made up
of two lines joined at a submerged buoy or sinker. The effects of current flow and
time dependent motions are not considered being averaged and balanced over time.
Elongation of the line is included by way of an uncertainty in the weight per unit
length.
It is shown here that the problem transforms to a single polynomial equation of
degree eight and this is solved by Laguerre’s iteration. The absence of a buoy and
sinker reduces the degree of the equation to four and this can be solved in closed
form. The problem can also be cast as coupled polynomial equations and solved by
global homotopy methods (Watson et al., 1987). This may prove in our ongoing
study to be better suited for more general arrangements of components, loads and
boundaries.

Fig. 1. A three component mooring line arrangement.


R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914 901

2. Catenary equations

The completely submerged mooring line configuration under consideration is


shown in Fig. 1. The lower line is a grounded catenary that connects to an ancillary
cable and anchor providing a horizontal restraint force. The upper line connects to
a floating platform through a fairlead and winch. The two lines are connected at a
point buoy or sinker. Loads on the mooring are the immersed self weights of the
components and end-point tensions. The lines are considered inextensible in the
present description and the effects of elongation are considered below.
Static equilibrium between the touchdown point at the origin of the axes x and y,
and distance s along the mooring is given by the summation of forces


s

⫺T0t̂0⫹ fdt⫹T(s)t̂(s)⫽0, (1)


0

where vectors are denoted by boldface, t̂ is the unit tangent in the direction of increas-
ing s, f is the external load distribution and T is the tension. The load f is given by
f⫽⫺f ĵ, f⫽w1⫹W1d(s⫺s1)⫹(w2⫺w1)H(s⫺s1), (2)
where w1 and w2 are uniform immersed line weights per unit length, W1 is a concen-
trated force at the connection point 1, and d and H are the Dirac impulse and Heavi-
side step functions. Resolving Eqs. (1) and (2) along the directions of the axes x
and y, gives
T(s) cos q(s)⫽T0⬅constant, T(s) sin q(s)⫽F(s) (3a-b)
where

冕 再
s
w1s, s⬍s1,
F(s)⫽ fdt⫽ (4)
w1s1+W1+w2(s−s1), s⬎s1
0

The horizontal component of the tension is constant along the mooring and rep-
resents the restraint force of an anchor positioned at s⬍0.
Manipulation of Eqs. (3) and (4) leads to the classical catenary equations (Symon,
1971; Meriam, 1975):
dy/dx⫽ sinh(wix/T0⫹ai), y⫽T0/wi cosh(wix/T0⫹ai)⫹bi, (5)
s⫽T0/wi sinh(wix/T0⫹ai)⫹gi, T⫽wi(y⫺bi), (6a,b)
T/T0⫽ cosh(wix/T0⫹ai), 冑T /T −1⫽ sinh(W x/T ⫹a ),
2 2
0 i 0 i (7)

where the subscript i=1,2 for mooring lines on the intervals 0=s0ⱕs⬍s1, s1ⱕsⱕs2
and ai, bi, and gi are integration constants to be determined from boundary conditions.
Eq. (6b) is obtained by a translation of co-ordinates from an origin located where
902 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914

dy/dx=0. The tension T0 is the main unknown and is sometimes referred to by the
ratio T0/wi called the catenary parameter (Avallone and Baumeister, 1986). Other
unknown parameters are s1, x1 and x2. The prescribed parameters for the problem
addressed here are H, L2 and T2.

3. Boundary conditions

3.1. Point of seabed contact

The boundary conditions at the point of seabed contact are


s⫽0, x⫽0, y⫽0, dy/dx⫽0, (8)
and this gives for the integration constants
a1⫽0, b1⫽⫺T0/w1, g1⫽0. (9)

3.2. Connection point

Immediately to the left of the connection point at s=s1 :


y−1 ⫽T0/w1(cosh(w1x1/T0)⫺1), (dy/dx)|−1⫽sinh(w1x1/T0), (10)
s ⫽T0/w1 sinh(w1x1/T0),

1

where the superscript denotes a negative sided limit. Continuity of distance and dis-
placement requires
s−1⫽s+1, y−1 ⫽y+1 (11)
where the superscripts denote negative and positive sided limits. The change in slope
across s=s1 is obtained from Eqs. (3) and (4) and noting dy/dx= tan q
(dy/dx)|+1⫺(dy/dx)|−1⫽(F(s+1)⫺F(s−1 ))/T0⫽W1/T0. (12)

Applying boundary conditions Eqs. (11) and (12) gives implicit relations for the
integration constants:
sinh(w2x1/T0⫹a2)⫽sinh(w1x1/T0)⫹W1/T0, (13)
b2⫽T0/w1(cosh(w1x1/T0)⫺1)⫺T0/w2cosh(w2x1/T0⫹a2), (14)
g2⫽s1⫺T0/w2 sinh(w2x1/T0⫹a2). (15)
R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914 903

3.3. Lower fairlead

The boundary conditions on the seaward side of the lower fairlead are
s⫽s2⫽s1⫹L2, y⫽H, T⫽T2, (16)
where L2, H and T2 have prescribed values. Substitutions for s, y, T, b2 andg2 from
above, gives


sinh(w1x1/T0)⫽ T 22/T 20−1⫺(W1⫹w2L2)/T0, (17)

b2⫽H⫺T2/w2, (18)
cosh(w2x1/T0⫹a2)⫺w2/w1 cosh(w1x1/T0)⫽(T2⫺w2H)/T0⫺w2/w1. (19)

4. Reduced boundary conditions

The boundary conditions reduce to a system of five coupled equations. Defining


a⫽w2/w1, b⫽T2⫺w2H, c⫽W1, d⫽w2L2, e⫽T2, t⫽1/T0, (20)
C1⫽ cosh(w1x1/T0), S1⫽ sinh(w1x1/T0), (21)
C2⫽cosh(w2x1/T0⫹a2), S2⫽sinh(w2x1/T0⫹a2), (22)
then the boundary conditions Eqs. (17), (13) and (19) and the identity cosh2()-
sinh2()=1, gives


S1⫽ e2t2−1⫺(c⫹d)t ⇒ S 21⫹2(c⫹d)S1t⫹((c⫹d 2)⫺e2)t2⫽⫺1, (23a,b)

S2⫽S1⫹ct, (24)

再 冑 冑 冎
C1⫽⫹ 1+S 21, C2⫽⫹ 1+S 22 ⇒ 兵C22⫺S 22⫽1, C21⫺S 21⫽1其, (25a-d)

C2⫹a(1⫺C1)⫺bt⫽0. (26)

The parameters a, b, c, d and e are prescribed and a solution is sought for t, C1,
S1, C2 and S2, from which horizontal and vertical spans, and end-point slopes are
readily obtained. This implicit approach gives a simpler set of boundary conditions
compared to directly solving for T0, x1, y1, x2, a2 and g2. The solution is also required
to satisfy the inequalities
t⬎1/e, S1⬎0, C1⬎1, C2⬎1. (27)
Combination of the first two inequalities and (23a), gives the reduced set
904 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914


t⬎1/ e2−(c+d)2, C1⬎1, C2⬎1. (28a-c)

5. Elongation

Central to the derivation of the catenary equations is the evaluation of

冕冑 冕
p x

1/ 1+t2dt⫽ wi / T0dt, (29)


pi xi

where p=dy/dx and if the mooring line elongates, wi is not uniform and is dependent
on tension. The left side of Eq. (29) is independent of elongation and is bound by
least and upper values of the right side
[wmin max
i (x⫺xi)/T0, wi (x⫺xi)/T0], (30)
where wmini and wmax
i are minimum and maximum values for wi. The ith component
line lies between two catenaries of uniform weights per unit length wmin
i and wmax
i

i x/T0⫹ai,1)ⱕdy/dxⱕ sinh(wi
sinh(wmin x/T0⫹ai,2),
max
(31)
T0/w min
i cosh(w min
i x/T0⫹ai,1)⫹bi,1ⱕyⱕT0/w max
i cosh(w max
i x/T0⫹ai,2)⫹bi,2, (32)
T0/wmin i x/T0⫹ai,1)⫹gi,1ⱕsⱕT0/wi
sinh(wmin x/T0⫹ai,2)⫹gi,2,
max
i sinh(wmax
i (33)
i x/T0⫹ai,1)ⱕT/T0ⱕ cosh(wi
cosh(wmin x/T0⫹ai,2)
max
(34)
where ai,j, bi,j and gi,j are integration constants derived in a similar manner to ai, bi
and gi,
The variation of wi along a line between values wmin i and wmax
i is partitioned as
wi⫽w0i⫹⌬wi, (35)
where the superscript 0 and ⌬ denote a nominal value and a (local) deviation due
to elongation. The line length is similarly partitioned,
Li⫽L0i⫹⌬Li. (36)

Influence of deviations ⌬wi, and ⌬Li on prescribed parameters of the reduced


boundary conditions is given by

a0 (w2/w1)0
⫽ 0 ⫺ 冉
⌬a ⌬(w2/w1) w02+⌬w2 w02 w02

w1+⌬w1 w01 w01 冊 (37)

⌬w1 ⌬w2 ⌬w1⌬w2 ⌬w1 2


⫽⫺ 0 ⫹ 0 ⫺ 0 0 ⫺ 0 ⫹O
w1 w2 w1 w2 w1
⌬w1 ⌬w1
w01
⫹ 0
w2 冉 冊 冉冉 冊冊3
, (38)
R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914 905

where O shows the term at which the Taylor series is truncated,


⌬b (⌬w2)H w02H/T2 ⌬w2 w02H ⌬w2
0
⫽ 0
⫽ 0 0
b T2+w2H 1+w2H/T2 w2

T2 w20
1⫹
w02H
T2
⫹O 冉 冊 冉冉 冊 冊 w02H
T2
3
⌬w2
w02
, (39)

⌬d ⌬(w2L2) (w02+⌬w2)(L02+⌬L2)−w02L02 ⌬w2 ⌬L2 ⌬w2 ⌬L2


⫽ ⫽ ⫽ 0⫹ 0⫹ 0 . (40)
d 0 (w2L2)0 w02L02 w2 L2 w2 L02

The ratios ⌬wi/w0i , w02H/T2 and ⌬Li/L0i are usually small and their products are of
second order. Hence (⌬a/a0, ⌬d/d 0) and ⌬b/b0 are respectively of first and second
orders.
An approximate relationship between ⌬wi, and ⌬Li is established by assuming the
overall immersed weight of a line is conserved under loading:
w0iL0i⫽(w0i⫹⌬wi)(L0i⫹⌬Li) (41)
⇔ ⌬wi/w0i⫽⫺ei/(1⫹ei)⫽⫺ei⫹e2i⫹O(e3i), ei⫽⌬Li/L0i, (42)
where the overbar denotes a mean value and ei is the strain based on the change in
Li0 Substituting Eq. (42) in Eq. (38) and Eq. (39) and Eq. (41) in Eq. (40), gives
⌬a/a0⬇e1⫺e2, ⌬b/b0⬇⫺(w02H/T2)e2, ⌬d/d 0⫽0, (43)
where, previously stated, ⌬b/b0 is of second order. Only a is significantly affected
by elongation and changes from the nominal value are of first order in strain.
The strain is likely to be between 0.001 and 0.1 for lines made of chain and wire
or synthetic rope. Since this cannot be easily measured during normal operational
procedures, the effect of elongation on static equilibrium can be determined to within
a small interval by assigning an uncertainty to the weight per unit length. The uncer-
tainty is of course real, along with those of the prescribed tension at the platform
T2 and water depth H.
The elongation relation Eq. (41) is less restrictive than Oppenheim and Wilson
(1982) and Ogawa (1984), where the immersed weight of an infinitesimal length is
conserved under loading. Further modelling of mooring line elongation is described
in Papoulias and Bernitsas (1988), including lateral contraction due to Poisson’s ratio.

6. Methods of solution

Four methods of solution are presented. The first is based on an iteration of Eqs.
(23)–(26) and is similar to Niedzwecki and Casarella (1976) and Orgill et al. (1985).
The second method introduces a polynomial approach for a special case of no point
buoy or sinker between the lines. The boundary conditions reduce to a single quartic
equation in 1/T0 where the roots have a closed form solution. The method is then
extended to include a buoy or sinker, and yields a polynomial equation in 1/T0 of
degree eight. The roots of the polynomial are obtained by Laguerre’s iteration which
is efficient and rapidly convergent in evaluating all the (complex) roots of a poly-
906 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914

nomial (Kincaid and Cheney, 1990; Press et al., 1992), and is also used as an alterna-
tive to the closed form solutions of cubics and quartics (NAG, 1993).
The set of extra solutions for T0 obtained by the polynomial method is an artifact
produced by twice or thrice squaring for the removal of radicals. Squaring an
algebraic equation is not an equivalence transformation as it generally doubles the
number of solutions. The choice of solution for T0 is not only restricted by Eq. (28)
but also requires lm(1/T0=0) and a final check by backward substitution into (23a),
(24), (25a,b) and (26). A slight modification is required in these checks for an
implementation where the arithmetic is of finite precision.
The reduced boundary conditions (23b)(24)(25c,d) and (26) form a polynomial
system and further reduction to a single polynomial may not be feasible for more
general mooring line arrangements. This is being addressed in our ongoing study by
globally convergent homotopy methods (Watson et al., 1987).

6.1. Method 1

On prescribing a to e, t is found by iterating through the sequence (23a), (24),


(25a,b) and (26) where Eq. (26) acts as an error function. A suitable starting value
for t, say t0, is given by (28a) and a monotonic search may proceed with
tn+1=1.01tn, where the subscript n denotes the number of iterations. More sophisti-
cated search strategies are described in Press et al. (1992).

6.2. Method 2: W1=0, w2⫽w1

This method is for a mooring with only two components lines of different weights
per unit length and yields a quartic equation in t. Substitutions from (23a), (24),
(25a,b) into Eq. (26), where now c=0, gives

(1⫺a) 冪1+(冑e t −1−dt) ⫽bt⫺a.


2 2 2
(44)

The removal of the square roots is aided by defining


a⫽1/(1⫺a), z1⫽(z2⫺dt)2, z2⫽ e2t2−1, 冑 (45)

then
冑1+z ⫽a(bt⫺a) ⇒ (z ⫺dt) ⫽a (bt⫺a) ⫺1
1 2
2 2 2
(46)

⇒4(e2t2⫺1)d 2t2⫽(a2(bt⫺a)2⫺e2t2⫺d 2t2)2 (47)


⇒r⬅r0t4⫹r1t3⫹r2t2⫹r3t⫹r4⫽0, (48)
where
r0⫽(a2b2⫺(d⫹e)2)(a2b2⫺(d⫺e)2), r1⫽⫺4a2ab(a2b2⫺d 2⫺e2), (49)
R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914 907

r2⫽2a2a2(3a2b2⫺(d 2⫺e2))⫹4d 2, r3⫽⫺4a4a3b, r4⫽a4a4 (50)


The roots for the quartic equation Eq. (48) have a closed form solution (Bronshtein
and Semendyayev, 1985), however, it offers no practical advantage over Laguerre’s
iteration which is also well suited for the more general problem of W1⫽⫽0.

6.3. Method 3

A derivation similar to the preceding method is applied with the inclusion of a


buoy or sinker between the lines. Substituting (23a) in Eq. (24) and both of these
in Eq. (26) by utilising Eq. (25), gives

冪1+(冑e t −1−dt) ⫺a冪1+(冑e t −1−(c+d)t) ⫽bt⫺a.


2 2 2 2 2 2
(51)

The removal of the multiplicity of square roots is aided by defining

z11⫽1⫹(z2⫺dt)2, z12⫽1⫹(z2⫺(c⫹d)t)2, z2⫽ e2t2−1. 冑 (52)

The outer square roots are removed by the sequence of operations,

冑 冑 冑
( z11)2⫽(bt⫺a⫹a z12)2 ⇒ (⫺2(bt⫺a)a z12)2⫽((bt⫺a)2⫺z11⫹a2z12)2 (53)

⇒ (bt⫺a)4⫺2(bt⫺a)2(z11⫹a2z12)⫹(z11⫺a2z12)2⫽0. (54)

As the removal of the square roots proceed, polynomials in t accumulate. These


are defined by qi, i=1,%,12 and the first two are, from Eq. (54):
q1⫽(bt⫺a)4, q2⫽⫺2(bt⫺a)2. (55)
The other qi are defined in the following and the coefficients are listed in Appendix
A. Substituting for z11 and z12 from Eq. (52) in Eq. (54), gives
q3z42⫹q4z32⫹q5z22⫹q6z2⫹q7⫽0, (56)
where q3, q4, q5, q6 and q7 are polynomials in t, respectively of degrees zero to four.
Substituting for z2 from Eq. (52) in Eq. (56) and isolating the square root, gives
q8⫽q4(e2t2⫺1)⫹q6, q9⫽q3(e2t2⫺1)2⫹q5(e2t2⫺1)⫹q7, (57)


⫺q8 e2t2−1⫽q9, (58)

where q8 and q9 are cubic and quartic in t. The final equation sequence for t is
q10⫽q28, q11⫽q29, q12⫽q11⫺q10(e2t2⫺1), (59)
q12⫽0, (60)
where q12 is a polynomial in t of degree eight.
908 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914

6.4. Method 4

An alternative solution sequence to Method 3 is now described. Symbolic equival-


ence between the polynomial coefficients is not readily apparent but numerical tests
have shown they are equal. It was for this confirmation that the method was derived.
Substituting for C1 from (25a) in Eq. (26) then for C2 from Eq. (26) and S2 from
Eq. (24) in (25c) gives

(a( 1+S 21⫺1)⫹bt)2⫺(S1⫹ct)2⫽1, (61)

and this expands to


a2(1⫹S 21)⫹(bt⫺a)2⫺(S1⫹ct)2⫺1⫽⫺2a(bt⫺a) (1+S 21). 冑 (62)

Squaring both sides of Eq. (62) and collecting terms yields a polynomial in S1, where
the coefficients are polynomials in t,
p1,0S 41⫹p2,1tS 31⫹(p3,2t2⫹p3,1t⫹p3,0)S 21⫹(p4,3t3⫹p4,2t2⫹p4,1t)S1⫹(p5,4t4 (63)
⫹p5,3t3⫹p5,2t2⫹p5,1t⫹p5,0)⫽0,
and the constants pi,j i=1%5, are defined in terms of a, b, c as listed in Appendix
A. Substituting for S1 from (23a) in Eq. (63) and isolating the square root, gives

p6,4t4⫹p6,3t3⫹p6,2t2⫹p6,1t⫹p6,0⫽(p7,3t3⫹p7,2t2⫹p7,1t) e2t2−1, (64)

where p6,j and p7,j are defined in terms of c, d, e, p1,j, p2,j, %, p5,j; see Appendix A.
Squaring both sides of Eq. (64) and collecting powers of t gives an equation of
degree eight,
p8,8t8⫹p8,7t7⫹p8,6t6⫹p8,5t5⫹p8,4t4⫹p8,3t3⫹p8,2t2⫹p8,1t⫹p8,0⫽0, (65)
where p8,j are defined in terms of e, p6,j and p7,j.

7. An example

A FORTRAN program has been written where the main part of the computation
is performed by a modified Laguerre method from the NAG library (NAG, 1993).
The execution time on an Intel-486 33 MHz processor is much less than one second.
The input parameters are partly based on data from Uittenbogaard and Pijfers (1996):
w1⫽100 kgf/m, W1⫽10 tef buoy, w2⫽20 kgf/m, L2⫽500 m, (66)
H⫽1000 m, T2⫽100 tef,
where kgf and tef respectively denote kilogram-force (9.81 N) and tonne-force (9810
N). This gives one solution for the mooring line problem:
T0⫽44.29 tef, s1⫽896.6 m, x1⫽644.2 m, H⫺y1⫽442.9 m, (67)
R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914 909

x2⫺x1⫽231.9 m, q2⫽63.7 deg.


Substitution of these values in the physical boundary conditions, Eqs. (11), (12) and
(16) shows the numerical error is negligible:

s+1⫺s11⫽⫺4⫻10−13 m, y+1⫺y−1⫽⫺1⫻10−12 m, 冉 冊冒 冉 冊
dy
dx
+
1
dy − W1
⫺ ⫽⫺3
dx 1 T0
(68)

⫻1016,
s2⫺(s1⫹L2)⫽2⫻10−13 m, y2⫺H⫽5⫻10−13 m, T 20+F2(s2)⫺T2⫽6 冑 (69)
⫻10−15 tef.
The error for the tension at the fairlead is evaluated from the vector sum of Eq.
(3a,b), rather than from Eq. (6b) which is trivially related to Eq. (18). Other solution
details are listed below. Table 1 lists all the roots of the mooring polynomial.

7.1. Integration constants

a2⫽1.059, b2⫽⫺4000 m, g2⫽⫺3086 m (70)

7.2. Transformed input parameters

a⫽0.2, b⫽80 tef, c⫽⫺10 tef, d⫽10 tef, e⫽100 tef (71)

Table 1
Roots for t tef ⫺1 and respective values of the polynomial

no Re(t) Im(t) Re(poly) Im(poly)

1 2.258×10⫺2 0.000 ⫺4.094×10⫺15 0.000


2 ⫺2.218×10⫺2 0.000 7.281×10⫺16 0.000
3 ⫺5.096×10⫺3 2.316×10⫺3 ⫺1.148×10⫺19 ⫺7.191×10⫺19
4 ⫺5.096×10⫺3 ⫺2.316×10⫺3 ⫺1.148×10⫺19 7.191×10⫺19
5 1.018×10⫺3 5.464×10⫺4 0.000 ⫺1.069×10⫺22
6 1.018×10⫺3 ⫺5.464×10⫺4 0.000 ⫺1.069×10⫺22
7 8.012×10⫺4 3.987×10⫺4 0.000 ⫺2.374×10⫺22
8 8.012×10⫺4 ⫺3.987×10⫺4 0.000 2.374×10⫺22
910 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914

7.3. Polynomial equation for t tef-1

⫺1.526⫻1014t8⫺9.389⫻1011t7⫹7.687⫻1010t6⫹5.104⫻108t5⫺5.059 (72)
⫻104t4⫺4.772⫻103t3⫹10.06t2⫺8.192⫻10−3t⫹5.260⫻10−6⫽0

8. Conclusions

There are many situations where an estimation scheme for a static catenary will
be sufficient or where a more capable discretisation scheme can readily be used.
There are, however, situations where the solutions presented here are valuable. It is,
for example, important in a stability measurement system (such as the MOSIS
system) to minimise sources of error and if the uncertainty of the solution can be
directly related to uncertainties in the prescribed parameters then that allows easier
acceptance of the results.
In addition there are deep water mooring configurations of the form in Fig. 1 that
are sensitive to initial estimates and solution tolerances. The techniques described
here provide alternative, more robust convergence and where there is no buoy or
sinker the solution is in closed form. Finally these solutions are computationally
efficient and there will always be situations where an efficient mooring kernel has
value.

Acknowledgements

The authors are grateful to MOSIS Systems Ltd for supporting this work.

Appendix A

Coefficients of the polynomials in Methods 3 and 4 are listed below. Subscripts


denote position in the sequence of derivation and degree of the term containing
the coefficient.

A.1. Method 3

q1⫽q1,4t4⫹q1,3t3⫹q1,2t2⫹q1,1t⫹q1,0 (73)
q1,4⫽b4, q1,3⫽⫺4ab3, q1,2⫽6a2b2, q1,1⫽⫺4a3b, q1,0⫽a4 (74)
q2⫽q2,2t ⫹q2,1t⫹q2,0
2
(75)
R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914 911

q2,2⫽⫺2b2, q2,1⫽4a,b, q2,0⫽⫺2a2 (76)

The coefficients of q1, q2 have been directly substituted in the following poly-
nomials
q3⫽q3,0 (77)
q3,0⫽(a2⫺1)2 (78)
q4⫽q4,1t (79)
q4,1⫽4(a2⫺1)(⫺a2(c⫹d)⫹d) (80)
q5⫽q5,2t ⫹q5,1t⫹q5,0
2
(81)
q5,2⫽2a2(c⫹d))(3a2(c⫹d)⫺(b2⫹c⫹5d⫹d 2)⫺2(b2⫺3d 2) (82)
q5,1⫽4ab(a ⫹1), q5,0⫽⫺2(3a ⫺1)
2 2
(83)
q6⫽q6,3t3⫹q6,2t2⫹q6,1t (84)
q6,3⫽4a (c⫹d)(⫺a (c⫹d) ⫹b ⫹cd⫹2d )⫹4d(b ⫹d )
2 2 2 2 2 2 2
(85)
q6,2⫽⫺8ab(a2(c⫹d)⫹d), q6,1⫽4(c⫹3d)a2⫺4d (86)
q7⫽q7,4t ⫹q7,3t ⫹q7,2t ⫹q7,1t⫹q7,0
4 3 2
(87)
q7,4⫽a2(c⫹d)2(a2(c⫹d)2⫺2(b2⫹d 2))⫹(b2⫺d 2)2, q7,3⫽4ab(a2(c⫹d)2⫺b2 (88)
⫹d ) 2

q7,2⫽2(2a2⫺1)(b2⫺d 2)⫺2a2(c⫹d)2, q7,1⫽4ab, q7,0⫽1⫺4a2 (89)


q8⫽q8,3t ⫹q8,2t ⫹q8,1t
2 2
(90)
q8,3⫽(q6,3⫹e2q4,1), q8,2⫽q6,2, q8,1⫽q6,1⫺q4,1 (91)
q9⫽q9,4t ⫹q9,3t ⫹q9,2t ⫹q9,1t⫹q9,0
4 3 2
(92)
q9,4⫽e4q3,0⫹q5,2e2⫹q7,4, q9,3⫽e2q5,1⫹q7,3, q9,2⫽e2(q5,0⫺2q3,0)⫺q5,2 (93)
⫹q7,2
q9,1⫽q7,1⫺q5,1, q9,0⫽q3,0⫺q5,0⫹q7,0 (94)
q10⫽q10,6t ⫹q10,5t ⫹q10,4t ⫹q10,3t ⫹q10,2t
6 5 4 3 2
(95)
q10,6⫽q28,3, q10,5⫽2q8,2q8,3, q10,4⫽2q8,1q8,3⫹q28,2, q10,3⫽2q8,1q8,2, q10,2 (96)
⫽q 2
8,1

q11⫽q11,4t4⫹q11,3t3⫹q11,2t2⫹q11,1t⫹q11,0 (97)
q11,8⫽q , q11,7⫽2q9,3q9,4, q11,6⫽2q9,2q9,4⫹q , q11,5⫽2(q9,1q9,4
2
9,4
2
9,3 (98)
⫹q9,2q9,3)
q11,4⫽2(q9,0q9,4⫹q9,1q9,3)⫹q29,2, q11,3⫽2(q9,0q9,3⫹q9,1q9,2), q11,2⫽2q9,0q9,2 (99)
912 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914

⫹q29,1
q11,1⫽2q9,0q9,1, q11,0⫽q29,0 (100)
q12⫽q12,8t ⫹q12,7t ⫹q12,6t ⫹q12,5t ⫹q12,4t ⫹q12,3t ⫹q12,2t ⫹q12,1t⫹q12,0
8 7 6 5 4 3 2
(101)
q12,8⫽q11,8⫺e2q10,6, q12,7⫽q11,7⫺e2q10,5, q12,6⫽q10,6⫺e2q10,4⫹q11,6 (102)
q12,5⫽q10,5⫺e q10,3⫹q11,5, q12,4⫽q10,4⫺e q10,2⫹q11,4, q12,3⫽q10,3⫹q11,3
2 2
(103)
q12,2⫽q10,2⫹q11,2, q12,1⫽q11,1, q12,0⫽q11,0 (104)

A.2. Method 4

p1⫽p1,0 (105)
p1,0⫽(1⫺a2)2 (106)
p2⫽p2,1t (107)
p2,1⫽4c(1⫺a2) (108)
p3⫽p3,2t ⫹p3,1t⫹p3,0
2
(109)
p3,2⫽⫺2(1⫹a2)b2⫹2(3⫺a2)c2, p3,1⫽4ab(1⫹a2), p3,0⫽2(1⫺3a2) (110)
p4⫽p4,3t ⫹p4,2t ⫹p4,1t
3 2
(111)
p4,3⫽4(c2⫺b2)c, p4,2⫽8abc, p4,1⫽4(1⫺2a2)c (112)
p5⫽p5,4t ⫹p5,3t ⫹p5,2t ⫹p5,1t⫹p5,0
4 3 2
(113)
p5,4⫽(b2⫺c2)2, p5,3⫽⫺4ab(b2⫺c2), p5,2⫽2(2a2⫺1)(b2⫺c2), p5,1 (114)
⫽4ab, p5,0⫽1⫺4a 2

p6⫽p6,4t4⫹p6,3t3⫹p6,2t2⫹p6,1t⫹p6,0 (115)
p6,4⫽p1,0(c⫹d) ⫺p2,1(c⫹d) ⫹p3,2⫹6e p1,0)(c⫹d) ⫺(p4,3⫹3e p2,1)(c
4 3 2 2 2
(116)
⫹d)⫹e2(e2p1,0⫹p3,2)⫹p5,4
p6,3⫽((c⫹d)2⫹e2)p3,1⫹p5,3⫺(c⫹d)p4,2 (117)
p6,2⫽(c⫹d)2(p3,0⫺6p1,0)⫹(c⫹d)(3p2,1⫺p4,1)⫹e2(p3,0⫺2p1,0)⫺p3.2⫹p5,2 (118)
p6,1⫽p5.1⫺p3,1, p6,0⫽p1,0⫺p3,0⫹p5.0 (119)
p7⫽p7,3t3⫹p7,2t2⫹p7,1t (120)
p7,3⫽⫺4(c⫹d)3p1,0⫹3(c⫹d)2p2,1⫺2(c⫹d)(2e2p1,0⫹p3,2)⫹e2p2,1⫹p4,3 (121)
p7,2⫽⫺2(c⫹d)p3,1⫹p4,2, p7,1⫽2(c⫹d)(2p1,0⫺p3,0)⫹p4,1⫺p2,1 (122)
R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914 913

p8⫽p8,8t8⫹p8,7t7⫹p8,6t6⫹p8,5t5⫹p8,4t4⫹p8,3t3⫹p8,2t2⫹p8,1t⫹p8,0 (123)
p8,8⫽p ⫺e p , p8,7⫽2(p6,3p6,4⫺e p7,2p7,3)
2
6,4
2 2
7,3
2
(124)
p8,6⫽2p6,2p6,4⫹p ⫺e (2p7,1p7,3⫹p ⫹p
2
6,3
2 2
7,2
2
7,3 (125)
p8,5⫽2((p7,3⫺e2p7,1)p7,2⫹p6,1p6,4⫹p6,2p6,3) (126)
p8,4⫽2(p6,0p6,4⫹p6,1p6,3)⫹p ⫹(2p7,3⫺e p7,1)p7,1⫹p
2
6,2
2 2
7,2 (127)
p8,3⫽2(p6,0p6,3⫹p6,1p6,2⫹p7,1P7,2), p8,2⫽2p6,0p6,2⫹p26,1⫹p27,1, p8,1 (128)
⫽2p6,0p6,1⫽2p6,0p6,1
p8,0⫽p26,0 (129)

References

Ansari, K.A., 1980. Mooring with multi-component cable systems. Transactions of the ASME — Journal
of Energy Resources Technology 102, 62–69.
Ansari, K.A., Khan, N.V., 1986. The effect of cable dynamics on the station-keeping response of a moored
offshore vessel Transactions of the ASME. Journal of Energy Resources Technology 108, 52–58.
Avallone, E.A., Baumeister, T. (Eds.), 1986. Marks’ Std Handbook for Mechanical Engineers, 9th ed.
McGraw-Hill, New York.
Bradley, M.S., MacFarlane, C.J., 1986a. Inclining tests in service. In: Kuo, C. (Ed.) Stationing and Stab-
ility of Semi-submersibles. Graham & Trotman (Kluwer), London, pp. 83–109.
Bradley, M.S., MacFarlane, C.J., 1986b. The in-service measurement of hydrostatic stability. In Proceed-
ings of the Eighteenth Annual Offshore Technology Conference, Houston, Texas, 4, 491–508.
Bradley, M.S., MacFarlane, C.J., 1995. Some lessons to be learned from the stability control of semi-
submersibles. In Offshore 95 — Design and Safety Assessment for Floating Installations — Confer-
ence Papers, Institute of Marine Engineers, London.
Bronshtein, I.N., Semendyayev, K.A., 1985. In: Hirsch, K.A. (Ed.), Handbook of Mathematics, 3rd ed.
van Nostrand Reinhold, Frankfurt and New York.
Brown, D.T., 1997. Convergence of non-linear wave loading on a catenary moored floating production
system design. International Shipbuilding Progress 44, 161–176.
Faltinsen, O.M., 1990. Sea Loads on Ships and Offshore Structures. Cambridge University Press, Cam-
bridge.
Jain, R.K., 1980. A simple method of calculating the equivalent stiffnesses in mooring cables. Applied
Ocean Research 2, 139–142.
Kincaid, D., Cheney, W., 1990. Numerical Analysis - Mathematics of Scientific Computing. Brooks-Cole,
Pacific Grove, California.
Koburt, M.D., Herbert, E.J., 1970. Some notes on static anchor chain curve. In Second Annual Offshore
Technology Conference, Houston, Texas, (pre-prints) 1, 147–160.
MacFarlane, C.J., 1996. Discussion of Brown D.T., Witz, J.A. (Eds.). Estimation of vessel stability at
sea using roll motion records, Transactions of the Royal Institution of Naval Architects, 138, 130–145.
Mavrakos, S.A. et al., 1996. Deep water mooring dynamics. Marine Structures 9, 181–209.
Meriam, J.L., 1975. Statics, 2nd ed. John Wiley & Sons, New York (SI-Version).
NAG, 1993. The NAG FORTRAN Library Manual - Mark 16, The Numerical Algorithms Group, Oxford.
Niedzwecki, J.W., Casarella, M.J., 1976. On the design of mooring lines for deep water applications.
Transactions of the ASME-Journal of Engineering for Industry 98, 514–522.
Ogawa, Y., 1984. Fundamental analysis of deep sea mooring line in static equilibrium. Applied Ocean
Research 6, 140–147.
914 R.J. Smith, C.J. MacFarlane / Ocean Engineering 28 (2001) 899–914

Oppenheim, B.W., Wilson, P.A., 1982. Static 2-D solution of a mooring line of arbitrary composition in
the vertical and horizontal operating modes. International Shipbuilding Progress 29, 142–153.
Orgill, G., Wilson, J.F., Schmertmann, G.R., 1985. Static design of cable mooring arrays for offshore
guyed towers. Applied Ocean Research 7, 166–174.
Papoulias, F.A., and Bernitsas, M.M., 1988. MOORINE: A Program for Static Analysis of MOORing
LINES. Report to The University of Michigan/Sea Grant/Industry Consortium in Offshore Engineer-
ing, and Department of Naval Architecture and Marine Engineering. The University of Michigan Ann
Arbour, Publication 309.
Press, W.H., Flannery, B.P., Teukolsky, S.A., Vetterling, W.T., 1992. Numerical Recipes in FORTRAN —
The Art of Scientific Computing, 2nd ed. Cambridge University Press, Cambridge.
Rothwell, A., 1979. A graphical procedure for the stiffness of a catenary mooring. Applied Ocean
Research 1, 217–219.
Skop, R.A., 1988. Mooring systems — a state of the art review. Transactions of the ASME — Journal
of Offshore Engineering and Arctic Engineering 110, 365–372.
Sun, Y., Leonard, J.W., 1998. Dynamics of ocean cables with low local-tension regions. Ocean Engineer-
ing 25, 443–463.
Symon, K.R., 1971. Mechanics, 3rd ed. Addison-Wesley, Reading, Massachusetts.
Uittenbogaard, R., Pijfers, J., 1996. Integrated asymmetric mooring and hybrid riser system for turret
moored vessels in deepwater. In Proceedings of the Twenty-eighth Annual Offshore Technology Con-
ference, Houston, Texas, 3, 67–78.
Watson, L.T., Billups, S.C., Morgan, A.P., 1987. Algorithm 652 — Hompack — A suite of codes for
globally convergent homotopy algorithms. ACM Transactions on Mathematical Software 13, 281–310.

You might also like