You are on page 1of 14

Applied Ocean Research 22 (2000) 169–182

www.elsevier.com/locate/apor

Development and validation of a lumped-mass dynamics model of a


deep-sea ROV system
F.R. Driscoll a,*, R.G. Lueck b, M. Nahon c
a
Department of Mechanical Engineering and School of Earth and Ocean Sciences, University of Victoria, P.O. Box 3055, Victoria, British Columbia,
Canada, V8W 3P6
b
Centre for Earth and Ocean Research, University of Victoria, P.O. Box 3055 Victoria, British Columbia, Canada, V8W 3P6
c
Department of Mechanical Engineering, University of Victoria,, P.O. Box 3055 Victoria, British Columbia, Canada, V8W 3P6
Received 23 March 1999; received in revised form 6 December 1999; accepted 13 December 1999

Abstract
A one-dimensional finite-element lumped-mass model of a vertically tethered caged ROV system subject to surface excitation is
presented. Data acquired during normal operation at sea are used with a least-squares technique to estimate the coefficients required by
the model. The model correctly predicts: (a) the motion of the cage and the tension in the tether at the ship; (b) the spectrum of cage
acceleration in the wave-band; (c) the transfer function between the vertical motion of the ship and cage; and (d) the natural frequency of the
system and its harmonics. A simple model of the wake of the cage was added to the model simulation and this reduced the error in the
calculated motion and tension by almost a factor of two and brought the calculated transfer function within the 95% confidence interval of the
measurements. By increasing ship motion slightly, the model accurately reproduced eight snap loads and their non-linear characteristics—a
regularly spaced series of rapid increases in the records of the acceleration of the cage and tension at the ship—that occurred during the
measurements used for validation. 䉷 2000 Elsevier Science Ltd. All rights reserved.
Keywords: Tethered marine systems; Finite element analysis; Offshore engineering; Remotely operated vehicle; Snap loading; Cable modelling

1. Introduction water sampling rosettes, tethered equipment being lowered


to the ocean floor, and piston coring tools. Reliable models
Caged remotely operated vehicles (ROVs) are robots that are needed to predict the behaviour of vertically tethered
effectively operate in the dangerous high-pressure environ- systems so that designers can minimise motions and
ment of deep water while their human controllers remain tensions and users can know which sea states will produce
safely aboard the support ship (Fig. 1). The tether connect- snap loads.
ing the ship and cage provides the ROV with the ability to Two tools are available to predict the response of verti-
stay at the underwater work site indefinitely and allows real- cally tethered systems—continuous (closed-form analytic)
time communication with the ROV. However, the tether and discrete models. Continuous models are powerful tools
also couples the ship motion to the cage, and, as a result, and they have been used to accurately calculate many of the
these systems cannot operate in rough seas. Rough seas dynamic characteristics of tethered systems, including
cause slack tether, large snap loads and erratic cage motion motion and tension spectra, transfer functions, natural
[1], which damage the tether and its internal electrical and frequency and its harmonics, and the onset of slack in the
optical conductors. The resulting degradation reduces the tether [2–5]. However, continuous models have limitations.
life of the tether and endangers the recovery of the ROV. Equivalent linearisation techniques are used to approximate
Many other vertically tethered systems are dynamically the quadratic drag for steady amplitudes, which enable
similar to caged ROVs, including CTDs (conductivity closed form solutions and frequency-domain analysis [3–
temperature and depth measuring platforms), drill strings, 6]. Thus, these models are not appropriate during unsteady
wave conditions. Continuous models are also invalid for a
slack tether, which is the precursor to snap-loads, and it is
* Corresponding author. Present address: Department of Ocean Engineer-
ing, Florida Atlantic University, 101 North Beach Road, Dania Beach,
difficult (if not impossible) to solve the differential equa-
Florida 33004, USA. Tel.: ⫹ 954-924-7221; fax: ⫹ 954-924-7233. tions governing a tether when hydrodynamic forces and its
E-mail address: rdriscol@oe.fau.edu (F.R. Driscoll). properties (stiffness, cross-section, material) change along
0141-1187/00/$ - see front matter 䉷 2000 Elsevier Science Ltd. All rights reserved.
PII: S0141-118 7(00)00002-X
170 F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182

Fig. 1. Diagrammatic representation of the ROPOS ROV system consisting of a support ship (C.S.S. John P. Tully), winch, umbilical tether, cage and vehicle.
ROPOS is investigating the actively venting sulfide structure Godzilla [20].

its length. Discrete models are not constrained by the limita- must be calibrated using measured data to obtain confidence
tions of continuous models and they are valid for a wider in a model. Additionally, discrete models are approxima-
range of operating conditions and system configurations. tions to real systems, and therefore, validation against real
Niedzwecki and Thampi [7] developed a discrete multi- data is essential in proving their accuracy and reliability. In
segment lumped-parameter representation of a drill string the literature search, we found one work that calibrates
connecting a ship to a subsea package and used it to inves- coefficients of a deep towed vehicle using data [8] and
tigate snap loading. They reproduced many of the snap load one that compares results of a discrete model of a towed
and motion characteristics that were observed by Driscoll et system against its actual motion [9]. However, no work was
al. [1] for a real system, namely, a repeating pattern of found that uses data to calibrate and validate discrete models
rapidly increasing tension during snap loads and vertical of vertically tethered systems.
motion at the platform that was greater than at the ship. In this paper, a finite-element lumped-mass model of a
The predicted behaviour can be sensitive to the model vertically tethered ROV system is developed by applying
coefficients, such as added mass [7], and choosing represen- Galerkin’s method to a continuous representation of the
tative values is difficult because many components are tether. Motion and tension measurements of the ROPOS
complex and asymmetric (armoured/faired tether, ROV ROV system [1] are used to accurately determine values
cages and Conductivity Temperature and Depth measuring for the model coefficients. In particular, a least-squares tech-
platforms). For example, Yoerger et al. [8] found that the nique is used to estimate the drag coefficient and the virtual
normal drag coefficient of a long vertical tow cable is ampli- mass (including the system mass, the hydrodynamic added
fied well above the value for a rigid cylinder and that it and the entrained mass) of the ROV cage. The vertical
depends on operating conditions. Therefore, coefficients motion of the top element is prescribed here along with
F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182 171

roll, pitch and yaw motions is small compared to the vertical


scale O(1:1000) of the tether and such motions are not
effectively transmitted by the tether. Therefore, a one-
dimensional model is sufficient to represent a caged ROV
system. For modelling purposes, the system is broken down
into three individual sub-systems: the support ship with its
A-frame, the tether, and the cage and ROV. A finite-element
lumped-mass model is used to represent the tether, a displa-
cement function to prescribe the ship motion and a second-
order differential equation to represent the cage and ROV
unit. The model is developed using a single inertial refer-
ence frame located at the mean sea surface, with downward
displacements defined as positive (Fig. 2).

2.1. Tether model

The tether is a long elastic structural member that can


only sustain tensile loads and may have length-varying
properties. We represent the tether as a long elastic cylinder
that is suspended directly below the ship’s A-frame. Defin-
ing s as the vertical unstretched co-ordinate along the tether,
the balance of forces (per unit length) at any point in the
tether is [10]:
2T
⫹ W ⫹ F ˆ ma …1†
2s
where
W ˆ …rc ⫺ rw †Ag …2†
is the in situ weight per unit length, T(s) is the tension in the
tether, F(s) is the hydrodynamic force per unit length, m is
the effective mass per unit length and includes the entrained
water, a is the inertial acceleration, g is the gravitational
acceleration, A(s) is the cross-sectional area of the tether,
rc …s† is the density of the tether and rw is the density of sea
water. The added mass forces are zero for flow tangent to a
cylinder [11], and therefore, F(s) is only composed of drag
forces. The hydrostatic pressure induces circumferential and
Fig. 2. Diagrammatic representation of the finite-element, including the radial stress and a depth dependent axial strain in the amour
reference and elastic vectors for node i. strands of the tether [12]. The ratio of the standard deviations
of the strain induced by the tension to the strain induced by the
the actual ship motion and comparing the calculated and pressure at any point in the tether has a numeric value of
measured motion of the cage and tension in the tether vali- approximately 1000 because the tether only oscillates short
dates the model. A simple representation of the wake of the vertical distances (⬍3 m). Thus, hydrostatically induced axial
cage is also developed and its addition significantly strain is considered constant in time and it is neglected in the
improves the model accuracy. Finally, we compare the analysis. The strain in the tether is less than 0.004 of the
predicted and actual response during snap loads. manufacturer’s maximum recommended working load and
the armour is wrapped around an incompressible core; there-
fore, the dilation of the tether will be very small and its cross-
2. Discrete model section is assumed to be constant. Further, assuming the water
is stationary in the far field, the hydrodynamic force per unit
Experimental measurements have shown that when a length, F(s), can be expressed as [13]:

typical deep-sea ROV system operates in the absence of F…s† ˆ ⫺ 12 …e ⫹ 1†rw DCDT U_ U_ …3†
time varying currents or large horizontal excursions of the
support vessels, the ship and cage are only coupled verti- where e is the local strain in the tether, D(s) is the diameter of
cally [1]. The displacement of the cage due to surge, sway, the tether, CDT is the tangential drag coefficient, ‘·’ overhead
172 F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182

represents its time derivative and U_ is the velocity of the tether. methods, such as collocation, least squares and least squares
The relationship between tension and axial strain is: collocation, because it is more prevalent in literature [16].
Knowing that rc and A are uniform or change very slowly
T…s† ˆ EAe …4†
[…li =rc †…2rc =2s† Ⰶ 1 or … li =A†…2A=2s† Ⰶ 1] along an element
where E is the effective Young’s modulus of the tether. and that the strain in the tether is small …e Ⰶ 1†; the elemen-
A finite-element implementation was chosen because it is tal equations of element i are:
versatile and can model the complicated non-linear charac- " # " #2 …1† 3
teristics of vertically tethered systems. For example, one can mi li Z 1=3 1=6 u
4 i 5
⫹ mi li
quickly assemble complex models of multi-component 2 Z 1=6 1=3 u …2†
i
systems connected by tethers with different and length depen-
" #2 …1† 3 2 …1† 3 2 …1† 3
dent properties because the elemental tether equations are EA 1 ⫺1 u f b
developed from a generalised linear-elastic tether subject to ⫹ i i 4 i 5 ˆ 4 i 5 ⫹ 4 i 5 …8†
unspecified boundary conditions. Components, such as ROV
li ⫺1 1 u…2†i f …2† i b…2† i
cages, weights/floats, and intermediate platforms can be easily where Ei is the effective Young’s modulus of the tether and
added because they are each represented as a point mass and s_i ˆ si ˆ 0: The consistent hydrodynamic, gravitational and
the hydrodynamic, gravitational and buoyancy forces acting buoyancy forces at upper and lower nodes are:
on them are described by one differential equation.
The tether is discretised by first dividing it into N fi…1† ˆ 1
2 …mi ⫺ rw Ai †gli ⫺ 1
2 rw Di CDT
i
li
segments (elements) which have the same properties as
the continuous tether. The end points of each element are  ‰ 13 …Z_ ⫹ u…1† _ _ …1†
i †jZ ⫹ u i j⫹
1
6 …Z_ ⫹ u_ …2† _ _…2†
i †jZ ⫹ u i j
the nodes of the system. Each element i is of length li and
has an associated body-fixed reference frame attached to the ⫺ 1
12 …u_…1†
i ⫺u_ …2† _…1†
i †ju _ …2†
i ⫺u i jŠ …9a†
upper node with the z-axis pointing downward along the
element. The top of the tether is attached to the ship’s A- fi…2† ˆ 1
2 …mi ⫺ rw Ai †gli ⫺ 1
2 rw Di CDT
i
li
frame. To facilitate the finite-element derivation, the loca-
tion of any point p in element i is defined using a coupled set  ‰ 16 …Z_ ⫹ u_ …1† _ _ …1†
i †jZ ⫹ u i j⫹
1
3 …Z_ ⫹ u_ …2† _ _…2†
i †jZ ⫹ u i j
of reference and elastic co-ordinates [14] (Fig. 2):
⫺ 1
12 …u_…1†
i ⫺u_ …2† _…1†
i †ju _ …2†
i ⫺u i j …9b†
zip ˆ Z ⫹ sip ⫹ uip …5†
where Di is the diameter of the tether and the boundary
where Z is the distance from the mean sea surface to the top conditions for the continuity of strain between nodes of
of the tether at node 1, in the inertial frame, sip is the connecting elements at the upper and lower nodes, respec-
constant length position vector from node 1 to a point p tively, are:
along the undeformed tether and uip is the cumulative elastic
2u~
displacement of the tether up to point p. An elemental trial b…1†
i ˆ Ei A i f…s…1†
i †i;1 …10a†
solution, u~ ip [15], is used to approximate the elastic displa- 2s
cement, uip : 2u~  …2† 
b…2†
i ˆ Ei A i f si …10b†
X
n
2s i;2
uip ⬇ u~ ip …s; t; d† ˆ d…i j† …t†fi; j …s† …6†
jˆ1
2.2. Ship model
where fi; j …s† is a shape function, n ⫺ 1 is the order of the
shape function and di… j† …t† is a time-dependent coefficient The vertical motion of the ship is assumed to be known
(yet to be defined). Each tether element has only one degree and is treated as a prescribed displacement function applied
of freedom and a linear shape function is sufficient to to the upper end (first node) of the tether, Z ˆ ZS …t†:
approximate the elastic displacement of any point p in Because the motion of the first node is prescribed, measured
element i, that is n ˆ 2 and (actual) ship motion data can be used as input. Thus, the
model can be validated by directly comparing the calculated
s…2†
i ⫺s s ⫺ s…1† cage motion and tension in the tether against the actual
fi;1 …s† ˆ fi;2 …s† ˆ i
si ⫺ s…1†
…2†
i s…2†
i ⫺ s…1†
i …7† measured values for the same ship motion.

d…1†
i ˆ u…1†
1 di
…2†
ˆ u…2†
i
2.3. Cage and ROV model

where … †…1† …2†


i and … †i represent the upper and lower nodes of The ratio of tether length to cage height is large, O(1000),
element i, respectively. Galerkin’s method of weighted resi- and thus, the cage can be represented as a single point mass
duals [15] is applied directly to the elemental force balance that is attached to the lower end of the tether and subject to
Eq. (1) to yield the discrete, finite-element, equations. the forces of gravity, buoyancy, drag and tension. When the
Galerkin’s method was chosen over other weighted residual ROV is in the cage, its mass, weight and buoyancy are
F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182 173

added to those of the cage. When the ROV is out of the cage, end points; therefore, the boundary forces, b…2† i and b…1†
i⫹1 ;
the tether connecting the ROV and cage is slack and MROV cancel when the elemental equations are assembled. The
and WROV are set to zero. Thus, the force balance for the cage and ROV are included by the addition of their mass
heave motion at the cage is: to that of node N ⫹ 1 and applying the forces acting on them
to that node. The boundary condition at the bottom end of
…MCG ⫹ MROV †…Z ⫹ u …2†
N †⫹
1
2 rw AC CD …Z_ ⫹ u_…2† _ _ …2†
N †jZ ⫹ u N j the tether is given by Eq. (11). The consistent-mass matrix is
replaced with a lumped-mass matrix to uncouple the accel-
ˆ WCG ⫹ WROV ⫺ T…L† …11†
eration terms [17]. The set of N second-order differential
where MCG and MROV are the virtual mass (including equations that govern the motion of the vertically tethered
entrained and added mass) of the cage and ROV, system are:
respectively, AC is the effective horizontal cross- Mz ⫹ Ku ˆ f …12†
sectional area of the cage, CD is the drag coefficient ÿ 
of the cage, WCG and WROV are the weight less buoyancy where the lumped-mass matrix, M 僆 RN× N⫹1
; is:

2 3
0 m1 l1 ⫹ m2 l2 0 … 0 0
6 7
60 m2 l2 ⫹ m3 l3 … 7
6 0 0 0 7
6 7
16 .. .. .. .. .. 7
Mˆ 6 7 …13†
26
6
. . . ] . . 7
7
6 7
60 0 0 … mN⫺1 lN⫺1 ⫹ mN lN 0 7
4 5
0 0 0 … 0 mN ⫹ 2MCG ⫹ 2MROV
ÿ 
the stiffness matrix, K 僆 R N× N⫹1
; is:
2 3
E 1 A1 E 1 A1 E A E2 A2
6⫺ l ⫹ 2 2 ⫺ … 0 0 7
6 l l2 l2 7
6 1 1
7
6 7
6 E A E2 A 2 E A 7
6 0 ⫺ 2 2 ⫹ 3 3 … 0 0 7
6 l2 l2 l3 7
6 7
6 7
6 .. .. .. .. .. 7
Kˆ6 . . . . . 7 …14†
6 ] 7
6 7
6 EN⫺1 AN⫺1 E A EN AN 7
6 … ⫹ N N ⫺ 7
6 0 0 0 7
6 lN⫺1 lN lN 7
6 7
6 7
4 EN AN EN AN 5
0 0 0 … ⫺
lN lN

of the cage and ROV, respectively, T…L† is the tension in the the external force matrix, f 僆 RN×1 ; is:
tether at the cage termination, u_ …2†
N is the elastic displacement
of the tether at the cage termination, and here ‘·’ and ‘··’ 2 3
overhead represent first and second time derivatives, f1…2† ⫹ f2…1†
6 7
respectively. 6 …2† 7
6 f2 ⫹ f3…1† 7
6 7
6 7
6 . 7
fˆ6 .. 7
2.4. Model assembly 6 7
6 7 …15†
6 …2† …1† 7
6 fN⫺1 ⫹ fN 7
The N tether elements are assembled end-to-end and 4 5
joined at their node points as shown in Fig. 2. Nodes are fN…2† ⫹ fCG
numbered consecutively from the top node, 1, to the bottom
node, N ⫹ 1: The heave acceleration at the ship is _ …2†
fCG ˆ WCG ⫹ WROV ⫺ 1
2 rw AC CD …Z_ ⫹ u…2†
N † Z ⫹ u N
prescribed as the acceleration of node 1. The tether is
contiguous at each node point ranging between 2 and N;
therefore, the trial functions of connected elements must the acceleration of the top node, 1, is:
be identically equal at their connecting nodes. Additionally,
continuity requires the strain boundary conditions of
connected elements to be equal and opposite at coincident Z ˆ z1 ˆ Z S …t† …16†
174 F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182

Fig. 3. ROPOS in the cage during a launch from the R/V Sonne, 1997.

and the position, z 僆 RN×1 ; and elastic displacement, u 僆 method takes small time steps during periods of rapid
RN×1 ; vectors are, respectively: change and large time steps when changes are slow [18].
2 3 2 3 Thus, good accuracy is achieved with only moderate
z1 Z
6 7 6 7 computational effort and errors and instabilities introduced
6 z 7 6 Z ⫹ s…1† ⫹ u…1† 7
6 2 7 6 2 2 7 by the previous time step information are significantly
6 7 6 7
6 7 6 …1† 7 reduced.
6 z3 7 6 Z ⫹ s…1† ⫹ u 3 7
6 7 6 3 7
zˆ6 6 . 7 6
7ˆ6
.
7
7 …17†
6 .. 7 6 .. 7
6 7 6 7 3. Data
6 7 6 7
6 7 6 …1† …1† 7
6 zN 7 6 Z ⫹ sN ⫹ uN 7
4 5 4 5 Motion data were collected for the ROPOS ROV system
zN⫹1 Z ⫹ s…2†
N ⫹ u …2†
N
(Fig. 3) to accurately determine the unknown model coeffi-
cients, to verify manufacturer’s specifications and to vali-
2 3 date the numerical model. ROPOS is a large ROV that dives
0
6 7 to 5000 m, and is attached to the cage by an umbilical tether.
6 z2 ⫺ s…1† ⫺ Z 7
6 2 7 The cage itself weighs 49 000 N in air and has dimensions
6 7
6 …1† 7 of 2.1 × 3.4 × 4.2 m (WLH) and is connected to the ship by
6 z3 ⫺ s3 ⫺ Z 7
6 7
uˆ6
6 .
7
7 …18† a triple armoured tether that is 0.03 m in diameter, 4200 m
6 .. 7 long and weighs 30.5 N m ⫺1 in air (Fig. 4).
6 7
6 7 Data were collected while ROPOS operated at depths of
6 …1† 7
6 zN ⫺ sN ⫺ Z 7 up to 1765 m. Sea conditions ranged from relatively calm to
4 5
…2† very rough with sheave displacements less than 1 m and
zN⫹1 ⫺ sN ⫺ Z
where s…1† …1†
1 and u1 are zero and R
m×n
denotes an m × n
matrix.

2.5. Numerical implementation

The N second-order differential equations given by Eqs.


(12) and (16) are rewritten as 2N first-order differential
equations, thus allowing the use of a conventional ODE
integration routine. This approach adds very little computa-
tional overhead because the additional N terms are simple
equality statements; however, previous time step informa-
tion is needed. A Runge–Kutta fourth/fifth-order numerical Fig. 4. A cross-section of the main umbilical tether used in the ROPOS
integration routine with adaptive step size is used. This ROV system.
F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182 175

Table 1 The wet mass, m; is 2% larger than the dry mass and its
Values of the model coefficients estimated using motion and tension numeric value is given in Table 1.
measurements of the ROPOS ROV system
The mass and drag coefficients of the cage were deter-
Coefficient Calculated value Manufacturer’s or mined by expressing Eq. (11) as:
empirical estimates
WCG ⫺ T…L† ˆ MCG …Z ⫹ u…2†
N †⫹
1
2 rw AC CD …Z_ ⫹ u_…2† _ _ …2†
N †jZ ⫹ u N j
⫺1 ⫺1
m 3.01 ^ 0.06 kg m 3.11 kg m
…19†
W 25.4 ^ 0.5 N m ⫺1 25.5 N m ⫺1
WCG 43 200 ^ 1000 N 42 000 N Since Eq. (19) is linear in MCG and CD ; a least-squares fit of
EA 45.5 ^ 1.4 × 10 6 N 33.4 × 10 6 N
tension in the tether to the velocity times its modulus and the
CDT 0.02 0.02
CVCG 8150 ^ 170 kg m ⫺1 7300 kg m ⫺1 acceleration of the cage yields values for the unknown coef-
MCG 7750 ^ 160 kg 9200 kg ficients. However,
ÿ  tension was not measured at the cage
termination,ÿ T L ; but it was assumed that it is well approxi-
greater the 4 m peak-to-peak, respectively. Two sites were mated by T 0 for ÿ L ⱕ 50ÿ m. For L ⱖ 50 m, the phase and
instrumented: the top of the ship’s A-frame (directly above magnitude of T L and T 0 are different [1] and therefore
the sheave that supports the tether and thus, at node 1), and records for these greater depths cannot be used.
the top of the ROV cage (node N ⫹ 1). Tri-axial, orthogonal Records from 28 and 32 m depth, measured on separate
accelerometers and rate gyros were used to measure all six days with different sea states, provide values and 95% con-
degrees of motion of the ship and cage. A pressure sensor fidence intervals of MCG ˆ 7750 ^ 160 kg and CVCG ˆ
was used to measure the depth of the cage and a load cell 1=2rw AC CD ˆ 8150 ^ 170 kg m⫺1 : The statistical error is
was used to measure the tension in the tether at the A-frame. less than 1% but the different values for the two depths
Data were sampled at 256 Hz to a resolution of 16 bits. A imply a bias and systematic error of 2% and this larger
detailed description of the instrumentation, data reduction value is used as the confidence interval. Agreement is within
and data analysis is given in Ref. [1]. For all measurements 10% for values of MCG and CVCG calculated from the data
used in the calibration and validation, the ROV was out of and values estimated using empirical tables (Table 1).
the cage, the tether connecting the ROV and cage was slack The axial stiffness, EA; is given by [4]:
and the ROV and cage were horizontally separated by at EA ˆ mc2 …20†
least 5 m. Thus, MROV and WROV are set to zero.
where c is the speed of strain waves along the tether. As will
be shown in Section 7 (see Fig. 9), echoes of a snap load are
spaced at constant intervals equal to the travel time of a
4. Calibration strain wave through the length of tether, and thus they
provide a very precise measure c for large L. A second
The ability of any model to accurately predict the
measure of c is provided by the higher order natural frequen-
response to specific input requires the numeric values of
cies, which are spaced at intervals of 0:5c=L Hz: Both meth-
the coefficients to be well known. However, values for
ods yield identical values of c ˆ 3870 m s⫺1 : Using this
most of the model’s coefficients are difficult to determine.
value in Eq. (20), the value of the axial stiffness was
The cage is a three-dimensional bluff body with two planes
found to be EA ˆ 45:5 ^ 1:4 × 106 N: The manufacturer’s
of symmetry and complex geometry. Its drag coefficient, CD
estimated value of EA is approximately 25% less (Table 1),
and added mass (included in MCG † cannot be accurately
indicating a large error in their calculations.
determined using empirical values of coefficients for bodies
of simple geometry. The linear stiffness, EA; of the tether
can be difficult to accurately determine using analytical 5. Validation
equations and empirical data because the tether consists of
an inner core of several different materials surrounded by The model was validated by directly comparing the real
layers of armour strands that are counter-helically wound. and simulated motion of the cage and tension in the tether
Thus, experimental data was used to determine accurate for the same forcing at the top node. This was done for
values of the coefficients. several depths, but for brevity this paper will focus on
For several different deployments, the weight less buoy- operation at 1730 m. Numerical results were calculated
ancy of the tether, W (per unit length), and cage, WCG ; were using 50 elements and the estimated coefficients (Table 1).
estimated from a least-squares linear fit of tension to depth The heave acceleration of node 1 was prescribed using a
for the range of 218 to 1730 m. Tether deployment was 5000 s time series of measured acceleration of the A-frame
halted to allow a steady length for each measurement. The (ship). Since data were collected during a one-month period,
slope W and offset WCG (Table 1) agreed well with the which included a large range of sea states and operating
manufacturer’s specifications. The mass of the tether is depths, different data were used for the coefficient estima-
approximated by its dry mass plus the mass of the water tion and model validation. For example, MCG and CD were
entrained into the void space between the strands of armour. estimated from two data sets collected two weeks before,
176 F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182

Fig. 5. Short time series of calculated (thick line) and measured (thin line) cage motion and tension in the tether for the system at 1730 m. The ship motion
(dotted line) is included for comparison.

and at an operating depth 1700 m shallower than the data which is well approximated by an odd Talyor series. Thus,
used for validation in this report. variance is produced at odd harmonics of the fundamental
The model provides a very good approximation of the frequency of excitation, 0.15 Hz (the peak of the spectrum
actual displacement of the cage as suggested by Fig. 5. in Fig. 6). The third and fifth harmonics are evident at 0.44
The ship displacement was included in Fig. 5a to highlight and 0.73 Hz (Fig. 6). The model over predicts the variance
the difference between input and response. Indeed, the at those frequencies; however, the system remains weakly
model correctly predicts that the displacement of the cage non-linear [1] because the harmonics only contributed a
lags behind and is greater than that of the ship. This parti- small fraction to the total variance. For example, at
cular section of data has a large range of amplitudes and 1730 m, the variance at the harmonics provided only 4.6
includes a snap load which starts at t ˆ 627 s: The larger and 0.3% of the total variance of the cage acceleration.
amplitudes are underestimated by the model and the snap The transfer function estimate (TFE) provides a measure
behaviour of the system, that is the sudden change of accel- of the motion amplification and phase difference between
eration at the cage, was not reproduced. Quantitatively, the the ship and cage and it is estimated using:
standard deviation of the difference between the simulated CZ S Z C
andC measured position of the cage (tension at the Cship) s is HZ S Z C ˆ …21†
Ts CZ S Z S
s diff
Z
ˆ 0:11 m …s diff ˆ 4000 N† and the ratio of s diff
Z
…s diff
T
†
to the rms of the measured values is 0.12 (0.28). where CZ S Z C is the cross-spectrum of ship-displacement, Z S ;
The spectra of the real and simulated vertical cage accel- and cage-displacement, Z C ; and CZ S Z S is the auto-spectrum
eration are compared for the same 5000 s record of ship of Z S : The analysis of the TFE is restricted to the waveband
motion in Fig. 6. The model predictions that include the (0.1–0.25 Hz) because the error in the real TFE is large and
wake effects are addressed later in the Section 7. The the modelled cage response is not significant at higher
frequency domain analyses focuses upon the waveband frequencies. At 1730 m, agreement between the real and
that extends from 0.1 to 0.25 Hz, because this band contains simulated TFE is good up to 0.2 Hz and the largest
90% of the variance of ship motion. The measured and difference in magnitude is only 8% at 0.13 Hz (Fig.
simulated acceleration spectra agree well within the wave- 7). The model over predicts the cage response at the
band at 1730 m. The largest difference is 13% at 0.14 Hz higher frequencies within the wavebound. The simulated
and the spectra are equal at 0.175 Hz. For L ˆ 975 m (not phase is always slightly larger than the real phase and
shown) the real and simulated spectra are almost identical the maximum difference is 15⬚ at 0.2 Hz (near the
and the largest difference is only 7% at 0.14 Hz. Above middle of the waveband). These discrepancies are attrib-
1 Hz, significant excitation is only produced by the impul- uted, in part, to the assumption of a constant drag coef-
sive snap loads and no wave excitation is transferred to the ficient and a constant added mass of the cage which is not
tether because the large inertia and length of the ship act as a completely accurate for oscillatory flow around a bluff
low-pass filter for short (high-frequency) waves. body. This issue is addressed in Section 7. At 975 m, the
The cage is a bluff body and is subject to quadratic drag magnitude and phase of the TFE (not shown here) agree
F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182 177

Fig. 6. Spectra of actual (thin solid line) and model calculated (dotted line—without wake, thick solid line—with wake) vertical acceleration records of the
cage at a depth of 1730 m. The arrows are the 3rd and 5th harmonic of the peak cage motion and the natural frequency, f1. Both axes have logarithmic scales.

exceptionally well. The largest TFE differences are 0.05 accurately known and thus, differences will exist between
(4%) at 0.19 Hz and 10⬚ at 0.25 Hz. estimated values and actual values as shown in Table 1.
The depth dependent natural frequencies, fn ; provide Therefore, the model output must be reasonably insensitive
another characteristic to validate the model. The predicted to variations of its coefficients to be a useful design tool. If
and observed first natural frequency, indicated by the first not, slight errors in the values of the estimated coefficients
90⬚ phase shift of HZ S Z C shown in Fig. 7, agree within 7%. could produce erroneous predictions. Nine coefficients are
Although very little excitation occurs above 1 Hz, the model used in the model: N, L, MCG, m, CVCG ; CDT, WCG, W, and EA.
exactly predicts the second, third, and fourth natural frequen- The tether drag can be ignored because the ratio of tether-to-
cies at 1.17, 2.24 and 3.36 Hz. This is shown in Fig. 6. cage drag force is 0.06. Gravitational and buoyancy forces
only affect the static solution, and the dynamic response of
6. Sensitivity the system in the waveband is independent of the number of
elements used when N ⬎ 5 (the effects of N on the predicted
During the design process, the model coefficients are not response will be addressed later in Section 7). Thus, only

Fig. 7. Magnitude (a) and phase (b) of the actual (stars) and calculated (dotted line—without wake, solid line—with wake) transfer function estimate between
the ship and cage motion at a depth of 1730 m.
178 F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182

Fig. 8. Sensitivities of the magnitude (a) and the phase (b) of the transfer function between ship and cage motion to changes in MCG (thin solid line), CVCG (dashed
line), and EA (thick solid line) for operating depth of 1730 m. The dotted line is the difference between the actual and simulated TFE.

five coefficients affect the dynamic response of the model: L, phase of the transfer function can be obtained using
MCG, m, CVCG ; and EA. In previous work [4], it was shown “rough” estimates of EA, MCG, and CVCG : As in the contin-
that the non-dimensional transfer function between ship and uous model [4], the sensitivities of the phase to MCG are zero
cage motion is only a function of two non-dimensional para- at 0.25 Hz and the phase is independent of MCG; and
meters, namely: between 0.22 and 0.25 Hz the magnitude of HZ S Z C is only
slightly sensitive to variations in CVCG : Also notable is that
CVCG MCG
; the signs of the sensitivities are preserved between the
mc mL continuous model and present finite-element model used
when the frequency is non-dimensionalised with: in this study.
r The bounds of uncertainty for the estimated values of
0 m L
v ˆL vˆ v MCG, CVCG and EA are less than 5% and thus all their prob-
EA c able values are contained within the ^15% sensitivities. At
where v 0 is the non-dimensional frequency and v is the most frequencies in the waveband, the difference between
dimensional frequency in rad s ⫺1. Therefore, there are the real and calculated TFE (the dotted line in Fig. 8) is
only three independent parameters that characterise the much larger than the ^15% sensitivities; therefore, the
dimensional response and only the dependence of the model results cannot be significantly improved with realistic
model on CVCG ; MCG, and EA need be examined. This changes in MCG, CVCG and EA. Additionally, because the
is done by calculating the relative changes in the sensitivities change sign (either in magnitude and phase)
magnitude and the absolute changes of the phase of at different frequencies (not coincident with the error) one
the transfer function between vertical motions of the cannot expect that changes in MCG, CVCG and EA will
ship and cage for variations of CVCG ; MCG, and EA, of improve the predictions at all frequencies within the
^15% about their nominal values (Fig. 8). The sensi- waveband.
tivities of the TFE to EA look similar to those of MCG
but the magnitude and phase do not scale. The dotted
line is the difference between the real and calculated TFE 7. Discussion
(using the nominal coefficient values) and it is addressed in
Section 7. Replacing the consistent-mass matrix with the lumped-
In the mid to upper wave band …0:15 ⱕ f ⱕ 0:25 Hz†; the mass matrix to uncouple the acceleration terms in the
changes in the TFE are, at most, of the same order as the elemental equations eliminated the inversion of a N × N ⫺
changes in the coefficients and the magnitude is most sensi- 1 matrix and 6N 2 operations. This simplification signifi-
tive to CVCG while the phase is most sensitive to EA. In the cantly reduced computation time, especially during periods
lower waveband, … f ⬍ 0:15 Hz† the magnitude and phase of of rapid system change and very small time stepping. In
HZ S Z C are not sensitive to EA, MCG, and CVCG : Therefore, particular, simulation times were reduced from hours to
reasonably accurate predictions of the magnitude and minutes for a 5000 s run. For comparison with the lumped
F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182 179

Fig. 9. Measured (thin line) and calculated (thick line) records of the tension in the tether at the A-frame (a) and acceleration of the cage (b) during a snap load.

mass model, the motion of the cage and tension at the ship required. An N-element model can only predict up to N
are calculated using the consistent-mass representation with natural frequencies and the accuracy of the predicted
50 elements and the same 5000 s record of ship motion. value increases by increasing N.
Results from the two representations are identical. Notice- The hydrodynamic forces on the cage have been charac-
able differences only occur when a small number of terised by a drag coefficient multiplied by the velocity of the
elements (⬍5) is used. For example, predictions of the cage. Although this formulation is common in practice, it is
second harmonic of the natural frequency differ by 0.26 based upon constant speed flow over a body. For steady
and 0.06 Hz for a two-element and a three-element model, external (far from the body) flows, the drag coefficient
respectively. changes with changing speed because of variations in the
The run time of the simulation increases quadratically point of boundary layer separation and the pattern of recir-
with the number of elements and, therefore, it is advanta- culation over the body and in its wake. The boundary layer
geous to use as few elements as possible. In the model, the and wake of an accelerating body that has reached a given
mass of each element is “lumped” at its nodes and, as a speed are never the same as those over the same but non-
result, the mass distribution and (therefore) the accuracy accelerating body at that speed. In addition, the boundary
of the static results increase with increasing number of layer during acceleration is not the same as during decelera-
elements. However, the dynamic motion of the cage and tion (for the same speed) [19].
tension at the ship were found to be nearly independent of The inertial forces on the cage have been represented by
the number of elements for predictions within the wave- its acceleration multiplied by its effective mass. This, too, is
band. Thus, the practical considerations for the lower common practice but this formulation obscures the physics
bound of N are as follows. The prediction of zero tension of unsteady flow around bluff bodies and only makes sense
(snap loads) depends on the total (static ⫹ dynamic) tension if the relative rate of change of velocity is small and the drag
in the tether and, therefore, a fine element break-up is coefficient is not sensitive to small changes in Reynolds
180 F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182

number. The notion of added mass becomes simplistic when (15⬚) to 0.07 (7⬚), respectively. Surprisingly, addition of the
the relative change of velocity is large and when the velocity wake model reduces the non-linearity of the predicted
changes direction. For example, the added mass can response and brings the predicted spectrum of acceleration
momentarily become negative when vortices detach from at the third and fifth harmonic of the fundamental frequency
the body during acceleration and very large when they reat- of excitation into close agreement with the measured spec-
tach during deceleration. trum. Similar improvements were found for an operation
The flow history of water near the cage has been ignored. depth of L ˆ 975 m.
Although this is also common in practice, it is based upon a An interesting phenomenon can be seen in the magnitude
body travelling in a constant direction. For an oscillating of the transfer function between ship and cage motion (Fig.
body, the effects of a disturbed flow become important when 7). The magnitudes of the TFE calculated for the models
significant momentum is transferred to the water and the with and without wake agree at 0.17 Hz. At lower (higher)
period of oscillation is less than the decay time of the frequencies within the waveband, the magnitude of the TFE
disturbed flow (wake). The heave motion of the cage is predicted by the model with wake is larger (smaller) than
unsteady, it oscillates between positive and negative values that for the model without wake. This phenomenon is easily
and the flow around the cage has been disturbed by previous explained by the momentum exchange between the cage
oscillations. Thus, one cannot expect hydrodynamic forces and wake. When the cage slows from its maximum velocity,
on the cage to be represented, with complete accuracy, by a the wake overtakes it and transfers momentum to the cage.
constant drag coefficient, a constant added mass and the When the cage changes direction, it must pass back through
exclusion of flow history. the wake and momentum is transferred from the cage to the
The sensitivity analysis indicates that differences between wake. At frequencies lower than 0.17 Hz, much of wake has
measured and modeled motions cannot be explained by decayed by the time the cage reverses direction and thus,
parameter uncertainty. The differences may be due to the there is a net momentum transferred to the cage. At higher
limitations of the concept of a constant drag coefficient, a frequencies, the slowing down and reversal is more rapid
constant added mass and excluding flow history. To test this and the wake is stronger as the cage passes back through. As
notion, a simple wake model was added that takes some a result, there is a net momentum transfer to the wake.
account of the history of motion. Only directional dependent Motion during a snap load is non-linear and different
drag coefficients and added mass values were used because from that during normal operation. Large and rapid down-
estimates as functions of velocity and acceleration could not ward displacements of the ship produce slack tether at the
be calculated from the data. The objective is to motivate cage. A snap load results when the tether is re-tensioned
additional work on the unsteady forces on bluff bodies because the velocities of the tether and cage are different.
attached to tethers rather than to promote this particular The induced strain wave travels along the tether with speed
model. It is assumed that, at each time step, the water c and it is reflected at its ends (the ship and cage). This
directly behind the cage (on its lee side) has the same velo- results in a series of rapid changes of cage acceleration
city as the cage. The position of these water parcels is (jerks) and tension at the ship and these changes are spaced
tracked and it is assumed that their momentum decays expo- at intervals equal to the round trip travel time of strain waves
nentially with time. When a water parcel overtakes the cage [1].
(typically when the cage decelerates) an additional force is Snap loads occurred in less than 1% of the motion cycles
added to the lea side that is proportional to the velocity in the 5000 s record that was used to validate the model.
difference between the cage and this parcel. The decay Such threshold conditions are very difficult to predict
time for momentum is 2.9 s and was chosen by trial and because they require the numerical model to be very precise.
error. The corresponding length scale is 2.9 m for typical Without the wake model, the simulation underestimated the
speeds and is comparable to both the height of the cage and large displacement and did not reproduce the slack tether—
the maximum amplitude of motion. the minimum tension was 10 000 N. With the inclusion of
Including the wake model significantly improves the the wake model, the calculated motion only slightly under-
results of the model for operation at 1730 m. The standard estimated the large vertical displacement and the tension in
deviation of the difference between the calculated and the tether at the cage approached zero. However, zero
measured position of the cage (tension in the tether) tension was still not achieved and snap loads were not
is C improved by S a factor of 1.6 (1.8) and it is predicted. Therefore the ship motion was increased by
s diff
Z
ˆ 0.07 m …s diff
T
ˆ 2220 N† vs 0.11 mC (4000 N) for the 10% so that slack tether would result and the response of
TS
model without wake. The new ratio of s diff Z
…s diff † to the rms the model to this condition could be investigated. Under
of the measured values is 0.08 (0.16). The calculated and these conditions, the model reproduced the observed snap
real spectrum and transfer function (both magnitude and loads and the associated acceleration and tension profiles
phase) between ship and cage motion are nearly identical, (Fig. 9). The rapid changes of acceleration and tension
and the agreement in the upper waveband is notably aligned with the measured values and agreed in magnitude.
improved as seen in Figs. 6 and 7, respectively. The error Most notable is that the slope of the real and calculated
in the magnitude (phase) of the TFE is reduced from 0.12 acceleration is nearly identical at the first jerk. It is therefore
F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182 181

concluded that the discrete model inherently has the ability motion by 10%, the model predicted the observed snap
to produce characteristics associated with snap loads. A loads and accurately reproduced their characteristics.
more realistic representation of the drag, added mass and These included rapid increases in the acceleration/tension
wake could improve the prediction of slack tether so that record that were spaced at regular intervals, equal to the
threshold conditions of snap loading could be better travel time of the strain wave along the length of deployed
predicted without artificially increasing the ship motion. tether.

8. Conclusions Acknowledgements

A one-dimensional finite-element lumped-mass model of We would like to thank Jim McFarlane and the staff of
a caged remotely operated vehicle system using a coupled International Submarine Engineering (ISE) for their contin-
set of reference and elastic co-ordinates was developed. The ual support and assistance in this project. We would also
elastic displacements were approximated with a linear trial like to thank Inna Sharf for sharing her expertise in finite-
solution. Using Galerkin’s method, the equation for the element modelling and Nedjib Djilali for sharing his exper-
balance of forces at any point in the tether was discretised. tise in fluid dynamics. F.R. Driscoll was financially
The resulting elemental equations were assembled to obtain supported by scholarships from the Natural Science and
the finite-element equations of motion for a deep-sea ROV Engineering Research Council (NSERC), the Science Coun-
system. The resulting model is a function of only nine coef- cil of British Columbia (SCBC) and ISE. This work was
ficients: N, L, MCG, m, CVCG CDT, WCG, W, and EA, when partially supported by the US Office of Naval Research
tether has constant properties along its length. under contract N00014-93-1-0362 and NSERC.
Values of the model coefficients were estimated using
motion and tension measurements of the ROPOS ROV
system. Although many of the coefficients are readily esti- References
mated, it is very difficult to determine accurate values of the
mass, MCG, and drag coefficient, CVCG ; of the cage and the [1] Driscoll FR, Lueck RG, Nahon M. The motion of a deep-sea remotely
stiffness of the tether, EA. A least-squares method was used operated vehicle system. Part 1: motion observations. Ocean Engi-
to determine values for MCG and CVCG to an accuracy of 2% of neering 2000;27:29–56.
[2] Hover FS, Grosenbaugh MA, Triantafyllou MS. Calculation of
their real values. The stiffness, EA, was determined to 3% dynamic motions and tensions in towed underwater cables. IEEE
accuracy using the speed of tensile waves in the tether, c, Journal of Oceanic Engineering 1994;19:449–57.
which was estimated from the time of successive tension [3] Grosenbaugh MA. On the dynamics of oceanographic surface moor-
maxima during a snap load. These methods are useful because ings. Ocean Engineering 1996;32:7–25.
they only used data acquired during normal operation of the [4] Driscoll FR, Lueck RG, Nahon M. The motion of a deep-sea remotely
operated vehicle system. Part 2: analytic model. Ocean Engineering
ROV system and did not require any special procedures. 2000;27:57–76.
The vertical motion of the top node was prescribed based [5] Lueck RG, Driscoll FR, Nahon N. A wavelet for predicting the time-
upon measurements of ship motion. These measurements domain response of vertically tethered systems. Ocean Engineering
were used to validate the model by comparing the output 2000 (in press).
against the corresponding measurements of cage motion and [6] Caughey TK. Equivalent linearization techniques. Journal of the
Acoustic Society of America 1963;35:1706–11.
tension at the ship. The model presented herein accurately [7] Niedzwecki JM, Thampi SK. Snap loading of marine cable systems.
predicts (a) the motion of the cage and the tension in the Applied Ocean Research 1991;13:210–8.
tether; (b) the spectrum of cage acceleration in the wave- [8] Yoerger DR, Grosenbaugh MA, Triantafyllou MS, Burgess JJ. Drag
band; (c) the transfer function between the vertical motion forces and flow-induced vibrations of a long vertical tow cable—Part
of the ship and cage; and (d) the natural frequency and 1: steady-state towing conditions. Journal of Offshore Mechanics and
Arctic Engineering 1991;113:117–27.
harmonics of the system. Since data used in the coefficient [9] Hover FS, Yoerger DR. Identification of low-order dynamics models
estimation and model validation are different, the model for deeply towed underwater vehicle systems. International Journal of
should be accurate for a larger range of ship motions than Offshore and Polar Engineering 1992;2:38–45.
presented in the validation. [10] Ablow CM, Schechter S. Numerical simulation of undersea cable
Sensitivities to the three independent parameters of the dynamics. Ocean Engineering 1983;10:443–57.
[11] Newman JN. Marine hydrodynamics. Cambridge MA: MIT Press,
model indicate that the mismatch between the simulated and 1989.
real motion cannot be significantly improved by adjusting [12] Sparks CP. The influence of tension, pressure and weight on pipe and
the values of those parameters. However, inclusion of a riser deformations and stresses. Transactions of the ASME
simple wake model improved the agreement between the 1984;106:46–54.
real and calculated results by almost a factor of 2. Eight [13] Folb R, Nelligan JJ. Hydrodynamic loading on armoured towcables.
DTNSRDC report 82/116, 1983.
snap loads also occurred in the measurements used to vali- [14] Shabana AA. Dynamics of multibody systems. New York: Wiley, 1989.
date the model; however, these were only threshold occur- [15] Burnett DS. Finite element analysis from concepts to applications.
rences. Using the wake model and increasing the ship Don Mills, Ontario: Addison-Wesley, 1987.
182 F.R. Driscoll et al. / Applied Ocean Research 22 (2000) 169–182

[16] Finlayson BA. The method of weighted residuals and variational ical recipes in C. 2nd ed. Cambridge: Cambridge University
principles. New York: Academic Press, 1972. Press, 1992.
[17] Logan DL. A first course in the finite element method. Boston, MA: [19] Sarpkaya T, Isaacson M. Mechanics of wave forces on offshore struc-
PWS-KENT, 1992. tures. New York: Van Nostrand Reinhold, 1981.
[18] Press WH, Teukolsky SA, Vetterling WT, Flannery BP. Numer- [20] Robigou V. Journal of Geophysical Research Letters 1993;20:1.

You might also like