You are on page 1of 52

August 18, 2009 17:48 9.75in x 6.

5in b684-ch26 FA

Chapter 26

Prediction of Squat for Underkeel Clearance

Michael J. Briggs
Coastal and Hydraulics Laboratory
US Army Engineer Research and Development Center
3909 Halls Ferry Road, Vicksburg, MS 39180-6199, USA
michael.j.briggs@usace.army.mil

Marc Vantorre
Ghent University, IR04, Division of Maritime Technology
Technologiepark Zwijnaarde 904, B 9052 Gent, Belgium
marc.vantorre@ugent.be

Klemens Uliczka
Federal Waterways Engineering and Research Institute
Hamburg Office, Wedeler Landstrasse 157
D-22559 Hamburg, Germany
klemens.uliczka@baw.de

Pierre Debaillon
Centre d’Etudes Techniques Maritimes Et Fluviales
2 bd Gambetta, BP60039, 60321 Compiegne, France
pierre.debaillon@equipement.gouv.fr

This chapter presents a summary of ship squat and its effect on vessel underkeel
clearance. An overview of squat research and its importance in safe and efficient
design of entrance channels is presented. Representative PIANC empirical for-
mulas for predicting squat in canals and in restricted and open channels are dis-
cussed and illustrated with examples. Most of these formulas are based on hard
bottoms and single ships. Ongoing research on passing and overtaking ships in
confined channels, and offset distances and drift angles is presented. The effect
of fluid bottoms or mud is described. Numerical modeling of squat is an area of
future research and some comparisons are presented and discussed.

723
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

724 M. J. Briggs et al.

26.1. Introduction

When a ship travels through shallow water it undergoes changes in its vertical
position due to hydrodynamic forces from the flow of water and wave-induced
motions of heave, pitch, and roll. The focus of this chapter is on the former mech-
anism of ship squat. Squat is the reduction in underkeel clearance (UKC) between
a vessel at-rest and underway due to the increased flow of water past the moving
body. The forward motion of the ship pushes water ahead of it that must return
around the sides and under the keel. This water motion induces a relative velocity
between the ship and the surrounding water that causes a water-level depression in
which the ship sinks. The effect of shallow water and channel banks only exacerbates
these conditions. The velocity field produces a hydrodynamic pressure change along
the ship similar to the Bernoulli effect in that kinetic and potential energy must be
in balance.1 This phenomenon produces a downward vertical force (sinkage, pos-
itive downward) and a moment about the transverse axis (trim, positive bow up)
that can result in different values of squat at the bow and stern (Fig. 26.1). This
combination of sinkage and change in trim is called ship squat.
Most of the time squat at the bow, Sb , represents the maximum value, especially
for full-form ships, such as supertankers. In very narrow channels or canals and
for high-speed (fine-form) ships, such as passenger liners and containerships, the
maximum squat can occur at the stern Ss . The initial trim of the ship also influences
the location of the maximum squat. The ship will always experience maximum squat
in the same direction as the static trim.2 If trimmed by the bow (stern), maximum
squat will occur at the bow (stern). A ship trimmed by the bow or stern when static
will remain that way and will not level out when underway to offset the sinkage at
the bow or stern due to squat.
So why do we care about ship squat? For one thing, ship squat has always
existed, but was less of a concern with smaller vessels and with relatively deeper
channels. The new supertankers and supercontainerships have smaller static UKC
and higher service speeds. Secondly, the goal of all ports is to provide safe and effi-
cient navigation for waterborne commerce. Since operation and maintenance costs
continue to escalate and can easily exceed $3M per vertical meter, it is imperative
to minimize required channel depths and associated dredging costs. Finally, even
though we have a pretty good handle on squat predictions, accidents continue to
occur. Barrass3 noted that there have been 12 major incidents between 1987 and
2004. In 2007, this number of ship incidents had increased to as many as 82 that
are partially attributable to ship squat.4 The luxury passenger liner QEII grounded

Fig. 26.1. Schematic of ship squat at bow and stern.


August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 725

off Massachusetts in 1992 with a repair cost of $13M and another $50M for lost
passenger bookings.
In the early 1990s, the Maritime Commission (MarCom) of the Permanent Inter-
national Association of Navigation Congresses (PIANC) formed a working group
(WG30) to provide information and recommendations on the design of approach
channels.5 In the past 10 years since the WG30 report, research in squat predic-
tions was a dynamic area in naval architecture with new experiments to study
the effects of fluid bottoms and passing and overtaking vessels, especially with the
increasing size of the shipping fleet. Time domain Reynolds Average Navier–Stokes
Equation (RANSE) numerical models are being developed to predict squat, but
these models are still being validated. In 2005, the PIANC MarCom formed a new
working group Horizontal and Vertical Dimensions of Fairways (WG49) to update
the WG30 report on design of deep draft navigation channels.6
A summary of ship squat is presented in this chapter. In the second section,
factors governing squat including ship characteristics, channel configurations, and
combined factors are discussed. Some empirical formulas from the PIANC WG30
report are presented and compared in the third section. The fourth section presents
some recent research on the effect of squat on passing and overtaking ships in con-
fined channels by the Federal Waterways Engineering and Research Institute (BAW)
in Hamburg, Germany, and the Flanders Hydraulic Research (FHR) Laboratory in
Antwerp, Belgium. It also includes numerical modeling by Delft University of Tech-
nology and laboratory modeling by FHR on the effect of ship offset and drift on
squat. The fifth section summarizes the recent studies at FHR on the effect of fluid
bottoms (i.e., mud) on squat. The development of numerical models to predict ship
squat is an ongoing research area. The current status of this development at Centre
d’Etudes Techniques Maritimes Et Fluviales (CETMEF), France, is discussed in the
sixth section. Finally, a summary and conclusions of ship squat issues is presented
in the last section.

26.2. Factors Governing Squat

Prediction of ship squat depends on ship characteristics and channel configurations.


These factors are often combined to create new normalized parameters to describe
the squat phenomenon.

26.2.1. Ship characteristics


The main ship parameters include ship draft, T , hull shape as represented by the
block coefficient, CB , and ship speed, VS (m/s) or VK (knots). Other ship parameters
include the length between forward and aft perpendiculars Lpp and the beam, B.
The CB is a measure of the “fineness” of the vessel’s shape relative to an equivalent
rectangular volume with the same dimensions. The range of values of CB is typically
between 0.45 for high-speed vessels and 0.85 for slow, full-size tankers and bulk
carriers. The most important ship parameter is its speed VS . This is the relative
speed of the ship in water, so fluvial and tidal currents must be included. In general,
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

726 M. J. Briggs et al.

squat varies as the square of the speed. Therefore, doubling the speed quadruples
the squat and vice versa.
There are two calculated ship parameters that are based on the basic ship dimen-
sions. The ship’s displacement volume ∇ (m3 ) is defined as

∇ = CB Lpp BT. (26.1)

The CB can be determined from the ∇ if the other ship dimensions are known. The
underwater midship cross-sectional area AS is generally defined as

AS = 0.98BT. (26.2)

The “0.98” constant accounts for reduction in area due to the keel radius.7 Some
researchers ignore this and use a constant of “1.00” since the error is small relative
to other uncertainties in the squat calculations.
Finally, the bulbous bow and stern-transom are two other characteristics of a
ship that affect squat. Many of the early squat measurements were made before
bulbous bows were in use. Newer designs of bulbous bows, although mainly to reduce
drag and increase fuel efficiency, also have an effect on squat. The newer “stern-
transoms” on some ships are “blockier” (i.e., wider and less streamlined) than earlier
ship designs and affect squat as they become more fully submerged with increases
in draft.8

26.2.2. Channel configurations


The main channel considerations are proximity of the channel sides and bottom, as
represented by the channel depth h and cross-sectional configuration. If the ship is
not in relatively shallow water with a small UKC, squat is usually negligible. Ratios
of water depth to ship draft h/T greater than 1.5–2.0 (i.e., relatively deepwater)
are usually considered safe from the influences of squat.
The main types of “idealized” channel configuration are (a) open or unrestricted
(U), (b) confined or restricted (R), and (c) canal (C). Figure 26.2 is a schematic of
these three types of entrance channels for ocean-going or deep draft ships. Unre-
stricted channels are in relatively larger open bodies of water and usually toward
the offshore end of entrance channels. Analytically and numerically, they are easier
to describe and were some of the first types studied. Sections of rivers may even

Fig. 26.2. Schematic of three channel types: unrestricted or open, restricted or confined, and
canal.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 727

be classified as unrestricted channels if they are wide enough. The second type
of channel is the restricted channel with an underwater trench that is typical of
dredged channels. The restricted channel is a cross between the canal and unre-
stricted channel type. The trench acts as a canal by containing and influencing the
flow around the ship, and the water column above the hT allows the flow to act
as if the ship is in an unrestricted channel. The last type of channel is the canal.
These channels are representative of channels in rivers with emergent banks. The
sides are idealized as one slope when in reality they may have compound slopes with
revetment to protect against ship waves and erosion. The canal may or may not
be exposed to tidal fluctuations. For instance, the Panama and Suez Canals have a
constant water depth.
Many channels can be characterized by two or three of these channel types as the
different segments or reaches of the channel have different cross-sections. Finally,
many real-world channels look like combinations of these three types as one side may
look like an open unrestricted channel and the other side like a canal or restricted
channel with side walls. Most of the PIANC empirical formulas are based on ships
in the center of symmetrical channels, so the user has to use “engineering judgment”
when selecting the most appropriate formulas. New data are being collected for
some of these more realistic channel shapes, so future formulas may account for
these differences in channel shapes.
Other important parameters necessary to describe restricted channels and canals
are the channel width at the bottom of the channel W , trench height hT from the
bottom of the channel to the top of the trench, and inverse bank slope n (i.e.,
run/rise = 1/ tan θ). The value of n, although not necessarily an integer, typically
has a value such as 1, 2, or 3 representing side slopes of 1:1, 1:2, and 1:3, respectively.
How does one define the width of an unrestricted or an open channel since there
are no banks or sides? In 2004, Barrass had defined an effective width Weff for the
unrestricted channel as the artificial side boundary on both sides of a moving ship
where the ship will experience changes in performance and resistance that affect
squat, propeller RPMs, and speed.3 His width of influence FB is defined for h/T
values from 1.10 to 1.40 as
 
7.04
FB = Weff = B. (26.3)
CB0.85

Mean values of FB are of the order of 8B to 8.3B for supertankers (CB range from
0.81 to 0.87), 9B to 9.5B for general cargo ships (CB range from 0.68 to 0.80), and
10B to 11.5B for containerships (CB range from 0.57 to 0.71).
The calculated cross-sectional area AC is the wetted cross-section of the canal
or the equivalent wetted area of the restricted channel by projecting the slope to
the water surface. It is given by

AC = W h + nh2 . (26.4)

For an unrestricted channel, use Barrass’s effective width Weff for channel width W
and set n = 0 in the equation for AC .
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

728 M. J. Briggs et al.

26.2.3. Combined ship and channel factors


Several dimensionless parameters are required in the PIANC squat prediction for-
mulas that are ratios of both ship and channel parameters. They include the depth
Froude number Fnh and the blockage factor S.
The most important dimensionless parameter is Fnh , which is a measure of the
ship’s resistance to motion in shallow water. Most ships have insufficient power to
overcome Fnh values greater than 0.6 for tankers and 0.7 for containerships. Most
of the empirical equations require that Fnh be less than 0.7. For all cases, the value
of Fnh should satisfy Fnh < 1, an effective speed barrier and the defining level for
the subcritical speed range. The Fnh is defined as
Vs
Fnh = √ (26.5)
gh
with gravitational acceleration g (m/s2 ).
The blockage factor S is the fraction of the cross-sectional area of the waterway
AC that is occupied by the ship’s underwater midships cross-section AS defined as
AS
S= . (26.6)
AC
Typical S values can vary from 0.03 to 0.25 or larger for restricted channels and
canals, and to 0.10 or less for unrestricted channels.3,9 Higher values may occur,
for example, the canal from Terneuzen (The Netherlands) to Ghent (Belgium) is
operated with a blockage factor S = 0.275, and higher values will be evaluated in
the near future.10 The value of S is a factor in the calculation of the ship’s critical
speed in canals and restricted channels (see Sec. 26.3.3.3 and Appendix 26.A).

26.3. PIANC Squat Formulas

26.3.1. Background
In 1997 the PIANC WG30 report included 11 empirical formulas and one graphical
method from nine different authors for the prediction of ship squat.5 They were
based on physical model experiments and field measurements for different ships,
channels, and loading characteristics. The formulas included the pioneering work of
Tuck11 , Tuck and Taylor,12 and Beck et al.,13 and the early research by Hooft,14
Dand,15 Eryuzlu and Hausser,16 Römisch,17 and Millward.18,19 The PIANC recom-
mends that channels be designed in two stages. The first is the “Concept” Design
where a “quick” or “ballpark” answer is desired. The WG30 report recommended the
International Commission for the Reception of Large Ships (ICORELS) formula20
in this phase. The second stage is the “Detailed” Design phase where more accurate
and thorough predictions and comparisons are required. The WG30 recommended
the formulas by ICORELS, Huuska,7 Barrass,21,22 and Eryuzlu et al.23 in this
second stage.
All of these formulas give predictions of bow squat Sb , but only the Römisch
formula gives predictions for stern squat Ss for all channel types. The Barrass
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 729

formula gives Ss for unrestricted channels, and for canals and restricted channels
depending on the value of CB . Each formula has certain constraints that it should
satisfy before being applied, usually based on the ship and channel conditions under
which it was developed. Caution should be exercised if these empirical formulas are
used for conditions outside those for which they were developed.
In 2005 the PIANC MarCom formed WG49,6 which is in the process of reviewing
and revising these formulas for an updated report on channel design (expected to be
completed in 2010). There have been some new formulations since the WG30 report
that are being evaluated. Barrass has continued to develop and refine his formulas
and now has predictions for both Sb and Ss . Ankudinov et al.24 proposed the Mar-
itime Simulation and Ship Maneuverability (MARSIM) 2000 formula for maximum
squat based on a midpoint sinkage and vessel trim in shallow water. It is one of the
most thorough and the most complicated formulas for predicting ship squat. The
St. Lawrence Seaway (SLS) Trial and Very Large Crude Carriers (VLCC) formulas
are based on the prototype measurements in the SLS by Stocks et al.25 Briggs26
developed a FORTRAN program to calculate squat using most of these formulas.
It is not possible to include all the formulas in this chapter. We have selected
a representative sample of formulas that can be used for both phases of design.
Some are the “old tried and true” formulas and some are based on new research.
The Concept Design phase is by definition the simplest, of course this does not
necessarily mean that these formulas are any less accurate than some of the more
complicated formulas. In the Detailed Design phase, it is usually a good practice to
evaluate the squat with several of the formulas and calculate some statistics such
as average and range of values. In some cases, the maximum squat values might be
used in design for the case of dangerous cargo and/or hard channel bottoms.
The user should always be mindful for the original constraints. Some of these
constraints are very restrictive (especially for the newer vessels coming on line) as
they are based on the limited set of conditions tested in physical models by the
individual researchers. This does not mean that the particular formula would not
be applicable if the constraints are exceeded by a reasonable amount. Therefore,
the user should exercise Engineering Judgment when deciding the applicability of
those predictions. Table 26.1 summarizes the applicable channel configurations and

Table 26.1. Channel configurations and parameter constraints for PIANC squat formulas.

Configuration Constraint
Code ID Code ID

Formulas U R C CB B/T h/T hT /h L/B L/h L/T

Barrass27 Y Y Y 0.5–0.85 1.1–1.4


Eryuzlu
et al.23 Y Y ≥ 0.8 2.4–2.9 1.1–2.5 6.7–6.8
Huuska7 Y Y Y 0.6–≥0.8 2.19–3.5 1.1–2.0 0.22–0.81 5.5–8.5 16.1–20.2
ICORELS20 Y
Yoshimura28 Y Y Y 0.55–0.8 2.5–5.5 ≥1.2 3.7–6.0
Römisch17 Y Y Y 2.6 1.19–2.25 8.7 22.9

Notes: 1. Huuska/Guliev originally for Fnh ≤ 0.7.


August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

730 M. J. Briggs et al.

parameter constraints according to the individual testing conditions for the formulas
in this chapter.

26.3.2. Concept design


26.3.2.1. ICORELS
The ICORELS formula20 for bow squat Sb is one of the original formulas from the
PIANC WG30 report.5 It was developed for unrestricted or open channels only, so
it should be used with caution if applied for restricted and canal channels. It is
similar to Hooft’s14 and Huuska’s7 equations and is defined as
2
∇ Fnh
Sb = CS  (26.7)
L2pp 1 − Fnh
2

where CS = 2.4 and the other factors have been previously defined.
The Finnish Maritime Administration (FMA) uses this formula with different
values of CS depending on the ship’s CB .29,30


 1.7 CB < 0.70
CS = 2.0 0.70 ≤ CB < 0.80 . (26.8)

 2.4 C ≥ 0.80
B

The BAW, however, recommends a value of CS = 2.0 for the larger containerships
of today which may have a CB < 0.70. Their research is based on many measure-
ments along the restricted channel (side slope n varies from 15 to 40), 100-km long,
River Elbe.31 The wider stern-transom ships (see Sec. 4.3) require CS = 3 because
of the increased bow squat. The FHR has found CS ≥ 2.0 for modern container-
ships. They typically travel at much higher speeds than the ICORELS formula was
originally developed, even in shallow and restricted waters. The Fnh are higher and
in this speed range the effect of blockage S on the critical ship speed is considerable.
For example, a very small S = 0.01 results in an important decrease in critical
speed.10

26.3.2.2. Barrass
The Barrass4,27 formula is one of the simplest and “user friendly” and can be
applied for all channel configurations. Based on his earlier work in 1979,21 1981,22
and 2004,3 the maximum squat SMax at the bow or stern is determined by the value
of ship’s CB and Vk as

KCB Vk2
SMax = . (26.9)
100
According to Barrass,2 the value of CB determines whether SMax is at the bow
Sb or stern SS (requires even keel when static). He notes that full-form ships with
CB > 0.7 tend to squat by the bow and fine-form ships with CB < 0.7 tend to
squat by the stern. The CB = 0.7 is an “even keel” situation with squat the same
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 731

at both bow and stern. Of course, for channel design, one is mainly interested in
the maximum squat and not necessarily whether it is at the bow or stern.
This formula is based on a regression analysis of more than 600 laboratory and
prototype measurements. Stocks et al.25 found that the Barrass formulas gave the
best results for New and Traditional Lakers in the Lake St. Francis area (unre-
stricted channel) of the SLS. The BAW feels that the Barrass restricted formula is
conservative for their restricted channel applications in the Elbe River.
The coefficient K 4 is defined in terms of blockage factor S as

K = 5.74S 0.76. (26.10)

A value of S = 0.10 is equivalent to a “wide” river (unrestricted or open water


conditions). The value of K = 1 and the denominator in the equation for SMax
remains 100. If S < 0.10, the value of K should be set to 1. For restricted channels,
a value of the order of S = 0.25 gives a value of K = 2, and the denominator
becomes 50. Thus, the effect of K is to modify the denominator constant between
values of 50 to 100. Constraints on these equations are 1.10 ≤ h/T ≤ 1.40 and
0.10 ≤ S ≤ 0.25. This equation can accommodate a medium width river with a
value of S between the limits of S above.
For ships in unrestricted channels that are at even keel when in a static condition
(i.e., moored), one can estimate the squat at the other end of the ship (either bow
or stern) based on SMax . Thus, if CB indicates the ship will squat by the bow, then
this formula will give the squat at the stern, and vice versa:

2 Sb CB ≤ 0.7
[1 − 40(0.7 − CB ) ]SMax = . (26.11)
SS CB > 0.7

26.3.2.3. Yoshimura
The Overseas Coastal Area Development Institute of Japan32 and Ohtsu et al.33
proposed the following formula for Sb as part of their new Design Standard for
Fairways in Japan. This formula was originally developed by Yoshimura28 for open
or unrestricted channels typical of Japan. The range of parameters for which this
formula is applicable is shown in Table 26.1. In 2007, Ohtsu34 proposed a small
change to the ship velocity term Vs (last factor in the equation is now Ve ) to include
S to improve its predictions in restricted channels and canals:

 Vs Unrestricted
Ve = Vs . (26.12)
 Restricted, canal
(1 − S)

Their Sb predictions generally fall near the average for most of the other PIANC
bow squat predictions, regardless of ship type:



3 2
1 CB 1 CB Ve
Sb = 0.7 + 1.5 + 15 . (26.13)
h/T Lpp /B h/T Lpp /B g
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

732 M. J. Briggs et al.

26.3.3. Detailed design


26.3.3.1. Eryuzlu
One of the more recent series of physical model tests and field measurements was
conducted by Eryuzlu et al.23 for cargo ships and bulk carriers with bulbous bows
in unrestricted and restricted channels. Their tests used self-propelled models with
bulbous bows. Many of the early PIANC formulas did not have ships with bulbous
bows. The range of ship parameters was somewhat limited with CB ≥ 0.8, B/T
from 2.4 to 2.9, and Lpp/B from 6.7 to 6.8. The Eryuzlu formula should not be
used for containerships unless they meet this CB criteria. They conducted some
supplemental physical model tests with an hT /h = 0.5 and n = 2 to investigate the
effect of channel width in restricted channels. The Canadian Coast Guard35 is using
the Eryuzlu et al.23 formula exclusively. Stocks et al.25 recommended the Eryuzlu
formula for the chemical tankers in the Lake St. Louis section (unrestricted channel)
of the SLS.
The Eryuzlu formula for Sb is defined as

2.289
−2.972
h2 Vs h
Sb = 0.298 √ Kb . (26.14)
T gT T
Note that the Ship Froude number rather than Fnh is used in their equation since
the ship draft T is used in the denominator instead of the channel depth h.
The Kb is a correction factor for channel width W relative to ship’s B given by

 3.1 W

 < 9.61
W/B B
Kb = . (26.15)

 W
1 ≥ 9.61
B
One should use the second value of Kb = 1 for unrestricted channels regardless
of effective width Weff since the channel has no boundary effects on the flow and
pressures on the ship.

26.3.3.2. Huuska/Guliev
The next empirical formula in the Detailed Design phase is by Huuska.7 This Finnish
professor extended Hooft’s work for unrestricted channels to include restricted
channels and canals by adding a correction factor for channel width Ks that Guliev36
had developed. The Spanish ROM 3.1-99 (Recommendations for Designing Mar-
itime Configuration of Ports, Approach Channels, and Floatation Areas37 ) and the
FMA recommend the Huuska/Guliev formula for all three channel configurations.
In general, this formula should not be used for Fnh > 0.7. The FMA29 also includes
some additional constraints for lower and upper limits as follows (Table 26.1):
• CB 0.60 to 0.80
• B/T 2.19 to 3.50
• Lpp /B 5.50 to 8.50
• hT /h 0.22 to 0.81
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 733

Fig. 26.3. Huuska/Guliev K1 versus S.

The Huuska/Guliev formula is defined as

2
∇ Fnh
Sb = CS  Ks . (26.16)
L2pp 1 − Fnh
2

The squat constant CS = 2.40 is typically used as an average value in this formula.
The value for Ks for restricted channels and canals is determined from

7.45s1 + 0.76 s1 > 0.03
Ks = (26.17)
1.0 s1 ≤ 0.03

with a corrected blockage factor s1 defined as

S
s1 = . (26.18)
K1

The correction factor K1 is given by Huuska’s plot of K1 versus S for different


trench height ratios hT /h shown in Fig. 26.3. One should use a value of hT = 0 for
unrestricted channels and hT = h for canals. Appendix 26.A contains a set of least
square fit coefficients for Fig. 26.3 if one wants to program these curves.26

26.3.3.3. Römisch
Römisch17 developed formulas for both bow and stern squat from physical model
experiments for all three channel configurations. His empirical formulas are some of
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

734 M. J. Briggs et al.

the most difficult to use, but seem to give good predictions for bow Sb and stern
squat Ss given by

Sb = CV CF K∆T T
Ss = CV K∆T T (26.19)

where CV is a correction factor for ship speed, CF is a correction factor for ship
shape, and K∆T is a correction factor for squat at ship critical speed. The value for
CF is equal to 1.0 for the stern squat. The values for these coefficients are defined as

2
4
V V
CV = 8 − 0.5 + 0.0625 (26.20)
Vcr Vcr

2
10CB
CF = (26.21)
Lpp /B

K∆T = 0.155 h/T . (26.22)

The ship critical or Schijf-limiting speed Vcr is the speed that ships cannot exceed
due to the balance between the continuity equation and Bernoulli’s law.9,38,39 For
economic reasons, maximum ship speeds are typically only 80% of Vcr . The Vcr
(m/s) varies as a function of the channel configuration given by


 CKU Unrestricted
Vcr = Cm KC Canal . (26.23)


CmT KR Restricted

The three-wave celerity parameters C, Cm , and CmT (m/s) are defined as


  
C= gh; Cm = ghm ; CmT = ghmT . (26.24)

The mean water depth hm (m) is a standard hydraulic parameter that is used for
canals and restricted channels. It is defined as
AC
hm = (26.25)
WTop

where WTop (m) is the projected channel width at the top of the channel equal to

WTop = W + 2nh. (26.26)

The relevant water depth hmT (m) is for restricted channels and is defined as

hT
hmT = h − (h − hm ).
h
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 735

Table 26.2. Römisch’s KC versus 1/S.

1/S 1 6 10 20 30 ∞
KC 0.0 0.52 0.62 0.73 0.78 1.0

Römisch’s correction factors KU , KC , and KR for unrestricted, canal, and


restricted channels, respectively, are defined as


0.125
h Lpp
KU = 0.58 (26.27)
T B

1.5
Arc sin(1 − S)
KC = 2 sin (26.28)
3
KR = KU (1 − hT /h) + KC (hT /h). (26.29)

Note that the KR for the restricted channel is a function of both KU and KC .
Table 26.2 lists Römisch’s limited dataset for KC as a function of 1/S (i.e., AC /AS ).
Appendix 26.A contains more detailed descriptions of KC and some additional
equations for defining it relative to Schijf’s limiting speed and his limiting Froude
number FHL .

26.3.4. Example problems


Three example problems are presented in this section to illustrate the different
formulas for several channel and ship types. All are for bow squat Sb . Comparisons
of the different formulas with the measured laboratory values are shown for each
example in Figs. 26.4–26.6, respectively. Appendix 26.B contains worked examples
for at least one Concept and one Detailed Design application for each example
problem.

26.3.4.1. Example 1: BAW Post-Panamax containership


in unrestricted channel
The first example is for a Post-Panamax containership traveling at Vk = 13.3 kt
(Vs = 6.84 m/s) in an unrestricted channel. This speed matches laboratory data
(Sb = 0.70 m) obtained at BAW by Flügge and Uliczka.40,41 This vessel is similar
to the last generation Emma Maersk containership (launched in August 2006), but
with a larger CB . The larger CB is not realistic for the newer containerships (most
have CB < 0.7), but was tested by BAW by “lengthening” an existing model during
design experiments. A comparable CB is of the order of 0.62 for a ship of this size.
The dimensions of the ship and channel are listed in Table 26.3.
Figure 26.4 shows comparisons among the Barrass, Eryuzlu, Huuska, ICORELS,
Römisch, and Yoshimura formulas and the measured BAW laboratory values. The
numerical values are from a numerical model described in Sec. 26.6. In general, the
best formulas are the Yoshimura (Concept) and Eryuzlu (Detail) as they are slightly
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

736 M. J. Briggs et al.

0.0
Bow Squat for BAW Hansa Container Ship - Unrestricted

R
Barrass
R Eryuzlu
N
R Huuska
-0.5 N R ICORELS
N R R Romisch
R
Yoshimura
N N Numerical
R BAW
Sb, m

N
-1.0 R

N
-1.5
Example R
N

Unrestricted Channel Bottom


-2.0
9 11 13 15 17 19
Vk, knots

Fig. 26.4. Comparison of BAW’s experimental measurements, empirical formulas, and numerical
model of bow squat for a Post-Panamax containership in an unrestricted channel (open water).

0.0
Bow Squat for FHR Tanker G, Condition C - Canal

Barrass
Huuska
R Romisch
Yoshimura
N N Numerical
-0.5 FHR
R
R N
S b, m

R
Example
R N

R
-1.0
R

N R

Canal Bottom
-1.5 R

7 8 9 10 11 12
Vk, knots

Fig. 26.5. Comparison of FHR’s experimental measurements, empirical formulas, and numerical
model of bow squat for a Tanker “G”, in Condition C in a canal with vertical sides.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 737

0.0
Bow Squat for Tothil Canadian Laker - Canal
R
R
N R
R
N R

R
N
-0.5
R
Sb, m

R
Barrass
Huuska
R Romisch
-1.0
Yoshimura
N Numerical
N
Tothil Example R

-1.5 Canal Bottom

4 5 6 7 8
Vk, knots

Fig. 26.6. Comparison of Tothil’s experimental measurements, empirical formulas, and numerical
model of bow squat for a Canadian Laker in a canal.

Table 26.3. BAW’s Post-Panamax containership


in unrestricted channel.

Lpp (m) B (m) T (m) CB h (m)

400 50 17 0.84 19

conservative (i.e., larger than measured). The Römisch is slightly smaller than the
measured values, but follows the trend very well. Appendix 26.B contains worked
examples for the Concept Design formulas of Yoshimura and ICORELS and the
Detail Design formulas of Eryuzlu and Römisch.

26.3.4.2. Example 2: FHR “G” Tanker in a canal with vertical side,


Condition C
The second example is for the “G” Tanker, Condition C in a canal with vertical sides
(similar to a restricted channel) from FHR and Ghent University.42 The 1:50 scale
laboratory experiments were performed in a 7.0-m-wide (350-m prototype) towing
tank. The measured Sb = 1.18 m for the ship sailing at Vk = 10 kt (Vs = 5.14 m/s).
The ship and channel characteristics are listed in Table 26.4.
Figure 26.5 shows comparisons among the Barrass, Huuska, Römisch, and
Yoshimura formulas and the measured FHR laboratory values for the canal with
vertical sides. The numerical values are from a numerical model that is described in
Sec. 26.6. In general, the best formulas are the Yoshimura (Concept) and Römisch
(Detail) as they are nearly exact or slightly conservative for the smaller ship speeds
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

738 M. J. Briggs et al.

Table 26.4. FHR “G” Tanker in restricted channel, Condition C.

Lpp (m) B (m) T (m) CB h (m) hT (m) W (m) WTop (m) n (deg)

180 33 13 0.85 14.5 14.5 350 350 0.0

(i.e., larger than measured). Appendix 26.B contains worked examples for the
Yoshimura, Barrass (Concept), and Huuska (Detail). The Römisch is not included
in the worked examples for this case as it has already been demonstrated. The
Barrass is a little small, especially for higher ship speeds. The Huuska formula is
conservative for all ship speeds.

26.3.4.3. Example 3: Tothil’s Canadian Laker in a canal


The third example is for a Canadian Laker in a canal with sloping sides (typical
canal). These data are from Tothil’s 1:48 scale model experiments.43 The measured
Sb = 0.93 m for the ship traveling at 6.98 kt (Vs = 3.59 m/s). Ship and channel
features are listed in Table 26.5.
Figure 26.6 shows comparisons among the Barrass, Huuska, Römisch, and
Yoshimura formulas and the measured Tothil laboratory values for the canal case.
The numerical values are from a numerical model that is described in Sec. 26.6. In
general, the best formulas are the Barrass (Concept), Huuska (Detail), and Römisch
(Detail). The Barrass is a good match for ship speeds less than 6.54 kt, but does
not follow the measured values for increasing speeds. The Huuska is on the low side,
but matches reasonably well until Vk exceeds 6.54 kt. The Römisch is on the low
side, but follows the measured trend of the data for all speeds. The Barrass and
Römisch formulas are included in worked examples in Appendix 26.B.

26.4. Recent Investigations of Ship Squat

So far we have discussed the PIANC empirical formulas for predicting ship squat.
These are based on “idealized” conditions with single vessels that are sailing along
the centerline of symmetrical channels. Unfortunately, real-world channels and ship
transits are seldom this simple. This section discusses some recent research in lab-
oratory and field measurements of ship head-on passing encounters and overtaking
maneuvers in two-way traffic, stern-transom effects, abrupt sills, and offset and drift
angle effects for ships sailing off the centerline with drift angles.
When two ships pass or overtake each other, the water flow and corresponding
squat is affected as a function of the other ship’s size, speed, and direction of
travel, and the channels configuration. Dand44 was one of the first to study this

Table 26.5. Tothil’s Canadian Laker in a canal.

Lpp (m) B (m) T (m) CB h (m) W (m) WTop (m) n

215.6 22.9 7.77 0.86 9.33 72.3 105.9 1.8 (29 deg)
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 739

phenomenon. He found increases in bow squat of 50–100% during passing and over-
taking encounters.
During the past 10 years, the BAW has conducted many field and labo-
ratory studies to investigate ship–waterway interactions, especially head-on passing
encounters and overtaking maneuvers of ships in restricted channels within German
federal waterways. Preliminary studies of the dynamic response of large container-
ships in laboratory models have shown tendencies of reduced squat.40,41,45 These
results were confirmed by additional model tests in restricted and unrestricted
channels and field measurements along the Elbe River.31 The FHR (in cooperation
with the Ghent University) has conducted laboratory experiments to study passing
and overtaking in their automated towing tank as part of a larger study to improve
their ship simulator for traffic in Flemish waterways.46 Finally, the Delft University
of Technology47 had conducted some numerical modeling of the effects of ship offset
and drift angles on ship squat. Thus, this section presents a summary of recent
laboratory, field, and numerical investigations of ship squat in real-world situations
including head-on passing encounters, overtaking maneuvers, wider stern-transoms,
and ships with offset and drift angles.

26.4.1. Head-on passing ship encounters


26.4.1.1. BAW laboratory experiments
Laboratory experiments were conducted at the BAW-DH shallow water basin to
study squat as a function of ship size, hull form, draft, speed, direction of travel,
and channel water level. This facility has approximate dimensions of 100-m length,
35-m width, and 0.7-m maximum water depth. Geometric and dynamic conditions
were accurately scaled according to dimensional analysis at a scale of 1:40. A section
of the River Elbe (i.e., restricted or confined channel) with a width of 1.0 km and
length of 1.5 km was modeled. The cross-section had a channel depth h = 18.5 m,
channel width of 265 m, and river width of 850 m. The results of a Panamax (PM)
containership (PM32) and a Post-Panamax (PPM) bulk carrier (MG58) during
head-on passing were investigated (Table 26.6).
Note that the MG58 is the larger vessel. The two ships passed each other at a
passing distance of 156 m (between course lines). The range of ship speeds for the
two ships was approximately 7–14 kt for the PM32 and 7–12 kt for the MG58.
A laser measuring system was installed on the self-propelled, cable-guided model
ships to record their vertical behavior. Measurements were recorded over a dis-
tance of approximately 90 m, including acceleration and braking phases of each
run. The velocity-independent precision of the laser system was ∆S <1 mm model,

Table 26.6. BAW ship head-on passing characteristics.

Code Description Lpp (m) B (m) T (m) CB

PM32 Panamax containership 280 32.2 12.8 0.68


MG58 Bulk carrier 349 58 14.5 0.80
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

740 M. J. Briggs et al.

Fig. 26.7. Laboratory measurement of the effect of head-on passing on bow and stern squat for a
PM containership passing a large bulk carrier in the River Elbe. The dark blue curves represent the
single runs of the containership; the light blue curves the encounters with the large bulk carrier.

corresponding to <4 cm in the prototype. Additional measurements with a point,


laser-geometric method allowed for correlation of squat as a function of ship speed,
with an accuracy of ∆S < 1 mm (model) for speeds up to 14 kt prototype.
Figure 26.7 illustrates the effect of passing on bow and stern squat for the smaller
PM32 containership. The measured squat for the single PM32 sailing by itself is
shown in dark blue. The effect of the larger MG58 bulk carrier on the PM32 squat
is shown in light blue. An additional increase in maximum bow squat for the PM32
(Vk = 14 kt) of ∆S ≈ 0.6 m was recorded due to the passing encounter with the
MG58 (Vk = 12 kt). The trim of the PM32 changed from even keel for single runs
in the channel and low speeds to bow trim at higher passing speeds during the
encounter situation. Figure 26.8 is the analogous figure for the larger MG58 bulk
carrier, but shown in red colors for ease of readability. The larger and slower MG58
experienced an additional squat of ∆S ≈ +0.2 m at the stern. The trim of the
MG58 changed only slightly at the stern from its original trim as a single ship in
the channel.

26.4.1.2. BAW field measurements


Field measurements of 12 transits on PPM containerships along the River Elbe were
made between April 2003 and June 2004. Meteorological conditions included very
calm to stormy (up to Beaufort Wind Scale 9). The shipping company Hapag Lloyd
Container Line GmbH (HLCL) supported eight journeys (transits) of Hamburg
Express Class ships (7506 TEU), and Yang Ming Marine Transport Corporation
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 741

Fig. 26.8. Laboratory measurement of the effect of head-on passing on bow and stern squat for
a large bulk carrier passing a PM containership in the River Elbe. The dark red curves represent
the single runs of the large bulk carrier; the light red curves the encounters with the containership.

Table 26.7. Characteristics of the PPM containerships in the BAW field measurements.

Vessel type Lpp (m) B (m) T (m) CB Capacity TEU

Hamburg Express Class 320.4 42.8 10.8–12.6 0.62–0.65 7506


5500 TEU YM Class 274.7 40 11.4–13.2 0.56–0.59 5500

(YM) supported four journeys on ships of the 5500 TEU Class with Tollerort Con-
tainer Terminals (TCT) acting as the intermediary.8 Table 26.7 presents selected
characteristics of the vessel types as well as the range of mean draft T and draft-
dependent CB during these journeys.
In Hamburg Harbor, the containerships were equipped with four autonomous
digital global positioning systems (DGPS) on the bow and the bridge and one data
collection system on the bridge. Vessel dynamics data were collected from Container
Terminal Altenwerder (CTA) or from TCT until just north of Scharhörn (about
120 km from Hamburg Harbor). The width of the channel in this section ranged
from 250 to 400 m. Current, temperature, and conductivity were measured by a very
fast, small ship at six cross-sections of the lower River Elbe just before the passing
encounters. Head-on and passing situations were recorded and documented.8 Vessel
movement, nautical maneuvers, local squat, trim, heel, and net maneuvering lane
were calculated using special water gage evaluations, precise DGPS measurements,
and calculations of virtual reference positions. Vessel data included propeller speed,
rudder position, etc. Maximum differences between water level interpolation and
DGPS zero measurements of <1 cm were obtained.48
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

742 M. J. Briggs et al.

Fig. 26.9. Cumulative distribution of the increase in squat for 125 head-on passing encounters of
large PPM containerships (HLCL and YM) at the channel of the lower and outer River Elbe.

Squat measurement errors were estimated to be ∆S = ±0.05 m for UKC determi-


nation. Given the quality of the digital terrain model from area and traffic soundings,
a precision of ∆UKC < ±0.2 m was estimated. Precisions of ∆VS = ±0.08 kt for
ship velocity (over ground) and ∆Φ = ±0.07 deg for ship heel were also estimated.48
During the seven-hour transits of the 12 ships from Hamburg to the sea, 125
head-on passing encounters were recorded. Figure 26.9 shows the increase in bow
(blue diamond) and stern (red square) squat due to these head-on passing encounters
and the corresponding cumulative distribution curve. The increase in squat was
about the same at both the bow and stern. For this limited set of large containerships
in the River Elbe, the maximum increase in squat was 0.44 m; 50% of the cases
experienced bow or stern squat less than 0.16 m, while 90% were less than 0.33 m.

26.4.1.3. FHR laboratory experiments


The FHR conducted comprehensive laboratory studies of head-on passing
encounters to improve the quality of their ship simulator.42 Figure 26.10 illustrates
the effect on the squat of a containership (LOA = 291.3 m, B = 40.3 m, T = 13.5 m)
sailing at a forward speed of 12 kt, caused by a head-on passing encounter with a
bulk carrier (LOA = 310.6 m, B = 37.8 m, and T = 13.5 m). The lateral distance
dy between the two centerlines was 114.5 m and the water depth h was 17.1 m.
The triangles indicate the sinkage fore and aft of the containership as a function of
the relative longitudinal position of both vessels if the bulk carrier approaches at
a speed of 8 kt, while the squares refer to an approach speed of 12 kt. The abscissa
takes values of −1 and +1, respectively, when the bows and the sterns are located
at the same longitudinal position. When the two bows meet, the ship’s bow sinkage
increases, whereas the stern is lifted, resulting in trim by the bow. The trim changes
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 743

RELATIVE LONGITUDINAL POSITION (-)


-2 -1.5 -1 -0.5 0 0.5 1 1.5 2
0

0.2

0.4
SINKAGE (m)

0.6

0.8

1
sinkage aft - VT = 12 knots
sinkage fore - VT = 12 knots
1.2 sinkage aft - VT = 8 knots
sinkage fore - VT = 8 knots
1.4

Fig. 26.10. Effect of passing encounter on ship bow and stern squat as a function of ship speed
in FHR tow tank for containership and bulk carrier.49

sign when the midship sections of both ships are at the same position. During the
second part of the meeting, the sinkage aft is increased while the bow is lifted. In
the given examples, the sinkage aft of the containership increases from an initial
value of 0.6 to about 0.9 m, if the bulk carrier has a speed of 8 kt and to about 1.2 m
when both ships have a speed of 12 kt. This corresponds to an increase in squat of
50% for the 8-kt case and 100% for the 12-kt case.

26.4.2. Overtaking ship maneuvers


26.4.2.1. BAW laboratory experiments
The squat interaction between overtaking Feeder (VG3) and General Cargo (VG4)
vessels (traveling in the same direction) were simulated in a 1:33.3 scale laboratory
model. A schematized trapezoidal cross-section of a portion of the western Kiel
Canal (approximate length of 100 km, channel width of 70–90 m, and depth of 11 m)
was modeled. Maximum overtaking squat values were measured only during the
time when both ships were aligned parallel to each other.
The lateral passing distance during the time when the ships were parallel was
54 m (between course lines). The two ships in this experiment had the properties
listed in Table 26.8.

Table 26.8. BAW model ships during overtaking maneuvers.

Code Description Lpp (m) B (m) T (m) CB

VG3 Feeder Containership 158 23 7.5 0.66


VG4 General Cargo ship 127 19 6.1 0.725
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

744 M. J. Briggs et al.

Fig. 26.11. Laboratory measurements of the effect of overtaking on bow and stern squat for
General Cargo (VG3) and Feeder containership (VG4) at the western Kiel Canal. The bow and
stern squat values for the VG3 are shown in red, and the VG4 are shown in blue.

Figure 26.11 is similar to Figs. 26.7 and 26.8 for head-on passing ships. It
shows bow and stern squat for both ships as single ships and during the over-
taking maneuver. Squat during overtaking is larger than for single ships. The
increase in stern squat for the VG3 was ∆S ≈ 0.6 m and ∆S ≈ 0.8 m for the
VG4 at a speed of Vk = 8.1 kt (15 km/h). This increase in squat is caused by
the effect of the additional hydrodynamic mass and channel blockage of each ship.
Since both ships experienced a common speed-dependent long wave, they had
the same order of magnitude of total stern squat Ss = 1.0 m at Vk = 8.1 kt
(15 km/h) (light blue and light red curves at left side of Fig. 26.11). The shorter
VG3 squatted with even keel in the long wave of the larger VG4 (light red
curves).

26.4.2.2. FHR laboratory experiments


Results from the FHR laboratory experiments on overtaking maneuvers are shown
in Fig. 26.12. It shows sinkage fore and aft of the same containership from the ship
passing experiments, sailing at a speed of 12 kt, while overtaking a bulk carrier
(LOA = 301.5 m, B = 46.7 m, and T = 15.5 m) sailing at 8 kt. The water depth h
was 18.6 m. Three lateral distances dy between centerlines were investigated: 84, 124,
and 205 m. Squat increased up to 0.3 m as the lateral distance decreased between
vessels during these overtaking experiments. This is equivalent to an increase in
ship squat of over 40%.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 745

Fig. 26.12. Effect of overtaking maneuver on bow and stern squat as a function of lateral distance
between ship centerlines for a containership and bulk carrier in the FHR tow tank.

26.4.3. Wide stern-transom effects


The BAW collected field measurements of bow and stern squat for the 12 large
containerships (HLCL and YM) during transits in the River Elbe from Hamburg to
the North Sea. Flow and density conditions covering the entire channel navigation
were obtained for nine of these runs. Fig. 26.13 shows the maximum bow squat
(HLCL = squares; YM = triangles) for these nine transits along the River Elbe.
They represent ranges of vessel types, channel configurations, and UKC for the two
shipping companies.
In general, the YM vessels experience a larger bow squat due to the design
of the hull with a wider transom-stern. Above a speed of 11 kt, the ship starts
to squat and trim strongly. However, once the transom-stern submerges below a
draft of 12 m, the ship experiences greater buoyancy, which causes it to trim by
the bow. This produces larger bow squat then ships without the wider transom-
stern. The HLCL ships exhibited a much weaker trim, as the narrower stern does
not immerse even with larger drafts. The buoyancy remains approximately equally
distributed along the hull compared to the wider transom-stern of the YM ships.
Therefore, in extremely shallow water, the trim behavior and the deepest point
of a vessel (here the bow squat) clearly depends on the overall design of the
underwater hull and especially on the buoyancy distribution in the longitudinal
direction. This result indicates that, for these wider transom-stern ships, the use
of the CB may not be as reliable an indicator of squat as has traditionally been
observed.8
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

746 M. J. Briggs et al.

v
vSS c.W.
d.W. [kn]
[Kn]
0 2 4 6 8 10 12 14 16 18 20
0,00

0,50

1,00

1,50 BERLIN EXPRESS 22.06.03


SHANGHAI EXPRESS 01.06.03
SHANGHAI EXPRESS 27.07.03
HAMBURG EXPRESS 15.06.03
BERLIN EXPRESS 01.02.04
2,00
MING COSMOS 22.02.04
MING PLUM 28.03.04
MING COSMOS 17.04.04
SMAX [m] MING COSMOS 12.06.04
2,50

Fig. 26.13. Field measurements of bow squat for nine container vessels of HLCL (blue/light blue
squares) and YM (red/yellow triangles) in the channel of the lower and outer River Elbe.8

26.4.4. Vertical variations in the channel


In real channels, the bathymetry is not constant, especially where the entrance
channel meets the offshore contours or enters more sheltered waters. Some channels
are characterized by undulating ripples along the channel bottom that can have
significant vertical rise above the bottom. An abrupt change in depth or sill due to
dredging can induce a significant transient squat that can be critical if the ship is
entering at deep water speeds. There has been little new research on these effects,
but the designer should be aware of their potential impact. The BAW has conducted
laboratory experiments on the effects of these ripples on ship squat. These results
should be available in the future, but additional research is recommended.

26.4.5. Ship offset and drift angle


Ships in the PIANC formulas are assumed to be sailing on the centerline of the
channel. When ships are offset from the centerline, they experience increased squat
because the hydrodynamic pressure is affected by the bank. The National Ports
Council50 showed that squat increases as the UKC and distance D between the ship
and the toe of the bank decrease relative to beam B. Squat increased in a restricted
channel from 16% to 47% for 1.1 ≤ h/T ≤ 1.2, 0.5B ≤ D ≤ B, and CB from 0.70 to
0.85. Squat increased even more in a canal due to the larger bank effect. The bank
effect became insignificant for D > 3B.
Similarly, a ship with a drift angle to the channel centerline experiences increased
water flow past the hull due to the increased blockage factor and a smaller gap
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 747

between the ship and the channel. The ship acts as a lifting surface as it moves
asymmetrically through the water. Drift angles are usually the result of trying to
compensate for large wind forces, especially on containerships.

26.4.5.1. Delft numerical model


The Delft University of Technology (Delft) has recently completed a limited set of
numerical modeling of ship squat for ships sailing with an offset and drift angle to
the channel centerline.47
A panel method was used in the tests for a 6500 TEU Post-Panamax container
ship with one draft but a range of offsets and drift angles. The potential flow model
includes inertial effects, but no viscosity that will cause vortices and increased squat
if included. The modeled ship had Lpp = 302 m, B = 42.9 m, T = 14 m, and
CB = 0.67. The canal had W = 300 m and h = 16 m. The UKC = 2 m with offsets
of 0 and ±20 m, and drift angles of 0, ±7.5 deg and ±15 deg.
They found that both offsets and drift angles increase squat, in a quadratic
manner. High drift angles should be avoided by using tugs if available. They recom-
mended additional research for a range of ships, channels, UKC, offsets, and drift
angles.

26.4.5.2. FHR laboratory experiments


The FHR has conducted towing tank experiments with containerships to study ship
offset and drift angle effects on squat. Figure 26.14 shows that moving the ship
laterally from the center of the channel (red) to the toe of the bank (blue) results
in an increase in squat of about 20%. At higher ship speeds, however, this effect
is amplified. A slight bow squat turns into a significant stern squat, and it is clear
that the ship sailing off-center will reach its critical speed much sooner.

Fig. 26.14. Influence of offset on squat of a containership (Lpp = 331.3 m, B = 42.8 m, and
T = 14.5 m) sailing at constant speed in a channel with h = 19.6 m. Scale 1:80 towing tank tests,
no propeller action. Open symbols: stern; closed symbols: bow.49
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

748 M. J. Briggs et al.

Fig. 26.15. Influence of drift angle on squat of a containership at constant speed in a rectangular
channel of 565-m width with h = 16 m. Ship test specifications same as Fig. 26.14.49

Figure 26.15 shows the bow and stern sinkage (squat) of a containership as a
function of speed for several drift angles (0, 5, and 10 deg). The bow sinkage increases
significantly with the drift angle, whereas the stern sinkage decreases slightly.

26.5. Muddy Navigation Areas

26.5.1. Governing effects


The presence of a fluid mud layer on the bottom of a channel has a significant
influence on ship behavior in general, and sinkage and trim in particular. Two effects
play a dominant role:

(1) The pressure field around the moving hull causes undulations of the water–
mud interface that themselves modify the distribution of vertical forces over the
length of the ship and, therefore, sinkage and trim.
(2) If ship’s keel penetrates into the mud layer, the hydrostatic (buoyancy) force
acting on the submerged hull increases due to the higher density of the mud.

The interface deformation is a function of many parameters, such as ship speed,


layer thickness, mud density and rheology, and (initial) UKC referred to the mud–
water interface. Contact between the ship’s keel and the mud layer depends mainly
on the UKC, but is also influenced by the interface undulations and the ship’s
sinkage. As a result, both effects are not independent. A general description of the
vertical interface motions on squat is presented. Most of the information available
on this subject is based on experimental research, mostly at model scale.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 749

26.5.2. Mud-water interface undulations


A ship navigating above fluid mud layers will cause vertical interface motions
(internal waves, internal undulations) that are influenced by the ship’s forward speed
(Fig. 26.16):

• At very low speed (first speed range), the interface remains practically
undisturbed.
• At intermediate speed (second speed range), an interface sinkage is observed under
the ship’s bow if the fluid mud layer is relatively thick. At a certain time, an
internal hydraulic jump, perpendicular to the ship’s longitudinal axis, is observed.
The front of this internal jump moves aft with increasing speed.
• At higher speeds, the internal or interface jump occurs behind the stern (third
speed range).

It can be shown by means of a simplified theory that the critical speed separating
the second and third speed ranges is a function of the mud ρ2 to water density ρ1
ratio and the water depth h1 (Fig. 26.17).
  
8 ρ1
Ucrit = gh1 1 − (1 − S1 )3 (26.30)
27 ρ2

where S1 = AS /Ac1 is similar to blockage factor S except that the Ac1 is the cross-
sectional area of the channel to the top of the mud layer. This equation is based on
ideal fluid assumptions, and appears to underestimate the critical speed for mud
layers of higher viscosity.

Fig. 26.16. Mud–water interface undulations for second speed range (top) and third speed range
(bottom).51
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

750 M. J. Briggs et al.

10

8 h1 = 25 m
h1 = 20 m
Ucrit (knots)

6 h1 = 15 m

4 h1 = 10 m

0
1 1.05 1.1 1.15 1.2 1.25 1.3
ρ2/ρ1

Fig. 26.17. Critical speed separating second and third speed ranges as a function of mud–water
density ρ2 /ρ1 ratio for different water depths h1 .51

The description above is typical for motions of the mud–water interface occurring
when a ship moves with a positive UKC above a fluid mud layer of low viscosity
(black water). In case of a negative UKC (i.e., when the keel penetrates the mud
layer), a second internal wave system, comparable to the Kelvin wave system in the
water–air interface, interferes with the hydraulic jump. This may result in either
an interface rising amidships or a double-peaked rising along the hull. Figure 26.18
illustrates the effect of speed (5 and 10 kt), UKC (−12% to +10%), and mud density
(1100–1250 kg/m3 ) on the interface undulation pattern.
Due to the vertical motion of the interface and the ship, contact between the
ship’s keel and the mud layer can occur even if, initially at rest, the UKC of the
ship is positive relative to the mud–water interface. Figure 26.19 shows the initial
UKC required to avoid contact between mud and keel as a function of Depth Froude
number Fn (speed) for different mud characteristics.

26.5.3. Effect of mud layers on sinkage and trim


The effect of the presence of a fluid mud layer covering the bottom on the ship’s
vertical motions is closely related to the interface deformation. If no contact between
the ship’s keel and the mud layer occurs [Figs. 26.20(a) and 26.20(c)], a rising
interface yields an increased velocity of the ship relative to the water and, as a
result, a pressure drop and a local water depression. A mud–water interface sinkage,
on the other hand, leads to a local decrease of the relative velocity and an increased
pressure, at least compared to the solid bottom case. In case of contact between
keel and a rising mud interface (Fig. 26.20(b)), the velocity of the mud relative to
the ship’s surface decreases. Contact with a lowered interface with negative UKC
(Fig. 26.20(d)) leads to an increased relative fluid velocity, with associated local
pressure fluctuations acting on the ship’s keel.
Figure 26.21 illustrates the effect of the presence of a mud layer on the sinkage
and trim of a containership for the case in which the initial UKC is sufficiently large
so that the interface undulations do not cause any contact between the keel and the
mud layer. The sinkage for a ship sailing in a muddy bottom condition is decreased
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 751

bottom (m)

-5

-4.5
solid bottom

-4
above solid

-3.5

-3
above

-2.5
position

-2
position

Layer thickness: 3.0 m -12%


interfaceinterface

-1.5

Density: 1100 kg/m³ -7%


-1 UKC to interface:
+4%
-0.5
Ship’s speed: 5 knots +10%
0
-1 -0.5 0 0.5 1

Interface motions
bottom (m)

-5

-4.5
solid bottom

-4
abovesolid

-3.5

-3
above

-2.5
position

-2
position

Layer thickness: 3.0 m -12%


interfaceinterface

-1.5

Density: 1100 kg/m³ -7%


-1
UKC to interface:
+4%
-0.5
Ship’s speed: 10 knots +10%
0
-1 -0.5 0 0.5 1

-5
(m) (m)

-4.5
bottom

-4
solid bottom

-3.5
abovesolid

-3
above

-2.5
position

-2
position

-1.5 Layer thickness: 3 3.0


.0 m 1100
Interfaceinterface

1150
-1 UKC to interface: 10 % Density (kg/m³): 1180
-0.5
1210
Ship speed: 5 5 knots 1250
0
-1 -0.5 0 0.5 1
(m)

-5
bottom

-4.5
(m)

-4
solidbottom

-3.5
above solid

-3
above

-2.5
position

-2
position

-1.5 Layer thickness: 3 3.0


.0 m 1100
Interfaceinterface

1150
-1 UKC to interface: 10 % Density (kg/m³): 1180
-0.5
1210
Ship's speed: 10 10 knots 1250
0
-1 -0.5 0 0.5 1

Fig. 26.18. Influence of speed, UKC, and mud density on undulations of the interface.52
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

752 M. J. Briggs et al.

(a)

(b)

Fig. 26.19. Critical UKC h1,crit /T for different mud layers with stopped propeller for model
ship D, where mud layer thickness to draft ratio h2 /T (a) 0.11 and (b) 0.22 as a function of Depth
Froude Number Fn .52

relative to the condition in which the mud layer is replaced by a solid bottom. This
is because the ship can “feel” the hard bottom more than the softer, less dense,
mud layer. If the mud layer is replaced by water (normal conditions without a mud
layer), however, the sinkage would decrease relative to the condition with the mud
layer. However, this does not take into account the effect of extra buoyancy (i.e.,
mud is denser than water), but this is only important in very dense mud layers
and/or important penetration. In general, the influence on trim is more important
than sinkage since the mud layer causes the ship to be dynamically trimmed by
the stern over its complete speed range. Thus, the effect of mud layers on average
sinkage is only marginal as trim is much more important.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 753

Fig. 26.20. Effect of mud layers on sinkage and trim (a) no interface contact, (b) contact with
mud interface, (c) no contact with interface, and (d) negative UKC. The blue line represents the
water surface, the brown line the mud layer interface, and black, the solid bottom.53

The effect of the decrease of UKC is shown in Fig. 26.22. In a range of small
positive to negative UKC, the trim is mostly affected in a moderate speed range
(second speed range, as defined above). A large negative UKC (keel into the bottom
mud–water interface) causes trim by the stern in the complete speed range. The
effect of mud on the average sinkage is less important, but the combination of trim
and sinkage results in an increase of the sinkage aft in some conditions.
Figures 26.21 and 26.22 are valid for slender ships (CB < 0.7) that tend to trim
by the stern above a solid bottom. Full-formed ships, on the other hand, usually
trim by the bow. In muddy navigation areas, such vessels will experience a reduced
trim by the bow — or even trim by the stern — when they have sufficient UKC in
the second speed range. In the third speed range, this effect will be reduced again.
Figure 26.23 shows this effect of midships sinkage and trim as a function of UKC
for a full-form trailing suction hopper dredge.

26.6. Numerical Models

Many different numerical methods can be used to calculate the ship squat. Their
only common point is that they calculate the velocity components and the pressure
of the flow surrounding the ship. Depending on whether the fluid is modeled as
viscous, a potential velocity function can be used or a more sophisticated flow model
has to be applied. Some models are based on slender body theory, whereas others
use the boundary elements method (BEM) or the finite element method (FEM).
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

754 M. J. Briggs et al.

(a) 0.6
0.5
sinkage FP (m)
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10 12
ship speed (knots)

(b) 0.6
0.5
sinkage AP (m)

0.4
0.3
0.2
0.1
0
0 2 4 6 8 10 12
ship speed (knots)
(c) 0.6
sinkage midships (m)

0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10 12
ship speed (knots)
(d) 0.2
0
trim (mm/m)

-0.2
-0.4
-0.6
-0.8
-1
0 2 4 6 8 10 12
V (knots)
solid 15% mud F mud G mud H
mud E mud C mud D solid 26%

Fig. 26.21. Sinkage (a) fore, (b) aft, (c) and midships, and (d) trim as a function of ship speed
for Containership D (LOA = 300 m, B = 40.3 m, h = 13.5 m) sailing above a mud layer of 1.5 m
thickness with 15% clearance referenced to mud-water interface (26% to solid bottom). Note the
legends are the same for all plots.52
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 755

Fig. 26.22. Sinkage (a) fore, (b) aft, (c) and midships, and (d) trim as a function of ship speed
for Containership D (LOA = 300 m, B = 40.3 m, h = 13.5 m) sailing above a mud layer of 3.0 m
thickness, ρB = 1,180 kg/m3 , ρD = 1,100 kg/m3 .52

26.6.1. Numerical modeling approaches


Most numerical models use the continuity equation to calculate the velocity compo-
nents and Bernoulli’s equation to obtain pressure that is integrated on the hull for
hydrodynamic forces. Then applying Archimedes’ principle, a vertical displacement
and a trim angle are calculated.

26.6.1.1. Slender body theory


Tuck11 established a mathematical expression for squat with a slender body theory.
The slender body theory assumes that the beam, draft, and water depth are very
small relative to ship length. This theory uses potential flow where the continuity
equation becomes Laplace’s equation. The flow is taken to be inviscid and incom-
pressible and is steady and irrotational. In restricted water, the problem is divided
into the inner and the outer problems, following a technique of matched asymp-
totic expansions to construct an approximate solution. The inner problem deals
with flow very close to the ship. The potential is only a function of y and z in
the Cartesian coordinate system. In the cross-flow sections, the potential function
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

756 M. J. Briggs et al.

SHIP SPEED (knots)


0 1 2 3 4 5 6
0
SOLID BOTTOM - UKC 10%
MUDDY BOTTOM - UKC 10%
MUDDY BOTTOM - UKC 0%
MEAN SINKAGE (% of draft)

1 MUDDY BOTTOM - UKC -10%

3 TRAILING SUCTION HOPPER DREDGER


LPP = 115.6 m; B = 23.0 m; T = 8.0 m

MUD LAYER: ρ2/ρ1 = 1.11 - h2/T = 0.175

4
SHIP SPEED (knots)
0 1 2 3 4 5 6
1
SOLID BOTTOM - UKC 10%
0.8 MUDDY BOTTOM - UKC 10%
MUDDY BOTTOM - UKC -4%
0.6

0.4
TRIM (mm/m)

0.2

-0.2

-0.4

-0.6 TRAILING SUCTION HOPPER DREDGER


LPP = 115.6 m; B = 23.0 m; T = 8.0 m
-0.8
MUD LAYER: ρ2/ρ1 = 1.225 - h2/T = 0.175
-1

Fig. 26.23. (a) Mean midships sinkage and (b) trim as a function of UKC for a full-form trailing
suction hopper dredger (115.6 × 23.0 × 8.0 m3 , scale 1/40) above a simulated mud layer (ρ2 /ρ1 =
1.22, h2 /T = 0.175). Positive trim is equivalent to increased stern squat.54

satisfies the two-dimensional (2D) Laplace’s equation with impermeable conditions


at the boundaries. The outer problem looks at the flow far from the ship where it
mainly depends on x and y directions.
Dand and Ferguson55 and Beck56 used slender body theory and found good
agreement with squat measurements for ratios of water depth to ship draft h/T > 2.
Dand used the cross-sectional strip theory of Korvin-Kroukovsky.57 The slender
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 757

body method works by vertical cross-sections of the flow, so it is also called the
one-dimensional (1D) theory of squat.
Gourlay58 extended the slender body theory of Tuck with the unsteady slender
body theory. This improvement allows one to consider a ship moving in a non-
uniform depth since the coordinate system is now earth-fixed, whereas it is ship-
fixed for classic numerical methods. The 1D system still uses vertical cross-sections
and decomposition into an inner and outer expansion. The pressure integration is
only made on the ship length based on the ship section B(x) at each x along the
hull. Resolution of the 1D equation is made with the finite difference method. Com-
parison with experimental results for soft squat situations (h/T > 4) showed good
agreement with numerical results. No tests were made for hard squat conditions
(i.e., shallow depths) where flow around the ship is affected.

26.6.1.2. Boundary element method


The BEM is really based on a particular numerical resolution. It is commonly
applied for wave-resistance calculations using Green’s function to calculate the
potential velocity function. Derivatives of the potential velocity function give the
velocity components in Cartesian coordinates. Bühring59 made a squat model called
fast boundary elements method (FBEM) based on this boundary element method.
The reliability of the model has to be verified, however, as no comparisons with ship
squat measurements was found.

26.6.1.3. Computational fluid dynamics models


A number of commercially available computational fluid dynamics (CFD) models
could be used for the prediction of squat. At the core of any CFD problem is a
computational grid or mesh where the solution is divided into thousands of elements.
These elements are usually 2D quadrilaterals or triangles; and three-dimensional
(3D) hexahedral, tetrahedral, or prisms. Mathematical equations are solved for each
element by the numerical model. For hydrodynamics the Navier–Stokes equations
(NSEs) can be solved to include viscosity and turbulence. The NSEs provide detailed
prediction (vortices) of the flow field, but require very thin meshes, high central
processing unit (CPU) time, and memory storage. Its resolution is also quite difficult
with numerical instabilities. Examples of commercial CFD models include Fluent
and Fidap.
Nowadays, CFD models can solve 3D problems, such as ship squat, but the
computation domain has to be relatively narrow using NSE. To extend the width
of the computation domain, some models solve the problem by zones. Far from the
ship, the model solves a potential function with a nonviscous fluid and, in the vicinity
of the ship, the model solves using the NSEs. The advantage of the potential flow
solution is that it requires low CPU time and less memory storage. The boundary
conditions for the NSE model are extracted from the potential flow solution. One
example of this kind of commercial model is ShipFlow.
In very restricted water, squat can substantially reduce the vertical cross-section
around the ship and can subsequently increase the flow velocity below the hull.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

758 M. J. Briggs et al.

According to Bernoulli’s principle, the pressure will decrease which will make the
ship sink more. Numerical models have to take into account this “over squat” to
precisely calculate ship squat in all channel configurations. So when a first squat
result has been found, the model has to check that this squat is not disturbing
the hydrodynamics in such a way that squat could increase more. This checking is
important to ensure a reliable result from the numerical model. As these commercial
numerical models do not perform squat checking, they may not be very efficient in
restricted water. The user has to be very careful and take the result with reservations
since the numerical model could in these conditions underestimate ship squat.

26.6.2. New modeling system to predict ship squat


In an unrestricted channel (i.e., open sea or large channel), “over squat” is negligible,
and the previous numerical models work well unless the UKC is relatively small
(h/T < 1.1). Under a 3-m UKC or in a very restricted channel, the numerical
model has to check that the hydrodynamics are not being modified by the squat.
Some empirical formulas try to make such a correction with a restriction factor that
multiplies the squat calculated for unrestricted water. For instance, the K coefficient
in Barrass and Ks for Huuska are examples of these types of correction factors.
Such a modeling system with squat checking was developed by Debaillon.60 The
basic principle is to reproduce the physical process of ship squat using a numerical
model coupling. As the ship is moving, a return flow is generated around the hull.
This induced velocity reduces the pressure under the hull. The ship sinks until
pressure forces balance the ship weight. As the ship position changes, flow around
it might be different, and it has to be updated with a new cycle of hydrodynamic
and equilibrium computations. The modeling system (Fig. 26.24) is thus composed
of (a) a hydrodynamic model to calculate the flow around the hull, (b) an equi-
librium model to move the ship with balanced force and momentum equations,
and (c) a mesh updating model to take into account the ship and the free surface
displacements.

26.6.2.1. Coupling principle


As shown on Fig. 26.24, the system starts from the rest position of the ship with a
sinkage equal to its draft T . A 3D mesh of finite elements (tetrahedral) is constructed
with the ship features (i.e., Lpp , B, T , and CB ) and the fluid domain (i.e., h, channel
shape, boundary conditions, etc.). A first run of the model is done with null velocity
of the ship. The equilibrium model is then calibrated with the ship weight (Wb )
and the position of the center of gravity (XG , YG ), as all hull nodes must have no
displacements with the hydrodynamic model results. Once these ship features are
set up, the system is ready to start. A small ship velocity ∆V is imposed in the
hydrodynamic model, which gives hull pressure to the equilibrium model. The latter
displaces the hull, so the mesh has to be updated by the third model. The system
checks the hull displacement. If it is negligible, the ship velocity is increased by ∆V
or the same velocity is retained and a new cycle is begun. The system stops when the
velocity has reached the velocity specified by the user or if the ship has grounded.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 759

Position V= ∆ V Hull Hull node Fluid mesh


Hydrodynamic Equilibrium
at rest pressure Displacement updating

no
Displacements < ε

yes
V=V+ ∆ V

Fig. 26.24. Coupling for the ship squat numerical modeling system.

Fig. 26.25. Example of 3D mesh of a ship in a channel.

26.6.2.2. Finite element method


The hydrodynamic model is based on the FEM. It requires a numerical mesh that is
a subset of the liquid volume around the ship. An example of the 3D mesh, with the
imprint of the ship, is shown in Fig. 26.25. It is important to note that numerical
results depend on the resolution of the mesh. If it is not sufficiently refined, the
model results will not be as accurate as possible.
The hydrodynamic model solves Laplace’s equation in 3D to obtain the velocity
potential function. The moving body is fixed, and an incoming flow is imposed with
the same velocity in the opposite direction at the far upstream boundary. A matrix
system is assembled with Laplace’s equation for each tetrahedron and the different
boundary conditions. As the problem is linear, the potential function Ψ is solved
by a matrix inversion represented by

{Ψ} = [K]−1 {F } (26.31)

where K is a matrix of Laplace’s equation for each tetrahedral and F is a vector


composed of the boundary conditions on the triangles.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

760 M. J. Briggs et al.

26.6.2.3. Three numerical models


In Fig. 26.24, the three numerical models are represented by shaded boxes. The
first model calculates the velocity components and pressure at each node of the
mesh. The second receives pressures at all nodes of the hull. The corresponding
vertical force FPZ is obtained by integration of pressure over the hull. Squat ∆Sb is
estimated as

Wb − FPZ
∆Sb = α (26.32)
ρgSf

where Sf is the floating surface of the ship, and α (fixed at 0.9) is a relaxation
coefficient to limit the variation of squat in a calculation cycle for the mesh updating
model.
Pitching θ1 and rolling θ2 angles are estimated in the second model from the x
and y components of the moment equations:

Wb (xG − xP )
tan θ1 = α   (26.33)
ρg x∈Sf y∈Sf x(x − x0 )dx dy

Wb (yG − yP )
tan θ2 = α   . (26.34)
ρg x∈Sf y∈Sf y(y − y0 )dx dy

The equilibrium (second) model has to calculate the position of the center of vertical
thrust P (xP , yP ) by integration of the pressure over the hull, and O(x0 , y0 ) the
center of the floating surface Sf . All nodes of the ship are then vertically translated
by ∆z and rotated by θ1 around the y-axis and by θ2 around the x-axis.
The third model moves all nodes of the free surface per the results of Bernoulli’s
relation, and the hull node displacements in x and y directions. Then the inner
nodes of the mesh are moved proportionally according to boundary modifications
and the distance from those boundaries.
Since 2006 the University of Compiègne has been working to improve Debaillon’s
squat system. As in Gourlay’s model, the coordinate system will be earth-fixed to
allow ship passing or crossing and bridge pile crossings. The system will also be able
to take into account a nonuniform water depth along the channel.

26.6.3. Numerical modeling examples


Figures 26.4–26.6 showed comparisons of the PIANC empirical formulas with the
measured laboratory measurements. These figures also included comparisons with
the numerical model predictions for each example. These examples included BAW’s
PPM containership in an unrestricted channel, FHR’s tanker in restricted water,
and Tothil’s Canadian Laker in a canal. In general, the numerical model matched
the measured values from the laboratory measurements very well. Details of the
individual examples are given in the following sections.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 761

26.6.3.1. BAW PPM containership in an unrestricted channel


The mesh contains 57,602 nodes, 255,447 tetrahedra, and 58,458 triangles. A general
comparison between squat measurements, empirical formulas, and the numerical
model was shown in Fig. 26.4. The numerical modeling match is very good, with a
maximum error smaller than 0.12 m.

26.6.3.2. FHR “G” Tanker in restricted water, Condition “C”


The mesh contains 35,149 nodes, 158,022 tetrahedra, and 33,532 triangles.
Figure 26.5 showed the comparisons. The numerical model had to rotate the hull
(trim) around the center of gravity. The agreement with the measured values was
reasonable.

26.6.3.3. Tothil’s Canadian Laker in canal


The mesh contains 27,191 nodes, 118,513 tetrahedra, and 28,300 triangles. A general
comparison between squat measurements, empirical formulas, and the numerical
model was shown in Fig. 26.6. The numerical model experienced some numerical
instability problems (rotate the hull (trim) around the center of gravity), but still
gave good predictions.

26.7. Conclusions and Outlook

This chapter has focused on some of the latest advances in predicting ship squat
and its effect on underkeel clearance for channel design. Several of the more popular
PIANC empirical formulas were presented for Concept and Detailed Design phases.
In general, the simpler and more “user friendly” formulas were recommended for
the Concept Design phase, but this does not preclude them being used in the Detail
Design phase and vice versa. Ultimately, the designer wants the maximum squat
value possible (bow or stern) in the Concept Design phase and a more realistic value
in the Detail Design phase. All empirical formulas have certain constraints based
on the field and laboratory data used in their development. It is up to the user to
exercise Engineering Judgment when applying these formulas as they give a range
of squat values.
The PIANC formulas were developed for “idealized” channel and sailing con-
ditions for single ships. Recent research has been conducted to investigate more
“real-world” conditions for the latest generation of larger ships. The BAW and FHR
have conducted extensive laboratory and field investigations of head-on passing
and overtaking maneuvers, where squat is a function of ship speed and lateral sep-
aration distance. Their results indicate that squat can increase 50–100% relative
to a single ship. The BAW measured maximum additional squat during passing
encounters of 0.6 m in the laboratory and 0.4 m in the field. Similarly, the FHR
measured maximum additional squat of 0.3–0.6 m in the laboratory. For overtaking
maneuvers in the laboratory, the BAW recorded additional squat of 0.6–0.8 m and
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

762 M. J. Briggs et al.

the FHR measured 0.2–0.3 m. These additional squat values represent the effect of
these more realistic ship and channel interactions for a range of conditions.
The BAW noted that ships with wider “transom” sterns experience more bow
squat than most ships with more streamlined transoms due to the additional
buoyancy of the stern. They measured additional bow squat of 0.2–0.5 m as a
function of ship speed.
The Delft and FHR have conducted research on ships sailing with offsets and
drift angle from the channel centerline. Delft found that both offsets and drift angles
increase squat, in a quadratic manner. The FHR found that offsets can increase
squat by 20% and drift angles produce significant increase in bow squat and slight
decrease in stern squat.
The average sinkage of a vessel navigating in muddy channels is generally
reduced by the presence of mud layers. The dynamic trim is affected significantly
by the generation of interface undulations. For ships navigating above mud layers,
the maximum sinkage is comparable to or slightly less than the values occurring if
the mud layer were replaced by a solid bottom. Compared to the situation in which
the mud layer is not present (i.e., replaced by water), the muddy bottom interface
always increases the maximum sinkage, even in case of contact with the mud layer.
This means that the mud layer will increase sinkage even if the ship “plows” through
it. The maximum sinkage (bow or stern) always increases when the lower part of
the water column is replaced by fluid mud.
The designer and harbor pilots should be aware that all of these special influences
can increase squat. Sometimes, it is not economically feasible to design a channel
for all of these eventualities, but it is always possible to slow down as the conditions
warrant.
Numerical modeling of ship squat is just beginning to be developed. Histori-
cally, ship modeling has been concentrated in the areas of wave resistance models.
Squat modeling requires time domain models that are very computer intensive. The
increasing cost of dredging and the larger ships coming on line have motivated many
institutions around the world (such as USACE, BAW, FHR, CETMEF, and FMA)
to begin a more active development of ship squat models. The improvements in
computer speed and storage have made these types of models much more promising.
The CETMEF numerical model matched measured laboratory values very well in
the three examples presented in this chapter.
Numerical models will continue to be improved, but field measurements and
laboratory models will still be necessary to investigate the highly nonlinear dynamic
behavior of the newer and larger ships. They will respond differently than existing
ships and continued study of passing and overtaking, bank and bottom effects, and
sailing alignment will insure optimum and safe navigation design of channels and
waterways.

Acknowledgments

The authors wish to acknowledge the Headquarters, US Army Corps of Engineers,


Ghent University, Flanders Hydraulics Research, Federal Waterways Engineering
and Research Institute, Centre d’Etudes Techniques Maritimes Et Fluviales, and the
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 763

PIANC WG49 for authorizing publication of this chapter. Particular thanks goes to
Bryan Barrass, the WG49 Chair Mark McBride, original WG30 and WG49 member
Werner Dietze, and WG49 vertical subcommittee members Martin Boll, Hans Moes,
Terry O’Brien, and Kohei Ohtsu who assisted with the ship squat and UKC research.
Other members of the WG49 making contributions to this effort included Larry Cao,
Don Cockrill, Rink Groenveld, Jarmo Hartikainen, Jose Iribarren, Susumu Naruse,
Sahil Patel, Carlos Sanchidrian, Esa Sirkiä, and Jos Van Doorn.

References

1. J. N. Newman, Marine Hydrodynamics (MIT Press, Cambridge, Massachusetts, 1977),


pp. 329–386.
2. C. B. Barrass, Ship Squat Seminar, The Nautical Institute, Humberside Branch,
September 1995, pp. 21–33.
3. C. B. Barrass, Thirty-Two Years of Research into Ship Squat, 2nd Squat Workshop
2004, Elsfleth, Germany, 3–4 March 2004.
4. C. B. Barrass, Ship Squat and Interaction for Masters, Private report (2007), www.
ship-squat.com.
5. PIANC, Approach Channels: A Guide for Design, Final Report of the Joint PIANC-
IAPH Working Group II-30 in cooperation with IMPA and IALA, Supplement to
Bulletin No. 95 (1997).
6. PIANC, Horizontal and Vertical Dimensions of Fairways, Maritime Navigation Com-
mission Working Group 49 (MarCom WG 49), July 2005.
7. O. Huuska, On the Evaluation of Underkeel Clearances in Finnish Waterways, Helsinki
University of Technology, Ship Hydrodynamics Laboratory, Otaniemi, Report No. 9
(1976).
8. K. Uliczka and B. Kondziella, Dynamic response of very large containerships in
extremely shallow water, Proc. 31st PIANC Cong., Estoril, Spain (2006).
9. U.S. Army Corps of Engineers (USACE), Engineering and Design: Hydraulic
Design Guidance for Deep-draft Navigation Projects, Engineer Manual 1110-2-1613,
Headquarters, Washington, DC (2004).
10. M. Vantorre, Personal communication (2007).
11. E. O. Tuck, Shallow-water flows past slender bodies, JFM 26(Part 1), 81–95 (1966).
12. E. O. Tuck and P. J. Taylor, Shallow water problems in ship hydrodynamics, Proc.
8th Symp. Nav. Hydrod., Pasadena, CA (1970).
13. R. F. Beck, J. N. Newman and E. O. Tuck, Hydrodynamic forces on ships in dredged
channels, J. Ship Res. 19(3), 166–171 (1975).
14. J. P. Hooft, The behavior of a ship in head waves at restricted water depth, Int.
Shipbuild. Prog. 21(244), 367–378 (1974).
15. I. W. Dand, Squat Estimation: A Graphical Method for Full Form Ships, National
Physical Laboratory, Report No. TM 348 (1975).
16. N. E. Eryuzlu and R. Hausser, Experimental investigation into some aspects of large
vessel navigation in restricted waterways, Proc. Symp. Asp. Navi. Const. Waterway.
Incl. Harb. Ent. 2, 1–15 (1978).
17. K. Römisch and Empfehlungen zur Bemessung von Hafeneinfahrten, Wasserbauliche
Mitteilungen der Technischen Universität Dresden, Heft 1, 39–63 (1989).
18. A. Millward, A preliminary design method for the prediction of squat in shallow water,
Mar. Tech. 27(1), 10–19 (1990).
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

764 M. J. Briggs et al.

19. A. Millward, A comparison of the theoretical and empirical prediction of squat in


shallow water, Int. Shipbuilding Prog. 39(417), 69–78 (1992).
20. ICORELS (International Commission for the Reception of Large Ships), Report of
Working Group IV, PIANC Bulletin No. 35, Supplement (1980).
21. C. B. Barrass, The phenomenon of ship squat, Int. Shipbuilding Prog. 26(294), 44–47
(1979).
22. C. B. Barrass, Ship Squat — A Reply, The Naval Architect, November 1981,
pp. 268–272.
23. N. E. Eryuzlu, Y. L. Cao and F. D’Agnolo, Underkeel requirements for large vessels in
shallow waterways, Proc. 28th Int. Navi. Cong., PIANC, Paper S II-2, Sevilla, Spain
(1994), pp. 17–25.
24. V. Ankudinov, L. L. Daggett, J. C. Hewlett and B. K. Jakobsen, Prototype mea-
surement of ship sinkage in confined water, Proc. Int. Conf. Mar. Sim. Ship Maneuv.
(MARSIM 2000), Orlando, FL, 8–12 May 2002.
25. D. T. Stocks, L. L. Dagget and Y. Page, Maximization of Ship Draft in the St.
Lawrence Seaway Volume I: Squat Study, prepared for Transportation Development
Centre, Transport Canada, June 2002.
26. M. J. Briggs, Ship Squat Predictions for Ship/Tow Simulator, Coastal and Hydraulics
Engineering Technical Note CHETN-I-72, U.S. Army Engineer Research and Devel-
opment Center, Vicksburg, MS (2006), http://chl.wes.army.mil/library/publications/
chetn/.
27. C. B. Barrass, Ship Squat — A Guide for Masters, Private report (2002), www.ship-
squat.com.
28. Y. Yoshimura, Mathematical model for the maneuvering ship motion in shallow water,
J. Kansai Soc. Nav. Arch. Japan 200, 41–51 (1986).
29. FMA (Finnish Maritime Administration), The Channel Depth Practice in Finland,
Bulletin, Waterways Division, Helsinki, Finland, 12 July 2005.
30. E. Sirkia, Economical efficiency to be achieved with a regulatory change only with
consideration for navigational risks, PIANC Magazine 129, 23–34 (2007).
31. K. Uliczka, B. Kondziella and G. Flügge, Dynamisches Fahrverhalten sehr großer
Containerschiffe in seitlich begrenztem extremen Flachwasser, HANSA, 141, Jahrgang,
Nr. 1 (2004) (in German).
32. Overseas Coastal Area Development Institute of Japan, Technical Standards and
Commentaries for Port and Harbor Facilities in Japan (2002).
33. K. Ohtsu, Y. Yoshimura, M. Hirano, M. Tsugane and H. Takahashi, Design standard
for fairway in next generation, Asia Navigation Conf. 26 (2006).
34. K. Ohtsu, Personal communication (2007).
35. Canadian Coast Guard, Safe Waterways (A Users Guide to the Design, Mainte-
nance and Safe Use of Waterways), Part 1(a) Guidelines for the Safe Design of
Commercial Shipping Channels, Software User Manual Version 3.0, Waterways Devel-
opment Division, Fisheries and Oceans Canada, December 2001.
36. U. M. Guliev, On squat calculations for vessels going in shallow water and through
channels, PIANC Bulletin 1(7), 17–20 (1971).
37. Puertos Del Estado, Recommendations for Maritime Works (Spain) ROM 3.1-99:
Designing Maritime Configuration of Ports, Approach Channels and Floatation Areas,
CEDEX, Spain (1999).
38. C. J. Huval, Lock Approach Canal Surge and Tow Squat at Lock and Dam
17, Arkansas River Project; Mathematical Model Investigation, Technical Report
HL-80-17, U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS
(1980).
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 765

39. V. Balanin et al., Peculiarities of navigation on canals and restricted channels, orig-
inating hydraulic phenomena associated with them and their effect on the canal
bed; measurements preventing slope deterioration, Paper S.I-3, 24th Int. Nav. Cong.,
Leningrad, Russia (1977).
40. G. Flügge and K. Uliczka, Dynamisches Fahrverhalten und Wechselwirkungen mit der
Fahrrinnensohle von sehr großen Containerschiffen unter extremen Flachwasserbedin-
gungen, Proceedings HTG-Kongress 2001, Hamburg (2001) (in German).
41. G. Flügge and K. Uliczka, Dynamisches Fahrverhalten mit der Fahrinnensohle
von sehr grossen Containershiffen unter extremen Flachwasserbedingungen. Hansa
(2001).
42. M. Vantorre, E. Verzhbitskaya and E. Laforce, Model Test Based Formulations of
Ship-Ship Interaction Forces, Ship Technology Research/Schiffstechnik, Band 49, Heft,
3 August 2002.
43. J. T. Tothil, Ships in Restricted Channels, A Correlation of Model Tests, Field Mea-
surements and Theory, National Research Council of Canada Mechanical Engineering
Report MB264, January 1966.
44. I. W. Dand, Some Measurements in Interaction between Ship Models Passing on
Parallel Courses, NMI R108, August 1981.
45. K. Uliczka and G. Flügge, Squat-Untersuchungen für sehr große Post-Panamax-
Containerschiffe, HTG/STG-Sprechtag FA Seeschifffahrststraßen, Hafen und Schiff,
Hamburg (2001) (in German).
46. M. Vantorre, E. Laforce, G. Dumon and W. Wackenier, Development of a probabilistic
admittance policy for the flemish harbours, 30th PIANC Cong., Sydney, September
2002.
47. H. J. de Koning Gans and H. Boonstra, Squat effects of very large container ships with
drift in a harbor environment, MTEC2007 Conference, Singapore, 26–28 September
2007.
48. Ch. Maushake and S. Joswig, Messung von Squat, Trimm und Krängung sehr
großer Containerschiffe im Rahmen von Grundsatzuntersuchungen auf der Elbe,
Hydrographische Nachrichten Nr. 072, Deutsche Hydrographische Gesellschaft (2004)
(in German).
49. M. Vantorre and G. Dumon, Model test based requirements for the under keel
clearance in the access channels to the flemish harbours, 2nd Squat Workshop Aspects
of Under Keel Clearance in Analysis and Application, Elsfleth, March 2004.
50. National Ports Council, Ship Behavior in Ports and their Approaches — Part 2:
Additional Sinkage Caused by Sailing in the Proximity of Channel Bank, Research
Transport Headquarters, London, U.K. (1980).
51. M. Vantorre, Nautical Bottom Approach – Application to the Access to the Harbour of
Zeebrugge, HANSA — Schiffahrt — Schiffbau — Hafen, 138. Jahrgang, Nr. 6 (2001),
pp. 93–97.
52. G. Delefortrie, Maneuvering behavior of container vessels in muddy navigation areas,
Ph.D. thesis, Ghent University (2007).
53. K. Van Craenenbroeck, M. Vantorre and P. De Wolf, Navigation in Muddy
Areas: Establishing the Navigable Depth in the Port of Zeebrugge, Proceedings
CEDA/PIANC Conference 1991 (incorporating CEDA Dredging Days): Accessible
Harbours, Paper No. E4, Amsterdam (1991).
54. M. Vantorre, Systematische proevenreeksen met het zelfaangedreven schaalmodel van
een sleephopperzuiger boven een mengsel petroleum¬trichloorethaan als slibsimulatie-
materiaal experimentele waarnemingen en theoretische interpretaties. Rijksuniver-
siteit Gent/Waterbouwkundig Laboratorium Borgerhout. Gent/Antwerpen (1990).
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

766 M. J. Briggs et al.

55. I. W. Dand and A. M. Ferguson, The squat of hull ships in shallow water, Roy. Inst.
Nav. Arch. 115, 237–247 (1973).
56. R. F. Beck, Hydrodynamic forces caused by ship in confined waters, Proc. Am. Soc.
Civ. Eng. 107, ASCE, EM3, June 1981, pp. 523–546.
57. B. V. Korvin-Kroukovsky, Investigations of ship motions in regular waves, Trans.
SNAME 63, 386–435 (1955).
58. T. Gourlay, Mathematical and computational techniques for predicting the squat of
ships. Thesis of the University of Adelaide, Australia (2000).
59. H. Bühring, Prediction of Squat by Fast Boundary Elements, 2nd Squat Workshop,
Oldenburg, Germany (2004).
60. P. Debaillon, Système de modélisation de l’enfoncement dynamique des bateaux,
Thesis of University of Compiègne, France (2005).

26.A. Appendix

26.A.1. Least square coefficients for Huuska K1 versus S


A least squares polynomial fit of Huuska’s Fig. 26.3 curves for K1 as a function of
S was calculated according to the formula:

K 1 = a0 + a1 S + a2 S 2 + a3 S 3 . (26.A1)

Table 26.A1 lists the correlation coefficient R2 and the polynomial coefficients for
each of the hT /h curves. This allows one to program the value for K1 without having
to manually read a plot.

26.A.2. Römisch Kc for canals and restricted channels


In Sec. 26.3.3.3, we presented an equation for Römisch’s Kc for canals and restricted
channels as a function of the blockage factor S. The Kc given in Sec. 26.3.3.3 is
equivalent to

1.5
π Arc cos(1 − S)
KC = 2 cos + . (26.A2)
3 3

Table 26.A1. Huuska K1 versus S least square fit coefficients.

Polynomial coefficients

hT /h R2 a0 a1 a2 a3

0.2 0.9988 1.0 2.7704 214.87 −569.42


0.4 0.9983 1.0 8.0885 89.87 −214.88
0.6 0.9963 1.0 −1.9528 137.6 −347.93
0.8 0.9991 1.0 1.9453 45.325 −129.48
1.0 1.0000 1.0 0.0 0.0 0.0
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 767

Both equations are the explicit solution to



1.5
VCr 2 V2
√ = 1 − S + Cr (26.A3)
ghm 3 2ghm
which has the critical or limiting ship speed VCr on both sides of the equation.
The KC can also be defined several other ways as it is related to Schijf’s limiting
speed VL and Schijf’s limiting froude number FHL .9 Schijf’s limiting Froude Number
is defined as


VL π Arc cos
FHL = √ = 8 cos3 + (1 − S) = KC . (26.A4)
gh 3 3

Note that the right-hand side is equivalent to KC for canals and restricted channels,
the same as Eqs. (26.28) and (26.A2).
Finally, Briggs26 determined the formula for KC from a least square fit of
Römisch’s limited set of discrete data points in Table 26.2 with an R2 = 0.97. It
gives the same results as the other formulas for KC :

KC = 0.2472 ln(1/S) + 0.02411. (26.A5)

26.B. Appendix: Worked Example Problems

Several worked examples are contained in this appendix. They are the same
examples described in Sec. 26.3. The input ship and channel characteristics were
described in the text and listed in Tables 26.3–26.5.

26.B.1. Example 1: BAW’s Post-Panamax Containership


in unrestricted channel
26.B.1.1. Constraint check
The first step is to examine the constraints for this case as listed in Table 26.1 and
the input values from Table 26.3. The Fnh is given by
Vs 6.84
Fnh = √ =  = 0.50. (26.B1)
gh 9.81(19)

Since Fnh ≤ 0.70, it is acceptable for all methods. The full form CB = 0.84 is
acceptable for all methods, but slightly exceeds Huuska’s upper limit of CB < 0.8.
The ratio of B/T = 50/17 = 2.94, slightly exceeds Eryuzlu and Römisch’s criterion.
Next, h/T = 19/17 = 1.12 is acceptable, but minimal UKC. The ratio Lpp /B =
400/50 = 8.0 is larger than Eryuzlu’s upper limit of 6.8. Finally, Lpp /T = 400/17 =
23.53 is slightly larger than the upper limits of Huuska and Römisch. Bottom line:
probably acceptable to use the different formulas for this case, but remember that
this is a much larger vessel than the criteria for which most of these formulas were
developed.
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

768 M. J. Briggs et al.

26.B.1.2. ICORELS application


This is a logical first choice for an unrestricted channel as it is one of the first for-
mulas developed as a guide. There are two steps involved in the ICORELS estimate
of bow squat Sb .
The first step is to calculate the ship’s displacement volume Λ using values from
Table 26.3

∇ = CB Lpp BT = 0.84(400)(50)(17) = 285, 600 m3. (26.B2)

The second step is to substitute this value into the equation for bow squat Sb
from Table 26.3 and Sec. 26.B.1.1.

∇ F2 285, 600 (0.50)2


Sb = CS 2
 nh = 2.4 2
 = 1.24 m. (26.B3)
2
Lpp 1 − Fnh (400) 1 − (0.50)2

This value is too large compared to the measured value of Sb = 0.70 m. Of course,
some institutions use a smaller value of the constant CS . Thus, this value could be
reduced by using a smaller CS , but it would need to be of the order of 1.4 which is
much smaller than commonly recommended.

26.B.1.3. Yoshimura application


The Yoshimura formula is a good Concept Design application for this case. It is
very straightforward and easy to apply. Substituting values from Table 26.3 and
Sec. 26.B.1.1 into Yoshimura’s equation for Sb



3 2
1 0.84 1 0.84 (6.84)
Sb = 0.7 + 1.5 + 15 = 1.10 m. (26.B4)
1.12 8.0 1.12 8.0 9.81

Thus, this Concept Design application is on the high side, but at least it is a
conservative estimate relative to the measured Sb = 0.70 m.

26.B.1.4. Eryuzlu application


This is an example of a Detailed Design application using Eryuzlu’s formula. The
steps are described in the following paragraphs.
The first step is to calculate the correction factor for channel width Kb . Since
this is an unrestricted channel with no side boundaries, a value of Kb = 1 was
selected since it is assumed that Eryuzlu meant the first part of his equation to be
for restricted channels only.
The second step is to substitute values from Table 26.3 and Sec. 26.B.1.1 into
Eryuzlu’s equation for bow squat Sb
 2.289
(19)2 6.84
Sb = 0.298  (1.12)−2.972 (1) = 1.06 m. (26.B5)
17 9.81(17)
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 769

This value is also larger than the measured value of Sb = 0.70 m, but it is conser-
vative and similar to Yoshimura’s value.

26.B.1.5. Römisch application


This is an example of a Detailed Design application using Römisch’s formula. It is
probably the most complicated of all the PIANC empirical formulas. The five steps
are described below using the values from Table 26.3 and Sec. 26.B.1.1.
The first step is to calculate the coefficients required to estimate the critical
speed Vcr in an unrestricted channel. They are phase speed or celerity C and the
correction factor KU .
 
C = gh = 9.81(19) = 13.65 m/s (26.B6)


0.125
h Lpp
KU = 0.58 = 0.58[(1.12)(8.0)]0.125 = 0.76 (26.B7)
T B
Vcr = CKU = 13.65(0.76) = 10.41 m/s. (26.B8)
The second step is to calculate the correction factor for ship speed CV

2
4
6.84 6.84
CV = 8 − 0.5 + 0.0625 = 0.22. (26.B9)
10.41 10.41

The third step is the calculation of the correction factor for ship shape CF

2
2
10CB 10(0.84)
CF = = = 1.10. (26.B10)
Lpp /B 8.0
The fourth step is the calculation of the correction factor for squat at the critical
speed K∆T
 √
K∆T = 0.155 h/T = 0.155 1.12 = 0.16. (26.B11)
The last step is to substitute these values into the equation for bow squat Sb
Sb = CV CF K∆T T = 0.22(1.10)(0.16)(17) = 0.67 m. (26.B12)
This value, although slightly small, is in excellent agreement with the BAW mea-
sured value of Sb = 0.70 m.

26.B.2. Example 2: FHR “G” Tanker in a canal with vertical


sides, Condition C
26.B.2.1. Constraint check
As in the first example, the first step is to determine the constraints for this case
using Table 26.1 and input values from Table 26.4. The Fnh is given by
Vs 5.14
Fnh = √ =  = 0.43. (26.B13)
gh 9.81(14.5)
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

770 M. J. Briggs et al.

Since Fnh ≤ 0.70, it is acceptable for all methods. The full form CB = 0.85 is
acceptable for all the methods, but is greater than Huuska’s upper limit of CB < 0.8.
The ratio of B/T = 33/13 = 2.54 is good for all methods. Next, h/T = 14.5/13 =
1.12 is acceptable. The ratio Lpp /B = 180/33 = 5.45 is acceptable, although a little
low for most methods. Finally, Lpp /T = 180/13 = 13.85 is on the low side, but
acceptable. Bottom line: probably acceptable to use the different formulas for this
case, but remember that some of the original constraints are exceeded.

26.B.2.2. Barrass application


The Barrass formula is a good Concept Design formula as it is relatively easy to
apply and gives reasonable estimates.
The first step is to calculate the mid-ship cross-sectional area AS using
Table 26.4.

AS = 0.98 B T = 0.98(33)(13) = 420.42 m2. (26.B14)

The second step is to calculate the channel cross-sectional area AC . The zero
value of slope n is to account for the vertical sides of the flume.

AC = W h + nh2 = 350(14.5) + 0.0(14.5)2 = 5075 m2. (26.B15)

The third step is to calculate the blockage factor S given by

AS 420.42
S= = = 0.083. (26.B16)
AC 5075

This relatively small value of S indicates that the channel should be considered as
an unrestricted channel for Barrass application.
The fourth step is to calculate the correction coefficient K given by

K = 5.74S 0.76 = 5.74(0.083)0.76 = 0.87 => 1.00. (26.B17)

The value of K = 0.87 is replaced by K = 1.00 since this is the minimum value
that Barrass intended for relatively wide channels.
The last step is to substitute the values above into the equation for Sb

KCB Vk2 1.0(0.85)(10)2


SMax = = = 0.85 m. (26.B18)
100 100
This value is a little low compared to the measured value of Sb = 1.18 m, but is a
good first estimate.

26.B.2.3. Yoshimura application


The first step in the Yoshimura application is to calculate the equivalent ship speed
Ve because it is a canal and not an unrestricted channel. The blockage factor S has
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 771

already been calculated for the Barrass application above. Therefore, Ve is given by

Vs 5.14
Ve = = = 5.61 m/s. (26.B19)
(1 − S) (1 − 0.083)

The second step is to substitute values from above, Table 26.4, and Sec. 26.B.2.1
into the Yoshimura equation for Sb



3
1 0.85 1 0.85 (5.61) 2
Sb = 0.7 + 1.5 + 15 = 1.19 m. (26.B20)
1.12 5.45 1.12 5.45 9.81

Thus, Yoshimura’s prediction is an excellent match to the measured Sb = 1.18 m.

26.B.2.4. Huuska application


The last application will be using Huuska’s Detailed Design formula. Römisch’s
formula gives very good agreement, but the Huuska formula will be demonstrated
here since Römisch’s was illustrated in the previous example. The Huuska formula
is more complicated to use than some, but not as difficult as the Römisch. It is
very similar to the ICORELS formula, but includes a correction factor for restricted
channels and canals.
The first step is to calculate the correction factor K1 that is used in the corrected
blockage factor s1 . Since hT = h for this case, which is similar to a canal, one can
use the graph from Fig. 26.3 or the least square coefficients in Appendix 26.A to
get the value of K1 = 1.0.
The second step is to calculate the corrected blockage factor s1

S 0.083
s1 = = = 0.083. (26.B21)
K1 1.0

The third step is to calculate the correction factor for channel width Ks , which
depends on the value of s1 . The first equation for Ks is used since s1 > 0.03.

Ks = 7.45s1 + 0.76 = 7.45(0.083) + 0.76 = 1.38. (26.B22)

The fourth step is to calculate the ship’s displacement volume Λ

∇ = CB Lpp BT = 0.85(180)(33)(13) = 65, 637 m3. (26.B23)

The last step is to substitute these values above into Huuska’s equation for Sb

∇ F2 65, 637 (0.43)2


Sb = CS 2
 nh K s = 2.4 2
 (1.38) = 1.38 m. (26.B24)
2
Lpp 1 − Fnh (180) 1 − (0.43)2
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

772 M. J. Briggs et al.

This value is a little large compared to the measured value. The values of both
Yoshimura and Römisch are Sb = 1.19 m, which are excellent matches to the mea-
sured value of Sb = 1.18 m.

26.B.3. Example 3: Tothil’s Canadian Laker in a canal


26.B.3.1. Constraint check
The first step is to determine the constraints for this case by comparing calculated
values with those in Table 26.1. Using values in Table 26.5, the Fnh is given by

Vs 3.59
Fnh = √ =  = 0.38. (26.B25)
gh 9.81(9.33)

Since Fnh ≤ 0.70, it is acceptable for all methods. The full form CB = 0.86 is
acceptable for all the methods, although it is slightly larger than the upper CB limit
of both Barrass and Huuska. The ratio of B/T = 22.9/7.77 = 2.95, slightly exceeds
Eryuzlu’s and Römisch’s upper limits. Next, h/T = 9.33/7.77 = 1.2 is acceptable.
The ratio Lpp /B = 215.6/22.9 = 9.41 is larger than most upper limits. Finally,
Lpp /T = 215.6/7.77 = 27.75 is slightly larger than all upper limits. Bottom line:
probably acceptable to use the different formulas for this case, but remember that
some of the original constraints are exceeded.

26.B.3.2. Barrass application


The Barrass is a good Concept Design formula for the canal example. It is one of
the best fits to the measured data, but is too low for higher ship speeds above 6.4 kt.
It is simple to apply, but there are several steps to follow. Again, the values are
substituted from Table 26.5 and Sec. 26.B.3.1.
The first step is to calculate the mid-ship cross-sectional area AS

AS = 0.98 B T = 0.98(22.9)(7.77) = 174.37 m2. (26.B26)

The second step is to calculate the channel cross-sectional area AC . The slope
n (i.e., run/rise = 1.8 = (105.9 − 72.3)/(2 ∗ 9.33)) is equivalent to an angle of
θ = 29 deg (i.e., θ = arctan(1/n)).

AC = W h + nh2 = 72.3(9.33) + 1.8(9.33)2 = 831.25 m2. (26.B27)

The third step is to calculate the blockage factor S given by

AS 174.37
S= = = 0.21. (26.B28)
AC 831.25

The fourth step is to calculate the correction coefficient K given by

K = 5.74 S 0.76 = 5.74(0.21)0.76 = 1.75. (26.B29)


August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

Prediction of Squat for Underkeel Clearance 773

The last step is to substitute the values above into the equation for Sb

KCB Vk2 1.75(0.86)(6.98)2


SMax = = = 0.73 m. (26.B30)
100 100

This value is a little low compared to the measured value of Sb = 0.93 m, but is a
good first estimate.

26.B.3.3. Römisch application


The Römisch formula is a good Detail Design formula, and this application will
illustrate how it is used for a canal. The Römisch results are slightly low, but
the trend follows the measured values reasonably well. The main difference with the
unrestricted case from before is that the correction factor KC replaces KU for the
canal application. Input values from Table 26.5 and Secs. 26.B.3.1 and 26.B.3.2 are
used in this application.
The first step is to calculate the coefficients required to estimate the critical
speed VCr in a canal. They are phase speed or celerity Cm that is based on the mean
water depth hm and the correction factor KC . Although Table 26.2 could be used,
the formula for KC [from Eq. (26.28)] is used in this example.

AC 831.25
hm = = = 7.85 m (26.B31)
WTop 105.9
 
Cm = ghm = 9.81(7.85) = 8.78 m/s (26.B32)

1.5
Arc sin(1 − S)
Kc = 2 sin
3

1.5
Arc sin(1 − 0.21)
= 2 sin = 0.46 (26.B33)
3
Vcr = Cm Kc = 8.78(0.46) = 4.06 m/s. (26.B34)

The second step is to calculate the correction factor for ship speed CV


2
4
3.59 3.59
CV = 8 − 0.5 + 0.0625 = 0.53. (26.B35)
4.06 4.06

The third step is the calculation of the correction factor for ship shape CF


2
2
10CB 10(0.86)
CF = = = 0.83. (26.B36)
Lpp /B 9.41
August 18, 2009 17:48 9.75in x 6.5in b684-ch26 FA

774 M. J. Briggs et al.

The fourth step is the calculation of the correction factor for squat at the critical
speed K∆T
 √
K∆T = 0.155 h/T = 0.155 1.2 = 0.17. (26.B37)

The last step is to substitute these values into the equation for bow squat Sb
Sb = CV CF K∆T T = 0.53(0.83)(0.17)(7.77) = 0.58 m. (26.B38)
This value is 0.35 m too small compared to the measured value of Sb = 0.93 m.
The Römisch trend is pretty good, however.

You might also like