08.02.2015 Views

Guidelines for Managing Wake Wash from High ... - PIANC USA

Guidelines for Managing Wake Wash from High ... - PIANC USA

Guidelines for Managing Wake Wash from High ... - PIANC USA

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

INTERNATIONAL NAVIGATION ASSOCIATION<br />

GUIDELINES FOR MANAGING WAKE WASH<br />

FROM HIGH-SPEED VESSELS<br />

Report of Working Group 41<br />

of the<br />

MARITIME NAVIGATION COMMISSION<br />

INTERNATIONAL NAVIGATION<br />

ASSOCIATION<br />

© COPYRIGHT <strong>PIANC</strong><br />

ASSOCIATION INTERNATIONALE<br />

DE NAVIGATION<br />

2003


<strong>PIANC</strong> has Technical Commissions concerned with inland waterways and ports (InCom), coastal and ocean<br />

waterways (including ports and harbours) (MarCom), environmental aspects (EnviCom) and sport and pleasure<br />

navigation (RecCom).<br />

This Report has been produced by an international Working Group convened by the Maritime Navigation<br />

Commission (MarCom). Members of the Working Group represent several countries and are acknowledged<br />

experts in their profession.<br />

The objective of this report is to provide in<strong>for</strong>mation and recommendations on good practice. Con<strong>for</strong>mity is not<br />

obligatory and engineering judgement should be used in its application, especially in special circumstances.<br />

This report should be seen as an expert guidance and state of the art on this particular subject. <strong>PIANC</strong> disclaims<br />

all responsibility in case this report should be presented as an official standard.<br />

© COPYRIGHT <strong>PIANC</strong><br />

<strong>PIANC</strong> General Secretariat<br />

Graaf de Ferraris-gebouw – 11th floor<br />

Boulevard du Roi Albert II 20, B.3<br />

B-1000 Brussels<br />

BELGIUM<br />

http://www.pianc-aipcn.org<br />

VAT/TVA BE 408-287-945<br />

ISBN 2-87223-142-0<br />

All rights reserved


CONTENT<br />

Working group members . . . . . . . . . . . . . . . . . . . . . . .4<br />

Technical expert . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4<br />

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5<br />

1.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . .5<br />

1.2 Method of undertaking the task . . . . . . . . . . . .5<br />

1.3 Definition of high-speed vessel . . . . . . . . . . . .5<br />

2. International experience with high-speed<br />

vessel wake . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5<br />

3. Wave generation and vessel wake . . . . . . . . . . . . . .6<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . .6<br />

3.2 Wave generation and wake characteristics . . . .6<br />

3.2.1 Conventional wake patterns –<br />

sub-critical wash . . . . . . . . . . . . . . . .6<br />

3.2.2 Near-critical wash . . . . . . . . . . . . . . .6<br />

3.2.3 Super-critical wave pattern . . . . . . . . .7<br />

3.2.4 General mechanisms . . . . . . . . . . . . .8<br />

3.3 Propagation and trans<strong>for</strong>mation<br />

of high-speed vessel wake . . . . . . . . . . . . . . . .9<br />

3.3.1 Direction of wave propagation . . . . . .9<br />

3.3.2 Composition of wash and its frequency<br />

components . . . . . . . . . . . . . . . . . . .10<br />

3.3.3 Wave trans<strong>for</strong>mation and<br />

wave decay . . . . . . . . . . . . . . . . . . .10<br />

3.4 <strong>Wake</strong> in coastal areas . . . . . . . . . . . . . . . . . .11<br />

3.4.1 General wave trans<strong>for</strong>mation . . . . . .11<br />

3.4.2 Propagation of vessel wake<br />

in coastal zone . . . . . . . . . . . . . . . . .11<br />

3.5 <strong>Wake</strong> generated by non-steady operation . . . .12<br />

3.5.1 Acceleration, deceleration,<br />

and extent of wash . . . . . . . . . . . . . .12<br />

3.5.2 Change of wake regime with<br />

constant speed . . . . . . . . . . . . . . . . .12<br />

3.5.3 Focusing the wash during course<br />

changes . . . . . . . . . . . . . . . . . . . . . .12<br />

3.5.4 Impact of bathymetry on<br />

operational procedures . . . . . . . . . . .13<br />

3.6 Bernoulli wake and soliton waves . . . . . . . . .13<br />

3.7 <strong>Wake</strong> prediction, analysis, and assessment . .13<br />

3.7.1 Numerical prediction of wake . . . . .13<br />

3.7.2 Full-scale trials and measurements . .15<br />

3.7.3 Model scale trials . . . . . . . . . . . . . . .17<br />

3.7.4 <strong>Wake</strong> analysis . . . . . . . . . . . . . . . . .17<br />

3.7.5 <strong>Wake</strong> assessment . . . . . . . . . . . . . . .17<br />

4.3 Environmental impact of wake . . . . . . . . . . .20<br />

4.3.1 Overview . . . . . . . . . . . . . . . . . . . . . . .20<br />

4.3.2 Potential impacts . . . . . . . . . . . . . . .20<br />

4.3.3 Differentiating between causes . . . . .22<br />

4.3.4 Predicting environmental impact . . .22<br />

5. <strong>Managing</strong> vessel wake . . . . . . . . . . . . . . . . . . . . . .22<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . .22<br />

5.2 Management measures . . . . . . . . . . . . . . . . .23<br />

5.2.1 Vessel design . . . . . . . . . . . . . . . . . .23<br />

5.2.2 Operational measures . . . . . . . . . . . .23<br />

5.2.3 Non-operational measures . . . . . . . .24<br />

5.3 Route assessment . . . . . . . . . . . . . . . . . . . . .24<br />

5.3.1 Overview . . . . . . . . . . . . . . . . . . . . .24<br />

5.3.2 Route characterization . . . . . . . . . . .24<br />

5.3.3 Impact identification . . . . . . . . . . . .25<br />

5.3.4 Developing and assessing potential<br />

management measures . . . . . . . . . . .27<br />

5.3.5 Monitoring . . . . . . . . . . . . . . . . . . . .28<br />

6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . .28<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . .29<br />

Annex A – Terms of reference . . . . . . . . . . . . . . . . . .32<br />

© COPYRIGHT <strong>PIANC</strong><br />

4. Impacts associated with vessel wake . . . . . . . . . . .17<br />

4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . .17<br />

4.2 Safety impacts . . . . . . . . . . . . . . . . . . . . . . . .18<br />

4.2.1 Safety of people . . . . . . . . . . . . . . . .18<br />

4.2.2 Safety of vessels . . . . . . . . . . . . . . . .18<br />

4.2.3 Structural damage . . . . . . . . . . . . . .19<br />

3 Report of Working Group 41 - MARCOM


WORKING GROUP MEMBERS<br />

Christian Aage, Technical University of Denmark,<br />

Denmark<br />

Adrian Bell (Chair until September 2002), Kirk McClure<br />

Morton, Northern Ireland, United Kingdom<br />

Lars Bergdahl, Chalmers University of Technology,<br />

Sweden<br />

Alan Blume (Chair), U.S. Coast Guard, United States of<br />

America<br />

Ernst Bolt, Ministry of Public Works, Transport and Water<br />

Management, Netherlands<br />

Hendrik Eusterbarkey, Federal Waterways and Shipping<br />

Administration, Germany<br />

Tetsuya Hiraishi, Port and Harbour Research Institute,<br />

Ministry of Transport, Japan<br />

Henrik Kofoed-Hansen, DHI Water & Environment,<br />

Denmark<br />

Denis Maly, Port Authority Bruges-Zeebrugge, Belgium<br />

Martin Single, University of Canterbury, New Zealand<br />

Jorma Rytkönen, VTT Manufacturing Technology,<br />

Finland<br />

Trevor Whittaker, Queen’s University, Northern Ireland,<br />

United Kingdom<br />

TECHNICAL EXPERT<br />

Björn Elsäßer, Queen’s University, Northern Ireland,<br />

United Kingdom<br />

© COPYRIGHT <strong>PIANC</strong><br />

Report of Working Group 41 - MARCOM<br />

4


1. INTRODUCTION<br />

1.1 SUMMARY<br />

The intent of this report is to provide an overview of the<br />

hydrodynamic and physical aspects of high-speed vessel<br />

wake and to provide guidance <strong>for</strong> its effective management.<br />

This guidance does not prescribe a solution. Rather,<br />

it provides a process that waterway management authorities<br />

and vessel operators can use to develop an appropriate<br />

solution <strong>for</strong> managing high-speed vessel wake. This guidance<br />

is consistent with the International Maritime<br />

Organization’s (IMO) <strong>High</strong>-Speed Craft Code (2000)<br />

and may be used to support the development of the route<br />

operational manual required by Regulation 18.2.2 of that<br />

Code.<br />

The report does not address other issues associated with<br />

the operation of high-speed vessels that are within the<br />

purview of other organisations (e.g., IMO, the<br />

International Association of Marine Aids to Navigation<br />

and Lighthouse Authorities (IALA)); nor does it address<br />

issues related to vessel design since they are beyond the<br />

expertise of <strong>PIANC</strong>.<br />

Although these guidelines were developed in response to<br />

concern about the potential impacts of high-speed vessel<br />

wake (see Annex A), they can be used as the basis <strong>for</strong><br />

managing potential impacts of wake generated by any vessel.<br />

1.2 METHOD OF<br />

UNDERTAKING THE TASK<br />

The working group accomplished its task through a series<br />

of three meetings and correspondence. Members of the<br />

working group also corresponded with experts in their<br />

respective countries.<br />

© COPYRIGHT <strong>PIANC</strong><br />

1.3 DEFINITION<br />

OF HIGH-SPEED VESSEL<br />

For the purpose of this report, the term “high-speed vessel”<br />

includes vessels that meet the definition of a highspeed<br />

craft in the IMO <strong>High</strong>-Speed Craft Code (a vessel<br />

capable of a maximum speed equal to or exceeding<br />

3.7 ∇ 0.1667 (m/s), where ∇(m 3 ) is the displacement of the<br />

vessel at the design waterline), or a definition adopted by<br />

a national maritime authority.<br />

2. INTERNATIONAL<br />

EXPERIENCE WITH<br />

HIGH-SPEED VESSEL WAKE<br />

Real or perceived safety and environmental impacts associated<br />

with high-speed vessel wake in confined waters<br />

have been reported in many locations including Canada<br />

(Sandwell, 2000), Denmark (Kofoed-Hansen, 1996;<br />

Danish Maritime Authority, 1997; Kirkegaard et al., 1998;<br />

Kofoed-Hansen and Mikkelsen, 1997), Great Britain<br />

(Marine Accident Investigation Branch, 2000), Ireland<br />

(Maritime and Coastguard Agency, 1998), Sweden (Strom<br />

and Ziegler, 1998; Allenström et al., 2003), The<br />

Netherlands (Anonymous, 2000), New Zealand (Croad<br />

and Parnell, 2002; Kirk and Single, 2000; Parnell, 1996;<br />

Single and Kirk, 1999), Finland and Estonia (Peltoniemi et<br />

al., 2002), and the United States (Anonymous, 1999;<br />

Stumbo et al., 1999). To date, much of the research undertaken<br />

on high-speed vessel wake wash has appeared only<br />

as unpublished reports <strong>for</strong> various authorities and management<br />

agencies.<br />

<strong>Wake</strong> effects on rivers, lakes, and other inland waters<br />

also can be substantial. For example, see Nanson et al.<br />

(1994), Gadd (1994), Pickrill (1985), and references<br />

therein.<br />

As a result of these reported impacts, vessel operators and<br />

waterway managers alike have tried a number of different<br />

approaches <strong>for</strong> managing high-speed vessel wake. These<br />

include establishing standards <strong>for</strong> maximum allowed wave<br />

heights, maximum allowed wave energy, speed limits, and<br />

risk assessments. Other measures that have been used<br />

include the installation of wave-absorbing materials on sea<br />

walls and other harbour structures as well as ef<strong>for</strong>ts to educate<br />

other waterway users about the potential impacts of<br />

high-speed vessel wake. Based on the experience to date<br />

<strong>from</strong> these different ef<strong>for</strong>ts, it is increasingly apparent that<br />

the effective management of high-speed vessel wake is a<br />

multi-faceted problem that defies a simple “one size fits<br />

all” solution. It is also apparent that a means of managing<br />

high-speed vessel wake wash that addresses legitimate<br />

concerns related to waterway safety and protection of the<br />

marine environment, while also not unduly restricting<br />

their operation, must be identified to realize the full contribution<br />

of high-speed vessels to the national and global<br />

marine transportation systems.<br />

5 Report of Working Group 41 - MARCOM


3. WAVE GENERATION AND<br />

VESSEL WAKE<br />

3.1 INTRODUCTION<br />

This section provides a description of the fundamentals of<br />

wake wave generation by vessels and wave mechanics. It<br />

contains the basic in<strong>for</strong>mation necessary to understand the<br />

effects of wake on the coastline and on other vessels. The<br />

description primarily deals with the free wave system<br />

sometimes referred to as Kelvin, Havelock, and Bernoulli<br />

wave systems and does not discuss near-vessel or bound<br />

wave system problems.<br />

3.2 WAVE GENERATION AND<br />

WAKE CHARACTERISTICS<br />

3.2.1 Conventional wake<br />

patterns – sub-critical wash<br />

© COPYRIGHT<br />

In 1887, Lord Kelvin described the wave system produced<br />

by a moving vessel in deep water. This wave pattern is<br />

confined to a wedge shape known as the Kelvin wedge<br />

(Fig. 3-1). Within this wedge, which has an apex angle of<br />

±19.5°, it is possible to distinguish between diverging<br />

waves, propagating at an angle of θ = 90º to 35º <strong>from</strong> the<br />

vessel’s track, and transverse waves, propagating at an<br />

angle of θ = 35º to 0º <strong>from</strong> the vessel’s track. Under<br />

steady conditions, the transverse waves along the vessel’s<br />

track travel at the same speed as the vessel.<br />

<strong>PIANC</strong><br />

V<br />

Fn l = ,<br />

gL<br />

where V (m/s) is the vessel speed, g (m/s 2 ) is the acceleration<br />

of gravity, and L (m) is the vessel waterline length.<br />

The Kelvin wave pattern is strictly correct only <strong>for</strong> a moving<br />

point source. The wash of a vessel will be the superposition<br />

of a number of these patterns <strong>from</strong> many sources,<br />

of which the bow and stern wave system will generally be<br />

dominant.<br />

When the length Froude number is 0.4 (Fn l = 2π ) in deep<br />

water, the wavelength of the transverse waves is equal to<br />

the vessel’s length. The pressure peaks at the bow and<br />

stern amplify each other while the wave-making resistance,<br />

which is the net longitudinal <strong>for</strong>ce due to the fluid<br />

pressure of the water acting on the hull, increases significantly<br />

(Lewis, 1988). This speed is called the hump speed<br />

and <strong>for</strong>ms a firm barrier <strong>for</strong> most ‘conventional’ vessels<br />

(Fig. 3-2). Basically the vessel continuously tries to move<br />

‘uphill’, which requires more energy. In practice the resistance<br />

hump speed most often occurs when length Froude<br />

numbers are between approximately 0.4 and 0.6. <strong>High</strong>speed<br />

vessels, the subject of these guidelines, are able to<br />

pass through this ‘hump speed’ in deep water. This<br />

requires a relatively high ratio of propulsion power to vessel<br />

displacement.<br />

Figure 3-2 Resistance of a typical ship hull<br />

in deep and shallow water<br />

Figure 3-1 Steady state Kelvin wave pattern<br />

(Ekman, 1906; Newman, 1977)<br />

All vessels operating in deep water produce a Kelvin type<br />

wave pattern. In deep water the wavelength is a function<br />

of the wave speed. The higher the vessel’s speed, the<br />

longer the generated waves become. The most important<br />

parameter <strong>for</strong> the characterisation of the Kelvin wave pattern<br />

is the vessel’s speed related to its waterline length<br />

given in dimensionless terms by the length Froude number<br />

(Fn l ). The length Froude number is defined as:<br />

The Kelvin wave pattern is only valid <strong>for</strong> deep-water<br />

conditions. Deep water in this context means that the<br />

vicinity of the bottom does not affect the propagation<br />

speed of the waves produced.<br />

3.2.2 Near-critical wash<br />

In 1908, Havelock investigated the wave pattern generated<br />

by a single point source in shallow water. He introduced<br />

the depth Froude number, stating that the characteristics of<br />

the wave pattern in depth-limited water are a function of<br />

Report of Working Group 41 - MARCOM<br />

6


vessel speed and water depth. The depth Froude number<br />

(Fn h ) is defined as the ratio of the vessel speed to the wave<br />

propagation speed in shallow water:<br />

V<br />

Fn h = ,<br />

gh<br />

where: V (m/s) is the vessel speed, g (m/s 2 ) is the acceleration<br />

of gravity, and h (m) is the water depth. The classical<br />

Kelvin wave pattern will be generated at depth Froude<br />

numbers under 0.57. The length of the transverse waves,<br />

which are the longest waves in the pattern, will increase as<br />

the depth Froude number becomes larger.<br />

As the depth Froude number approaches 1.0, the vessel’s<br />

speed becomes equal to the maximum wave propagation<br />

speed in the given depth of water. This speed is often<br />

referred to as the critical speed. At this stage, all transverse<br />

waves are left behind the vessel and a wave builds perpendicular<br />

to the vessel as shown in Figure 3-3. If the vessel<br />

stays at critical speed, this wave will extend further <strong>from</strong><br />

the vessel’s track and build in height. All diverging waves<br />

are found behind this critical wave. The conventional<br />

Kelvin wave pattern is often referred to as sub-critical.<br />

Where the water depth starts to change the wave pattern<br />

significantly (Fn h > 0.85), the range is often referred to as<br />

the near-critical speed range.<br />

3.2.3 Super-critical wave pattern<br />

At higher depth Froude numbers (Fig. 3-4) the transverse<br />

waves disappear. They simply cannot keep up with the<br />

vessel since the water depth limits their speed. In water of<br />

constant depth, the first wave crest in the pattern is straight<br />

and the crest length is related to the time the vessel has<br />

been travelling at the particular depth Froude number. The<br />

following waves have curved crests and troughs and are<br />

also continuous.<br />

Figure 3-4 Super-critical wave pattern<br />

in constant water depth<br />

The angle of propagation θ <strong>for</strong> the first wave is determined<br />

by:<br />

cos θ = 1 / Fn h .<br />

There<strong>for</strong>e, as the depth Froude number increases, the<br />

waves bend further backwards and their angle of propagation<br />

becomes increasingly perpendicular to the vessel’s<br />

track. For example, whereas at a depth Froude number of<br />

1.5 the angle of propagation is approximately 48º; at a<br />

depth Froude number of 2.0 the angle of propagation is<br />

60°.<br />

Figure 3-3 Critical wave pattern<br />

in constant water depth<br />

At critical speed there is one transverse wave that moves<br />

with the vessel and dramatically increases the wave-making<br />

resistance. This is because the energy in the longest<br />

wave can no longer disperse generating more waves<br />

behind, as was the case at lower speeds. This is the main<br />

reason most conventional vessels are not capable of<br />

exceeding the critical speed. It also needs to be mentioned<br />

that the critical wave pattern is unsteady and changes with<br />

time. The shape of the first wave is a function of the distance<br />

the vessel has been travelling at critical speed.<br />

The wash generated by a high-speed vessel operating<br />

above the critical depth Froude number in shallow water<br />

has some unique peculiarities. A typical wave time trace<br />

of wash <strong>from</strong> a high-speed vessel operating in shallow<br />

coastal waters measured 2700 meters <strong>from</strong> its track is<br />

shown in Figure 3-5. The trace has been divided into<br />

zones according to wave period and grouping. The first<br />

group, Zone I, comprises the long period super-critical<br />

wash waves and is peculiar to high-speed vessels operating<br />

above the critical depth Froude number. The second<br />

group, Zone II, comprises waves similar in size and height<br />

to the wash waves produced by conventional vessels of<br />

comparable displacement and length. Finally there is a<br />

group of tail waves in Zone III. These are peculiar to highspeed<br />

vessels with transom sterns.<br />

7 Report of Working Group 41 - MARCOM


Figure 3-5 Typical wave trace of a large high-speed vessel at super-critical depth<br />

Froude number with wash zones indicated<br />

©<br />

The time series was measured at a distance of 2700 meters<br />

<strong>from</strong> a catamaran. The average wave periods are 11 seconds<br />

(s), 5s, and 3s, <strong>for</strong> Zone I, II and III respectively.<br />

3.2.4 General mechanisms<br />

The basic mechanism of wash propagation is discussed in<br />

the following section. For simplicity, linear wave theory is<br />

used, though there is an argument to use higher order wave<br />

theory in very shallow water.<br />

With regards to the propagation of surface water waves the<br />

water depth plays a significant role. In deep water the<br />

wave celerity is a function of the wavelength, but in shallow<br />

water the water depth limits the wave celerity. Water<br />

depths of more than half the wavelength can be considered<br />

as deep water, whereas water depths of less than 1/20 th the<br />

wavelength is commonly regarded as shallow water.<br />

COPY-<br />

RIGHT<br />

a) b)<br />

Figure 3-6 a) illustrates the deep-water propagation of<br />

waves generated by a point source moving <strong>from</strong> A to B at<br />

speed V. Waves generated at A would have travelled as<br />

far as the outer circle if they had propagated at the individual<br />

wave phase speed (celerity). The fastest and<br />

longest waves would have moved along the track to B<br />

while shorter and slower waves would have travelled to<br />

locations such as C and D. However, in deep water the<br />

wave energy travels at only one-half the phase speed. As a<br />

result, the locus of wave energy, and there<strong>for</strong>e the largest<br />

wave, moves only as far as the inner circle, to points such<br />

as B’, C’, and D’. In the meantime, the point source has<br />

generated new waves along its way to B. The energy <strong>from</strong><br />

these waves propagates outward in progressively smaller<br />

inner circles. The resulting limit of wave propagation <strong>for</strong><br />

all waves generated as the ship moves <strong>from</strong> A to B is tangent<br />

to the inner circle <strong>for</strong>ming the Kelvin wedge. For any<br />

value of V, this <strong>for</strong>ms an apex angle of ±sin -1 (1/3) =<br />

±19.5°.<br />

<strong>PIANC</strong><br />

Figure 3-6 a) Wave rays <strong>for</strong> sub-critical operation b) Wave rays <strong>for</strong> super-critical operation at Fn h = 1.3<br />

Report of Working Group 41 - MARCOM<br />

8


The visible wave pattern then lies mainly inside and along<br />

the wedge. As individual waves propagate outward at<br />

their phase speed and attempt to outrace the wedge, they<br />

die out and disappear. The largest waves, there<strong>for</strong>e, <strong>for</strong>m<br />

along the wedge and wave amplitudes diminish within a<br />

short distance outside the wedge (Lighthill, 1978;<br />

Newman, 1977).<br />

A similar wave ray diagram can be drawn <strong>for</strong> a shallow<br />

water wave pattern and a depth Froude number greater<br />

than 1 with the important difference, however, that in shallow<br />

water the celerity and speed of energy propagation are<br />

equal. This is shown in Figure 3-6 b) <strong>for</strong> a depth Froude<br />

number of 1.3. There are no waves travelling <strong>from</strong> A at an<br />

angle θ of less than 39.7° in this example. All waves radiated<br />

at A have travelled as far as the outer semicircle. As<br />

waves in shallow water are non-dispersive, the energy of<br />

the wave ray A – C has travelled to C. The wave travelling<br />

in the direction of D is in intermediate water depth.<br />

Hence, the energy has travelled more than half the distance<br />

A – D, as this wave is partially dispersive. The wave travelling<br />

towards F is a deep-water wave, as its wavelength is<br />

short compared to the water depth and, hence, it is fully<br />

dispersive. As a result the energy has travelled only half<br />

the distance to F'.<br />

3.3 PROPAGATION AND<br />

TRANSFORMATION OF<br />

HIGH-SPEED VESSEL WAKE<br />

The following sections focus on propagation of the free<br />

wave field generated by a vessel at super-critical speed.<br />

© COPYRIGHT<br />

3.3.1 Direction of wave propagation<br />

The direction of propagation <strong>for</strong> the leading wave in the<br />

super-critical wash is only a function of the depth Froude<br />

number. This was originally proposed by Havelock (1908)<br />

and proven by various other authors. This relationship is<br />

plotted in Figure 3-7.<br />

Whittaker et al. (2000) studied the divergence between the<br />

first and the following waves and determined that this<br />

angle is dependent on the depth Froude number. The theoretical<br />

divergence between the first and second leading<br />

crest is shown in Figure 3-8. The divergence angle is<br />

defined as the difference in propagation direction between<br />

successive waves. Whittaker et al. (2000) has shown that<br />

this relationship is valid <strong>for</strong> real wave patterns. At high<br />

super-critical Froude numbers the waves are more parallel<br />

than at lower super-critical depth Froude numbers. The<br />

divergence decreases with perpendicular distance <strong>from</strong> the<br />

vessel; the rate of decrease also depends on the water<br />

depth. In deeper water the change in wave pattern is less<br />

than with the same distance in shallow water.<br />

wave direction [°]<br />

<strong>PIANC</strong><br />

Figure 3-7 Angle of wave propagation<br />

at outer edge of stationary pattern (sub-critical)<br />

and leading wave (super-critical)<br />

Figure 3-8 Divergence of second wave compared to first with varying depth<br />

Froude number and perpendicular distance <strong>from</strong> vessel normalised by water depth<br />

9 Report of Working Group 41 - MARCOM


With each successive wave the divergence angle decreases<br />

at constant perpendicular distance <strong>from</strong> the vessel’s track.<br />

Thus, the waves further back in the pattern become<br />

increasingly parallel to each other. Finally, the pattern<br />

contains waves whose crests are almost parallel to the vessel's<br />

track. The slight change in the propagation direction<br />

of the waves leads to an apparent lengthening of the waves<br />

farther <strong>from</strong> the vessel. Hence, the peak-to-peak distance<br />

of the wave increases with distance <strong>from</strong> the vessel’s track.<br />

Waves with a period in excess of 40 seconds have been<br />

measured 2.7 kilometres <strong>from</strong> the vessel (MCA, 1998).<br />

Though the wash measured as a time series often appears<br />

as a series of sinusoidal waves each with a slightly shorter<br />

period, the surface elevation is in fact a superposition of an<br />

infinite number of waves. Using the zero crossing period<br />

on its own to calculate the wave speed will lead to slightly<br />

smaller values of wave celerity. Most importantly it has<br />

to be recognised that the period of the waves in a vessel’s<br />

wake is not a function of hull design, but rather of vessel<br />

speed, water depth, and the distance the wave travels <strong>from</strong><br />

the point where it was generated. In contrast, the height of<br />

the waves in the wake is a function of the hull design. In<br />

general, <strong>for</strong> vessels with a similar hull <strong>for</strong>m but different<br />

lengths, the height of waves generated by the shorter vessel<br />

will be lower than those generated by the longer vessel.<br />

The implication is that the period of the waves generated<br />

by vessels of different lengths operating at the same depth<br />

Froude number will be the same. However, the waves<br />

generated by the shorter vessel will be less pronounced<br />

when compared to the waves generated by the longer vessel<br />

since they will not be as high.<br />

3.3.2 Composition of wash and<br />

its frequency components<br />

So far these guidelines have only focused on the location<br />

of crests and troughs in the wave pattern. The wave-making<br />

resistance, which is a significant part of the total resistance<br />

of a vessel, corresponds to the energy used to generate<br />

the wave pattern. This resistance is spread as a continuous<br />

spectrum over the wave rays radiated <strong>from</strong> the vessel.<br />

The distribution of the wave-making resistance is often<br />

referred to as the free wave spectrum or the amplitude<br />

function. In deep water this distribution is a function of<br />

the vessel speed, the hull length and displacement, and the<br />

hull shape.<br />

Figure 3-9 a) shows such a distribution <strong>for</strong> a typical ship<br />

hull in deep water. There is a different distribution <strong>for</strong><br />

each vessel speed <strong>for</strong> a given hull, or to be more precise,<br />

<strong>for</strong> each length Froude number. In shallow water the distribution<br />

is a function of the vessel speed, water depth, hull<br />

length, displacement, and shape. A shallow water distribution<br />

<strong>for</strong> the same vessel speed is given in Figure 3-9 b)<br />

<strong>for</strong> a depth Froude number of 1.3. Note that there are no<br />

wave resistance components between θ = 0° and 39.7°,<br />

and a large amount of energy is concentrated between<br />

39.7° and 60°. These resistance components <strong>for</strong>m the<br />

leading waves in the super-critical wave pattern. The area<br />

underneath these curves is proportional to the wave pattern<br />

resistance. In this case, the area under both curves is<br />

almost equal.<br />

3.3.3 Wave trans<strong>for</strong>mation and wave decay<br />

Havelock (1908) investigated the decay of wave height<br />

with distance at sub-critical depth Froude numbers. He<br />

found that the decay of both the transverse and diverging<br />

waves is proportional to ϖ n , where ϖ is the distance <strong>from</strong><br />

the vessel and n is a constant value. He showed that the<br />

theoretical decay inside the Kelvin wedge has the exponent<br />

n = -1/2 and the waves along the outer edge of the<br />

wedge decay have an exponent of n = -1/3.<br />

Figure 3-10 Decay of wave amplitude<br />

as investigated by Havelock<br />

© COPYRIGHT <strong>PIANC</strong><br />

Figure 3-9 a) Distribution of wave-making resistance over propagation angle (θ)<br />

b) the resistance is here made dimensionless as resistance coefficient (C w )<br />

Report of Working Group 41 - MARCOM<br />

10


Kofoed-Hansen et al. (1999) suggested a decay rate of<br />

n = -0.55 based on a best fit through a wide range of wake<br />

measurements <strong>for</strong> catamarans operating at super-critical<br />

speeds. They also stated that a decay rate of n = -1/3 close<br />

to the vessel (less than 3 vessel lengths) would be more<br />

appropriate. Whittaker et al. (2001) concluded <strong>from</strong> a<br />

series of towing tank tests in shallow water that the decay<br />

rate could be substantially less. The lowest decay rate<br />

observed was as low as n = - 0.2. While Havelock compared<br />

the wave height at a straight ray <strong>from</strong> the vessel (Fig.<br />

3-10), the later comparisons are based on the maximum<br />

wave height found in wave cuts at different distances perpendicular<br />

<strong>from</strong> the vessel’s track.<br />

3.4 WAKE IN COASTAL AREAS<br />

3.4.1 General wave trans<strong>for</strong>mation<br />

There are a number of processes that can affect the wave<br />

as it propagates into coastal and shallow water areas. The<br />

most important are listed as follows:<br />

• Refraction - the bending of wave fronts and wave height<br />

reductions as they travel into shallower water<br />

• Diffraction - the lateral transfer of energy along the<br />

wave crest<br />

• Shoaling - the alteration in wave height as waves propagate<br />

into shallow water<br />

• Wave-current interaction - the alteration of wave celerity<br />

and height due to currents<br />

• Breaking - energy dissipation due to increased wave<br />

steepness<br />

• Friction - energy dissipation due to friction on seabed or<br />

obstacles or due to percolation through a porous seabed<br />

• Reflection - alteration in wave height due to full or<br />

partial reflection <strong>from</strong> structures or seabed<br />

© COPYRIGHT <strong>PIANC</strong><br />

All these processes are general coastal processes and are<br />

not discussed in detail as part of these guidelines. For<br />

more details see Vincent et al. (2002).<br />

3.4.2 Propagation of vessel wake<br />

in coastal zone<br />

The propagation of waves <strong>from</strong> a vessel to the shore can be<br />

divided into two processes:<br />

• Vessel speed - depth Froude number based trans<strong>for</strong>mation<br />

of wake. As described in Section 3.2.3, the first<br />

wave crest of the supercritical wave pattern is straight in<br />

constant water depth, and the angle of propagation<br />

depends only on the ratio of vessel speed to maximum<br />

wave speed. The following waves trans<strong>for</strong>m, and as a<br />

result the wave period changes with distance <strong>from</strong> the<br />

vessel in water of constant depth.<br />

• Trans<strong>for</strong>mation of wake due to change in bathymetry.<br />

Variations in waterway bathymetry can cause the<br />

height, propagation direction, and celerity of waves to<br />

change. These are common coastal processes as listed<br />

in Section 3.4.1.<br />

Both the depth Froude number based trans<strong>for</strong>mation and<br />

the bathymetric trans<strong>for</strong>mation need to be combined to<br />

predict the wash. The leading wave can be treated as a<br />

monochromatic shallow water wave. The following waves<br />

are a superposition of an infinite number of wave components<br />

(see Section 3.3.1). Even in shoaling water, the<br />

waves <strong>for</strong>ming the pattern have different celerities, which<br />

results in the dispersion of energy <strong>from</strong> one wave to the<br />

next further back in the trace.<br />

Once in the breaker zone the linear wave theory is no<br />

longer valid. This means that the waves can no longer be<br />

considered as being sinusoidal. Several wave theories<br />

have been developed in the past century, which deal with<br />

shallow water waves like Boussinesq wave theory, or<br />

higher order wave theory (Demirbilek and Vincent, 2002).<br />

In some cases the long waves propagating into areas with<br />

extremely shallow water, e.g., on mud flats or salt<br />

marshes, can be approximated as solitary waves.<br />

In some studies the leading group of long-period waves,<br />

which are often the most critical waves in terms of risk,<br />

have been treated as waves with a very narrow spectrum<br />

(Kofoed-Hansen et al., 1996 and 1999; MCA, 1998). The<br />

correlation between the modelling and some full-scale<br />

measurements proved to be sufficient <strong>for</strong> most practical<br />

problems.<br />

<strong>Wash</strong> and wind waves will superimpose causing local<br />

steepening that can lead to breaking and energy loss.<br />

Otherwise, they will pass through each other unchanged.<br />

In general, high-speed wash is very small compared to<br />

fully developed wind waves. Thus, high-speed wash can<br />

hardly be detected in wash measurements with the presence<br />

of pronounced wind waves or swells. Vessel wash<br />

can, however, still be a problem on the shoreline even<br />

under storm conditions as the very long period waves are<br />

still present but masked. No evidence has been found that<br />

suggests that wind affects the long wave components in<br />

high-speed wash because the interactive time is too short<br />

and the steepness is too small.<br />

11 Report of Working Group 41 - MARCOM


3.5 WAKE GENERATED BY<br />

NON-STEADY OPERATION<br />

3.5.1 Acceleration, deceleration,<br />

and extent of wash<br />

During acceleration in water of constant depth the wave<br />

pattern and the wave resistance will change. When accelerating<br />

<strong>from</strong> sub-critical to super-critical speeds the vessel<br />

has to pass through a resistance hump at critical speed (see<br />

Fig. 3-2). While at near-critical or critical speeds, the<br />

wake generated is more energetic vis-à-vis the wake generated<br />

at either sub- or super-critical speeds. It is, there<strong>for</strong>e,<br />

desirable to pass through the near-critical speed<br />

range as quickly as possible. The same applies to deceleration,<br />

where the operator should aim <strong>for</strong> a quick decrease<br />

in speed near the critical wake regime. The different wake<br />

regimes are shown in Figure 3-11. If possible, the operator<br />

should avoid operation at depth Froude numbers<br />

between approximately 0.85 and 1.1.<br />

unacceptable waves were being generated when the vessel<br />

was abeam of location A and reduced speed immediately,<br />

it is likely that substantial wash would still propagate to<br />

location A.<br />

Figure 3-12 Outer envelope and extent of wash<br />

3.5.2 Change of wake regime<br />

with constant speed<br />

Figure 3-11 <strong>Wake</strong> regime depending<br />

on speed and water depth<br />

When accelerating the vessel will produce a wave pattern<br />

with a transitional zone. This transitional zone is nonsteady<br />

and will change its position and area with time. A<br />

simplified wave pattern generated by a high-speed vessel<br />

with instantaneous acceleration in constant water depth is<br />

shown in Figure 3-12. It is assumed the vessel started its<br />

passage at the left and instantly accelerated to a super-critical<br />

speed. Three zones can be identified in the graph. An<br />

object at location A will encounter the full wash waves.<br />

An object at location B will experience the same leading<br />

waves as location A; however, with reduced wave heights.<br />

An object at location C may be subjected to little or no<br />

wave action <strong>from</strong> the leading waves. As the vessel continues<br />

to operate, some waves will continue to travel toward<br />

locations A, B, and C with the boundary of the wash continuously<br />

moving outward. Finally, location C will<br />

encounter the short tail waves (Fig. 3-5, Zone III). The<br />

implication is that even if the operator recognised that<br />

Report of Working Group 41 - MARCOM<br />

© COPY-<br />

RIGHT <strong>PIANC</strong><br />

12<br />

Although a high-speed vessel may continue to operate at<br />

constant speed, the wake regime may change with variation<br />

of water depth along the vessel’s route. A vessel<br />

approaching a fairway <strong>from</strong> deep water may operate at<br />

sub-critical speeds. Continuation of operation at high<br />

speeds into shallow water results in the transition to critical<br />

speed. Depending on the bathymetry the vessel might<br />

operate <strong>for</strong> a considerable time at near-critical speeds and<br />

finally proceed to super-critical speed or decelerate. It is<br />

extremely important to recognise that the wash progressing<br />

toward the shoreline may have been caused by operating<br />

at near-critical depth Froude numbers along an earlier<br />

part of the passage.<br />

3.5.3 Focusing the wash during course changes<br />

When a vessel changes course, the energy density on the<br />

inside of the turn will be higher than the energy density on<br />

the outside of the turn. The difference is more pronounced<br />

<strong>for</strong> larger course changes than smaller changes as well as<br />

when the vessel is operating at higher speeds. This is<br />

because although the energy transferred into the wake is<br />

equal on both sides of the vessel’s track, it becomes<br />

focused on the inside of the turn insofar as it is transferred<br />

into a smaller area. As a result, the waves generated by the<br />

vessel will also be higher on the inside of the turn than<br />

those generated when the vessel is operating on a straight<br />

course. Similarly, the waves on the outside of the turn will<br />

be lower.


To mitigate the higher energy density and waves on the<br />

inside of a turn, it may be necessary to increase the radius<br />

of the turn or to make the change in a number of smaller<br />

changes with sufficient distance between them to avoid<br />

creating areas of increased wash concentration. It may<br />

also be necessary to reduce speed during the change of<br />

course. Particular care should be taken where the focused<br />

wash of the inner bend will propagate onto shoaling areas,<br />

such as banks or headlands, as a further increase in wave<br />

height can be expected. On the other hand, turns can be<br />

used to decrease wash heights on the outer bend, if the<br />

operational area allows such manoeuvres.<br />

confined water are capable of producing large Bernoulli<br />

wakes. Figure 3-13 is an example of a Bernoulli wake generated<br />

by a large container vessel in water approximately<br />

15 meters deep operating at sub-critical speeds.<br />

3.5.4 Impact of bathymetry<br />

on operational procedures<br />

While the variation of water depth along the passage may<br />

change the wake regime it also has a significant impact on<br />

the power requirement of the vessel. As described in<br />

Section 3.2.2 and illustrated in Figure 3-2, the resistance<br />

of the vessel increases close to the critical depth Froude<br />

number. Hence, the vessel needs sufficient propulsion to<br />

overcome this resistance peak. A vessel approaching shallow<br />

water might, there<strong>for</strong>e, proceed at near-critical depth<br />

Froude number with constantly decreasing speed while<br />

generating a wake with higher energy compared to supercritical<br />

operation at the same water depth. This is most<br />

likely if the vessel’s speed is reduced <strong>for</strong> some reason, e.g.,<br />

extra drag due to hull fouling, the vessel is loaded higher<br />

beyond its normal service load, or partial power loss<br />

(Cain, 2000).<br />

Sudden bathymetric changes can be used to transition<br />

quickly <strong>from</strong> either a sub- or super-critical wake regime to<br />

the other without prolonged operation at near-critical<br />

speed. One means of accomplishing this is to construct a<br />

steep depth contour in the channel as described by<br />

Feldtmann and Garner (1999).<br />

3.6 BERNOULLI WAKE<br />

AND SOLITON WAVES<br />

The dynamic displacement of water caused by the <strong>for</strong>ward<br />

movement of a vessel through water results in a velocity<br />

field around the hull (bound wave field). While this velocity<br />

field is responsible <strong>for</strong> the pressure distribution along<br />

the surface and, hence, the generation of waves, it has little<br />

effect on the far field propagation and trans<strong>for</strong>mation of<br />

the Kelvin wave pattern (free wave field). However, in a<br />

confined environment, e.g., fairway, shallow water or<br />

canal, the flow field can be restricted by the surrounding<br />

boundaries. This is very distinct in inland canals and is<br />

directly related to the blockage.<br />

© COPY-<br />

RIGHT <strong>PIANC</strong><br />

Theoretically any type of vessel will produce such a displacement<br />

wave. However, in particular large vessels in<br />

Figure 3-13: Bernoulli wake and Kelvin wake<br />

generated by a large container vessel<br />

A vessel operating close to critical speed (i.e., depth<br />

Froude number between 0.85 and 1.1) is capable of generating<br />

a solitary wave, which in fact can be faster than the<br />

vessel. Both conventional vessels and high-speed vessels<br />

can produce solitary type waves, which are of very long<br />

period and can travel several vessel lengths ahead in very<br />

shallow open water. Large displacement vessels operating<br />

in shallow water are particularly prone to generating this<br />

type of wave (Scott-Russell, 1865; Dand et al., 1999;<br />

Whittaker et al., 2001).<br />

3.7 WAKE PREDICTION,<br />

ANALYSIS, AND ASSESSMENT<br />

3.7.1 Numerical prediction of wake<br />

A numerical non-linear time-domain model capable of<br />

calculating vessel-generated waves, wave propagation,<br />

and wave trans<strong>for</strong>mation in non-homogeneous media is<br />

the ultimate tool <strong>for</strong> evaluating the potential <strong>for</strong> vessel<br />

wake to have adverse safety or environmental impacts.<br />

Besides predicting the unstationary flow field and the<br />

associated wave pattern around the vessel hull, the same<br />

model (preferably) will be able to calculate the dynamics<br />

of the transient waves in the surf zone including the maximum<br />

wave height be<strong>for</strong>e wave breaking as well as run-up<br />

on beaches and river banks. Such models do not exist yet;<br />

however, if they did, they would be far too computationally<br />

demanding <strong>for</strong> practical use.<br />

In recent years, the use of computational fluid dynamic<br />

(CFD) codes has provided a valuable supplement to more<br />

classic methods <strong>for</strong> design and optimisation of vessel<br />

hulls. Comparisons between results <strong>from</strong> various CFD<br />

codes, full-scale measurements, and towing tank tests have<br />

increased confidence in using these models <strong>for</strong> prediction<br />

of wake wash caused by multi-hull vessels at varying<br />

13 Report of Working Group 41 - MARCOM


water depths. A general limitation of these models is that<br />

they do not permit calculations of the far-field wave pattern.<br />

Most often, the calculation is limited to an area within<br />

approximately three to five vessel lengths.<br />

Furthermore, the codes usually only provide a stationary<br />

solution of the potential flow field at a constant depth.<br />

Predicting the impact of the wake <strong>for</strong> a certain vessel operating<br />

on a given route requires, in most cases, the solution<br />

of an unsteady problem in a three-dimensional domain.<br />

This is, in most cases, not feasible with one particular<br />

numerical model. Hence, the numerical prediction of the<br />

wake problem is divided into two different problems:<br />

• Prediction of the wave field near the vessel.<br />

• Prediction of the wave trans<strong>for</strong>mation and propagation<br />

in the far field and coastal zone.<br />

Very few numerical models predict both problems; those<br />

that have been developed to date are coupled methods<br />

(Raven, 2000; Kofoed-Hansen et al., 2000).<br />

3.7.1.1 Prediction of the wave field near the vessel<br />

The wave field generated by a vessel can be derived by<br />

numerical methods in several ways. These methods can be<br />

grouped in principle in two categories: panel codes and<br />

thin vessel theory models. Because of the inherent limitations<br />

associated with each type of model, several research<br />

institutions are working on hybrid models.<br />

Potential Flow Panel Codes<br />

These codes use a non-linear free-surface potential flow<br />

method, where the hull and water surfaces are represented<br />

by a large number of panels. As the pressure along the hull<br />

is calculated, sinkage and trim can be predicted. Since the<br />

boundary conditions are non-linear the problem cannot be<br />

solved directly. The actual wave surface is derived by<br />

means of iteration starting with a flat surface and an estimated<br />

trim and sinkage. Shallow water effects can be<br />

implemented. However, as in most codes, the grid moves<br />

with the vessel and only steady state or quasi-steady state<br />

conditions can be computed. Only panel codes with timestepping<br />

treatment have the potential of solving fully nonsteady<br />

conditions. The computational ef<strong>for</strong>t is rather large<br />

and, thus, the computation time can be considerable. Most<br />

models calculate the wave elevation or free surface profile<br />

up to a few vessel lengths distance <strong>from</strong> the vessel. A<br />

review of the current capabilities of panel codes can be<br />

found in Raven (2000) and Hughes (2001) as well as more<br />

detailed literature.<br />

Thin ship theory code<br />

This theory, which is often referred to as slender body theory,<br />

assumes the vessel hull(s) to be slender compared to<br />

their length. Most theories represent the body of the vessel<br />

as a series of Kelvin sources along the centreline of the<br />

hull assuming a linear free surface condition. As such the<br />

theory is limited to linear wave theory and in particular<br />

small waves compared to the wavelength. The theory<br />

behind these programs originates <strong>from</strong> work carried out by<br />

Lord Kelvin, Havelock (1908), Mitchell (1898) and<br />

Eggers et al. (1967). The strength of each of the sources<br />

is derived <strong>from</strong> the local slope of the hull at a number of<br />

water lines. The locations where the local slope is determined<br />

are often referred to as panels. The underlying<br />

equations are multiple integrals and are solved directly by<br />

numerical integration. The code can compute only steady<br />

state (constant speed) conditions and shallow water problems.<br />

Some programs include the input of reflective<br />

boundaries to simulate tank walls and, thus, allow the<br />

results to be compared with narrow tank experiments.<br />

Trim and sinkage is usually user defined. More recent<br />

codes derive the sinkage and trim by means of calculating<br />

the <strong>for</strong>ces on the hull <strong>from</strong> the surrounding surface elevation<br />

through iteration. Difficulties occur with transom<br />

sterns, as basic thin vessel theory cannot deal with flow<br />

separation. Adjustments to the code like artificial appendices<br />

(virtual stern) or the use of potential flow methods<br />

have helped to improve the predictions significantly. Very<br />

high length Froude numbers seem to produce numerical<br />

instabilities with certain codes or the accuracy decreases.<br />

Thin ship theory still is heavily used and produces good<br />

results <strong>for</strong> many vessel wash problems. There is no limitation<br />

to the distance <strong>from</strong> the vessel at which a wave elevation<br />

can be computed. However, some programs can<br />

only derive entire wave fields and the computation is<br />

restricted to constant water depth. Other effects like surface<br />

tension, wave breaking, and seabed shear stress are<br />

usually neglected. In particular, due to numerical damping<br />

and other numerical inaccuracies, the error can become<br />

too large in the far field to be acceptable.<br />

Most thin ship theory programs can be run on a normal<br />

desktop computer and results are obtained within seconds.<br />

A description of current programs can be found in Gadd<br />

(1999), Molland et al. (2000), Tuck et al. (2001) and<br />

Doctors (1997) as well as more detailed literature.<br />

© COPYRIGHT <strong>PIANC</strong><br />

3.7.1.2 Prediction of wave trans<strong>for</strong>mation and propagation<br />

in the far field and coastal zone<br />

Once the generated waves are no longer influenced by the<br />

vessel, other coastal engineering tools can be used to compute<br />

the wash propagation over a variable bathymetry. To<br />

date only a few successful approaches have been published:<br />

Report of Working Group 41 - MARCOM<br />

14


• Kofoed-Hansen et al. (1996, 1999 and 2000) and MCA<br />

(1998) used a spectral wave model, which describes the<br />

propagation, growth, and decay of short-period waves<br />

in near shore areas to predict the wave height contours<br />

over a variable bathymetry. The model includes the<br />

effects of refraction and shoaling due to varying depth,<br />

wave generation due to wind, and energy dissipation<br />

due to bottom friction and wave breaking. The basic<br />

equations in the model are derived <strong>from</strong> the conservation<br />

equation <strong>for</strong> the spectral wave action density.<br />

• The model boundary allows <strong>for</strong> the variable spectral<br />

<strong>for</strong>ms and permits the wave conditions to vary along the<br />

boundary. For wash studies, waves were specified in<br />

terms of mean wave height, mean wave period, mean<br />

wave direction as well as parameters <strong>for</strong> spreading <strong>from</strong><br />

the mean direction. The code also can be modified to<br />

allow <strong>for</strong> the input of empirical decay rates derived<br />

<strong>from</strong> experiments. A linear refraction-diffraction<br />

model also has been used to calculate the wave disturbance<br />

patterns of the time history as the wave propagates<br />

<strong>from</strong> the vessel to the shore (MCA, 1998).<br />

• Raven (2000) suggested a number of analytical techniques<br />

<strong>for</strong> calculating wave propagation in the far field,<br />

where effects <strong>from</strong> banks or bottom topography can be<br />

neglected. He also demonstrated the successful coupling<br />

of a steady state panel code with a space domain<br />

Boussinesq-type model. A similar approach is discussed<br />

in Kofoed-Hansen et al. (2000). The principle<br />

behind the Boussinesq-type models is to eliminate the<br />

vertical dimension in the flow description without losing<br />

important effects like the influence of the vertical<br />

acceleration on the wave propagation. The non-uni<strong>for</strong>m<br />

distribution of the velocity profile is responsible<br />

<strong>for</strong> the frequency dispersion. The idea is to couple the<br />

two types of models in order to utilise the computational<br />

cost effectiveness of the two-dimensional Boussinesq<br />

model where it is possible, and to apply the detailed,<br />

but time-consuming Navier-Stokes (or Euler/CFD)<br />

based free surface model only in areas where the flow<br />

field is three-dimensional.<br />

© COPYRIGHT <strong>PIANC</strong><br />

To calculate the essential boundary conditions <strong>for</strong> the<br />

Boussinesq model, the full three-dimensional velocity<br />

field and the free surface elevation is required. The free<br />

surface elevation is not generally sufficient as pointed out<br />

by Kofoed-Hansen et al. (2000).<br />

The great advantage of using a Boussinesq-type model is<br />

the possibility of simulating the wash sequence including<br />

both Bernoulli and Kelvin wake. The tool has proven to<br />

be very powerful. However, it still has its restrictions and<br />

requires further development.<br />

• Chen and Sharma (1995) investigated the wave generation<br />

of a slender vessel in a shallow channel at near critical<br />

speed using a combination of thin-ship theory and<br />

simplified Boussinesq equations. In particular the periodic<br />

generation of solitary waves was presented, which<br />

required the use of a non-steady model. The methods<br />

produced good results <strong>for</strong> a shallow channel with variable<br />

cross section. Jiang (2000) combined the thin-ship<br />

approximation with an enhanced Boussinesq method<br />

with improved results, and Jiang et al. (2002) expanded<br />

the model <strong>for</strong> a moving ship, accelerating or decelerating,<br />

in an arbitrary bottom topography, although with<br />

simplified Boussinesq equations.<br />

In general there is some agreement that Boussinesq-type<br />

models will give more realistic results, particularly if the<br />

phase relationship of the successive waves is modelled<br />

correctly. More complex wave models based on<br />

Boussinesq-type equations with a wider applicability in<br />

terms of wavelengths are currently being developed.<br />

Together with improved algorithms and faster desktop<br />

computers these models will be more applicable to a wider<br />

range of cases.<br />

Regardless of what type of numerical model is determined<br />

to be the most appropriate, it is necessary to calibrate it<br />

against either full- or model-scale experiments and then<br />

validate its output with field data be<strong>for</strong>e it can be used as<br />

a practical tool <strong>for</strong> managing wake wash.<br />

3.7.2 Full-scale trials and measurements<br />

3.7.2.1 Measurement techniques<br />

A range of measurement techniques can be used to obtain<br />

wave elevations generated by a vessel. It is beyond the<br />

scope of these guidelines to discuss all the techniques in<br />

detail. An overview of the techniques applied so far is<br />

given and some of the difficulties and criticisms are discussed.<br />

The basic types of point measurement techniques<br />

are listed in Table 3-1. Three different types of techniques<br />

are listed in Table 3-2 that can be used to map a whole<br />

wave field. Not all devices are suitable <strong>for</strong> all locations,<br />

i.e., near field, far field or shoreline. The basic difficulty<br />

is accessibility and range as listed in Table 3-1. The equipment<br />

used to measure the wake should undergo a rigorous<br />

calibration procedure. It is important that the device is<br />

capable of measuring long-period waves (over 20 seconds)<br />

with small amplitude. If possible, validation should be<br />

undertaken with a different type of equipment, particularly<br />

<strong>for</strong> those devices not measuring elevation directly.<br />

15 Report of Working Group 41 - MARCOM


Table 3-1 Single point measurement techniques <strong>for</strong> ship wave measurements<br />

Type: Quantity measured: Technique: Requirements: Constrains: Reference: (examples)<br />

Wave staff Water elevation Capacitance Mounting pole, Shallow water Parnell & Kofoedlogger<br />

only Hansen (2001)<br />

Surveying rod Water elevation Optical Accessibility, Very labour Hannon & Varyani<br />

& camera solid structure intensive (1999)<br />

Altimetry Distance <strong>from</strong> Laser, Solid structure Location Kirk McClure Morton<br />

above surface Ultra-sound, <strong>for</strong> transducer (1998), Koushan<br />

Radar et al. (2001)<br />

Subsurface Water elevation / Echo sound Watertight -<br />

water depth<br />

device<br />

Subsurface Water pressure Piezo transducer Watertight Depth limitation Whittaker (2001),<br />

device <strong>for</strong> short waves Stumbo et al. (1999)<br />

Aqua Particle velocity Acoustic Doppler Watertight device Depth limitation Fissel et al. (2001)<br />

Doppler<br />

<strong>for</strong> short waves<br />

Wave Acceleration Accelerometer Floating buoy Small signal <strong>for</strong> Koushan et al. (2001)<br />

riding Buoy with adequate very long waves<br />

size/weight<br />

Table 3-2 Area measurement techniques (3 dimensional)<br />

Type: Technique: Reference:<br />

Stereo photogrammetry Optical, two pictures taken at different locations Inui (1962).<br />

of same area at same time<br />

LIDAR Fast scanning laser beam across surface Bolt (2001)<br />

and time delay of reflection measured<br />

RADAR<br />

Microwave beam scanning surface<br />

and time delay of reflection measured<br />

3.7.2.2 Monitoring and trial procedures<br />

Full-scale trials can serve two different purposes and as a<br />

result the necessary procedures <strong>for</strong> undertaking such<br />

assessments may differ.<br />

General vessel per<strong>for</strong>mance measurements<br />

Measurement of the wave generation by the vessel and the<br />

wave profile at a given distance <strong>from</strong> the sailing line may<br />

be carried out. In such cases a steady state wave pattern is<br />

of interest. These trials are similar to tank tests and, hence,<br />

are subject to similar procedures. However, it has to be<br />

recognised that conditions are not as controlled as in a<br />

towing tank. Apart <strong>from</strong> the vessel speed over ground,<br />

course over ground, and continuous recording of the vessel's<br />

position after constant time intervals, it is recommended<br />

to record still water trim, overall displacement,<br />

and operation / control of trim flaps, stabilisers, and foils.<br />

It has to be ensured that the vessel has been operating at<br />

constant speed <strong>for</strong> a distance long enough (in particular<br />

close to critical speed). The vessel needs to continue <strong>for</strong><br />

sufficient distance with constant speed after passing the<br />

monitoring location (persistence of wash). It is also desirable<br />

that the seabed is smooth and constant in water depth<br />

along the vessel's track unless deep-water trials are carried<br />

out. The slope of the seabed perpendicular to the vessel’s<br />

track should be uni<strong>for</strong>m. However, constant water depth<br />

would be desirable. Preferably trials should be undertaken<br />

with no or minimal current. In any case a repetition of<br />

Report of Working Group 41 - MARCOM<br />

16


at least one trial in opposite directions is recommended <strong>for</strong><br />

repeatability (assuming the vessel is symmetrical). The<br />

background wave climate, with particular attention to<br />

other vessels’ wake, wind generated waves and swells,<br />

needs to be minimal to reduce post-processing of the data.<br />

<strong>Wake</strong> assessment of a vessel along its operational route<br />

It may be necessary to monitor the wake of a vessel along<br />

a given route <strong>for</strong> risk assessment or numerical modelling<br />

purposes. Due to the variability of the operation, several<br />

passages may need to be monitored to cover different tidal<br />

stages, variable loading of the vessel or different operational<br />

procedures. The monitoring might, there<strong>for</strong>e, be<br />

undertaken over a period of several days. The length of<br />

the monitoring window needs to be sufficient to cover the<br />

full extent of wash (possibly more than 40 minutes). It is<br />

recommended to start monitoring well be<strong>for</strong>e the vessel<br />

passes the monitoring location. The primary reason is that<br />

waves could travel ahead of the vessel or the change of<br />

operation further out causes larger disturbance than the<br />

nearby passage. Monitoring should be continued after the<br />

passage of the concerned vessel until no significant waves<br />

have been measured <strong>for</strong> a considerable time. It is recommended<br />

to record at least still water trim, overall displacement,<br />

and operation of trim flaps, stabilisers, and foils.<br />

3.7.3 Model scale trials<br />

<strong>Guidelines</strong> on tank testing have been published by the<br />

International Towing Tank Conference and are not discussed<br />

in detail in these guidelines. In<strong>for</strong>mation can be<br />

obtained through the Naval Architecture Institutions and<br />

major naval research institutes and departments. Attention<br />

may be drawn to the effects caused by the narrow tank<br />

width in shallow water tests, which are discussed in Inui<br />

(1954). Close to critical speed the towed model is capable<br />

of producing a number of solitary waves depending on the<br />

distance towed at such speed Dand et al. (1999). However,<br />

this has limited practical relevance. The propulsion system<br />

has also significant effect on the wave generation by the<br />

vessel as shown by Taatø et al. (1998). It was found that<br />

the wave amplitude generated by a water jet propelled<br />

model increased by 10 to 40 percent compared to the wave<br />

amplitude generated by a towed model. Several research<br />

groups have used wide shallow water tanks with a tank<br />

width of up to 10 vessel lengths to establish an unreflected<br />

free wave pattern behind the vessel. Again the towing<br />

length needs to be long enough to generate a steady-state<br />

pattern (persistence of waves).<br />

(moreover the event is transient and not periodic). Modern<br />

wavelet analysis might, however, be a good tool <strong>for</strong> mapping<br />

both frequency and time in<strong>for</strong>mation.<br />

• Signal preparation: For the removal of high frequency<br />

noise, a filter method of higher order is recommended.<br />

A more efficient filter technique that does not result in<br />

a phase lag is to make a fast Fourier trans<strong>for</strong>m, cut the<br />

high frequency noise and make an inverse trans<strong>for</strong>m.<br />

Moving averages may also be used, but are less<br />

favourable as the signal is not removed but only<br />

smoothed. Tidal variations or varying offsets may be<br />

best removed with moving averages or high frequency<br />

pass filter.<br />

• Wave cuts in the space domain derived <strong>from</strong> numerical<br />

steady state simulations may be converted into the time<br />

domain.<br />

• Wave-by-wave analysis: Complying with standard<br />

engineering procedures the zero-crossing technique<br />

should be adopted. Thus, the wake can be characterised<br />

in terms of zero-crossing period and zero-crossing<br />

wave height. Otherwise a peak-to-peak analysis can be<br />

used.<br />

3.7.5 <strong>Wake</strong> assessment<br />

The wave generation of the vessel can be assessed using<br />

the wave-making resistance (denoted using either R w or<br />

the wave resistance coefficient C w ) as well as the ratio of<br />

the distribution of resistance R w to the propagation direction<br />

θ (∆R w /∆θ).<br />

As the wave trace is transient, the use of significant wave<br />

height (H 1/3 or H m0 ) and average wave energy (energy per<br />

unit area) is ambiguous. While significant wave height<br />

deceptively suggests an average height throughout the<br />

wash trace, both values depend on the sample length and<br />

become smaller with large sample lengths <strong>for</strong> the same<br />

wash event. In general, using maximum wave height<br />

(H max ) and relating maximum wave period (T max ) and distribution<br />

of wave height and period through the entire<br />

wave trace, e.g., Fig. 3-5, are good measures to characterize<br />

the wake at a given location <strong>for</strong> risk assessment purposes.<br />

© COPYRIGHT <strong>PIANC</strong><br />

4. IMPACTS ASSOCIATED<br />

WITH VESSEL WAKE<br />

3.7.4 <strong>Wake</strong> analysis<br />

Due to the short period of the event and the large spreading<br />

of periods, wave data needs to be analysed in the fulltime<br />

domain on a wave-by-wave basis. Frequency domain<br />

analysis of super-critical wash using Fourier trans<strong>for</strong>ms is<br />

not recommended due to the low-frequency content of<br />

super-critical wash relative to the required sampling time<br />

4.1 OVERVIEW<br />

The objective of this section is to highlight some of the<br />

potential safety and environmental impacts associated<br />

with wake <strong>from</strong> high-speed vessels. Many of these impacts<br />

are based on reported incidents that have been attributed to<br />

17 Report of Working Group 41 - MARCOM


Some examples are discussed in Kofoed-Hansen and<br />

Mikkelsen (1997) and Marine Accident Investigation<br />

Branch report (2000). Although potential impacts associated<br />

with wake <strong>from</strong> high-speed vessels are a legitimate<br />

concern, it is important to note that many of the impacts<br />

discussed in this section are not unique to these vessels.<br />

<strong>Wake</strong> <strong>from</strong> other types of vessels can have the same, or<br />

very similar, impacts. Although every ef<strong>for</strong>t has been<br />

made to identify the most significant impacts, it is possi<br />

ble that there may be others.<br />

The potential <strong>for</strong> wake wash to have an adverse impact on<br />

safety or the environment is related to the physical characteristics<br />

of the waterway and adjacent shoreline as well as<br />

the characteristics of the wake. It is also related to how the<br />

wake interacts with the people and vessels that use the<br />

waterway, the structures that are built on, or near the<br />

waterway, and the near-shore flora and fauna. The implication,<br />

whether wake will have an adverse impact on safety<br />

or the environment, is site specific.<br />

4.2 SAFETY IMPACTS<br />

Safety impacts associated with wake generally involve:<br />

• People on or near the shoreline;<br />

• Vessels underway or moored; or,<br />

• Structures located in, on, or adjacent to the waterway.<br />

4.2.1 Safety of people<br />

<strong>Wake</strong> characteristics in combination with shoreline features<br />

and waterway topography can be used to identify<br />

potential impacts to the safety of people on or near the<br />

shoreline. In general, impacts involving people on or near<br />

the shoreline are primarily associated with transverse<br />

waves generated by vessels operating at or near critical<br />

speed and the long-period waves that comprise the first<br />

group of waves generated by vessels operating at supercritical<br />

speeds. Impacts that can be attributed to these<br />

waves based on shoreline characteristics are summarised<br />

in Table 4-1. Some of these impacts may be of more concern<br />

during certain times of the year, e.g., summer months<br />

when beaches are being used by bathers, or more pronounced<br />

during certain environmental conditions, e.g.,<br />

high or low tide or during calm weather, when large waves<br />

may be unexpected.<br />

4.2.2 Safety of vessels<br />

4.2.2.1 Overview<br />

Just as the characteristics of the wake generated by a vessel<br />

is related to its physical characteristics (e.g., length,<br />

beam, draft, displacement), how another vessel will be<br />

impacted by wake is related to its own physical characteristics<br />

(Bolt, 2002). How a vessel responds to wake also<br />

depends on how it is operated when wake is encountered<br />

and is there<strong>for</strong>e related to the knowledge and skill of the<br />

vessel’s operator. Consequently the specific risk to each<br />

vessel must be assessed individually. This is true regardless<br />

of the type of vessel that generated the wake.<br />

However, there are several general observations that can<br />

be made about wake generated by high-speed vessels or<br />

other vessels operating at or near critical speeds (MCA,<br />

2001):<br />

• All vessels are affected by some part of the near- or<br />

super-critical wash due to the wide range of wave periods<br />

(see Fig. 3-5).<br />

• Small craft are particularly at risk of being swamped,<br />

broached, or capsized by the steep, near breaking waves<br />

produced by vessels operating in the near-critical zone.<br />

© COPYRIGHT <strong>PIANC</strong><br />

Table 4-1 Impacts related to the safety of people based on shoreline characteristics<br />

(Kofoed-Hansen, 1996; MCA, 1998)<br />

Shoreline Characteristics Potential Risk Probable Cause<br />

Shallow sloping beach People may be caught in Rapid inundation of large<br />

water or knocked down areas <strong>from</strong> run up of long-period waves<br />

Moderate / steep People knocked down Plunging and breaking waves<br />

beaches, boat ramps<br />

Damage to boats and<br />

vehicles on ramp<br />

Shorelines with sea walls People on narrow beaches Rapid inundation of exposed beach<br />

trapped against sea wall <strong>from</strong> run up of long-period waves<br />

Breaking waves that overtop the seawall<br />

Report of Working Group 41 - MARCOM<br />

18


• Long period wash waves may cause large vessels to<br />

yaw and alter course, which in a confined channel can<br />

increase the risk of grounding, or when passing other<br />

vessels in close proximity can increase the risk of collision.<br />

• Maximum pitch motions are expected <strong>for</strong> most vessels<br />

when the super-critical wash is encountered as either<br />

head or following waves.<br />

• Maximum heave motions are expected <strong>for</strong> bigger vessels<br />

when the super-critical wash is encountered as<br />

beam waves.<br />

• <strong>Wash</strong> induced roll can increase the risk of grounding at<br />

the turn of the bilge of deep draft, wide beam vessels,<br />

such as container vessels, operating in confined channels<br />

with small under-keel clearances.<br />

• Orbital motions of fluid particles in shallow water are<br />

elliptical rather than circular so that a vessel’s horizontal<br />

motion is larger than its vertical motion.<br />

4.2.2.2 Vessels underway<br />

© COPYRIGHT<br />

<strong>Wake</strong> characteristics, in combination with the topography<br />

of the waterway, can be used to identify potential impacts<br />

to the safety of vessels. In general, impacts related to the<br />

safety of vessels are primarily associated with transverse<br />

waves generated by vessels operating at or near critical<br />

speed as well as the long-period waves comprising the first<br />

<strong>PIANC</strong><br />

group of waves and the short-period waves comprising the<br />

third group of waves generated by vessels operating at<br />

super-critical speeds. Impacts to vessels that can be attributed<br />

to these waves are summarised in Table 4-2. Some of<br />

these impacts may be more pronounced during certain<br />

environmental conditions (e.g., calm weather, high or low<br />

tide, or certain times of year when large numbers of small,<br />

recreational vessels are using a waterway).<br />

4.2.2.3 Vessels on moorings and in harbours<br />

The wave pattern reaching moored vessels is usually more<br />

complicated compared to open water due to the influence<br />

of coastal structures and the local bathymetry. The angle<br />

of propagation is due to wave refraction, reflection, and<br />

diffraction around solid obstacles and, unless the berth is<br />

open to the incoming waves, these changes need to be considered.<br />

The wavelength can be shorter due to diffraction,<br />

which can increase the height at the same time. Potential<br />

impacts of wake on moored vessels are summarized in<br />

Table 4-3.<br />

4.2.3 Structural damage<br />

The effect of wake on floating structures depends on the<br />

response characteristics of the object. Consequently the<br />

risk to each structure must be assessed individually. Of<br />

particular concern is the possibility that the waves will<br />

cause excessive movement of floating structures. There is<br />

also the possibility that the impact of breaking waves may<br />

undercut pier footings.<br />

Table 4-2 Impacts related to the safety of vessels (Kofoed-Hansen, 1996; MCA, 1998)<br />

Waterway Topography Potential Impact Probable Cause<br />

Open water Small craft may be swamped, Waves generated by vessel operating at<br />

broached or capsized<br />

critical speed, or short-period, high-amplitude<br />

waves in the third set of waves<br />

Larger vessels may have<br />

difficulty maintaining course<br />

Long-period waves can cause<br />

larger vessels to yaw<br />

Harbour or estuary Vessels with small under keel Long-period waves result in considerable<br />

entrances with clearances may ground seiching over large areas with shallow water<br />

shallow bars<br />

Small craft may have<br />

difficulty maintaining course<br />

<strong>Wash</strong> generated by vessels operating at<br />

trans- and super-critical speeds can create an<br />

oscillating current over bars and shallow banks<br />

Shallow banks (


Table 4-3 Potential impacts on moored vessels (MCA, 2001)<br />

Activity Potential Impact Probable Cause<br />

Moored vessels Excessive movement Transverse waves generated by vessels operating<br />

(surge, heave and roll) may at or near trans-critical speeds and both<br />

cause moorings to fail, damage long-period and tail waves generated by vessels<br />

to vessels as well as docks and operating at super-critical speeds<br />

piers, vessels to touch bottom<br />

Vessels with small under keel<br />

clearance may ground<br />

Long-period waves result in considerable<br />

seiching over large areas with shallow water<br />

Loading / discharging cargo Disruption of cargo operations Long-period waves generated by vessels operating<br />

that are sensitive to vessel at super-critical speeds<br />

movement<br />

Vessels alongside a moored Damage to either or both vessels Transverse waves generated by vessels operating<br />

vessel (tugs, bunker barges, due to large relative motion due at or near trans-critical speeds and both<br />

etc.) to different response to waves long-period and tail waves generated by vessels<br />

operating at super-critical speeds<br />

4.2.3.1 Historic and archaeological monuments<br />

The possibility of damage to sites of cultural significance,<br />

including historic monuments and archaeological sites, on<br />

the seabed (such as old historical sites and wrecks), and<br />

the shoreline (<strong>for</strong>ts and other historical sites) is prevalent<br />

in several countries including Denmark, New Zealand, and<br />

Sweden. For example, in the archipelago of Göteborg,<br />

Sweden, many bays are sheltered <strong>from</strong> large wind waves<br />

and the tide is minimal. After the introduction of highspeed<br />

ferries a several-hundred-year-old burial ground <strong>for</strong><br />

sailors close to the sea was threatened by wake erosion.<br />

Although the erosion was halted in 1997 when the speed<br />

of the fast ferries was limited, there is evidence suggesting<br />

that wake <strong>from</strong> large container ships may be a heavier<br />

threat (Svensson, 1999).<br />

4.3 ENVIRONMENTAL<br />

IMPACT OF WAKE<br />

© COPY-<br />

4.3.1 Overview<br />

All waves, whether generated by the wind or a vessel, can<br />

have some impact on the marine environment. As indicated<br />

in Section 4.3.2, there are characteristics of high-speed<br />

vessel wake that contribute to its higher propensity vis-àvis<br />

wake <strong>from</strong> other vessels to impact the marine environment.<br />

However, <strong>for</strong> the reasons raised in Section 4.3.4 it<br />

cannot be assumed a priori that high-speed vessel wake is<br />

the cause of any adverse impacts to the marine environment<br />

that might be observed in areas where these vessels<br />

are operating. Determining the cause of adverse impacts<br />

RIGHT<br />

<strong>PIANC</strong><br />

to the marine environment that are observed in a given<br />

area requires detailed, local studies.<br />

The severity of any environmental impact caused by wake<br />

will depend on how the wash regime differs <strong>from</strong> the natural<br />

wave climate. It is also dependent on the susceptibility<br />

of the recipient shores to wave attack. Naturally sheltered<br />

environments and soft sedimentary shores are more<br />

likely to be adversely impacted than naturally exposed<br />

environments with rocky shores.<br />

The available literature on the ecological impacts of highspeed<br />

vessel wake is limited insofar as most studies have<br />

focused on the wash generated by watercraft other than<br />

high speed vessels, i.e. conventional ferries, boats, and<br />

personal water craft (jet skis and wave riders). However,<br />

the studies that have been conducted are useful since they<br />

highlight potential effects of high-speed vessel wake on<br />

the marine environment.<br />

4.3.2 Potential impacts<br />

4.3.2.1 Physical change<br />

In general, an increase in wave action, whether <strong>from</strong> natural<br />

causes (e.g., storm events) or vessel wake will result<br />

in higher energy within the coastal system. This increase<br />

can result in an adjustment to the beach environment<br />

including beach orientation, erosion, accretion, and<br />

increasing the envelope of dynamic change in the attainment<br />

of a new equilibrium to the wave conditions. One<br />

factor influencing the vulnerability of a shoreline to being<br />

changed by waves is its morphological state. In general,<br />

Report of Working Group<br />

41 - MARCOM<br />

20


a stable shoreline is less vulnerable to attack than one that<br />

has been previously disturbed, either because of prior<br />

storm events or human activity (e.g., shoreline construction<br />

or sediment removal). The vulnerability of a coastal<br />

zone to wave attack is also dependent upon its material<br />

composition and typical particle size.<br />

The addition of energy to the system <strong>from</strong> wake can cause<br />

sediment mobilisation and accelerated weathering of<br />

rocky shores. The probable wake related causes of these<br />

impacts are summarised in Table 4-4.<br />

© COPY-<br />

Changes to the physical environment, with particular<br />

emphasis on sediment transport, can cause a variety of different<br />

environmental impacts. Mobilisation of larger sediments<br />

can result in rapid changes to biological communities.<br />

Sedentary organisms may be relocated, crushed, and<br />

damaged as the rocks and boulders on which they are<br />

attached are rolled around. Resuspension of finer sediments<br />

can create an abrasive environment that may damage<br />

soft-bodied animals and algae and prevent spores <strong>from</strong><br />

settling. When sediments do settle out, they can potentially<br />

smother benthic organisms and cover fish eggs and<br />

spawning grounds. Aquatic plants can be physiologically<br />

RIGHT<br />

impaired if their surfaces are covered with silt.<br />

Resuspension of sediments also increases the turbidity of<br />

the water, and by blocking the light that reaches the bottom<br />

can have an adverse impact on benthic organisms.<br />

<strong>PIANC</strong><br />

4.3.2.2 Impacts on flora and fauna<br />

The flora and fauna that inhabit coastal environments are<br />

subject to inundation as well as the hydrodynamic <strong>for</strong>ces<br />

associated with both wind and vessel generated waves.<br />

Waves are a primary agent of ecological disturbance that<br />

affect individuals, populations, and communities to varying<br />

degrees. The potential biological impacts of wake on<br />

coastal and shoreline habitats, which are summarised in<br />

Table 4-5, are a consequence of changes to the natural<br />

physical environment or natural wave climate. Without<br />

external influences, near-shore and intertidal habitats may<br />

exist in a state of long-term equilibrium. Although an<br />

increase in the energy within a coastal system by introducing<br />

vessel wake can be sufficient to upset the equilibrium<br />

(Kirk McClure Morton, 2000), it is usually difficult<br />

to assess changes caused by a changed wave regime due to<br />

vessel wake.<br />

Table 4.4 Physical impacts of vessel wake<br />

(Bell et al., 2000; Kirk McClure Morton, 2000; Single, 2002)<br />

Potential Impact<br />

Probable Cause<br />

Sediment Transport<br />

Sediment resuspension<br />

and associated increase<br />

in turbidity<br />

Cross-shore and long-shore<br />

transport of bottom sediment<br />

and associated accretion<br />

or erosion<br />

Weathering of Rocky Shores<br />

Long-period waves penetrate deeper into the water column than<br />

shorter-period waves and can disturb bottom sediments farther<br />

off-shore than would likely occur under normal conditions.<br />

Large, steep waves similar to storm generated waves can move sediment<br />

seaward; if they attack the beach at an angle, the material will be transported<br />

along the shore. Short, steep waves may transport material onto higher<br />

sections of the beach. Depending on the angle of attack, wake wash<br />

encountering the shore can result in either cross-shore or long-shore transport<br />

of sand and other shoreline material. Large, steep waves, similar to those<br />

generated by storm waves, can move material seaward and <strong>for</strong>m a bar<br />

in front of the beach. If the waves attack the beach at an angle,<br />

the material will be moved along the beach. Short, steep waves can move<br />

material higher up on the beach. <strong>High</strong>-speed vessel wake wash has been<br />

observed to result in both accretion and erosion.<br />

The impact of the incoming wave and the resulting fluctuations in cracks<br />

can separate layers of rock <strong>from</strong> the bedrock. Laboratory measurements suggest<br />

that regular waves, e.g., vessel wake wash, generate higher crack pressures<br />

than mixed waves, e.g., wind generated waves, of similar magnitude.<br />

Hence the weathering of rocky shores by the leading waves of the wake<br />

<strong>from</strong> high-speed vessels may be larger than the weathering induced<br />

by comparable natural seas.<br />

21 Report of Working Group 41 - MARCOM


Table 4-5 Biological impacts of vessel wake<br />

(Danish Maritime Institute, 1997; Bell et al., 2000; Hotchkiss et al., 2002)<br />

Potential Impact<br />

Probable Cause<br />

Fixed and mobile organisms may be dislodged and stranded;<br />

plants may be broken and or uprooted. Overtime species<br />

distribution within the intertidal and near shore zone may change;<br />

some species may be removed<br />

Smothering of plants, sedimentary animals, and spawning grounds<br />

Loss of sea grasses<br />

Altered distribution of species within the intertidal zone<br />

Localized algal blooms and associated increase in turbidity<br />

Reduced productivity and loss of bird nesting areas; change in<br />

feeding behaviour of birds, lower over wintering survival rates<br />

Reduced productivity of seal rookeries; loss of haul-out areas<br />

Waves with higher than ambient energy<br />

due to wake wash breaking on individual<br />

plants and animals. Sediment mobilisation<br />

can damage individual organisms as well as<br />

cause habitat loss.<br />

Settling out of fine, suspended sediment<br />

Turbidity <strong>from</strong> sediment mobilisation can<br />

reduce the amount of light reaching<br />

the bottom<br />

Increased immersion due to regular<br />

inundation of intertidal zone<br />

Resuspension of nutrients associated with<br />

sediment mobilisation<br />

Beach nesting and shoreline feeding areas<br />

inundated by run up of long period waves;<br />

cliff nesting areas inundated by<br />

breaking waves<br />

Inundation by breaking waves<br />

4.3.3 Differentiating between causes<br />

Although high-speed vessels (e.g., high-speed ferries)<br />

have been in operation <strong>for</strong> over ten years, relatively little is<br />

known about the impact of their wake on the marine environment<br />

vis-à-vis the wake <strong>from</strong> other types of vessels.<br />

This is mainly due to the difficulty of distinguishing<br />

between impacts that result <strong>from</strong> different causes including<br />

normal wind generated waves, storms, high-speed and<br />

non high-speed vessel wake, and other human activities.<br />

Making such a distinction is particularly difficult since the<br />

coastal environment naturally is always undergoing some<br />

change (Kirk McClure Morton, 2000).<br />

4.3.4 Predicting environmental impact<br />

© COPY-<br />

Even though most coastal areas are always undergoing<br />

some change, the physical and biological environment of<br />

a beach or a coastline generally is shaped by the predominant<br />

wave climate, which is the product of both regular<br />

wind and storm generated waves as well as vessel wake.<br />

Hence, if the wave climate changes significantly, it is reasonable<br />

to expect that the coastline and its physical and<br />

biological properties will also change, although it is not<br />

always immediately apparent how the change will become<br />

Report of Working Group 41<br />

- MARCOM<br />

RIGHT<br />

manifest. The implication is that any ef<strong>for</strong>t to assess how<br />

high-speed vessel wake may impact the marine environment<br />

will require the comparison of data obtained both<br />

be<strong>for</strong>e and after a particular shoreline is exposed to wake<br />

<strong>from</strong> these vessels. It should be noted that in many cases<br />

the physical impact could be initially large and then<br />

become more stable; there<strong>for</strong>e, any assessed impacts<br />

based on initial trends could possibly result in overestimating<br />

the impact. Although models can be developed to<br />

help predict how wake might impact a particular shoreline,<br />

given the complexity of shoreline systems, model results<br />

should not be generalized to similar shorelines in other<br />

areas (Bell et al., 2000).<br />

5. MANAGING VESSEL WAKE<br />

<strong>PIANC</strong><br />

5.1 INTRODUCTION<br />

Concern about the potential safety and environmental<br />

impact of high-speed vessel wake are causing high-speed<br />

vessel operators and waterway managers to establish<br />

either voluntary or mandatory management measures.<br />

Establishing effective management measures <strong>for</strong> highspeed<br />

vessel wake requires an understanding of the causal


elationship between a vessel’s wake and its actual or<br />

potential impact. As discussed in Sections 3 and 4, establishing<br />

this causal relationship is a multi-faceted problem<br />

that requires understanding how wake interacts with other<br />

vessels, people, structures, and near-shore flora and fauna<br />

as well as the bottom and shoreline of the waterway. This<br />

requires knowledge of the physical characteristics of the<br />

wake generated by the vessel as well as how the wake is<br />

affected by the physical characteristics of the waterway as<br />

it propagates away <strong>from</strong> the vessel’s line of travel and<br />

shoals to break at the shore. It also requires understanding<br />

the effect of the management measures that are considered<br />

on wake generation to ensure that the measures implemented<br />

will be effective and that they will not result in<br />

unintended consequences. The implication is that there is<br />

not a simple, universal solution <strong>for</strong> managing high-speed<br />

wake insofar as how wake interacts with the physical environment<br />

and human activities is site specific. In addition,<br />

there are differences in the wake generated by different<br />

hull <strong>for</strong>ms and sizes and operating speeds – differences<br />

that may permit alternative ways of mitigating the potential<br />

impact.<br />

5.2 MANAGEMENT MEASURES<br />

Mitigation measures can be divided into three categories:<br />

vessel design, operational measures, and non-operational<br />

measures. It is likely that the management regime adopted<br />

<strong>for</strong> any given route will involve a combination of operational<br />

and non-operational measures.<br />

5.2.1 Vessel design<br />

Insofar as the wake generated by a vessel is directly related<br />

to its hull <strong>for</strong>m, vessel design is a primary means of<br />

managing wake. Although it is possible to modify a vessel<br />

after it is constructed (e.g., increasing its length or<br />

installing trim tabs or interceptors) to improve the characteristics<br />

of the wake that is generated, the cost of doing so<br />

may be prohibitive. Similarly, modifications may be prohibited<br />

by regulatory requirements or physical constraints.<br />

There<strong>for</strong>e, naval architects should understand the potential<br />

implications of the wake generated by a given hull design<br />

<strong>for</strong> wake management and consider any wash criteria<br />

related to the intended route including design requirements,<br />

along with emissions, speed, and other limiting factors.<br />

There are, however, key features of high-speed vessel<br />

wake that cannot be reduced or removed by optimising<br />

the hull <strong>for</strong>m and design ratios. An example is the wave<br />

period, which generally increases with vessel speed and<br />

distance <strong>from</strong> the navigation line and is a particularly<br />

important parameter in wave impact in shallow water<br />

areas.<br />

5.2.2 Operational measures<br />

Operational measures can be used to reduce both the safety<br />

and environmental impacts of wake. Because the potential<br />

impacts of wake are related to the physical characteristics<br />

of both the vessel and the waterway, operational<br />

measures are generally related to the route or the vessel’s<br />

operational profile. Insofar as ferry routes are a function<br />

of geography, it is usually not possible to reduce potential<br />

impacts by selecting an alternate route. However, there are<br />

some changes that might be possible, including<br />

• Moving a route farther <strong>from</strong> shore to increase the distance<br />

between the vessel’s track and the area where<br />

wake impacts are of concern.<br />

• Establishing route segments so that changes in water<br />

depth do not cause the vessel to transition <strong>from</strong> subcritical<br />

or super-critical speeds into the critical speed<br />

range.<br />

• Relocating where course or speed changes are made to<br />

avoid focusing wake at a particular location or to avoid<br />

generating wash associated with the critical speed<br />

range.<br />

• Altering the orientation of a route relative to the shoreline<br />

to change the angle at which the wake encounters<br />

an area of concern.<br />

• Modifying the schedule to reduce the potential <strong>for</strong><br />

impacts that may be associated with predicable shoreline<br />

use or environmental factors (e.g., tide or sustained<br />

winds <strong>from</strong> a particular direction).<br />

Although the characteristics of the wake generated by a<br />

vessel are a function of the hull <strong>for</strong>m and cannot be altered<br />

unless the vessel undergoes modifications after it is constructed,<br />

there are several operational measures related to<br />

the vessel’s operational profile that can be used to reduce<br />

the potential impacts of vessel wake. These include<br />

© COPYRIGHT <strong>PIANC</strong><br />

• Training vessel masters and mates so that they understand<br />

the relationship between the navigation of the<br />

vessel and the generation of wake.<br />

• Ensuring that the navigation of the vessel con<strong>for</strong>ms<br />

with the courses and speeds established <strong>for</strong> each leg of<br />

the route.<br />

• Ensuring that the vessel is trimmed on each run to minimize<br />

the wake that is generated.<br />

• Establishing contingency plans <strong>for</strong> situations (e.g., loss<br />

of engine power), or other situations when it may not be<br />

possible to follow normal operating procedures.<br />

23 Report of Working Group 41 - MARCOM


5.2.3 Non-operational measures<br />

The intent of non-operational management measures is to<br />

reduce safety related impacts of wake by lessening the<br />

potential <strong>for</strong> people and small craft to interact directly<br />

with wake. Some non-operational measures include:<br />

• Posting signs on shore or including notices on navigation<br />

charts in areas where high-speed vessel wake<br />

might reasonably be encountered.<br />

• Engaging in outreach activities to ensure the public is<br />

aware of the potential impacts associated with wake.<br />

• Designing new sea walls and quay walls or retro-fitting<br />

existing ones with wave absorbing materials to reduce<br />

wave amplification by reflection. And,<br />

• Coordinating with other operators and harbour authorities<br />

or owners to identify impacts and means of mitigation.<br />

Non-operational measures intended to in<strong>for</strong>m the public<br />

are particularly important in areas where wake <strong>from</strong> highspeed<br />

vessels is not actively managed or is a new activity.<br />

5.3 ROUTE ASSESSMENT<br />

5.3.1 Overview<br />

Developing appropriate management measures <strong>for</strong> highspeed<br />

vessel wake requires identifying the potential safety<br />

and environmental impacts the wash may have. This<br />

requires ensuring that the interrelationship of wake generation,<br />

trans<strong>for</strong>mation, and impact is understood. Since the<br />

impacts that may occur are the result of a number of factors,<br />

including wake generation and trans<strong>for</strong>mation, the<br />

physical characteristics of the waterway and its shoreline,<br />

as well as how the waterway and shoreline are used, this<br />

interrelationship is route dependent. The route assessment<br />

is an objective, systematic process <strong>for</strong> identifying potential<br />

impacts and <strong>for</strong> developing management measures that<br />

may be implemented by vessel operators and waterway<br />

managers that are appropriate <strong>for</strong> a particular route.<br />

© COPYRIGHT <strong>PIANC</strong><br />

A route assessment is intentionally flexible so that it can<br />

be applied to different waterways and different vessels. It<br />

is also flexible with regard to the outcome. In other words,<br />

it does not favour one measure <strong>for</strong> reducing potential wake<br />

impacts over any other. Similarly, whereas in some<br />

instances a single management measure may be sufficient<br />

to prevent an impact <strong>from</strong> occurring, there may be other<br />

cases where a suite of different measures may be needed.<br />

Another aspect of the model is that it is scaleable so the<br />

level of ef<strong>for</strong>t is appropriate <strong>for</strong> the safety and environmental<br />

impacts being considered. The process involves<br />

four steps:<br />

1. Characterizing the route.<br />

2. Identifying the potential impacts and severity of wake.<br />

3. Developing and assessing potential management measures.<br />

4. Implementing and monitoring management measures.<br />

Conducting a route assessment requires an understanding<br />

of specific vessel and route variables. The in<strong>for</strong>mation<br />

needed to conduct the route assessment can be grouped<br />

into five general areas:<br />

1. The characteristics of the vessel’s wake in its entire<br />

operational envelope.<br />

2. The bathymetry and shoreline topography of the proposed<br />

route.<br />

3. Waterway activities along the route.<br />

4. Shore activities along the route.<br />

5. Properties of the physical and biological environment<br />

of the adjacent coastline.<br />

5.3.2 Route characterization<br />

Route characterization involves dividing the planned route<br />

or routes, if alternatives are available, into segments.<br />

Although there are a number of ways this can be accomplished,<br />

the most practical approach is to segment the<br />

route according to where the vessel will be operating at<br />

sub-critical, near-critical, and super-critical speeds based<br />

on the ranges of depth Froude numbers described in<br />

Figure 5-1. In addition, deep-water operation might be<br />

segmented by ‘conventional low speed’ and ‘high speed<br />

sub-critical’ speeds (MCA, 1998). Stumbo et al. (2000)<br />

have observed that in deep water the transition <strong>from</strong> conventional<br />

low speed to high speed sub-critical speed<br />

occurs when vessels have a length Froude number<br />

between 0.65 and 0.9. Using depth Froude numbers, and<br />

length Froude numbers when appropriate, to segment the<br />

route provides a means of conducting a first order assessment<br />

of potential wake impacts along a planned route. It<br />

may identify places where some change to the route (e.g.,<br />

a speed or course change), may eliminate or at least significantly<br />

reduce some of the potential wake impacts.<br />

Report of Working Group 41 - MARCOM<br />

24


Figure 5-1 Operational zones <strong>for</strong> route characterization based on wash regimes<br />

©<br />

5.3.3 Impact identification<br />

Once the route is divided into segments, the type and location<br />

as well as the likelihood and severity of the potential<br />

impacts that wake may have on each segment of the route<br />

must be identified. It is also necessary to identify the<br />

aspect of wake that may cause the impact and then determine<br />

whether wake <strong>from</strong> high-speed vessels is a primary,<br />

secondary, or tertiary cause of the identified impact.<br />

5.3.3.1 Identifying potential impacts<br />

Identifying potential wake impacts requires first ascertaining<br />

who, (e.g., swimmers or people near the shore), or<br />

what, (e.g., small-craft underway or moored in marinas,<br />

near-shore or coastal marine habitats, or waterfront structures)<br />

may be impacted. It also requires determining the<br />

locations where wake impacts can reasonably be expected<br />

to occur. The description of where an impact may occur<br />

should include some reference to the shoreline (e.g., offshore,<br />

near-shore, on-shore), as well as a geographic reference<br />

(e.g., the name of a location or other landmark).<br />

Although latitude and longitude may be useful <strong>for</strong> providing<br />

more precise locations, the use of common names provides<br />

a means to quickly identify areas where wake may<br />

have an adverse impact.<br />

COPY-<br />

RIGHT<br />

Identifying potential wake impacts requires understanding<br />

of how the waterway is used as well as in<strong>for</strong>mation about<br />

the characteristics of the waterway’s natural and physical<br />

environment. There are any number of ways that this can<br />

be accomplished, including site surveys and meeting with<br />

representatives of different waterway users, natural<br />

resource agencies, and land owners. It is likely that the<br />

level of ef<strong>for</strong>t required to identify potential safety and<br />

environmental impacts will vary between routes and<br />

between different segments of the same route. Regardless<br />

of the means and level of ef<strong>for</strong>t employed, it is important<br />

that the result is a complete picture of the type and location<br />

of the potential safety and environmental impacts that<br />

wake may have on each segment of the route.<br />

When identifying potential impacts, it is helpful to establish<br />

when a particular impact may occur. Determining<br />

when an impact might occur is needed <strong>for</strong> determining the<br />

frequency that high-speed vessel wake might cause a<br />

potential impact. It also might help to highlight possible<br />

causal relationships that have to exist <strong>for</strong> an impact to<br />

occur (e.g., state of the tide or sustained winds <strong>from</strong> a<br />

given direction). Several different timeframes can be used<br />

when identifying when an impact might occur. Whereas<br />

some potential impacts might occur multiple times a day,<br />

others might occur infrequently, <strong>for</strong> example, those that<br />

are related to seasonal weather patterns. Establishing the<br />

timeframe might also help highlight impacts that are not<br />

exclusively related to wake <strong>from</strong> high-speed vessels.<br />

<strong>PIANC</strong><br />

Report of Working Group 41 - MARCOM


Table 5-1 Likelihood and consequence scores<br />

Likelihood<br />

Score Description Examples<br />

5 Very Likely Every passage<br />

4 Likely Most passages<br />

3 Quite Possible <strong>High</strong> tide during passage<br />

2 Possible During storms or other periodic<br />

environmental conditions<br />

1 Unlikely Only during unusual /<br />

unpredictable circumstances<br />

Consequence<br />

Score Description Examples<br />

5 Very <strong>High</strong> Unacceptable impact, people may receive<br />

serious injuries or die, small craft cannot<br />

navigate safely, waterfront structures cannot<br />

be occupied / used, marine environment<br />

disrupted<br />

4 <strong>High</strong><br />

3 Moderate Noticeable impact, people may be injured,<br />

moored vessels or waterfront structures may<br />

be damaged, marine environment likely<br />

to be damaged<br />

2 Slight<br />

1 Minimal No noticeable impact, use of shoreline /<br />

waterway not interrupted, any damage<br />

to marine environment is minimal<br />

©<br />

5.3.3.2 <strong>Wake</strong> components<br />

To develop appropriate management measures it is necessary<br />

to determine both the component of wake (i.e., wave<br />

height, wave length, energy, and the magnitude of the<br />

high-speed vessel’s wake) in terms of characteristic wave<br />

height/energy, wave period/length and wave direction that<br />

might be expected to cause each of the potential impacts<br />

that are identified. In some instances there may be more<br />

than one component that might be of concern. Failure to<br />

identify the component, or components, of wake that reasonably<br />

can be expected to cause a potential impact can<br />

result in developing management measures that are not<br />

appropriate.<br />

5.3.3.3 Likelihood and consequence of potential impacts<br />

COPY-<br />

After the potential impacts and the responsible component<br />

of high-speed vessel wake have been identified, it is necessary<br />

to assess both the likelihood that a particular impact<br />

will occur and its potential consequence. Suggested scales<br />

<strong>for</strong> assigning likelihood and consequence scores are<br />

shown in Table 5-1. It is recognized that the descriptions<br />

and definitions <strong>for</strong> the different scores are somewhat subjective.<br />

This is unavoidable since, <strong>for</strong> the most part, there<br />

RIGHT<br />

Report of Working Group 41 - MARCOM<br />

will be limited site-specific data available upon which to<br />

base assessment. There<strong>for</strong>e, the assessment must be based<br />

on the best available in<strong>for</strong>mation and the experience of<br />

those who are conducting it.<br />

<strong>PIANC</strong><br />

26<br />

5.3.3.4 How important is wake<br />

Once the preceding steps have been completed, a determination<br />

should be made whether high-speed vessel wake is<br />

a primary, secondary, or tertiary cause of the potential<br />

impacts that are identified. Making such a determination<br />

is necessary to help provide some basis <strong>for</strong> establishing the<br />

extent to which a high-speed vessel operator should reasonably<br />

be expected to be responsible <strong>for</strong> preventing a<br />

potential impact <strong>from</strong> occurring. Making this determination<br />

will require having an understanding of the impacts<br />

that may be caused by other vessels as well as those that<br />

result <strong>from</strong> natural processes. It will also require an<br />

understanding of the physical differences of waves generated<br />

by other vessels as well as those generated by natural<br />

processes vis-à-vis those generated by high-speed vessels.<br />

As discussed in Section 4.3.3, it is recognized that it may<br />

not always be possible to make this determination with a<br />

high degree of certainty.


5.3.3.5 Priority<br />

Since it may not be reasonable to develop mitigation measures<br />

<strong>for</strong> every potential impact, it is necessary to establish<br />

a prioritised list of those that are identified. An initial prioritised<br />

list can be established based on an index number<br />

(I n ) calculated using the following equation:<br />

L<br />

I n = s x C s<br />

,<br />

CF s<br />

where L s is the likelihood score, C s is the consequence<br />

score and CF s is the causal factor score. CF s values are<br />

provided in Table 5-2. An example of potential management<br />

actions based on the index number is provided in<br />

Table 5-3.<br />

5.3.4 Developing and assessing potential<br />

management measures<br />

<strong>PIANC</strong><br />

© COPYRIGHT<br />

5.3.4.1 Overview<br />

The route assessment is the foundation of any measures<br />

that are developed to managing wake. To be effective, the<br />

management measures that are employed should reduce<br />

the identified impact by<br />

• Focusing on the component of wake that is responsible<br />

<strong>for</strong> the potential impact; and<br />

•<br />

Ensuring the magnitude of the component that is of<br />

concern is maintained within acceptable limits.<br />

In addition, management measures should be appropriate<br />

<strong>for</strong> the location where they will be implemented.<br />

Management measures may be mandated by government<br />

agency regulations, company operating policies, or both.<br />

Establishing management measures should involve agencies<br />

responsible <strong>for</strong> waterways management, vessel operators,<br />

and effected stakeholders.<br />

5.3.4.2 Establishing management measures<br />

The first step in the process of developing a management<br />

regime <strong>for</strong> wake is to determine whether there is adequate<br />

data available to establish a maximum value or standard<br />

<strong>for</strong> the component or components that are of concern that<br />

Table 5-2 Causal factor scores<br />

Causal Factor Score (CF s )<br />

Description<br />

1 <strong>High</strong>-speed vessel wake wash is the primary cause of the impact<br />

2 <strong>High</strong>-speed vessel wake wash is a secondary cause of the impact<br />

3 <strong>High</strong>-speed vessel wake wash is a tertiary cause of the impact<br />

Table 5-3 Index numbers and recommended actions (MCA, 1998)<br />

Index Number (I n )<br />

Recommended Action<br />

1-5 Minimal No mitigation required<br />

6 – 10 Acceptable No mitigation required. Monitoring is required to ensure controls<br />

are maintained.<br />

11 - 15 Moderate Mitigation is required, but the costs of prevention may be taken into<br />

account. Mitigation measures should be implemented within a defined<br />

period. Monitoring is also required. Where moderate risk is associated<br />

with high or very high consequences further assessment may be necessary<br />

to establish more precisely the likelihood of harm as basis <strong>for</strong> determining<br />

the need <strong>for</strong> mitigation.<br />

16 - 20 Significant Vessel should not operate on route until risk has been reduced.<br />

21 -25 Unacceptable Vessel should not operate on route until risk has been reduced.<br />

If it is not possible to reduce risk even with unlimited resources,<br />

the vessel should be prohibited <strong>from</strong> operating on the route.<br />

27 Report of Working Group 41 - MARCOM


cannot be exceeded when measured at a specified location<br />

(e.g., a minimum water depth or distance <strong>from</strong> the shoreline).<br />

The ultimate responsibility <strong>for</strong> making this determination<br />

resides with the agencies responsible <strong>for</strong> waterways<br />

management. However, vessel operators and effected<br />

stakeholders should participate in this process.<br />

Insofar as the route assessment may identify a number of<br />

different wake related impacts that are of concern, a single<br />

standard may not be appropriate <strong>for</strong> the entire route.<br />

There<strong>for</strong>e, to avoid overprotecting some portions of a<br />

route while not providing enough protection along other<br />

portions, it will be necessary to establish different standards<br />

<strong>for</strong> different portions of a route. This is particularly<br />

true if the identified impacts are related to different components<br />

of wake. An advantage of standards is that compliance<br />

can be verified by measuring the wake. However,<br />

it is probable that in many areas sufficient data may not be<br />

available to establish a standard or set of standards. In<br />

these instances, it is necessary to either establish a value<br />

based on the best available data <strong>for</strong> the component of wake<br />

that is of concern or, if the data cannot support a standard,<br />

establish general criteria to minimize the potential that<br />

wake will have an adverse impact on safety or the marine<br />

environment.<br />

Many areas where high-speed vessels have been introduced<br />

have had conventional commercial vessel traffic <strong>for</strong><br />

some time, and the environment is likely to have been<br />

modified by wakes <strong>from</strong> those vessels as well as by<br />

numerous other human activities and natural processes. In<br />

most cases the wakes of conventional vessels generally<br />

cause few complaints, presumably due to familiarity over<br />

time. For such locations the wake <strong>from</strong> conventional commercial<br />

vessels may be used as a reference <strong>for</strong> establishing<br />

a standard <strong>for</strong> high-speed vessel wake.<br />

Despite the fact that wake generated by high-speed vessels<br />

is substantially different <strong>from</strong> wash caused by conventional<br />

ships and ferries, the choice of a standard should reflect<br />

a connection between these two types of wake.<br />

Characteristic measures such as wave energy (Stumbo et<br />

al., 1999), maximum wave height prior to breaking<br />

(Kofoed-Hansen, 1996; Parnell and Kofoed-Hansen,<br />

2001), particle velocity at the seabed, and wave run-up on<br />

the shorelines can be used in a <strong>for</strong>mulation of standard.<br />

Standards based on this approach have been implemented<br />

in Denmark, Sweden, New Zealand, and the United States.<br />

© COPYRIGHT <strong>PIANC</strong><br />

After acceptable standards or criteria have been established,<br />

the next step is to identify an appropriate suite of<br />

management measures. It is likely that this suite will<br />

include both operational and non-operational measures.<br />

Since the measures that may be employed to comply with<br />

the established standards and criteria are linked directly to<br />

the vessel and its operation, vessel operators should have<br />

the primary responsibility <strong>for</strong> identifying specific management<br />

measures. However, it is expected that the management<br />

measures will be subject to review by the agency<br />

responsible <strong>for</strong> waterways management.<br />

Once potential management measures are identified, new<br />

likelihood, consequence, and causal factor scores should<br />

be assigned and the index number recalculated to determine<br />

whether it is within an acceptable level <strong>for</strong> the vessel<br />

to operate on the route in question.<br />

5.3.5 Monitoring<br />

A monitoring program should be developed to determine<br />

whether the management measures that are implemented<br />

are effective in preventing or reducing impacts associated<br />

with wake. Although the details of the monitoring program<br />

are site specific, at a minimum the program should<br />

provide a means of determining or verifying whether wake<br />

<strong>from</strong> high-speed vessels is having adverse impacts on<br />

safety or the marine environment. In some instances, this<br />

will require having access to baseline data so that impacts<br />

that occur over time (e.g., shoreline erosion or habitat<br />

degradation), can be detected. It may also be necessary <strong>for</strong><br />

the monitoring program to provide data that can be used to<br />

differentiate between impacts associated with high-speed<br />

vessel wake and impacts associated with wake <strong>from</strong> other<br />

vessels or natural processes. This is particularly important<br />

<strong>for</strong> impacts where high-speed vessel wake is considered to<br />

be a secondary or tertiary factor.<br />

6. CONCLUSION<br />

The characteristics of wake <strong>from</strong> high-speed vessels are<br />

fundamentally different <strong>from</strong> the wake generated by ‘conventional’<br />

displacement vessels. Although wake <strong>from</strong> any<br />

type of vessel can potentially have adverse impacts on<br />

waterway safety or the marine environment, wake <strong>from</strong><br />

high-speed vessels has received a significant amount of<br />

attention in countries around the world. In addition to<br />

concerns about the potential safety and environmental<br />

impacts of high-speed vessel wake, there is also concern<br />

that ef<strong>for</strong>ts to manage wake might result in the development<br />

of standards that are inappropriate and which might<br />

also unduly restrict the operation of these vessels. As a<br />

result, vessel operators and waterway managers alike have<br />

tried a number of different approaches <strong>for</strong> managing highspeed<br />

vessel wake.<br />

Based on the experience to date <strong>from</strong> these different<br />

ef<strong>for</strong>ts, it is apparent that the effective management of<br />

high-speed vessel wake is a multi-faceted problem that<br />

defies a simple “one size fits all” solution. It is also apparent<br />

that a means of managing high-speed vessel wake that<br />

addresses the legitimate concerns related to waterway<br />

safety and protection of the marine environment while also<br />

not unduly restricting their operation must be identified <strong>for</strong><br />

the full benefits of high-speed vessels to be realized.<br />

Waterway managers and high-speed vessel operators are<br />

encouraged to follow the guidelines <strong>for</strong> conducting a route<br />

assessment and developing management measures that are<br />

outlined in this report.<br />

Report of Working Group 41 - MARCOM<br />

28


REFERENCES<br />

Allenström et al., 2003. “The interaction of large and highspeed<br />

vessels with the environment in archipelagos - Final<br />

report.” SSPA Research Report No 122, Göteborg,<br />

Sweden.<br />

Anonymous, 1999. “Speed restriction fails to halt<br />

<strong>Wash</strong>ington State Ferries.” Fast Ferry International,<br />

38(8):22-25.<br />

Anonymous, 2000. “Early problems <strong>for</strong> Dutch service.”<br />

Fast Ferry International, 39(1):47.<br />

Bell, A., Elsaesser, B., and Whittaker, T. J. T., 2000.<br />

“Environmental Impact of Fast Ferry <strong>Wash</strong> in Shallow<br />

Water.” Hydrodynamics of <strong>High</strong> Speed Craft <strong>Wake</strong> <strong>Wash</strong><br />

& Motions Control, London, 1-14.<br />

Bolt, E., 2001. “Fast Ferry <strong>Wash</strong> Measurement and<br />

Criteria.” Fast 2001 The 6th International Conference on<br />

Fast Sea Transportation, Southampton, 135-148.<br />

Croad, R. and Parnell, 2002. “Proposed Controls on shipping<br />

activity in the Marlborough Sound - A review under<br />

s32 of the Resource Management Act.” Report to the<br />

Marlborough District Council, September 2002. Available<br />

at http://www.marlborough.govt.nz/documents.html<br />

Cain, C., 2000. “<strong>Wake</strong> <strong>Wash</strong> - An Operators Viewpoint -<br />

Passage Plans & Risk Assessment,” Hydrodynamics of<br />

<strong>High</strong> Speed Craft <strong>Wake</strong> <strong>Wash</strong> & Motions Control. The<br />

Royal Institution of Naval Architects, London, pp. 1-16.<br />

Chen, X.-N. and Sharma, S.D., 1995. “A Slender Ship<br />

Moving at a Near-critical Speed in a Shallow Channel.”<br />

Journal of Fluid Mechanics, 291: 263-285.<br />

Dand, I.W., Dinham-Peren, T.A. and King, L., 1999.<br />

“Hydrodynamic Aspects of a Fast Catamaran Operating in<br />

Shallow Water”, Hydrodynamics of <strong>High</strong> Speed Craft. The<br />

Royal Institution of Naval Architects, London, pp. 1-17.<br />

© COPYRIGHT <strong>PIANC</strong><br />

Danish Maritime Authority, 1997. Report on the Impact<br />

of <strong>High</strong>-Speed Ferries on the External Environment (in<br />

Danish).<br />

Demirbilek, Z., and Vincent, L., 2002. “ Water waves<br />

mechanics.” Coastal Engineering Manual, Part II,<br />

Hydrodynamics, Chapter II-1, L. Vincent, ed., U.S. Army<br />

Corps of Engineers, <strong>Wash</strong>ington, DC., 1 - 115.<br />

Doctors, L.J., 1997. "Resistance Predictions <strong>for</strong> Transomstern<br />

Vessels.” Fast 2001 The 6th International Conference<br />

on Fast Sea Transportation, Southampton.<br />

Eggers, K.W.H., Sharma, S.D. and Ward, L.W., 1967. “An<br />

Assessment of Some Experimental Methods <strong>for</strong><br />

Determining the Wavemaking Characteristics of a Ship<br />

Form.”<br />

Ekman, V.W. 1906., “On Stationary Waves in Running<br />

Water.” Arkiv för Matematik, Astronomi och Fysik, Vol. 3,<br />

No. 2, Stockholm, pp. 1-30.<br />

Feldtmann, M. and Garner, J., 1999. “Seabed<br />

Modifications to Prevent <strong>Wake</strong> <strong>Wash</strong> <strong>from</strong> Fast Ferries,<br />

Coastal Ships and Inland Waterways.” Royal Institution of<br />

Naval Architects, London.<br />

Fissel, D., Billenness, D., Lemon, D., and Readshaw, J.,<br />

2001. “Measurement of the Wave <strong>Wash</strong> Generated by Fast<br />

Ferries with Upward Looking Sonar Instrumentation.”<br />

17th International Fast Ferry Conference, New Orleans<br />

<strong>USA</strong>.<br />

Gadd, G.E., 1994. “The <strong>Wash</strong> of Boats on Recreational<br />

Waterways.” Transactions of the Royal Institution of<br />

Naval Architects.<br />

Gadd, G.E., 1999. “Far Field Waves Made by <strong>High</strong> Speed<br />

Ferries,” Hydrodynamics of <strong>High</strong> Speed Craft. The Royal<br />

Institution of Naval Architects, London, pp. 1-9.<br />

Hannon, M. A., and Varyani, K. S., 1999. “The <strong>Wash</strong><br />

Effect of <strong>High</strong> Speed Ferries in Coastal and Inland<br />

Waterways.” International Conference on Coastal Ships<br />

and Inland Waterways, London, 1-12.<br />

Havelock, T.H., 1908. “The Propagation of Groups of<br />

Waves in Dispersive Media, with Application to Waves on<br />

Water produced by a Travelling Disturbance.”<br />

Proceedings of the Royal Society of London, Series A, Vol.<br />

LXXXI, pp. 398-430.<br />

Hotchkiss, S., Braithwaite, R., Fletcher, B., Elsaesser, B.,<br />

and Whittaker, T. J. T., 2002. “Assessing the<br />

Environmental Impact of Fast Ferry <strong>Wash</strong> on Rocky Shore<br />

Communities.” Ship Design and Operation <strong>for</strong><br />

Environmental Sustainability, London, 1-13.<br />

Hughes, M., 2001. “CFD Prediction of <strong>Wake</strong> <strong>Wash</strong> in<br />

Finite Water Depth.” HIPER'01 2nd International<br />

EuroConference on <strong>High</strong> Per<strong>for</strong>mance Marine Vehicles,<br />

Hamburg, pp. 200-211.<br />

<strong>High</strong>-Speed Craft Code, 2000. “International Code of<br />

Safety <strong>for</strong> <strong>High</strong>-Speed Craft.” International Maritime<br />

Organization, London.<br />

Inui, T., 1954. “Wave-Making Resistance in Shallow Sea<br />

and in Restricted Water, with Special Reference to its<br />

Discontinuities.” Journal of the Society of Naval<br />

Architects of Japan, 76: 1-10.<br />

29 Report of Working Group 41 - MARCOM


Jiang T., 2000. “Investigation of Waves Generated by<br />

Ships in Shallow Water.” In Twenty-Second Symposium on<br />

Naval Hydrodynamics. The National Academies Press,<br />

<strong>Wash</strong>ington, D.C., United States.<br />

Jiang, T., Henn, R. and Sharma, S.D., 2002. “<strong>Wash</strong> waves<br />

generated by ships moving on fairways of varying topography.”<br />

In Twenty-Forth Symposium on Naval<br />

Hydrodynamics, Fukuoka Japan 8-13 July 2002.<br />

Kirk McClure Morton, 2000. “Investigation of Coastal<br />

Processes in Loch Ryan.” Loch Ryan Advisory<br />

Management Forum, Dumfries.<br />

Kirk, R. M. and M. B. Single, 2000. “Coastal Effects of<br />

New Forms of Transport: the Case of the Interisland Fast<br />

Ferries.” Chap. 21 in Environmental Planning and<br />

Management in New Zealand, eds. P.A. Memon and H.C.<br />

Perkins , Palmerston North, NZ: Dunmore Press Ltd.<br />

Kirkegaard, J., Kofoed-Hansen, H., and B. Elfrink, 1998.<br />

“<strong>Wake</strong> <strong>Wash</strong> of <strong>High</strong>-Speed Craft in Coastal Areas.” In<br />

Proceedings of the 26th International Coastal Engineering<br />

Conference, 22-26 June 1998, Copenhagen, Denmark.<br />

Kofoed-Hansen, H., 1996. “Technical Investigation of<br />

<strong>Wake</strong> <strong>Wash</strong> <strong>from</strong> Fast Ferries Summary & Conclusions.”<br />

5012, Danish Hydraulic Institute, Hørsholm.<br />

Kofoed-Hansen, H., Jensen, T., Kirkegaard, J. and Fuchs,<br />

J., 1999. “Prediction of <strong>Wake</strong> <strong>Wash</strong> <strong>from</strong> <strong>High</strong>-Speed<br />

Craft in Coastal Areas,” Hydrodynamics of <strong>High</strong> Speed<br />

Craft. The Royal Institution of Naval Architects, London,<br />

pp. 1-10.<br />

Kofoed-Hansen, H., Jensen, T., Sørensen, O.R. and Fuchs,<br />

J., 2000. “<strong>Wake</strong> <strong>Wash</strong> Risk Assessment of <strong>High</strong>-Speed<br />

Ferry Routes - A Case Study and Suggestions <strong>for</strong> Model<br />

Improvements.” The Royal Institution of Naval Architects,<br />

London.<br />

Kofoed-Hansen, H. and Mikkelsen, A.C., 1997. “<strong>Wake</strong><br />

<strong>Wash</strong> <strong>from</strong> Fast Ferries in Denmark.” In Proceedings of<br />

the 4th International Conference on Fast Sea<br />

Transportation, Sydney, Australia.<br />

© COPYRIGHT <strong>PIANC</strong><br />

Koushan, K., Werenskiold, P., Zhao, R., and Lawless, J.,<br />

2001. “Experimental and Theoretical Investigation of<br />

<strong>Wake</strong> <strong>Wash</strong>.” Fast 2001 The 6th International Conference<br />

on Fast Sea Transportation, Southampton, 165-179.<br />

Lighthill, J., 1978. “Waves in Fluids.” Cambridge<br />

University Press, Cambridge.<br />

Marine Accident Investigation Branch, 2000. “Report on<br />

the Investigation of the Man Overboard Fatality <strong>from</strong> the<br />

Angling Boat PURDY at Shipwash Bank, off Harwich on<br />

17 July 1999.” MAIB Report No. 17/2000.<br />

Maritime and Coastguard Agency (MCA), 1998.<br />

“Research Project 420 - Investigation of <strong>High</strong> Speed Craft<br />

on Routes Near to Land or Enclosed Estuaries.” Maritime<br />

and Coastguard Agency.<br />

Maritime and Coastguard Agency (MCA), 2001.<br />

“Research Project 457 - A Physical Study of Fast Ferry<br />

<strong>Wash</strong> Characteristics in Shallow Water.” Final Report,<br />

Maritime and Coastguard Agency.<br />

Mitchell, J.H., 1998. “The Wave Resistance of Ships.”<br />

Philosophical Magazine, 45 (Series 5): 106-123.<br />

Molland, A.F., Wilson, P.A. and Chandraprabha, S., 2000.<br />

“The Prediction of Ship Generated Near-Field <strong>Wash</strong><br />

Waves Using Thin Ship Theory,” Hydrodynamics of <strong>High</strong><br />

Speed Craft. The Royal Institution of Naval Architects,<br />

London, pp. 1-13.<br />

Nanson, G. C., Von Krusenstierna, A., and Bryant, E. A.,<br />

1994. “Experimental Measurements of River-Bank<br />

Erosion Caused by Boat-Generated Wave on the Gordon<br />

River, Tasmania.” Regulated River: Research &<br />

Management, 9:1-14.<br />

Newman, J. N., 1977: “Marine Hydrodynamics”. The<br />

MIT Press, Cambridge, Massachusetts, pp. 270-278.<br />

Parnell, K. 1996. “Monitoring Effects of Ferry <strong>Wash</strong> in<br />

Tory Channel and Queen Charlotte Sound.” Unpublished<br />

report to Marlborough District Council, April 1996.<br />

Parnell, K. E. and H. Kofoed-Hansen, 2001. “<strong>Wake</strong>s <strong>from</strong><br />

Large <strong>High</strong>-Speed Ferries in Confined Coastal Waters:<br />

Management Approaches with Examples <strong>from</strong> New<br />

Zealand and Denmark,” Coastal Management, Vol. 29, pp.<br />

217 – 237.<br />

Peltoniemi, H., A. Bengston, J. Rytkönen and T. Kõuts,<br />

2002. “Measurements of Fast Ferry Waves in Helsinki -<br />

Tallinn Run.” The Changing State of the Gulf of Finland<br />

Ecosystem Symposium, Tallinn, Estonia.<br />

Lewis, E. V., 1988. “Principles of Naval Architecture,”<br />

Vol. II: Resistance, Propulsion and Vibration. Society of<br />

Naval Architects and Marine Engineers, Jersey City, New<br />

Jersey.<br />

Pickrill, R.A. 1985 “Beach Changes on Low Energy Lake<br />

Shorelines, Lakes Manapouri and Te Anau, New Zealand.”<br />

Journal of Coastal Research, Vol 1, No 4 pp 353-363.<br />

Report of Working Group 41 - MARCOM<br />

30


Raven, H.C., 2000. “Numerical <strong>Wash</strong> Prediction Using a<br />

Free-surface Panel Code”, Hydrodynamics of <strong>High</strong> Speed<br />

Craft <strong>Wake</strong> <strong>Wash</strong> & Motion Control. The Royal Institution<br />

of Naval Architects, London, pp. 1-12.<br />

Sandwell Engineering, Inc., 2000. British Columbia Ferry<br />

Corp. Fast Ferry Program - <strong>Wake</strong> and <strong>Wash</strong> Project.<br />

Scott-Russell, J., 1965. “The Modern System of Naval<br />

Architecture,” London,Vol.1.<br />

Single, M.B. and Kirk, R.M., 1999. “Coastal Change and<br />

Human Processes in Tory Channel, Marlborough Sounds.”<br />

in P.C. Forer and P.J. Perry (eds) Proceedings of the 18th<br />

New Zealand Geography Society Conference 1995.<br />

Single, M. B., 2002. “Effects on the Shoreline of Fast<br />

Ferries in Tory Channel, New Zealand.” Proceedings of<br />

the 30th <strong>PIANC</strong> Congress, Sydney, Australia.<br />

Ström, K. and F. Ziegler, 1998. “Environmental Impacts of<br />

<strong>Wake</strong> <strong>Wash</strong> <strong>from</strong> <strong>High</strong> Speed Ferries in the Archipelago of<br />

Göteborg.” Environmental Office, Göteborg (In Swedish).<br />

Stumbo, S., Fox, K., Dvorak, F., and Elliot, L., 1999. “The<br />

Prediction, Measurement, and Analysis of <strong>Wake</strong> <strong>Wash</strong><br />

<strong>from</strong> Marine Vessels,” Marine Tech., Vol. 36, No. 4, pp.<br />

248 – 260.<br />

Stumbo, S., Fox, K., and Elliot, L., 2000. “An Assessment<br />

of <strong>Wake</strong> <strong>Wash</strong> Reduction of Fast Ferries at Supercritical<br />

Froude Numbers and at Optimised Trim.” RINA, The<br />

Royal Institution of Naval Architects, London.<br />

Svensson, U., 1999. “Environmental Impacts of Large and<br />

Fast Ships.” SSPA, Maritime Consulting Report, No<br />

984798-01 (in Swedish).<br />

Taatø, S.H., Aage, C. and Arnskov, M.M., 1998. “Waves<br />

<strong>from</strong> Propulsion Systems of Fast Ferries.” 14th Fast Ferry<br />

International Conference, Copenhagen.<br />

© COPYRIGHT <strong>PIANC</strong><br />

Tuck, E.O., Scullen, D.C. and Lazauskas, L., 2001. “Ship-<br />

Wave Patterns in the Spirit of Michell.” In: Y.<br />

Shikhmurzaev (Editor), IUTAM Conference on Free<br />

Surface Flows. The University of Birmingham,<br />

Birmingham.<br />

Whittaker, T.J.T., Doyle, R. and Elsaesser, B., 2000. “A<br />

Study of the Leading Long Period Waves in Fast Ferry<br />

<strong>Wash</strong>,” Hydrodynamics of <strong>High</strong> Speed Craft <strong>Wake</strong> <strong>Wash</strong><br />

& Motions Control. RINA, The Royal Institution of Naval<br />

Architects, London.<br />

Whittaker, T.J.T., Doyle, R. and Elsaesser, B., 2001. “An<br />

Experimental Investigation of the Physical Characteristics<br />

of Fast Ferry <strong>Wash</strong>.” In: V. Bertram (Editor), HIPER'01<br />

2nd International EuroConference on <strong>High</strong> Per<strong>for</strong>mance<br />

Marine Vehicles, Hamburg, pp. 480-491.<br />

Vincent, L., Bratos, S., Demirbilek, Z. and Weggel, J.R.,<br />

2002. “Estimation of Nearshore Waves.” L. Vincent<br />

(Editor), Coastal Engineering Manual, Part II,<br />

Hydrodynamics, Chapter II-3. Engineer Manual 1110-2-<br />

1100. U.S. Army Corps of Engineers, <strong>Wash</strong>ington, DC.,<br />

pp. 1-34.<br />

31 Report of Working Group 41 - MARCOM


ANNEX A<br />

TERMS OF REFERENCE<br />

The introduction of high-speed ferries in the mid-nineties<br />

has led to wake conditions not known earlier.<br />

The vessels in question are characterised by a travelling<br />

speed of 35 to 45 knots. They may be catamaran types but<br />

can also be single hull vessels.<br />

According to observations by the general public in<br />

Denmark, the waves created by the wake are more dangerous<br />

than the waves created by traditional ferries. They<br />

are higher and they may appear without warning and suddenly<br />

rise in shallow waters, then break and cause run-up<br />

on the coast. Danish newspapers and television news have<br />

shown dangerous situations <strong>for</strong> small boats and <strong>for</strong><br />

bathers, and large numbers of complaints and protests<br />

<strong>from</strong> coastal house owners and landowners, <strong>from</strong> fishermen,<br />

anglers, sailors, rowers, hunters, divers, swimmers/holidaymakers<br />

were received by the authorities.<br />

Anxieties were also voiced regarding coastal erosion, subsidence<br />

of natural stone reefs and disturbance of bird sanctuaries<br />

and dike safety.<br />

In 1995 the Danish Maritime Authority commissioned the<br />

Danish Hydraulic Institute to carry out a technical investigation<br />

of wake wash <strong>from</strong> fast ferries.<br />

This investigation concluded that<br />

• the new high speed vehicle ferries generate waves,<br />

which, in coastal and shallow water regions, are considered<br />

by the general public as being considerably different<br />

to the waves caused by conventional vessels.<br />

• this view has been widely confirmed by the measurements<br />

and calculations which have been made.<br />

The report also mentions that wake wash problems have<br />

appeared in Sweden, Ireland, Australia, New Zealand,<br />

England, Portugal, and U.S.A. In Denmark regulations<br />

were issued in 1998 concerning the effects of waves <strong>from</strong><br />

high-speed ferries on leisure activities, on the environment<br />

in general and in particular on erosion and dike safety<br />

problems.<br />

• categorize problems met<br />

• analyse in which ways such problems may be <strong>for</strong>eseen<br />

and avoided<br />

• gather in<strong>for</strong>mation on regulations established in various<br />

countries and regions<br />

• coordinate study with IMO working groups on problems<br />

in relations to high-speed ferries;<br />

all in order to set up guidelines <strong>for</strong> traffic planners, ship<br />

owners, and relevant public authorities, including waterway<br />

managers.<br />

© COPYRIGHT <strong>PIANC</strong><br />

<strong>PIANC</strong> started in 1995 a working group that deals with<br />

harbour facilities <strong>for</strong> high-speed ferries. The task <strong>for</strong> this<br />

new working group is there<strong>for</strong>e limited to high-speed ferries<br />

in motion. This working group was requested to<br />

• gather in<strong>for</strong>mation internationally about experiences<br />

with high-speed ferries<br />

Report of Working Group 41 - MARCOM<br />

32

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!