Download as pdf or txt
Download as pdf or txt
You are on page 1of 191

1

PORT WORKS MANUAL


Design, Construction and Maintenance

Civil Engineering Department

Hong Kong Government First published, December 1992. Second Edition, December 1996. Prepared by : Civil Engineering Office, Civil Engineering Department, Civil Engineering Building, 101 Princess Margaret Road, Homantin, Kowloon, Hong Kong. This publication is available from : Government Publications Centre, Ground Floor, Low Block, Queensway Government Offices, 66 Queensway, Hong Kong. Overseas orders should be placed with : Publications (Sales) Office, Information Services Department, 28th Floor, Siu On Centre, 188 Lockhart Road, Wan Chai, Hong Kong. Price in Hong Kong : HK$50 Price overseas : US$12 (including surface postage) An additional bank charge of HK$50 or US$6.50 is required per cheque made in currencies other than Hong Kong dollars. Cheques, bank drafts or money orders must be made payable to HONG KONG GOVERNMENT.

() 101

66

188 28 50 12() 506.5

FOREWORD

The Port Works Manual offers general guidance on the design, construction and maintenance of marine structures in Hong Kong. The Manual was first published in English in December 1992 after consultations with practitioners. Since then it has been widely circulated as a useful reference for port engineering works in the territory. This second edition presents the original English version and its Chinese counterpart in a parallel format. The opportunity has been taken to rectify minor typographic errors in the first edition and to up-date the environmental data and basic characteristics of vessels. These data were respectively provided by the Royal Observatory, Marine Department and local ferry operators whose constributions are gratefully acknowledged. This second edition was prepared by Mr Leung Kin-man, Mr Ma Pak-fai and Mr Tso Man-lun, staff of the Civil Engineering Office under the supervision of Mr Shiu Wing-yu. The Chinese translation work was assisted by Mr Jiang Ju-yao and Ms Li Ling-ling of the Tianjin Research Institute of Water Transport Engineering.

Practitioners are encouraged to offer comments at any time to the Civil Engineering Office on the contents of this Manual, so that improvements can be made to future editions.

C C Chan Principal Government Civil Engineer

CONTENTS

Page No.

TITLE PAGE FOREWORD CONTENTS 1. INTRODUCTION 1.1 1.2 2 Scope Definitions and Symbols

2 3 5 11 11 11 13 13 13 14 14 16 16 17 17 17 17 17 18 18 19 19 19 19 20 21 21 24 24 24 26 27 27 29 29 29

OPERATIONAL CONSIDERATIONS 2.1 2.2 2.3 2.4 2.5 2.6 General Design Life Ship Data Approach Channels Berthing Conditions Currents

3.

ENVIRONMENTAL DATA 3.1 3.2 General Winds 3.2.1 Collection Stations 3.2.2 Mean Hourly Wind Speeds 3.2.3 Mean Wind Speeds for Durations Exceeding One Hour 3.2.4 Mean Wind Speeds for Durations Less than One Hour 3.2.5 Maximum Gusts 3.2.6 Pictorial Summaries of Wind Direction and Speed Waves 3.3.1 General 3.3.2 Ship Observations 3.3.3 Wave Recording and Analysis 3.3.4 Wave Prediction from Local Wind Records 3.3.5 Transformation of Offshore Wave Characteristics 3.3.6 Selection of Wave Parameters Tides and Water Levels Currents Joint Probability Wave Overtopping

3.3

3.4 3.5 3.6 3.7 4.

LOADS 4.1 4.2 General Loading Conditions and Combinations

Page No. 4.2.1 Normal Loading Conditions 4.2.2 Extreme Loading Conditions 4.2.3 Temporary Loading Conditions 4.2.4 Accident Loading Conditions Dead Loads Superimposed Dead Loads Live Loads 4.5.1 Determination of Imposed Live Loads 4.5.2 Determination of Continuous Live Loads Hydrostatic Loads Soil Pressures Temperature Variations Tides and Water Level Variations Winds Currents 4.11.1 Steady Drag Forces 4.11.2 Flow-induced Oscillations Waves 4.12.1General 4.12.2 Design Wave Parameters 4.12.3 Calculation of Average Maximum Wave Height 4.12.4 Depth-limited Situations 4.12.5 Calculation of Wave Forces in General 4.12.6 Wave Forces for Reflective Conditions 4.12.7 Wave Forces Using Morison's Equation 4.12.8 Wave Uplift Pressures Berthing 4.13.1 General 4.13.2 Assessment of Energy to Be Absorbed 4.13.3 Berthing Reactions Mooring Earthquakes Movements and Vibrations Morison 30 31 32 33 34 34 34 34 35 36 36 37 37 39 40 41 41 41 41 42 43 44 45 45 46 49 49 49 50 51 51 52 53 55 55 56 56 56 59 59 59 60

4.3 4.4 4.5

4.6 4.7 4.8 4.9 4.10 4.11

4.12

4.13

4.14 4.15 4.16 5.

DESIGN OF FOUNDATIONS 5.1 5.2 5.3 5.4 Introduction Site Investigations Properties of the Ground Piled Foundations

6.

DESIGN OF SUSPENDED DECK STRUCTURES 6.1 6.2 6.3 Introduction Load Combinations and Factors Superstructure

Page No. 6.4 6.5 Piles Durability 6.5.1 Reinforced and Prestressed Concrete 6.5.2 Steelwork 60 60 60 61 63 63 63 65 65 65 65 66 66 67 67 67 69 69 69 70 71 71 72 75 75 76 77 78 78 79 79 81 83 83 83 84 87 87 87 87 87

7.

DESIGN OF SHEET PILED STRUCTURES 7.1 7.2 General Corrosion Protection

8.

DESIGN OF GRAVITY STRUCTURES 8.1 8.2 General Concrete Blockwork Walls 8.2.1 General 8.2.2 Ground Water Levels and Profiles 8.2.3 Consideration of Settlement

9.

DESIGN OF RUBBLE STRUCTURES 9.1 9.2 9.3 General Design Wave Stability 9.3.1 General 9.3.2 Design of Armour Units Using Hudson's Formula 9.3.3 Structure Head Conditions 9.3.4 Model Testing 9.3.5 Design of Armour Units Using Van Der Meer's Formulae Crest Level

Hudson Van der Meer

9.4 10.

DESIGN OF RECLAMATIONS 10.1 10.2 10.3 10.4 General Extent and Layout Reclamation Level Reclamation Method 10.4.1 General 10.4.2 Marine Deposit Removal 10.4.3 Marine Deposit Displacement 10.4.4 Controlled Thin Layer Fill Placement Miscellaneous 10.5.1 General 10.5.2 Piling 10.5.3 Culvert Foundations

10.5

11.

CONSTRUCTION MATERIALS 11.1 11.2 11.3 General Armour Rock Fill 11.3.1 General

Page No. 11.3.2 Public Dump Material 11.3.3 Selected Fill 11.3.4 Crushed Rock 11.3.5 Marine Fill Concrete 11.4.1 Reinforced and Prestressed Concrete in General 11.4.2 Durability 11.4.3 Unreinforced Concrete 11.4.4 Underwater Concrete 11.4.5 Design Steel 11.5.1 Structural Steel in General 11.5.2 Corrosion Protection 11.5.3 Use of Stainless Steel 11.5.4 General Guidance Timber 11.6.1 Types of Material 11.6.2 Design Stress 11.6.3 Loading Factors Rubber Piles 11.8.1 General 11.8.2 Driven Concrete Piles 11.8.3 Tubular Steel Piles 11.8.4 Bored Piles 11.8.5 Fender Piles 11.8.6 Sheet Piles Protective Measures 11.9.1 General 11.9.2 Life of Protective Coatings 11.9.3 Important Points to Be Considered 11.9.4 Corrosion Protection of Steel Tubular Piles 88 90 91 91 91 91 92 93 94 95 95 95 95 96 96 97 97 97 98 100 100 100 100 101 101 102 102 102 102 103 104 105 107 107 107 108 111 111 112 113 114 114 114 114 115 116

11.4

11.5

11.6

11.7 11.8

11.9

12.

TYPES OF STRUCTURE 12.1 12.2 12.3 12.4 General Breakwaters Seawalls Piers 12.4.1 General 12.4.2 Public Piers 12.4.3 Ferry Piers Dolphins Pumphouses 12.6.1 General 12.6.2 Layout and Location 12.6.3 Structure and Design 12.6.4 Ties and Waterstops

12.5 12.6

Page No. 12.6.5 Screens, Guides and Fittings Slipways and Ramps 12.7.1 General 12.7.2 Location and Basic Dimensions 12.7.3 Slipway Design 12.7.4 Ramp Design 12.8 Navigation Aids 12.9 Outfalls and Intakes 12.10 Miscellaneous 12.7 13. CONSTRUCTION 13.1 13.2 General Dredging 13.2.1 General 13.2.2 Preparation and Execution of Works 13.2.3 Sampling of Dredged Materials 13.2.4 Surveys for Dredging 13.2.5 Dumping of Dredged Material Breakwaters and Seawall Foundations 13.3.1 General 13.3.2 Preparation and Execution of Filling 13.3.3 Surveys for Fill Materials 13.3.4 Tolerances for Fill Materials Concrete Blockwork Walls 13.4.1 Levelling Stones and Blocks 13.4.2 Bermstones 13.4.3 Facing Stones and Copings Piers and Dolphins 13.5.1 Preparation of Works 13.5.2 Piling 13.5.3 Fendering 13.5.4 Works by Others Reclamations 13.6.1 Sequence of Reclamation 13.6.2 Precautions to Be Taken During Reclamation Underwater Blasting Material Inspection and Testing 13.8.1 General 13.8.2 Marine Fill 13.8.3 Rock Fill 13.8.4 Rock Armour 13.8.5 Timber for Fenders 13.8.6 Rubber Fenders Completion of Works 13.9.1 General 13.9.2 Completion Certificates 13.9.3 As-constructed Drawings 117 118 118 118 118 120 120 121 123 125 125 125 125 125 126 128 130 130 130 130 132 132 133 133 133 134 134 134 135 136 136 137 137 137 138 138 138 139 139 140 140 140 141 141 141 141

13.3

13.4

13.5

13.6

13.7 13.8

13.9

10

Page No. 14. MAINTENANCE 14.1 Maintenance Inspections 14.1.1 Routine Inspections 14.1.2 Special Inspections 14.1.3 Inspection Procedures Dredging 14.2.1 General 14.2.2 Sampling and Surveys 14.2.3 Maintenance Dredging Adjacent to Structures 14.2.4 Maintenance Dredging of Rivers and Nullahs Piers and Dolphins 14.3.1 Piles 14.3.2 Decks 14.3.3 Fendering Systems 14.3.4 Steps and Landings 14.3.5 Miscellaneous Items Blockwork Seawalls 14.4.1 General 14.4.2 Repairs to Rubble Mounds 14.4.3 Repairs to Concrete Blocks 14.4.4 Repairs to Granite Facing 14.4.5 Repairs to Concrete Coping Rubble Seawalls and Breakwaters Ramps and Slipways Pumphouses Navigation Aids Estimates of Future Maintenance 14.9.1 General 14.9.2 Estimates for Dredging 14.9.3 Estimates for Maintenance of Piers 14.9.4 Estimates for Maintenance of Seawalls 14.9.5 Estimates for Maintenance of Breakwaters

143 143 143 143 143 144 144 144 145 145 146 146 148 150 152 152 154 154 155 155 156 156 157 158 159 159 159 159 160 160 161 161 163 169 171 173 187 189 191

14.2

14.3

14.4

14.5 14.6 14.7 14.8 14.9

REFERENCES TABLES List of Tables Tables FIGURES List of Figures Figures

11

1. INTRODUCTION

1.

1.1

Scope

1 . 1

This Manual offers guidance on the design, construction and maintenance of those marine works and structures which are normally constructed by the Hong Kong Government. Such works and structures include breakwaters, seawalls, public piers, ferry piers, dolphins, reclamations, pumphouses, slipways, ramps and navigation aid supports. However, it should also provide a source of useful data and design guidance for other marine works in Hong Kong. The recommendations in this Manual are for guidance only. Because of the nature of the loads involved, the design of marine structures relies particularly on the use of sound engineering judgement and experience.

1.2

Definitions and Symbols

1 . 2 BS6349: Part 1 (BSI, 1984a) 1.3 Shore Protection Manual (CERC, 1 9 8 4 ) BS6349: Part 1

The definitions given in BS 6349:Part 1 (BSI, 1984a), Section 1.3 apply. The Glossary of Terms given in the Shore Protection Manual (CERC, 1984) Volume II, Appendix A may also be useful for reference. Where symbols are not defined in the text, the standard meanings for the symbols given in BS 6349:Part 1 and other quoted references, as appropriate to the context, apply.

12

13

2. OPERATIONAL CONSIDERATIONS

2.

2.1

General

2 . 1

This Chapter gives guidance on general aspects such as the design life of structures, ship data, approach channels and other operational considerations. Appropriate advice should be obtained from the Director of Marine, Commissioner for Transport, other concerned Government Departments as appropriate, and the ferry operators, on all operational matters.

2.2

Design Life

2 . 2 50

The design life of a structure is taken to be its intended useful life, and will depend on the purpose for which it is required. The choice of design life is a matter to be decided in relation to each project. Unless special circumstances apply, the design life for all permanent marine structures covered by this Manual should be taken to be 50 years. Design life and return period are not the same and should not be confused. If the return period for an event is numerically equal to the design life, then there is about a 63% chance of the event occurring within the design life. Recommended return periods are covered in other sections. The design life is significant when assessing :

6 3 %

(a) (b) (c)

(a)

time dependent factors affecting the stability of the structure, such as fatigue loading, corrosion, marine growth and soil strength reductions,

(b) probability levels for limit state design and for design condition return periods, and (c) economic feasibility of the project and future developments.

14

2.3

Ship Data

2 . 3

Where possible, details and dimensions should be obtained from the Director of Marine, the client, owners and operators of the vessels to be accommodated, and those likely in the anticipated lifetime of the structure. Vessel characteristics which should be considered include type, size and shape, ship handling requirements, cargo or passenger handling requirements, and vessel servicing requirements. Basic characteristics of local vessels taken from the Local Craft Registry provided by the Director of Marine are given in Table 1. Basic characteristics of the ferries owned by three of the major ferry operators at the time of going to press are given in Tables 2 to 4. All values should be checked with the Director of Marine or the ferry operators as appropriate before being used for design purposes. Information on other vessels using Hong Kong as a port of call should be sought from the appropriate authorities when required.

1 2 4

2.4

Approach Channels

2 . 4

The depth and width of approach channels should be specified or approved by the Director of Marine. The required depth can be calculated taking into account the following factors : (a) the design vessel loaded draft,

(a) (b) (c) (d) (e) (f)

(b) the tide, (c) vessel squat,

(d) vessel pitching and rolling, (e) (f) vessel trim, and an empirical factor giving an under-keel clearance to facilitate manoeuvrability, economic propeller efficiency and a factor of safety.

15

The required channel width depends upon the following factors : (a) the beam, speed and manoeuvrability of the design vessel,

(a) (b) (c) (d) (e) (f)

(b) whether the vessel is to pass another vessel, (c) the channel depth,

(d) the channel alignment, (e) (f) the stability of the channel banks, and the winds, waves, currents and crosscurrents in the channel.

The above factors are covered in detail by Bruun (1981), the National Ports Council (1981) and the Department of Transport (1982a, b & c). Where the bottom of the channel consists of mud it is usual in international ports to define the depth for navigation as being that between low water level and the level at which the density of the bottom sediment is equal to or greater than 1200 kg/m3, since research elsewhere has shown that mud layers of lower density do not significantly impede the passage of a ship. It should be noted that, depending on its operating frequency, an echo sounder will identify muds of significantly lower density at the seabed, whereas a sounding lead will sink until supported by muds of greater density than 1200 kg/m3. This aspect is covered in more detail in other chapters, but it should be noted that such measurement of density requires specialist equipment which may not be readily available. When planning the location of approach channels, and approaches or fairways in general, account should be taken of future siltation and maintenance. Dredging should be carried out to a depth greater than the minimum required navigation depth, with the intention of eliminating the need for maintenance dredging for at least two and preferably five years after completion of initial dredging.

Bruun (1981), National Ports Council (1981), Department of Transport (1982a b c ) 1200 kg/m3 1200 kg/m3

16

2.5

Berthing Conditions

2 . 5 B S 6 3 4 9 : P a r t 1 ( B S I , 1 9 8 4 a ) 30 31 Binnie & Partners (1987)

Acceptable wave conditions at berths for ferries and public vessels or within cargo handling basins and typhoon shelters can only be determined after consultation with the Director of Marine, and ferry and other vessel operators. Guidance on acceptable wave conditions for moored vessels is given in BS 6349:Part 1 (BSI, 1984a), Sections 30 and 31, and some suggestions on port operation criteria have been made by Binnie & Partners (1987).

2.6

Currents
2 . 6

Major sea defence, reclamation or dredging work may cause changes in the pattern of tidal flow and consequently affect navigation, mooring and berthing forces, siltation and water quality in the vicinity of the reclamation or dredging site, and possibly some distance from the site. During planning of the project, advice should be sought from the Civil Engineering Department and Environmental Protection Department on whether detailed mathematical or physical modelling studies will be necessary.

17

3. ENVIRONMENTAL DATA

3.

3.1

General

3 . 1

This Chapter gives information related to the environmental records available for Hong Kong conditions with regard to winds, waves, currents and sea levels, and gives guidance on the assessment of extreme values. Five day means of meteorological elements for Hong Kong from 1961 to 1990 are given in Table 5. These have been taken from Surface Observations in Hong Kong 1990 (RO, 1990).

5 Surface Observations in Hong Kong 1990 (RO, 1990)

3.2

Winds

3.2 3 . 2 . 1 1 Chin & Leong (1978)

3.2.1 Collection Stations The locations of the stations at which the Observatory has made wind observations for 10 years or more are shown in Figure 1. For details of these stations, covering such aspects as date of first observation, type of anemometer and height above mean sea level, reference may be made to Chin & Leong (1978).

3.2.2 Mean Hourly Wind Speeds Mean hourly wind speeds in metres/second for return periods of 5, 10, 20, 50, 100 and 200 years for four of the main stations, the Observatory, Kai Tak Airport (SE), Waglan Island and Cheung Chau are given in Tables 6 to 9. The assessment was carried out by the Observatory by applying Gumbel's method to the annual maximum mean hourly wind speeds for the eight directions for each of the stations. The period of records used for each station is given in Tables 6 to 9. The figures in Table 6 should be used with care because of the effects of urbanisation on winds after 1959; reference may be made to Chin & Leong (1978) for details of this effect.

3 . 2 . 2 6 9 5 1 0 2 0 5 0 1 0 0 2 0 0 m / s Gumbel 6 9 6 Chin & Leong (1978)

18

The mean hourly wind speeds given in Tables 6 to 9 may be considered for design purposes to have been corrected to the standard height of 10 m above mean sea level. No corrections have been applied to the speeds recorded at the Observatory for the reasons stated by Chin & Leong, although the recording height varies from 47 to 72 m above mean sea level, or to the Kai Tak Airport (SE) speeds, with a recording height of 10 to 16 m above mean sea level. Although the Observatory station has the longest series of wind observations, the data are by no means homogenous due to the building development and changes in the instrumentation and height of the anemometer over the years. Hence, the wind information for the Observatory station should be used with due care and discretion. Minor corrections to the Waglan Island and Cheung Chau speeds, which have recording heights of 75 and 92 m above mean sea level respectively, have been made in line with the guidelines given by Chin & Leong, following advice from the Observatory. It should be noted that the normal wind-height adjustment formulae, including the power law referred to by the Shore Protection Manual (SPM) (CERC,1984), are not recommended by the Observatory for use in Hong Kong conditions.

6 9 10 47 72 Chin & Leong 1016 Chin & Leong 7 5 9 2 S h o r e P r o t e c t i o n M a n u a l ( SPM) (CERC, 1984)

3.2.3 Mean Wind Speeds for Durations Exceeding One Hour Mean wind speeds for durations of 2, 3, 4, 6 and 10 hours and return periods of 5, 10, 20, 50, 100 and 200 years for Waglan Island NE, E, SE, S and SW directions and for Cheung Chau SE, S and SW directions are given in Tables 10 to 17. The assessment was carried out by the Observatory, applying Gumbel's method to the annual means for each duration and direction. Again, the wind speeds given can be considered for design purposes to have been corrected to the standard height of 10 metres above mean sea level.

3 . 2 . 3

10 17 510 2 0 5 0 1 0 0 2 0 0 2 3 4 610 Gumbel 10

3.2.4 Mean Wind Speeds for Durations Less than One Hour For conversion of the mean hourly wind speeds

3 . 2 . 4

19

to mean speeds corresponding to durations of less than one hour, the following conversion factors are recommended for design purposes : Duration 1 minute 5 minutes 20 minutes 1 hour Conversion Factor 1.19 1.11 1.05 1.00

1 5 20 1

1.19 1.11 1.05 1.00

3.2.5 Maximum Gusts For information on maximum gusts, reference may be made to Chen (1975) and Poon (1982). Additional useful information concerning the frequency and duration of tropical cyclone warning signals for the period from 1946 to 1990 is given by the RO (1991).

3 . 2 . 5 Chen (1975) Poon (1982) (RO, 1991)

3.2.6 Pictorial Summaries of Wind Direction and Speed Pictorial summaries of the frequency distribution of wind direction and speed measurements, in the form of wind roses, are given for the four stations referred to in Section 3.2.2 on an annual basis in Figure 2.

3 . 2 . 6

2 3.2.2

3.3

Waves

3 . 3 3.3.1 B S 6 3 4 9 : P a r t 1 ( B S I , 1 9 8 4 a ) 21.121.221.3 SPM 2 - 3 2 - 11

3.3.1 General General notes on wave characteristics and properties are given in BS 6349:Part 1 (BSI, 1984a), Sections 21.1, 21.2 and 21.3, and the SPM, pages 2-3 to 2-11. Estimates of extreme wave conditions at a site should ideally be obtained by extrapolating a series of wave measurements made at or close to the site. However, because of the relatively high cost of setting up a wave recording system, and the need for records covering a suitable period (not less than one year) to enable sufficiently reliable extrapolation, no direct wave record will be available for the vast

20

majority of marine structures covered by this Manual. In Hong Kong waters, the most severe wave conditions are usually associated with storm waves and, in the absence of wave records, wave forecasting from wind records can be used to predict such conditions, as outlined in later sections. In some situations, particularly where there is direct exposure to the South China Sea and longer period waves are therefore considered important, swell waves from distant storms should be taken into account during design. If no direct wave records are available, information on swell waves can be obtained from existing wave records elsewhere or from visual wave observations made by ships' crews. Because of the complex geographical features in Hong Kong waters, waves propagating into such waters are likely to be transformed by processes such as refraction, diffraction, reflection and seabed friction. These processes may have significant influence on the wave climate in the area to be studied. The designer has to assess these factors at an early stage to ascertain whether more sophisticated analysis has to be carried out. Computer models are available for such analysis and are recommended for use in studying the wave transformation in complex areas. These models have to be calibrated to make sure that they are suitable for that particular study area.

3.3.2 Ship Observations It is generally accepted that predictions based on a large number of observations from ships made by different people can give a useful estimate of wave conditions. Records of ship observations within the area of the South China Sea bounded by longitudes 100E and 120E and by latitudes 0N and 25N are kept by the Observatory. Reference may be made to Lam (1979, 1980), although it should be noted that the areas covered by these studies include some relatively protected inshore regions. If information on swell waves is required from ship observations for a particular project, an open area of sea should be considered when approaching the Observatory for

3 . 3 . 2 1 0 0 o 1 2 0 o 0 o 25o Lam (1979, 1980) Hogben, Dacunha & Olliver (1986)

21

details of records held. Wave data for the South China Sea can also be obtained from Hogben, Dacunha & Olliver (1986). Offshore wave climate has been derived from ship observation records given in Binnie & Partners (1989). Data were taken from the sea area 18.8N to 21.9N and 112.0E to 115.0E, comprising some 24,000 observations from 1949 to 1985. The data were divided into 30 directional sectors centred upon 60, 90, 120, 150, 180, 210, 240 and 270. The eight resulting distributions of significant wave height have been extrapolated to extremes using the Weibull fitting.
Binnie & Partners (1989) 112.0o 115.0o 18.8o 21.9o 24,000 30o 6 0 o 9 0 o 1 2 0 o 1 5 0 o 180o210o240o 270o Weibull

3.3.3 Wave Recording and Analysis For general information on wave recording and analysis, reference may be made to BS 6349:Part 1, Section 26. Wave records at Waglan Island have been kept by the Observatory since 1971. For details, reference may be made to Apps & Chen (1973), Chen (1979a, 1979b) and Cheng (1986). It should be noted from the latest of the above publications that there were several recording system breakdowns when waves reached heights of nine to ten metres and that no return period/wave height analysis was carried out due to the limited number of records available for analysis. The return period/wave height information given by Chen (1979b) should not be used without checking with the Observatory concerning the latest wave records. Details of wave height observations made at Waglan Island between 1959 and 1966 are given by Cuming (1967).

3 . 3 . 3 BS6349: Part 1 26 Apps & Chen (1973)Chen (1979a1979b) Cheng (1986) Cheng (1986) 910 Chen (1979b) C u m i n g ( 1 9 6 7 )

3.3.4 Wave Prediction from Local Wind Records For wave prediction from local wind records, it is recommended that the charts given in the SPM should be used : Figure 3-23 should be used for deep-water waves and Figures 3-27 to 3-36 for

3 . 3 . 4

SPM 3 - 23 3 - 273 - 36 37 SPM

22

shallow-water waves. Figures 3 to 7 of this Manual are based on Figures 3-23, 3-30, 3-32, 3-34 and 3-36 of the SPM, but with the range of wind stress factor extended up to 100 m/s to allow for local conditions. When using these figures, the significant wave period can be taken as 0.95 times the peak spectral period for design purposes, see the SPM, page 3-47. The estimation of fetch length for the direction being considered should be based on the method given in Section V, page 3-42 of the SPM, which considers a 24 sector centred on that direction, and involves the arithmetic averaging of a number of equally spaced radials within that sector. Before converting any wind speed from Tables 6 to 9 or Tables 10 to 17 to a wind stress factor by applying equation 3-28a of the SPM, the application of correction factors for stability and location in accordance with page 3-30 of the SPM is required. For the prediction of waves generated by tropical storms, a stability correction factor of 1.1 may be used for air/sea temperature adjustment. A location correction factor of 1 is recommended, as all anemometer stations considered are relatively close to shore.

3 - 233 - 303 - 323 - 343 - 36 1 0 0 m / s 0.95SPM 3 47

SPM 3 . 4 2 V 24o SPM3 - 28a6 9 10 17 SPM3 - 30 1.1 1

For 'fetch-limited' situations, mean hourly wind speeds from Tables 6 to 9 may be used for initial wave prediction, after correction for stability and location but, for final wave prediction, the corrected mean hourly wind speed should be adjusted using the conversion factor appropriate for the actual critical duration, as read from the wave prediction curves. Conversion factors for durations of less than one hour are given in Section 3.2. Linear interpolation between the figures given will be sufficiently accurate for other durations. As an example, for a corrected mean hourly wind speed of 48 m/s and a fetch of 2 km, the corresponding wind stress factor of 83.0 m/s gives a deep-water significant wave height of 1.9 m and a peak spectral period of 3.4s (significant wave period 3.2s) from Figure 3, with a critical duration of 20 min. As this critical duration is different from one hour, the mean hourly wind speed must be adjusted for this duration. From Section 3.2, the conversion factor for a 20 minute duration is 1.05. For the

6 9 3 . 2 48 m/s 2 83.0 m / s 3 1 . 9 3 . 4 3.220 3.2 2 0 1 . 0 5 50.4 m/s 2

23

corrected mean 20 minute wind speed of 50.4 m/s and the same fetch of 2 km, the corresponding wind stress factor of 88.2 m/s gives a deep-water significant wave height of 2.0 m and a peak spectral period of 3.5s (significant wave period 3.3s). For durations greater than one hour, the conversion factor for the adjustment of the mean hourly wind speed can be assessed from Tables 10 to 17.

88.2 m/s 2 . 0 3 . 5 3.3 1 0 17

For fetch-limited situations, several directions should be considered, as the most severe wave conditions will not necessarily correspond with the direction which has the longest fetch. Mean hourly wind speeds for directions other than those given in Tables 6 to 9 should be interpolated by an appropriate method. For situations which are not fetch-limited, a series of durations and corresponding mean wind speeds, taken from Tables 10 to 17, should be considered for each direction, before the most critical wave conditions can be determined. The mean wind speed/duration distributions given in Tables 10 to 17 differ for different locations and directions. When using Figure 3 for wave prediction from the wind speed/duration figures for some directions, it will be found that there is a maximum off-shore significant wave height corresponding to a particular duration less than about ten hours, and within the range of two to six hours. However, for other directions, it will be found that the wave height continues to increase as the duration increases from 2 to 25 hours, and there is no apparent maximum. For durations greater than about 10 hours, particular care should be taken when assessing wave heights. It is recommended that the Observatory is consulted in such circumstances, and other methods of wave prediction and assessment should be used. The use of Figure 3 for wave prediction is based on the assumption that winds measured in Hong Kong waters apply equally with regard to strength, duration and direction over the respective fetch being considered. This assumption is reasonable for relatively short durations of several hours, which correspond to fetches of less than about 100 km for 100 year return period winds. However, for durations exceeding ten hours, which correspond

6 9

1 0 17 10 17 3 2 5 3 100 100 100 200

24

to fetches greater than about 200 km for 100 year return period winds, there is increasing doubt about this assumption, particularly as such extreme winds will be associated generally with tropical storm conditions.

3.3.5 Transformation of Offshore Wave Characteristics For the transformation of offshore wave characteristics into inshore wave characteristics, covering such aspects as refraction, shoaling and diffraction, reference should be made to BS 6349:Part 1, Sections 23.2 and 29 and the SPM, pages 2-60 to 2-108. Breaking waves are covered by BS 6349:Part 1, Section 23.4, and the SPM, pages 2-129 to 2-136. Further useful comments on breaking and nonbreaking waves and design wave conditions are given in the SPM, pages 7-1 to 7-16. When investigating whether a structure will be subjected to breaking or nonbreaking waves, all possible sea levels should be considered. For some structures in relatively shallow water, breaking waves at low sea levels will be critical, whereas nonbreaking wave conditions will apply at higher sea levels.

3 . 3 . 5

B S 6 3 4 9 : Part 1 2 3 . 2 2 9 S P M 2 - 602 - 108BS6349: Part 1 2 3 . 4 S P M 2 - 1 2 9 2 - 1 3 6 SPM7 - 1 7 - 16

3.3.6 Selection of Wave Parameters Guidance on the selection of design wave parameters for particular types of structure is given in other sections. Wave height parameters in terms of significant wave height are given in the SPM, page 72. Notes on the average maximum wave height and its relationship with the significant wave height, and the number of waves in the duration of the design condition, are given in BS 6349:Part 1, Section 27.3.2, and also in Section 4.12 of this Manual.

3 . 3 . 6 S P M 7 - 2 B S 6 3 4 9 : P a r t 1 2 7 . 3 . 2 4.12

3.4

Tides and Water Levels

3 . 4 8

The locations of tide gauges under the control of the Observatory at the time of going to press are given in Figure 8. Tide Tables are published each year by the Observatory. These give predicted times and

25

heights of high and low waters at the tide stations. Tide levels are given in terms of Chart Datum (CD), which is 0.146 m below Principal Datum (PD).

(CD) (PD) 0.146

Useful information on storm surges in Hong Kong during the period from 1906 to 1982, and the assessment of extreme sea levels for the tide gauge locations at North Point (since replaced by Quarry Bay in 1986), Tai Po Kau and Chi Ma Wan for various return periods, is given by Chan (1983). Updated extreme sea level analyses have been carried out by the Observatory for Quarry Bay/North Point, Tai Po Kau, Ko Lau Wan, Tsim Bei Tsui and Waglan Island. Extreme sea levels for return periods of 2, 5, 10, 20, 50, 100 and 200 years for these five locations are given in Tables 18 to 22. The period of records used in each case is given in the Tables. At each location, the assessment was carried out by fitting a Gumbel distribution to the annual maximum sea levels and using the method of moments in parameter estimation. Lowest still levels observed at the tide stations shown in Figure 8 are as follows : Station North Point Quarry Bay Chi Ma Wan Ko Lau Wan Lok On Pai Tai O Tamar Tsim Bei Tsui Tai Po Kau Waglan Island Sea Level (mCD) -0.16 -0.12 -0.13 -0.13 -0.28 -0.52 -0.17 -0.21 -0.26 -0.17

Chan (1983) 18 22 2 5 1 0 2 0 5 0 1 0 0 200 Gumbel

8 (mCD) -0.16 -0.12 -0.13 -0.13 -0.28 -0.52 -0.17 -0.21 -0.26 -0.17

Probable minimum sea levels at North Point have been estimated by the Observatory using Gumbel's method, and are as follows :

Gumbel

26

Return Period (years) 2 5 10 20 50 100 200

Sea Level (mCD) 0 -0.10 -0.15 -0.20 -0.30 -0.35 -0.40

2 5 10 20 50 100 200

(mCD) 0 -0.10 -0.15 -0.20 -0.30 -0.35 -0.40

3.5

Currents

3 . 5 Admiralty Tidal Stream Atlas, Hong Kong (Taunton, 1975)

Indications of current speeds for Hong Kong waters are given in the Admiralty Tidal Stream Atlas, Hong Kong (Taunton, 1975). Additional information on currents in specific locations can be obtained from reports prepared by consultants for various Government departments over the last few years. The Civil Engineering Department, Territory Development Department, Environmental Protection Department and Marine Department should be consulted in the first instance for details of studies carried out in any areas for which information on currents is required. It is possible to predict representative currents within Hong Kong waters using the WAHMO suite of models held by Government. The computerbased mathematical models cover the whole of Hong Kong waters and include simulation of the two-layer effect in the wet season due to the Pearl River flow. In addition, the physical tidal model may be used for predictions inside Victoria Harbour for the dry season. For locations where no information on existing currents is available, it may be necessary to carry out measurements on site. General notes on current meter observations and float tracking are given in Sections 11.2.2 and 11.2.3 of BS 6349: Part 1. The period of measurement should depend on flow characteristics in the area of interest and aspects of the study for which the current data is needed. It should also be set in such a way as to give an indication of velocity distribution with

(WAHMO M o d e l s )

BS6349: Part 1 11.2.211.2.3

27

depth below water level, and of direction as in some areas, stratification may be significant in the wet season.

3.6

Joint Probability

3 . 6

Some of the worst conditions that should be considered in relation to the design of coastal defence works are those which involve the combination of extreme sea level with severe wave action. This combination does not necessarily mean adding the predicted extreme still water level directly to the separately estimated extreme wave height value. There is a need to determine the degree of correlation between these two variables so that an appropriate choice can be made in relation to each type of structure in the light of the purpose for which it is being designed. The astronomical tide is considered an independent factor while the meteorological factors of storm surge and wave attack are linked. The degree of correlation can vary between different sites. Joint probability analysis is therefore recommended to establish such correlation, especially for large projects for which the cost implications can be very significant.

3.7

Wave Overtopping

3 . 7 F u k u d a , U n o & I r i e ( 1 9 7 4 ) G o d a ( 1 9 7 1 )

Assessment of the amount of overtopping is needed to determine the crest level. As guidance, the permissible volumes of overtopping water can be found in Fukuda, Uno & Irie (1974) and Goda (1971). However, the designer should examine each individual case to decide whether more conservative figures or otherwise, have to be adopted. The factors to be considered should include consequences of excessive overtopping, drainage behind the seawall and reliability of design parameters. The amount of overtopping can be estimated by various methods. Two of the most commonly used ones are detailed in Goda (1971) and Hydraulics Research Station (1980). These methods have been

Goda (1971) Hydraulics Research Station (1980)

28

based on results of model tests using irregular waves. It is recommended that model tests should be carried out as far as is possible to verify the predicted amount of overtopping.

29

4. LOADS

4.

4.1

General

4 . 1

This Chapter describes the loading conditions which should be considered in the design of marine structures (excluding rubble structures, which are covered in Chapter 9) and includes information on the loads to be taken into account. Guidance is given on the selection of relevant design parameters and methods of calculation to derive the resulting direct forces on structures, taking into account the nature and characteristics of the structures. In addition to dead loads, superimposed dead loads, hydrostatic loads and soil pressures, the other forces which may act on marine structures are environmental, arising from such natural phenomena as winds, temperature variations, tides, currents, waves and earthquakes, and those imposed loads due to operational activities. General imposed loads cover live loads from pedestrians, vehicles, cargo storage and handling. Vessel imposed loads cover berthing, mooring and slipping. Unless stated otherwise, the design loads given in this Chapter are unfactored and can be used directly where appropriate in checking geotechnical instability modes, in design of piled foundations, see Section 5.4, and of steel structures using the working stress method of design, see Section 6.3. Guidance on appropriate partial factors for limit state design is to be found in other sections.

5 . 4 6.3

4.2

Loading Conditions and Combinations

4 . 2

The structure as a whole, or any part or section, should be designed and checked for at least the loading conditions given below. If it is expected that other loading conditions could be critical, they should also be investigated. Various types of load should be combined in a manner consistent with the probability of their simultaneous occurrence.

30

4.2.1 Normal Loading Conditions These loading conditions are those in which normal operations continue unaffected by environmental conditions. A combination of the following should be considered : (a) dead loads,

4 . 2 . 1

(a) (b) (c)

(b) superimposed dead loads, (c) live loads due to normal working operations (the most severe arrangement likely to occur simultaneously),

(d) vessel imposed loads (berthing and mooring), (e) normal environmental loads (winds, currents and waves), soil pressures, and

( d ) ( e )

(f)

(f) (g)

(g) hydrostatic loads. Guidance on the calculation of environmental loads associated with normal working operations is given later in this Chapter under each type of loading condition. It should be assumed that maximum imposed live loads can occur simultaneously with maximum vessel imposed loads from either berthing or mooring, whichever gives the most severe effect, unless the size or geometry of the structure indicates that it is possible for certain mooring loads to occur at the same time as berthing. In this latter case, the most severe combination of berthing and mooring loads should be determined by the designer and this combination assumed to occur simultaneously with maximum imposed live loads. For normal environmental loads, it should be assumed that maximum loads from winds, currents and waves can occur simultaneously. All directions should be considered when assessing the most severe effects from these loads. When combining the effects of maximum imposed loads due to normal working operations with

31

maximum normal environmental loads, the possible shielding effect caused by a vessel when berthing or mooring should be considered. For certain structures, e.g. dolphins, it may not be realistic to assume that full berthing or mooring loads can occur simultaneously with full wind, current and wave loads.

4.2.2 Extreme Loading Conditions

4 . 2 . 2

These loading conditions are associated with the most severe environmental conditions which the structure is designed to withstand. It is assumed that under these conditions most normal operations, such as vessel berthing and mooring, pedestrian and vehicle movements, and cargo storage and handling, will have ceased. A combination of the following should be considered : (a) dead loads (same values as for Normal Loading Conditions),

(a)

(b) superimposed dead loads (these may be different from Normal Loading Conditions), (c) reduced live loads (if any at all) due to continuing operations,

(b) (c) (d) (e) (f) (g)

(d) reduced vessel-imposed loads (if any) due to continuing operations, (e) extreme environmental loads (winds, currents, waves and temperature variations), soil pressures (same values as for Normal Loading Conditions), and

(f)

(g) hydrostatic loads (in some cases, these will be different from Normal Loading Conditions).

It should be assumed for extreme environmental loads that maximum effects from winds, currents and waves can occur simultaneously, but maximum effects from temperature variations should be considered separately. Vessel-imposed loads can be ignored under

32

extreme environmental conditions from winds, currents and waves, as these will occur during tropical cyclone conditions when normal vessel movements will have ceased. However, vessel-imposed loads should be combined with maximum effects from temperature variations. Guidance on live loads to be considered under extreme environmental conditions from winds, currents and waves is given in Section 4.5. Normal maximum live loads should be combined with maximum effects from temperature variations, as these variations will not occur during tropical cyclone conditions.

4.5

Unless stated otherwise, the extreme environmental conditions for structures having a design life of 50 years should be taken as those having return periods of 100 years. This is a conservative assumption intended to take account of the occurrence of tropical cyclones and the relatively short periods for which most records are available. Where special circumstances apply, resulting in a shorter or longer design life, the return period should be adjusted accordingly. It should be noted that for an event with a return period of 100 years, there is approximately a 1% chance of occurrence in any one year and approximately an 18%, 40% and 63% chance of occurrence in a 20 year, 50 year and 100 year period respectively. Reference can be made to Section 21.4 of BS 6349:Part 1 (BSI, 1984a) for the method of calculation. In any event, under Hong Kong conditions, where at present the periods of environmental records are relatively short, resulting in some degree of lack of confidence in assessed extremes, the true chance of occurrence may well be underestimated by using the formula.

50 100 100 1 20 50 100 1 8 4 0 63 BS6349: Part 1 (BSI, 1984a) 21.4

4.2.3 Temporary Loading Conditions Temporary Loading Conditions are those which arise during construction, towing, installation or the carrying out of unusual but foreseeable operations, such as the application of a test load. For these conditions, a combination of the appropriate dead and maximum temporary loads, together with the associated environmental loads, should be considered.

4 . 2 . 3

33

Temporary design and environmental conditions should be appropriate for the location, and for the time of year, when the construction or operation will be carried out.

4.2.4 Accident Loading Conditions Accident Loading Conditions are those which occur during accidental impact by a vessel. For these conditions, a combination of dead, superimposed dead and hydrostatic loads, soil pressures, live loads and normal environmental loads, together with the appropriate accident berthing load, should be considered. Guidance on accident berthing loads is given later in this Chapter. The above combination is to some extent artificial, as an accident can occur at a time of normal or extreme environmental loading conditions. However, it is not normally necessary to combine accident berthing loads with maximum imposed loads and extreme environmental loads because of the low probability of their simultaneous occurrence. It is not necessary for all marine structures to be designed or checked for accident loading conditions. The need for checking will depend on :

4 . 2 . 4

(a)

the importance of the structure,

(a) (b)

(b) the location with respect to normal ferry routes and fairways, (c) the degree of exposure environmental conditions, to adverse

(c) (d) (e)

(d) the expected degree of use if the structure is a pier, and (e) the susceptibility to damage of the type of design used.

Public and ferry piers, except those expected to have a particularly low level of use and in a particularly protected location, should generally be designed or checked for accident loading conditions. For such accident loading conditions, damage to minor structural members which can be readily repaired, and

34

to such items as fenders, can be accepted at the discretion of the designer.

4.3

Dead Loads

4 . 3 4.6

The dead load is the weight of the structural elements of the structure, including any substructure, piling and superstructure. The weight of the structure is its weight in air. Where parts are wholly, partially or intermittently immersed in water, upthrust on those parts should be calculated separately, as recommended in Section 4.6.

4.4

Superimposed Dead Loads

4 . 4

The superimposed dead load is the weight of all materials imposing loads on the structure that are not structural elements, and should include surfacing, fixed equipment, fenders, bollards, handrails, ladders, walkways, stairways, services, fittings and furniture. For all loading conditions, the possibility of any of the superimposed dead loads being removed should be considered.

4.5

Live Loads

4 . 5 BS6399: P a r t 1 ( B S I , 1 9 8 4 b )

The imposed live loads include all loads which the structure has to withstand except dead, superimposed dead, hydrostatic, soil, vessel-imposed and environmental loads and, in the case of concrete structures, loads from such effects as prestress and creep, which are treated separately in other sections. Where appropriate, the imposed live loads should be taken as equal to those defined in, and calculated in accordance with, BS 6399:Part 1 (BSI, 1984b).

4.5.1 Determination of Imposed Live Loads The determination of imposed live loads due to normal working operations for various structures should be in accordance with the following. (1) Public Piers. The live load for the main decks of

4 . 5 . 1

(1)

35

public piers, to include for the movement of pedestrians, hand luggage, ship provisions and temporary stacking, should be taken as 10 kPa. Where access to the pier by vehicles is not physically prevented, the live load should also allow for possible emergency vehicular access by an ambulance, police vehicle and/or fire engine as appropriate and agreed with the relevant authority. The live load for any upper deck should be taken as 5 kPa, and 2.5 kPa for a roof with no general access. (2) Ferry Piers. The live loads for pedestrian ferry piers should be no less than those given above for public piers, but should in addition be checked and agreed with the prospective ferry operators. The live loads for vehicular ferry pier waiting areas and ramps will depend on the types of vehicles allowed or expected to use the services, and should be agreed with the prospective ferry operators. (3) Other Piers. The live loads for other piers should be determined after consultation with the prospective users, taking into account the proposed use, possible cargo storage, cargo handling equipment and vehicular access. (4) Seawalls. Where no specific use has been designated for the area of land to be formed immediately behind a seawall at the time of design, the live load for such land for the design of the seawall should be taken as 10 kPa. Where a specific use has been designated, the live load should be determined taking such use into account, but this load should not be less than 10 kPa.

1 0 k P a 5 kPa 2.5 kPa

(2)

(3) (4) 1 0 k P a 10 kPa

4.5.2 Determination of Continuous Live Loads Guidance on the determination of the live load due to continuing operations under extreme environmental conditions from winds, currents and waves, and of the live loads to be used in accident loading conditions referred to in Section 4.2, is given below.

4 . 5 . 2 4.2

(1) Live Loads under Extreme Environmental Conditions. The live loads due to continuing operations under extreme environmental conditions from winds, currents and

(1)

36

waves may be taken as zero for piers unless there is a specific need or requirement for the pier to be used during tropical cyclone conditions, e.g. for emergencies or storage. For seawalls, the maximum live loads on the adjacent land due to continuing operations under extreme environmental conditions should be taken as 50% of the live loads due to normal working operations under normal environmental conditions, unless it can be ensured with reasonable certainty that the land behind the seawall will not be used for the storage or temporary stacking of materials. In this latter case, the live load can be taken as zero. For other structures, the live loads due to continuing operations under extreme environmental conditions should be assessed by the designer. Normal maximum live loads should be considered to apply under extreme environmental conditions relating to temperature variations. (2) Live Loads under Accident Conditions. The live loads to be used in Accident Loading Conditions for normal structures can be taken as 50% of the live loads due to normal working operations under normal environmental conditions. At the discretion of the designer, this percentage may be reduced to 25% for structures expected to be loaded infrequently, or increased to 75% for structures expected to have particularly heavy usage such as ferry piers on major routes with exceptionally frequent services.

5 0

(2) 5 0 2 5 7 5

4.6

Hydrostatic Loads

4 . 6

When considering the effects of buoyancy, it is preferable to represent the buoyancy and gravitational loads as separately applied loading systems. In this way, the effect of changes in water level can be seen more clearly, and it is possible in limit state design to apply different load factors to dead loads and hydrostatic loads as appropriate.

4.7

Soil Pressures

4 . 7 B S 6 3 4 9 : P a r t 1 6 G e o g u i d e 1 (GEO, 1993)

Guidance on the calculation of soil pressures is given in BS 6349:Part 1, Section 6. Reference can also be made to Geoguide 1 (GEO, 1993).

37

For the purposes of calculating soil pressures : (a) extreme water levels should be derived as described in Section 4.9,

(a) 4.9 (b)

(b) ground pore water pressures should be determined with reference to tidal range, soil permeability, drainage provisions, and any artesian and sub-artesian ground water conditions, and (c) allowance should be made for reduced passive resistance due to overdredging or scour.

(c)

4.8

Temperature Variations

4 . 8 5 0

The loads or load effects arising from thermal expansion or contraction of the structure and from temperature gradients in the structure will usually be minor in relation to other loads for marine structures with a maximum length between joints of 50 metres, and need not be considered. The loads arising from thermal expansion or contraction of the structure for marine structures with a length between joints exceeding 50 metres should be assessed. This is particularly important for piers and similar suspended deck structures where thermal movements of the deck induce loads in the supporting piles. Where no specific information is available concerning the temperatures of the structure at the time of construction, and the extremes expected during the design life of the structure, for design purposes an effective maximum temperature drop of 25C and an effective maximum temperature rise of 20C can be assumed for concrete deck structures under extreme environmental conditions. Under normal loading conditions, the effects of temperature variations may be ignored.

50 2 5 2 0

4.9

Tides and Water Level Variations

4 . 9 3.4

Information on tides and the extreme range of still water levels is given in Section 3.4. Such information is

38

required for the evaluation of :

(a)

overtopping, pressures, including buoyancy

(a) (b)

(b) hydrostatic effects, (c)

soil pressures, and

(c) (d)

(d) levels of action of mooring, berthing and wave forces. In addition, the effect of waves and wave run-up should be considered in relation to overtopping and hydrostatic pressures. For structures with a design life of 50 years, for the loading conditions referred to in Section 4.2, maximum still water levels due to tide and surge effects corresponding to the following return periods can be assumed : Loading Conditions Normal Extreme Accident Return Period 2 years 100 years 2 years

50 4.2

2 100 2

For structures where a different design life applies, the return period for extreme loading conditions should be adjusted accordingly. The observed lowest levels given in Section 3.4 can be used for Normal and Accident Loading Conditions. For Extreme Loading Conditions, the lowest water level can be assumed to be 0.3 m below the levels given in Section 3.4. For information on tide levels at locations not covered by the tide stations, the Observatory should be consulted. The range of still water level to be considered for Temporary Loading Conditions should be assessed by the designer for each individual case. Structures should be designed to withstand safely the effects of the extreme range of still water level

3.4 3 . 4 0 . 3

39

referred to above for each loading condition. It should be noted that for different types of structure, different loading cases, and different conditions, the critical still water level may be the minimum, maximum or some intermediate level; the full range must be investigated by the designer.

4.10 Winds For the assessment of wind loads on marine structures, for the loading conditions referred to in Section 4.2, the following design wind pressures may be assumed : Loading Conditions Normal Extreme Accident Design Wind Pressure 1.2 kPa 3.0 kPa 1.2 kPa

4 . 1 0 4.2

1.2 kPa 3.0 kPa 1.2 kPa

For Temporary Loading Conditions, the design wind pressure should be assessed by the designer for each individual case, taking into account the following points. (a) The design wind pressure of 1.2 kPa for Normal and Accident Loading Conditions corresponds to a gust of about 44 m/s, which is the maximum gust expected to occur with a mean hourly wind speed of 17 m/s (33 knots). This by definition is the maximum mean hourly wind speed likely to occur while Tropical Cyclone Signal No. 3 is hoisted or within the first few hours of the hoisting of Tropical Cyclone Signal No. 8. The above assumes a gustiness factor (ratio between maximum gust and mean hourly wind speed) of about 2.6, which is not normally exceeded under Hong Kong conditions. For details of gustiness factors, reference may be made to Chen (1975) and Poon (1982).

(a) 1 . 2 k P a 4 4 m / s 17 m/s (33) 1 7 m / s 2 . 6 Chen (1975) Poon (1982)

(b) The design wind pressure of 3.0 kPa under extreme environmental conditions corresponds to a gust of about 70 m/s (136 knots), which is the maximum gust expected to occur with a

(b) 3 . 0 k P a 7 0 m/s (136 ) 50

40

return period of about 50 years in Hong Kong waters. (c) Wind forces on structures and elements of structures should be calculated in accordance with Sections 4 and 5 of BDD (1983). Further guidance may be obtained from CP3 (BSI, 1972), Chapter V, Part 2.
(c) BDD (1983) 45 CP3 (BSI, 1972) 2

4.11

Currents

4 . 1 1 1 5 1 m / s 1 5 1 m/s

Where no detailed information or records are available at a site where a structure is to be located, the design current velocity for Normal, Extreme, Temporary and Accident Loading Conditions may be taken as a constant 1 m/s at a depth of 15 metres below the water surface. Below 15 metres water depth, the current may be ignored. For most locations, particularly within the harbour area, the above will be conservative, as current forces are assumed to act simultaneously with wave and wind forces. For locations near channels such as Kap Shui Mun, Urmston Road, Tolo Channel, Rambler Channel and Lei Yue Mun, where above average currents are encountered, the figure of 1 m/s should not be used without a detailed investigation. Where measurements are available, the designer should assess design current velocities for the various loading conditions. The direction of the design current for locations where no information or records are available should be determined by the designer. For locations close to the shore, the direction may normally be assumed to be parallel to the shore line. For isolated locations remote from the shore, it should normally be assumed that the design current can occur in all directions.

For the assessment of current forces on piles and other parts of structures, for all loading conditions other than for temporary conditions during construction, the area normal to flow should include an allowance for marine growth. Where no other information or site measurements are available, a uniform effective thickness of 100 mm of marine

100

41

growth for all surfaces below mean sea level can be assumed. Loads imposed by currents on marine structures may be classified as either drag forces parallel to the flow direction, or cross-flow forces transverse to the flow direction. Current drag forces are principally steady; the oscillatory component is only significant when its frequency approaches the natural frequency of the structure. Cross-flow forces are entirely oscillatory for bodies symmetrically presented to the flow.

4.11.1

Steady Drag Forces

4 . 1 1 . 1 BS6349: Part 138.2

Steady drag forces can be calculated using the formula given in Section 38.2 of BS 6349:Part 1. For the assessment of drag coefficients for circular cylinders, the values corresponding to moderate marine growth should be used unless special circumstances apply.

4.11.2

Flow-induced Oscillations

4 . 1 1 . 2 B S 6 3 4 9 : P a r t 1 3 8 . 3

Notes on flow induced oscillations for piles are given in Section 38.3 of BS 6349:Part 1. During construction, restraint should be provided to pile heads immediately after driving to prevent the possibility of oscillation in the cantilever mode. For completed structures in Hong Kong conditions, in typical water depths and with the types of pile normally used, it is not usually necessary to check critical flow velocities. Checks should be made for structures in particularly deep water where slender piles are being considered, and at locations where high design current velocities apply.

4.12 Waves 4.12.1 General

4 . 1 2 4 . 1 2 . 1

Wave loads on a structure are dynamic in nature, but when the design wave period is much higher than the structure's fundamental period, as will be the case for the vast majority of structures covered by this

42

Manual, these loads may be adequately represented by their static equivalents. Dynamic responses and vibrations are covered in Section 4.16. The crest of any design wave should be positioned relative to a structure such that the wave forces have their maximum effect on the structure. It should be noted that the maximum stress in elements of the structure may occur for wave positions, directions and periods other than those causing the maximum force on the structure.

4 . 1 6

4.12.2

Design Wave Parameters

4 . 1 2 . 2

Design wave parameters depend very much on the design life and return period, which have to be assessed for each individual structure. Design life is the intended useful life of the structure taking into account, amongst other things, changes in circumstances and operational practices which may make the structure redundant. Return period may vary according to the consequence of failure, availability of data and financial considerations.

As a guide, for extreme environmental conditions for a structure having a design life of 50 years, (except for rubble structures which are covered in Section 9.2), the design wave for the assessment of wave loads should be taken as the average of the highest 1% of all waves (H1), which is approximately equal to 1.67 times the significant wave height, having a return period of 100 years. For structures with different design lives, the return period should be adjusted accordingly. The design wave for normal and accident loading conditions, for the assessment of wave loads, should be taken as the average of the highest 1% of all waves corresponding to a mean hourly wind speed of 17 m/s, or the equivalent wind speed adjusted for duration, as appropriate for the fetch being considered. The reason for selecting this particular mean hourly wind speed is given in Section 4.10. Where, for the particular location and direction being considered, the mean hourly wind speed for a 5 year return period from Tables 6 to 9 is less than 17 m/s, a reduced mean hourly wind speed equal to this figure may be used in place of 17 m/s. For Temporary Loading Conditions,

5 0 100 1 ( H 1 ) 9 . 2 1 . 6 7

17 m/s 1 14.10 6 9 1 7 m / s 17 m / s

43

the designer should assess the design wave parameters for each situation, taking into account the likely mean hourly wind speed or its equivalent expected to be experienced.

Wave loading due to swells should be checked in the design, particularly for structures exposed to the South China Sea. For Normal and Accident Loading Conditions, swells having a return period of five years can be used in the design. A return period of 100 years should be used for Extreme Loading Conditions.

100

4.12.3

Calculation of Average Maximum Wave Height

4 . 1 2 . 3

The designer may, in some cases, choose to adopt the average maximum wave height for extreme environmental conditions. Information on the method of calculation of average maximum wave height is given in Section 3.3. The ratio between the average maximum wave height and significant wave height depends on the number of waves in the design event, as follows : Number of Waves 200 400 600 1000 2000 4000 Ratio 1.72 1.81 1.87 1.94 2.02 2.11

3.3

200 400 600 1000 2000 4000

1.72 1.81 1.87 1.94 2.02 2.11

In each case, the number of waves to be considered in the design event, and the resulting ratio between the average maximum wave height to be used for design and the significant wave height, must be decided by the designer. As a guide, the ratio should be within the range 1.8 to 2.0, which corresponds to a number of waves between about 400 and 2000. Normally, the average maximum wave height may be taken as 1.9 times the significant wave height. This corresponds to about 750 waves in the design event which, with an assumed average period of five seconds (typical for

1 . 8 2 . 0 400 2000 1.9 750 5

44

relatively protected locations in Hong Kong), in turn corresponds to a duration of the design event or storm of just over one hour. For design purposes, the wave period corresponding to the average maximum wave height may be taken to be equal to the significant wave period. It should be noted that, although the average maximum wave height is close to being the most probable value of the maximum wave height, by definition the actual maximum wave height is quite likely to exceed the average maximum wave height. As an example, using the probability formula given in Section 27.3.2 of BS 6349:Part 1, for a group of 750 waves there is about a 22% probability of the maximum wave height exceeding 2.0 times the significant wave height and about a 10% probability of it exceeding 2.1 times the significant wave height.

BS6349: Part 1 27.3.2 750 2 . 0 2 2 2 . 1 10

4.12.4

Depth-limited Situations

4 . 1 2 . 4 3 . 3

For comments on breaking and non-breaking waves in relation to the range of water levels to be considered, reference should be made to Section 3.3. It should be noted that, for some structures, the design wave will be 'depth-limited' and the design wave parameters will not correspond to those referred to above for the various loading conditions. For these situations, particular care should be taken with the design, as the structure will be subject to breaking waves and will be more likely to have to withstand many waves similar in magnitude to the design wave during its design life. For depth-limited designs, the design wave height will be dependent on the water depth at the structure, but the full range of possible periods should be investigated before determining the design wave period. As an upper bound, under depth-limited conditions, the design wave period may be taken to be the period of the design wave referred to above for the loading condition being considered, assuming the conditions are not depth-limited.

45

4.12.5

Calculation of Wave Forces in General

4 . 1 2 . 5 B S 6 3 4 9 : P a r t 1 3 9 . 4 S h o r e Protection ManualSPM (CERC, 1984) 3 T o m l i n s o n (1987) Bruun (1981)4.11

Guidance on the calculation of wave forces is given in Section 39.4 of BS 6349:Part 1 and Chapter 7, Section III of the Shore Protection Manual (SPM) (CERC, 1984). Reference may also be made to Tomlinson (1987) for wave forces on piles and to Bruun (1981) for wave forces on vertical walls. As for current loads in Section 4.11, allowance should be made in calculations for the build-up of marine growth on the structures. Notes on when to use reflection theory, diffraction theory and Morison's equation for the assessment of wave loads are given in Section 39.4.1 of BS 6349:Part 1. For the range of structure width to wavelength where diffraction theory applies, it is suggested that calculations are carried out separately for reflective conditions and using Morison's equation, and a weighted average used for the design wave load. For some structures, it will be necessary to separate the structure into different elements and apply different theories to different elements in order to assess the total wave load on the structure. For a normal pier consisting of a relatively open concrete deck with timber fenders suspended on piles, the deck should be considered to consist of a solid concrete deck edge, with effective depth to be assessed by the designer, for which reflective conditions will apply if the deck length is sufficient. Below this solid concrete deck edge, wave loads on the piles and fenders should be assessed separately using Morison's equation. It should normally be assumed that maximum wave forces on the deck edge, front piles and fenders can occur simultaneously. However, it should be noted that maximum wave forces will not occur simultaneously at all piles in a pile bent.

BS6349: Part 1 39.4.1 M o r i s o n M o r i s o n M o r i s o n

4.12.6

Wave Forces for Reflective Conditions

4 . 1 2 . 6 B S 6 3 4 9 : P a r t 139.4.2 SPM7 - 1617 - 180

For reflective conditions, the method of calculation of wave forces given in Section 39.4.2 of BS 6349:Part 1 may be used for breaking and nonbreaking waves, but care should be taken to cater for possible high local wave pressures if breaking waves apply. For non-breaking waves, the method of

46

calculation given in pages 7-161 to 7-180 of the SPM may be used as an alternative. For the situation referred to above, with a solid concrete deck edge suspended on piles, it is conservative to assume that the pressure distribution will be the same as that for a solid vertical wall replacing the piles, and to use either of the methods given above, allowing for a rubble foundation in place of the piles under the deck, see Figure 23(a) of BS 6349:Part 1 and Figure 7-98 of the SPM.

B S 6 3 4 9 : P a r t 1 2 3 ( a ) SPM7 - 98

For background information on the development of the formulae given in Section 39.4.2 of BS 6349:Part 1 for the calculation of wave forces for reflective conditions, reference may be made to Goda (1974). These formulae were developed experimentally for both breaking and non-breaking waves, and calibrated with prototype breakwaters. From the information and conclusions given by Goda (1974), it is suggested that Minikin's method for the assessment of wave forces under reflective conditions for breaking waves, as given in pages 7-181 to 7-192 of the SPM, should not normally be used for long vertical-face structures, as the predicted wave pressures appear generally to be far larger than those measured. However, Minikin's method may still be used as an upper bound check on :

BS6349: Part 1 39.4.2 G o d a ( 1 9 7 4 ) Goda ( 1 9 7 4 ) S P M 7 - 1817 - 192 Minikin Minikin

(a)

(a)

wave forces from breaking waves for particularly critical structures,

(b) (c) Morison

(b) local wave pressures, and (c) for structures of intermediate width, where diffraction theory applies and both reflection formulae and the use of Morison's equation will be required, as explained above.

4 . 1 2 . 7 M o r i s o n

4.12.7

Wave Forces Using Morison's Equation

For the assessment of wave forces on piles and other elements or structures which extend from water level to sea-bed level, the method of calculation given in pages 7-101 to 7-160 of the SPM based on

SPM7 - 1017 - 160 M o r i s o n

47

Morison's equation may be used. For the assessment of wave loads on elements or structures such as timber fenders and dolphin pile caps which do not fully extend from water level to sea-bed level, the following methods are suggested : (a) Maximum total wave force, assuming the element or structure extends fully from water level to sea-bed level, can be assessed as above from Figures 7-76 to 7-83 of the SPM.

(a) SPM 7-767-83 (b) SPM7 -717 - 72 BS6349: Part 1 39.4.4

(b) The degree of predominance of inertia/drag force components can be determined by assessing separately maximum total inertia and drag forces using Figures 7-71 and 7-72 of the SPM and comparing these forces with the maximum total wave force assessed above. The degree of predominance can be checked using the criteria given in Section 39.4.4 of BS 6349:Part 1 relating the width of the submerged part of the element or structure to the orbit width of the water particles at the surface.

(c)

The maximum total wave force on the element or structure, as a percentage of the maximum total wave force assuming the element or structure extends fully from water level to seabed level, can be estimated from Figures 9 to 20. Figures 9 and 10 give the variation of maximum wave force with depth, from 0.4 d above still water level to sea-bed level, where d is the depth from still water level to sea bed level, for inertia and drag components, and have been prepared for linear (Airy) wave theory using equations 7.25 and 7.26 from the SPM. Figures 11 to 20 give the variation of the percentage of total maximum wave force with depth, from water level to sea bed level, for inertia and drag components, and have been prepared directly from Figures 9 and 10. In each case, sets of curves are given for five different water levels from 0.4 d above still water level to still water level. The maximum wave crest elevation above still water level can be assessed from Figure 7-69 of the SPM. The actual water level at the

(c) 9 2 0 9 10 0.4d d ( A i r y ) S P M 7.25 7.26 11 20 9 10 0.4d SPM 7 - 69 100 100

48

element or structure can be estimated on the assumption that 100% inertia force predominance corresponds to still water level at the element or structure, and 100% drag force predominance corresponds to maximum wave crest level at the element or structure. (d) The actual maximum total wave force on the element or structure can be calculated by multiplying the maximum total wave force assessed in (a) above by the percentage estimated in (c) above. It should be noted that this method is only approximate, the major limitation being that the total wave loads are based on Dean's stream-function theory and the distribution of wave load with depth is based on linear (Airy) wave theory. Figure 7-75 of the SPM give details of the regions of validity of various wave theories and can be used to assess the degree of non-linearity of a wave.

(d) ( a ) ( c ) Dean ( A i r y ) S P M 7 - 7 5

It is particularly important when assessing wave forces for suspended deck structures, where reflective conditions may apply for one part and Morison's equation for another part of the structure, to check wave forces for different still water levels. The critical still water level for wave loads on different elements of the structure will not always be the same, and will not always correspond to the critical water level for wave loads for the structure as a whole.

M o r i s o n

It should be noted that the formulae given in Section 39.4.4 of BS 6349:Part 1 for the calculation of wave forces using Morison's equation are derived from linear (Airy) wave theory, and it is recommended that these formulae are not directly used for the calculation of wave forces without first checking that this theory applies for the situation being considered. This point is covered by pages 7-101 to 7-112 of the SPM. For suggested values of inertia and drag coefficients, reference may be made to Figure 24 and Table 5 of BS 6349:Part 1. It is recommended that the values corresponding to moderate marine growth in

B S 6 3 4 9 : P a r t 1 39.4.4 Morison ( A i r y ) SPM 7 - 1017 - 112

BS6349: Part 1245 2 4

49

Figure 24 are used unless special circumstances apply. 4.12.8 Wave Uplift Pressures

4 . 1 2 . 8 2 0 French (1979) Hydraulic Research Station (1971a b)

The standard references given above do not cover wave uplift pressures on suspended deck structures in any detail. For a deck whose soffit is above still water level, uplift pressures are characterised by an initial peak pressure of relatively high magnitude but short duration, followed by a relatively slowly varying uplift pressure of lower magnitude but of considerable duration, and which is first positive and then negative. Where no other information is available, for normal structures in open water and for wide structures fronting sloping rock-armoured seawalls, where it is possible for still water level to coincide approximately with the deck soffit level, the deck should be designed for a uniform uplift and down drag pressure corresponding to one half of the maximum wave height, with an additional uplift pressure corresponding to the average maximum wave height covering a one metre wide strip parallel to the wave front. For narrow structures (width less than 20 metres) fronting sloping rock-armoured seawalls, for all structures fronting vertical seawalls, and for structures in particularly exposed locations, the above wave pressures may not be adequate for design, and further research should be carried out by the designer. If possible, model studies should be carried out. For additional information on wave uplift pressures, reference may be made to French (1979), and the Hydraulic Research Station (1971a & b).

4 . 1 3

4.13 Berthing 4.13.1 General

4 . 1 3 . 1

In the course of berthing, loads will be generated between the vessel and the berthing structure from the moment at which contact is first made until the vessel is finally brought to rest. The magnitude of the loads will depend, not only on the size and velocity of the vessel, but also on the nature of the structure, including any fendering, and the degree of resilience it presents under impact.

50 B S 6 3 4 9 : P a r t 4 ( B S I , 1 9 9 4 ) 3

Berthing loads transmitted to a structure comprise impact loads normal to the berthing face and friction loads parallel to the berthing face. The impact load normal to the berthing face depends upon the berthing energy and the load/deflection characteristics of the vessel, structure and fender system. The design friction load parallel to the berthing face should be taken as the coefficient of friction between the two faces in contact multiplied by the maximum design impact load. Guidance on this coefficient is given in Table 3 of BS 6349:Part 4 (BSI, 1985a).

4 . 1 3 . 2

4.13.2

Assessment of Energy to Be Absorbed

A method of assessing the total amount of energy to the absorbed, either by the fender system alone or by a combination of the fender system and the structure itself for a structure with some flexibility, is given in Section 4.7 of BS 6349:Part 4. For this assessment, the eccentricity coefficient can normally be taken as 0.5, the softness coefficient as 1.0, and the berth configuration coefficient as 0.8 and 1.0 for a solid quay wall and open piled jetty respectively. The design velocity of the vessel normal to the berth depends on the vessel size and type, frequency of arrival, possible constraints on movement approaching the berth, and wave, current and wind conditions likely to be encountered at berthing. Where no other information is available, for the Normal Loading Conditions referred to in Section 4.2, the following transverse velocities may be used as a guide : Vessel Displacement Under 100 t 100 to 200 t 200 to 500 t 500 to 1500 t Transverse Velocity 0.40 m/s 0.35 m/s 0.30 m/s 0.25 m/s

B S 6 3 4 9 : P a r t 4 4 . 7 0.5 1.0 0 . 8 1 . 0 4.2

100 100 200 500 t 200 t 500 t 1500 t

0.40 0.35 0.30 0.25 m/s m/s m/s m/s

The transverse velocities suggested above relate to structures located at sites with normal exposure to environmental conditions without excessive frequency of use, and assume that berthing may continue after the raising of Tropical Cyclone Signal No. 3, and for the

51

first few hours after the raising of Tropical Cyclone Signal No. 8. Before any velocity is finally adopted for detailed design, advice should be sought from the Director of Marine and other users or ferry operators as appropriate.
4.2.4 50 100

For Accident Loading Conditions, general comments are given in Section 4.2.4. The vessel displacement and transverse velocity for such conditions must be decided by the designer for the individual structure being considered, but as a general rule the total energy to be absorbed for accident loading should be at least 50% greater than for normal loading. For particularly critical structures or for structures with expected heavy use and exposure, this may need to be increased to 100%.

4 . 1 3 . 3

4.13.3

Berthing Reactions

Berthing reactions to be taken by the structure can be assessed from the manufacturer's reaction/deformation/energy curves once the type of fender to be used has been determined. For information on types of fenders, see BS 6349:Part 4, Quinn (1972) and manufacturers' catalogues. The ultimate energy capacity of each fender should in general be at least 50% greater than that calculated for normal loading conditions to allow for accidental occurrences such as vessel engine failure, breaking of mooring or towing lines, sudden changes of wind or current conditions, and human error. Because of the non-linear energy/deflection and reaction/deflection characteristics of most fender systems, the effects of both normal and abnormal impacts on the fender system and berth structures should be examined.

BS6349: Part 4 Quinn ( 1 9 7 2 ) 5 0

4 . 1 4

4.14 Mooring Mooring loads comprise those loads imposed on a structure by a vessel tied up alongside, both through contact between the vessel and structure or its fendering system, and through tension in mooring ropes. These loads are principally caused by winds and

(1500

52

currents and, in more exposed locations, by waves. For the normal structures covered by this Manual and the relatively small vessels considered for berthing (displacement generally less than 1500 t), it is not usually necessary to carry out specific calculations to determine the probable maximum mooring loads, as these loads generally will not be critical for structural design.

Mooring bollard locations and normal maximum working loads should be agreed with the Director of Marine, user departments and the ferry operators as appropriate. For Normal Loading Conditions, mooring loads may be assumed to be equal to the normal maximum bollard working loads. Allowance should be made for the mooring lines not being horizontal; if no other information is available, a maximum angle to the horizontal of 30 (up and down) may be assumed. The direction of each mooring load should be taken as that having the most adverse effect on the structure, and in general it should be assumed that all mooring loads on a structure can act simultaneously. Because of the relatively small vessels considered for berthing, the loads imposed on a structure by direct contact between the vessel and the structure, or its fendering system, need not be considered for Normal Loading Conditions, as these will usually be minor in relation to the combined effects of other imposed loads such as those from winds, currents, waves and berthing. Where it is considered necessary to calculate the forces acting on moored vessels in order to check bollard loads or loads imposed directly by vessels on a structure, reference may be made to BS 6349:Parts 1 & 4, Quinn (1972) and Bruun (1981).

3 0 o

BS6349: Part 1 BS6349: Part 4 Quinn (1972) Bruun (1981)

4 . 1 5 GCO (1991)

4.15 Earthquakes For the marine structures covered by this Manual, seismic forces in Hong Kong can be assumed to be minor in relation to the combined effects of other

53

imposed loads, and it can be assumed that these structures can adequately withstand such forces without a specific check. Further information on seismicity may be obtained from GCO (1991).
4 . 1 6 BS6349: Part 1 47

4.16 Movements and Vibrations For guidance on movements and vibrations, reference may be made to Section 47 of BS 6349:Part 1. For the marine structures covered by this Manual and the relatively shallow-water depths normally applying, movement and vibration problems should not be expected and usually can be effectively ignored. Movements between different parts of structures, and between new and existing structures, should be assessed in the usual way in order to fix joint sizes and locations. Where vessel berthing occurs, movements of flexible and even relatively inflexible structures can be important in assisting with energy absorption.

54

55

5. DESIGN OF FOUNDATIONS

5.

5.1

Introduction

5 . 1

This Chapter gives guidance on the design of foundations for marine structures and covers such aspects as site investigation, soil properties and foundations. The structural design of piles is covered by Chapter 6. Guidance on the design of sheet piled, gravity and rubble structures is given in Chapters 7, 8 and 9 respectively, although many comments in this Chapter apply also to these types of structures. The structure and its foundation should be designed so that, during its intended life, foundation displacements and movements are kept within the limits that the structure can tolerate without affecting its structural integrity and functional capability. Consideration of the interaction between structure and soil, and the need to limit foundation movements, may determine the most suitable type of structure for a particular location. The performance of the structure and the sea bed should be considered together. BS 8004 (BSI, 1986) relates to the foundations of buildings and general engineering structures, but many of the recommendations are equally applicable to marine structures. One particularly important aspect to be investigated for foundations for marine structures relates to the stability of the adjacent sea bed under wave and current action, and the possibility of scour and undermining. It is also necessary to check the overall stability of marine structures against potential shear failure in the supporting ground. It is recommended that global factors of safety should be used when designing foundations for marine works. Loads used should be unfactored values covered by Chapter 4, with no allowance for partial safety factors. When considering the interaction between structure and soil, all of the appropriate loading conditions described in Section 4.2 should be examined. If it is expected that other loading conditions could be critical, they should also be investigated. Guidance on factors of safety is given in Section 5.4 for piled foundations and in later sections

B S 8 0 0 4 ( B S I , 1986)

4.2 5 . 4

56

for other types of foundations.

5.2

Site Investigations

5 . 2 Geoguide 2 (GCO, 1987) Geoguide 3 (GCO, 1988) G e o g u i d e 2 1 4 1 0 . 7 . 7 3 3 . 3 B S 6 3 4 9 : P a r t 1 ( B S I , 1 9 8 4 a ) 4 9

Reference should be made to Geoguide 2 (GCO, 1987) for guidance on good site investigation practice and Geoguide 3 (GCO, 1988) for guidance on description of rocks and soils in Hong Kong. Chapter 14 of Geoguide 2, which covers ground investigations over water, is particularly relevant, together with Sections 10.7.7 and 33.3. Further guidance relevant to marine situations is given in Section 49 of BS 6349:Part 1 (BSI, 1984a).

5.3

Properties of the Ground

5 . 3 B S 6 3 4 9 : P a r t 1 5 0 BS8004 (BSI, 1986)

Information on average properties for preliminary design and the selection of parameters for working design is given in Chapter 50 of BS 6349:Part 1. Further guidance on the properties of the ground in relation to various structures is given in BS 8004 (BSI, 1986).

5.4

Piled Foundations

5 . 4 BS8004 Tomlinson (1987) B S 6 3 4 9 : P a r t 2 (BSI, 1988) 6.12.4 6.12.5 6 . 1 2 . 66 . 1 2 . 8 6 . 1 2 . 1 2 2 . 0 3.0

Guidance on the design of piled foundations is given in BS 8004 and Tomlinson (1987). Sections 6.12.4, 6.12.5, 6.12.6, 6.12.8 and 6.12.12 of BS 6349:Part 2 (BSI, 1988) are also relevant.

For any pile, the working load should be not greater than the ultimate bearing or pull-out capacity, as appropriate, divided by a factor of safety. Wherever possible and practical, ultimate bearing or pullout capacities should be assessed from loading tests. Where ultimate bearing or pullout capacities have been assessed from a number of loading tests, a global factor of safety of 2.0 is recommended. Where only one or a small number of loading tests has been carried out, an increase in the factor of safety should be considered. Where no loading test has been carried out, and ultimate bearing or pullout capacity has been assessed from application of a dynamic driving

57

formula, use of stress wave analysis, soil tests, or combination of these, a global factor of safety of 3.0 is recommended. Where reduced resistance has been found during redriving, or where no combination referred to above has been used, or where inconsistent results have been found, an increase in the factor of safety should be considered. The working load for a pile may be taken as the maximum of the computed axial bearing and pullout loads for the various loading conditions referred to in Section 4.2 multiplied by the following factors: Loading Condition Normal Extreme Temporary Accident Factor 1.00 0.80 0.80 0.75
4 . 2

1.00 0.80 0.80 0.75

It should be noted that the above factors imply an acceptable level of overstress compared with normal conditions of 25% for extreme and temporary conditions and 33.3% for accident conditions.

2 5 33.3

There are a number of definitions for the ultimate capacity of piles. However, for a loading test carried out in accordance with the General Specification for Civil Engineering Works (Hong Kong Government, 1992a), the ultimate bearing or pullout capacity of a pile may be taken as the maximum test load for which the permanent settlement or upward movement, on completion of the load test, is D/50 or 8 mm, whichever is the lesser, where D is the diameter of a circular pile or the least dimension of a rectangular pile.

General Specification for C i v i l E n g i n e e r i n g W o r k s ( H o n g K o n g Government, 1992a) D / 5 0 8 D

58

59

6. DESIGN OF SUSPENDED DECK STRUCTURES

6.

6.1

Introduction

6 . 1

This Chapter gives guidance on the design of suspended deck structures, including dolphins. It covers structures of concrete, steel and a combination of these materials. The design of sheet pile and gravity structures located immediately behind marginal quays, and of rubble mound structures located under marginal quays, is covered in Chapter 7, 8 and 9 respectively. Comments on general design methods are given in Section 6.10 of BS 6349:Part 2 (BSI, 1988). Particular note should be taken of Clause 6.10.3.1 of BS 6349:Part 2, which relates to the distribution of transverse loads.

B S 6 3 4 9 : P a r t 2 ( B S I , 1 9 8 8 ) 6 . 1 0 B S 6 3 4 9 : P a r t 2 6 . 1 0 . 3 . 1

6.2

Load Combinations and Factors

6 . 2 4 . 2 4.34.16

The structure should be designed to resist all combinations of loads which may realistically be assumed to act on the structure simultaneously, either directly on the superstructure or indirectly via the piles. As a minimum, the load cases referred to in Section 4.2 should be considered. Guidance on nominal loads for both limit state and working stress design methods are given in Sections 4.3 to 4.16. For limit state design, it is recommended that the load factors given in Clause 6.11.4.2 of BS 6349:Part 2 are adopted. It is suggested that, for Accident Loading Conditions (see Section 4.2.4), the partial load factors given in Table 2 of BS 6349:Part 2 for Extreme Loading Conditions (Section 4.2.2) should be used. For working stress design, under Extreme and Temporary Loading Conditions (Sections 4.2.2 and 4.2.3), it is suggested that normal permissible working stresses may be increased by 25%; under Accident Loading Conditions, an allowable increase of 33.3% is

BS6349: Part 2 6.11.4.2 4.2.4 B S 6 3 4 9 : P a r t 2 2 4.2.2 4.2.2 4 . 2 . 3 2 5 3 3 . 3 B S 6 3 4 9 : P a r t 2 6 . 1 1 . 4 . 3

60

suggested. Reference is made to Section 6.11.4.3 of BS 6349:Part 2. For all loading conditions other than Accident Loading Conditions, the partial safety factors for materials given in Table 2.2 of BS 8110:Part 1 (BSI, 1985a) should be used. For accident loading conditions, the factors for concrete in flexure, and for steel reinforcement, may be reduced from 1.5 to 1.3 and from 1.15 to 1.0 respectively.
BS8110: Part 1 (BSI, 1985a) 2.2 1.51.31.151.0

6 . 3

6.3

Superstructure
BS8110: Part 1 6.2 B S 4 4 9 : P a r t 2 ( B S I , 1 9 6 9 ) 6.2

It is recommended that BS 8110:Part 1 should form the basis for the design of concrete deck structures, using the partial load factors referred to in Section 6.2. For steel deck structures, the use of BS 449:Part 2 (BSI, 1969) is recommended, with increased permissible stresses for certain loading conditions as suggested in Section 6.2.

6 . 4

6.4

Piles
BS6349: Part 2 6.12 6.3

For general comments on the design of piles, see Section 6.12 of BS 6349:Part 2. It is recommended that limit state design is used for concrete piles, and working stress design for steel piles, as for concrete and steel deck structures in Section 6.3 above, using similar load factors and increased permissible stresses as appropriate.

6 . 5

6.5

Durability
6 . 5 . 1

6.5.1 Reinforced and Prestressed Concrete General requirements for reinforced or prestressed concrete can be found in Section 11.4. Compliance with the bar spacing rules given in Section 3.12.11 of BS 8110:Part 1 will generally ensure that, for the most severe combination of loads under the serviceability limit state, crack widths anywhere in a concrete structure will be limited to a maximum of 0.3 mm. For concrete within the
11.4 B S 8 1 1 0 : P a r t 1 3 . 1 2 . 1 1 0 . 3 0.1

61

intertidal and splash zones, it is recommended that crack widths under typical average long term loading conditions should be limited to 0.1 mm. The typical average long term loading conditions for each element of a structure will depend on the type of structure and its use, and must be assessed by the designer for each case. As a guide, for the normal types of suspended deck structure covered by this Manual, typical average long term loading should cover full dead and superimposed dead loads, combined with 50 to 75% of full live loads, using nominal or characteristic loads in each case. Normally, berthing, mooring, wind and wave loads need not be considered, because of their relatively short duration. For the assessment of crack widths, reference can be made to Section 3.8 of BS 8110:Part 2 (BSI, 1985b).

5 0 7 5 BS8110: Part 2 (BSI, 1985b) 3.8

6 . 5 . 2

6.5.2 Steelwork Guidance on corrosion protection and allowance for metal losses can be found in Chapter 11.
11

62

63

7. DESIGN STRUCTURES

OF

SHEET

PILED

7.

7.1

General

7 . 1 BS6349: Part 1 (BSI, 1984a) 51 BS6349: Part 2(BSI,1988) 4 GCO (1990) BS8002(BSI, 1994c)

General information on the design of sheet piled structures is given in Section 51 of BS 6349:Part 1 (BSI, 1984a), and information on the design of sheet piled walls in quay and jetty construction is given in Section 4 of BS 6349:Part 2 (BSI,1988). Reference may also be made to GCO (1990) and to BS 8002 (BSI, 1994c).

7 . 2

7.2

Corrosion Protection
11.5.2 0.5

For steel sheet piled structures, the basic comments regarding corrosion loss and protection given in Section 11.5.2 for steel suspended deck structures apply. Full corrosion protection is recommended above sea bed level for normal permanent structures. For temporary structures, with a design life not greater than ten years, where there is no corrosion protection, an allowance for corrosion loss of 0.5 mm/year per surface is suggested for the zone between sea bed level and Chart Datum. Above Chart Datum, full corrosion protection is strongly recommended, even for temporary structures.

64

65

8. DESIGN OF GRAVITY STRUCTURES

8.

8.1

General

8 . 1 B S 6 3 4 9 : P a r t 2 ( B S I , 1 9 8 8 ) 5 BS8002 (BSI, 1994c) G e o g u i d e 1 ( G E O , 1 9 9 3 )

Information on the design of gravity walls in quay and jetty construction is given in Section 5 of BS 6349:Part 2 (BSI, 1988). Reference may also be made to ISE (1951), which is currently under revision as BS 8002 (BSI, 1994c), and Geoguide 1 (GEO, 1993). Certain aspects of the design of concrete blockwork walls, which are the most commonly constructed gravity structures in Hong Kong, are expanded upon in this Chapter.

8.2

Concrete Blockwork Walls

8 . 2 8 . 2 . 1 G e o g u i d e 2 ( G C O , 1 9 8 7 ) G e o g u i d e 2 1.3 1.2 2.0 1.5 1 . 7 5 1.5

8.2.1 General Whenever possible, checks against soil shear failure should be based on ground investigation in accordance with Geoguide 2 (GCO, 1987) and tests on construction materials. Where soil properties have been tested in accordance with Geoguide 2, minimum factors of safety against soil shear failure of 1.3 for Normal Loading Conditions and 1.2 for Extreme Loading Conditions are recommended. For overturning, under Normal Loading, a factor of safety of 2.0 is recommended provided that the resultant is within the middle third of the base width for the interface between the blocks and the rubble foundation when transient loads are ignored. For the interface between the blocks, the resultant force may lie within the middle half. For Extreme Loading, the resultant may fall outside the middle third provided the minimum factor of safety against overturning is 1.5. Recommended minimum factors of safety against sliding of the base, or at horizontal block interfaces, are 1.75 for Normal Loading and 1.5 for Extreme Loading Conditions. For information on coefficients of friction, to supplement that given in Sections 5.3.1.4 and 5.4.8.5 of BS 6349:Part 2, see Table 7-16 of the Shore

BS6349: Part 2 5.3.1.4 5.4.8.5 Shore Protection Manual

66

Protection Manual (CERC,1984) and Appendix B, Part 3, of Bruun (1981). For friction between two precast concrete blocks and between a precast concrete block and a levelled rubble foundation, a coefficient of friction of 0.6 is suggested. For guidance on the stability against wave attack of rubble foundations for concrete blockwork walls, reference should be made to Chapter 9.

(CERC,1984) 7 - 16 Bruun (1981) B 0.6

8.2.2 Ground Water Levels and Profiles The situation with a wave trough at the seawall combined with the expected range of still water level should be investigated for both Normal and Extreme Loading Conditions. The ground water level immediately behind the concrete block seawall can be assumed to be the same as still water level. The ground water profile in the fill behind the seawall should be assessed, taking into account expected fill permeability, tidal lag and the flow of surface or subsoil water from landward sources. As a guide, for normal conditions, where the land behind the seawall is not paved and where the fill is highly variable, a ground water profile rising from behind the seawall at a slope of one vertical to two horizontal to a level of 0.5 m above MHHW is suggested; for extreme conditions, a slope of one vertical to one horizontal rising to ground level is considered reasonable.

8 . 2 . 2 1 2 0.5 1 1

8.2.3 Consideration of Settlement For all concrete blockwork walls, the total settlement expected during the design life of the structure should be assessed to ensure that this is acceptable in relation to the proposed use. In many cases, the depth of dredging and width of dredged trench for a rubble foundation of a concrete blockwork wall will be determined by the need to limit long term settlement rather than the need to have an adequate factor of safety against a deep rotational failure in the underlying soil.

8 . 2 . 3

67

9. DESIGN OF RUBBLE STRUCTURES

9.

9.1

General

9 . 1 Shore Protection Manual SPM (CERC,1984) 7 - 202 7 - 2 4 9

Information on the design of rubble structures is given in pages 7-202 to 7-249 of the Shore Protection Manual (SPM) (CERC,1984). Further guidance is given in this Chapter on some aspects of the design of rubble structures, including comments on the choice of design wave, stability formulae and the determination of crest level. It should be noted that, although this Chapter mainly covers rubble mound seawalls and breakwaters, many of the comments, particularly those covering toe protection and scour, apply equally to rubble foundations for gravity structures such as concrete blockwork seawalls.

9.2

Design Wave

9 . 2

The designer should first assess the type of waves critical for design of the structure. In general, for sheltered areas inside the harbour, short period locally generated storm waves will be predominant for the extreme case. In areas exposed to offshore waves, the effect of attack from long period swells has to be checked. In some cases, the combined effect of the two types of waves has to be taken into account. Notes on the selection of design wave for a rubble structure are given in 7-2 to 7-4 and 7-203 of the SPM. It is recommended that the design wave height for rubble seawalls and breakwaters, for use in Hudson's formula referred to in Section 9.3, should normally be taken as the average of the highest 10% of all waves, H10, which is approximately equal to 1.27 times the significant wave height. Hudson's formula does not include the effect of wave period and is not recommended for use in structures subject to attacks from swells. In Van der Meer's Formulae, the significant wave height is adopted.

SPM7 - 27 - 47 - 203 9 . 3 H u d s o n H10 1 0 1.27 Hudson Van der Meer

68

The designer should consider the use of a higher design wave if one or more of the following factors apply :(a) The structure is particularly important and major damage or failure would result in loss of life.

(a)

(b) The location is such that access for maintenance is exceptionally difficult and any maintenance would be expensive. (c) The design wave is from the easterly sector, which corresponds to the prevailing direction for both normal and tropical cyclone winds.

(b)

(c)

(d) There is some doubt about the assessed wave heights, for example if the waves have been forecast from wind records with a long or unlimited fetch. (e) The structure is located in relatively shallow water and regularly exposed to breaking waves of similar magnitude to the design wave. The site is exposed to long period swell waves of height similar to the design wave.

(d)

(e)

(f)

(f)

The use of a lower design wave height, such as the significant wave height, should only be considered if, say, three or more of the following factors apply: (a) The structure is of minor importance and major damage or failure would not result in loss of life.

(a)

(b) Access for maintenance is easy and any maintenance would be relatively inexpensive. (c) The site is not exposed to long period swell waves and the fetch is limited.

(b) (c) (d)

(d) Wave heights assessed from wind records are supported by wave records.

69

9.3

Stability

9 . 3

9.3.1 General For the stability of rubble structures against wave attack, covering the design of armour, underlayers, bedding layers, core and toe protection, reference may be made to pages 7-204 to 7-249 of the SPM. Comments on soil shear failure are given in Section 8.2 for rubble foundations of gravity structures; these comments apply also to rubble seawalls and breakwaters.

9 . 3 . 1 S P M 7 - 2 0 4 7 - 2 4 9 8.2

9.3.2 Design of Armour Units Using Hudson's Formula The design of armour units given by the SPM is based on the use of Hudson's formula, see Eq. 7-116. Figures 21 and 22 in this Manual have been prepared using this formula and the stability coefficients given in Table 7-8 of the SPM to assist in the preliminary design of armour units consisting of rough angular quarrystone with random placement and layer thickness comprising two units. A unit weight for the armour unit corresponding to a specific gravity of 2.6 has been assumed, and curves are given for structure slopes of 1 on 1.5, 1 on 2 and 1 on 3 for breaking and non-breaking wave conditions for structure trunk and structure head. The curves cover design wave heights up to about 6 m, with corresponding armour weights up to about 10 t; this corresponds approximately to the range of rock size normally available locally, although rock in the upper part of this range will usually only be available in small quantities. For armour design, it is recommended that the specific gravity of the rock, if obtained locally, should be taken as 2.6. A figure higher than this value should not be used for design without extensive testing, both prior to construction, where a rock source has been identified, and during construction for quality control.

9 . 3 . 2 H u d s o n

S P M H u d s o n 7 - 116 2 1 2 2 S P M 7 - 8 2.61: 1.51: 2 1: 3 6 1 0 1 0

2 . 6 2.6

70

9.3.3 Structure Head Conditions It is recommended that, when using Hudson's formula for armour design, and in particular the stability coefficients given in Table 7-8 of the SPM, the term "structure head" should be applied, not only to the ends of breakwaters, but also to all other locations/discontinuities, where normal "trunk" conditions no longer exist, including the following :

9 . 3 . 3 Hudson SPM 7 - 8

(a)

where breakwaters or rubble seawalls have sharp changes in direction,

(a) (b)

(b) at the ends of breakwaters or rubble seawalls where there is a junction with a vertical-face seawall, and (c) where other types of construction/ structure, such as a vertical-face seawall with landing steps or extensive culvert wing walls, have been incorporated into a length of breakwater or rubble seawall.

(c)

The length of structure to be considered as corresponding to "head" conditions will be dependent on site conditions, crest level and armour slope, and must be decided by the designer in each case. As a guide, at any junction/discontinuity, head conditions should be taken for design purposes to extend at least 20 metres past the location where the normal trunk section exists. It is usual to cater for head conditions by maintaining the same basic geometry of the structure used for trunk conditions and increasing the armour size to suit the reduced stability coefficient. As an alternative, if larger armour is not available, the armour size can remain unchanged for head conditions and the armour slope adjusted instead. For smaller structures with significant junctions/discontinuities where head conditions apply, for simplicity it may be justified to use stability factors corresponding to head conditions for the full structure length.

2 0

71

9.3.4 Model Testing Where possible, armour design using Hudson's formula and other aspects of rubble seawall and breakwater design taken from the SPM should be checked by model testing. For the majority of rubble structures in Hong Kong, with maximum armour size in the range of, say, 6 to 8 t, model testing will not normally be practical based on economic and programming considerations. Where required maximum rock armour sizes exceed the range given above, the use of precast concrete armour units will normally be necessary; in this case model testing is considered essential and account must be taken of this when planning a project, with regard to the effect on costs and programming.

9 . 3 . 4 Hudson S P M 6 8

9.3.5 Design of Armour Units Using Van Der Meer's Formulae The Van der Meer's Formulae have been established based on the result of a series of model tests using irregular waves. They are considered the most widely applicable of the prediction methods currently available and are based on the widest set of model data. These formulae take into account the wave period, number of waves, permeability of the core and distinguish between breaking and non-breaking conditions. They are described as practical design formulae for rock armour, although experience in their use is limited at present.

9 . 3 . 5 V a n d e r M e e r

Van der Meer

Details of the formulae are given in Van der Meer & Pilarczyk (1987). When using Van der Meer's Formulae, the core permeability factor, P, has to be assessed carefully. The values of P suggested range from 0.1 for a relatively impermeable core, up to 0.8 for a virtually homogeneous rock structure. However, it should be noted that values are only assumed and have not yet been related to the actual core permeability.

Van der Meer & Pilarczyk (1987) Van der Meer PP 0 . 1 0 . 8 0.1 0.8 P 0.3 0.1

72

Permeability coefficients of 0.3 for normal breakwaters and 0.1 to 0.2 for normal seawalls or revetments are suggested. Ultimately the choice of P must depend on the designer's judgement, and it is strongly recommended that the permeability be underestimated rather than over-estimated, if in doubt. Sensitivity of the final calculated rock weight to the assumed value of P should always be checked. Unless data available allow more detailed assessment to be made, it has been suggested that the following values be used :N = 3000 to 5000 S = 1 to 3 (roughly equivalent to 5% damage) It should be noted that experience in the application of these formulae is limited. Their use should not be extended beyond the experimental limits. Except for relatively sheltered areas and for small or unimportant structures, physical model tests are strongly recommended to verify the design.

0 . 2P P P

N= 30005000 S = 135

9.4

Crest Level

9 . 4 SPM7 - 2297 - 233 7 - 16 7 - 43 0.2 0.3

Notes on crest elevation and width are given in pages 7-229 to 7-233 of the SPM , and wave runup is covered in pages 7-16 to 7-43. For the majority of rubble structures covered by this Manual, where minor overtopping during the design event can be tolerated, the design wave for the assessment of runup can be taken to be equal to the significant wave height applicable at the structure. The crest level of the rubble structure will be the extreme water level plus the runup corresponding to this design wave, plus an allowance for structure settlement during the design life, plus a nominal allowance for freeboard of, say, 0.2 to 0.3 metre. An indication of the extent of overtopping to be expected, if the design wave for assessing runup is taken to be the significant wave height, can be obtained from Figure 7-23 of the SPM. The runup height exceeded by 10%, 5%, 2% and 1% of the

SPM7 - 23 1 0 5 2 1 8 2 3 3 9 5 2

73

waves is approximately 8%, 23%, 39% and 52% respectively greater than the runup for the significant wave, and the runup from approximately 13% of the waves will exceed that for the significant wave. Notes on concrete caps, including wave walls, for rubble structures is given in pages 7-235 and 7-236 of the SPM. For both breakwaters and rubble seawalls it is recommended that concrete caps/wave walls should be avoided where possible. The need for a wave wall can often be avoided simply by increasing the crest level of the rubble structure. If a wave wall is considered essential, for example for a rubble seawall to reduce minor overtopping to the minimum, the wave wall should be founded on an underlayer or the core, not the primary armour layer, and should be set back from the primary armour units. Any wave wall itself should consist of substantial concrete blocks, preferably keyed and unreinforced, and should be capable of taking some differential settlement; a rigid wave wall design founded on a flexible rubble structure should be avoided.

13

SPM7 - 2357 - 236

74

75

10. DESIGN OF RECLAMATIONS

10.

10.1

General

1 0 . 1

This Chapter gives brief comments and guidance on the design of reclamations, covering such aspects as extent and layout, reclamation level and methods, fill materials, and miscellaneous related matters including piling through reclamations, and culvert foundations.

The design of a reclamation should be carried out with the aim of fulfilling the requirements of the planned use and programme for development, particularly with regard to time/settlement characteristics of the reclamation and foundation materials. The timing of the future development is of particular importance, together with the location of future roads, drains, buildings and areas of open space. It should be noted that sometimes it may be necessary to design reclamations for the main initial purpose of spoil disposal; in such circumstances, future land use and development is likely to be known only in general terms at the time of reclamation design. It is relatively common for planned use and development of reclamations to change with time, due to changing needs, and account should be taken of this where possible by designing reclamations for flexibility of use. Site investigations prior to the design of reclamations are particularly important in order to be able to arrive at the most appropriate design, as reclamation projects are usually extensive, carrying major financial implications if problems develop during or after the construction stage. In addition to the normal investigations required for marine structures, the investigations may need to cover potential sources of fill materials. In particular, insitu and laboratory testing of soil samples within the proposed reclamation area should be carried out with the aim of determining settlement characteristics of underlying marine and alluvial soils.

76

10.2 Extent and Layout The extent of a reclamation will be governed largely by the shape of the existing shoreline, where an existing bay or indentation is to be reclaimed, and by strategic planning for more extensive reclamations, where major changes to the land/sea boundary result. The major limitations on the extent of reclamations will be water depth, and the need to maintain fairways, moorings and other marine traffic channels and clearances to the approval of the Director of Marine.

1 0 . 2

For all major reclamations, and for minor reclamations where there is a significant change in the shape of the land/sea boundary, it will be necessary to carry out detailed investigations, including physical and mathematical modelling, to ensure that the changes in currents, waves and storm surge characteristics expected to be caused by the reclamation have no unacceptable effects with respect to :

(a)

navigation of large and small vessels,

(a) (b)

(b) tidal flushing and water quality, (c) siltation and seabed scour,

(c)

(d) the operation of piers, wharves and cargo-handling areas, and (e) flooding due to tides combined with storm surge.

(d)

(e)

It must be recognised that the effects of a reclamation on water flow in general, and currents, waves and storm surge characteristics in particular, can cover a far more extensive area than the area immediately adjacent to the reclamation itself, and the detailed investigations referred to above should take this into account. The layout of a reclamation will largely be governed by town planning considerations. However, marinefrontage related uses on the planned reclamation, such as piers, wharves, cargo-handling areas, seawater

77

intakes, stormwater and sewerage outfalls, need to be identified at an early stage. The proposed locations of such structures, areas and facilities will need to be known, at least in general terms, at the time of the detailed investigations referred to above. The layout of a reclamation, with regard to the locations of roads, bridges, buildings, stormwater culverts and major sewers, is important when considering the settlement characteristics of the fill material to be used in different areas of the reclamation and the proposed reclamation method. However, as pointed out in Section 10.1, reclamations should be designed for full flexibility of land use where possible.

1 0 . 1

10.3 Reclamation Level Before deciding on the finished level for a reclamation, the following aspects should be considered : (a) the availability and cost of fill,

1 0 . 3

(a) (b) (c) (d)

(b) the urgency of the land development, (c) existing ground, road and drain levels for adjacent developments,

(d) post-construction settlement of the fill and underlying marine and alluvial soils, (e) the type of seawall and the extent of wave overtopping expected, normal and extreme still water levels due to tides combined with storm surge,

(e)

(f)

(f)

(g) land use of sea frontage, and (h) possible long term increase in mean sea level.

(g) (h)

78

It is relatively easy in most cases to ensure that a seawall cope level and reclamation level is higher than extreme still water level corresponding to a reasonable return period of, say, 100 years, or even 200 years where flooding would cause substantial damage to lives and property. However, it is generally not possible or practical to design a seawall, even with the addition of a wave wall, to effectively prevent overtopping from waves during extreme events. A certain degree of overtopping, particularly with vertical-face seawalls, is always to be expected at times of extreme water level from the higher waves in the spectrum, and the drainage immediately behind the seawall should be designed to cater for this if major flooding is to be avoided. Wave walls are likely to be of only limited use in controlling or preventing extensive overtopping unless careful detailing is carried out at discontinuities such as piers, wharves, pumphouses and landing steps. It has been common practice in the past to use a nominal figure of +4.0 mCD for reclamation and seawall cope levels within the general harbour area. This figure should not be used for future reclamations without detailed consideration being given to the points listed above. All reclamations should be considered on an individual basis, and the consequences of adopted reclamation level fully investigated, before determining the most appropriate level in each case.

100 200

4.0

10.4 Reclamation Method 10.4.1 General

1 0 . 4 1 0 . 4 . 1

The design of a reclamation with regard to settlement is dependent on the method used to form the reclamation. Three main reclamation methods have been used in the past under local conditions, each one different in terms of reduction or control of settlement. These are : full removal of marine deposits, displacement of marine deposits, and controlled fill placement, combined with the use of vertical drains as necessary. Each of these methods is described briefly below, with comments on particular aspects, advantages and disadvantages in each case. Permanent

79

land usage is a major factor which should be considered before deciding which method of reclamation is to be adopted.

10.4.2

Marine Deposit Removal

1 0 . 4 . 2

Settlement can be controlled effectively by removing all marine deposits by dredging. Normally, dredging will stop when alluvium or weathered rock has been reached, as determined by detailed ground investigation. This can be expensive where thick marine deposits exists, and disposal of marine mud, particularly contaminated mud, may be severely problematic. However, the basic method is relatively simple, both dredging and filling operations can be carried out with minimal restrictions, and a relatively short overall construction programme can be achieved combined with an economical cost, even when large quantities of mud and fill are involved.

The removal of upper marine deposits only, with the lower, stiffer or stronger deposits remaining in place, has the advantage of reduction in dredging and fill quantities, and may be adequate in certain circumstances. Better control is required than for full marine deposit removal for the final trimming in dredging and for initial fill placement, to avoid future differential settlements, compared with full marine deposit removal. The remaining relatively thin layer of marine deposits will consolidate relatively quickly under vertical drainage through the fill during the filling operation, with the major part of the total settlement probably being completed during the reclamation formation period. The extent of marine deposits to be left should be subject to investigation and detailed design, and should depend on the magnitude of differential settlement which can be tolerated in the particular situation. Post-construction problems may arise if undetected deep pockets of mud are left in place.

10.4.3

Marine Deposit Displacement

1 0 . 4 . 3

This method involves direct tipping of fill, possibly

80

including public dump material (see Section 10.5.2), from trucks onto the surface of the marine deposits. Initial filling is often carried out along the lines of future roads and drainage reserves; as filling progresses, the softer upper marine deposits are displaced to the sides and in front of the reclamation bund or line. The displaced marine deposits or mud waves can often be controlled by forming the initial reclamation in bunds on grids. Where displaced marine deposits are not sufficiently dry and stable to be eventually covered by fill, they are removed by land based plant, transported to a barging point by truck and disposed of at a marine dumping site. Such removal is relatively expensive on a unit cost basis because of the multiple handling, but in relation to the overall project cost is usually not of major consequence; the total quantity of marine deposits required to be removed is usually small in relation to the total quantity covered and the total fill quantity.

1 0 . 5 . 2

Where initial filling is carried out on a wide front rather than in bunds, it is often possible for excess displaced marine deposits in front of the reclamation line to be removed by grab dredger and bottomdump barges, provided adequate access and water depths can be maintained. The unit cost for such removal will be significantly lower than for removal by land based plant, but the total quantity required to be removed will be larger, as less material will be covered by the fill. The reclamation method referred to above gives a very low initial cost for the reclamation, particularly where public dump material is used for filling. Reclamation by marine deposit displacement can be the most appropriate and economical method where public dump material is available, low initial cost is important and permanent development of the land so formed is not urgently required. However, where thick marine deposits occur, the method can result in long term settlement problems due to consolidation of the trapped marine deposits under the fill. Where initial reclamation is carried out by end-tipping on a wide front or in bunds, it is never possible to ensure that all marine deposits are displaced. Any remaining mud trapped under the fill will be of unknown and variable

81

thickness, and can give rise to differential settlement problems, related notonly to different settlement magnitudes but also to different periods for consolidation to take place. Where initially displaced marine deposits are isolated within bunds and later partially or completely covered by fill, the settlement problems are likely to be even greater, as these marine deposits will be from the upper layers and remoulded, and therefore of lower strength, and the overall thickness of marine deposits within the bunds will be greater than the original thickness. Differential settlements between areas formed as bunds and areas within bunds will usually be significant.

10.4.4

Controlled Thin Layer Fill Placement

1 0 . 4 . 4 1 2 5

This method leaves the existing marine deposits undisturbed, and involves the controlled placement of relatively thin layers of fill to avoid shear failure in the underlying soils and ; 'mud waves', usually using bottom-dump barges, hydraulic filling or grab. The initial capping or blanket layer is particularly important and should consist of free-draining granular material to ensure satisfactory vertical drainage from the marine deposits. Initial thicknesses of one to two metres are usually required, with subsequent layer thicknesses increased as appropriate. Layers must be staggered, with control being exercised to ensure a sufficient leading edge for the underlayer in relation to the layer being formed. When a total thickness of about 5 m of fill has been placed, the degree of control may be reduced, and it is often possible to complete the reclamation using end-tipping by truck, again in layers if necessary. Where the thickness of marine deposits is too great to ensure that settlements will have been substantially completed before final development of the reclamation is required, special measures to increase the rate of consolidation will be required. Such measures include the use of vertical drains, surcharging, or a combination of these measures.

Vertical drains are drainage conduits inserted into compressible soils of low permeability (typically, the

82

marine deposits) to enable the pore water in the soil matrix to drain horizontally to the conduits, and then vertically through the conduits to the free draining layer provided on top of the deposits, and to more permeable underlying soils. The rate of consolidation of the marine deposits will depend on the drain size, drain spacing and the properties of the deposits. Prefabricated paper or plastic band drains are the most commonly used in local conditions because of their resistance to shearing (Premchitt & To, 1991). A triangular spacing in the range of 1.5 to 3.0 m is normal, with primary consolidation, (which can be in the order of 20 to 30% of the thickness of the marine deposits), being completed within one to two years for typical sites with 5 to 10 m thickness of marine deposits, compared with 10 to 20 years if no vertical drains are used. To avoid the need for preboring through the reclamation fill, the band drains are commonly installed over water using special plant, through the blanket layer of free draining granular material, prior to the fill being placed. For additional information, reference may be made to the various published papers related to the test embankment at Chek Lap Kok.

(Premchitt & To, 1991) 1 . 5 3 . 0 5 1 0 20 30 1 2 1 0 2 0

The rate of consolidation of marine deposits can be increased by the application of a temporary surcharge load. Surcharge loads are usually applied immediately on completion of a part of the reclamation to design level, using fill. By careful planning, after completion of consolidation, the surplus surcharge fill can be used elsewhere in the reclamation, so reducing double handling costs to the minimum. When combined with the use of vertical drains, surcharging can be a very effective method of increasing the rate of consolidation of thick marine deposits.

Reclamation using controlled thin layer fill placement, combined with the use of vertical drains, surcharging or both where appropriate, can be the most suitable and economical method for large projects on thick marine deposits, where fill is costly, disposal of dredged mud is problematic, and development of the land is required within a year or so of the filling completion date. However, it should be noted that the techniques involved with thin layer fill

83

placement and vertical drain installation are relatively sophisticated compared with other reclamation methods, and require tight control and experienced contractors. With controlled thin layer fill placement, close monitoring on site is necessary and instrumentation, including piezometers, inclinometers, sub-surface and surface settlement measurement systems, and monitoring will be needed.

10.5 MISCELLANEOUS 10.5.1 General

1 0 . 5 1 0 . 5 . 1

Comments on particular aspects of reclamation design related to piling and culvert foundations are given below.

10.5.2

Piling

1 0 . 5 . 2 250 H

Obstructions to piling installed at reclamation sites can cause serious problems, with cost overruns and programme delays. The largest rock or boulder size encountered when driving piles must be breakable by impact or be capable of displacement. For the latter case there must be sufficient voids in the fill to cope with the pile volume and allow the displacement. Experience has shown that 250 mm is the approximate upper limit in rock or boulder size within the fill to ensure no major problems with the installation of the types of piles normally used, including driven concrete, steel H-section and steel tubular piles, and bored piles.

For bored piles, permanent steel liners are usually necessary to avoid necking, particularly in crushed rock fill or where pockets of unconsolidated marine deposits are suspected, as would be the case if marine deposit displacement had been used for the reclamation method. Special shoes or reinforced sections or points are normally required for driven piles, particularly in public dump material or crushed rock fill.

84

For all piles installed in reclamations, the possible effects of negative skin friction on the pile design should be assessed. The magnitude of the negative skin friction on any site will depend on the pile characteristics, fill characteristics, pile movement and the time since the reclamation was completed.

1 0 . 5 . 3

10.5.3

Culvert Foundations

Existing stormwater outfalls will need to be temporarily diverted away from the line of the future culvert extensions while reclamation is being carried out. The type of foundation to be used for a culvert extension across the new reclamation will depend on the method of reclamation used and the degree of consolidation of any remaining marine deposits. Where the marine deposits have been fully removed or where consolidation of any remaining deposits has effectively been completed before culvert construction commences, a normal foundation for the culvert, consisting of a nominal crushed rock or rubble layer, will usually be acceptable. Where marine deposit displacement has been used as the reclamation method, with a bund being formed along the line of the future culvert, a normal foundation as described above can usually be used following surcharging. Suitable surcharging would be carried out with surplus fill over a strip two to three times the width of the culvert and about four metres high for a period of three to six months until settlement has stabilised. Where consolidation of any remaining marine deposits has not effectively been completed or where variable thicknesses of underlying marine deposits are suspected, surcharging as outlined above can be of assistance in controlling settlement problems, but it is recommended that special measures are taken to allow for some differential settlement at the culvert joints. The use of piled foundations may be considered where differential settlement problems are expected to be particularly severe, but such a solution will be relatively expensive on an initial cost basis, and can result in other forms of settlement problems, as outlined below.

85

Normally culvert outfalls are formed in concrete block seawalls using special concrete outfall blocks. Such seawalls usually have rubble foundations and will experience some settlement with time, causing differential movement to develop at the junction with a piled culvert. In addition, differential settlement will also develop with time along each side of a piled culvert, as the adjacent reclamation material continues to settle over the underlying consolidating marine deposits.

86

87

11. CONSTRUCTION MATERIALS

11.

11.1

General

11.1

This Chapter gives comments and guidance on particular matters considered of importance for the design of marine structures concerning material selection, use and specification. The materials covered are armour rock, filling, concrete, steel, timber and rubber. For general information on these materials and any other materials used in marine structures, reference should be made to the General Specification for Civil Engineering Works (GS) (Hong Kong Government, 1992a). Comments on piles and protective measures in relation to material selection and use are also given in this Chapter.

General Specification for Civil Engineering Works GS (Hong Kong Government, 1992a)

11.2

Armour Rock

1 1 . 2 9.3.2 2 . 6 G S 2 1 G S B S 6 3 4 9 : P a r t 1 ( B S I , 1984a) 57.2 BS6349: Part 1 57.1 9.3.4 1 0 6 8

It is recommended in Section 9.3.2 that a specific gravity of 2.6 should normally be used for armour design; this figure corresponds with the minimum requirement for armour rock given in Section 21 of the GS. The requirements given in the GS for armour rock are in agreement with those given in Clause 57.2 of BS 6349:Part 1 (BSI, 1984a), although in this clause it is noted that lower standards of quality may be acceptable for stone to be used in layers other than the primary armouring. Clause 57.1 of BS 6349:Part 1 recommends that individual stones should be prismoidal in shape, with the maximum dimension generally not exceeding twice the minimum dimension, and never exceeding three times the minimum dimension. As stated in Section 9.3.4, the normal maximum armour size available locally in reasonable quantities is in the range 6 to 8 t, although sizes up to about 10 t may be available in small quantities.

11.3 11.3.1

Fill General

1 1 . 3 11.3.1

When placing fill under water, the material and

88

method of placement should be capable of achieving a relatively high density fill untreated, as external compaction is expensive. Care must be taken with the choice of bedding and filter materials to prevent loss of material from wave or current action and groundwater movements. Fill material placed immediately behind seawalls above Chart Datum should be free draining to avoid the unnecessary build up of water pressures due to tidal lag and ground water flow.

The choice of fill materials for use in a reclamation will be greatly dependent on availability and cost. The four main types of fill are public dump material, selected fill from land sources, crushed rock, and marine fill. Pulverized Fuel Ash (PFA) is another possible material, but the quantities available are relatively small, and it contains toxic contaminants. Use of PFA as fill is subject to the control of the Director of Environmental Protection. For most reclamations only one of the above types will normally be used, but there is no reason in principle why two or more could not be used on the same project, depending on availability with regard to the required programme and cost. Brief notes on the four main types are given below, with comments on particular aspects, advantages and disadvantages. Some general comments on fill suitability are also given herein. In order to conserve and ensure the most economic and effective use of fill from land and marine sources in Hong Kong, the Fill Management Committee was established in 1989 under the chairmanship of the Director of Civil Engineering to identify and manage the supply and demand of land and marine fill sources for all Government, quasi-Government and major private sector projects.

(PFA)

11.3.2

Public Dump Material

11.3.2

Public dumps are operated as a service to the general community to provide sites for the disposal of surplus material from private development and building demolition sites and to ensure that good use is made of such material, rather than it becoming a

89

wasted resource. The material brought to a dump is restricted by the conditions of the dumping licence to earth, inert building debris and broken rock and concrete. Since rock and concrete over 250 mm would impede subsequent piling, it has to be broken down to this size or dumped in areas where no building development will take place, e.g. in areas zoned for open space, road and drainage reserves, immediately behind seawalls etc. Where there is land access to the site, public dump material will be directly end-tipped by the licensed trucks onto the reclamation, under the control of dump supervisors, with spreading also carried out under the control of dump supervisors using hired plant. Normally, no compaction is carried out other than that caused directly by the passage of the dumping vehicles and spreading plant. Where possible, the tipping is carried out in two stages, firstly to a general level just above high water mark and later to the finished reclamation level.

250

Where direct land access to the site is not available, public dump material will be conveyed via ramps into barges, under the direction of dump supervisors, for transport to the reclamation site by contractor. Contract arrangements should ensure that barges are available at all times, to avoid the need for temporary stockpiling of dump material at the barge loading site. At the reclamation site, material will either be bottomdumped by barge or placed by grab, depending on the available water depth. Usually, completion of the reclamation to the required finished level is not possible using marine plant and direct grab placement; the use of temporary stockpiles, and distribution and final placement using trucks and spreading plant is common.

Disadvantages of using public dump material include the extreme variability of the material. The material is governed by the licence conditions but the quality is difficult to control, being mainly dependent on the type of development and demolition work from which it originates. Invariably, despite strict control, timber and other floating material are

90

discharged from trucks, particularly those laden with building debris. Supervisors should be especially vigilant, and should warn drivers of such trucks. Repeat offenders should have their licences revoked. Also, sampans should be employed to pick up floating materials from the water within the reclamation site. Such materials can be hazardous to shipping and are unsightly if allowed to drift into open waters; where possible, floating refuse booms should be used to surround the site to contain the debris and ease removal by the sampans. Another disadvantage is that the rate of supply of material to a public dump, being dependent mainly on private development projects, is largely not controllable, and planning work is particularly difficult if completion dates are critical. During the operation of a public dump it is quite usual for the number of trucks per day to vary widely. The number per day can typically vary from lows of 100 to 200 to highs of 1,000 to 1,500, with an average load per truck of 5 to 8 m3. This leads to another disadvantage in that traffic congestion at entrances and exits is unavoidable. The main advantage of using public dumping in reclamations, other than the benefit to the community, is low cost. The only costs relate to the provision of wheel washing facilities at the exit to the site, road cleaning adjacent to the site, provision of booms and sampans, provision of plant to spread the material, and supervisory staff.

1 0 0 2 0 0 1 0 0 0 1 5 0 0 5 8

11.3.3

Selected Fill

11.3.3

Selected fill material from a land source can be provided either by a contractor from his own source, with the quality subject to the specification in the contract, or from designated Government borrow areas or site formation projects. In the latter case some materials from the designated source may not comply with the specified quality. Weathered granites from such sources are generally considered to be suitable as fill for reclamations. However, weathered volcanic rocks should normally be used only above tide level, and preferably when mixed with other granular

91

materials, because of their typically finer grading and higher plasticity.

11.3.4

Crushed Rock

11.3.4

Crushed rock, uniformly graded and with maximum dimension about 250 mm, is suitable as a fill material for reclamations, but normally will be available only at a reasonable cost when provided as a by-product of a major public or private site development project involving large quantities of rock excavation and removal. For reclamations comprising crushed rock, consideration should be given to reduce the maximum size and mix the crushed rock with fine material in the top metre or so, and on the lines of drainage reserves, to allow easier future installation of services.

250

11.3.5

Marine Fill

11.3.5

The Fill Management Committee has taken over responsibility for identifying and investigating marine borrow areas, and for allocation of fill to individual projects. The use of marine fill can be economically viable for a reclamation project where a marine borrow area can be allocated by the Fill Management Committee within a reasonable distance of the site and where the size of the project justifies the use of sophisticated dredgers with high mobilisation costs. Such plant can dredge marine deposits at relatively low unit cost. Trailer dredgers can deposit marine sand in the reclamation by bottom dumping or by pumping, depending on the access and water depth available. The rate of formation of the reclamation can be extremely rapid compared with the use of other types of fill, and often several trailer dredgers can be used on the same site if necessary.

11.4 11.4.1

Concrete Reinforced and Prestressed Concrete in General

1 1 . 4 11.4.1

Recommendations for reinforced and prestressed

92

concrete, except where protected from direct exposure to the marine atmosphere, are in line with those given in Chapter 6 for suspended deck structures and are as follows : (a) (b) Minimum characteristic strength 45 MPa

(a) (b)

45MPa 0.40 60 mm

Maximum free water/cement ratio 0.40 60 mm

(c)(i) Nominal concrete cover fully immersed, and in tidal and splash zones (ii) Nominal concrete cover above splash zone

(c)(i)

50 mm

(ii)

50 mm

It is considered that criteria (a) and (b) above should apply irrespective of whether the concrete is fully immersed, within the tidal or splash zones or located above the normal splash zone. For concrete within the tidal and splash zone, crack widths under typical average long term loading conditions, as described in Chapter 6, should be limited to 0.1 mm. Where protected from direct exposure to the marine atmosphere, reinforced concrete should comply with the recommendations given in BS 8110:Part 1 (BSI, 1985a) for "moderate" exposure conditions.

(a) (b) 0.1 B S 8 1 1 0 : P a r t 1 ( B S I , 1 9 8 5 a )

11.4.2

Durability

11.4.2

The durability of reinforced concrete depends fundamentally on the quality and impermeability of the concrete, which prevent steel reinforcement from corroding. Normally concrete provides an alkaline environment which slows corrosive reactions. The best method of ensuring that this natural alkaline protective mechanism is maintained is by providing concrete which has the lowest possible permeability. This can be obtained by adopting mixes designed to produce concrete of the highest practicable density. However, care should be taken that requirements for high strength and low permeability do not result in the workability being restricted to less than that required for adequate compaction. The recommendation in

(b) 0.40 / /

93

criterion (b) above for a maximum free water/cement ratio of 0.40 is taken from the Interim Concrete Specification for Reinforced Concrete Structures in Marine Environment issued by the Standing Committee on Concrete Technology. In practice, the use of such a low water/cement ratio will require the use of a water reducing admixture, or superplasticizer, for many structures or elements, particularly those heavily reinforced, to ensure adequate workability and compaction. Particular care needs to be taken when detailing reinforcement to avoid congestion so that the concrete can be easily placed and subsequently compacted. It has to be noted that adequate curing of the concrete is essential to achieve the desired durability. For reinforced or prestressed concrete work in the tidal zone, precast units should be used wherever possible, with the minimum of in situ concrete connections; the use of epoxy coated reinforcement may be considered, and prestressed concrete should be avoided. Correct use of PFA as a cement replacement can reduce permeability and increase resistance to sulphate attack.

11.4.3

Unreinforced Concrete

11.4.3

For unreinforced concrete in massive sections, such as precast concrete seawall blocks and backing concrete for granite facing in seawalls, the use of concrete with a minimum characteristic strength of 20 MPa has been shown to be successful, with no significant maintenance problems. The continued use of such concrete for massive sections is recommended, irrespective of whether the concrete is fully immersed or within the tidal or splash zones, provided the concrete is actually placed 'in the dry'. It should be noted that Clause 6.2.4.2 and Table 6.2 of Part 1 of BS 8110:Part 1 recommend the use of higher grades of concrete to 'ensure long service life' for unreinforced concrete under all exposure conditions other than 'mild'. It is suggested that these recommendations should be followed for unreinforced concrete except for massive sections referred to above.

20 MPa B S 8 1 1 0 : P a r t 1 6 . 2 . 4 . 2 6 . 2

94 5 0

Partial replacement of cement by PFA in thick sections will reduce the effects of heat of hydration. Replacement of up to 50% of the cement may be considered if early strength is not critical.

11.4.4

11.4.4

Underwater Concrete
B S 6 3 4 9 : P a r t 1 5 8 . 1 1

Guidance on underwater concrete is given in Section 58.11 of BS 6349:Part 1. The use of epoxy coated reinforcement and a waterproofing admixture in the concrete mix are recommended. Reinforced concrete placed underwater should only be used where absolutely necessary, because of the difficulties of ensuring sound results and the problems of inspection. In particular, the use of concrete placed by tremie for forming heavily reinforced small elements such as pile caps within the fully immersed or tidal zones should be avoided, by making full use of precast units or requiring the use of watertight steel shutters, extended in height as necessary, to enable the concrete to be placed 'in the dry'. For driven steel tubular piles which are intended to be filled with reinforced concrete to one or two metres below sea-bed level, it is usually possible, after excavation, to form a plug in the bottom and pump the inside of the pile dry. This approach is preferred to the use of concrete placed by tremie. For bored concrete piles with a steel tubular casing, it will not usually be possible to pump the inside of the pile dry before concreting, and there will be no alternative to using reinforced concrete placed by tremie. For such piles, it is recommended that a minimum of three galvanised steel tubes, of a diameter suitable for possible future sonic testing, are cast in the concrete from founding to cut off level. It should be noted that concrete placed under water should not be designed for a characteristic strength greater than 25 MPa. It is recommended that this limitation should apply to bored piles formed by reinforced concrete placed by tremie due to the defects which can occur, but a higher grade of concrete should be specified in order to achieve this condition.

1 2

2 5 M P a

95 11.4.5

11.4.5

Design
B S 8 1 1 0 : P a r t 1 6 . 2

It is recommended that BS 8110:Part 1 should form the basis for the design of concrete structures. Information on partial load factors for suspended deck structures is given in Section 6.2; this can be used as a guide for other types of concrete structure.

1 1 . 5

11.5 11.5.1

Steel
11.5.1

Structural Steel in General


BS

Structural steel in marine structures should normally be weldable structural steel complying with BS 7668:1994 (BSI, 1994d), BS EN 10137, Parts 1, 2 & 3:1996 (BSI, 1996a, b & c), BS EN 10113 Parts 1, 2 & 3:1993 (BSI, 1993a, b & c), BS EN 10029:1991 (BSI, 1991a), BS EN 10155: 1993 (BSI, 1993d), or BS EN 10210-1 (BSI, 1994b) as appropriate. The use of BS 449:Part 2 (BSI, 1969) for the design of steel elements and structures is recommended, with increased permissible stresses for certain loading conditions, in line with the suggestions given in Chapter 6 for suspended deck structures.

7668:1994 (BSI, 1994d), BS EN 10137, Parts 1, 2 & 3:1996 (BSI, 1996a, b & c), BS EN 10113 Parts 1, 2 & 3:1993 (BSI, 1993a, b & c), BS EN 10029:1991 (BSI, 1991a), BS EN 10155: 1993 (BSI, 1993d), or BS EN 102101 (BSI, 1994b)
BS449: P a r t 2 ( B S I , 1 9 6 9 )

11.5.2

11.5.2

Corrosion Protection
B S 6 3 4 9 : P a r t 1 2 2 0 . 0 5

For steel structures and major steel elements, corrosion protection or allowances for metal losses due to corrosion, or both, will be a major consideration. It should be noted that the advice in Table 22 of BS 6349:Part 1, which gives typical upper rates of corrosion for structural steels in maritime conditions in the United Kingdom, is not recommended for use in Hong Kong. Hong Kong waters are relatively warm, and contain various pollutants whose effect on steel is generally unknown. In many sites, the presence of anaerobic sulphatereducing bacteria, which can greatly increase normal steel corrosion rates, is also suspected. In the absence of full scale long-term tests covering metal loss from corrosion in Hong Kong waters, it is recommended that all structural steelwork above sea-bed level,

96

whether fully immersed, within the tidal or splash zones, or generally above the splash zone, is fully protected against corrosion for the design life of the structure. Below sea-bed level, an allowance for corrosion loss of 0.05 mm per year on the outside face of steel piles is considered reasonable, if no corrosion protection is carried out within this zone. For guidance on protective measures which can be taken against corrosion, see Section 11.9.

11.9

11.5.3

11.5.3

Use of Stainless Steel


G S 2 1 B S 9 7 0 : P a r t 1 ( B S I , 1 9 9 6 d ) B S 1 4 4 9 : P a r t 2 ( B S I , 1 9 8 3 ) BS EN 10088: Parts 1, 2 & 3 (BSI, 1995a, b & c) 316 304

Section 21 of the General Specification for Civil Engineering Works (Hong Kong Government, 1992a) requires stainless steel for elements in marine works such as chains, railings, cat ladders, pumphouse screens and screen guides, mooring eyes and other fittings to be austenitic stainless steel grade 316 complying with BS 970:Part 1 (BSI, 1996d), BS 1449:Part 2 (BSI, 1983), or BS EN 10088: Parts 1, 2 & 3 (BSI, 1995a, b & c) as appropriate. It should be noted that the commonly available grade 304 stainless steel is not suitable for use in a marine environment due to the presence of chlorides. The selection of the correct grade of stainless steel at the design stage is most important, as corrosion in stainless steel members and fasteners may not be readily evident. In stress corrosion cracking, corrosion occurs along grain boundaries, and there may be no corrosion product evident, or only slight staining. A visual examination may not show this cracking, even though the member or fastener is about to fail.

11.5.4

11.5.4

General Guidance
B S 6 3 4 9 : P a r t 1 5 9

General guidance on the use of structural steel and other metals in marine structures is given in Clause 59 of BS 6349:Part 1. Important points to note are as follows : (a) Fabrication details should be kept as simple as possible and should be designed to avoid corrosion and facilitate maintenance.

(a)

(b) Tolerances for on-site connections should be

(b)

97

generous because of the difficulties associated with working in a marine environment. (c) As much prefabrication as possible should be undertaken, taking advantage of mechanised welding and early painting under factorycontrolled conditions.
(c)

(d) Steel embedded in concrete is cathodic relative to the same steel in seawater, and rapid corrosion will therefore occur at the interface of a partly embedded member unless special treatment is carried out, e.g. use of sacrificial anodes and impressed currents.

(d) (e)

(e)

Chemical composition of steels has less influence on corrosion rates in a marine environment than physical factors such as the roughness of the surface finish of the steel and the presence of holes and re-entrant corners, all of which tend to promote the formation of galvanic corrosion cells.

1 1 . 6

11.6 11.6.1

Timber
11.6.1

Types of Material
G S 2 1 ( S e l a n g a n B a t uY a c a l B a l a u ) B S 5 7 5 6 ( B S I , 1 9 8 0 ) (HS) BS5268: Part 2 (BSI, 1991b) SC8SC9

Section 21 of the GS requires timber for fenders to be "Selangan Batu" (also known as "Yacal" and "Balau") or similar species of hardwood, visually stress graded to the Hardwood Structural (HS) grade of BS 5756 (BSI, 1980), and to satisfy the strength requirements for strength class SC 8 or SC 9 of BS 5268:Part 2 (BSI, 1991b).

11.6.2

11.6.2

Design Stress
B S 5 2 6 8 : P a r t 2 9 S C 7 BS5268: Part 2 16 K2 9

For the design of timber fenders and walings, it is recommended that the grade stresses and moduli of elasticity for strength class SC 7 given in Table 9 of BS 5268:Part 2 should be used, after application of the modification factor K 2 for wet exposure conditions given in Table 16 of BS 5268:Part 2. It

98

should be noted that the grade stresses given in Table 9 apply to long term loading, and Table 17 of BS 5268:Part 2 gives details of the modification factor K3 by which these stresses can be multiplied for various other durations of loading. It is recommended that any combination of loading with berthing or wave loads may be considered 'very short term' for the purposes of assessment of a value for K3 from Table 17. Note should be taken of Clause 14.6 of BS 5268:Part 2, which covers the depth modification factor K7 which is used for adjusting the grade bending stresses given in Table 9, where the depth of the member being designed is other than 300 mm.

BS5268: Part 2 17 K317 K3

3 0 0 BS5268: Part 2 14.6 K7 9

Clause 2.8 of BS 5268:Part 2 defines the grade stress as the stress which can safely be permanently sustained by material of a specific section size and of a particular strength class or species and grade. By implication, grade stresses, adjusted where appropriate using the modification factors, should be used for the design of timber structures or timber members, where it is required that the structure or member should be able to last without replacement for the design life of the structure. It is not considered necessary or economical to design a timber fendering system for the full design life of a suspended deck structure, because replacement of individual members is relatively simple and inexpensive, and damage or failure of part of the fendering system will normally result in normal minor damage, if any, to the suspended deck structure itself. For timber, the variation in material properties within any strength class is particularly large, and to design timber fendering systems based on the grade stress, which is based on the 5% lower exclusion values of strength, would not make the best use of the material, and would result in larger timber sections, steel bolts and fixings being required than have been used in the past and are commonly available.

B S 5 2 6 8 : P a r t 2 2 . 8 5

11.6.3

11.6.3

Loading Factors
BS5268:

For the reasons given above, it is recommended

99

that the permissible stresses obtained from the grade stresses and modification factors in BS 5268:Part 2 should be increased by multiplying by a design factor for the design of timber fendering systems. Suggested values for this design factor for different members and loading conditions are as follows :

Part 2

1.6 2.0 2.0 2.4

Loading Condition Normal Extreme Temporary Accident

Factor 1.6 2.0 2.0 2.4


The factors proposed above relate to fendering systems for typical suspended deck structures. At the discretion of the designer, these factors may be increased or decreased to suit the structure use and location. For a public pier with relatively low use, for example, where the design vessel is not expected to berth frequently, an increase in the factors of up to 20% may be considered, whereas for a heavily used ferry pier, where the design vessel will berth frequently, a decrease in the factors of up to 20% would be more reasonable. When designing timber fenders for public piers and ferry piers, for vessels with displacement not exceeding about 1500 t, it is normal to assume for the design condition that the vessel contacts two vertical fenders simultaneously, provided that the clear distance between the fenders is not greater than about 800 mm. It is not normal, when designing timber fenders for public piers or ferry piers, to make an allowance for material loss due to wear of the faces of the fenders by the vessels during berthing. However, such an allowance should be considered by the designer when designing a ferry pier which is expected to be heavily used by vessels with steel-faced rubbing or berthing strips. In such a situation, the fender design should be checked with the ferry operator, but in any event it is not recommended that the fender size should be increased to an extent that this results in a size of fender and fixing outside the normal range.

2 0 20

1500 800

100 1 1 . 7

11.7

Rubber

Section 21 of the GS requires rubber for fenders to be resistant to aging, weathering and wearing, to be homogeneous, free from any defects or impurities, pores or cracks and to have certain defined properties as covered by parts of BS 903 (BSI, several parts, 1976 to 1995). Sources of information on types of rubber fenders are given in Section 4.13. Information given in the major manufacturers' catalogues concerning fender reaction, deformation and energy characteristics may generally be accepted with confidence. Before finalising a rubber fender design, advice should always be sought from one or more of the major reputable suppliers regarding suitability for the project. Wherever possible, rubber fenders should be selected or specified to match existing fenders, to minimise the different types of fenders required to be kept in stock for future maintenance.

G S 2 1 BS903 (BSI, several parts, 1976 1 9 9 5 ) 4 . 1 3

11.8

11.8 11.8.1

Piles General

11.8.1

Section 8 of the GS covers piling works. For general information on types of piles and their suitability for different ground conditions, locations and types of structure, see Section 7.4 of BS 8004 (BSI, 1986) and Chapter 2 of Tomlinson (1987). Useful information on piles in maritime works is given in Clause 61 of BS 6349:Part 1.

GS 8 BS8004 (BSI, 1986) 7.4 T o m l i n s o n ( 1 9 8 7 ) BS6349: P a r t 1 6 1

11.8.2

11.8.2

Driven Concrete Piles

For the types of marine structure covered by this Manual, the choice of bearing pile will generally be limited to driven prestressed concrete, driven tubular steel and bored cast in situ concrete piles, and minipiles. Driven prestressed concrete piles have been successfully used in local conditions for many years, but have disadvantages related to loading and size/weight/length limitations, problems with

/ /

101

extensions, driving through boulders and obstructions, and can give programming problems due to the need for precasting. The provision of a suitably large works area for a precasting yard with marine frontage is becoming increasingly difficult to arrange as pressure for waterfront sites increases.
11.8.3

11.8.3

Tubular Steel Piles

Driven tubular steel piles are becoming increasingly widely used, due to their flexibility and general ease of use related to range of load, diameter, length, ease of extension and storage. Steel tubular piles can be either fully protected against corrosion or filled with reinforced concrete after driving, with the steel tube in this case often being left unprotected above seabed level. To reduce the possibility of long term maintenance problems, preference is for steel tubular piles to be infilled with concrete to at least below seabed level and for the steel tube above seabed level to be considered as sacrificial and ignored for design purposes, the length of pile above seabed level becoming in effect a reinforced concrete cast in situ pile. Such reinforced concrete should follow the recommendations of Section 11.4, with the increase in durability provided by the steel 'casing' considered an additional benefit. With driven tubular steel piles, it is possible to overcome problems with obstructions, by chiselling inside the pile if driven open ended, or by temporary extraction, followed by the installation of an oversize casing through the obstruction and subsequent redriving within the casing.

1 1 . 4

11.8.4

11.8.4

Bored Piles

Bored cast in situ concrete piles should be avoided where possible because they generally require the use of a tremie, as discussed in Section 11.4, although any loss of durability caused by the use of concrete placed by tremie will be offset to some extent by the presence of the 'sacrificial' steel casing above seabed level. The advantages of this type of pile relate to the ability to carry high loads and the relative ease with which boulders and obstructions can be dealt with. High

11.4

102

cost, due to the need to use specialist plant to install a large diameter casing over water, and to excavate within this casing, is usually a major disadvantage.
11.8.5

11.8.5

Fender Piles

Fender piles should, where possible, be dispensed with by designing structures and fendering systems in such a way that berthing loads are distributed directly to the main structure. Where fender piles are considered essential, they should be of steel or prestressed concrete and should be capable of being replaced during the design life of the structure, because of the possibility of accidental damage. Such replacement can best be arranged by locating the fender piles in sockets at or slightly below seabed level, rather than by driving the fender piles into the seabed as for normal driven bearing piles. Such sockets can be formed in precast concrete blocks surrounded by and resting on rubble. Replacement of driven fender piles can be extremely difficult or unsatisfactory due to the proximity of the structure, often with an overhanging or projecting roof or upper deck, and to disturbance of the seabed due to the driving and extraction of the original fender piles.

11.8.6

11.8.6

Sheet Piles

For the types of marine structure covered by this Manual, permanent sheet piles will not commonly be used. Where they are used, the common types of interlocking steel sheet piles will usually be suitable. Comments are given in Section 7.2 regarding material loss and corrosion protection for steel sheet piles.

7.2

1 1 . 9

11.9 11.9.1

Protective Measures General

11.9.1

Information on protective measures which can be used to stop or reduce deterioration in marine

B S 6 3 4 9 : P a r t 1 6 8 6.5 7.2 11.5

103

structures is given in Section 68 of BS 6349:Part 1 (BSI, 1984a). Comments on corrosion losses for steel in local conditions are given in Sections 6.5, 7.2 and 11.5. BS 5493 (BSI, 1977) gives valuable guidance on the choice and specification of coating systems available, although it should be noted that the definitions of environment and recommendations for coatings are primarily related to conditions in the United Kingdom; local conditions are likely to be more corrosive, due mainly to higher air and sea temperatures, and humidity.

BS5493 (BSI, 1977)

11.9.2

11.9.2

Life of Protective Coatings

Where access for repair or maintenance is not possible or extremely difficult, the initial protective coating will be required to have the same life as the structure. Where access for repair or maintenance is readily available, the life requirement of the initial protective coating is based on the time which can elapse before major or general maintenance of the coating becomes necessary. That time is referred to as the 'time to first maintenance' in BS 5493. BS 5493 gives the following recommendations for common protective coatings in United Kingdom conditions for the seawater splash zone, frequent salt spray, or immersion in seawater :

B S 5 4 9 3 BS5493

( )

Typical Time to First Maintenance

General Description with Total Nominal Thickness (m) (a)Galvanising (85 min.) plus Coal Tar Epoxy(150) (b)Coal Tar Epoxy (450)

(1020)

a) (85) (150) b) (450)

Long (10 to 20 years)

(510)

a) (140) b) (350)

Medium (5 to 10 years)

(a)Galvanising (140) (b)Coal Tar Epoxy (350)

(5)

a) (85) b) (250)

Short

(a)Galvanising (85 min.)

104

(less than 5 years)

(b)Coal Tar Epoxy (250)

11.9.3

11.9.3

Important Points to Be Considered

B S 6 3 4 9 : P a r t 1 6 8 . 1 (a) BS5493 A (b)

The following notes, which are taken partly from Clause 68.1 of BS 6349:Part 1, are particularly important when considering protective systems : (a) Potential corrosion hazards can be eliminated by careful design. Reference should be made to Appendix A of BS 5493.

(b) The costs of protective measures are repetitive in that the protective materials themselves deteriorate, and regular maintenance and renewal of coatings will be necessary for all structures except those with relatively short design lives. (c) For important, heavily used structures, the need for regular maintenance and renewal of coatings should not be allowed to restrict normal use of the structures.

(c)

(d) Corrosion does not proceed at a uniform rate over the whole structure or member, and at certain corrosion points, loss of the original material can be much more rapid than expected; any estimate of a corrosion allowance is likely to be excessive for some parts while being inadequate for others. (e) The cost of renewing a protective system is likely to be much more than the initial protection due to the need to remove marine growth and old paint prior to renewal of the system, and the fact that access will usually be more difficult than during construction. Marine growth is prevalent on structures below mean high water level. Evidence exists that such growth can be protective against corrosion and therefore generally should not be removed, as it may be more effective and durable than a paint system which might replace it. Normally the only exposure zones which might usefully be

(d)

(e)

(f)

(f)

105

repainted are the splash and atmospheric zones.

11.9.4

11.9.4

Corrosion Protection of Steel Tubular Piles

For the corrosion protection of steel tubular piles, covering the immersed, tidal and splash zones, it is unlikely that coal tar epoxy, even when applied with a total dry film thickness of 400 to 500 m, will have an effective life of more than 15 to 20 years, under the most favourable conditions and assuming no damage during handling and driving. The use of polyethylene sheeting for coating steel tubular piles may be considered, and in theory such a system should be able to offer full corrosion protection for the normal structure design life of 50 years. The polyethylene sheeting is normally several millimetres thick and is applied under controlled factory conditions by heatshrinking on to the outside of the steel tube, which has been treated with undercoat/primer and an adhesive layer. As for all pile coatings, extreme care is necessary not to cause damage during handling and driving. The main disadvantages relate to lack of experience with the material amongst local contractors and lack of full scale long term durability test data under local conditions.

400500 15 20 50 /

The use of normal site-applied wrappings consisting of mastics or primers wrapped with petroleum jelly impregnated tapes and PVC or polyethylene outerwraps are generally limited to tie rods, pipes and tubular steel piles in the splash zone. Wrapping systems are available for treating tubular steel piles and other steel members below water level, but these systems are relatively expensive and need only be considered for critical repair work where no other solution is apparent. For all proprietary coatings and wrappings, where site application is unavoidable, the advice of the manufacturer, particularly with regard to surface preparation, should be strictly followed and close supervision maintained. The electrochemical processes which accompany corrosion of submerged steel elements in seawater are

BS7361 (BSI, 1991c)

106

described in BS 7361 (BSI, 1991c), which also gives details of the way in which cathodic protection, both by sacrificial galvanic anodes or by impressed current systems, should be applied to combat corrosion. Cathodic protection is usually considered to be fully effective up to about half-tide mark. The main disadvantage of using a cathodic protection system for marine structures covered by this Manual, apart from possible increased costs compared with other designs and protection methods, relates to a general lack of expertise with regard to monitoring work after installation. It is recommended that the detailed design for any cathodic protection system should be entrusted to a suitably qualified specialist company and an operating and maintenance manual should be provided. For monitoring work after installation, consideration should be given to arranging a maintenance contract with a suitably qualified specialist.

107

12. TYPES OF STRUCTURE

12.

12.1

General

12.1

This Chapter gives guidance on types of structure and includes recommendations on preferred forms of construction and choice of materials where appropriate. Notes on design criteria are given where these have not been covered in other Chapters. Types of structure covered include breakwaters, seawalls, piers, dolphins, pumphouses, slipways, ramps, navigation aids, outfalls, intakes and miscellaneous minor structures.

12.2 Breakwaters Information on types of breakwater, with examples, is given in pages 6-88 to 6-94 of the Shore Protection Manual (SPM) (CERC, 1984) and Chapter 3 of Bruun (1981). Vertical wall breakwaters are rigid structures; they are generally impermeable and cause full wave reflection, thereby attracting large wave forces and increasing wave activity adjacent to the breakwater. Main types of vertical wall breakwaters include precast concrete blockwork, concrete caissons and cellular steel sheet piles. Rubble mound breakwaters have sloping faces and are flexible structures; they have significant permeability, are able to dissipate wave energy and greatly reduce wave reflection. Outer protective armour is of rock or precast concrete units. Composite breakwaters consist of a rigid vertical wall construction on a rubble mound foundation. Depending on the relative depth of the base of the vertical wall, some wave dissipation usually occurs, resulting in only Partial wave reflection. Where the base of the vertical wall is relatively high, problems can occur with wave scour at the top of the rubble foundation. For local conditions, rubble mound breakwaters are normally preferred to vertical wall or composite breakwaters. Rubble mound breakwaters have the following advantages :

12.2

Shore Protection Manual ( SPM)(CERC, 1984) 6 - 886 94Bruun (1981)

108

(a)

Cost is relatively low due to availability of local rock for core, underlayers and armour, except for the most exposed locations.

(a) (b)

(b) Wave conditions are improved due to wave absorption and minimal reflection, both behind and in front of the breakwater. (c) Overtopping is easier to control due to greatly reduced runup compared with vertical wall construction.

(c)

(d) Damage is progressive and sudden failure is unlikely, due to the flexible form of construction. (e) Repairs and maintenance can be carried out relatively easily.

(d)

(e)

One disadvantage is that floating refuse can be trapped easily in the rubble, and this is environmentally undesirable. Guidance on the design of rubble mound breakwaters is given in Chapter 9. As stated in Section 9.4, concrete caps or wave walls should be avoided where possible. Navigation lights should be provided on breakwaters to the requirements of the Director of Marine. Access steps, in the form of precast concrete blocks, should be provided for light posts down to low tide level on the protected faces. For typhoon shelters and other similar breakwaters, bollards or mooring eyes may be required by the Director of Marine to be constructed at locations on the protected inner faces. Innovative designs may be required where conventional breakwater design would result in unacceptable reduction of water quality due to obstruction of currents.

9.4

12.3 Seawalls It is important to distinguish between seawalls whose only function is to protect the adjacent land

12.3

109

from erosion by the sea, and seawalls which, in addition, have a secondary, and sometimes major, function of providing a berth for vessels. The term 'seawall' should strictly only be applied to the first type, where no berthing is possible or allowed for, and the term 'quay wall' should be used where berthing is possible or a primary requirement. Details of types of seawall are given in pages 6-1 to 6-14 of the SPM. The structure may be rigid, flexible or of some intermediate type, the choice of type being heavily dependent on the degree of exposure of the site to wave attack. For local conditions, seawalls of rubble mound construction are preferred. Similar advantages over rigid impermeable structures apply as given in Section 12.2 for rubble mound breakwaters. Guidance on design is given in Chapter 9; comments on concrete caps and wave walls are given in Section 9.4. The structure will be permeable and flexible, and any handrail and pavement edge should be set well back from the seawall crest to minimise maintenance problems.

S P M 6 - 1 6 - 1 4

12.2 9 . 4

Details of various types of quay wall, with examples, are given in Chapter 4 of Bruun (1981). Quay walls can be divided into two main types : sheet walls and gravity walls. Types of sheet wall structure and sheet wall are given in Sections 4.3 and 4.4 of BS 6349:Part 2 (BSI, 1988). Types of gravity wall are given in Section 5.2 of BS 6349:Part 2. In cases where excavation of soft material from the seabed is not practicable, alternative foundation designs, for example employing in-situ strengthening of soft material should be considered. For local conditions, with required water depth alongside not exceeding about 7 metres, gravity quay walls of concrete block construction, using standard concrete blocks and standard seawall sections where appropriate, are preferred. Such concrete block quay walls have the following advantages :

Bruun (1981) BS6349: Part 2 (BSI, 1988) 4.34.4 BS6349: Part 25.2

(a)

Cost is relatively low due to experience of local contractors with this type of construction and

(a)

110

efficient use of materials. (b) The form of construction has a long history of satisfactory performance with negligible need for maintenance. (c) The construction has some flexibility and can cope with some differential foundation settlement without risk of sudden failure.

(b)

(c)

(d) Modification work is relatively simple, and damage from vessels in accidents is usually minor. (e) Incorporation of landings, pumphouses and drainage outfalls into lengths of wall is relatively simple. The form of construction can be used for a large range of sea bed depths and subsoil conditions by varying the dredging depth and extent of rubble foundation to suit.

(d)

(e)

(f)

(f)

Disadvantages of concrete block quay walls relate mainly to the relatively long construction period required and the need for a large casting yard and stacking area with marine frontage; however, these disadvantages can generally be reduced in significance with adequate project planning. Another disadvantage is that vertical walls reflect waves, with the consequence that adjacent wave activity is increased. It should be noted that standard quay wall sections (CED, 1991) with toe levels at approximately +0.30, -1.05, -2.40, -3.75, -5.05 and -6.40 mCD incorporating standard concrete blocks are suitable only for use in relatively protected locations not exposed to severe waves, where the required ground level is approximately +4.0 mCD and the required live load behind the seawall is limited to 10 kPa. For other locations, ground levels and loadings, it may still be possible to use the standard sections, but the design must be checked in detail. Where checks show that the standard sections are not satisfactory, factors of safety against sliding and overturning can be improved by :

(CED, 1991) + 0 . 3 0- 1 . 0 5- 2 . 4 0- 3 . 7 5 -5.05 -6.40 mCD +4.0 mCD 1 0 k P a

111

(a)

incorporating keys at horizontal joints,

(a) (b) (c)

(b) widening of the section at critical levels, and (c) improving the quality of the fill normally placed immediately behind the seawall.

For all concrete block quay walls, whether standard sections are used or not, care should be taken to check the stability of the bermstones and adjacent rubble foundation against possible scour and undermining from wave attack; this is particularly important for walls with shallow toe levels in exposed locations.

Toe levels of concrete block quay walls should be determined following consultation with the Director of Marine, who will also advise on locations of public landings, bollards, mooring eyes and ladders. It is normal to provide standard timber fendering systems for public landings, but for public water fronts and public cargo working areas, fenders need not be provided. Mooring eyes, ladders and handrails should preferably be of stainless steel, particularly where located within the tidal, splash or spray zones. Notice boards should be provided as required by the Director of Marine.

12.4 Piers 12.4.1 General

12.4 12.4.1

Details of types of piers, with examples, are given in Chapter 4 of Bruun (1981). The majority of piers are suspended deck structures; examples of types of suspended deck structure are given in Section 6.4, Table 1 and Figure 41 of BS 6349:Part 2. Figure 44 of BS 6349:Part 2 gives examples of structural arrangements used to resist berthing loads, and Figure 45 gives examples of pile and deck connections. For local conditions, preference is for reinforced or prestressed concrete suspended decks, with maximum use of precasting for individual members, and prestressed concrete piles or steel tubular piles, infilled

Bruun (1981) BS6349: P a r t 2 6 . 4 1 4 1 BS6349: Part 2 44 4 5

112

with reinforced concrete to below seabed level, with the steel casing above seabed level being treated as non-structural. The advantages of such construction relate mainly to minimising maintenance costs.

For comments on standards for ferry piers and the siting of ferry piers, see Volume 9 of the Transport Planning & Design Manual (Transport Department, 1986). Vessel design requirements and layout requirements for ferry piers and public piers should be agreed with the Commissioner for Transport, the Director of Marine, and the proposed ferry operators as appropriate. For public piers, it is suggested that the pier structure and fendering system should be designed for use by vessels up to about 400 t displacement; this will allow for possible use by existing double deck ferries in an emergency.

Transport Planning & Design Manual (Transport Department, 1986) 400

12.4.2

12.4.2

Public Piers
0.6 0.9 230 4 0 1 . 5 2 . 0 +2.5 +1.0 mCD +0.3 mCD +3.8 mCD 81010

Timber fenders at about 0.6 to 0.9 m centres, with timber walings and cylindrical rubber end-on buffers, are considered most appropriate for use for public piers, because of the need to cater for a large range of vessel size. Sets of landing steps, preferably 2 m wide, should be provided at minimum 30 to 40 m centres; each set should have two landings, each 1.5 to 2.0 metres long, at approximate levels of +2.5 and +1.0 mCD. Fenders should extend from about +0.3 mCD to at least +3.8 mCD. These levels are applicable to tidal conditions in Victoria Harbour. At locations which have different tidal ranges, these levels should be reviewed. A top capping piece should be provided to prevent fouling by ropes, and timber step blocks provided at the landing steps. Standard 10 t bollards at about 8 to 10 m centres are considered appropriate, with a reasonable number of mooring eyes or cleats at each set of landing steps. Handrails, preferably of stainless steel, should be provided at landing steps. Navigation lights should be provided to the requirements of the Director of Marine.

113 12.4.3

12.4.3

Ferry Piers

Fendering systems for ferry piers should be designed on an individual basis to suit the expected vessel use, after consultation with the proposed ferry operator, as the operator will generally be responsible for maintenance of the fendering system under the terms of his franchise or licence. Locations of bollards and other fixtures and fittings should be decided in a similar manner after consultation with the proposed ferry operator. For ferry piers and single user piers where vessels have berthing or rubbing strips, and the range of vessel size using each pier is restricted, although a timber fendering system may still prove to be satisfactory, consideration should be given to providing other types of fendering system, which may be more cost effective in the long term. One such suitable type of system consists of steel frontal frames with nylon or polyethylene wearing pads supported on rubber cell fenders. The main advantages of such a system are :

(a)

The deflection/energy/reaction characteristics are such that relatively constant low reactions will apply for a range of energy absorption.

(a) / / (b) (c) (d) 1 1 . 8

(b) For relatively light, thin hulled vessels such as hydrofoils, the reaction on the vessel berthing or rubbing strips can be kept within acceptable limits by selecting suitably wide frontal frames. (c) The lengths of the frontal frames can be selected to cover the full required tidal range.

(d) By rotating the nylon or polyethylene wearing pads, even wear rates can be maintained and efficient use made of the pads. This is not possible with timber fenders, where material wear is concentrated. Comments on the use of fender piles are given in Section 11.8. Particular attention should be paid to the possible need to replace fender piles during the design life of the pier.

114 12.5

12.5 Dolphins Useful general comments on dolphins and examples of different types of flexible and rigid structures are given in Section 7 of BS 6349:Part 2. The comments given in Section 12.4 regarding preferred materials for construction of piers apply equally to dolphins. Fenders should be provided for mooring dolphins because of the possibility of accidental vessel impact. Fendering systems for both berthing (breasting) dolphins and mooring dolphins should be simple and robust, with the aim of minimising future maintenance. The Director of Marine should be consulted on the need for navigation lights on tops of dolphins.
BS6349: Part 2 7 1 2 . 4

12.6

12.6 Pumphouses
12.6.1

12.6.1

General

Pumphouses covered by this Manual include sets of individual small units, interconnected small units and larger units for installation of pumps for providing salt water for buildings, usually for air-conditioning purposes. 12.6.2 Layout and Location

12.6.2

In the design and construction of pumphouses covered by this Manual, the requirements of the client for such details as size, layout, facilities and fittings should be provided by or agreed in advance with the client. The following points should be noted when selecting a site for a pumphouse : (a) The intake should be remote from sewage outfalls and other sources of contamination and debris, and also from salt water outlets which discharge heated water.

(a)

(b) The seabed should be sufficiently deep to accommodate the intake, after allowance for silting. (c) The water in front of the intake should not be

(b)

(c)

115

stagnant and the adjacent seawall should not be used by vessels for berthing.

12.6.3

12.6.3

Structure and Design


BS8007 (BSI,1987)

Pumphouses normally consist of reinforced concrete units, precast where placed below general water level and cast in situ above water level. To ensure that the units are watertight, it is recommended that the design of all walls and base slabs in contact with seawater should be in accordance with BS 8007 (BSI, 1987). Pumphouse units are usually constructed as Part of a concrete blockwork quay wall founded on concrete blocks or on an extended rubble seawall foundation. To avoid possible future settlement problems, it is important that the underlying ground is consolidated by preloading before the setting of the pumphouse units. This preloading is particularly important where a pumphouse is to be constructed as an extension to, or immediately behind, an existing quay wall. The pumphouse units are connected to the sea by intakes formed in special precast concrete blocks. To ensure satisfactory operation of the pumps in all tidal and wave conditions, it is recommended that the crown of the intake should generally be at a level not higher than -0.6 mCD, see Section 12.9. For ease of construction and to minimise the number of joints, precast pumphouse units should be individual self-contained units with walls formed to as high a level as possible, subject to weight limitations, and preferably to a level between MSL (+1.4 mCD) and MHHW (+2.1 mCD) for harbour locations. For larger pumphouses, sets of units can be interconnected above the junction between the precast and in situ concrete level. It is usual for precast pumphouse units to be cast on a waterfront site, lifted by crane or crane barge, transported to the pumphouse site by barge, and set in position by crane or crane barge. For this method of construction, the weight of an individual unit is limited by the lifting capacity of available plant; units within the weight range of 50 to 100 t are relatively common. Another method of construction is

- 0 . 6 mCD12.9

(+1.4 mCD) (+2.1 m C D ) 50 100

116

for the unit to be launched on a slipway after casting, floated, towed to the pumphouse site and set in position by crane or crane barge.
2 4

When using the construction method referred to above, involving transport by barge, it is usual to test each unit at the casting yard for watertightness by filling the unit with water and leaving it filled for at least 24 hours. Although this method of testing does not fairly reflect normal water pressures during pumphouse operation, it is far simpler and less expensive than immersing the unit in water. Whichever method of testing is adopted, it is important that, during the design stage, the test loading condition is also checked and the reinforcement designed and detailed accordingly. Water or sand is usually used as ballast during the placing of the precast pumphouse units to guard against buoyancy. Such ballast should not be removed until a careful design check is made on the buoyancy of the structure.

12.6.4

12.6.4

Ties and Waterstops


BS8007 B S 8 0 0 7 BS8007 D BS8007

When drafting Particular Specification clauses, it should be noted that BS 8007 recommends that ties used to secure and align formwork should not pass completely through any liquid-retaining Part of the structure, unless effective precautions can be taken to ensure watertightness after their removal. The ends of any embedded ties should have cover equal to that required for the reinforcement. The gap left from the end of the tie to the face of the concrete should be effectively sealed. Although it has been common practice to provide central waterstops and keys at construction joints between the precast units and in situ concrete sections, BS 8007 states that waterstops are not usually required for construction joints with complete continuity in water-retaining structures and, in Appendix D of BS 8007, an example of a construction joint with no waterstop or key is shown. Central waterstops can be difficult to fix and hold in position during concreting, and problems can be experienced when placing and compacting concrete around the waterstop. As explained in BS 8007,

117

whether or not a centre waterstop is used, extreme care should be taken during surface preparation for construction joints in pumphouse unit walls.

12.6.5

12.6.5

Screens, Guides and Fittings


Pumphouse intake screen guides may be stainless steel or cast iron sections bolted onto the outside of the concrete intake blocks, or formed directly as a recess in the concrete intake blocks themselves. For the former case, the guides should be protected from damage by vessel impact using securely fixed timber fenders. For the latter case, the concrete nib between the recess and the outer block face should be detailed with care, with stainless steel sections being used as necessary to protect and line the recess. Problems have been experienced with such nibs due to inadequate thickness and concrete cover, combined with inappropriate reinforcement detailing.

Screens may be of stainless steel, PVC tubing, hardwood, or a combination of materials as specified by, or agreed with, the client. For sets of pumphouse units constructed for possible future use by Parties other than the client, it is suggested that fabrication of the screens should be made the responsibility of the eventual users. The time between construction of the pumphouse and handover can often be several years; if the screens are placed in the guides immediately after construction, later removal will be difficult due to marine growth, and the alternative of storage for an indefinite period is not efficient or appropriate.

Internal and external steel fittings and fixtures, such as ladders, gratings, guide covers and runway beams, should be stainless, galvanised or painted with coal tar epoxy, as agreed with the user. To protect the internal fittings and to guard against the entry of silt and other deposits, a temporary stopper should be provided to block the intake pipe.

118 1 2 . 7

12.7 Slipways And Ramps


12.7.1

12.7.1

General
Grove & Little (1951)

A slipway is a structure, consisting of a rail track, cradle and haulage device, used in ship building and ship repair work for the movement of vessels to and from the sea. The cradle is used to support the vessel and runs along the rail track, usually of standard flatbottomed rails in two, three or four parallel lengths. Wire ropes are usually used to haul the vessel by means of a winch. Useful information on slipways is given by Grove & Little (1951).

12.7.2

12.7.2

Location and Basic Dimensions


1.5 1 : 1 0 1 : 2 5 1 : 1 5

Slipways should be located, where possible, at sites well protected from wave action. The slipway dimensions will depend on the size of the largest vessel to be slipped; in general the length of track above high water should exceed the vessel length, and the lower end of the track should extend to a depth adequate to allow the cradle to clear the vessel at lowest tide. The overall slipway width should be at least one and a half times the width of the largest vessel, and the gradient of the track within the range 1 in 10 to 1 in 25, with about 1 in 15 being normal.

12.7.3

12.7.3

Slipway Design

To a large extent, slipway design will depend on the method of construction; construction in the dry within a cofferdam may be more expensive in terms of initial cost than construction underwater, but will enable better quality of construction and tighter tolerances, resulting in a significant reduction in likely long term maintenance costs. With piled foundations, differential settlement will be controlled. With rubble mound foundations, it is essential that pre-loading is carried out to limit future differential settlement. Track support

119

beams should be connected by cross-ties to maintain track gauge. Rail track fixing details should allow for possible relevelling and realignment during the design life of the structure, and also possible replacement of the upper lengths due to corrosion; it is common for the track to be laid on a hardwood timber runner and fixed in position using special track spikes at the edges of the rail bottom flange. Setting tolerances for line and level will depend on the cradle design, but will normally be significantly tighter than for general marine works, and Particular Specification clauses should be drafted to reflect this; a tolerance of 10mm for line and level is considered typical, but is often difficult to achieve for underwater work. For the design of the rail track support beams, the main problem relates to the assessment of the load distribution as the vessel ceases to be waterborne and becomes carried on the loading cradle. At the start of slipping, with the cradle at the bottom of the slipway, the vessel is warped into position until bearing is obtained on the first section of the cradle. As slipping commences, by hauling up the cradle, gradually more and more weight is taken by the first section, and this load reaches the maximum just as the second section begins to take a share of the weight. Thereafter, all sections progressively take some load until the vessel is clear of the water and bearing uniformly over the whole cradle length. The exact value of the maximum load bearing on the first section, or 'sue' load, depends on the draft and outline of the vessel concerned, but as a guide can be taken to be about one third of the vessel weight. Since the sue load is only effective over a relatively short length, it is unnecessary to design the full slipway length for this load. The lowest length need only be designed to carry the weight of the cradle, and the upper length to carry the weight of the cradle plus vessel uniformly distributed. The intermediate length should be designed for the full sue load, or a proportion of the full sue load increasing from the lower end to the full sue load at the upper end, as appropriate. Care should be taken in estimating the cross distribution of load; with a cradle carried on two rails only, it is safe to regard the load as being equally divided between them, but where three or four rails are

10

120

involved, such an assumption is not recommended due to possible rail settlement causing the cradle to carry loads unevenly, and it is recommended that each rail should be designed for at least one half of the load.

12.7.4

12.7.4

Ramp Design
5 10 12 81 : 12

In comparison with a slipway, a ramp is a relatively simple structure. It consists essentially of a concrete slab sloping from about lowest tide level to above high tide level, for the movement of vehicles, usually from vessels to the shore. Design criteria should be agreed with the client. Design axle loads are typically 5 to 10 t with a maximum of about 12 t, with a normal ramp width of about 8 m and a slope of about 1 in 12. A simple rubble foundation, at least 3 m thick, is usually satisfactory for a ramp, as settlement problems are not usually significant. The section within the lower tidal range is usually constructed using precast concrete blocks, for ease of construction and additional stability from possible wave scour and vessel underside/ramp fouling. The upper section is usually a normal in situ concrete slab, typically 0.3 m thick, either reinforced for crack control or unreinforced with joints at 4 to 5 m centres. Care should be taken to ensure that the rubble foundation at the lower end and sides is trimmed, and checked by a diver, to ensure no projection of rubble above the slab line which might cause damage to a vessel approaching the ramp.

3 0.3 4 5

12.8

12.8 Navigation Aids


Aids to navigation are used to mark limits of structures such as piers, quay walls, breakwaters and dolphins, channel entrances, boundaries and turns, and hidden dangers such as shoals and rock outcrops, to act as a guide for vessels and to assist with their safe movement. The type, size, location and details of fittings and fixtures for navigation aids should be to the requirements of the Director of Marine.

121

Navigation aids covered by this Manual include lit and unlit beacons located offshore, on the foreshore or rock outcrops, and on land, and navigation lights on marine structures. Lights can be mains- or batterypowered as appropriate to the location, and as required by the Director of Marine. A beacon located offshore can either be a piled structure, similar to a dolphin in design, or a precast reinforced concrete gravity structure with enlarged base and rubble foundation, depending on the sea-bed conditions and water depth. Generally, single pile dolphins are not recommended because of their susceptibility to accidental damage. Beacons located on the foreshore or rock outcrops can usually be simple precast or cast in situ concrete structures doweled to underlying sound rock where possible. The above beacons will all be topped with steel light posts for final light connection for lit beacons, or simple steel/concrete marker posts for unlit beacons. Beacons located on land and navigation lights on structures will generally only be subjected to dead and wind loads, and simple mass concrete foundations for the light posts or marker posts will usually be adequate. Ladders, fenders and mooring eyes as appropriate should be provided for beacons located offshore. Beacons located on the foreshore, rock outcrops and land should be provided with landing facilities, either incorporated into the beacon structure or built separately. Fittings and fixtures such as ladders, handrails and mooring eyes should be stainless steel. Steel light posts and marker posts should preferably be galvanised after fabrication, and painted with a paint appropriate for the location, taking into account the ease of access for future maintenance (see Section 14.3.5).

14.3.5

12.9

12.9 Outfalls And Intakes Stormwater outfalls in seawalls and quaywalls should preferably be located with invert levels no higher than +0.3 mCD to reduce visual impact, staining, and possible problems with adjacent vessels. In the case of larger sized culverts constructed through new reclamations, this requirement may need to be
+ 0 . 3 m C D

122

waived in order to produce efficient hydraulic designs. In some recent engineering studies, reclamation levels have been raised, partly to permit higher invert levels to reduce backwater effects in the drainage systems. Outfalls should be located well clear of pumphouses, intakes and landing steps, and where possible should not be located immediately adjacent to suspended deck structures, because of possible future dredging access problems during desilting.

Outfalls through rubble seawalls usually consist of precast concrete units with wing walls on the outer face. Outfalls through concrete blockwork quaywalls are formed in special precast concrete blocks to suit the size of drainage pipe or culvert. For large box culverts, it is often necessary to form two units with a horizontal joint at about mid-wall height in order to reduce unit weights to a reasonable level. Wherever possible, lifting hooks for precast concrete outfall units should be formed in recesses which can be filled with suitable grout or concrete after unit setting; in this way, lifting hooks need not be removed and are available for future use during demolition or future modification. Seals between outfall units are not usually necessary but shear keys are often provided. Where outfalls are constructed in advance of drainage pipes or box culverts, the outfalls should be temporarily sealed by timber boards, brickwork, concrete or steel plates as appropriate for the opening size; the loads on the temporary seals from waves, water pressure and soil pressure should be assessed. Intakes are usually formed in concrete blockwork quaywalls to provide seawater for pumping stations, and are usually constructed concurrently with the quaywall. Size and location of the intake will be determined by the client. The invert level should be designed to ensure a continuous supply of water, unaffected by waves, tides, currents and water temperature variations. The usual method of construction is to use precast concrete units for the base slab and lower walls, and cast in situ concrete for the upper walls and roof slab. Joints between precast concrete units are usually required by the client to be sealed.

123

It is recommended that reinforced concrete for outfalls and intakes should have a characteristic strength of 45 MPa, and should be in accordance with Section 11.4.

45 MPa11.4

12.10

12.10

Miscellaneous

Measured mile markers or transit markers are marker posts erected in conspicuous locations on land, and used by the Marine Department for carrying out vessel speed trials. Two pairs of marker posts are used, each pair separated by a known fixed distance, and the lines joining the marker posts in each pair perpendicular to the known fixed distance. It is recommended that the ratio between the distance between each marker post in each pair, and the distance between the vessel trial line and the nearest marker post, should be not less than one third. Details of the size, height, and locations of marker posts are usually determined by the Director of Marine. Steel tubular marker posts are usually used with steel guy ropes as necessary, and mass concrete foundations.

Concrete mooring blocks are normally precast in sizes of , 1, 2, 3, 5, 10, 15, 25, 50 and 90 t. Combined lifting and anchor hooks are provided for the five smallest sizes, and separate lifting and anchor hooks for the larger sizes. The 90 t mooring block consists of a 50 t main block and a separate 40 t saddle block which can be placed on top. Concrete for mooring blocks is usually unreinforced for the smaller sizes, and nominally reinforced for the larger sizes, excluding lifting and anchor hooks; concrete with a characteristic strength of 30 MPa is considered appropriate, as the blocks are fully immersed during use.

1 / 2 1 2 3 51 01 52 55 0 9 0 90 50 4 0 30 MPa

124

125

13. CONSTRUCTION

13.

13.1

GENERAL

1 3 . 1

This Chapter covers current practice in the supervision and the keeping of site records for different types of work and construction, including dredging, breakwaters and seawall foundations, concrete blockwork walls, piers and dolphins, reclamations and underwater blasting. Comments are given on aspects which require particular attention.

13.2 DREDGING

13.2

13.2.1

General

13.2.1

For capital works dredging, it is important to distinguish between foundation dredging which is required for breakwaters, seawall foundations and submarine pipelines, and navigation dredging which is required for new fairways and approaches to piers and ferry terminals. Maintenance dredging is covered separately in Chapter 14, although many aspects covered in this Chapter related to navigation dredging also apply to maintenance dredging. For foundation dredging, which is usually carried out in trenches, the quality of the remaining material at foundation level is of primary importance; for navigation dredging, the finished dredging level is the most important consideration. For foundation dredging, the stability of trench side slopes is of only secondary importance as these are required to be stable only temporarily until the trench has been filled with foundation material. For navigation dredging, the stability of any side slopes is of major importance as these are required to be stable under long term conditions covering current and wave extremes.

1 4

13.2.2

Preparation and Execution of Works

13.2.2

Before any dredging works commence on site, it is

126

important to ensure that the contractor has erected sufficient land marks, and for the location of all land marks and the levels of all temporary tide gauges to be checked by land surveyor. The use of land marks and tide gauges will not be possible for some types of navigation dredging remote from the shore; in these cases, position checking will usually be by electronic distance measurement from shore stations, with the possible assistance of buoys for on-site cross checking, and level checking will be by interpolation from the nearest standard tide gauge stations. Before dredging commences, a detailed initial sounding survey of the seabed should be undertaken jointly with the contractor and agreed for record and measurement purposes. Throughout the dredging works, a daily record of dredging should be maintained. Dredged quantities for interim payment purposes are usually estimated by assessing the quantity in each barge and applying a reduction factor, which should be agreed with the contractor, to take account of bulking and possible overdredging. An initial reduction factor for marine deposits of 0.6 is suggested. Reduction factors should be regularly checked and adjusted as necessary using interim surveys, or final surveys, as sections of the works are completed. Upon completion of dredging, a detailed final sounding survey of the dredged area should be undertaken jointly with the contractor and agreed. Calculation of final quantities for payment purposes is usually based on the initial and final sounding surveys, as proposed by the Standard Method of Measurement for Civil Engineering Works (Hong Kong Government, 1992b).

0.6

Standard Method of Measurement for Civil Engineering Works (Hong Kong Government, 1992b)

13.2.3

Sampling of Dredged Materials

13.2.3

Samples of dredged materials should be taken at regular intervals, and at any change in stratum or material quality in general. As a guide, for foundation dredging in a trench, with uniform material, samples should be taken at about 15 m centres along the line of the trench and 2 m centres vertically. Each sample should have a mass of about 1 kg and should

15 2

127

preferably be taken from the centre of a grab or bucket load; for a trailer suction dredger, the sample should be taken from the pipe discharging into the hopper. Samples should initially be placed in clear plastic bags, all air expelled and the bags sealed with plastic tape. Labels should be attached with the following information included : (a) contract no.,

(a) (b) (c) (d) (e) (f) (g) ()

(b) location, (c) depth and level,

(d) time and date, (e) (f) dredging method, material description, and

(g) sand percentage (if available). Samples need normally only be kept until the completion of all dredging work and finalisation of dredging quantities. Where it is necessary to keep samples for more than a few months, for example where there may be contractual claims, the samples should be transferred to air-tight clear glass jars for storage. All samples should be kept in the site office, together with a summary in table form of the information included on the labels attached to the samples. When carrying out foundation dredging, for samples taken from depths close to the proposed dredge profile, the sand content should be determined on site. As a guide for checking the suitability of material at the bottom of a dredged trench for a seawall or quaywall foundation, common practice is to accept material with a sand content not less than 70% by weight, subject to checking with the design calculations in each case. The basis for this rule of thumb is that normal specifications for sand filling and underwater filling material in general allow the use of material with up to 30% by weight passing a BS No. 200 (63 micron) sieve. In general, it is not considered necessary to remove material from the

70 BS 200 63 3 0

128

bottom of a dredged trench if this material is to be replaced by filling material with a similar sand content. Of course, the above guide should not be used when the foundation material at the bottom of a dredged trench is a stiff clay. Such conditions should normally be identified at the design stage and alternative measures to determine the suitability of the founding material should be specified in the contract.

Reference should be made to the design calculations to ensure that the material properties at the bottom of a dredged foundation trench assumed in the design are in fact attained. When the clay content is low, it may be possible to reduce the depth of dredging if material with a sand content exceeding 70% is reached before the required depth shown on the original profile. If the required depth is reached and the material has a sand content of less than 70%, it may be necessary for the trench to be deepened or widened, or both. A design check should be carried out before such variations are implemented. It is recommended that important design assumptions are indicated on the drawings to assist in such checks.

7 0 7 0

13.2.4

Surveys for Dredging

13.2.4

Section 21 of the General Specification for Civil Engineering Works (GS) (Hong Kong Government, 1992a) requires the final survey for dredging to be carried out within 30 days after completion of dredging, and states that dredging should be carried out in such a manner and sequence that semi-fluid or disturbed seabed and foundation material will not accumulate in the dredged areas. For foundation dredging in a trench, it is important to check that there has been no significant deposition or accumulation of soft deposits in the bottom of the trench between completion of dredging and the start of filling with foundation material. This is particularly important when there has been a period of high waves during a storm. Such checking can be carried out by diver, grab sampling or repeat survey, or preferably a combination of these.

General Specification for Civil Engineering Works GS (Hong Kong Government, 1992a) 2 1 3 0

129

Section 21 of the GS states that surveys for dredging should be carried out by echo sounders of 200 to 220 kHz frequency or other methods specified by the Engineer. It is important to arrange for echo sounding results to be checked using chain sounding, particularly for foundation dredging in trenches and where the presence of mud in suspension is suspected. Mud in suspension can be a major problem in deep trenches where soft marine deposits have been recently dredged; in some cases, the rate of settlement of the mud is extremely slow, possibly due to continuous disturbance due to the passage of vessels, wave effects and currents. Differences of several metres between echo sounder and chain sounding results in deep trenches have been experienced, even after several days following the completion of dredging. For such cases it has been necessary to ignore the echo sounder results and to use chain sounding for the final trench survey.

G S 2 1 200220 kHz

Problems with echo soundings being reflected from mud suspensions can be investigated by taking samples by diver at different depths and by using a gravity corer. For navigation dredging, information on the density of bottom sediments in relation to 'nautical depth' is given in Section 2.4.

2.4

Even where there are no apparent problems with echo soundings being reflected from mud suspensions, there will often be difficulties in interpreting echo sounding traces, due to aspects such as wave movements and beam widths. Each trace should be discussed with the surveyor, and contractor if appropriate, for each case, before determining the most appropriate interpretation under the circumstances. For different situations, it may be appropriate to use the high, mean or low reading from the trace at a particular point to represent the bed level at that point, or to use the mean of all the high and low readings from the trace between two points to represent the bed level midway between the two points. For payment purposes, the use of mean readings is considered generally to be the most reasonable and appropriate interpretation. For acceptance purposes, particularly for navigation

130

dredging where isolated high spots may be critical, the use of the high readings from the trace will generally be more appropriate. For foundation dredging, when a section of trench has been completed and the profile checked by sounding survey and sampling, approval of the profile should be recorded and agreed with the contractor. No filling material should be placed in the section of trench under consideration until the relevant approval has been recorded.

13.2.5

13.2.5

Dumping of Dredged Material


It is not usually necessary to physically ensure that dredged materials are dumped at gazetted marine dumping grounds, as this is a legal requirement of the dumping licence issued by and policed by the Environmental Protection Department. However, periodic checks should be made that the contractor's barges are properly licensed and have appropriate dumping licences. The periods between the barges leaving full from the site and returning empty to the site should also be checked to ensure they are compatible with the time the trip to the appropriate dumping ground should take.

13.3

13.3 BREAKWATERS AND SEAWALL FOUNDATIONS


13.3.1

13.3.1

General

This Section covers sand and rock filling and the placing of rock armour for breakwaters and seawall foundations, although many aspects covered also apply to foundations for other marine structures and rubble mound construction in general.

13.3.2

13.3.2

Preparation and Execution of Filling


Prior to any filling or placing work, as for dredging, it is important to ensure that sufficient land marks and stagings have been erected and that the location of these and the levels of all temporary tide gauges have

131

been checked, if this has not already been done for prior dredging work. A daily record of filling and placing should be prepared. As for dredging, quantities for interim payment purposes can be estimated by assessing the quantity in each barge and applying a reduction factor, in this case to take account of waste, loss, consolidation and possible dumping outside the required profile. Initial reduction factors for sand filling and rock filling of 0.7 and 0.65 respectively are suggested and should be agreed with the contractor. Reduction factors should be checked regularly and adjusted as necessary using interim surveys, or final surveys as sections of the works are completed. Upon completion of filling or placing each material, final detailed surveys should be taken jointly with the contractor and agreed. These surveys form the basis for measurement and payment, in accordance with the Standard Method of Measurement for Civil Engineering Works (Hong Kong Government, 1992b).
0.70.65 Standard Method of Measurement for Civil Engineering Works (Hong Kong Government, 1992b)

Section 21 of the GS requires rock in armour layers and underlayers in rubble mound construction to be placed working from the bottom to the top of a section, in such a manner and sequence that the individual rock pieces interlock and do not segregate and the interstices are kept free of small rock fragments. These requirements are particularly important as they relate directly to design assumptions covering stability against wave attack and wave run-up. There should be no free pieces on the surface of a completed layer, and all pieces should be wedged and locked together so that they are not free to move without disturbing adjacent pieces in the same layer. For rock armour layers and underlayers above water level, final visual inspections from the top of the slope and by boat from the bottom of the slope should be carried out in addition to the normal profile check by survey. Below water level, a final visual inspection by diver is recommended where possible, depending on visibility, particularly for rock armour layers. If any significant holes or areas with infilled interstices are detected, whether above or below water level, it will be difficult for these to be satisfactorily rectified without almost complete reconstruction of the adjacent areas.

GS 2 1

132

Where filling or placing for a section of the work has been completed, and the profile checked by survey and inspection, approval of the profile should be recorded in the Site Instruction Book and countersigned by the contractor. No overlying work should be started for the section under consideration until the relevant approval has been recorded.

13.3.3

13.3.3

Surveys for Fill Materials


GS 2 1 200 220 kHz 1 3 . 2

Section 21 of the GS states that surveys for the deposition of fill material should be carried out by echo sounders of 200 to 220 kHz frequency or other methods specified by the engineer. As for dredging, echo sounding results should be checked using chain sounding, although mud in suspension is less likely to be a problem in this case. Difficulties in interpreting echo sounder traces are covered in Section 13.2. For fill material, the use of mean readings for payment purposes and low readings for acceptance purposes is suggested, but this should be discussed with the surveyor, and contractor if appropriate, for each particular case. Rock filling and rock armour above water level will need to be surveyed using a levelling staff. The method of survey should be agreed with the surveyor before work starts, to ensure that readings are taken at truly representative points but that any high and low spots are also identified.

13.3.4

13.3.4

Tolerances for Fill Materials


GS21 100

Tolerances for the deposition of sand fill, rock fill and rock armour are covered in Section 21 of the GS. It is usual practice, particularly for concrete blockwork walls, for the level of the top of the rock fill in foundation construction to be raised above the required design level to allow for subsequent settlement during the construction period. The amount of set-up should be proposed by the contractor, and checked and agreed by the engineer. The amount depends on

133

many variables, including the characteristics of the underlying foundation material, the thickness of the sand and rock filling, the mass of the works to be constructed on the foundation, and the expected construction period. As a guide, for normal concrete blockwork wall construction, a set-up of between about 100 and 300 mm would be expected.

300

13.4

13.4 CONCRETE BLOCKWORK WALLS


13.4.1

13.4.1

Levelling Stones and Blocks


Normally, no packing pieces are allowed at vertical or horizontal joints between precast concrete blocks; therefore the ease and accuracy with which a concrete blockwork wall can be constructed is greatly dependent on the accuracy in shape and size of the blocks being used, and the accuracy and consistency of the levelling stones on top of the rubble mound foundation. It is important for the levels of the rails or other profile marks to be checked by surveyor before the laying of the levelling stones starts, and for the levelling stones to be inspected by diver before any block setting. Daily records for the casting and setting of blocks should be kept. In addition, record drawings giving the date of setting of each block should be kept in the site office. After the setting of each layer of blocks has been completed, a diving inspection should be carried out to check such matters as the accuracy of setting, joint widths, infilling of channels between adjacent blocks and cleanliness of the top surface for receiving the next layer of blocks.

13.4.2

13.4.2

Bermstones
GS21

Section 21 of the GS requires bermstones to be placed as soon as practicable after the setting of toe blocks. Bermstones are required to prevent scour of foundation material from immediately in front of the toe blocks due to waves and currents, and early placement is particularly important when one or more of the following conditions apply :

134 (a)

(a)

the location is subject to strong currents,


(b)

(b) the location is exposed to wave attack, (c) block setting is to be carried out in the season when tropical storms may be frequent, or
(c) (d) -3 mCD

(d) the toe level of the wall is higher than -3 mCD. The placing of bermstones should be checked by diver. It is important that the bermstones should extend over the foundation width required, that the gaps between bermstones are kept to the minimum, and that no part of any bermstone extends above the level of the top of the adjacent toe block.

13.4.3

13.4.3

Facing Stones and Copings


The construction of in situ concrete copings and the pointing of facing stones should preferably be carried out as late as possible in the construction programme for each section of wall in order to minimise the effects of wall settlement. Subject to user requirements, it is common to allow the contractor to delay these two items of work until towards the end of the maintenance period. If it is necessary for any bollards to be used during the maintenance period, at least the sections of coping adjacent to each bollard foundation should be constructed together with the bollard before a certificate of completion for that section of the works is issued.

13.5

13.5 PIERS AND DOLPHINS


13.5.1

13.5.1

Preparation of Works

The location of all temporary setting-out marks and the levels of all temporary tide gauges should be checked by surveyor before any piling or other construction work starts on site. For a driven pile, the location and rake of the pile should be checked after pitching before driving starts, and for a bored pile, the location of the casing should be checked before any excavation is carried out.

135

Tidal work requires careful planning to ensure that all available working time at low tide is efficiently used. Charts should be drawn up from the tide tables, taking into account past records, to assist with planning work. The contractor should be reminded to apply for a Construction Noise Permit, if required, well in advance of the planned date of the work, if tidal work is required to be carried out outside normal working hours. When checking any contractor's temporary stagings and temporary works designs, care should be taken to take into account the effects of the following :

(a)

(a)

wind pressure,
(b) (c)

(b) wave loads including possible uplift,

(c)

construction materials,
(d)

(d) formwork and falsework,


(e)

(e) (f)

plant, equipment and workmen, and


(f)

dynamic (impact) loads.


13.5.2

13.5.2

Piling

Temporary stagings for pile driving plant and equipment should take particular account of pile weights during pitching and during the driving of raking piles. Section 8 of the GS requires all marine piles to be driven from fixed stagings unless approved otherwise by the engineer. There is the likelihood of damage to precast concrete piles driven from a barge, especially at exposed sites. Under certain circumstances, pile driving from a barge may be acceptable for relatively protected sites, particularly where steel piles are to be used. Large piling barges should be used, and the contractor should demonstrate that pile damage during driving due to barge movements is very unlikely.

GS 8

136 GS 8

The provision of temporary supports for driven piles during driving, and until incorporation into the superstructure, is covered in Section 8 of the GS. For marine piles it is important to ensure that bracing to pile heads, in two directions at right angles, is provided immediately after driving, to prevent the possibility of oscillation in the cantilever mode due to current and wave forces. In addition to ensuring that the pile driving or installation records are kept in accordance with Section 8 of the GS, a drawing showing the location and reference mark of each pile, the date of driving or installation, sea bed and pile toe levels, and final set for each pile if applicable, should be prepared and kept up to date on site. For precast concrete piles, a record giving the pile reference mark, size, length and date of casting should be kept on site, together with the dates of tensioning and strand cutting if applicable. It is normal practice for the contractor and engineer jointly to keep a record of all relevant data relating to piling in the contract. It is important to check that precast concrete piles are lifted and supported at the specified points and generally handled with care.

GS 8

13.5.3

13.5.3

Fendering

It is normal to delay final fender fixing until as late as possible in the construction period, to reduce the possibility of damage by construction plant to the minimum. It is important that any cast-in fixing bolts are adequately protected against corrosion and damage, prior to the fenders being fixed.

13.5.4

13.5.4

Works by Others

When facilities such as lights, cables, pipes, lifts and ramps are to be installed by other agencies during the works, particular attention should be paid to coordinating the different parties. The overall construction programme should take into account

137

advance notice or lead time required by the various agencies, and any changes to the programme should be transmitted to all parties without delay. Wherever possible, a set of drawings giving details of all work to be carried out by others should be kept on site for reference.

13.6

13.6 RECLAMATIONS
13.6.1

13.6.1

Sequence of Reclamation
100

Where reclamation is being carried out at the same time as seawall construction, regular seabed surveys should be carried out to ensure that no mud waves affect the seawall under construction. In general, no filling should be carried out within 100 m of an uncompleted seawall. For a rubble or concrete block seawall with rubble foundation, it is usual to form a bund, using selected granular material where possible, along the line of and immediately behind the seawall, after completion of block setting and placement of filter material. Such a bund assists with stabilising the seawall and its foundation, reducing the possibility of mud waves adversely affecting the seawall stability, and helps to induce initial settlement before continuation of the seawall.

13.6.2

13.6.2

Precautions to Be Taken During Reclamation


During reclamation, care should be taken to check for existing outfalls, intakes and drains along the lines of existing shorelines, particularly at low tide. Any outfalls, intakes or drains which do not appear on the reclamation construction drawings should be recorded immediately and investigated with the appropriate authority; temporary channels should be provided as necessary across the reclamation as a short term measure to avoid any delays. Care should also be taken to minimise dust generated by trucks whilst inside the reclamation site. This can be achieved by providing water lorries to regularly water the access roads on the site. Wheel washing facilities should also be provided at exits to

138

ensure that dumping trucks do not take out and deposit mud onto public roads.
13.7

13.7 UNDERWATER BLASTING Requirements for blasting trials, blasting in general, and submissions of particulars of the proposed blasting procedures are given in Section 6 of the GS. Where underwater blasting is envisaged, additional requirements should be included in the Particular Specification. The following aspects require special attention : (a) Underwater blasting can only be carried out by a diver who possesses a valid Mine Blasting Certificate issued by the Commissioner of Mines.
(a) GS 6

(b) All vessels within 150 m of the blast area should be cleared before the start of blasting operations. (c) Requirements concerning the possession and use of explosives are included in the Dangerous Goods Ordinance and subsidiary legislation.

( b ) 1 5 0

(c)

(d) Adequate prior notice should be given to the Director of Marine before any underwater blasting is carried out. (e) When explosives are being transported by sea and when underwater blasting operations are under way, signals should be displayed as prescribed by the Director of Marine.

(d)

(e)

13.8

13.8 MATERIAL INSPECTION AND TESTING


13.8.1

13.8.1

General
GS

This Section gives guidance on inspection and testing of materials and should be read in conjunction with the requirements given in the various sections of

139

the GS. Only materials which are directly related to, and generally peculiar to, marine works are covered. The inspection and testing of materials such as concrete, steel reinforcement and structural steelwork used in marine works should be no different from that for the same materials used in general civil engineering works, and are therefore not covered.

13.8.2

13.8.2

Marine Fill
G S 2 1

The sampling and testing of marine fill material is covered in Section 21 of the GS; particle size distributions for samples taken from the proposed source of supply are required before any source is approved, and regular checks on the sand content of samples should be made before deposition on site. Regular checks should also be made that the material is coming from an approved source. It should be noted that weathered granite from land sources can be used as an alternative to sand in the foundations of seawalls and breakwaters. Reference may be made to Lo (1980).

Lo (1980)

13.8.3

13.8.3

Rock Fill
G S 2 6 GS 2 1

No particular requirements for sampling and testing rock fill are included in Section 26 of the GS. Different types of rock fill should be checked regularly for compliance with the relevant clauses of Section 21 of the GS by visual inspection, preferably at stockpiles before loading onto barges for dumping at site. To assist with these visual inspections, specimens of pieces of rock complying with the upper and lower specified bounds for size and mass should be kept available for comparison. In extreme circumstances, particularly when there is major disagreement with a contractor, it may be necessary to carry out a full scale check on the material on site before compliance with the specification can be assured.

140 13.8.4

13.8.4

Rock Armour
G S 2 1 GS21 G S 21

Requirements for laboratory testing and field checking of rock armour are covered in Section 21 of the GS. In general, it is not considered necessary for laboratory testing to be carried out where rock armour consists of sound granite. Field checking for mass need normally only be carried out by visual comparison with the specimen samples of the upper and lower bounds for each rock armour type covered in Section 21 of the GS. The 'dropping test' described in Section 21 of the GS should be used as a field check for all rock types.

13.8.5

13.8.5

Timber for Fenders


G S 2 1 BS5756 (BSI, 1980) BS5756

The sampling and laboratory testing of timber for fenders are covered in Section 21 of the GS. Where the three pieces of timber selected for preparation of test specimens are not taken from a supply of timber already delivered to site, additional specimens from each piece should be taken at the time of preparation of the test specimens for future visual comparison with the timber actually delivered to site and purported to be from the same supply. All timber delivered to site should be checked for the marking required by for Tropical Hardwoods Graded for Structural Use in BS 5756 (BSI, 1980) to identify each piece as visually stress graded hardwood. In addition to the above checking, every piece should be inspected for major defects including knots, fissures, resin pockets, insect holes and fungal decay, before use on site. Any major defects should be checked against the requirements of BS 5756.

13.8.6

13.8.6

Rubber Fenders
G S 2 1

Requirements for testing and certification of rubber fenders are covered in Section 21 of the GS. On delivery to site and before incorporation in the works, each rubber fender should be inspected for defects such as surface inclusions, pores and cracks.

141 13.9

13.9 COMPLETION OF WORKS


13.9.1

13.9.1

General

This Section gives comments and guidance on certain aspects related to the completion of works, covering the issue of completion certificates and the preparation of as-constructed drawings. Only matters particularly related to marine works are considered.

13.9.2

13.9.2

Completion Certificates

For concrete blockwork seawall contracts, it is common for the construction of the coping and the pointing of the granite facing, and in some circumstances the installation of bollards, fenders and other fittings, to be carried out in the maintenance period; the works can therefore be generally considered to be substantially complete without these items. When considering the possibility of issuing a partial completion certificate for a section of seawall not immediately required for permanent occupation or use, the following should normally apply :

(a)

(a)

The seawall section should be substantially completed, i.e. complete except for coping, pointing and fittings as appropriate.
(b)

(b) The length of the section should be not less than one standard panel length, which is about 24 metres.
(c)

24

(c)

The land formed behind the section of seawall should be capable of being put to the use for which it is intended. As-constructed Drawings

13.9.3

13.9.3

As-constructed drawings should be prepared and signed by the engineer for the contract. In addition to checking the drawings before signature, the engineer should ensure that all relevant information related to the use and maintenance of the works is extracted from

142

the design calculations and incorporated, by way of notes or otherwise, into the general arrangement and other as-constructed drawings, as appropriate. Such information should include, but should not necessarily be limited to : (a) details of design loadings, in particular live loads and any vessel design characteristics,

(a) (b) (c) (d)

(b) safe working loads for bollards, cranes and other equipment, (c) corrosion allowances with design lives for any exposed steel members, and inspection or maintenance

(d) any special requirements.

Pile record drawings should be included. These should give full details, including pile types, toe levels, sizes, hammer type, final sets and any load testing results. Sounding survey results, before and after construction, should be included where appropriate. Foundation trench dredging drawings should be amended to suit, where the actual final dredged profile differs from that shown on the original construction drawings. As-constructed drawings for reclamations should give final levels, and the outlines of any special dumping areas such as rock dumps. Temporary drains should be included where these still remain at the date of preparation of the as-constructed drawings. It should be noted that as-constructed drawings are intended to provide a true and accurate record of the works as completed, for use by the client or user and the maintenance authority, and for possible later use when planning or undertaking modification, extension or demolition work. As such, the as-constructed drawings need not necessarily exactly correspond to the drawings used for payment purposes, particularly where different methods of construction from those shown on the tender drawings have been proposed by the contractor and accepted, with payment being based on the original tender details.

143

144

14. MAINTENANCE

14.

14.1 14.1.1

Maintenance Inspections Routine Inspections

14.1 14.1.1

Regular routine inspections of all piers and landings should be carried out. The frequency of inspection depends on the importance and vulnerability to damage of the pier or landing. In general, inspections of the structures above low water level should be carried out at least twice a year for facilities in the urban area, and at least once a year for facilities in the New Territories. For heavily used piers and landings in the urban area, it is necessary to carry out inspections of the fendering systems, particularly step-blocks, at least once a month. The frequency of underwater inspections depends on the availability of divers, weather conditions and access constraints. The aim should be to keep the period between underwater inspections to about 12 to 18 months for the urban area and 18 to 24 months for the New Territories.

12 18 1824

14.1.2

Special Inspections

14.1.2

In addition to regular routine inspections, special inspections are needed in the following circumstances : (a) after a severe tropical storm or typhoon (priority should be given initially to inspection of the fendering systems),

(a)

(b) after an accident such as a vessel collision, (c) when a complaint or notice about any damage is received, and

(b) (c)

(d) after completion of repairs.

(d)

14.1.3

Inspection Procedures

14.1.3

Wherever possible, all above-water inspections should be arranged to be carried out at low tide. The

145

above-water inspections should include all structural members, fendering systems, landings, steps, handrails, pier light posts and mooring facilities. Underwater inspections should include, where possible, all piles, pedestals and foundations, and particular attention should be paid to possible undermining of pedestal bases and toe blocks at shallow depths, i.e. higher than about -3 mCD. In many cases, it will not be possible or practical to carry out an underwater inspection of all piles and pedestals due to time and access constraints. In this situation, typical piles and pedestals at different parts of the structure should be inspected, and this should be noted on the inspection report. When carrying out underwater inspections, marine growth generally should not be removed from piles and pedestals, except locally when investigating possible defects or voids.

-3 mCD

Where any inspection identifies areas of deterioration or damage, the inspection report should be accompanied by sketches showing the location of the deterioration or damage and the extent of the repair required.

14.2 Dredging 14.2.1 General

14.2 14.2.1

Many aspects covered in Section 13.2, and in particular those concerning navigation dredging, are also relevant when considering maintenance dredging. Section 13.2 should be consulted for basic guidance on such matters as the checking of land marks and tide gauges, the keeping of daily records, reduction factors for interim surveys, material sampling and the interpretation of echo sounder results.

1 3 . 2 1 3 . 2

14.2.2

Sampling and Surveys

14.2.2

Normally, the sand content of samples need not normally be obtained for maintenance dredging. Because the thickness of material required to be removed by dredging is generally less than one to two

1 2

146

metres, problems are not usually experienced with mud in suspension affecting echo sounder results; however, checks using chain sounding should be made where echo sounding is used. In certain cases, vessel congestion, access problems or the shape of the area to be dredged will mean that chain sounding in place of echo sounding will need to be used in any event. For maintenance dredging in rivers and nullahs, current practice is to use chain sounding, with the lead modified with a steel disc because of the very soft deposits; echo sounding is sometimes used for check surveys at higher tides and over deeper sections, but not for payment purposes.

14.2.3

Maintenance Dredging Adjacent to Structures

14.2.3

Subject to the need to ensure no undermining of existing adjacent structures, maintenance dredging should preferably be carried out to a level deeper than the minimum depth required for navigation or other purposes, to allow for future siltation. Wherever possible, subject to the above limitations, there should be no need for future maintenance dredging for at least two years, and preferably three to five years, after completion. For maintenance dredging immediately adjacent to existing structures, particularly pumphouse intakes and slipways, and where it is necessary to fully remove deposits from the tops of toe blocks, bermstones, rock fill or rock armour structures, grab dredgers cannot be used. In these circumstances, it is usually necessary to carry out initial removal of the material to a more accessible location by air-lift equipment for later removal by grab dredger, or to an adjacent location where it is not necessary to remove the material due to partial dispersal or the presence of low spots.

14.2.4

Maintenance Dredging of Rivers and Nullahs

14.2.4

Maintenance dredging of rivers and nullahs may be hampered by the following difficulties :

147

(a)

restricted access for marine plant due to shallow water and existing bridges with limited headroom,

(a)

(b) limitations on dredging depths due to high channel design invert and bank levels, (c) land access only available at certain locations for removal of material by land plant,

(b) (c) (d) (e)

(d) the fluid nature of deposits, resulting in leakage from barges and trucks, and (e) existing gas pipe and other services within and across channels, requiring additional control on dredging location and level.

The type of equipment to be used to overcome the above difficulties should be left to the contractor to develop. Cutter suction dredgers as well as grab dredgers have been successfully used in the Shing Mun River, using specifically modified tugs and barges within the river channel itself. Maintenance dredging in fairways, e.g. the Northern Fairway, may require the use of trailer suction dredgers, where the Director of Marine stipulates that stationary dredging plant would be hazardous to shipping and therefore not permitted.

14.3 Piers and Dolphins 14.3.1 Piles

14.3 14.3.1

For precast reinforced and prestressed concrete piles, usually the only sections showing defects and requiring maintenance are in the tidal and splash zones; for repairs to these sections, reference may be made to the advice given for concrete pile caps under Section 14.3.2, as the work will be of a similar nature. For uncoated tubular steel piles, corrosion of the steel should only be ignored when the as-constructed drawings indicate that the piles have been infilled with reinforced concrete to well below bed level, and the design calculations show that the steel is sacrificial and

1 4 . 3 . 2

148

has not been taken into account in the design.

For coated or wrapped tubular steel piles, it should be assumed that the coating or wrapping is essential for the full protection of the piles, and no loss of steel thickness has been allowed for in the design, unless evidence to the contrary is available from the asconstructed drawings or the design calculations.

Where it is necessary for the corrosion rate for a steel pile to be checked, it is recommended that this should be carried out by a suitably qualified company using specialist equipment. For each pile, a sufficient number of measurements should be made at each level, to ensure confidence in the assessed maximum, minimum and average thicknesses. For uncoated steel piles, where the design calculations allow a certain yearly corrosion loss, it is recommended that the first corrosion check should be carried out between three and five years after completion of pile driving. Intervals between subsequent checks will depend on the losses measured in relation to the design assumptions, but it is suggested that the time between checks should not exceed five years. Where maintenance work to tubular steel piles is necessary, because of damage to the existing coating or wrapping, or steel corrosion rate greater than allowed for in the design calculations, it is suggested that discussions are held with specialist suppliers, and subcontractors if necessary, before the method and type of repair is finalised. In all cases, the most appropriate method and type of repair will depend on the site conditions, location, type and extent of the damage. The application of a coal tar reinstatement epoxy, which can be applied underwater by glove, may be economical for small areas, and a proprietary underwater wrapping system for larger areas. In all cases, surface preparation must be in accordance with the material supplier's recommendations, and there may be a need to use a specialist subcontractor.

Repairs to timber and steel H-section fender piles are not usually practical or economical, and it is

149

common to replace damaged piles with new ones. Where full replacement of steel piles is difficult or impossible due to time or headroom constraints, temporary repairs may be made for corroded or damaged sections above low water level by cutting off the section in question and welding or splicing on a replacement section at low tide.

For cathodic protection systems, it is recommended that all inspection, monitoring and maintenance work should be carried out by a suitably qualified company which specialises in such work, with all maintenance work being carried out in accordance with BS 7361 (BSI, 1991c). It should be noted that cathodic protection is usually considered to be fully effective up to about the mid-tide level, provided the system is well designed and maintained.

BS7361 (BSI, 1991c)

14.3.2

Decks

14.3.2

For reinforced concrete decks, most maintenance work requires the treatment of defects which have resulted in reinforcement corrosion. Such corrosion is usually apparent first as rust staining at cracks, followed in due course by concrete spalling, exposing corroded reinforcement. It is important to try to distinguish between different types of cracks and their causes, and to determine, where possible, the reasons for any reinforcement corrosion. Certain types of cracks, for example those caused by structural or temperature movements and shrinkage, may result in reinforcement corrosion, while other cracks may be caused by expansion of reinforcement as it corrodes. For the majority of reinforced concrete deck structures covered by this Chapter, the main problems with reinforcement corrosion occur about 15 to 25 years after construction, in the upper tidal and splash zones, and appear to be due to chloride movement from the surfaces of the pile caps, beam sides and soffits, and slab soffits to the concrete immediately surrounding the reinforcing bars. Corrosion problems are usually at first only apparent in areas with higher

15 25

150

concrete permeability or lower cover. Repairs should be carried out as soon as cracks with rust stains are detected, in order to reduce the possibility of significant steel loss from the reinforcing bars due to corrosion. Concrete spalling can indicate that significant steel loss has already taken place, in which case costly repairs will be necessary. For susceptible areas, it is preferable for chloride penetration to be monitored, and protective coatings applied where necessary. Cracks can be repaired temporarily by breaking out to expose the affected reinforcing bars, removing loose rust, and patching with epoxy mortar or concrete, after cleaning and priming as recommended by the manufacturer. Such repairs will usually last several years until further cracks develop, usually at the edges of the previous repair. More permanent repairs, which can usually last for at least five years and sometimes up to eight or ten years, can be carried out by : (a) breaking out all concrete adjacent to the affected bars, including behind the bars,

(a)

(b) removing all loose rust, (c) scabbling the adjacent surfaces,

(b) (c) (d) (e) 13 /

(d) fixing a galvanised wire mesh, and (e) guniting in layers using a 1 : 3 cement/sand mix over the full affected area.

Maintenance problems with reinforced concrete pile caps and beams in the lower tide and submerged zones do not usually relate to normal reinforcement corrosion, due possibly to the reduced supply of oxygen and protection by marine growth. Within these zones, problems are more likely to be due to poorly compacted concrete or the use of concrete placed by tremie, combined with congested reinforcement. The use of concrete placed by tremie in low level pile caps appears to be a major cause of defects, and often results in effective wash-out of the concrete constituents over periods of months or years,

151

particularly when combined with high currents, waves and turbulence from passing or berthing vessels. Concrete repairs within the lower tide and submerged zones are generally difficult, and an effective life of only about five to eight years should be expected. The use of epoxy mortar and gunite is usually not possible, and the usual method of repair is to provide an additional collar of concrete to cover the area of defective concrete or washout. Care must be taken to remove all marine growth and or loose defective concrete by hand tools, air jet or water jet as appropriate, with work being carried out at low tide or by diver. Any exposed reinforcement must be cleaned and supplemented as necessary. Adjacent concrete surfaces should be scabbled, and dowel bars provided to key the new concrete collar, which should be nominally reinforced, to the existing concrete. Concrete can either be placed by tremie or 'in the dry' using extended watertight shutters, depending on the location, space and access constraints. If neither of the above methods of placing is considered suitable, the use of 'underwater' concrete, placed 'in the wet', using a proprietary admixture to increase cohesion, may be attempted; in this case, the manufacturer's instructions must be followed, and site trials must be carried out.

For comments on repairs to steel members where these comprise part of a deck, reference can be made to Section 14.3.1 for below-deck work similar in nature to that required for steel H-section fender piles, and Section 14.3.5 for work generally above tide level.

H 1 4 . 3 . 1 14.3.5

14.3.3

Fendering Systems

14.3.3

Repairs to normal timber fendering systems relate generally to loose, worn or missing step blocks, capping pieces and fenders. Repair work is usually relatively straightforward. Spikes, bolts and other steel fixtures and fittings should be replaced as a matter of course during repair work, where corrosion has caused a significant loss in section. In general, timber fenders should be replaced before the reduction in thickness at

2 0 50

152

any section due to wear exceeds 20%. It should be noted that such a reduction in thickness is equivalent to a section modulus or strength loss of almost 50%. When arranging for repair work, the opportunity should be taken to carry out minor improvements to design details, for example fixing, stiffening and bracing arrangements. For comments on the maintenance of timber and H -section steel piles, see Section 14.3.1.

H 14.3.1

Routine maintenance for fendering systems consisting of rubber cell fenders with steel frontal frames will normally be limited to rotation and occasional replacement of the nylon and polyethylene wearing pads. Touching up or complete repainting of the steel frontal frames should be carried out when damage to, or significant deterioration of, the original coal tar epoxy paint is detected. Removal of the frontal frames will be required for most touch up and repainting work; immediate replacement of the frontal frame with a spare frame or frame moved from a less critical location will usually be necessary. For touch up and repainting work, surface preparation should be carried out strictly in accordance with the manufacturer's recommendations. A minimum dry film thickness of 350 m for repair work using coal tar epoxy paint is recommended.

350

Damage to rubber cell fenders, rubber arch fenders and cylindrical rubber fenders will normally be limited to misuse or accident conditions. To avoid delays in replacing such fenders due to delivery time, a sufficient stock of spare fenders of each type and size should be kept ready for immediate use. Sway and sag chains for cell fender frontal frames and support chains for cylindrical fenders, together with their shackles, are usually of galvanised steel. Corrosion of the chains and shackles can normally be accepted until an effective reduction in diameter of about 10% at any section has occurred; in this case the full length of chain with shackles should be replaced. Such a reduction in diameter corresponds to a tension strength loss of almost 20%.

1 0 10 20

153

U-anchors for the above chains are usually also of galvanised steel. Because the loads on such anchors are usually a combination of tension and bending, and replacement of the anchors, which are usually cast in, is particularly difficult, it is recommended that any corrosion is treated as soon as it is detected. Corrosion will normally first be apparent at the junction between the galvanised steel and the concrete support member, and treatment will be far easier and more effective if carried out before serious pitting and concrete spalling occurs. Treatment with coal tar epoxy paint to a minimum dry film thickness of 350 m is suggested, after surface preparation strictly in accordance with the manufacturer's recommendations.

U 350

14.3.4

Steps and Landings

14.3.4

Repairs to concrete steps and landings are relatively common due to heavy use and the adverse tidal location; these relate mainly to : (a) damage to edges, due to step block/capping strip impact, and impact from heavy goods being carried to or from a vessel,

(a) (b)

(b) regular renovation or replacement of the rough-cast finish due to wear, and, in the longer term, and (c) cracks and possible spalling of concrete due to reinforcement corrosion.

(c)

Wherever possible, repairs should be carried out in such a way as to avoid the need for closing more than a part of any steps or landing at any one time. To avoid the need for complete replacement of any step or landing, cracks due to reinforcement corrosion should be treated as soon as they are detected. A suggested method of treatment is given in 14.3.2 above.

14.3.2

14.3.5

Miscellaneous Items

14.3.5

Repairs and maintenance to miscellaneous items

154

such as bollards, mooring eyes, ladders, light posts and handrails are covered below. Regular maintenance to standard cast iron concrete-infilled bollards is not usually necessary, as corrosion of the cast iron is not normally significant. Maintenance is limited to occasional repainting with bitumen paint to improve appearance where corrosion of the cast iron is noticeable; this is normally limited to bollards on public piers and at heavily used landing steps within the urban area. For the older mooring eyes and ladders, which were often only of steel treated with bitumen paint, normal practice has been to carry out no routine maintenance, but to wait until material loss from corrosion has become significant and then to replace the items with new sections. It is recommended that, for the older items, this practice should be continued because of the difficulty in carrying out an effective paint treatment to such items after corrosion is well underway. As a guide, mooring eyes and ladders should be replaced when an effective reduction in diameter or thickness of the steel section of about 20% has occurred; replacement items should be of galvanised or stainless steel.

2 0

For the newer and replacement mooring eyes and ladders, which will be of galvanised or stainless steel, it is recommended that regular maintenance covering treatment of any detected corrosion is carried out as for fender chain U-anchors under Section 14.3.3 above. Such treatment is particularly important at the junction between the steel section and the concrete. Regular maintenance will significantly extend the life of galvanised and stainless steel mooring eyes and ladders, and will avoid unsightly rust stains on the adjacent concrete faces. Heavily corroded mooring eyes can cause additional wear on mooring lines and heavily corroded ladders can be uncomfortable and dangerous for users. Light posts and light support steelwork should be regularly treated and repainted to ensure that no

14.3.3U

155

corrosion is allowed to develop, and that any colour code required by the Director of Marine is maintained. The steel structures are usually located above tidal zones and surface preparation followed by repainting is generally not difficult. The importance of adequate surface preparation and the need to carry out all work in favourable weather conditions to avoid contamination from moisture in the air and sea spray cannot be overstressed. Steel handrails should be regularly maintained to ensure maximum life and to avoid an unsightly appearance, particularly from rust staining. Unpainted galvanised steel sections should be treated with an appropriate paint system whenever corrosion is detected. Painted galvanised steel sections should be repainted whenever breakdown of the existing paint is detected and the underlying galvanised finish is visible. Surface preparation in accordance with the paint manufacturer's recommendations is important, particularly when painting or repainting galvanised surfaces. For handrails at lower landings and steps which are continuously within the lower tidal and splash zones, satisfactory surface preparation for painting or repainting will often not be possible. In this situation, the only practical solution is to replace any heavily corroded sections with new stainless steel or galvanised members, prepainted if possible.

14.4 Blockwork Seawalls 14.4.1 General

14.4 14.4.1

Regular routine maintenance to concrete blockwork seawalls of standard design is usually not necessary. However, repairs to certain parts of the seawall structure, including rubble foundations, outfall blocks, granite facing and coping, can be required under particular circumstances, and these are dealt with below. For comments on repairs to fendering systems, steps and landings, bollards and miscellaneous steelwork items, such as mooring eyes, ladders and handrails, see Sections 14.3.3, 14.3.4 and 14.3.5.

14.3.314.3.414.3.5

156

14.4.2

Repairs to Rubble Mounds

14.4.2

Damage to rubble mound foundations due to currents and wave attack can result in undermining of bermstones and toe blocks, which in turn can result in loss of fill material from behind the seawall. The first warning of such damage will usually be the development of cracks in the paving behind the seawall caused by voids forming under the paving. This type of damage usually only occurs in locations exposed to relatively severe wave attack and where the seawall toe block level is relatively high, generally above -2.0 mCD. Undermining is particularly likely to occur with sections of older seawalls which have been founded on rock outcrops where the normal standard bermstone/toe block detail has not been used. A diving inspection will confirm the presence of voids at the interface of the concrete blocks and rubble mound foundation when undermining is suspected; there will also usually be evidence of fill material in the region in front of the wall. It is important that repair work to any undermining is carried out as soon as it is detected to avoid additional loss of fill material, possible complete collapse of the paving behind the seawall, and possible settlement, rotation or even collapse of the seawall itself. The method of repair will depend on the circumstances in each case. One common method which is used, subject to navigation requirements immediately in front of the seawall, is to use concrete placed by tremie to fill voids between the concrete blocks and rubble, after constructing a concrete bagwork wall about one metre high doweled to the rubble mound, one metre or so in front of the concrete blocks, to retain the concrete. The existing rubble mound in front of the concrete bagwork wall should be supplemented as appropriate, and additional filling placed as necessary behind the seawall with the paving repaired to suit, after completion of the concreting.

-2.0 mCD

14.4.3

Repairs to Concrete Blocks

14.4.3

In general, the only concrete blocks in seawall

157

construction which contain reinforcement are outfall blocks. For older seawalls, problems can be found with cracking and later concrete spalling due to reinforcement corrosion within the tidal and splash zones. Treatment should be carried out as described in Section 14.3.2, preferably as soon as cracks with rust staining are detected.

14.3.2

14.4.4

Repairs to Granite Facing

14.4.4

Problems with granite facing relate to breakdown or loss of the mortar between the facing stones, apparently due to the chemical action of seawater or wave and current attack. Any sections where the original pointing has been damaged or lost should be repointed immediately in order to prevent further loss of mortar and eventual loss of individual facing stones. In older walls, where individual facing stones have already been lost, repairs can usually be carried out using in situ concrete doweled to the existing backing concrete, where appearance is not considered of major importance. Where it is considered essential that a uniform granite facing finish should be maintained, it is not normally possible to satisfactorily replace individual facing stones without effectively reconstructing the adjacent section of granite facing, including the backing concrete.

14.4.5

Repairs to Concrete Coping

14.4.5

Repairs to mass concrete copings relate mainly to damage from rope scrapes and vessels berthing, and to horizontal displacement, usually soon after construction of the coping and before final development and paving of the land behind the seawall has been carried out. Minor damage from rope scrapes and vessels need not normally be repaired unless the length of seawall is particularly heavily used by the public and appearance is considered important. More major damage can sometimes be effectively repaired using epoxy mortar, but it is usually necessary to break out and recast the full damaged length to produce a satisfactory result. Horizontal displacement of sections of coping should be treated on an individual basis,

158

depending on the circumstances. For newly completed seawalls, for displacement away from the sea, it is often not necessary to carry out immediate repairs, but to arrange for future repairs, including breaking out and recasting as necessary, to be carried out at the time of development and paving of the adjacent area.

14.5 Rubble Seawalls and Breakwaters Damage to rubble seawalls and breakwaters will normally only occur from wave attack during severe storms and typhoons. Minor damage to random rubble armoured structures, involving displacement, slippage or general movement of armour stones, can usually be repaired by placing additional stones of similar mass to infill any low regions and to build up the overall profile back to the existing design profile. Minor damage to older smooth faced rubble armoured structures, which normally consist of a single close fitting armour layer on rubble or quarry spalls, can most economically be repaired by infilling any holes with carefully selected and fitted armour stones, after levelling up the underlayer.

14.5

Repairs to major damage to rubble structures should not be carried out until a full design check has been carried out to determine whether the existing design, in particular the armour size, is in accordance with recommended practice. When carrying out repairs, the opportunity should be taken to modify or improve the design as appropriate. When major damage has occurred to older smooth faced rubble armour structures, it may be quicker and more economical to overlay the existing damaged slope with random rubble armour, subject to the existing slope not being too smooth and regular, rather than to completely reconstruct the existing armour layer, which will often involve difficult excavation and sorting work and will result in the underlayers being further exposed during the reconstruction period. For comments on repairs to minor items, such as steps, bollards and light posts, see Sections 14.3.4 and 14.3.5.

14.3.414.3.5

159

14.6 Ramps and Slipways Repairs to ramps are rarely required. In exceptional circumstances, after a severe storm or typhoon, retrimming of the adjacent rubble foundation to avoid fouling by landing craft, or the placing of additional rubble to avoid undermining the ramp slabs and toe blocks, may be required. Major repairs to slipways, related to rail misalignment, material deterioration and desilting, are regularly required. Repair work is generally difficult, slow and expensive, being mostly underwater with poor visibility. Close co-ordination with the slipway user is necessary as the slipway will need to be temporarily taken out of service while repair work is being carried out. Misalignment problems generally relate to differences in level, and occur for slipways with more than two parallel rails when the centre rail or rails settle more than the outer rails, causing the cradle to no longer run freely. The usual solution is to provide additional packs, of steel or hardwood, for the rails or to replace the existing timber runners with thicker sections.

14.6

During any realignment work, the opportunity should be taken to replace any timber runners, steel rails, fixtures and fittings which have badly deteriorated. Corrosion of the webs of the steel rails within the tidal zone will usually be particularly severe. As a guide, steel rails should be replaced when an average effective reduction in web thickness of 10% over any length can be measured. For checking horizontal alignment during realignment work underwater, a template based on the measured cradle roller dimensions, taking into account tolerances, should be used. When rail realignment work has been completed, any cradle deformation or defects, including sticking rollers, should be reported to the slipway user, for action by the appropriate maintenance authority. For comments on desilting for slipways, see Section 14.2. During the course of any slipway repair work, the opportunity should be taken to carry out any necessary desilting work, even if such work has not

1 0

14.2

160

been requested specifically by the slipway user, to reduce to the minimum the number of times the slipway needs to be taken out of service.

14.7 Pumphouses Repairs to pumphouses usually relate to particular operational problems and are requested specifically by the user, who should specify all technical requirements. The refixing and replacement of screen guides and the cleaning of the insides of screen guides and intakes are relatively common items of work, and can usually be carried out by divers without undue difficulty. For comments on desilting in front of pumphouse intakes, see Section 14.2.

14.7

14.2

14.8 Navigation Aids Maintenance and repair work to navigation aids, which generally consist of beacons and markers, is mainly limited to repainting and related treatment of miscellaneous steelwork items, such as light posts, ladders, mooring eyes and handrails. Reference may be made to Section 14.3.5 for comments on the general maintenance and repair of such items.

14.8

14.3.5

14.9 Estimates for Future Maintenance 14.9.1 General

14.9 14.9.1

This Section gives comments and guidance on the preparation of cost estimates for the future maintenance of various types of works, including dredging, piers, seawalls and breakwaters. It should be noted that all projects are individual, and the information given below should be used as a general guide only. When preparing estimates for a particular project, all special circumstances should be taken into account and, if possible, proposed maintenance estimates should be cross-checked against actual maintenance costs for similar works.

161

14.9.2

Estimates for Dredging

14.9.2

Estimates of future maintenance costs are usually difficult to make with any confidence, being dependent on many factors often for which little or no information is available. Estimates may be required for future maintenance costs for capital dredging works carried out as part of a project, or maintenance dredging costs for a project not necessarily involving capital dredging, e.g. the construction of a typhoon shelter. Factors which may influence estimates include :

(a)

general siltation and its effects on the works,

(a) (b)

(b) effects of currents and waves on the seabed within the area of the works and the possible changes in these caused by the construction of the works, (c) deposits collecting at existing and planned future stormwater drain, culvert and nullah outfalls,

(c)

(d) effects of construction of other works within the general area, and (e) possible illegal dumping of construction and waste materials.

(d) (e)

Past records and studies indicate siltation rates within the general harbour area, including fairways but remote from outfalls, in the order of 50 to 100 mm per year. However, rates measured around some fairly recent works are in the order of 200 to 300 mm per year, increasing markedly to 1000 mm per year or more in particular locations, e.g. near outfalls.

50100 200300 1 0 0 0

14.9.3

Estimates for Maintenance of Piers

14.9.3

Records for existing structures indicate that annual recurrent repair and maintenance costs for a public or ferry pier is in the order of 0.5% of the capital construction or replacement cost. This figure is based on a typical reinforced concrete suspended deck structure with a timber fendering system and

0.5

162

prestressed concrete or other similar 'maintenance-free' piles, in the harbour or reasonably protected location. An allowance has been included for the regular maintenance of miscellaneous minor items such as handrails, light posts and notice boards.

14.9.4

Estimates for Maintenance of Seawalls

14.9.4

The annual recurrent repair and maintenance costs for a length of vertical seawall has been estimated to be in the order of 0.1 to 0.3% of the capital construction or replacement cost. These figures are based on typical concrete blockwork structures, with rubble mound foundations and granite facing above about mid-tide level. The lower end of the range is applicable for walls with foundations below about -20 mCD dredged level, toe block levels lower than about -4 mCD and no or relatively few landing steps, bollards, notice boards and miscellaneous fixtures and fittings. The upper end of the range is applicable for walls with foundations higher than about -10 mCD dredged level, toe block levels higher than about -2 mCD, and a relatively large number of landing steps, bollards and miscellaneous items. For sloping rubble seawalls, annual repair and maintenance costs are in the order of 0.2 to 0.3% of the capital construction or replacement cost for relatively protected locations, and 0.4 to 0.6% for relatively exposed locations. The lower ends of the ranges are applicable for seawalls with foundations above about -10 mCD dredged level.

0 . 1 0 . 3 -20 mCD -4 mCD -10 mCD -2mCD

0 . 2 0 . 3 0.40.6 -10 mCD

14.9.5

Estimates for Maintenance of Breakwaters

14.9.5

Annual recurrent repair and maintenance costs for typical random placed rubble armoured breakwaters are in the order of 0.4 to 0.7% of the capital construction or replacement cost for relatively protected locations, and 0.8 to 1.1% for relatively exposed locations. As for sloping rubble seawalls in Section 14.9.4 above, the lower ends of the ranges are applicable for structures with relatively deep

0 . 4 0 . 7 0 . 8 1 . 1 14.9.4

163

foundations and the upper ranges for structures with relatively shallow foundations.

164

REFERENCES Apps, R.F. & Chen, T.Y. (1973). Sea Waves at Waglan Island. Royal Observatory, Hong Kong, Technical Note No. 36, 56 p. BDD (1983). Code of Practice on Wind Effects, Hong Kong. Building Development Department, Hong Kong, 14 p. Binnie & Partners (1987). Hydraulic and Water Quality Studies in Victoria Harbour. Port and Ship Studies - Initial Report . Territory Development Department, Hong Kong, 93 p. (Unpublished). Binnie & Partners (1988). Hydraulic and Water Quality Studies in Victoria Harbour. Part II Calibration and Validation Reports, Volume 10 - Wave Modelling Studies. Territory Development Department, Hong Kong, 56 p. Binnie & Partners (1989). Hydraulic and Water Quality Studies in Victoria Harbour. Mathematical Model Report, Part 1 - Overview. Territory Development Department, Hong Kong, 76 p. (Unpublished). Bruun, P. (1981). Port Engineering (Third Edition). Gulf Publishing Company, 787 p. BSI (several parts, 1957 to 1995). Physical Testing of Rubber (BS903). British Standards Institution, London. BSI (1969). Specification for the Use of Structural Steel in Building (BS 449:1969). Part 2 Metric Units. British Standards Institution, London, 116 p. BSI (1972). Code of Basic Data for the Design of Buildings (CP 3:1972). Chapter V - Loading, Part 2 - Wind loads. British Standards Institution, London, 49 p. BSI (1977). Code of Practice for Protective Coating of Iron and Steel Structures against Corrosion (BS 5493:1977). British Standards Institution, London, 114 p. BSI (1980). Specification for Tropical Hardwoods Graded for Structural Use (BS 5756:1980). British Standards Institution, London, 12 p. BSI (1983). Steel Plate, Sheet and Strip (BS 1449:1983). Part 2 - Specification for Stainless and Heat-Resisting Steel Plate, Sheet and Strip. British Standards Institution, London, 16 p. BSI (1984a). Code of Practice for Maritime Structures (BS 6349:1984). Part 1 - General Criteria. British Standards Institution, London, 179 p. BSI (1984b). Loading for Buildings (BS 6399:1984). Part 1 - Code of Practice for Dead and Imposed Loads. British Standards Institution, London, 21 p.

165

BSI (1985a). Structural Use of Concrete (BS 8110:1985). Part 1 - Code of Practice for Design and Construction. British Standards Institution, London, 126 p. BSI (1985b). Structural Use of Concrete (BS 8110:1985). Part 2 - Code of Practice for Special Circumstances. British Standards Institution, London, 50 p. BSI (1986). Code of Practice for Foundations (BS 8004:1986). British Standards Institution, London, 157 p. BSI (1987). Code of Practice for Design of Concrete Structure for Retaining Aqueous Liquids (BS 8007:1987). British Standards Institution, London, 31 p. BSI (1988). Code of Practice for Maritime Structure (BS 6349:1988). Part 2 - Design of Quay Walls, Jetties and Dolphins. British Standards Institution, London, 110 p. BSI (1991a). Specification for Tolerances on Dimensions, Shape and Mass for Hot Rolled Steel Plates 3mm Thick or Above. (BS EN 10029:1991). Part 1 - Code of Practice for Land and Marine Applications. British Standards Institution, London, 16 p. BSI (1991b). Structural Use of Timber (BS 5268:1991). Part 2 - Code of Practice for Permissible Stress Design, Materials and Workmanship. British Standards Institution, London, 130 p. BSI (1991c). Code of Practice for Cathodic Protection (BS 7361:1991). Part 1 - Code of Practice for Land and Marine Applications. British Standards Institution, London, 115 p. BSI (1993a). Hot-rolled Products in Weldable Fine Grain Structural Steels (BS EN 10113:1993). Part 1 - General Delivery Conditions. British Standards Institution, London, 24 p. BSI (1993b). Hot-rolled Products in Weldable Fine Grain Structural Steels (BS EN 10113:1993). Part 2 - Delivery Conditions for Normalized/Normalized Rolled Steels. British Standards Institution, London, 16 p. BSI (1993c). Hot-rolled Products in Weldable Fine Grain Structural Steels (BS EN 10113:1993). Part 1 - Delivery Conditions for Thermomechanical Rolled Steels. British Standards Institution, London, 16 p. BSI (1993d). Structural Steels with Improved Atmospheric Corrosion Resistance. Technical Delivery Conditions. (BS EN 10155:1993). British Standards Institution, London, 30 p. BSI (1994a). Code of Practice for Maritime Structures (BS 6349:1994). Part 4 - Design of Fendering and Mooring Systems. British Standards Institution, London, 51 p.

166

BSI (1994b). Hot Finished Structural Hollow Sections of Non-alloy and Fine Grain Structural Steels (BS EN 10210:1994). Part 1 - Technical Delivery Requirements. British Standards Institution, London, 28 p. BSI (1994c). Code of Practice for Earth Retaining Structures (BS 8002:1994). British Standards Institution, London, 113 p. BSI (1994d). Specification for Weldable Structural Steels. Hot Finished Structural Hollow Sections in Weather Resistant Steels (BS 7668:1994). British Standards Institution, London, 16 p. BSI (1995a). Stainless Steel (BS EN 10088:1995). Part 1 - List of Stainless Steel. British Standards Institution, London, 24p. BSI (1995b). Stainless Steel (BS EN 10088:1995). Part 2 - Technical Delivery Conditions for Sheet/Plate and Strip for General Purposes. British Standards Institution, London, 48p. BSI (1995c). Stainless Steel (BS EN 10088:1995). Part 3- Technical Delivery Conditions for Semi-finished Products, Bars, Rods, and Sections for General Purposes. British Standards Institution, London, 44p. BSI (1996a). Plates and Wide Flats Made of High Yield Strength Structural Steels in the Quenched and Tempered or Precipitation Hardened Conditions (BS EN 10137:1996). Part 1 - General Delivery Conditions. British Standards Institution, London, 20 p. BSI (1996b). Plates and Wide Flats Made of High Yield Strength Structural Steels in the Quenched and Tempered or Precipitation Hardened Conditions (BS EN 10137:1996). Part 2 - Delivery Conditions for Quenched and Tempered Steels. British Standards Institution, London, 20 p. BSI (1996c). Plates and Wide Flats Made of High Yield Strength Structural Steels in the Quenched and Tempered or Precipitation Hardened Conditions (BS EN 10137:1996). Part 3 - Delivery Conditions for Precipitation Hardened Steels. British Standards Institution, London, 18 p. BSI (1996d). Specification for Wrought Steels for Mechanical and Allied Engineering Purposes (BS 970:1996). Part 1 - General Inspection and Testing Procedures and Specific Requirements for Carbon, Carbon Manganese, Alloy and Stainless Steels. British Standards Institution, London, 50 p. CED (1991). Civil Engineering Department Standard Drawings. Civil Engineering Department, Hong Kong. CERC (1984). Shore Protection Manual, Volume I & II (4th Edition). Coastal Engineering Research Centre, Vicksburg, 1222 p.

167

Chan, Y.K. (1983). Statistics of Extreme Sea-Levels in Hong Kong. Royal Observatory, Hong Kong, Technical Note (Local) No. 35, 24 p. Chen, T.Y. (1975). Comparison of Surface Winds in Hong Kong. Royal Observatory, Hong Kong, Technical Note No. 41, 43 p. Chen, T.Y. (1979a). Comparison of Waves recorded at Waglan Island and at Kwun Mun. Royal Observatory, Hong Kong, Technical Note No. 48, 51 p. Chen, T.Y. (1979b). Spectral Analysis of Sea Waves at Waglan Island. Royal Observatory, Hong Kong, Technical Note No. 50, 39 p. Cheng, T.S. (1986). Tropical Cyclone Wave Statistics at Waglan Island. Royal Observatory, Hong Kong, Technical Note (Local) No. 37, 74 p. Chin, P.C. & Leong, H.C. (1978). Estimation of Wind Speeds near Sea-Level during Tropical Cyclone Conditions in Hong Kong. Royal Observatory, Hong Kong, Technical Note No. 45, 31 p. Cuming, M.J. (1967). Wave Heights in the Southeast Approaches to Hong Kong Harbour. Royal Observatory, Hong Kong, Technical Note No. 25, 25 p. Department of Transport (1982a). Port Approach Design - A Survey of Ship Behaviour Studies. Part 5 - The Determination of Channel Width Requirements allowing for Interactive Forces. Department of Transport, United Kingdom, 19 p. Department of Transport (1982b). Port Approach Design - A Survey of Ship Behaviour Studies. Part 6 - The Determination of Channel Width Requirements allowing for the Effects of Wind. Department of Transport, United Kingdom, 19 p. Department of Transport (1982c). Port Approach Design - A Survey of Ship Behaviour Studies. Part 7 - A Compendium Guide to the Design of Navigation Channels. Department of Transport, United Kingdom, 72 p. French, J.A. (1979). Wave Uplift Pressures on Horizontal Platforms. Proceedings of the Symposium on Civil Engineering in the Oceans IV , ASCE, pp 187-201. Fukuda N., Uno T. & Irie I (1974). Field Observations of Wave Overtopping of Wave Absorbing Revetment. Coastal Engineering in Japan, Vol. 17, Pg. 117-128. GCO (1984). Geotechnical Manual for Slopes, Second Edition. Geotechnical Control Office, Hong Kong, 295 p. GCO (1987). Guide to Site Investigation (Geoguide 2). Geotechnical Control Office, Hong Kong, 365 p.

168

GCO (1988). Guide to Rock and Soil Descriptions (Geoguide 3). Geotechnical Control Office, Hong Kong, 191 p. GCO (1990). Review of Design Methods for Excavations (GCO Publication No. 1/90). Geotechnical Control Office, Hong Kong, 192 p. GCO (1991). Review of Earthquake Data for the Hong Kong Region (GCO Publication No. 1/91). Geotechnical Control Office, Hong Kong, 115 p. GEO (1993). Guide to Retaining Wall Design (Geoguide 1), Second Edition. Geotechnical Engineering Office, Hong Kong, 267 p. Goda, Y. (1971). Expected Rate of Irregular Wave Overtopping of Sea Walls. Coastal Engineering in Japan, vol. 14, pp 43-51. Goda, Y. (1974). New Wave Pressure Formulae for Composite Breakwaters. Proceedings of the Fourteenth Coastal Engineering Conference, Copenhagen, vol. III, pp 1702-1720. Grove, G.C. & Little, D.H. (1951). The Design and Construction of Some Slipways up to 1,200 tons. The Institution of Civil Engineers Maritime Paper No. 17, 35 p. Hong Kong Government (1992a). General Specification for Civil Engineering Works. Hong Kong Government, 3 volumes. Hong Kong Government (1992b). Standard Method of Measurement for Civil Engineering Works. Hong Kong Government, 207 p. Hogben, N., Dacunha, N. & Olliver, G. (1986). Global Wave Statistics. British Maritime Technology, 661 p. Hydraulics Research Station (1971a). H.K. Container Terminal - Model Study of Uplift Forces due to Wave Action Beneath the Deck of a Piled Wharf Structure (Ex 536). Hydraulics Research Station, Wallingford, 17 p. Hydraulics Research Station (1971b). H.K. Container Terminal - Study of Uplift Pressures and Retaining Rock Stability in a Model Representation of the Junction with Berth No. 1 (Ex 559). Hydraulics Research Station, Wallingford, 14 p. Hydraulics Research Station (1980). Design of Sea Walls Allowing for Wave Overtopping (Ex 924). Hydraulics Research Station, Wallingford, 39 p. Lam, C.Y. (1979). Wind, Visibility, Sea and Swell over Coastal Waters of Eastern Guangdong 1961 - 1970. Royal Observatory, Hong Kong, Climatological Note No. 4, 37 p. Lam, C.Y. (1980). Wind, Visibility, Sea and Swell over Coastal Waters of Western Guangdong 1961 - 1970. Royal Observatory, Hong Kong, Climatological Note No. 5, 37 p.

169

Lo, Y.C. (1980). Behaviour of Decomposed Granite used in Marine Works. MSc(Eng.) thesis, University of Hong Kong, 42 p. (Unpublished). National Ports Council (1981). Port Approach Design - A Survey of Ship Behaviour Studies. Part 3 - The Prediction of Squat for Vessels in Shallow Waters. National Ports Council, United Kingdom, 19 p. Poon, W.C. (1982). Tropical Cyclones causing Persistent Gales at the Royal Observatory 1884 - 1957 and at Waglan Island 1953 - 1980. Royal Observatory, Hong Kong, Technical Note No. 66, 33 p. Premchitt, J. & To, P. (1991). A review of prefabricated band drains. Reclamation - Important Current Issues. Proceedings of the seminar organised by the Geotechnical Division of the Hong Kong Institution of Engineers on 14 May 1991, edited by P. Blacker, pp 75-93. Quinn, A.D. (1972). Design and Construction of Ports and Marine Structures. McGraw-Hill Inc., 611 p. RO (1990). Surface Observations in Hong Kong 1990. Royal Observatory, Hong Kong. RO (1991). Tropical Cyclones in 1990. Royal Observatory, Hong Kong. RO (annual). Tide Tables for Hong Kong. Royal Observatory, Hong Kong. Taunton (1975). Admiralty Tidal Stream Atlas, Hong Kong. Hydrographic Department, Taunton, 16 p. Tomlinson, M.J. (1987). Pile Design and Construction Practice. Viewpoint Publications, 378 p Transport Department (1986). Transport Planning & Design Manual, Volume 9 - Public Transport, Chapter 7 - Ferries. Transport Department, Hong Kong, 15 p. Van der Meer, J.W., Pilarczyk K.W. (1987), Stability of Breakwater Armour Layers Deterministic and Pro babilistic Design. Delft Hydraulics Communication No. 378, 34 p.

170

Table

171

172

LIST OF TABLES
Table No 1 2 3 4 5 Basic Characteristics of Vessels Local Craft Registry (September 1996) Basic Characteristics of Vessels H.K.F. Ferry Fleet (September 1996) Basic Characteristics of Vessels The Star Ferry Fleet (September 1996) Basic Characteristics of Vessels Discovery Bay Ferry Fleet (September 1996) Five-day Normals of the Meteorological Elements for Hong Kong 1961 - 1990 Mean Hourly Wind Speeds - The Observatory 1885 - 1939, 1947 - 1994 Mean Hourly Wind Speeds - Kai Tak Airport (South-east) 1968 - 1994 Mean Hourly Wind Speeds - Waglan Island 1975 - 1994 Mean Hourly Wind Speeds - Cheung Chau 1970 - 1991 Mean Wind Speeds - Waglan Island (North-east) 1975 - 1994 Mean Wind Speeds - Waglan Island (East) 1975 - 1994 Mean Wind Speeds - Waglan Island (South-east) 1975 - 1994 Mean Wind Speeds - Waglan Island (South) 1975 - 1994 Mean Wind Speeds - Waglan Island (South-west) 1975 - 1994 Mean Wind Speeds - Cheung Chau (South-east) 1970 - 1991 Mean Wind Speeds - Cheung Chau (South) 1970 - 1991 Mean Wind Speeds - Cheung Chau (South-west) 1970 - 1991 (19969) (19969) (19969) (19969) 1961 - 1990 - 1885 - 1939, 1947 - 1994 Page No. 173 174 175 176 177

6 7 8 9 10 11 12 13 14 15 16 17

179

- () 1968 - 179 1994 - 1975 - 1994 - 1970 - 1991 - () 1975 - 1994 - () 1975 - 1994 - () 1975 - 1994 - () 1975 - 1994 - () 1975 - 1994 - () 1970 - 1991 - () 1970 - 1991 - () 1970 - 1991 180 180 181 181 181 182 182 182 183 183

173 Table No 18 19 20 21 22 Extreme Sea Levels - Quarry Bay/North Point 1954 - 1994 Extreme Sea Levels - Tai Po Kau 1962 - 1994 Extreme Sea Levels - Ko Lau Wan 1974 - 1994 Extreme Sea Levels - Tsim Bei Tsui 1974 - 1994 Extreme Sea Levels - Waglan Island 1976 - 1994 1954 - 1994 / Page No. 183 184 184 184 185

- 1962 - 1994 - 1974 - 1994 - 1974 - 1994 - 1976 - 1994

173

Table 1

Basic Characteristics of Vessels - Local Craft Register (September 1996) - (19969)

No. of Vessels Licenced Type/ Category of Vessel Mech. NonMech. 118 1838 241 2853 2 13 3 5068 Total

Approximate Dimensions (m) Length O.A. Beam Depth Draft (m) Displacement (t) 15 - 3000 4 - 25 25 - 3200 13 - 6200 20 - 3200 20 - 40 16 - 350 150 - 58000 0.5 - 800 7 - 1400 pass. 1 - 5 pass. 25 - 2200 t 15 - 2200 t 5-6t 1 - 200 pass. Carrying Capacity Service Speed (m/s) 3 - 17 2-4 2-5 5-7 3 - 15 2 - 30

1. 2. 3. 4. 5. 6. 7. 8. 9.

Motor Launches & Ferry Vessels Class I I Class II II Class III III Class IV IV Class V V Class VI VI Class VII VII Class VIII VIII Passenger Sampans Wooden, Steel Cargo / Work Barges Stationary Craft Fishing, Misc. Craft Trading Vessels Fishing Vessels Floating Dock / Workshops Float Restaurant

1077 126 437 2005 72 3606 5860 Total 13183

1077 244 2275 241 4858 72 3608 13 3 5860 18251

6.5 - 62.0 5.0 - 7.8 6.2 - 133.6 4.6 - 45.5 3.7 - 132.3 7.7 - 36.1 6.7 - 39.3 23.2 - 270.0 52.5 - 76.2 1.7 - 62.5

2.4 - 19.1 2.1 - 3.5 1.8 - 29.9 2.5 - 24.6 1.0 - 30 3.1 - 8.5 2.8 - 19.3 8.5 - 59.2 15.1 - 22.7 1.7 - 11.3

1.0 - 4.0 0.7 - 1.0 2.0 - 4.0 1.0 - 4.0 0.7 - 4.0 1.0 - 3.0 1.0 - 2.0 2.0 - 3.0 0.3 - 5.0

0.5 - 3.0 0.3 - 0.7 1.0 - 3.0 0.7 - 3.0 0.3 - 3.0 0.7 - 2.5 0.7 - 1.5 1.0 - 2.5 0.2 - 4.5

10. Various Classes Pleasure Vessels

Mech. Non-Mech. Pass.

= Mechanical = Non-mechanical = Passenger

( ) ( ) ( )

174 Table 2 Basic Characteristics of Vessels - H.K.F. Ferry Fleet (September 1996) - (19969)

Type of Vessel

Length O.A. (m)

Beam (m)

Draft (m)

Displacement (t)

Carrying Capacity Pass. Veh.

Service Speed (m/s)

VEHICULAR FERRIES / 1. Double End Double Deck (Displacement < 1000t) / ( <1000t) Double End Double Deck (Displacement > 1000t) / ( >1000t) Double End Triple Deck (Night Club/Restaurant) / ( / ) 52.0 12.8 2.9 - 3.0 907 - 953 436-897 30-55 5.3-5.4

2.

64.0

13.4

3.2

1166 - 1366

440

65-70

5.2-5.4

3.

64.0

13.4

3.3 - 3.4

1326 - 1379

482-492

60

5.4

PASSENGER FERRIES 4. Single End Double Deck (Displacement < 200t) ( <200t) Single End Double Deck (Displacement 200 - 300t) ( 200-300t) Single End Double Deck (Displacement > 300t) ( >300t) Single End Triple Deck Double End Double Deck Double End Double Deck (Night Club/Restaurant) ( / ) HM2 HM5 30.0 7.0 1.5 107 436 6.1

5.

31.6 - 44.2

7.3 - 8.2

1.9 - 2.8

234 - 299

469 - 676

4.8 - 7.4

6.

43.1

8.2

2.4

301 - 309

596 - 598

6.9

7. 8. 9.

49.4 - 65.5 35.7 - 64.0 64.0

10.1 - 11.6 8.4 - 13.4 13.4

2.2 - 2.8 2.2 - 3.2 3.2

534 - 965 357 - 1366 1113

1071- 1646 717 - 1697 536

6.7 - 8.2 5.1 - 6.1 5.9

10. Hovercraft 11. Hoverferry

15.7 - 18.2 26.2 40.0

5.8 9.7 10.1

1.6 - 1.7 2.6 1.7

21 - 28 144 136

77 - 108 200 433

13.4 14.4 17.0

12. Waterjet Catamaran

Pass. Veh.

= Passenger = Vehicle

( ) ( )

175 Table 3 Basic Characteristics of Vessels - The Star Ferry Fleet (September 1996) - (19969)

Name of Vessel 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. Celestial Star Meridian Star Solar Star Northern Star Night Star Shinning Star Day Star Twinkling Star Morning Star Silver Star Pacific Princess (launch) Golden Star World Star

Length O.A. (m)

Beam (m)

Draft (m)

Displacement (t)

Passenger Carrying Capacity

Service Speed (m/s)

33.8

8.6

2.4

241

566

5.7

39.2

7.9

3.8

251

100

6.2

44.4

9.3

3.6

383

762

6.0

176 Table 4 Basic Characteristics of Vessels - Discovery Bay Ferry Fleet (September 1996) - (19969)

Type & Name of Vessel 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. HM218 Sidewall HM218 Sidewall HM218 Sidewall Monohull GRP Monohull GRP Monohull GRP Waterjet Monohull Waterjet Monohull Waterjet Monohull Waterjet Monohull Waterjet Monohull Waterjet Monohull Waterjet Monohull Catamaran Catamaran Catamaran HC (RTS201) HC (RTS202) HC (RTS203) PB (DB8) PB (DB9) PB (DB23) PB (DB12) PB (DB15) PB (DB16) PB (DB19) PB (DB20) PB (DB21) PB (DB22) PB (DB1) PB (DB2) PB (DB3)

Length O.A. (m)

Beam (m)

Draft (m)

Displacement (t)

Passenger Carrying Capacity

Service Speed (m/s)

18.3

6.1

1.7

28.4

100

14.4

19.7

5.4

1.7

43

130

9.8

35.5

7.7

1.2

93

300

12.8

42.0

11.5

1.3

168

500

17.0

177

Table 5

Five-day Normals of the Meteorological Elements for Hong Kong 1961 - 1990
Mean Daily Maximum ( C) 18.5 18.9 18.5 18.5 19.1 18.7 17.3 17.8 18.8 19.6 19.2 18.3 19.7 20.6 21.8 22.1 21.4 22.7 22.8 24.0 24.2 25.4 26.6 26.6 27.6 28.6 28.6 28.8 28.8 29.4 29.5 29.2 30.0 30.6 31.2 31.2 31.3 Air Temperature Mean ( C) 15.7 16.1 15.6 15.6 16.2 16.0 14.8 15.3 16.0 16.8 16.7 15.8 16.9 17.9 18.8 19.2 18.8 20.0 20.2 21.3 21.6 22.6 23.6 23.8 24.7 25.7 25.9 26.0 26.2 26.8 26.9 26.8 27.5 28.1 28.5 28.5 28.7 Mean Daily Minimum ( C) 13.5 13.9 13.3 13.3 14.0 13.9 12.8 13.3 13.9 14.7 14.7 13.8 14.8 16.1 16.6 17.2 16.8 18.0 18.2 19.3 19.6 20.6 21.5 21.9 22.7 23.6 23.9 24.0 24.3 24.8 25.0 25.0 25.6 26.2 26.6 26.6 26.7 Wet-Bulb Temperature ( C) 12.6 13.0 12.6 12.8 13.6 13.6 12.4 13.1 13.9 14.6 14.7 13.7 14.8 15.9 16.8 17.3 16.9 18.0 18.3 19.4 19.5 20.6 21.5 21.7 22.6 23.4 23.8 23.6 23.9 24.5 24.8 24.6 25.1 25.6 26.0 26.0 26.1 The Observatory Dew Point ( C) 9.3 9.8 9.5 10.1 11.2 11.3 10.1 11.0 12.0 12.7 13.1 11.7 12.8 14.3 15.3 16.0 15.5 16.7 17.0 18.2 18.3 19.5 20.3 20.7 21.6 22.3 22.8 22.4 22.9 23.5 23.8 23.7 24.2 24.6 25.0 25.0 25.1 Relative Humidity (% ) 68 68 69 72 74 76 75 77 78 78 81 78 79 80 81 82 82 82 83 83 83 83 83 83 84 82 84 81 83 83 84 84 83 82 82 82 81 Rainfall (Mean Daily) () (mm) 1.0 0.4 0.5 0.7 0.6 0.9 1.0 1.7 1.3 1.3 2.2 2.8 2.4 0.9 1.7 2.4 3.4 2.1 6.3 6.3 4.6 3.4 4.8 6.9 8.0 6.6 12.1 12.3 7.6 14.3 15.2 12.9 12.8 16.0 7.7 11.2 11.3 Amount of Cloud (% ) 56 51 52 55 62 69 71 72 69 70 80 78 75 76 71 76 81 78 84 81 79 79 71 76 75 72 76 70 73 75 78 77 75 74 71 73 69 Sunshine (Mean Daily) () (hr) 5.1 5.7 5.3 5.4 4.6 3.8 3.5 3.6 3.9 4.0 2.6 2.9 3.3 3.2 3.7 3.1 2.5 3.0 2.5 2.9 3.1 3.9 5.1 4.4 4.7 5.4 4.7 5.4 4.8 4.6 4.2 4.5 5.3 5.6 6.6 5.9 6.8 Kings Park Prevailing Direction (deg.) 010 070 070 070 070 070 070 070 070 070 070 010 070 070 070 080 070 080 080 080 080 080 080 080 080 090 090 080 090 090 090 090 090 220 100 190 230 Wind Mean Speed (m/s) 7.2 6.2 6.6 6.7 6.4 6.6 6.9 6.8 6.1 6.1 6.4 7.1 6.3 6.2 5.5 6.0 6.8 5.5 6.2 5.6 5.6 5.2 4.9 4.9 4.8 4.5 5.1 6.3 5.4 5.5 5.9 5.9 5.7 6.3 5.8 6.1 5.5

M.S.L. Pressure (hPa) Jan 1 1- 5 6 - 10 11 - 15 16 - 20 21 - 25 26 - 30 31 - 4 5- 9 10 - 14 15 - 19 20 - 24 25 - 1 2- 6 7 - 11 12 - 16 17 - 21 22 - 26 27 - 31 1- 5 6 - 10 11 - 15 16 - 20 21 - 25 26 - 30 1- 5 6 - 10 11 - 15 16 - 20 21 - 25 26 - 30 31 - 4 5- 9 10 - 14 15 - 19 20 - 24 25 - 29 30 - 4 Observed at 1020.5 1020.6 1020.5 1020.3 1019.8 1019.3 1019.9 1019.6 1019.1 1018.0 1017.5 1018.5 1018.1 1017.5 1016.2 1014.8 1015.5 1014.9 1014.5 1013.8 1013.7 1012.6 1012.0 1011.9 1010.7 1010.6 1009.7 1008.8 1008.2 1007.1 1006.8 1006.9 1006.4 1005.6 1005.7 1005.3 1005.2

Feb 2

Mar 3

Apr 4

May 5

Jun 6

Waglan Island

178

Table 5

Five-day Normals of the Meteorological Elements for Hong Kong (Continued) () 1961 - 1990
Mean Daily Maximum ( C) 31.8 31.7 31.7 31.5 31.3 31.2 31.5 31.1 31.4 31.4 31.4 31.1 31.0 30.9 30.6 30.2 29.6 29.1 29.1 28.6 28.4 27.4 26.9 26.0 26.1 25.0 24.3 23.9 22.9 21.9 21.7 21.6 20.3 20.3 20.0 18.9 Air Temperature Mean ( C) 29.1 28.9 28.8 28.7 28.5 28.5 28.7 28.4 28.4 28.4 28.5 28.2 28.2 27.9 27.8 27.5 27.0 26.6 26.4 25.9 25.6 24.8 24.3 23.3 23.3 22.5 21.7 21.1 20.1 18.8 18.7 18.6 17.5 17.6 17.2 16.2 Mean Daily Minimum ( C) 27.0 26.7 26.5 26.5 26.3 26.3 26.5 26.3 26.2 26.2 26.4 26.0 25.9 25.7 25.7 25.3 24.9 24.6 24.3 23.9 23.5 22.8 22.3 21.3 21.2 20.5 19.6 18.8 17.9 16.2 16.4 16.3 15.2 15.4 14.9 14.0 Wet-Bulb Temperature ( C) 26.2 26.1 26.0 26.1 25.8 26.0 26.1 25.9 25.9 25.8 25.8 25.6 25.5 24.9 24.6 24.3 24.0 23.6 22.8 22.3 22.1 21.5 20.9 19.9 19.8 19.2 18.5 17.5 16.2 14.8 15.1 15.3 13.9 14.4 14.1 13.2 The Observatory Dew Point ( C) 25.1 25.0 24.9 25.0 24.7 25.0 25.1 24.9 24.8 24.7 24.6 24.5 24.4 23.6 23.0 22.8 22.6 22.1 20.9 20.4 20.2 19.5 18.9 17.8 17.7 17.1 16.2 14.8 12.9 11.2 11.9 12.4 10.6 11.4 11.1 10.2 Relative Humidity (% ) 80 80 80 81 81 82 82 82 82 81 80 81 81 78 76 77 78 77 73 72 73 74 73 72 72 72 72 69 65 63 67 69 66 69 70 69 Rainfall (Mean Daily) () (mm) 5.8 10.8 11.3 9.7 12.8 14.9 8.7 14.7 14.7 14.3 9.3 12.1 12.6 12.9 5.8 6.9 10.3 7.9 4.6 3.5 7.9 5.3 3.0 3.3 0.9 1.8 1.6 1.2 0.6 0.8 1.7 0.6 0.2 1.1 0.6 1.1 Amount of Cloud (% ) 62 65 63 67 65 67 66 69 66 65 64 66 65 60 63 63 64 63 53 53 55 56 60 53 50 61 63 51 49 44 44 44 44 52 53 60 Sunshine (Mean Daily) () (hr) 8.0 7.7 8.1 7.4 6.8 6.6 6.9 6.6 6.8 6.8 6.8 6.1 6.5 6.6 6.2 6.0 5.5 5.5 6.6 6.8 6.5 6.2 5.7 6.3 6.7 5.4 5.0 6.4 6.1 6.7 6.3 6.5 6.3 5.5 5.4 4.8 Kings Park Prevailing Direction (deg.) 230 230 230 230 230 230 230 090 090 230 090 090 090 090 090 090 090 090 090 090 090 080 080 090 080 080 080 080 080 360 080 080 080 080 070 070 5.4 5.5 5.5 5.5 5.6 5.4 4.7 5.7 5.1 5.5 4.7 4.6 5.7 5.0 5.5 6.4 6.9 7.6 6.8 7.5 8.1 8.2 7.8 7.1 7.1 8.1 7.6 7.3 7.7 7.5 6.7 6.8 7.4 6.8 6.9 7.3 Wind Mean Speed (m/s)

M.S.L. Pressure (hPa) Jul 7 5- 9 10 - 14 15 - 19 20 - 24 25 - 29 30 - 3 4- 8 9 - 13 14 - 18 19 - 23 24 - 28 29 - 2 3- 7 8 - 12 13 - 17 18 - 22 23 - 27 28 - 2 3- 7 8 - 12 13 - 17 18 - 22 23 - 27 28 - 1 2- 6 7 - 11 12 - 16 17 - 21 22 - 26 27 - 1 2- 6 7 - 11 12 - 16 17 - 21 22 - 26 27 - 31 Observed at 1005.7 1005.8 1005.6 1005.7 1004.3 1004.4 1004.8 1004.8 1004.8 1004.8 1006.1 1006.4 1006.9 1007.6 1008.9 1009.9 1010.6 1011.3 1012.4 1013.3 1013.4 1014.6 1015.3 1016.4 1016.7 1016.5 1017.6 1018.2 1019.0 1020.1 1020.0 1019.7 1020.7 1020.3 1019.8 1020.3

Aug 8

Sep 9

Oct 10

Nov 11

Dec 12

Waglan Island

179

Table 6

Mean Hourly Wind Speeds - The Observatory - 1885 - 1939, 1947 - 1994 (m/s)

Return Period (yr) 5 10 20 50 100 200

N 16 19 22 26 29 31

NE 21 25 29 34 38 42

E 24 27 31 36 39 43

SE 21 25 29 34 38 42

S 15 18 21 24 27 30

SW 16 19 22 25 28 30

W 13 16 18 21 23 26

NW 13 16 18 22 24 27

Table 7

Mean Hourly Wind Speeds - Kai Tak Airport (South-east) - () 1968 - 1994 (m/s)

Return Period (yr) 5 10 20 50 100 200

N 15 17 19 22 24 26

NE 16 18 20 22 24 26

E 21 24 27 31 34 37

SE 19 22 25 29 32 35

S 16 18 21 24 27 29

SW 17 20 24 28 31 34

W 16 18 20 23 25 28

NW 13 15 16 18 19 21

180

Table 8

Mean Hourly Wind Speeds - Waglan Island - 1975 - 1994 (m/s)

Return Period (yr) 5 10 20 50 100 200

N 21 24 26 30 32 35

NE 23 27 31 35 38 41

E 28 32 36 41 45 48

SE 24 29 35 41 45 50

S 22 26 30 36 39 43

SW 22 27 32 37 41 45

W 19 22 26 31 34 37

NW 14 16 18 21 23 25

Table 9

Mean Hourly Wind Speeds - Cheung Chau - 1970 - 1991 (m/s)

Return Period (yr) 5 10 20 50 100 200

N 17 19 21 23 25 27

NE 19 22 25 29 32 35

E 24 28 31 36 39 42

SE 25 30 35 41 45 49

S 21 27 32 38 43 48

SW 19 23 28 34 38 42

W 17 21 25 30 34 37

NW 17 20 23 27 30 34

181

Table 10

Mean Wind Speeds - Waglan Island (North-east) - () 1975 - 1994 (m/s) Duration (hr) 3 4 22 20 25 23 28 26 32 29 35 32 38 35

Return Period (yr) 5 10 20 50 100 200

1 23 27 31 35 38 41

2 22 26 29 33 36 39

6 19 21 24 28 31 34

10 17 20 23 27 30 33

Table 11

Mean Wind Speeds - Waglan Island (East) - () 1975 - 1994 (m/s) Duration (hr) 3 4 26 26 31 30 34 34 39 38 43 42 46 45

Return Period (yr) 5 10 20 50 100 200

1 28 32 36 41 45 48

2 27 32 35 40 44 47

6 22 24 26 30 32 34

10 21 23 25 28 30 33

Table 12

Mean Wind Speeds - Waglan Island (South-east) - () 1975 - 1994 (m/s) Duration (hr) 3 4 22 26 31 37 42 46 21 25 30 36 40 44

Return Period (yr) 5 10 20 50 100 200

1 24 29 35 41 45 50

2 23 28 33 39 43 48

6 18 23 27 32 36 40

10 13 15 18 21 23 25

182

Table 13

Mean Wind Speeds - Waglan Island (South) - () 1975 - 1994 (m/s) Duration (hr) 3 4 17 20 23 28 31 34 16 19 22 27 30 33

Return Period (yr) 5 10 20 50 100 200

1 22 26 30 36 39 43

2 21 25 29 34 38 42

6 14 17 20 24 27 30

10 11 14 17 20 22 25

Table 14

Mean Wind Speeds -Waglan Island (South-west) - () 1975 - 1994 (m/s) Duration (hr) 3 4 17 20 23 26 29 32 15 18 20 23 25 27

Return Period (yr) 5 10 20 50 100 200

1 22 27 32 37 41 45

2 18 21 24 28 31 34

6 14 16 19 21 23 26

10 12 13 15 18 19 21

Table 15

Mean Wind Speeds - Cheung Chau (South-east) - () 1970 - 1991 (m/s) Duration (hr) 3 4 18 20 23 26 29 32 17 20 22 26 28 31

Return Period (yr) 5 10 20 50 100 200

1 25 30 35 41 45 49

2 23 28 33 39 43 48

6 15 18 20 23 25 27

10 12 14 16 18 20 22

183

Table 16

Mean Wind Speeds - Cheung Chau (South) - () 1970 - 1991 (m/s) Duration (hr) 3 4 18 17 22 21 27 26 33 32 38 36 42 40

Return Period (yr) 5 10 20 50 100 200

1 21 27 32 38 43 48

2 20 26 31 38 43 47

6 14 19 23 29 33 37

10 9 11 13 17 19 21

Table 17

Mean Wind Speeds - Cheung Chau (South-west) - () 1970 - 1991 (m/s) Duration (hr) 3 4 14 12 18 15 21 18 25 22 29 25 32 28

Return Period (yr) 5 10 20 50 100 200

1 19 23 28 34 38 42

2 15 19 23 28 32 35

6 11 14 16 19 22 24

10 9 10 11 13 14 16

Table 18

Extreme Sea Levels - Quarry Bay/North Point - / 1954 - 1994 Sea Level (mCD) 2.8 3.1 3.2 3.3 3.5 3.6 3.8

Return Period (yr) 2 5 10 20 50 100 200

184

Table 19

Extreme Sea Levels - Tai Po Kau - 1962 - 1994 Sea Level (mCD) 3.1 3.5 3.8 4.0 4.3 4.6 4.8

Return Period (yr) 2 5 10 20 50 100 200

Table 20

Extreme Sea Levels - Ko Lau Wan - 1974 - 1994 Sea Level (mCD) 2.8 3.0 3.1 3.2 3.4 3.5 3.6

Return Period (yr) 2 5 10 20 50 100 200

Table 21

Extreme Sea Levels - Tsim Bei Tsui - 1974 - 1994 Sea Level (mCD) 3.2 3.5 3.6 3.8 4.0 4.2 4.3

Return Period (yr) 2 5 10 20 50 100 200

185

Table 22

Extreme Sea Levels - Waglan Island - 1976 - 1994 Sea Level (mCD) 2.6 2.9 3.0 3.2 3.4 3.6 3.7

Return Period (yr) 2 5 10 20 50 100 200

186

187

Figure

188

189

LIST OF FIGURES
Figure No. 1 2 3 4 5 6 7 8 9 10 11 Wind Record Stations Annual Wind Rose Deep-water Wave Prediction Curves Shallow-water Wave Prediction Curves Constant Depth 6.0 metres Shallow-water Wave Prediction Curves Constant Depth 9.0 metres Shallow-water Wave Prediction Curves Constant Depth 12.0 metres Shallow-water Wave Prediction Curves Constant Depth 15.0 metres Tide Gauge Locations Morison's Equation, Linear (Airy) Wave Theory Variation of Inertia Wave Force with Depth Morison's Equation, Linear (Airy) Wave Theory Variation of Drag Wave Force with Depth Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Inertia Wave Force with Depth Water Level at 0.4d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Inertia Wave Force with Depth Water Level at 0.3d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Inertia Wave Force with Depth Water Level at 0.2d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Inertia Wave Force with Depth Water Level at 0.1d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Inertia Wave Force with Depth Water Level at Still-water Level 6.0 9.0 12.0 15.0 Page No. 191 192 193 194 194 195 195 196

Morison (Airy) 197 Morison (Airy) 198 Morison (Airy) 199 0.4d Morison (Airy) 200 0.3d Morison (Airy) 201 0.2d Morison (Airy) 202 0.1d Morison (Airy) 203

12

13

14

15

190 Figure No. 16 Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Drag Wave Force with Depth Water Level at 0.4d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Drag Wave Force with Depth Water Level at 0.3d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Drag Wave Force with Depth Water Level at 0.2d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Drag Wave Force with Depth Water Level at 0.1d above Still-water Level Morison's Equation, Linear (Airy) Wave Theory Variation of Percentage of Total Drag Wave Force with Depth Water Level at Still-water Level Rock Armour Design Curves - Structure Trunk Rock Armour Design Curves - Structure Head Page No. Morison (Airy) 204 0.4d Morison (Airy) 205 0.3d Morison (Airy) 206 0.2d Morison (Airy) 207 0.1d Morison (Airy) 208 - - 209 210

17

18

19

20

21 22

You might also like