Sunteți pe pagina 1din 7

Hydrodynamic Heave Damping

Estimation and Scaling for Tension


Leg Platforms
K. P. Thiagarajan Resonant heave excitation of tension leg platform (TLP) tendons is typically of
Research Fellow, high frequency and small amplitude. The response of the tendons to this excitation
Marine Hydrodynamics Laboratory. is non-negligible due to a very small drag coefficient of the structure in this mode
of oscillation. Small values of the drag force complicate experimental estimation in
a laboratory due to the dominating inertial force. Model tests conducted at the
A. W. Troesch University of Michigan investigating the damping experienced by a cylinder of 0.457
Associate Professor. m (1.5 ft) diameter and 1.219 m (4.0 ft) draft are described here. The cylinder is
vertical and surface-piercing, and oscillates parallel to its axis. The amplitude of the
Department of Naval Achitecture and Marine forcing is varied to give a Keulegan-Carpenter (KC) number range of 0.1-1.0. The
Engineering, frequency parameter /3 is 89236, corresponding to an oscillation frequency of 0.41
The University of Michigan Hz. From these experiments, a definite nonlinear trend is observed between the drag
Ann Arbor, Ml 48109-2145 force and velocity conflicting with some of the results reported by Huse (1990) and
Chakrabarti and Hanna (1991). The heave damping coefficients of individual struc-
tural components of a TLP follow different scaling laws. Rules are presented for
scaling friction and form drag components from model to full scale. Results from
experiments are used to obtain a scaling law for vertical columns of a TLP. Previously
published results are used for horizontal pontoons. An example TLP calculation
shows that the heave damping ratio of horizontal cylinders is approximately 0.049-
0.078 percent, depending upon cylinder shape, and that for vertical cylinders is in
the range 0.025-0.171 percent, depending upon KC.

1 Introduction
Offshore structures like the tension leg platform (TLP) ex- shear stresses and normal stresses or pressures, each of which
perience small-amplitude, high-frequency oscillations in the may follow different scaling laws. For this reason, model test-
vertical mode—the so-called springing vibration (Demirbilek, ing of a complete geosim in a model test basin will not provide
1989). The high-frequency component of the tendon tension useful answers that can be extrapolated to full scale (Couch
is influenced by the heave, pitch, and roll motions of the TLP et al. 1984). Since it is not possible to satisfy all the relevant
hull. Generally, these three hull motions are coupled, and their scaling laws in one test, various common structural shapes are
natural frequencies are close together, A simplified analysis tested separately; see Fig. 1. Based upon engineering necessity,
is performed by considering the heave motion only, see e.g., the most significant quantities are then measured and scaled
Huse (1990), and Chakrabarti and Hanna (1991). The total properly. It is assumed that the total system hydrodynamic
damping of the system in heave is typically small in magnitude. damping can be expressed as the net sum of the individual
Nevertheless, knowing the damping is crucial in determining members, each evaluated separately from the other structural
the resonant response of the structure. Experimental investi- parts. This is clearly an approximation, for all three of the
gation of the damping coefficient is difficult, however, due to foregoing damping components are interrelated and, in gen-
the small magnitude of the damping forces. eral, not independent. However, these interaction effects are
The various components of hydrodynamic damping for bod- expected to be small in magnitude relative to the sum of the
ies such as TLPs are generally due to: 1) time-depending damping contributions from each member treated independ-
boundary layer flows, 2) separation, vortex formation and ently.
shedding, and 3) free surface effects such as wave generation The two main governing parameters for small amplitude
or wave diffraction. These components are influenced by var- oscillatory flow or rigid body vibration of vertical and hori-
ious factors including wall roughness, incident currents, back- zontal cylinders (Figs. 1(b) and (c)) are the Keulegan-Carpenter
ground turbulence, and by the effects of hull motions in other number (i.e., KC) and the frequency parameter (i.e., (3). These
modes. The resultant damping forces are integrations of both are defined as
Contributed by the OMAE Division for publication in the JOURNAL OF OFF- KC = f
SHORE MECHANICS AND ARCTIC ENGINEERING. Manuscript received by the OMAE
Division, December 22, 1993; revised manuscript received January 24, 1994. Iff
Associate Technical Editor: A. N. Williams. (1)

70/Vol. 116, MAY 1994 Transactions of the ASME

Copyright © 1994 by ASME


Downloaded From: http://offshoremechanics.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jmoeex/28093/ on 07/17/2017 Terms of Use: http://www.asme.org
trend of the laminar, attached flow theory. A least squares
(a)
linear fit of the data on log-log plots has the same slope as
the theoretical laminar line, but a higher offset value. This
behavior is consistent with observations made by Sarpkaya
(1986) and Bearman (1992) on the drag forces of circular cyl-
inder at increasing /3 values.
After flow separation occurs, a semi-empirical formula for
the drag force such as Morison's equation is used. This equa-
tion models the drag force as a quadratic function of velocity.
Marthinsen (1989) has combined Morison's drag formulation
and the laminar boundary layer theory to produce a revised
drag coefficient valid at both low and moderate KC numbers.
(b) (c) The inherent assumption in the formation of Morison's equa-
tion is that the flow be two-dimensional. Hence, the applic-
ability of this equation for three-dimensional flows is not
+ C D obvious. If one were to assume that a three-dimensional os-
cillatory flow possess similar gross characteristics as its two-
dimensional counterpart (for example, significant vortex shed-
ding off sharp corners), then one could expect the form of
Fig. 1 Component problems of a TLP
Morison's equation to be valid for a body in a three-dimen-
sional flow.
For cylinders oscillating parallel to their axes (i.e., cylinders
Symbols are defined in the Nomenclature. It can be seen that in heave, see Fig. 1(b)), KC and 13 in Eq. (1) can be defined
the product of @ and KC yields a Reynolds number based on based on the cylinder diameter (£>„). Huse (1990) and Chak-
the oscillation velocity (CUA:). rabarti and Hanna (1990, 1991) have conducted model tests
For oscillating horizontal cylinders (Fig. 1(c)), the charac- on axially oscillating vertical cylinders. Huse (1990) reports
teristic length, D, is the diameter Dh. Anaturk (1991) and that in a KC range of 0.0005-0.01, and /3 approximately 5 x
Anaturk et al. (1992) present regions in the KC-/3 plane where 106, the drag force varies linearly with velocity. He supports
two-dimensional flows possess different characteristics. For this trend with a theory based on wake and momentum con-
K C < 1 and /3 in the range of 103-104, the boundary layer is siderations. Chakrabarti and Hanna (1990, 1991) also have a
lammar and the flow remains attached to the body. Kim and similar observation, in a/3 range of 25 x 1 0 4 - 1 x 106. These
Troesch (1989) have observed that the flow remains attached authors measured the damping ratio at model scale to be in
at very low KC, even when the body has sharp corners such the range of 0.3-1 percent. Petrauskas and Liu (1987) have
as a square section perpendicular to the flow. For offshore measured the springing force on a vertical cylinder and on a
applications, /3 is in the range of 106 (Anaturk, 1991). At these TLP model due to incident waves. They estimate the damping
/3 values, when K C < 0 . 1 , the flow is attached and may be ratio of the model to be in the range 0.11-1.6 percent. Scaled
laminar. As KC increases, the turbulence in the boundary layer as the inverse square root of Reynolds number, the authors
grows and the flow separates. The sequence of events depends expect the full-scale damping ratio to be at most 0.2 percent.
upon the particular value of /3. The reader is referred to Honji To measure the hydrodynamic damping forces, model test-
(1981), Hall (1984), and Sarpkaya (1986) for futher details on ing can be conducted in a laboratory by oscillating a body in
transition and boundary layer stability. For three-dimensional a still fluid or by oscillating the flow past a stationary body.
flows associated with three-dimensional bodies, the hydro- Experiments performed by oscillating a body in a still fluid
dynamics may be expected to have similar characteristics, are usually one of the two types:
through the values of parameters where flow transition occurs
may vary.
As long as a two-dimensional oscillatory flow does not sep- 1 The body oscillates freely, and damping is estimated from
arate from the body, the flow can be satisfactorily described decrement curves (e.g., Chakrabarti and Hanna, 1990, 1991;
by a laminar attached boundary layer theory, (Riley, 1965; and Bearman, 1992).
Stuart, 1966; Wang, 1968; Batchelor, 1973; Kim and Troesch, 2 The body is forced to oscillate at a certain frequency,
1989). This theory predicts the drag resulting from the com- and drag coefficient is estimated from some type of Fourier
bined normal and tangential stresses to be a linear function of analysis (e.g., Troesch and Kim, 1991).
velocity. Troesch and Kim (1991) show that even for sharp-
edged cylinders at very small KC, drag coefficients follow the For lightly damped systems, both schemes have difficulties of

Nomenclature
Bh = equivalent linear damping coef- Dv = diameter of vertical cylinder ai = offset of f„-KC curve
ficient for horizontal cylinder Dh = diameter of horizontal cylinder «2 = slope of f„-KC curve
Bv = equivalent linear damping coef- / = frequency of oscillation (Hz) P = dimensionless frequency pa-
ficient for vertical cylinder KC = Keulegen-Carpenter no., 2-KX/ rameter, D2f/v
Cdh = drag coefficient of Morison's D ^ = dynamic viscosity
equation for horizontal cylin- L = length of horizontal cylinder P = water density
der m = mass of cylinder V = kinematic viscosity
Cdv = drag coefficient of Morison's S = frontal area of vertical cylin- CO = circular frequency
equation for vertical cylinder der, 7r£>„2/4 r„ = damping
damping ratio (ratio of system
Cm = inertia coefficient of Morison's T = draft of vertical cylinder to critical damping)
equation u = flow velocity or velocity of of vertical cylinder
D = characteristic body length, typ- body relative to still water f» = damping ratio of horizontal
ically cylinder diameter x - amplitude of oscillation cylinder

Journal of Offshore Mechanics and Arctic Engineering MAY 1994, Vol. 116/71

Downloaded From: http://offshoremechanics.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jmoeex/28093/ on 07/17/2017 Terms of Use: http://www.asme.org


Table 1 Parameters of model testing 20

Diameter of cylinder
Draft of cylinder
0.457 m (1.5 ft)
1.219 m (4.0 ft) - X
Surface area of water tank 100.5 m x 6.7 m (330 ft x 22 ft)
15 // A
Water depth 3.0 m (10 ft)
Frequency of oscillation 0.41 Hz
Amplitudes of oscillation
KC range
0.25-6.35 cm (011-2.5 in)
0.1-1.0
fi
^c?

X
0 89236 </3
10
: y A experiments
their own. In decrement tests, extreme care must be taken to m
ensure that the system damping is properly accounted for. See Y
— friction drag
Chakrabarti and Hanna (1990, 1991) and Bearman (1992). In •A
the case of forced oscillations, straightforward estimation of
damping is extremely unreliable, since the total force is inertia
dominated. Troesch and Kim (1991) conducted damping meas- o— —T—r—T—r- ——
,—
,—
,—
, — , — , — j — , — , —r—i—i—r— i i—i—r—

urements on cylinders in forced oscillation at or near reso- 0.2 0.4 0.6 0.8 1
nance. In this case, the restoring forces nearly offset the system
inertia forces. The resulting force measurements give a good KC
Fig. 2 Damping coefficient (B„) for a vertical cylinder versus KC; (1
estimate of the drag coefficients. Different springs were used 89236
to obtain resonances at several frequencies of interest. This
idea of forced oscillation at resonance is used in the present
work. 10

2 Description of Experiments
The experiments reported here were conducted at the Marine
Hydrodynamics Laboratory, University of Michigan. A smooth
surface-piercing vertical cylinder was forced to oscillate parallel A experiments
to its axis. Table 1 gives the dimensions of the test cylinder, — friction drag
and the parameter ranges tested. The frequency parameter, /?,
is based upon the cylinder diameter. The cylinder was attached
to a Vertical Motion Mechanism (VMM) and oscillated in
heave. The VMM produces a uniform vertical motion and
negligible motion in the other modes of freedom. The oscil-
lation frequency of 0.41 Hz was so chosen to correspond closely
to the natural frequency of the system in heave. At this resonant
I
condition, the inertia effect of the cylinder and attached dy-
namometers is nearly offset by the restoring force associated
with the fluctuating buoyancy force. As mentioned before, due
to the lack of inertia dominance, the force measurements give 0.2 0.4 0.6 0.8 1
a reliable estimate of the drag coefficient. The repeatability of KC
the data was verified during the experiments. Fig. 3 Drag coefficient (Crf„) for a horizontal cylinder versus KC; 0
For typical forced oscillation measurements, forces are 89236
measured using four load cells, and cylinder displacements are
measured using two potentiometers. At each run, data over
16 cycles is collected and sampled at a rate of 128 points/cycle. Alternatively, for sinusoidal motions the drag forces can be
This data is then Fourier transformed to identify the magni- interpreted in a Fourier-averaged sense as
tudes and phases of the forces and displacements. The drag
force is identified as the out-of-phase component of the total Fd(t) = Bvu{t) (3)
force at the oscillating frequency. The equivalent linear damping coefficient B„ can be related to
Cdv in a straightforward manner (Sarpkaya and Isaacson,
3 Results and Discussion 1981). For a vertical cylinder, this relationship is found to be
Two options are present for normalizing the force coefficient
for an axially oscillating vertical cylinder. One option would Bv = -^DvKCCdv (4)
use the vertical wetted surface area (i.e., -wDvT), while the other
would use the cross sectional area (irDv2/4). Using the cross- Some of the results from the experiments are shown in Figs.
sectional area would give a formulation similar to Morison's 2 and 3. Figure 2 is a plot of the damping coefficient Bv (in
equation, and has been used, for example, by Graham (1980) units of N-s/m) versus KC. From this figure, a definite linear
in relation to vortex shedding due to an isolated edge. For a trend is observed. A linear curve fit to the data gives a slope
vertical cylinder, vortex shedding off the sharp corner at the of 15.627 N-s/m, an offset of 3.424 N-s/m, and a correlation
bottom is expected to be the dominant part of the flow. Hence, coefficient of R2 = 0.995. In Fig. 3, a plot of Cd0 versus KC is
the latter normalization option is adopted. KC and jS are de- derived using Eq. (4).
fined based on the cylinder diameter (Z)„) for the same reason. Drag force component analysis commonly used in model-
Following Morison's equation, the vertical drag force is full-scale ship drag analysis is described in Couch et al. (1984).
modeled as Using a similar concept, the drag force on the cylinder is
decomposed into its friction and form drag components. The
Fdit) = -pSCdou(t) \u(t)l friction drag is mainly due to the viscous tangential stresses
(2)
acting along the walls of the cylinder. Vortex formation and

72 / V o l . 116, MAY 1994 Transactions of the ASME

Downloaded From: http://offshoremechanics.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jmoeex/28093/ on 07/17/2017 Terms of Use: http://www.asme.org


shedding at the b o t t o m edge of the cylinder are the main sources Fd = Fd(Dv,T,u>,x,v,p) (6)
of form drag. Define an equivalent linear damping ratio as
The magnitude and trend of the friction drag force expe-
rienced by the oscillating cylinder can be estimated approxi- B„ Bv
mately. Velocity profiles in the laminar b o u n d a r y layers of r„= Cc 2/wco
oscillating infinite flat plates are well k n o w n in the literature; where Cc is the critical damping of the system. Inserting an
see, for example, Batchelor (1973) or Telionis (1981). T h e shear expression for the mass, the damping ratio becomes
stress evaluated from this velocity variation can be integrated
over the cylinder surface to obtain an expression for the friction 1 (B,
drag. Chew and Liu (1989) note that when the radius of cylinder L-- •K C„, 0- (7)
to boundary layer thickness ratio
For a given geometry, f„ is directly proportional t o B„ at a
given 15. Since f„ is dimensionless, it can be used to rewrite Eq.
(6) in a nondimensional form

the transverse curvature of the cylinder can be neglected, i.e., t»=C^f(^,KC,0) (8)
the flow is locally two-dimensional. In the present case this
ratio is estimated to be 374. For the model cylinder, from Fig. 2 this relationship is iden-
For the case of a vertical cylinder in heave, the friction drag tified to be linear for fixed geometry and /3, i.e.,
components are given by the following expressions, using the iv=Cm\a, + a2KC) (9)
velocity profiles given by Batchelor (1973) or Telionis (1981):
Equation (9) can be used for estimating the damping ratio of
^ ( f r i c t i o n ) = 7T1'57>|80 the prototype, if at and a-i can be scaled appropriately. Assume
the following:

1 1 a i is scaled based on Eq. (5).


Q„(friction) = 6ir ^
KC7 ^
4TT/3
(5)
2 Cd„ (vortex shedding) is approximately a constant. Gra-
ham (1980) and Bearman et al. (1985) obtain this result for a
It can be seen from the equation for Cdv that the combination two-dimensional section whose included angle is 90 deg. This
KC (j30,5) is equivalent t o a Reynolds n u m b e r based on the is equivalent to letting Bv vary linearly with KC.
amplitude of oscillation. Kamphuis (1975) shows that transi-
tion to turbulence occurs in an oscillatory flat bed at an am- The resultant scaling laws are then obtained
plitude Reynolds number of about 10 4 . For the (3 and the range
ai«(/r0-5)
of K C tested, the b o u n d a r y layer is expected to be laminar,
and hence Eq. (5) is valid. The two expressions of Eq. (5) are
plotted in Figs. 2 and 3. Bu (friction) is independent of K C .
«2C (10)
Its value of 1.995 N - s / m (0.136 lb-sec/ft) is approximately 60
percent of the offset 3.424 N - s / m (0.234 lb-sec/ft) of the linear
fit to experimental d a t a in Fig. 2. T h e difference in the values
may be due to wall roughness, a n d / o r any pressure drag due
5 S c a l i n g L a w s f o r Oscillating H o r i z o n t a l Cylinders
to the b o t t o m of the cylinder. Based on the foregoing calcu-
lation, the friction drag, being linear with velocity, is the most Similar to the vertical cylinder case, the drag force for os-
significant c o m p o n e n t in the axis intercept of the least squares cillating horizontal cylinders depends o n the following vari-
curve fit of Fig. 2. The form drag due to vortex shedding then ables and parameters:
contributes t o t h e slope of the curve, i.e., it is linear with K C . Fd = Fd(Dh,L,oi,x,v,p)
Representing d a t a in terms of the equivalent linear d a m p i n g
coefficient Bv clearly illustrates the functional dependence of Defining the damping ratio for horizontal cylinders as
drag on K C .
In Fig. 3, for values of K C > 0 . 5 , the least squares fit to Cdv
f* =ir C,„
(11)
data almost becomes a flat line, i.e., independent of K C . For
an equation similar to E q . (8) is obtained
K C < 0 . 1 , it can be expected that the least squares curve will
be parallel to the friction drag line, since the form drag con-
tribution will be small. F r o m their experimental results, Chak- %h — Cm / ( — , K C ,
rabarti and H a n n a (1991) and Huse (1990) conclude that the
damping coefficient (B„) is independent of the amplitude of At small values of K C , the trend proposed by W a n g ' s (1968)
oscillation, and hence the K C n u m b e r . T h e K C range chosen theory can be used as scaling law. T h e range of applicability
by Huse (1990) is from 0.0005-0.01. F r o m Figs. 2 and 3, it of this laminar theory is restricted, however, due to occurrence
appears that for this K C range, the drag force is approximately of flow transition and separation. The value of K C at which
linear with velocity, i.e., Bv is a constant. However, H u s e (1990) transition occurs depends not only on the value of /3, but also
has concluded t h a t the vortex shedding force is also linear with on the experimental setup used. F r o m his experiments, Sarp-
velocity, in contrast to the present results. The theoretical kaya (1986) found that transition occurs at approximately KC
expression of Huse (1990) applied to the present case gives an •= 0.75 at /3 = 1035. Troesch and Kim (1991) report transition
equivalent Bv due to vortex shedding of 0.699 N - s / m (0.048 for circular cylinders at K C = 0.07 - 0.2 and 0 = 23200. These
lb-sec/ft). T h e total value for Bv using H u s e ' s (1990) method authors have observed that the measured drag coefficients fall
is 2.694 N - s / m (0.184 lb-sec/ft). in a line which has the same slope as the laminar theory line,
but at a higher intercept value on a log-log scale. Bearman
(1992) mentions n o transition at different /3 values (14371-
4 Scaling L a w s f o r Oscillating Vertical Cylinders 29014) in a K C range of 0 . 1 - 1 , t h o u g h all of his curves had
The damping force experienced by an oscillating vertical higher intercept values than laminar predictions. At (3 = 48600,
cylinder can be identified to be a function of the following Troesch and Kim (1991) see no transition over a KC range of
variables and parameters: 0.08-0.4, but still find a higher intercept. Sarpkaya (1986)

Journal of Offshore Mechanics and Arctic Engineering MAY 1994, Vol. 116/73

Downloaded From: http://offshoremechanics.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jmoeex/28093/ on 07/17/2017 Terms of Use: http://www.asme.org


10 Table 2 Example proptotype TLP (Petrauskas and Liu, 1987)

No of vertical columns 4
• Column diameter 15.85 m (52 ft)
Column draft 33.52 m (110 ft)
) No. of horizontal pontoons 4
&
* Pontoon diameter 8.53 m (28 ft)
Pontoon length 45.11 m(148 ft)
Total displacement 37724 t (2.584 x 106 slugs)
p 1 Estimated pre-tension 20 percent
CJ • • A Free-floating displacement 31437 t (2.154 x 106 slugs)
Free-floating draft 25.75 m (84.51 ft)
s$>2 Mass of each column 5212 t (3.571 X 105 slugs
Mass of each pontoon 2646 t (1.813 x 105 slugs)
Period of oscillation 2s

0.1
1E+2 1E+3 1E+4 1E+5 1E+6 f*=- (14)
P An effective eddy viscosity may be used in order to take the
intercept difference into account. The results of Troesch and
• Troesch & Kim (1991) experiment Kim (1991) for circular cylinders give
B Troesch & Kim (1991) extrapolated i>eff = 2 1 v

— Theory (Wang, 1968) The parameter /3 in Eq. (14) may be replaced by /3eff defined
as
« Bearman et al. (1985) experiment
(15)
n Sarpkaya (1986) experiment "eff

to obtain the scaling law


O Bearman (1992) experiment
A Anaturk (1991) experiment (16)
eff

Fig. 4 Drag coefficient (Cd„) for a horizontal cylinder versus 0; KC For square section cylinders, Troesch and Kim (1991) estimated
0.5 that
Ceff=51 V (17)
presents the idea of an effective eddy viscosity to explain this
difference in intercept. Troesch and Kim (1991) suggest that 6 Example TLP Calculation
an experimentally found effective eddy viscosity can be used
with laminar theory to derive empirical drag coefficients. The scaling laws described in the two previous sections are
applied to a prototype tension leg platform given by Petrauskas
Figure 4 is a plot of Cdh versus /3 at a KC value of 0.5.
and Liu (1987). Table 2 gives the dimensions of this prototype
Shown in the figure are results obtained by many authors,
TLP. Though this prototype does not scale perfectly with the
along with the laminar theory line of Wang (1968). Troesch
model dimensions of Table 1, the scaling laws can still be
and Kim's (1991) result shows an extrapolated point obtained
applied, since all quantities used in the calculations are di-
from the pre-transition least squares curve. Around /3 = 20000,
mensionless.
the drag coefficient suddenly increases, and decreases gradually
afterward. An effective eddy viscosity would generate lines Using the experimental data from Fig. 1, the scaling law for
parallel to the curve denoted as Wang's theory. Figure 4 sug- the vertical cylinder damping ratio (Eq. (9)) becomes
gests that an effective eddy viscosity correction, if valid, would
be applicable only over small ranges of /3. Alternatively, it may Jv ~~ C , „ ~0'5 + 0.0394 — KC (18)
be possible that at large values of /3, Cdh becomes independent
of (3. Evidence from this figure is not entirely conclusive, and For the vertical columns of the prototype
more results are required at higher /3 values to identify the
trend of Cdh. ^ = 0.473
Most authors agree that at low KC, the trend of Cdh is
satisfactorily represented by the laminar theory. Based upon
j3 = 9.324 xlO 7
this assumption, the theoretical expression for the drag coef-
ficient becomes (Wang, 1968; Troesch and Kim, 1991) Using Cm = 1.15, Eq. (18) yields the following scaling law for
the prototype:
Crf„=^"(KC)-/3- (12) f„ (prototype) = 8.718 x 10" 5 + 0.0162(KC) (19)
The friction drag contribution to the total damping is sur-
Following the procedure for Eq. (4), we obtain for a horizontal prisingly low for the prototype in comparison with the model.
cylinder This is to be partially expected, since the prototype /3 is much
larger than that of the model. At /3 = 9 x 107 and for very
low KC (say 0.01-0.1), the amplitude Reynolds number is in
(13) the range 103-105, and the boundary layer is laminar or in
3ir
transition. Based on the deductions of Kamphuis (1975), the
Inserting Eqs. (12) and (13) into the expression for ft, (Eq. effect of turbulence on friction drag is negligible over this range
(11)) of Reynolds number.

74 / V o l . 116, MAY 1994 Transactions of the ASME

Downloaded From: http://offshoremechanics.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jmoeex/28093/ on 07/17/2017 Terms of Use: http://www.asme.org


Table 3 Damping ratio (percent critical) of prototype TLP Table 4 Damping ratio (percent critical) of the prototype TLP
components (4 columns and 4 pontoons)
Component Method Damping ratio Pontoon K C = 0.01 K C = 0.1
(percent) section (percent) (percent)
Horizontal cylinder- Troesch and Kim 0.049 Circular 0.032 0.100
circular section (1991) Square 0.043 0.111
Horizontal cylinder- Troesch and Kim 0.078
square section (1991)
Vertical cylinder- Huse (1990) 0.062
circular section
Vertical cylinder- Present calculation 0.025 (KC = 0.01)The TLP example calculation demonstrates the need for
circular section additional research. Experimental data is required on damping
0.171 (KC = 0.1)
coefficients of both vertical and horizontal cylinders oscillating
at high values (in the range of 105 to 106) of/3. This will address
The damping ratio for the full scale TLP using Eq. (19) the following issues: (a) friction drag scaling of vertical cyl-
ranges from 0.0249 to 0.1708 percent for KC = 0.01-0.1. If inders at highft,and {b) ft dependence of the damping ratio
a designer were to use theory (Eq. (5)) to scale friction, and of horizontal cylinders. Several factors like wall roughness,
Huse's (1990) method for form drag, then the following values pitch/roll motions and interaction between structural members
would be obtained for the prototype: of a TLP have not been considered in the present analysis.
f„ (friction drag) = 0.005 percent Further research is required to investigate the influence of these
effects on damping.
f„(form drag) = 0.057 percent The TLP in the springing mode of oscillation is a lightly
giving a total damping ratio of 0.062 percent for oscillating damped system, and the resonant characteristics in this mode
vertical cylinders at small KC (<0.01). In comparison with have to be carefully evaluated in a good design. It is hoped
the present calculations, this method of scaling overestimates that this paper provides a useful design methodology for es-
the damping ratio at very low KC and underestimates it at low timating the damping coefficients of the oscillating compo-
to moderate KC. nents of a TLP.
For horizontal circular members, using Eqs. (15) and (16)
and Cm = 2, the results are Acknowledgments
6 This work was funded by the University of Michigan/Mich-
(3eff= 1.287 xlO
igan Sea Grant/Industry Consortium in Offshore Engineering
fA = 0.049 percent under Michigan Sea Grant College Program, projects number
If cylinders of square section replace the circular members, R/T-23 and R/T-29, under Grant No. NA85AA-D-SG045C
Eq. (17) yields from the Office of Sea Grant, National Oceanic and Atmos-
pheric Administration (NOAA), U. S. Department of Com-
5
ft,
e f f :5.149xl0
;
merce, and funds from the State of Michigan. Industry
participants include the American Bureau of Shipping; ARCO
ft, = 0.078 percent Gas and Oil Company; Conoco, Inc.; Exxon Production Re-
From these numerical values, it appears that as KC increases, search; Mobil Research and Development; Shell Companies
the vertical members contribute more to the total damping Foundation; and the U.S. Coast Guard. The authors also ap-
than the horizontal members. The results of this section are preciate the insightful comments of Professor T. Sarpkaya,
summarized in Table 3. Naval Postgraduate School. The help of the Marine Hydro-
When evaluating the damping ratio for the horizontal and dynamics Laboratory Staff during the experiments deserves
vertical cylinders, the local mass (of the column or pontoon) mention.
is used. This facilitates straight forward extrapolation from
model scale. Hence, one cannot sum the f contributions from
the individual structural components to obtain the total damp- References
ing ratio of the structure. Alternately, Eqs. (7) and (11) can Anaturk, A. R., 1991, "An Experimental Investigation to Measure Hydro-
be used to evaluate individual Bu and Bh, which can then be dynamic Forces at Small Amplitudes and High Frequencies," Applied Ocean
summed to give the total damping coefficient for the TLP. Research, Vol. 13, No. 4, pp. 200-208.
Anaturk, A. R., Tromans, P. S., van Hazendonk, H. C , Sluis, C. M., and
The damping ratios for the TLP obtained this way are tabulated Otter, A., 1992, "Drag Forces on Cylinders Oscillating at Small Amplitude: A
in Table 4. New Model," ASME JOURNAL OF OFFSHORE MECHANICS AND ARTIC ENGINEER-
ING, Vol. 114, pp. 91-103.
Batchelor, G. K., 1973, An Introduction to Fluid Dynamics, Cambridge Uni-
versity Press, London, U.K.
7 Conclusions Bearman, P. W., Downie, M. J., Graham, J. M. R., and Obasaju, E. D.,
1985, "Forces on Cylinders in Viscous Oscillatory Flow at Low Keulegan—
Experiments are conducted on a smooth vertical cylinder Carpenter Numbers," Journal of Fluid Mechanics, Vol. 154, pp. 337-356.
oscillating in heave. The frequency parameter (ft) is fixed at Bearman, P. W., 1992, "Fluid Loading of Cylinders With Application to
89236, and the Keulegan-Carpenter number (KC) varied from Risers: Results From Model Tests," NSF Workshop on Riser Mechanics, Ann
0.1-1.0. Measured values of the equivalent linear damping Arbor, MI.
Chakrabarti, S. K., and Hanna, S. Y., 1990, "Added Mass and Damping of
coefficient (Bv) have two components: (a) a KC-independ- a TLP Column Model," Proceedings 22nd Annual OTC, Paper 6406, pp. 559-
ent term due to friction drag, and (b) a term linear with KC '571.
due to form drag. Chakrabarti, S. K., and Hanna, S. Y., 1991, "High-Frequency Hydrodynamic
Scaling laws are presented for oscillating horizontal and Damping of a TLP Leg," Proceedings 10th International Offshore Mechanics
and Arctic Engineering, Vol. 1-A, pp. 147-151.
vertical cylinders. Damping calculations on a prototype tension Chew, Y. T., and Liu, C. Y., 1989, "Effects of Transverse Curvature on
leg platform described in Petrauskas and Liu (1987) show that Oscillatory Flow Along a Circular Cylinder," AIAA Journal, Vol. 27, No. 8,
at low KC, damping ratio (ft,) for horizontal circular pontoons pp. 1137-1139.
is about 0.049 percent, and for pontoons of square section, Couch, R. B., Troesch, A. W., and Ashcroft, F. H., 1984, "The Use of
Model Basins in the Design of Ships and Marine Structures," SNAME Spring
0.078 percent. For circular vertical columns, the damping ratio STAR Symposium, Los Angeles, CA, pp. 93-115.
(f„) at KC = 0.01 is estimated to be about 0.025 percent, Demirbilek.Z., ed., 1989, Tension Leg Platform: A Stale of the Art Review,
increasing to 0.171 percent at KC = 0.1. ASCE, New York, NY.

Journal of Offshore Mechanics and Arctic Engineering MAY 1994, Vol. 116/75

Downloaded From: http://offshoremechanics.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jmoeex/28093/ on 07/17/2017 Terms of Use: http://www.asme.org


Graham, J. M. R., 1980, "TheForces on Sharp-Edged Cylinders in Oscillatory Petrauskas, C , and Liu, S. V., 1987, "Springing Force Response of a Tension
Flow at Low Keulegan-Carpenter Numbers," Journal of Fluid Mechanics, Vol. Leg Platform," Proceedings 19th Annual OTC, Paper 5458, pp. 333-341.
97, Part 1, pp. 331-346. Riley, N., 1965, "Oscillating Viscous Flows," Mathematika, Vol. 12, pp. 161-
Hall, P., 1984, "On the Stability of the Unsteady Boundary Layer on a 175.
Cylinder Oscillating Transversely in a Viscous Fluid," Journal of Fluid Me- Sarpkaya, T., 1986, "Force on a Circular Cylinder in Viscous Oscillatory
chanics, Vol. 146, pp. 347-367. Flow at Low Keulegan-Carpenter Numbers," Journal of Fluid Mechanics, Vol.
Honji, H. 1981, "Streaked Flow Around an Oscillating Circular Cylinder," 165, pp. 61-71.
Journal of Fluid Mechanics, Vol. 107, pp. 509-520. Sarpkaya, T., and Isaacson, M., 1981, "Mechanics of Wave Forces on Off-
Huse, E., 1990, "Resonant Heave Damping of Tension Leg Platforms," shore Structures, van Nostrand Reinhold, New York, NY.
Proceedings, 22nd Annual OTC, Paper 6317, pp. 431-436. Stuart, J. T., 1966, "Double Boundary Layers in Oscillatory Viscous Flow,"
Kamphuis, J. W., 1975, "Friction Factor Under Oscillating Waves," Journal Journal of Fluid Mechanics, Vol. 24, Part. 4, pp. 673-687.
of Waterways Port Coastal Ocean Engineering, Vol. 101, WW2, pp. 135-144. Telionis, D. P., 19S1, Unsteady Viscous Flows, Springer-Verlag, New York,
Kim, S. K., and Troesch, A. W., 1989, "Streaming Flows Generated by High- NY.
Frequency Small-Amplitude Oscillations of Arbitrarily Shaped Cylinders," Troesch, A. W., and Kim, S. K., 1991, "Hydrodynamic Forces Acting on
Physics of Fluids, Vol. A 1, No. 6, pp. 975-985. Cylinders Oscillating at Small Amplitudes," Journal of Fluids and Structures,
Marthinsen, T., 1989, "Hydrodynamics in TLP Design," Proceedings 8th Vol. 5, pp. 113-126.
International Conference of Offshore Mechanics and Arctic Engineering, pp. Wang,C.-Y., 1968, "On High Frequency Oscillatory Viscous Flows," Journal
127-133. of Fluid Mechanics, Vol. 32, Part 1, pp. 55-68.

/TT\
s WORKING IN JAPAN ¥ ¥ AN INSIDER'S GUIDE FOR ENGINEERS
edited by Hiroshi Honda, R. C. Vonderau,
K. TaKaiwa, D. Day, and S. Fuhuda

C ontaining a wealth of useful information for any engineering professional planning to relocate to or visit Japan,
. this concise paperback gives practical answers to all of your questions about Japanese employment prac-
tices, culture, and society. Written by engineering professionals of many nationalities and backgrounds, WORKING IN
JAPAN is packed with insights and strategies that will help close the culture gap — from language barriers to
differences in the work ethic. Specifically, this book of practical insights features the following:

• guidance on which industries arc hiring foreign engineers • explanations of cultural differences you need lo Know
• practical information on hiring practices and salaries • tips for adjusting to daily life
• insights into future trends in employment • a useful glossary of Japanese terms

Prepare yourself now. Hake sure you know the hows and whys of working in Japan—whether you are already there,
making plans to go, or just thinking about possibilities — order your copies of WORKING IN JAPAN today.

TO ORDER :
In the U.S./Canada: \V* V , Fax: (201) 8 8 2 1 7 1 7 or (201) 8 8 2 5 1 55
( 8 0 0 ) T H E A S M E , ext. 618
(800 843 2763)
Write: ASME, Dept. M I S , 22 L a w Drive,
In Mexico:
9 5 , ( 8 0 0 ) 8 4 3 2 7 6 3 , ext. 618
Outside North America:
p^Tj B o x 2 9 0 0 , Fairfield, N J 0 7 0 0 7 - 2 9 0 0
ASME
(201)882 1 167. ext.618 <jQ Email: CompuServe 73302,1017 I KLlob

76/Vol. 116, MAY 1994 Transactions of the ASME

Downloaded From: http://offshoremechanics.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jmoeex/28093/ on 07/17/2017 Terms of Use: http://www.asme.org

S-ar putea să vă placă și