13.01.2013 Views

Mitigation of Marine Aggregate Dredging Impacts ... - Cefas - Defra

Mitigation of Marine Aggregate Dredging Impacts ... - Cefas - Defra

Mitigation of Marine Aggregate Dredging Impacts ... - Cefas - Defra

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Marine</strong><br />

<strong>Aggregate</strong> Levy<br />

Sustainability Fund<br />

MALSF<br />

<strong>Mitigation</strong> <strong>of</strong> <strong>Marine</strong> <strong>Aggregate</strong><br />

<strong>Dredging</strong> <strong>Impacts</strong> – Benchmarking<br />

Equipment, Practices and Technologies<br />

against Global Best Practice<br />

MEPF Ref No: MEPF 08/P33 Project Date: April 2010


Benchmarking Equipment, Practices and Technologies<br />

�<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

��������������������������������������������������<br />

�<br />

�<br />

����������������������������������������<br />

��������������������������������������������<br />

����������������������������������������������<br />

�<br />

�����������������������������������<br />

�<br />

�����������<br />

�<br />

�<br />

������������������ � ��������������� � �������������� � �<br />

�<br />

�<br />

�<br />

� ������������������������������������������������������������������������<br />

� �����������������������������������������������������������������������������<br />

� ������������������������������������������������������������������������������������������<br />

i


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

ii


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

iii


Benchmarking Equipment, Practices and Technologies<br />

�����������������������<br />

�����������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������<br />

���������������������<br />

�������������������������������������������������������������������������������������������������<br />

������������������������<br />

�����������������������<br />

�������������������������������������������������<br />

������������������������������������������������������������������������������������������������������<br />

����������������������������������<br />

�<br />

������������������������<br />

���������������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������������<br />

���������������������������������������������������<br />

��������������������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������<br />

�������������������������������������������������<br />

�<br />

�����������<br />

������������������������������������������������������������������������������������������������������<br />

���� ������ ��� ���� ������ ��� ������� � ��� ��� ������ ������ ���� ������ ��� ���� ��� ���� ����������� ���������<br />

��������������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������������������������<br />

������������������������������������������������������<br />

�<br />

��������������������<br />

���������������������������������������������������������� � � � �<br />

�����������������������<br />

������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������<br />

���������������������������������������������<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

iv


Benchmarking Equipment, Practices and Technologies<br />

����������������<br />

���������������������������������������������������������������������������������������������<br />

�����������������������������������������������<br />

�<br />

�����������������������<br />

���������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������<br />

��������������������������������������������������������������������������<br />

�����������<br />

��������������������������������������������������������������������������������������������<br />

�������������������������������������<br />

�<br />

������������������<br />

�����������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

v


Benchmarking Equipment, Practices and Technologies<br />

���������<br />

��� ������������ ����������������������������������������������������������������������������������������������������������� �� 3<br />

Report No: 10/J/1/06/1309/0996<br />

1.1� Context........................................................................................................................... 5� 3<br />

1.2� Approach........................................................................................................................5� 3<br />

1.3� Outputs .......................................................................................................................... 6� 5<br />

��� ��� ��������� ��� �������������� �������� ����������� ����� ������� ����������<br />

����������������������������������������������������������������������������������������������������������������� �� 6<br />

2.1� Introduction ....................................................................................................................9� 6<br />

2.2� The dredging cycle......................................................................................................... 9� 6<br />

2.2.1� Review <strong>of</strong> impacts general to all shipping.................................................... 11� 8<br />

2.2.2� <strong>Impacts</strong> associated with transit phase <strong>of</strong> dredging cycle............................. 14� 11<br />

2.2.3� <strong>Impacts</strong> associated with loading phase <strong>of</strong> dredging cycle ........................... 16� 13<br />

2.2.4� <strong>Impacts</strong> Associated with discharging phase <strong>of</strong> the dredging cycle.............. 28� 25<br />

2.3� Conclusions ................................................................................................................. 29� 26<br />

��� ������� ���������� ��������� ��� ���� ��� �� �������������� ������������� ����<br />

��������������������������������������� �������������������������������������������������������������� 28 ���<br />

3.1� Description <strong>of</strong> the current English fleet ........................................................................ 30� 28<br />

3.1.1� Introduction .................................................................................................. 30� 28<br />

3.1.2� Vessel Age................................................................................................... 33� 31<br />

3.1.3� Design.......................................................................................................... 35� 33<br />

3.1.4� Crew............................................................................................................. 38� 36<br />

3.1.5� Size.............................................................................................................. 39� 37<br />

3.1.6� Power and Productivity Rates...................................................................... 40� 38<br />

3.2� UK Technology ............................................................................................................ 42� 40<br />

3.2.1� Dragheads ................................................................................................... 42� 40<br />

3.2.2� Dredge pumps, impellers and pipes ............................................................ 44� 42<br />

3.2.3� Screening and Loading................................................................................ 52� 50<br />

3.2.4� Overflow control........................................................................................... 55� 53<br />

3.2.5� Discharge Systems...................................................................................... 60� 58<br />

3.3� <strong>Dredging</strong> Practices, <strong>Mitigation</strong> and Regulatory Framework......................................... 63� 61<br />

3.3.1� <strong>Dredging</strong> Practices....................................................................................... 63� 61<br />

3.3.2� <strong>Mitigation</strong>...................................................................................................... 67� 65<br />

3.3.3� Regulatory Framework................................................................................. 70� 68<br />

��� ��������� ������� ���������� ����������������������������� ���� ���������� �������<br />

���������������������� ����������������������������������������������������������������������������������������� 72<br />

���<br />

4.1� Description <strong>of</strong> current world fleet ................................................................................. 73� 72<br />

4.1.1� Introduction .................................................................................................. 73� 72<br />

4.1.2� Vessel Age................................................................................................... 73� 72<br />

4.1.3� Design.......................................................................................................... 74� 73<br />

4.1.4� Crew............................................................................................................. 76� 75<br />

vi


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

4.1.5� Size.............................................................................................................. 76� 75<br />

4.1.6� Power and Productivity Rates...................................................................... 79� 78<br />

4.2� World Technology ........................................................................................................ 79� 78<br />

4.2.1� Dragheads ................................................................................................... 79� 78<br />

4.2.2� Dredge pumps, impellers and pipes ............................................................ 92� 91<br />

4.2.3� Screening and Loading................................................................................ 97� 96<br />

4.2.4� Overflow control........................................................................................... 97� 96<br />

4.2.5� Discharge systems..................................................................................... 104� 103<br />

4.3� <strong>Dredging</strong> Practices, <strong>Mitigation</strong> and Regulatory Frameworks ..................................... 105� 104<br />

4.3.1� <strong>Dredging</strong> Practices..................................................................................... 105� 104<br />

4.3.2� <strong>Mitigation</strong> and dredging practices .............................................................. 110� 109<br />

4.3.3� Regulatory Frameworks............................................................................. 117� 116<br />

��� �������������������������������������������������������������������������������������������������� ���� 122<br />

5.1.1� Technology ................................................................................................ 124� 123<br />

5.1.2� <strong>Dredging</strong> Practices, <strong>Mitigation</strong> and Regulatory Framework....................... 127� 126<br />

Conclusions................................................................................................................................ 132� 131<br />

���������� ���������������������������������������������������������������������������������������������������������������������� ���� 134<br />

���������� Methods to estimate overflow losses........................................................................ ���� 144<br />

������������ Air pollution mitigation................................................................................................���� 148<br />

�<br />

��������<br />

Figure 1-1: Summary <strong>of</strong> project approach.......................................................................................................6� 4<br />

Figure 2-1: Schematic description <strong>of</strong> an entire dredging cycle .....................................................................10� 7<br />

Figure 2-2: Sound spectrum <strong>of</strong> dredging noise recorded at varying distances and ambient noise level (from<br />

<strong>Defra</strong>, 2003) ..................................................................................................................................................17� 14<br />

Figure 2-3: Summary <strong>of</strong> direct and indirect impacts associated with dredging .............................................19� 16<br />

Figure 2-4: Schematic diagram showing the likely re-colonization rates for benthic communities <strong>of</strong> estuarine<br />

mud, sand, gravel and rocky reefs (Newell and Seiderer, 2003) ..................................................................21� 18<br />

Figure 2-5: Direct impact <strong>of</strong> the dredging process ........................................................................................22� 19<br />

Figure 2-6: Indirect impacts <strong>of</strong> the dredging process ....................................................................................27� 24<br />

Figure 3-1: Typical 1950s Sand Dredger. .....................................................................................................32� 30<br />

Figure 3-2: The newest vessel dredging English Licence Areas – DEME’s vessel “Charlemagne” (Image<br />

courtesy <strong>of</strong> BMAPA)......................................................................................................................................34� 32<br />

Figure 3-3: Ages <strong>of</strong> vessels in the current fleet (years).................................................................................34� 32<br />

Figure 3-4: The "Donald Redford" - a converted grab dredger. (Image courtesy <strong>of</strong> BMAPA).......................35� 33<br />

Figure 3-5: Example <strong>of</strong> TSHD with an aft configuration - the "Sand Falcon". (Image courtesy <strong>of</strong> BMAPA). 36� 34<br />

Figure 3-6: Example <strong>of</strong> TSHD with forward configuration - the "City <strong>of</strong> Westminster". (Image courtesy <strong>of</strong><br />

BMAPA). .......................................................................................................................................................36� 34<br />

Figure 3-7: The TSHD “City <strong>of</strong> Chichester” – a 2,300 Tonne Vessel serving South Coast ports. (Image<br />

courtesy <strong>of</strong> BMAPA)......................................................................................................................................40� 38<br />

Figure 3-8: Engine room on the "Sand Fulmar". (Image courtesy <strong>of</strong> BMAPA)..............................................41� 39<br />

Figure 3-9: Simple single visor draghead. (Image courtesy <strong>of</strong> BMAPA). ......................................................42� 40<br />

Figure 3-10: California style draghead. (Image courtesy <strong>of</strong> BMAPA). ..........................................................43� 41<br />

Figure 3-11: Inboard double casing pump. (Image courtesy <strong>of</strong> BMAPA)......................................................45� 43<br />

Figure 3-12: Vertical hydraulic transport .......................................................................................................46� 44<br />

Figure 3-13: Mixture density as a function <strong>of</strong> the mixture velocity at minimum pump inlet pressure ............48� 46<br />

Figure 3-14: Mixture density and production at the vacuum limit..................................................................49� 47<br />

Figure 3-15: Influence <strong>of</strong> pump depth on density and production for two cases <strong>of</strong> dredging at 50m depth..50� 48<br />

Figure 3-16: Specific energy for hydraulic transportation (suction)...............................................................51� 49<br />

Figure 3-17: Fixed single screen on the “City <strong>of</strong> Cardiff”. (Image courtesy <strong>of</strong> BMAPA). ...............................53� 50<br />

Figure 3-18: Rotating screening tower on the “Arco Dijk”. (Image courtesy <strong>of</strong> BMAPA)...............................53� 50<br />

vii


Benchmarking Equipment, Practices and Technologies<br />

Figure 3-19: Sand level in hopper .................................................................................................................54� 52<br />

Figure 3-20: Determining optimal loading time .............................................................................................55� 53<br />

Figure 3-21: The three phases <strong>of</strong> loading and overflow................................................................................56� 54<br />

Figure 3-22: Velocity and concentration distribution in a hopper ..................................................................57� 55<br />

Figure 3-23: Settling velocity as a function <strong>of</strong> the particle size .....................................................................58� 56<br />

Figure 3-24: Screen and spillway overflows on the "Arco Dijk". (Image courtesy <strong>of</strong> BMAPA). .....................59� 57<br />

Figure 3-25: Rejects from the underside <strong>of</strong> the screen deck combine in the vertical overflow weir on the<br />

“City <strong>of</strong> Chichester”. The combined flow then exits through the bottom <strong>of</strong> the vessel. (Image courtesy <strong>of</strong><br />

BMAPA). .......................................................................................................................................................59� 57<br />

Figure 3-26: The "Donald Redford" being discharged by shore crane. (Image courtesy <strong>of</strong> BMAPA). ..........60� 58<br />

Figure 3-27: Modern drag scraper system on the "City <strong>of</strong> London". (Image courtesy <strong>of</strong> BMAPA)................61� 59<br />

Figure 3-28: Bucket wheel discharger on the "Sand Heron". (Image courtesy <strong>of</strong> BMAPA). .........................61� 59<br />

Figure 3-29: Grab discharger on the "City <strong>of</strong> Chichester". (Image courtesy <strong>of</strong> BMAPA). .............................62� 60<br />

Figure 3-30: 26m forward boom conveyor. (Image courtesy <strong>of</strong> BMAPA)......................................................63� 61<br />

Figure 3-31: Example <strong>of</strong> Microplot display showing Active Area boundary (red line); dredging lanes (blue<br />

boxes); contaminant symbols (individual red crosses and diamonds) and exclusion areas (solid red).<br />

(Image courtesy <strong>of</strong> HAML). ...........................................................................................................................64� 62<br />

Figure 3-32: Example <strong>of</strong> dredging intensity plot from the EMS system (from The Crown Estate, 2009). .....65� 63<br />

Figure 3-33: Regulatory procedure for obtaining an English aggregate licence (<strong>Marine</strong> and Fisheries<br />

Agency, 2009). ..............................................................................................................................................72� 70<br />

Figure 4-1: Design configurations <strong>of</strong> TSHDs <strong>of</strong> the world fleet .....................................................................74� 73<br />

Figure 4-2: Practical results <strong>of</strong> improved hull shapes: top the "Lange Wapper" (without bulb); bottom the<br />

"Charlemagne" (with bulb). (Image courtesy <strong>of</strong> BMAPA)..............................................................................75� 74<br />

Figure 4-3: Operator position on TSHD "Brabo". (Image from IHC, 2007). ..................................................76� 75<br />

Figure 4-4: Modern IHC draghead. ...............................................................................................................80� 79<br />

Figure 4-5: The cutting process in granular sediments .................................................................................81� 80<br />

Figure 4-6: Jet production and density as a function <strong>of</strong> the jet pressure.......................................................84� 83<br />

Figure 4-7: Comparison between specific energy for cutting and jetting ......................................................85� 84<br />

Figure 4-8: The Wild Dragon draghead on board the TSHD “Xin Hai Long” demonstrating the high power<br />

water jets (from IHC, 2005). ..........................................................................................................................87� 86<br />

Figure 4-9: Yield indicators comparing the Wild Dragon draghead (left) and conventional draghead (right)<br />

when dredging the Yangtze (from IHC, 2005)...............................................................................................87� 86<br />

Figure 4-10: 7.2m wide draghead on the TSHD "Seriyu-maru" (from Yano et al., 2006). ............................89� 87<br />

Figure 4-11: The Vic Vac dredge head (image used with permission <strong>of</strong> J.F. Brennan Co., Inc.). ................90� 89<br />

Figure 4-12: Isometric view from underneath the Vic Vac dredge head (image used with permission <strong>of</strong> J.F.<br />

Brennan Co., Inc.). ........................................................................................................................................90� 89<br />

Figure 4-13: Schematic <strong>of</strong> the Tornado Motion eddy pump (image used with permission <strong>of</strong> Tornado Motion<br />

Technologies)................................................................................................................................................91� 90<br />

Figure 4-14: Tornado Motion suction head (image used with permission <strong>of</strong> Tornado Motion Technologies).<br />

......................................................................................................................................................................91� 90<br />

Figure 4-15: Cross-section through a conventional dredge pump (from IHC, 2004). ...................................93� 92<br />

Figure 4-16: Cross-section through a high efficiency pump (from IHC, 2004). .............................................93� 92<br />

Figure 4-17: Increased output when using IHC high efficiency pumps and impellers in different substrates<br />

(from IHC, 2004). ..........................................................................................................................................94� 93<br />

Figure 4-18: The Punaise Submerged <strong>Dredging</strong> Pump. ...............................................................................95� 94<br />

Figure 4-19: The Aft Centre Dredge Pipe on the TSHD “Seriyu-maru”.........................................................96� 95<br />

Figure 4-20: Telescopic overflow pipe (HAM 317) ........................................................................................98� 97<br />

Figure 4-21: Loading production as a function <strong>of</strong> the inflow concentration ...................................................99� 98<br />

Figure 4-22� Influence <strong>of</strong> water temperature on settling velocity.......................................................99� 98<br />

Figure 4-23: Anti-turbidity overflow system (from Ofuji and Ishimatsu, 1976).............................................100� 99<br />

Figure 4-24: Sketch showing the low turbidity valve (Jan de Nul, 2003).....................................................101� 100<br />

Figure 4-25� Water level in hopper with or without the influence <strong>of</strong> the green valve .......................102� 101<br />

Figure 4-26: Sketch showing the "Green Pipe" system (Jan de Nul, 2003)................................................103� 102<br />

Figure 4-27: Comparison tubidity plume dispersions (Jan de Nul, 2003) ...................................................104� 103<br />

Figure 4-28: Screen display <strong>of</strong> the IHC Dredge Track Presentation System (from Mallee, 2000)..............106� 105<br />

Figure 4-29: Example <strong>of</strong> EMS record from a Dutch monitoring system (from Sutton and Boyd, 2009)......108� 107<br />

Figure 4-30: Analysis <strong>of</strong> dredging intensity in Køge Bugt, Denmark (from Sutton and Boyd, 2009)...........108� 107<br />

Figure 4-31: On board dredge display for the Dredge Specific System (from Minerals Management Service,<br />

2004) ...........................................................................................................................................................109� 108<br />

Figure 4-32: Silt curtain in Alberta, Canada, showing high sediment concentrations contained on the right<br />

hand side <strong>of</strong> the curtain...............................................................................................................................112� 111<br />

Figure 4-33: An example <strong>of</strong> the cumulative effect <strong>of</strong> multiple environmental windows applied to the same<br />

dredging project ..........................................................................................................................................114� 113<br />

Report No: 10/J/1/06/1309/0996<br />

viii


Benchmarking Equipment, Practices and Technologies<br />

Figure 4-34: Principles <strong>of</strong> feedback monitoring (from Jensen, 2006) ............................................116� 115<br />

Figure A-1 Ideal settling basin according to Camp ................................................................... 144 145�<br />

Figure A-2� Removal ratio according to Camp including the influence <strong>of</strong> turbulence ................... 145 146�<br />

Figure A-3� Velocity and concentration distribution during loading (2DV model )..........................148� 147<br />

Figure B-1� NOx limits (from www.dieselnet.com) ........................................................................150� 149<br />

Figure B-2� Power consumption <strong>of</strong> various transport means<br />

� (from Gabrielli and Von Karman, 1950) ...................................................................... 153 152<br />

Figure B-3� Comparison <strong>of</strong> Energy Index for different means <strong>of</strong> transportation. ...........................154� 153<br />

�������<br />

Table 2-1: General and dredging specific impacts generated by TSHDs .....................................................11� 8<br />

Table 2-2: Composition <strong>of</strong> World Fleet (2007). .............................................................................................12� 9<br />

Table 2-3: MARPOL Annexes, Waste Categories and Types <strong>of</strong> Waste .......................................................13� 10<br />

Table 2-4: Number <strong>of</strong> recorded marine pollution incidents involving the English fleet (BMAPA, 2009)........13� 10<br />

Table 2-5: Indicative port to port times (does not include discharge) and transit times for three regional<br />

cases <strong>of</strong> the English dredging industry. ........................................................................................................14� 11<br />

Table 2-6: Proportion <strong>of</strong> total fuel consumed during transit for long haul and short haul cases ...................15� 12<br />

Table 2-7: Typical vessel capacities and loading times for UK aggregate dredging regions. .......................16� 13<br />

Table 2-8: Approximate fuel consumption during dredging...........................................................................16� 13<br />

Table 2-9: Processes contributing to dredging noise during loading (Thomsen et al., 2009) .......................17� 14<br />

����������������������������������������������������������������������������������������������� Pennekamp, 1996)........................................................................................................................................25 22<br />

Table 2-11: Approximate fuel consumption during discharge .......................................................................28� 25<br />

Table 2-12: Matrix <strong>of</strong> relative sensitivities associated with the different phases <strong>of</strong> the dredging cycle.........29� 26<br />

Table 3-1: Operators and ships <strong>of</strong> the UK Fleet............................................................................................33� 30<br />

Table 3-2: Examples <strong>of</strong> third-party vessels that dredge on English licence areas........................................33� 31<br />

Table 3-3: Fuel usage and CO2 emissions for the English fleet (BMAPA, 2009)..........................................37� 35<br />

Table 3-4: General crew structure <strong>of</strong> English marine aggregate dredgers ...................................................38� 36<br />

Table 3-5: Sizes <strong>of</strong> vessels in the English dredging fleet (Clarkson Research Services Ltd, 2009) .............39� 37<br />

Table 3-6: Approximate production rates for two English dredging regions..................................................41� 39<br />

Table 3-7: Reduction in dredging hours 2006-2008 (BMAPA, 2009)............................................................66� 64<br />

Table 3-8: Production and dredging hours 2006-2008 (BMAPA, 2009)........................................................66� 64<br />

Table 4-1: Total world dredging fleet (Clarkson Research Services Ltd, 2009)............................................73� 72<br />

Table 4-2: Mean sizes <strong>of</strong> the world fleet and the maximum and minimum examples (Clarkson Research<br />

Services Ltd, 2009) .......................................................................................................................................77� 76<br />

Table 4-3: Currently commissioned and proposed megadredgers, as <strong>of</strong> November 2009...........................78� 77<br />

Table 4-4: Reported maximum and minimum engine, generator and dredge pump powers (Clarkson<br />

Research Services Ltd, 2009).......................................................................................................................79� 78<br />

Table 4-5: Examples <strong>of</strong> high productivity rates .............................................................................................79� 78<br />

Table 4-6: Results <strong>of</strong> including knives and water jets on a draghead (from Brogdon Jr et al., 1994). ..........88� 87<br />

Table 4-7:Suspended sediment concentrations with and without an Anti-turbidity Overflow System on the<br />

TSHD "Kairyu Maru" (from Herbich and Brahme, 1991). ............................................................................101� 100<br />

Report No: 10/J/1/06/1309/0996<br />

ix


Benchmarking Equipment, Practices and Technologies<br />

��� ���������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������<br />

�������������������������������<br />

��� ����������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������<br />

��� ���������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������<br />

��������������������<br />

��� �������������������������������������������������������������������������� � ���������� � ��<br />

��������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

1


Benchmarking Equipment, Practices and Technologies<br />

��������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������<br />

��� �������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������<br />

�������������������������������������������������<br />

��� ����������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������<br />

��������������������������<br />

��� ���������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������<br />

������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

2


Benchmarking Equipment, Practices and Technologies<br />

��� �������������<br />

���� ��������<br />

Report No: 10/J/1/06/1309/0996<br />

<strong>Marine</strong> aggregate dredging in English waters has been carried out in the same way, on a similar<br />

scale and using similar equipment for around 3 decades. Production <strong>of</strong> marine aggregates during<br />

this period rose initially, has remained relatively static for the last 20 years and has the potential<br />

to rise in the future. Elsewhere in the world during this time dredging engineering, materials and<br />

practices have changed; dredging activity has intensified and understanding <strong>of</strong> the marine<br />

environment and dredging impacts has advanced significantly. Currently dredging best practice is<br />

not necessarily shared between developers, or regulators and their environmental assessors, as<br />

a result <strong>of</strong> competition, culture and knowledge. In addition the English marine aggregate industry<br />

is relatively isolated – commonly new information on dredging practice or technology is not<br />

circulated amongst marine aggregates businesses which are firmly focussed in England.<br />

Consequently the review <strong>of</strong> practices and technologies will result in a range <strong>of</strong> new information<br />

for all stakeholders.<br />

This project will therefore review current marine aggregate dredging in English waters, and<br />

current English marine aggregates practices and technologies are benchmarked against<br />

dredging techniques used in other parts <strong>of</strong> the world. This project examines dredging best<br />

practice, case studies and equipment from around the world and assesses whether they are<br />

applicable, necessary or realistic for the English industry.<br />

Quantitative information about novel technologies is typically commercially confidential, difficult to<br />

obtain and quantitative comparisons are therefore <strong>of</strong>ten impossible. This project is not intended<br />

to provide solutions; however it will indicate where opportunities for developing and improving<br />

practice may exist. It is also important to note that this report will not look at the mitigation <strong>of</strong><br />

potential effects far from dredging areas. These can not be mitigated by alteration <strong>of</strong> dredging<br />

technology and are beyond the scope <strong>of</strong> this study.<br />

The equipment and methods used to achieve recovery and transport <strong>of</strong> marine aggregates are<br />

intimately associated with the environmental impacts <strong>of</strong> the process. This project will review and<br />

benchmark long standing English equipment, technologies and methods and identify<br />

opportunities to develop aggregate dredging mitigation and management techniques where<br />

appropriate. Benchmarking therefore <strong>of</strong>fers a clear potential to evolve marine aggregate<br />

extraction and promote environmentally sensitive practices through a clear understanding <strong>of</strong><br />

global best practice and available technologies. This project will assist the industry in assessing<br />

the ‘fitness for purpose’ <strong>of</strong> its current technology and mitigation strategies. It will also provide<br />

regulators and advisors with confidence regarding the best practice <strong>of</strong> aggregate dredging and<br />

the proportionality <strong>of</strong> dredging proposals.<br />

���� ���������<br />

This project will produce a position statement <strong>of</strong> the technologies and techniques in use by the<br />

English dredging industry and compare these with worldwide technologies and techniques.<br />

The approach the project will follow is summarized in Figure 1-1.<br />

3


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

4


Benchmarking Equipment, Practices and Technologies<br />

Outputs include:<br />

Report No: 10/J/1/06/1309/0996<br />

1. A summary <strong>of</strong> the technologies, capabilities and practices <strong>of</strong> the current English and<br />

world dredging fleets.<br />

2. Enhanced understanding <strong>of</strong> global practice and available technologies and<br />

confidence in whether current English technology and mitigation is proportionate<br />

3. Potential future investigation areas for development.<br />

4. Identification <strong>of</strong> risks - a global view <strong>of</strong> dredging will assist regulators and their<br />

advisors in assessing marine aggregate dredging proposals, and the industry in<br />

proposing ‘fit for purpose’ mitigation strategies<br />

5. Improving business performance through improved environmental and industry<br />

performance, potential efficiency improvement, investment guidance and application<br />

acceptability.<br />

This report will provide a high level review <strong>of</strong> the opportunities available and identifies knowledge<br />

gaps for further consideration.<br />

5


Benchmarking Equipment, Practices and Technologies<br />

��� ������������������������������������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

������������������������������<br />

���� �������������<br />

<strong>Marine</strong> aggregate dredging results in a range <strong>of</strong> impacts on the environment and on ecosystem<br />

goods and services (Austen et al., 2009). Some <strong>of</strong> these impacts are general to all types <strong>of</strong><br />

shipping while some are specifically a result <strong>of</strong> the operation <strong>of</strong> a marine aggregate dredging<br />

vessel. However, not all <strong>of</strong> the impacts specific to dredgers occur at all phases <strong>of</strong> the dredging<br />

cycle. This chapter will therefore define the dredging cycle, and give an overview <strong>of</strong> the<br />

environmental impacts associated with all shipping and specific to dredging vessels. It will<br />

provide an overview <strong>of</strong> which dredging specific impacts occur at which phase <strong>of</strong> the dredging<br />

cycle, and assesses the relative importance <strong>of</strong> these at the different phases.<br />

���� �������������������<br />

The English marine aggregate business involves the movement <strong>of</strong> a relatively low value product<br />

(sand and gravel) from where it occurs <strong>of</strong>fshore, to markets where it is required. The total<br />

aggregate dredging cycle may be divided into components, each component <strong>of</strong> which will have<br />

different environmental impacts. These components are:<br />

� Transit;<br />

� Loading;<br />

� Discharge.<br />

The cycle is shown schematically in Figure 2-1. English marine aggregate dredging involves, in<br />

most cases, the dredger sailing between multiple wharves and multiple licences.<br />

6


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

��������������������������������������������������������������<br />

The pr<strong>of</strong>it equation for the English marine aggregate industry involves maximising the individual<br />

cargo size and the number <strong>of</strong> cargoes delivered per year and minimising the costs, particularly<br />

fuel. The optimisation <strong>of</strong> cycle time and speed whilst maintaining good environmental<br />

performance is therefore critical to the economics <strong>of</strong> the business. However, with the exception <strong>of</strong><br />

deliveries to some Continental ports, most dredging cycles are governed by the tides and so<br />

most are defined in blocks <strong>of</strong> 12 hours. Because cycle times are governed so strongly by tides<br />

improving production rate is not such an issue for the English industry since this would <strong>of</strong>ten<br />

simply increase the amount <strong>of</strong> time a vessel was idle waiting to enter a tidal port. Improving<br />

production rates is a much stronger driver for the maintenance and capital dredging sectors<br />

where minimising time equates much more strongly with minimising costs.<br />

7<br />

��������<br />

Empty vessel sails from<br />

wharf to Licence Area<br />

��������<br />

Sediment is dredged from<br />

the Licence Area and<br />

loaded into the vessel<br />

��������<br />

Full vessel sails from<br />

Licence Area to wharf.<br />

Often a different wharf from<br />

its origin<br />

����������<br />

Offloading the cargo from the<br />

vessel at the wharf


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Smaller vessels with short steaming distances achieve 12 hour cycles and deliver over 400<br />

cargoes per year. Larger vessels with longer steaming times can sometimes achieve 24 hour<br />

cycles if the dredging time is limited to approximately 4 hours as sometimes occurs on the East<br />

Coast when sandy cargoes are required. More typically they require 36 hour cycles, and the<br />

largest vessel with very long steaming times can take 48 hours to complete a cycle – for example<br />

if delivering a cargo into a Thames port from Licence Areas around the Isle <strong>of</strong> Wight. Typically<br />

these larger vessels will deliver between 200 and 230 cargoes per year.<br />

������ ������������������������������������������<br />

This section <strong>of</strong> the report will give an overview <strong>of</strong> the environmental impacts associated with the<br />

dredging process. The environmental impacts <strong>of</strong> dredgers can be divided into impacts that are<br />

common to all types <strong>of</strong> shipping, and those that are specific to dredgers (Table 2-1).<br />

������������������������������������� �����������������������������<br />

Emissions Direct impacts <strong>of</strong> substrate removal<br />

Waste Indirect impacts <strong>of</strong> the sediment plume<br />

Noise generated by propulsion Noise generated by dredging process<br />

��������������������������������������������������������������������<br />

General shipping impacts are present during all phases <strong>of</strong> the dredging cycle, while impacts<br />

associated with substrate removal and generation <strong>of</strong> sediment plumes will only occur during the<br />

loading phase <strong>of</strong> the cycle.<br />

� ����������<br />

<strong>Marine</strong> fuels generate CO2, nitrogen oxide (NOx) and sulphur oxide (SOx) which are all<br />

greenhouse gasses. Vessel exhaust gasses can be scrubbed to remove SOx, and NOx<br />

emissions can be reduced through altering combustion process in the engine by changing the<br />

temperature in the engine. Adjusting the engine specification has the potential to reduce<br />

the NOx emissions by up to 20% while the introduction <strong>of</strong> selective catalytic recombiners (SCRs)<br />

can potentially reduce the levels by 90% while optimizing engine performance. The currently<br />

available NOx and SOx abatement technologies may be mutually exclusive – particularly<br />

the use <strong>of</strong> catalytic recombiners which require inlet gas temperatures much higher than the<br />

exhaust temperature <strong>of</strong> scrubbers. Further information on controlling pollution can be found<br />

in Appendix B.<br />

The IMO working group on Bulk, Liquid and Gas (BLG) has estimated the quantities <strong>of</strong> fuel used<br />

and emissions generated by the world’s shipping fleet based on the Lloyds Register, and Table<br />

2-2 below shows the composition <strong>of</strong> the world fleet in 2007.<br />

8


Benchmarking Equipment, Practices and Technologies<br />

���������������<br />

������������������<br />

Report No: 10/J/1/06/1309/0996<br />

����������<br />

�<br />

9<br />

������������<br />

Number Gross Tons Number Gross Tons<br />

Shipping 9,732 181,668,000 51,687 770,980,000<br />

<strong>Dredging</strong> 371 1,018,000 1,309 2,719,000<br />

Offshore supply 514 1,175,000 3,626 4,678,000<br />

Tugs 1,863 564,000 12,544 1,328,000<br />

����������������������������������������������<br />

The BLG estimates total fuel consumption in 2007 by the world fleet <strong>of</strong> approximately 370 million<br />

tonnes – composed <strong>of</strong> 285 million tonnes <strong>of</strong> HFO and 85 million tonnes <strong>of</strong> marine distillates, and<br />

this translates into approximately 1100 million tonnes <strong>of</strong> CO2. Table 2-2 shows that the numbers<br />

<strong>of</strong> dredging vessels relative to general fleet is small.<br />

� ������<br />

The International Convention for the Prevention <strong>of</strong> Pollution from Ships (MARPOL) was adopted<br />

on 2 November 1973 at IMO and covered pollution by oil, chemicals, harmful substances in<br />

packaged form, sewage and garbage.<br />

The MARPOL Convention is the main international convention covering prevention <strong>of</strong> pollution <strong>of</strong><br />

the marine environment by ships from operational or accidental causes. It is a combination <strong>of</strong> two<br />

treaties, adopted in 1973 and 1978 respectively, and updated by amendments through the years.<br />

Regulations set by MARPOL (73/78) and Directive 2000/59/EC <strong>of</strong> the European Parliament on<br />

port reception facilities for ship-generated waste and cargo residues also stipulate that ports<br />

should provide reception facilities for vessels to safely dispose and manage the various types <strong>of</strong><br />

waste designated. MARPOL designates marine waste under 6 Annexes (Table 2-3).<br />


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

������ ������������������ ����������������<br />

I Oil Covers all types <strong>of</strong> wastes from the carriage <strong>of</strong> oil: As<br />

II Noxious liquid<br />

substances in bulk<br />

III Harmful substances<br />

carried by sea in<br />

packaged form<br />

fuel, engine room slops, cargo (tank washings) or dirty<br />

ballast water<br />

Chemical wastes derived from bulk chemical<br />

transportation, including residues and mixtures<br />

containing noxious substances<br />

Hundreds <strong>of</strong> specific substances defined by the<br />

International Maritime Dangerous Goods Code<br />

IV Sewage from ships Raw sewage retained in holding tanks for disposal in<br />

port or outside 12 nm. Partially treated sewage retained<br />

in holding tanks for disposal in port or outside 4 nm<br />

V Rubbish from ships Rubbish includes domestic (food and packaging) and<br />

operational (maintenance, cargo and miscellaneous)<br />

wastes<br />

VI Air pollution from ships Sets limits on SOx and NOx emissions from ship<br />

exhausts and prohibits deliberate emissions <strong>of</strong> ozone<br />

depleting substances<br />

���������������������������������������������������������������<br />

National regulations require port authorities and terminal operators to provide reception facilities<br />

for ships waste to allow vessels to manage their obligations under MARPOL. All dredging vessels<br />

working on English licence areas fulfil their obligations under the regulations.<br />

Information on recorded marine pollution incidents involving English marine aggregate vessels<br />

has been published for the last three years as part <strong>of</strong> the industry’s commitment to environmental<br />

reporting (BMAPA, 2009). Table 2-4 shows the number <strong>of</strong> recorded marine pollution incidents<br />

from 2006-2008.<br />

� ����� ����� �����<br />

Incidents (all minor hydraulic leaks) 6 0 6<br />

���������������������������������������������������������������������������������������������������<br />

10


Benchmarking Equipment, Practices and Technologies<br />

� ������<br />

General noise generated by shipping is dominated by the ship engine and propeller and these<br />

are <strong>of</strong> a low frequency relative to higher frequency dredging specific noise generated by<br />

aggregates rising up through the suction pipe, movement <strong>of</strong> the draghead on the seabed and<br />

splashing from the spillways (Parvin ��� ���� 2008).<br />

������ ��������������������������������������������������������<br />

� �<br />

Report No: 10/J/1/06/1309/0996<br />

<strong>Marine</strong> aggregates in the UK are produced by 11 companies – the majority <strong>of</strong> which are<br />

members <strong>of</strong> the British <strong>Marine</strong> <strong>Aggregate</strong>s Producers Association. The BMAPA member<br />

companies have recognized that “the marine environment in which it operates is sensitive, and<br />

accepts that it has a responsibility to manage its operations in ways that minimise any effects on<br />

the marine environment and on its other users” (BMAPA, 2006).<br />

Transit times for UK aggregate dredging areas vary on the geographical location <strong>of</strong> the dredging<br />

area compared with the originating / deliver wharves. Long transit times can mean an increased<br />

impact with increased fuel consumption relative to the tonnage dredged. Table 2-5 below shows<br />

indicative 2-way transit times for three regional case studies. These times are port to port times<br />

and do not include time in port when the vessel is discharging. This is covered in Section 2.2.3.<br />

������������������� ����������� ����� ��� ��������������������� ����������� ������ ���<br />

�����������<br />

���������<br />

Bristol Channel 12 50 6<br />

South Coast 12 70-80 8-9.5<br />

East Anglia/Eastern<br />

English<br />

London<br />

Channel &<br />

25 70 20<br />

������ ����� ����������� ����� ��� ����� ������ ������ ���� �������� ����������� ���� �������� ������ ���� ������ ���������<br />

����������������������������������������<br />

� �������������������������������<br />

Fuel consumption for a dredging vessel varies depending on the scale <strong>of</strong> the vessel and the<br />

phase <strong>of</strong> the dredging cycle. The English dredging industry has, for the last three years, been<br />

reporting their fuel consumption figures as part <strong>of</strong> their ongoing ‘Strength From The Depths’<br />

sustainable development report (BMAPA, 2009) and energy consumption has been studied as<br />

part <strong>of</strong> The Crown Estate’s research programme (Kemp, 2008). Based on data from these<br />

reports Table 2-6 shows the estimated proportion <strong>of</strong> fuel consumed during transit for two<br />

dredging cases – long haul (larger ships) or short haul (smaller ships).<br />

11


Benchmarking Equipment, Practices and Technologies<br />

� �������<br />

Report No: 10/J/1/06/1309/0996<br />

���������������� �������������� �����������������<br />

>3000 tonnes Long haul e.g.<br />

East Anglia / EEC<br />


Benchmarking Equipment, Practices and Technologies<br />

������ ��������������������������������������������������������<br />

� �<br />

Report No: 10/J/1/06/1309/0996<br />

The time taken by a dredging vessel to load a cargo is controlled by a range <strong>of</strong> variables<br />

including the size (capacity) <strong>of</strong> the vessel, the power <strong>of</strong> the dredge pump, the amount <strong>of</strong><br />

screening taking place and the composition <strong>of</strong> the seabed sediment. Given this, Table 2-7<br />

summarizes some indicative values for typical vessel capacities and loading times for five<br />

regions <strong>of</strong> activity for the English aggregate industry.<br />

������� �������� ������� ���������<br />

���������<br />

Bristol Channel 750 2-4<br />

South Coast 750 2-4<br />

East Anglia 2750 5-8<br />

Eastern English Channel 5000 5-8<br />

Humber 4500 5-8<br />

13<br />

���������������������������<br />

������������������������������������������������������������������������������������������<br />

� �������������������������������<br />

� �������<br />

Based on data from BMAPA (2009) and Kemp (2008) the proportion <strong>of</strong> fuel consumed during<br />

loading has been estimated for two dredging cases – long haul (larger ships) or short haul<br />

(smaller ships) (Table 2-8).<br />

���������������� �������������� �����������������<br />

�����������������<br />

��������<br />

>3000 tonnes Long haul e.g.<br />

East Anglia / EEC<br />


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

����������� ������������<br />

Collection<br />

noise<br />

Pump<br />

noise<br />

Transport<br />

noise<br />

Deposition<br />

noise<br />

Ship/machinery<br />

noise<br />

Noise arising from collection <strong>of</strong> sediment from the sea floor e.g. the operation<br />

<strong>of</strong> the drag head<br />

Noise from the pump driving suction through the dredge pipe<br />

Noise <strong>of</strong> the sediment being lifted from the sea floor to the dredger i.e. for a<br />

TSHD the noise <strong>of</strong> the material in the dredge pipe<br />

Noise associated with the placement <strong>of</strong> sediment within the hopper<br />

Noise associated with the ship machinery, including propeller and thruster<br />

noise.<br />

������������������������������������������������������������������������������������������<br />

Noise associated with dredging is predominantly <strong>of</strong> low frequency – below 1 kHz (Thomsen et al.,<br />

2009); however within this, the noise derived from aggregates rising up through the suction pipe,<br />

the movement <strong>of</strong> the draghead on the seabed and splashing from the spillways are <strong>of</strong> higher<br />

frequencies than those generated by the ship engine and propeller (Parvin ������� 2008).<br />

Greene (1987) and Richardson et al. (1995) published studies describing the noise emitted by<br />

five vessels during dredging in the Beaufort Sea. These measurements were made in the early<br />

1980’s. In these studies the TSHDs produced the loudest noise during loading and the greatest<br />

range at which the dredging noise could be detected above background noise was 25 km.<br />

<strong>Defra</strong> (2003) took hydrophone measurements <strong>of</strong> noise produced by the TSHD Arco Adur<br />

operating at Licence Area 328 to the east <strong>of</strong> Great Yarmouth (Southern North Sea) and at<br />

Licence Area 366 on the Hastings Shingle Bank in the Eastern English Channel. The noise<br />

produced was predominantly <strong>of</strong> low frequency, below 500 Hz (Figure 2-2).<br />

���������������������������������������������������������������������������������������������������������<br />

�������������<br />

14


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Their findings also showed that:<br />

� At low frequencies (2 kHz), higher levels <strong>of</strong> noise were generated by full suction<br />

dredging activities.<br />

Parvin et al. (2008) found that the 100 m long TSHD City <strong>of</strong> Westminster, powered by two 1950<br />

kW engines, loading at 2.5 m 3 s -1 and dredging at an average speed <strong>of</strong> 2 knots, increased<br />

underwater noise in close proximity to the vessel over a broad range <strong>of</strong> frequencies (20 Hz – 80<br />

kHz approx.) and that the dredging noise would fall below the ambient noise level (and so not be<br />

audible) at a range <strong>of</strong> 6 km. This distance is substantially less than that measured by Richardson<br />

et al. (1995) and Thomsen et al. (2009) suggests that this is because the English Channel has<br />

significantly higher levels <strong>of</strong> background noise than the Beaufort Sea, Alaska.<br />

Thomsen et al. (2009) also concluded that the available data indicated that the noise generated<br />

during the loading phase was not as noisy as seismic surveys, pile driving or sonar; but louder<br />

than ship transits (including TSHD transits) and should be viewed as a medium impact activity. It<br />

was also concluded that the noise generated during loading is potentially <strong>of</strong> bigger relevance for<br />

seals and marine fish than for cetaceans, as the overlap between the emitted frequency<br />

spectrum and the bandwidth <strong>of</strong> hearing is bigger in seals and fish than for cetaceans (Thomsen<br />

et al., 2009).<br />

A new MALSF funded project (MEPF 09 / P108) has been commissioned which aims to produce<br />

a baseline dataset for underwater noise associated with a range <strong>of</strong> UK dredger types, deposit<br />

types and screening conditions.<br />

� ������������������������������������������<br />

The impacts <strong>of</strong> loading <strong>of</strong> marine aggregates in English waters are based upon the physical<br />

disturbance impacts related to the removal <strong>of</strong> coarse sediment substrate and can be divided into<br />

the direct impacts <strong>of</strong> substrate removal and the indirect impacts <strong>of</strong> the generated sediment<br />

plume. These are summarized in Figure 2-3.<br />

15


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

����������������������������������������������������������������������������<br />

16


Benchmarking Equipment, Practices and Technologies<br />

� ���������������<br />

The direct impacts <strong>of</strong> dredging result from the removal <strong>of</strong> the substrate in the licensed dredging<br />

area and are summarised in Figure 2-5. <strong>Dredging</strong> will alter the local bathymetry by directly<br />

removing sediment, and the length <strong>of</strong> time that dredging features (dredge trails) will remain<br />

distinct on the seabed depends on the sediment type and the transport potential <strong>of</strong> local tides<br />

and waves (Diesing et al., 2006). For example, at an experimental dredged gravel site <strong>of</strong>f Norfolk<br />

in 25 m <strong>of</strong> water, dredge tracks appeared to have been completely eroded within three years <strong>of</strong><br />

the cessation <strong>of</strong> dredging (Kenny and Rees, 1994, 1996; Kenny et al., 1998) while erosion <strong>of</strong><br />

dredge tracks in areas <strong>of</strong> moderate wave exposure and tidal currents have been observed to<br />

take from three to more than seven years in gravelly sediments (Limpenny et al., 2002; Boyd et<br />

al., 2003a, 2005; Cooper et al., 2005). In relict deposits, such as those typically dredged by the<br />

English industry, the gross alteration <strong>of</strong> the seabed topography and bathymetry is likely to be<br />

permanent and this means that while individual dredge trails will be eroded over time, the wider<br />

depressions will typically remain.<br />

Report No: 10/J/1/06/1309/0996<br />

Alteration <strong>of</strong> local seabed topography can potentially impact the hydrodynamics, and before a<br />

licence is granted for aggregate extraction, potential changes in local waves and current patterns<br />

are assessed through site-specific modelling studies. Changes in wave heights and current<br />

direction after dredging can result in localized changes in erosional and depositional patterns<br />

(nearfield effects), and possibly even in shoreline changes (farfield effects). It was concluded by<br />

Van Rijn et al. (2005) that, in general, dredging has little influence on the macroscale current<br />

pattern and in most cases the current pattern is only changed in the direct vicinity <strong>of</strong> the<br />

dredged area.<br />

A further direct impact <strong>of</strong> English dredging practices is the potential alteration <strong>of</strong> the substrate<br />

over time. This is because screening <strong>of</strong> sediments is common in order to meet a specific sand<br />

and gravel requirement for the market. Hitchcock and Drucker (1996) and Newell et al. (1998)<br />

estimated that, during typical loading <strong>of</strong> a dredged cargo at some extraction areas in the UK, up<br />

to 1.6 – 1.7 times the total cargo is discharged into the surrounding water column as a<br />

consequence <strong>of</strong> the screening process. Clearly, estimates such as these are site�specific and<br />

will vary in relation to the grain size <strong>of</strong> seabed sediments, the grading required for the cargo, and<br />

the efficiency <strong>of</strong> the dredger. However over time, the progressive removal <strong>of</strong> the original sandy<br />

gravel or coarse sands, and their replacement by finer sandier sediment fractions through<br />

screening activities, may result in a gradual fining <strong>of</strong> the sediment within the extraction areas. In<br />

some instances, this increase in fine sand may be temporary because <strong>of</strong> the reworking<br />

capabilities <strong>of</strong> tides and waves. By altering the size distribution <strong>of</strong> the seabed from coarser to<br />

finer particles, the seabed has the potential to provide alternative habitats to other fauna which<br />

may play a similar functional role to the fauna that have been lost through the dredging process<br />

(Cooper et al., 2008).<br />

<strong>Dredging</strong> also has an obvious direct impact due to removal <strong>of</strong> benthic organisms and habitat<br />

within the dredging areas, along with infaunal and epifaunal organisms that are incapable <strong>of</strong><br />

avoiding dredging. A number <strong>of</strong> studies have investigated the recovery <strong>of</strong> benthic communities<br />

following dredging (Blake et. al., 1996; Newell et al, 1998, 2004; Boyd et al., 2003b, 2004) and<br />

17


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

have shown that community diversity and abundance can recover within several years. Newell<br />

and Seiderer (2003) also summarize recovery rates <strong>of</strong> benthic communities post-dredging for<br />

different substrate types (Figure 2-4) with sandy substrates typically recovering within 2 to 4<br />

years.<br />

������� ����� ���������� �������� �������� ���� ������� ���������������� ������ ���� �������� ������������ ���<br />

������������������������������������������������������������������������<br />

The Minerals Management Service (2004) suggests that though re-colonized communities may<br />

be similar in terms <strong>of</strong> total abundance and species diversity, their taxonomic composition, in<br />

terms <strong>of</strong> dominant species and species abundance, is <strong>of</strong>ten very different from pre- to postdredging.<br />

Benthic resources are important in the food web for commercially and recreationally<br />

important fishes and invertebrates, and they contribute to the biodiversity <strong>of</strong> the pelagic<br />

environment. Their removal can therefore lead to reduced species richness, biodiversity and food<br />

sources within the dredging areas – and subsequent potential effects on spawning, stock<br />

recruitment and displacement <strong>of</strong> organisms.<br />

18


Benchmarking Equipment, Practices and Technologies<br />

�<br />

������������������<br />

�<br />

��������������<br />

Report No: 10/J/1/06/1309/0996<br />

������������������������������������<br />

�<br />

���������������������������������<br />

�<br />

���������������������<br />

�����������������<br />

������������������<br />

�����<br />

�<br />

�������������<br />

�<br />

���������������������������<br />

��������������������������������������������������<br />

19<br />

����������������<br />

����������<br />

�������������<br />

�<br />

��������������������<br />

�<br />

��������������<br />

������������<br />

��������������������<br />

����������������<br />

���������


Benchmarking Equipment, Practices and Technologies<br />

� ����������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

<strong>Dredging</strong> leads to the production <strong>of</strong> plumes <strong>of</strong> suspended material, partly from the disturbance <strong>of</strong><br />

the seabed by the draghead but mainly from the outwash <strong>of</strong> sediment from the dredger’s<br />

screening towers and the hopper spillways. These plumes can have indirect impacts (i.e. impacts<br />

outside the dredging area itself) on the environment and these are summarized in Figure 2-6.<br />

The spatial extent and transport <strong>of</strong> the plumes depends on the sediment particle size, total<br />

quantity <strong>of</strong> material suspended, velocity <strong>of</strong> discharge, and the local hydrodynamics (Hitchcock<br />

and Drucker, 1996; Hitchcock and Bell, 2004).<br />

The water flowing overboard from the vessel contains sediment but because the coarse particles<br />

settle faster than fine sediments, the overflow discharge contains relatively more fines that the<br />

sediment loaded. The largest part <strong>of</strong> the sediment flux flowing overboard will settle close to the<br />

dredging vessel because the sediment water mixture has a density larger than the ambient seawater.<br />

This creates a density driven plume (also called the dynamic plume) which transports<br />

sediment directly towards the seabed. The remainder <strong>of</strong> the overflow sediment that remains in<br />

suspension (the so-called source term) is the passive plume. Reliable and calibrated models are<br />

not yet available for these plumes, however a PhD research study is being undertaken at<br />

Technical University Delft to develop a model to quantify the magnitude <strong>of</strong> this source term (De<br />

Wit, 2009).<br />

The dynamic plume is driven by initial momentum or buoyancy due to density differences<br />

between the plume and ambient water. The density and velocity in the plume is governed by the<br />

bulk behaviour <strong>of</strong> the plume and not by the individual fall velocity <strong>of</strong> the particles. In contrast, for<br />

a passive plume the density difference between the plume and surrounding water is negligible<br />

and the individual settling velocities <strong>of</strong> the particles determine the settling velocity <strong>of</strong> the plume<br />

itself. The dynamic plume behaviour is mostly present in the near field <strong>of</strong> the dredger, while the<br />

passive behaviour is observed in the far field (Spearman et al. 2004). An exact location <strong>of</strong> the<br />

transition between near field and far field cannot be given and depends on the operational<br />

process conditions. From an experimental study (Winterwerp, 2002) it appeared that near field<br />

overflow plume can be either dynamic or passive depending on a bulk Richardson number and a<br />

velocity ratio, defined as:<br />

� �<br />

Where :<br />

� � �<br />

� �� �<br />

��<br />

� �<br />

� � � � � �<br />

� � � ������<br />

�<br />

�<br />

� � �<br />

��<br />

�<br />

relative density difference [-]<br />

� �<br />

density <strong>of</strong> overflow [kg/m 3 ]<br />

� �<br />

density <strong>of</strong> ambient water [kg/m 3 ]<br />

� velocity <strong>of</strong> the mixture in the overflow [m/s]<br />

�<br />

� velocity <strong>of</strong> the ambient water [m/s]<br />

�<br />

20


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

For high values <strong>of</strong> the Richardson number in combination with low values <strong>of</strong> the velocity ratio<br />

dynamic plumes and density currents along the bed developed. For the opposite situation (low<br />

Richardson number and a high velocity ratio), the overflow mixed over the total water column and<br />

a passive plume developed.<br />

The (vertical) flow velocity <strong>of</strong> the sediment particles in a dynamic plume is governed by the bulk<br />

velocity <strong>of</strong> the total plume and is much higher than the individual settling velocities <strong>of</strong> the<br />

sediment particles in a passive plume. The velocity in a dynamic plume is <strong>of</strong> the order <strong>of</strong> [m/s]<br />

while the velocity in a passive plume will be order <strong>of</strong> [mm/s]. The higher transport capacity <strong>of</strong> the<br />

dynamic plume transports more sediment towards the seabed compared with the passive plume,<br />

hence measured turbidity in the vicinity <strong>of</strong> the dredger will be lower in that case (Van Parys<br />

et al., 2001).<br />

Apart from the flow situation in the overflow as described in Equation 2-1 the following near-field<br />

processes are also important in the vicinity <strong>of</strong> the vessel:<br />

1. Influence <strong>of</strong> entrained air on the plume behaviour.<br />

Entrained air can cause a lift effect by upward buoyant forces (air-lift effect), this can lift fine<br />

particles to the water surface, where passive plumes <strong>of</strong> fine sediment particles are formed.<br />

2. Influence <strong>of</strong> the TSHD vessel itself.<br />

When the plume enters the propeller wake <strong>of</strong> the vessel the mixing <strong>of</strong> the sediment in the plume<br />

is enhanced because <strong>of</strong> the high turbulence level in the wake. The transition from active to<br />

passive plume is therefore accelerated.<br />

3. Instationary behaviour <strong>of</strong> the plume<br />

The overflow discharge is far from stationary, this generates separate sediment puffs or clouds<br />

instead <strong>of</strong> a continuous plume. These individual clouds are more prone to passive behaviour<br />

than a continuous plume. The instationary behaviour therefore accelerates the transition from<br />

dynamic to passive behaviour.<br />

Field studies <strong>of</strong> plume concentrations and behaviour are limited, however some data have been<br />

reported. Wakeman et al. (1975) described turbidity field studies during the 1974 maintenance<br />

work at Mare Island Strait (San Francisco Bay, USA). He reported concentrations 50 m behind<br />

the TSHD <strong>of</strong> 210 mg/l at the surface without overflow and 75-350 mg/l with overflow. 10 m below<br />

the water surface those values were 230 mg/l without overflow and 165-870 mg/l with overflow.<br />

Bernard (1978) suggested that plumes with concentrations <strong>of</strong> 200-300 mg/l may extend behind a<br />

TSHD for distances up to 1200 m.<br />

Willoughby and Crabb (1983) found concentrations close to the dredger <strong>of</strong> about 500 mg/l near<br />

the bed and 50 mg/l near the surface when dredging sand (0.25 mm) when background sediment<br />

concentrations were approx. 5 mg/l. Within approximately one hour <strong>of</strong> dredging ceasing the<br />

concentrations in the plume fell to background values and about 90% <strong>of</strong> this reduction occurred<br />

within the first 20 minutes. Given the local current velocity <strong>of</strong> 0.6 m/s, the major proportion <strong>of</strong> the<br />

dredge suspended material settled within 600-700 m downcurrent <strong>of</strong> the TSHD.<br />

21


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Pennekamp (1996) and Kirby and Land (1991) performed field measurements in several Dutch<br />

harbour basins which had small ambient currents


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

Dieppe, where the indirect effects <strong>of</strong> sand discharged from the dredger were as great as the<br />

direct effects <strong>of</strong> extraction on macrobenthic species (Desprez, 2000). Newell et al. (2002, 2004)<br />

and Robinson et al. (2005) both found evidence <strong>of</strong> suppression <strong>of</strong> benthic biomass beyond the<br />

margins <strong>of</strong> dredging areas. They conclude that this is due to the indirect impact <strong>of</strong> sediment<br />

generated by screening.<br />

It should also be noted that the natural background conditions will also have an influence on the<br />

impacts associated with generated plumes – where habitats are subjected to natural disturbance<br />

by bedload transport (such as <strong>of</strong>f the East Coast <strong>of</strong> England) impacts <strong>of</strong> plumes will be minimised<br />

relative to areas that are otherwise stable (such as the Eastern English Channel). Similarly,<br />

sedimentation and resuspension caused by dredging in deposits <strong>of</strong> clean, mobile sands are<br />

generally thought to be <strong>of</strong> less concern, because the fauna inhabiting such areas tend to be<br />

adapted to naturally high levels <strong>of</strong> suspended sediment resulting from wave and tidal current<br />

action (Millner et al., 1977; Newell et al., 2002; Cooper et al., 2005).<br />

In certain circumstances fine sediments liberated by dredging may actually result in more positive<br />

impacts. One study <strong>of</strong> a fine�sediment site in Moreton Bay, Australia, demonstrated enhanced<br />

abundance <strong>of</strong> benthic invertebrates adjacent to dredged subtidal sandbanks, which may have<br />

been linked to sedimentation <strong>of</strong> plume material (Pointer and Kennedy, 1984).<br />

23


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

���������������������������������<br />

�������������<br />

�<br />

��������<br />

�<br />

����������<br />

�����������������������������<br />

�<br />

�����������������<br />

�<br />

��������������<br />

�<br />

���������������<br />

�<br />

���������<br />

�����������������������������������������������������<br />

24<br />

�<br />

����������<br />

�����������<br />

�<br />

��������<br />

���������<br />

�<br />

��������


Benchmarking Equipment, Practices and Technologies<br />

�<br />

2.2.4 <strong>Impacts</strong> Associated with discharging phase <strong>of</strong> the dredging cycle<br />

� �������������������������������������<br />

As in the cases for transit and dredging described above, the fuel consumption for the dredging<br />

vessel during discharge varies and based on data from BMAPA (2009) and Kemp (2008) the<br />

proportion <strong>of</strong> fuel consumed during discharge has been estimated for two dredging cases – long<br />

haul (larger ships) or short haul (smaller ships) (Table 2-11).<br />

� ������<br />

Report No: 10/J/1/06/1309/0996<br />

���������������� �������������� �����������������<br />

���������������<br />

��������<br />

>3000 tonnes Long haul e.g.<br />

East Anglia / EEC<br />


Benchmarking Equipment, Practices and Technologies<br />

���� ������������<br />

Report No: 10/J/1/06/1309/0996<br />

Table 2-12 below is a matrix summarizing the relative sensitivities associated with the different<br />

phases <strong>of</strong> the dredging cycle.<br />

� � ������<br />

������������ Transit Loading Discharge<br />

Time as<br />

proportion <strong>of</strong><br />

cycle<br />

High Moderate Low<br />

Fuel Consumption<br />

/ Emissions<br />

High Moderate Low<br />

Noise Level Low High Low<br />

Direct impact Seabed None High None<br />

Benthos None High None<br />

Water Column None Moderate None<br />

Indirect impact Seabed None Moderate None<br />

Benthos None Moderate None<br />

Water Column None High None<br />

Interaction with<br />

other users<br />

Moderate High None<br />

���������������������������������������������������������������������������������������������������������<br />

Table 2-12 shows that the transit time relative to the overall dredging cycle time is high; while fuel<br />

consumption and emissions are high for both transit and loading but relatively low on discharge.<br />

Noise levels are relatively high during loading due to the additional noise generated by the<br />

draghead and dredging equipment but relatively low during transit and discharge when the noise<br />

generated by a dredging vessel is no different to that generated by ships <strong>of</strong> similar size. Direct<br />

and indirect impacts on the seabed only occur during the loading phase <strong>of</strong> the dredging cycle<br />

while the potential sensitivity for interaction with other users is also highest during the loading<br />

phase. Those sensitivities that can be most influenced by technological and practical changes,<br />

and hence the focus for the benchmarking process, are the direct and indirect impacts caused by<br />

the dredging process itself.<br />

26


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

27


Benchmarking Equipment, Practices and Technologies<br />

��� ���������������������������������������������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

����������������������������������������<br />

���� ������������������������������������������<br />

������ �������������������������<br />

Whilst the dredging <strong>of</strong> marine aggregates has been carried out in various ways for hundreds <strong>of</strong><br />

years, it is only over the last 35 years or so that it developed on an industrial scale, to the point<br />

today where the business produces around 25 million tonnes per year from UK waters. Originally<br />

dredging rights and pr<strong>of</strong>its from the trade belonged to the Lord High Admiral <strong>of</strong> England but in<br />

1594 the privilege was transferred by Lord Howard to Trinity House. In the river Thames, an act<br />

<strong>of</strong> Elizabeth’s reign gave the Corporation exclusive rights to supply dredged ballast to vessels in<br />

the River Thames from London Bridge to the sea.<br />

<strong>Dredging</strong> from the seabed around the UK coast has been taking place for at least two hundred<br />

years – on 14 th June 1736 the Reverend William Gosling <strong>of</strong> Canterbury recorded “...I have got a<br />

piece <strong>of</strong> a huge bone (I suppose from an elephant’s thigh bone) petrified, which the dredgers <strong>of</strong><br />

copperas stones fishes out <strong>of</strong> the sea.” Another cleric, the Reverend James Parker <strong>of</strong> Northfleet<br />

in Kent, patented Roman Cement (sometimes called Parker’s Cement) in 1796 which was a very<br />

rapidly setting hydraulic cement particularly suitable for wet or under-water work. Roman cement<br />

was made from copperas stone being broken up, burned in kilns and ground into powder and<br />

was shipped to all parts <strong>of</strong> the UK and Northern Europe.<br />

In 1826, the Topographical Dictionary <strong>of</strong> England recorded in its entry for Harwich “...about one<br />

hundred small vessels and boats are employed in or near the harbour dredging for stone for<br />

making cement. The manufacture <strong>of</strong> copperas from stones, which are found in abundance on the<br />

shore, was carried on here in the seventeenth century...” and by 1835 there were five cement<br />

factories in Harwich with “some five hundred men employed in dredging the stone and<br />

manufacturing the cement”. The removal <strong>of</strong> “several hundred thousand tons” <strong>of</strong> stone from<br />

Beacon Cliff caused changes in the strength <strong>of</strong> tides which threatened to silt up to the mouth <strong>of</strong><br />

Harwich Harbour and in 1845 removal <strong>of</strong> stone was immediately stopped by The Commission on<br />

Harbours <strong>of</strong> Refuge. The consequence <strong>of</strong> this was to see a marked increase in dredging until, by<br />

1850, up to 400 smacks from Kent ports, each with a crew <strong>of</strong> three or four, were dredging stone<br />

from the West Rocks <strong>of</strong>f Walton with some <strong>of</strong>f Brightlingsea and <strong>of</strong>f Hythe. However, by 1890 the<br />

industry had died out as Roman cement was replaced by the cheaper chalk based Portland<br />

cement.<br />

In the early days, both aggregate and maintenance dredging was the by way <strong>of</strong> “spoon and bag”<br />

which both Holland & Italy claim to have originated but is thought more likely to have been<br />

introduced in western Europe and Britain by the Phoenicians or Romans who, as their empire<br />

expanded, brought the practice <strong>of</strong> dredging with them. The 1829 edition <strong>of</strong> the London<br />

Encyclopaedia gives an exhaustive description <strong>of</strong> spoon and bag dredging or “ballast heaving”<br />

which allowed two to four men lift up to sixty tons <strong>of</strong> ballast in a tide from a depth <strong>of</strong> some three<br />

fathoms. The ballast or stone was scooped up in a leather bag which had its mouth held open by<br />

28


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

an iron hoop attached to a pole, which pivoted, allowing the bag to be swung inboard with its<br />

load. Vessels were limited in the water depth they could safely work in to about 10 metres and<br />

<strong>of</strong>ten suffered from stability problems that resulted in a number <strong>of</strong> vessels capsizing and sinking.<br />

This eventually led the regulating authorities to produce the dredger stability criteria that are still<br />

in use today.<br />

The first suction dredger used in the UK sailed from Bristol on the 15 th June 1912. This was the<br />

49 ton, steam driven suction dredger “City <strong>of</strong> York” and she dredged a cargo <strong>of</strong> sand from the<br />

Bristol Channel and returned to Bristol and thus marked the start <strong>of</strong> the modern British aggregate<br />

dredging industry. The industry’s first purpose built aggregate dredger was named “Portway”<br />

after the riverside road linking Bristol with Avonmouth and launched at Charles Hill & Sons’<br />

Bristol yard in 1926. The innovative design <strong>of</strong> the 297 ton vessel saw her fitted with wing tanks<br />

for fuel oil, allowing her to double up as a bunkering barge. In 1930 the 1920-built coaster “Jolly<br />

Marie” was converted to a suction dredger and was the first to have a hydraulic discharge<br />

capability. Re-named the “Sandholm”, with a cargo capacity <strong>of</strong> 300 tons, she continued in the<br />

trade until 1962.<br />

The development <strong>of</strong> maintenance or ‘capital’ dredgers for such tasks as channel clearing and<br />

land reclamation progressed separately from that <strong>of</strong> that <strong>of</strong> the aggregate dredger by way <strong>of</strong> all<br />

manner <strong>of</strong> craft such as the water harrow, mud mill, scrapper dredger, grab dredger, & bucket<br />

dredger. The first hydraulic (suction) dredger using a centrifugal pump for dredging spoil was<br />

thought to have been that invented by M. Bazin in 1864 and it is this dredging method which was<br />

eventually adopted by the “City <strong>of</strong> York” for aggregate dredging. A paper published in the<br />

Scientific American <strong>of</strong> 1882 records ���������������������������������������������������������<br />

���������������������������������������������������������������������������������������������<br />

��������������������������������������������������������������������������������������������.<br />

Technical advances in the trade have continued apace since 1912 with over 250 ships and some<br />

50 companies having been involved in the industry which has variously been centred in the<br />

Bristol Channel, Solent / South Coast, Southern North Sea, Liverpool Bay, River Dee, River<br />

Clyde and a well established trade on Lough Neagh in Northern Ireland. By the 1950s, slurry<br />

pumps which allowed water to be pumped at sufficient velocity to entrain large volumes <strong>of</strong><br />

sediment <strong>of</strong> 150 mm diameter were common and the modern marine aggregate dredging<br />

industry began to develop (Figure 3-1).<br />

29


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

����������������������������������������<br />

Until 1965 vessels were, in the main, reliant on shore cranes to discharge them but then the first<br />

successful dry self discharging systems were developed, using the dragline principle. By the<br />

early 1970s vessel capacities has increased above 2,000 tonnes and steady development has<br />

continued to the present day. This has resulted in the current UK fleet <strong>of</strong> 26 vessels <strong>of</strong> various<br />

sizes, having a variety <strong>of</strong> discharge systems, and the most modern vessels capable <strong>of</strong> dredging<br />

the deepest approved seabed deposits. Currently there are 8 UK based operators <strong>of</strong> these<br />

vessels (Table 3-1).<br />

��������� ������� ����������� �������������� ���������<br />

���������<br />

Britannia<br />

<strong>Aggregate</strong>s<br />

Britannia Beaver 1991 2775 4800<br />

CEMEX UK<br />

<strong>Marine</strong><br />

DEME Building<br />

Materials<br />

Sand Falcon 1998 5022 8700<br />

Sand Fulmar 1998 4000 6920<br />

Sand Harrier 1990 2700 4671<br />

Sand Heron 1990 2700 4671<br />

Sand Weaver 1974 2400 4152<br />

Welsh Piper 1987 790 1367<br />

Charlemagne 2002 5000 8650<br />

Hanson<br />

Arco Adur 1988 2890 5000<br />

<strong>Aggregate</strong>s Arco Arun 1987 2890 5000<br />

30


Benchmarking Equipment, Practices and Technologies<br />

<strong>Marine</strong><br />

Northwood<br />

Report No: 10/J/1/06/1309/0996<br />

Arco Avon 1986 2890 5000<br />

Arco Axe 1989 2890 5000<br />

Arco Beck 1989 2600 4500<br />

Arco Dart 1990 700 1250<br />

Arco Dee 1990 700 1250<br />

Arco Dijk 1992 5100 8800<br />

Arco Humber 1972 4800 8000<br />

Donald Redford 1981 510 880<br />

(Fareham) Norstone 1971 1075 1860<br />

Severn Sands Argabay 1985 620 900<br />

Tarmac <strong>Marine</strong><br />

<strong>Dredging</strong><br />

Westminster<br />

City <strong>of</strong> Cardiff 1997 1300 2300<br />

City <strong>of</strong><br />

Chichester<br />

1997 1300 2300<br />

City <strong>of</strong> London 1990 2775 4800<br />

City <strong>of</strong><br />

Westminster<br />

1990 3000 5200<br />

Sospan 1990 700 1200<br />

<strong>Dredging</strong> Sospan Dau 1978 1400 2400<br />

������������������������������������������������<br />

A small number <strong>of</strong> third party vessels are also used on an occasional basis by English aggregate<br />

dredging companies when operational need dictates. Table 3-2 shows an example <strong>of</strong> third party<br />

vessels that also dredge on English licence areas.<br />

������� ����������� ����������� � �� ���������<br />

���������<br />

Reimerswaal 1994 1600 2740<br />

Orisant (now renamed as DC<br />

Vlaanderen 3000)<br />

2002 2600 4460<br />

Brabo 2007 11630 20000<br />

��������������������������������������������������������������������������������<br />

������ �����������<br />

The mean age <strong>of</strong> the fleet is currently 21.8 years with vessels ranging between 8 years old in the<br />

case <strong>of</strong> the “Charlemagne” (Figure 3-2) to 39 years old in the case <strong>of</strong> the “Norstone”. It should<br />

however be noted that the “Charlemagne” is not an exclusively English operating vessel – <strong>of</strong><br />

these the youngest are CEMEX UK <strong>Marine</strong>’s vessels “Sand Falcon” and “Sand Fulmar” which<br />

are both 12 years old.<br />

31


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

��������������������������������������������������������������������������������������������<br />

���������������������������<br />

A large number <strong>of</strong> the vessels were built in the late 1980s and early 1990s and the majority <strong>of</strong> the<br />

vessels in the fleet are a few years younger than the mean (Figure 3-3).<br />

����������������������������������������������������������<br />

32


Benchmarking Equipment, Practices and Technologies<br />

������ �������<br />

Report No: 10/J/1/06/1309/0996<br />

The English marine aggregates dredging fleet is comprised entirely <strong>of</strong> Trailer Suction Hopper<br />

Dredgers (TSHDs). All but two <strong>of</strong> the vessels currently in service were built as dedicated marine<br />

aggregate dredgers – the smallest vessel, the “Donald Redford”, was originally built as a<br />

maintenance grab dredger for the Manchester Ship Canal (Figure 3-4), while the newest (and<br />

one <strong>of</strong> the largest ships) the “Charlemagne” (Figure 3-2), was built as a multipurpose vessel<br />

capable <strong>of</strong> undertaking maintenance work as well as dredging marine aggregates.<br />

����������������������������������������������������������������������������������������<br />

There are two typical configurations that have been favoured by English TSHDs - 60% <strong>of</strong> the<br />

vessels have the accommodation and wheelhouse aft (Figure 3-5) and the remainder have them<br />

located forward (Figure 3-6). This latter arrangement has the accommodation well separated<br />

from the main propulsion and discharge machinery noise, gives the possibility <strong>of</strong> a much more<br />

flexible discharge system arrangement and unrestricted forward vision but has the disadvantage<br />

<strong>of</strong> somewhat restricted accommodation space and more uncomfortable vessel motions for the<br />

crew. They are designed to dredge on relatively shallow licence areas, close to markets and<br />

operate on quick turnarounds<br />

33


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

������������������������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������<br />

���������������������������<br />

The “City <strong>of</strong> Westminster” (Figure 3 6) is an example <strong>of</strong> an ‘A’ Class equivalent vessel. There are<br />

7 ‘A’ Class equivalent vessels and they form the backbone <strong>of</strong> the English dredging fleet. They<br />

comprise the Hanson <strong>Aggregate</strong>s <strong>Marine</strong> vessels “Arco Adur”, “Arco Arun”, “Arco Avon” and<br />

“Arco Axe”, the Tarmac <strong>Marine</strong> <strong>Dredging</strong> vessels “City <strong>of</strong> London” and “City <strong>of</strong> Westminster” and<br />

the Britannia <strong>Aggregate</strong>s vessel “Britannia Beaver”. There are minor differences in engine<br />

configurations, however all the ‘A’ Class equivalents have capacities between 2700 and 3000 m 3<br />

34


Benchmarking Equipment, Practices and Technologies<br />

� ������<br />

Report No: 10/J/1/06/1309/0996<br />

and were all built between 1986 and 1991 at the Appledore Shipyard in North Devon. The<br />

CEMEX UK <strong>Marine</strong> vessels “Sand Harrier” and “Sand Heron” show similar capacities but are<br />

slightly different and are not designated as ‘A’ Class equivalents.<br />

Generally, a ship’s fuel consumption varies as the third power <strong>of</strong> the speed up to about 12 knots<br />

but beyond this it can vary as the forth power or more so is a significant disincentive to increasing<br />

output by faster speeds, thus making increasing size the best solution to increased output when<br />

the wharves can accept such vessels.<br />

The smaller vessels serving the South Coast and South Wales areas have service speeds <strong>of</strong><br />

only 9/10 knots but even so, because the dredging areas are <strong>of</strong>ten less than 20 miles away and<br />

the loading and discharging times are around two hours, they are capable <strong>of</strong> delivering two<br />

cargoes per day – one on each tidal cycle.<br />

The largest vessels however can be steaming over 130 miles each way to and from the dredging<br />

grounds, take over 4 hours to load and discharge and <strong>of</strong>ten have to fit into 12 hour tidal windows.<br />

As steaming is a large proportion <strong>of</strong> their cycle time they usually need to have a service speed <strong>of</strong><br />

at least 12 knots to match the tidal cycles when the wharves, such as most <strong>of</strong> them on the<br />

Thames, are tidally restricted. On the Continent, however, such restrictions do not apply so this<br />

gives greater freedom to choose the most economical speed.<br />

� �����������������<br />

BMAPA (2009) reports fuel consumption and emissions figures for English dredging vessels as<br />

part <strong>of</strong> an ongoing sustainable development and environmental reporting programme. Reported<br />

figures separate out smaller vessels with cargo capacities <strong>of</strong> less than 3000 tonnes from larger<br />

vessels with capacities greater than 3000 tonnes. Smaller vessels typically supply local wharves<br />

from near-shore licence areas on short cycle times. Larger vessels typically dredge cargoes from<br />

more <strong>of</strong>fshore areas and supply more distant wharves and typically supply a cargo every 24- to<br />

48-hours (BMAPA, 2009). Table 3-3 shows the fuel consumed and CO2 emitted for the two<br />

categories <strong>of</strong> vessel.<br />

�������������� ������������������<br />

�������������������������<br />

3000 tonnes 2.308 7.363<br />

����������������������������������������������������������������������������<br />

� ������������<br />

35<br />

�� ������������<br />

������ ������������������<br />

Fuel has now become a very significant operating cost. With widely varying unit costs and the<br />

diverse vessel operating patterns it is difficult to generalise, but it would not be unusual to have<br />

fuel costs being between 25 and 40 % the operating cost <strong>of</strong> a vessel. All UK operators, however,<br />

are using high grade gas oil or marine diesel fuel despite the fact that savings <strong>of</strong> up to 40% could


Benchmarking Equipment, Practices and Technologies<br />

������ �����<br />

Report No: 10/J/1/06/1309/0996<br />

be achieved by the use <strong>of</strong> low grade residuals fuels. However, there are issues to be addressed<br />

before this could be considered viable – the availability <strong>of</strong> low sulphur fuel for use in coastal<br />

waters, the increased capital equipment cost, somewhat increased maintenance, and the ability<br />

to deliver the fuel in the required quantities and to locations where the vessels operate.<br />

All vessels <strong>of</strong> the UK fleet operate on a 24 hour a day, 7 day a week basis, with the majority <strong>of</strong><br />

the crews working a 2 or 3 week on/<strong>of</strong>f rota. The work can be very intense, with some vessels<br />

delivering more than 400 cargoes per year, so modern vessels have a good standard <strong>of</strong> crew<br />

accommodation, comparable with “deep sea” vessels <strong>of</strong> similar size.<br />

Crew numbers are not proportional to vessel size but relate to the number required to effectively<br />

maintain and safely operate each particular vessel and vary from 7 to 8 on the smaller vessels to<br />

11 to 12 on the larger vessels.<br />

Table 3-4 shows the general crew structure <strong>of</strong> English aggregate vessels:<br />

������������ �����������������������������������<br />

Master<br />

1 st Mate<br />

2 nd Mate Additional Mate<br />

Chief Engineer<br />

1 st Engineer 2 nd Engineer<br />

2 x ABS Additional ABS<br />

Cook<br />

�����������������������������������������������������������������������<br />

36


Benchmarking Equipment, Practices and Technologies<br />

������ �����<br />

Report No: 10/J/1/06/1309/0996<br />

The fleet ranges in capacity from about 880 tonnes to 8800 tonnes (Table 3-1) while Table 3-5<br />

summarizes the lengths, breadths and dredging draughts for the English fleet. It should be noted<br />

that several vessels in the English fleet have been modified, post-build, to maximise their<br />

capacity. Sand Falcon was lengthened, and some <strong>of</strong> the Hanson A class vessels have had<br />

raised cargo combings installed (together with retrospective bottom dump valves to mitigate<br />

stability concerns) (Russell, �����������).<br />

������� ����������� ������������ ���������<br />

Britannia Beaver 100 17.4 6.3<br />

Sand Falcon 120 19.5 7.9<br />

Sand Fulmar 100 19.5 7.9<br />

Sand Harrier 99 16.5 6.4<br />

Sand Heron 99 16.5 6.4<br />

Sand Weaver 96 16.7 6.1<br />

Welsh Piper 69 12.4 4.4<br />

Charlemagne 100 20.8 8.5<br />

Arco Adur 98 17.4 6.7<br />

Arco Arun 98 17.4 6.7<br />

Arco Avon 98 17.4 6.7<br />

Arco Axe 98 17.4 6.7<br />

Arco Beck 100 17 6.5<br />

Arco Dart 68 13 4.1<br />

Arco Dee 68 13 4.1<br />

Arco Dijk 113 20.1 7.7<br />

Arco Humber 107 20 8.2<br />

Donald Redford 53 11 3.4<br />

Norstone 67 12.5 4.5<br />

Argabay 58 9.4 3.9<br />

City <strong>of</strong> Cardiff 72 15 5.1<br />

City <strong>of</strong> Chichester 72 15 5.1<br />

City <strong>of</strong> London 100 17.4 6.3<br />

City <strong>of</strong> Westminster 100 17.4 6.7<br />

Sospan 57 10 3.6<br />

Sospan Dau 70 14.3 3.2<br />

37<br />

�����������<br />

�������������������������������������������������������������������������������������������������


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

As the industry has matured the vessel size and configuration has become better matched to the<br />

commercial requirements and the physical constraints <strong>of</strong> the receiving wharves. For example,<br />

ships serving Langstone Harbour on the South coast and some South Wales ports have been<br />

designed around the largest practical ship dimensions for these ports - 72m length and 5m<br />

draught resulting in a maximum cargo capacity <strong>of</strong> about 2,300 tonnes (Figure 3-7).<br />

���������������������������������������������������������������������������������������������<br />

���������������������������<br />

This contrasts with some Thames and Continental wharves which can accept vessels up to 120m<br />

long and with 8m draughts. Vessels serving these locations have, therefore, been built around<br />

these parameters and can deliver in excess <strong>of</strong> 9,000 tonnes per cargo to these locations.<br />

������ �����������������������������<br />

The smaller, older, vessels such as the “Norstone” have single propulsion engines <strong>of</strong> about 500<br />

kilowatts (kW), two small generators to supply the domestic load and an inboard mounted diesel<br />

driven dredge pump <strong>of</strong> about 350 kW.<br />

As vessels have become bigger, the main propulsion load and the powering for the dredge pump<br />

and discharge system have inevitably greatly increased so that today the largest vessels such as<br />

the “Chalemagne” and the “Sand Fulmar” have two 2,800 kW main engines, each coupled to a<br />

propeller and a large alternator <strong>of</strong> up to 2,000 kW capable <strong>of</strong> supplying all the vessel’s power<br />

requirements. During normal steaming each engine drives a propeller but when dredging or<br />

discharging, one <strong>of</strong> the main engines also supplies the power via its generator for that duty<br />

(Figure 3-8).<br />

38


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

�������������������������������������������������������������������������<br />

Even the smaller modern vessels like the “City <strong>of</strong> Cardiff” and the “City <strong>of</strong> Chichester” have had a<br />

very significant uplift in power with these 2,300 tonne vessels now having two 1,360 kW engines.<br />

The large power requirements <strong>of</strong> these relatively small ships are a result <strong>of</strong> their relatively<br />

inefficient hull design which was necessary to obtain the required annual capacity within the<br />

length and draught constraints <strong>of</strong> their trading area. They also have relatively powerful dredge<br />

pumps to minimize cycle times and also have two 400 kW thrusters that are necessary for quick<br />

and safe manoeuvring in the constricted channels they have to navigate.<br />

Production rates for dredgers depend on a number <strong>of</strong> variables including the pump power and<br />

the nature <strong>of</strong> the sediment dredged. Productivity rates can be estimated for different dredging<br />

regions based on information presented earlier in Chapter 2.2 and two cases are presented in<br />

Table 3-6 below.<br />

�������<br />

�������� ������ �����<br />

���������<br />

39<br />

�������� ��������<br />

�������������<br />

Bristol Channel 750 2 – 4 190 – 375<br />

�������� �����������<br />

�������������������<br />

Eastern English Channel 5000 5 – 7 625 – 1000<br />

�������������������������������������������������������������������������


Benchmarking Equipment, Practices and Technologies<br />

�<br />

���� ��������������<br />

������ ����������<br />

Report No: 10/J/1/06/1309/0996<br />

With the exception <strong>of</strong> the sand dredged in the Bristol Channel, the North West, and one area on<br />

the South Coast where the material tends to be in thick, clearly defined deposits, dredging is<br />

performed by trailing the draghead at the end <strong>of</strong> the dredge pipe over the seabed at speeds<br />

varying from 0.5 to 2 knots. The primary source <strong>of</strong> production is erosion <strong>of</strong> sediment around the<br />

edges <strong>of</strong> the draghead caused by water being sucked into the system. No consolidated deposits<br />

are dredged by the English industry, although some may be compacted.<br />

The UK aggregates industry uses dragheads such as the single visor draghead (Figure 3-9) for<br />

dredging fine, free-flowing material such as sand where the sediment is easily entrained in the<br />

water flow and no special arrangements are required to maximise production. The California<br />

draghead (Figure 3-10) however is required when the grain size is larger and the sediment may<br />

be slightly compacted on the sea bed. It has two independently hinged visors that allow it to<br />

better follow the seabed contours. These also provide an increased area where water can enter<br />

the system and provide the necessary erosion to entrain sediment from the sea bed, into the<br />

water flow, which is subsequently pumped up the dredge pipe. Production at these dragheads is<br />

therefore based completely on the principle <strong>of</strong> erosion and vertical hydraulic transport (Figure<br />

3-12): a flow <strong>of</strong> water is created between the movable visor <strong>of</strong> the draghead and the sea bed,<br />

and this flow erodes the seabed sediment which is subsequently pumped up the dredge pipe. In<br />

the single visor draghead this water flow enters mainly on backside <strong>of</strong> the visor, while in the<br />

California draghead it mainly enters the sides <strong>of</strong> the visor.<br />

���������������������������������������������������������������������<br />

40


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

�<br />

�������������������������������������������������������������������<br />

The sediment production as a result <strong>of</strong> hydraulic entrainment is proportional to the momentum <strong>of</strong><br />

the water flow<br />

� ��� ��� �� �� � � � � � � ������<br />

�<br />

Where � is the momentum ([kg/s]), � is a non dimensionless constant ([s/m]) depending on the<br />

particle size. � is the discharge through the suction pipe ([m 3 /s]) and � the entrance flow rate into<br />

the draghead ([m/s]).<br />

The discharge � is a given quantity for a certain hopper and depends on the power <strong>of</strong> the dredge<br />

pumps installed. The entrance flow rate � can be increased by decreasing the inflow area below<br />

the edges <strong>of</strong> the draghead. Decreasing this area does however have an influence on the<br />

pressure difference over the draghead. Inside the draghead an under pressure must be present<br />

to accelerate the entering fluid. According to Bernouilli’s equation this pressure difference �� is<br />

proportional to the square <strong>of</strong> the velocity <strong>of</strong> the entering fluid:<br />

�<br />

� �<br />

���� � ��<br />

� � � � � � � � ������<br />

There are limits to the pressure difference over the draghead. Due to the pressure difference the<br />

draghead is pushed onto the seabed and if this force is too high the draghead is more or less<br />

anchoring the hopper on the seabed. Too much propulsion power is then needed to drag the<br />

head over the seabed at sufficient speed. Propulsion can even be insufficient to move the ship<br />

41


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

forward at all. Another disadvantage <strong>of</strong> an excessive value <strong>of</strong> pressure difference is the negative<br />

effect on hydraulic transport. The pressure difference over the draghead can be seen as an<br />

hydraulic loss in the system. More pumping power is needed and it will affect the pressure at the<br />

inlet side <strong>of</strong> the dredge pump and can reduce production due to the vacuum limit (see 3.2.2<br />

below).<br />

� �������������������<br />

The natural movements <strong>of</strong> a vessel due to the action <strong>of</strong> waves and the natural undulations <strong>of</strong> the<br />

seabed mean that a mechanism to control the draghead is needed to prevent it being lifted from<br />

its position. If the draghead is lifted too high from the seabed the sediment / water mixture tends<br />

to contain too much water while if it is too low or at full weight, the draghead does not extract the<br />

optimum quantities <strong>of</strong> bottom material at the best solids/water mix.<br />

The swell compensator on board a TSHD is designed to cope with uneven bottom pr<strong>of</strong>iles and to<br />

compensate vertical vessel motion at the gantry e.g. the IHC swell compensator is designed to<br />

compensate vertical motion <strong>of</strong> up to 6 meters (IHC, 2009a). Every swell compensation system is<br />

designed specifically for the individual vessel, hardware and range <strong>of</strong> conditions in which it will<br />

work and is basically a spring device integrated in the hoisting wire system. It consists <strong>of</strong> a set <strong>of</strong><br />

pressurised air vessels connected to a hydraulic cylinder between sheaves and the<br />

compensation levels necessary are prepared especially for different suction tube arrangements<br />

on individual dredgers.<br />

������ ����������������������������������<br />

Until the mid 1980s all dredge pumps were fitted inside the vessel and the largest had a power<br />

input <strong>of</strong> about 1,000 kW, pumping through 850 mm dredge pipes. Because <strong>of</strong> the resistance to<br />

flow between the draghead and the pump the dredging depth with this arrangement is limited to<br />

about 33 metres i.e. above this length resistance to flow in the pipe makes it impossible for the<br />

pump to produce sufficient flow to entrain and transport the sediment. These large pumps were<br />

also <strong>of</strong>ten used for “pump out” discharging and were occasionally <strong>of</strong> “double walled” construction<br />

allowing the main pump casing to be made <strong>of</strong> an extremely hard, wear resistant, material which<br />

unfortunately is also very brittle. The consequence <strong>of</strong> this could be to flood the ship if it fractured<br />

– hence the need for a secondary casing (Figure 3-11).<br />

42


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

��������������������������������������������������������������������<br />

The inboard mounted pumps are capable <strong>of</strong> pumping very large quantities <strong>of</strong> water but at a<br />

relatively low bulk density <strong>of</strong> 1.25 tonnes per m 3 from >50m water<br />

depth using 700 mm dredge pipes, saving weight, space and costs. Such pumps are capable <strong>of</strong><br />

pumping sandy sediment at rates in excess <strong>of</strong> 2500 tonnes per hour and give a depth capability<br />

not achievable with inboard pumps, or alternatively give 40% higher pumping rates in shallower<br />

water.<br />

The capacity <strong>of</strong> the dredge pump is dependant on the required loading rate which is affected by a<br />

number <strong>of</strong> factors such as water depth, particle size and the degree <strong>of</strong> screening. As a general<br />

rule, however, the loading time will vary from less than 2 hours for an “all in” sand cargo in<br />

shallow water to more than 6 hours for a heavily screened gravel cargo in deep water.<br />

The majority <strong>of</strong> dredge areas <strong>of</strong>f the East Coast, along the South Coast and on the West Coast<br />

lie in water depths <strong>of</strong> 15 to 33 m and therefore just within the theoretical capability <strong>of</strong> inboard<br />

43


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

dredge pumps. The recently permitted Eastern Channel licenses have depths in some places in<br />

excess <strong>of</strong> 50 m and require vessels with high power underwater pumps. In order to dredge these<br />

deposits vessels have undergone modification with longer dredge pipes and new outboard<br />

pumps fitted.<br />

The ability to entrain sediment at the draghead is governed by the pump vacuum limit and its<br />

pumping power. The energy needed to transport the sediment / water mixture towards the<br />

hopper can be divided into the energy needed for lifting the mixture towards the surface against<br />

gravity and the pipeline flow resistance.<br />

���� �<br />

�<br />

�<br />

44<br />

�<br />

�<br />

�<br />

��<br />

������������������������������������������<br />

� �������������<br />

The pressure at the inlet side <strong>of</strong> the dredge pump can be calculated as follows. At the water<br />

surface (Figure 3-12 point 1) the atmospheric pressure ���� is present. At the seabed (Figure<br />

3-12 point 2) the pressure is equal to the atmospheric pressure plus the pressure due to the<br />

water depth : ������������ Moving from point 2 to 3 (Figure 3-12) the pressure in the flow<br />

changes due to a difference in height and flow resistance. The pressure at the inlet side <strong>of</strong> the<br />

pump is:<br />

�<br />

In which:<br />

� � � � � �� �� �� � � � � �� �� �� �� �� � � � � ������<br />

� � � �<br />

� � � � � � � ��� � � � � �<br />

� pressure [N/m 2 ]<br />

� � mixture density [kg/m 3 ]<br />

� water depth [m]


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

� � suction depth [m]<br />

� mixture velocity [m/s]<br />

� friction factor [-]<br />

The friction factor includes the pipe wall friction and additional friction due to entrance losses,<br />

bends, etc.<br />

This equation shows that the pressure at the suction side <strong>of</strong> the pump will decrease with<br />

increasing density and velocity <strong>of</strong> the mixture sucked from the seabed and with an increasing<br />

level <strong>of</strong> difference between the pump and the seabed. If the pressure at the inlet side <strong>of</strong> the<br />

pump drops below a certain level the pump is cavitating, the efficiency <strong>of</strong> the pump will decrease<br />

and production reaches a limit. This is called the vacuum limit. If the minimum pressure is<br />

denoted as ���� the following equation for the density as a function <strong>of</strong> the other quantities can be<br />

derived:<br />

�<br />

� � � ����<br />

� �<br />

�<br />

��� ��� �<br />

� � � � � � � ������<br />

� �<br />

��� ��<br />

� �<br />

In Figure 3-13 the density is plotted as a function <strong>of</strong> the mixture velocity u for two different values<br />

<strong>of</strong> the suction depth. For this example the following values are chosen:<br />

� � � � ��<br />

: 4 [m]<br />

� ��� : 15 [kPa]<br />

� ��� : 100 [kPa]<br />

� : 2.5 [-]<br />

45


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

���������������<br />

1600<br />

1500<br />

1400<br />

1300<br />

1200<br />

1100<br />

1000<br />

0 2 4 6 8 10<br />

���������������<br />

46<br />

50 m 30 m<br />

��������������������������������������������������������������������������������������������������<br />

Figure 3-13 shows the mixture density decreases with increasing dredging depths and increasing<br />

velocity. From this figure however it can not be concluded that production decreases with mixture<br />

velocity since the production is proportional to the product <strong>of</strong> density and velocity. The suction<br />

production using the above example data is shown in Figure 3-14 which shows that production<br />

reaches a maximum at a particular velocity value, and subsequently begins to decline. The<br />

maximum production is decreased at greater suction depths.<br />

All the points in the lines in this graph have one thing in common i.e. pressure at the inlet side <strong>of</strong><br />

the pump is minimal and cannot be reduced further. Therefore increasing the available power <strong>of</strong><br />

the pump drive system will not have an effect on this maximum production – it is minimum<br />

pressure (in dredging <strong>of</strong>ten called ‘vacuum’) that will limit production.


Benchmarking Equipment, Practices and Technologies<br />

�������� �������<br />

Report No: 10/J/1/06/1309/0996<br />

1600<br />

1500<br />

1400<br />

1300<br />

1200<br />

1100<br />

1000<br />

0 2 4 6 8 10<br />

���������������<br />

50 m 30 m prod 50 m prod 30 m<br />

����������������������������������������������������������������<br />

Despite the fact that the vacuum limits production there are other ways <strong>of</strong> optimizing production.<br />

The first important factor is to ensure that the dredger is working at the maximum possible<br />

vacuum production - <strong>of</strong>ten mixture velocity is too high and by lowering the velocity mixture<br />

density increases and likewise production. Many modern TSHDs are equipped with process<br />

control systems that can regulate pump speed to arrive at the optimal vacuum limit production.<br />

Another factor that has a strong effect on the production and that follows directly from the mixture<br />

density equation is the submerged depth <strong>of</strong> the pump �. Increasing the depth <strong>of</strong> the pump<br />

increases the absolute pressure in the pump and therefore the pressure drop up to minimal<br />

pressure due to friction and lifting <strong>of</strong> the mixture is much larger. This is indicated in Figure 3-15<br />

which shows the vacuum production at a dredging depth <strong>of</strong> 50 m for two values <strong>of</strong> pump depth.<br />

Increasing the pump depth from 4 to 15 m below water level has a very large effect on<br />

production. In the 4m depth case the suction pump was placed on board the vessel and its<br />

maximum depth was therefore limited by the draught <strong>of</strong> the vessel. The only way to increase the<br />

submerged depth � is to place the pump on the suction pipe which explains the increasing<br />

popularity <strong>of</strong> under water dredge pumps on recent TSHDs.<br />

47<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

�����������


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

�������� �������<br />

1800<br />

1700<br />

1600<br />

1500<br />

1400<br />

1300<br />

1200<br />

1100<br />

1000<br />

0 2 4 6 8 10<br />

���������������<br />

k = 4 m k = 15 m<br />

������������������������������������������������������������������������������������������������������<br />

� ������������������������������������������������������<br />

The specific energy required can be determined for the hydraulic transportation process from the<br />

seabed to the hopper and is again defined as the ratio between power and production. For<br />

hydraulic transportation the power needed is:<br />

� ���� � � � � � � � � ������<br />

The production is [kg/s]:<br />

���� � � �� � � � � � � � � ������<br />

The ratio is the specific energy and is equal to:<br />

� ��<br />

� � �<br />

���� � �<br />

�<br />

�<br />

48<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

�����������<br />

� � � � � � � �����<br />

�����������������������������������������������������������������������������������������������<br />

��������� ������� ��� ����� ���� �������� ��� �� ��������� ���������� ���� ���� ����� ��������� ���� ��� ����������<br />

���������������� ���� ��������� ����������� ��� �� �������������� �������� ���� ��� ����������� ������<br />

������������������������������������������������������������������������������������������������<br />

����������������������������������������������������������������<br />

� � � � ��<br />

������ ������ ������ ����<br />

�<br />

���� ��<br />

� � � � �������<br />


Benchmarking Equipment, Practices and Technologies<br />

In which:<br />

Report No: 10/J/1/06/1309/0996<br />

� water depth [m]<br />

� length <strong>of</strong> suction pipeline [m]<br />

� friction factor [-]<br />

� �� empirical constant depending on particle size [s/m]<br />

The pressure loss is proportional to the transport distance, so specific energy will linearly<br />

increase with water depth. If we normalize the specific energy with water depth (� [Jkg -1 m -1 ]), this<br />

quantity can be written as:<br />

�<br />

��<br />

����� � � � ��<br />

� � ��<br />

� � � � �� � � �<br />

� � � � ����<br />

�<br />

� � �<br />

49<br />

� � � � � �<br />

The first term is the contribution due to vertical transportation against gravity, which is<br />

independent <strong>of</strong> the process parameters. The second term is the friction resistance. This term is<br />

dependent on the operational parameters <strong>of</strong> sediment concentration and mixture velocity.<br />

��������������������������������������������������������������������<br />

The specific energy according to equation (3-9) is shown in Figure 3-16 as a function <strong>of</strong> the<br />

mixture velocity and volume concentration. This shows that a low specific energy is obtained at<br />

������


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

high concentration and that for particular concentrations the energy is lowest at an optimal<br />

velocity.<br />

������ ����������������������<br />

Common to almost all marine aggregate dredgers is the need to have the ability to screen <strong>of</strong>f<br />

either the over or undersize from the dredged material and reject it back into the water.<br />

Screening is the mechanical separation <strong>of</strong> particles by physical size, generally by passing the<br />

base material over a uniformly perforated surface. Dredged sediment is passed over a steel or<br />

plastic mesh and is a static version copied from the land based minerals industry.<br />

Screening is important as it allows the dredger to modify the sand:gravel ratio in a cargo to<br />

satisfy customer demand. It also allows vessels to economically work marginal deposits by<br />

increasing the gravel ratio through increased rejection <strong>of</strong> sand. This reduces the environmental<br />

impact <strong>of</strong> dredging by reducing the amount <strong>of</strong> licensed seabed required, but reduces productivity<br />

<strong>of</strong> the vessels, increases the fuel consumption per tonne <strong>of</strong> aggregate landed and increases the<br />

wear and tear. The screening process can also be turned around to reject large particles instead<br />

<strong>of</strong> sand.<br />

There are two types <strong>of</strong> screen – the original arrangement and still used on the “Arco Dee”, “Arco<br />

Dart”, “City <strong>of</strong> Cardiff” and “City <strong>of</strong> Chichester” is a fixed screen box at one end <strong>of</strong> the hold and a<br />

longitudinal chute which distributes the aggregate along the length <strong>of</strong> the hold (Figure 3-17). The<br />

second arrangement consists <strong>of</strong> screens mounted on towers on one side <strong>of</strong> the hold that rotate to<br />

distribute the sediment. Some vessels in the English fleet have had screen towers retro-fitted<br />

from the original static screen boxes e.g. all the Hanson <strong>Aggregate</strong> <strong>Marine</strong> Ltd ‘A’ Class ships.<br />

The screen tower is much more flexible and does not interfere with the discharge system but is<br />

more complicated and expensive to maintain (Figure 3-18). It does, however, allow more efficient<br />

filling <strong>of</strong> the hold and results in greater screening efficiencies on account <strong>of</strong> the greater screen<br />

area.<br />

50


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

�<br />

��������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������������������������<br />

51


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

The design <strong>of</strong> the equipment (especially the hopper) has an influence on the settling <strong>of</strong> sediment.<br />

The most important factors are:<br />

� Hopper Length L<br />

� Hopper Width B<br />

� Hopper Depth H<br />

� Loading system<br />

� Overflow System<br />

The length � and width � are most important regarding the geometrical quantities. The product <strong>of</strong><br />

these geometrical quantities is the hopper area which forms with the discharge � the overflow<br />

rate. The ratio <strong>of</strong> the overflow rate and the settling velocity is the most decisive factor regarding<br />

the overflow losses as will be explained in detail in Appendix A. The depth <strong>of</strong> the hopper is less<br />

important. This is counter-intuitive since it might be argued that the flow velocity in the hopper<br />

would decrease with increasing depth. This would be true if the flow distribution in the hopper<br />

was uniform, which is not so because the flow is driven by gravity (density currents) and<br />

concentrated near the bed. The mixture is distributed into the hopper by either the static screen<br />

box or by rotating loading towers. The loading system must be efficient when loading fine or<br />

coarse grained sediments.<br />

The settling velocity <strong>of</strong> fine-grained sediments is low, therefore the effect on settling efficiency is<br />

most important. The turbulence level produced by the loading system must be as low as possible<br />

and the inflowing mixture should be evenly spread over the width <strong>of</strong> the hopper and with a low<br />

flow velocity. In addition the distance between the loading structure and overflow should be as<br />

large as possible.<br />

In the case <strong>of</strong> loading coarse sediments overflow losses do not play an important role since<br />

settling velocity <strong>of</strong> this type <strong>of</strong> sediment is large. Due to this high settling velocity the sediment<br />

settles at the location where it enters the hopper and spreading the mixture over the length <strong>of</strong> the<br />

hopper becomes a more important issue to avoid large piles <strong>of</strong> sediment in the hopper with<br />

enclosed pools <strong>of</strong> water (which reduce the effective load) (Figure 3-19).<br />

����������������������������������<br />

52


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

���������������������<br />

During the loading phase overflow losses <strong>of</strong>ten increase and loading production decreases with<br />

time. A typical graph <strong>of</strong> the load in a hopper during a hopper cycle is shown in Figure 3-20. The<br />

cycle starts when the hopper is fully loaded and starts sailing to the discharge area. Next the load<br />

is discharged (red line drops to zero). The third phase in the cycle is sailing empty to the<br />

dredging area. The last phase in the cycle is the hopper loading. After reaching the overflow level<br />

loading production decreases due to increasing overflow losses.<br />

����<br />

�<br />

�<br />

� �������� �<br />

���� ��<br />

����� �<br />

�<br />

� � �� ��<br />

����� ���� � �<br />

����������<br />

53<br />

�����������<br />

����������������������������������������������<br />

� � ��������<br />

The cycle production is defined as the discharged load divided by the cycle time and minimum<br />

costs are associated with a maximal cycle production. The cycle production is equal to the<br />

tangent <strong>of</strong> the angle <strong>of</strong> a line from the origin to the point on the loading graph where the loading<br />

phase is ended. On the graph two lines are drawn – one line is drawn to the load associated with<br />

the maximum draught (the end <strong>of</strong> the loading graph), the other to a point where the line is<br />

tangential to the loading curve. In this case the angle � is larger, indicating that the cycle<br />

production is larger. It is therefore not always economical to load the hopper to maximum<br />

draught. The maximum cycle production is sometimes reached at less than maximum load<br />

depending on the shape <strong>of</strong> the loading curve. An optimum as shown in Figure 3-20 can only be<br />

determined if this line shows a curved appearance. The maximum cycle production is reached at<br />

maximum draught in case the loading curve is a straight line (which is not uncommon).<br />

������ �����������������<br />

During dredging a mixture <strong>of</strong> sediment and water is discharged in the vessel hopper with the<br />

intention being that the sediment settles in the hopper and the excess water flows overboard to<br />

maintain vessel stability. Generally a part <strong>of</strong> the incoming sediment does not settle in the hopper<br />

during dredging, but flows overboard with the excess water. Depending on the particle size<br />

distribution (PSD) <strong>of</strong> the sediment, the hopper geometry and other process parameters this


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

overflow loss can reach values up to 30-40 % <strong>of</strong> the total volume <strong>of</strong> sediment pumped into the<br />

hopper. The loading time <strong>of</strong> a dredging vessel increases with increasing overflow losses, and the<br />

finer fractions <strong>of</strong> the particle size distribution (PSD) are present in the overflow mixture and a<br />

sediment plume can be generated.<br />

The loading process can be divided into three phases.<br />

������������������������������������������������������<br />

� �<br />

� ���������������������������������������<br />

When loading starts on board <strong>of</strong> a TSHD, the hopper is generally partly filled with water, with the<br />

water level inside the hopper equal to outside. During the first phase the level <strong>of</strong> the mixture in<br />

the hopper rises until the overflow level is exceeded and overflowing starts. The flow pattern<br />

inside the hopper is strongly influenced by density currents with the density <strong>of</strong> the mixture<br />

discharged in the hopper being greater than that <strong>of</strong> the fluid inside the hopper. The mixture flows<br />

down towards the hopper bed below the inlet section due to this density difference. In the inflow<br />

section, flow velocity and the level <strong>of</strong> turbulence is high near the hopper bed and, as a result,<br />

sedimentation in this section is lower compared with downstream (in direction <strong>of</strong> overflow)<br />

locations. As a result a scour hole develops at the inflow section after some time and from the<br />

scour hole the mixture flows to the other side <strong>of</strong> the hopper as a density current. This implies that<br />

the flow is concentrated near the bed and flow velocities are much higher than for a situation<br />

where uniform flow would exist. This phenomenon is shown in Figure 3-22<br />

54


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�����������������������������������������������������������������<br />

� �����������������������������<br />

During this phase the level <strong>of</strong> the mixture in the hopper is <strong>of</strong>ten constant because the level <strong>of</strong> the<br />

overflow is <strong>of</strong>ten constant. The discharge through the overflow is equal to the inflow and a<br />

density current is present near the bed for the duration <strong>of</strong> the second phase. Sediment settles<br />

from this current and the sediment bed rises gradually. Normally a certain proportion <strong>of</strong> the<br />

particles will not settle and remain in suspension. Therefore concentration in the suspension will<br />

increase during time and likewise the concentration in the overflow. Often both concentration in<br />

the suspension in the hopper and overflow mixture will increase during time until the settled bed<br />

reaches the overflow level and maximum load is reached.<br />

� ����������������������<br />

In the final phase the top <strong>of</strong> the density current reaches the water surface. Soon after that<br />

moment a free water flow (like a river flow) develops and the flow velocity increases<br />

considerably. The increased velocity reduces sedimentation and the outflow concentration<br />

increases strongly. If loading is continued the outflow concentration becomes equal to the inflow<br />

concentration, and the total influx <strong>of</strong> sediment disappears in the overflow.<br />

Overflow losses are influenced by the characteristics <strong>of</strong> the equipment, sediment properties and<br />

operational conditions. The losses are generally defined by the term cumulative overflow loss�<br />

����� which is the ratio between the total mass <strong>of</strong> sediment flowing overboard and the total mass<br />

<strong>of</strong> sediment discharged into the hopper:<br />

��<br />

���<br />

�<br />

�<br />

�<br />

�<br />

�<br />

����<br />

� �<br />

����<br />

� �<br />

�� � � � � � ��������������������<br />

55


Benchmarking Equipment, Practices and Technologies<br />

Where:<br />

Report No: 10/J/1/06/1309/0996<br />

�� Inflow discharge [m 3 /s]<br />

�� Overflow discharge [m 3 /s]<br />

�� Volumetric inflow concentration [-]<br />

�� Volumetric overflow concentration [-]<br />

� Loading time [s]<br />

Another term <strong>of</strong>ten encountered is loading efficiency � , defined as:<br />

� �������� � � � � � � � ����������������������<br />

During loading the sediment particles should settle in the hopper, the settling velocity increases<br />

with particle diameter and Figure 3-23 shows the settling velocity as a function <strong>of</strong> the particle<br />

diameter. The line ‘CD iteration’ results from direct solving <strong>of</strong> the equilibrium equation <strong>of</strong> a<br />

submerged particle. Natural sediment does not consist <strong>of</strong> one particular particle size, but <strong>of</strong><br />

different sizes and the particle size distribution is therefore important. The shape <strong>of</strong> particles also<br />

influences the settling velocity with flattened particles, like shells, having a large surface area<br />

relative to their volume. This results in more resistance and therefore settling velocity will be<br />

lower than a spherical particle <strong>of</strong> the same size.<br />

������������������������������������������������������������������<br />

56


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Unlike screening, overspill is a process necessary to maintain the vessel stability and the<br />

resulting discharges result in a silt plume developing around the vessel (Figure 3-24). On many<br />

English TSHDs the overflow is over the side <strong>of</strong> the hopper. Although this is a simple structure it<br />

has some disadvantages. The level in the hopper is less adjustable and the overflow location<br />

relative to the loading position is more difficult to adjust and <strong>of</strong>ten too small leading to large<br />

overflow losses. On three <strong>of</strong> the more recently built vessels, however, (“City <strong>of</strong> Cardiff”, “City <strong>of</strong><br />

Chichester” (Figure 3-25) and “Charlemagne”) both the screening system rejects and the hold<br />

overflows pass out through the bottom <strong>of</strong> the vessel. This encourages a density plume to form,<br />

which takes more <strong>of</strong> the returned sediment back to the seabed, rather than allowing it to disperse<br />

passively – thus, potentially, reducing indirect impacts.<br />

������������������������������������������������������������������������������������������<br />

�<br />

���������������������������������������������������������������������������<br />

����������������������������������������������������������������������������<br />

�������������������������������������������������������������������<br />

57


Benchmarking Equipment, Practices and Technologies<br />

������ ��������������������<br />

Report No: 10/J/1/06/1309/0996<br />

Only one vessel in the UK fleet is not fitted with some form <strong>of</strong> self discharging system (the<br />

“Donald Redford”) and this is discharged by shore cranes (Figure 3-26). All the other vessels are<br />

fitted with “dry” discharge systems with 35% <strong>of</strong> these also fitted with wet “pump out” systems.<br />

These have a high power demand and require the bottom <strong>of</strong> the hold to be tapered resulting in a<br />

loss <strong>of</strong> volume and raising <strong>of</strong> the cargo centre <strong>of</strong> gravity but this matches well with a bucket<br />

wheel dry discharge system which requires a similar shape. Whilst the wet system is now little<br />

used for discharging to wharves, mainly due to the problems <strong>of</strong> dealing with the resultant silty<br />

water, it is regularly used for beach replenishment work.<br />

�<br />

����������������������������������������������������������������������������������������������<br />

The first dry discharge systems were drag scrapers based on those used in land based mineral<br />

operations. They proved to be relatively simple and robust but result in high wear rates on the<br />

ships’ structure, have a high energy demand and are labour intensive as they are difficult to<br />

automate. Notwithstanding this, 8 systems remain in use today and a typical system has an<br />

average discharge rate <strong>of</strong> 1,800 tonnes per hour (Figure 3-27).<br />

58


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

��������������������������������������������������������������������������������������������<br />

Subsequently, the bucket wheel system was developed which was also adapted from land-based<br />

systems. It gives an almost constant discharge rate <strong>of</strong> 2,000 tonnes per hour and has a relatively<br />

low energy demand. The required shaping <strong>of</strong> the hold bottom is not a disadvantage when<br />

combined with a pump out system and the system causes little structural damage to the ship but<br />

it is labour intensive as it has not been successfully automated. There are currently 9 <strong>of</strong> these<br />

systems in service (Figure 3-28).<br />

�<br />

�������������������������������������������������������������������������������������<br />

59


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

One older vessel, the “Norstone”, has been retr<strong>of</strong>itted with an excavator based system<br />

discharging to a receiving hopper and a short ship to shore boom conveyor. This has proved to<br />

be a cost effective modification by decreasing discharge time, reducing damage to the vessel<br />

and removing the reliance on shore-based facilities and labour.<br />

The most recent development has been the discharger with a grab the full width <strong>of</strong> the hold and a<br />

shuttling receiving hopper. Introduced in its original form in 1990 on two 1,300 tonnes vessels<br />

(the “Arco Dee” and the “Arco Dart”), it was refined into a fully automated 1,200 tonne per hour<br />

system on two 2,300 tonnes vessels in 1999 (the “City <strong>of</strong> Cardiff” and the “City <strong>of</strong> Chichester”<br />

(Figure 3-29)). In 2002 a 2,500 tonnes per hour system was installed on the “Charlemagne” with<br />

another order placed for a sister vessel in Q2 2008. This system is easily automated, has little<br />

impact on the vessels’ structure and allows a variety <strong>of</strong> ship to shore conveyor arrangements.<br />

�<br />

�������������������������������������������������������������������������������������<br />

The ship to shore conveyor arrangements on most vessels are generally designed to match the<br />

wharf receiving facilities they expect to operate into and vary between a single 26m conveyor<br />

forward (Figure 3-30), to a single 15m conveyor aft to two 20m conveyors aft.<br />

60


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

�������������������������������������������������������������������<br />

���� ��������������������������������������������������������<br />

������ �������������������<br />

� ������������<br />

Positioning on all English aggregate dredgers utilizes Differential GPS, which is a satellite based<br />

positioning system. Differential GPS is the use <strong>of</strong> differential correction data for the satellites in<br />

view, to remove the errors in the position due to atmospheric effects and satellite clock errors.<br />

The correction data is generated by base stations, which are sited at accurately known locations.<br />

The UK base stations are operated by Trinity House and comprise 14 ground-based reference<br />

stations providing transmissions with coverage <strong>of</strong> at least 50 nautical miles around the coasts <strong>of</strong><br />

the United Kingdom and Republic <strong>of</strong> Ireland. It is an open system - available to all mariners - and<br />

is financed from light dues charged on commercial shipping and other income paid into the<br />

General Lighthouse Fund. The correction data is formatted in the industry standard RTCM SC-<br />

104 and is available on a continuous basis. Positional accuracy within 1 or 2 metres is achieved.<br />

� �������������������������������������<br />

All vessels dredging English licence areas are fitted with an AIS (Automatic Identification System)<br />

which the International Maritime Organization (IMO) requires for all vessels over 299GT. The AIS<br />

transponder transmits the vessel name, position, speed and course is intended to help ships<br />

avoid collisions, as well as assisting port authorities to better control sea traffic.<br />

<strong>Dredging</strong> management and navigation within <strong>Dredging</strong> Licence Areas on board all English<br />

dredgers is controlled through a Microplot system which allows the boundaries <strong>of</strong> dredging<br />

licence areas, active areas and dredging lanes to be displayed electronically (Figure 3-31).<br />

61


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

��������������������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������������������������<br />

��������������������������<br />

The Microplot is linked to the ship’s DGPS system and the ship position can be displayed on<br />

Microplot, in real time. It is also possible to include <strong>of</strong>fsets to the positional data from the DGPS<br />

so that the position <strong>of</strong> the draghead is displayed, rather than the ship position, which allows the<br />

vessel’s Master to accurately dredge within previously defined boundaries. Known targets such<br />

as wrecks or areas <strong>of</strong> contamination can be added as symbology allowing these to be avoided.<br />

The vessel Master is also able to electronically note any new areas <strong>of</strong> contamination or target<br />

contacts and inform the dredging fleet via the shore-based <strong>of</strong>fice.<br />

� �����������������������������<br />

All vessels dredging on UK licence areas are required, by The Crown Estate, to have onboard an<br />

Electronic Monitoring System (EMS) for recording vessel activity. The EMS consists <strong>of</strong> a secure<br />

computer on the ship’s bridge, which logs data from sensors on the dredge gear and positional<br />

data from the GPS. The EMS automatically records the date, time and position <strong>of</strong> all dredging<br />

activities and provides a secure read out <strong>of</strong> the location <strong>of</strong> the dredging vessel every thirty<br />

seconds when the dredging equipment is deployed, and every thirty minutes otherwise.<br />

These dredging data are supplied monthly to The Crown Estate for analysis and approximately<br />

500,000 data points are analysed for irregularities (i.e. time gaps in the data or indications <strong>of</strong> out<br />

<strong>of</strong> area dredging) every month. During 2008 approximately 20,000 hours <strong>of</strong> dredging were<br />

recorded and approximately 0.03% <strong>of</strong> this dredging was out <strong>of</strong> area/zone (The Crown Estate,<br />

2009). Since the introduction <strong>of</strong> EMS over one and a quarter million kilometres <strong>of</strong> dredging tracks<br />

62


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

have been analysed. In addition to analyzing dredging performance the EMS data can be used to<br />

measure the annual extent <strong>of</strong> dredging activities, dredging intensities within individual licence<br />

areas (Figure 3-32) and the cumulative footprint (i.e. the total extent <strong>of</strong> dredging activity) over a<br />

period <strong>of</strong> time.<br />

���������������������������������������������������������������������������������������������������<br />

BMAPA and The Crown Estate are committed to transparency and have been reporting the area<br />

licensed for dredging under the ‘Area Involved’ initiative. This included a commitment to publish<br />

an annual report detailing the extent <strong>of</strong> dredging within regional licensed areas using analysis <strong>of</strong><br />

EMS data. Every year since 1998, therefore, they have jointly produced the Area Involved report.<br />

In addition a five-year review report summarizing the period 1998-2002 was published in 2005,<br />

while a ten-year review report (summarizing the period 1998-2007) was published in 2009.<br />

� ����������������������������������������������<br />

Over the last decade vessels <strong>of</strong> the English dredging fleet have altered on-board dredging<br />

practices in order to reduce the environmental effects <strong>of</strong> the dredging. Examples include limiting<br />

the number <strong>of</strong> vessels able to dredge in a region at any one time and loading as efficiently as<br />

possible, in order to decrease loading time and hence reduce the time the environment is directly<br />

63


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

impacted and that noise is generated. Table 3-7 shows the total hours dredged by vessels in the<br />

fleet for the period 2006-2008.<br />

� ����� ����� �����<br />

Hours dredged 28,686 26,340 22,985<br />

���������������������������������������������������������������<br />

Reduced (or prohibited) screening has been instigated when dredging some license areas in<br />

order to minimize the effects <strong>of</strong> the plume. Table 3-8 combines data on total marine aggregate<br />

production and hours dredged from to calculate the tonnages landed per hour dredged.<br />

� ����� ����� ����� ����������<br />

<strong>Marine</strong><br />

aggregate<br />

production<br />

64<br />

����������<br />

20.29mt 20.64mt 19.75mt -4.31%<br />

Hours dredged 28,686 26,340 22,985 -12.74<br />

Tonnes landed<br />

per hour dredged<br />

707.41tph 783.57tph 859.12tph +9.64<br />

������������������������������������������������������������������<br />

A reduction in dredged hours <strong>of</strong> nearly 13 percent at a time when total production dropped by just<br />

over 4 percent suggests less intensive screening, and as a result the tonnage landed per hour<br />

dredged increased by over 9 percent (BMAPA, 2009). BMAPA members are also committed to<br />

reviewing all production licence areas over a rolling five year period, and to surrender areas no<br />

longer containing useful resources <strong>of</strong> sand and gravel.<br />

� ���������������������������<br />

Key to mitigating the impacts <strong>of</strong> dredging on other sea users is an effective mechanism for liaison<br />

e.g. the success <strong>of</strong> the majority <strong>of</strong> mitigation measures related to the commercial fishing industry<br />

in the UK relies heavily on communication between the two industries (Royal Haskoning, 2004).<br />

As part <strong>of</strong> the fisheries liaison process BMAPA, in conjunction with The Crown Estate, produces<br />

twice-yearly active dredge area charts which define the areas where dredging will take place.<br />

These are widely distributed in association with local marine and Fisheries Agency <strong>of</strong>ficers and<br />

Active Dredge Zones are not altered until advance notices to local groups have been made. Biannual<br />

fisheries liaison meetings also take place on the south and east coast.<br />

An area where potential conflicts with other users may become more apparent is in the proximity<br />

<strong>of</strong> dredging licences to regions where <strong>of</strong>fshore marine renewable energy interests will be<br />

developed. There is a long-term issue with the routing <strong>of</strong> cables from developing wind farms to<br />

the coast, but there is also a potential for operational impacts on dredging vessels (BMAPA,<br />

2009). <strong>Dredging</strong> vessels require access to and from their licence areas and the ability to navigate


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

within the area while dredging. To mitigate against operational impacts BMAPA has generated a<br />

series <strong>of</strong> charts, available to <strong>of</strong>fshore wind developers, showing the extent <strong>of</strong> dredger transit<br />

routes between licence areas and ports relative to round 1, 2 and 3 <strong>of</strong>fshore wind sites.<br />

� �����������������������������������������������������<br />

To reduce potential archaeological impacts dredging companies observe a code <strong>of</strong> practice on<br />

marine aggregate dredging and the historic environment developed jointly by BMAPA and<br />

English Heritage. This protocol means that finds <strong>of</strong> archaeological interest are reported to the<br />

Secretary <strong>of</strong> State, the Crown Estate Commissioners, DCMS, English Heritage and to relevant<br />

County Archaeologists. In the event that features <strong>of</strong> archaeological interest are encountered in<br />

the course <strong>of</strong> dredging, companies comply with the Merchant Shipping Act 1995 in respect <strong>of</strong><br />

reporting and ownership <strong>of</strong> wrecks including notification <strong>of</strong> the Receiver <strong>of</strong> Wrecks. <strong>Dredging</strong><br />

exclusion zones are implemented around features <strong>of</strong> acknowledged archaeological importance<br />

identified from the assessment <strong>of</strong> existing geophysical data and these may be reviewed in light <strong>of</strong><br />

archaeological assessment <strong>of</strong> new data. If any previously unreported wrecks become apparent<br />

within the boundaries <strong>of</strong> the dredging permission area precautionary exclusion zones, defined in<br />

consultation with independent marine archaeological consultants and English Heritage, are<br />

instituted around them.<br />

������ �����������<br />

The environmental impacts <strong>of</strong> English dredging may potentially be reduced by development <strong>of</strong><br />

practices and mitigation techniques. These techniques can involve spatial and temporal<br />

mitigation and application <strong>of</strong> suitable dredging practices.<br />

� �������������������<br />

Potential impacts <strong>of</strong> dredging can be mitigated by reducing the total area <strong>of</strong> seabed dredged i.e.<br />

through spatial mitigation. At a high level, all licence areas managed by English aggregate<br />

dredging companies are reviewed on an annual basis and unproductive zones are relinquished<br />

to minimise the area under licence. Thus the total area <strong>of</strong> seabed licensed decreased by 387.09<br />

km 2 between 1998 and 2007, with the greatest reductions <strong>of</strong> area occurring within 12 nm <strong>of</strong> the<br />

coast. This net change was due to a relinquishment <strong>of</strong> 749.84 km 2 combined with 362.44 km 2 <strong>of</strong><br />

new Licenses (The Crown Estate, 2008). The industry reports the extent <strong>of</strong> licensed and dredged<br />

area through the ongoing ‘Area Involved Initiative’ which started in 1999. Data from 2007<br />

represented the 10 th anniversary <strong>of</strong> the initiative and an additional review document was<br />

produced to illustrate trends in the licence and dredged areas since 1998. The annual documents<br />

also include data on the area <strong>of</strong> cumulative dredge footprint (BMAPA, 2009).<br />

<strong>Dredging</strong> within individual licence areas is also controlled through voluntary zoning to minimise<br />

the areas <strong>of</strong> seabed actually being worked (spatial footprint directly impacted). Thus Active<br />

Dredge Zones (ADZs) are defined within a License Area, and these ADZs are themselves<br />

subdivided into dredging lanes. Zoning is a requirement <strong>of</strong> the <strong>Marine</strong> Mineral Guidance Note 1:<br />

Guidance on the Extraction by <strong>Dredging</strong> <strong>of</strong> Sand, Gravel and Other Minerals from the English<br />

Seabed (also known as MMG1; Office <strong>of</strong> the Deputy Prime Minister, 2002). MMG1 describes the<br />

policies and procedures for marine minerals dredging in English waters and provides guidance to<br />

65


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

regulators, their scientific advisors, industry and wider stakeholders as to how marine minerals<br />

extraction should be undertaken so that it is consistent with the principles <strong>of</strong> sustainable<br />

development. The zoning guidance states that “the intention is to limit the extent <strong>of</strong> groups <strong>of</strong><br />

active dredge Areas whilst at the same time minimise the impact <strong>of</strong> multiple dredging activities<br />

where necessary” (ODPM, 2002). Zoning plans and ADZs are agreed with the regulators.<br />

Zoning an area potentially mitigates the effects <strong>of</strong> the dredging by reducing the area <strong>of</strong> direct<br />

removal <strong>of</strong> biomass. It also allows the majority <strong>of</strong> a License Area to be undredged, or undergoing<br />

recolonisation, at any particular time. MMG1 also requires an operator to work areas to economic<br />

exhaustion – so allowing uninterrupted recovery once extraction has ended (ODPM, 2002). Initial<br />

recolonisation <strong>of</strong> an area may be more rapid than if a larger area was being dredged. <strong>Dredging</strong> in<br />

lanes means that strips <strong>of</strong> undredged habitat may be left between adjacent lanes and this may<br />

mitigate the effects <strong>of</strong> the dredging since colonising species may have shorter distances to travel<br />

and recolonisation rates may increase. The rate <strong>of</strong> recovery <strong>of</strong> species diversity depends on the<br />

complexity <strong>of</strong> the fauna and the controls on recruitment <strong>of</strong> larvae and settlement. Longest<br />

recovery rates would be for slow-growing species that do not have planktonic larvae and rates<br />

are also long for sites that are repeatedly dredged. Minimizing the area dredged will therefore<br />

minimize these effects.<br />

Another means <strong>of</strong> mitigating the effect <strong>of</strong> dredging on marine habitats, and a standard condition<br />

on all English licenses and dredging permissions, is to leave the seabed post-dredging in a<br />

similar state to its pre-dredged condition e.g. if the seabed was a sandy gravel pre-dredging<br />

attempts will be made to leave the seabed as a sandy gravel following cessation <strong>of</strong> dredging. In<br />

addition sediments should not be dredged completely but 0.5 m thickness <strong>of</strong> resource sediment<br />

should be left after cessation <strong>of</strong> dredging.<br />

The English dredging industry makes efforts to maintain seabed condition through targeted<br />

screening and dredging and regular monitoring. Newell and Seiderer (2003) report that many<br />

species do not re-colonize regularly, and even though the seabed may remain in a similar<br />

condition to its pre-dredged condition there may still be a significant interval before all the species<br />

components are present in the community.<br />

Zoning reduces the area over which screened sediments are returned to the seabed and zones<br />

are positioned, where possible, parallel with the directions <strong>of</strong> tidal currents to ensure that bedload<br />

and plume transport is rapidly incorporated into any regional sediment transport pathway. Zoning<br />

to target sediments with low silt and clay contents also mitigates against the potential effects <strong>of</strong><br />

plumes by reducing the amount <strong>of</strong> fine sediment introduced into the water column. This mitigation<br />

technique, however, can only be properly implemented following commencement <strong>of</strong> dredging,<br />

when further knowledge about the location <strong>of</strong> fine sediment has been gained. Zoning <strong>of</strong> English<br />

License Areas means that <strong>of</strong> the approximately 1300 km 2 <strong>of</strong> licensed seabed between 222.6 km 2<br />

(1998) and 134.5 km 2 (2004) was actually dredged in any one year (The Crown Estate, 2009).<br />

<strong>Dredging</strong> also has the potential to create conflict with other potential users <strong>of</strong> the sea area<br />

particularly fishermen. Such conflicts may result in diminished access to fishing areas and a<br />

reduced catch. Spatial mitigation actions to reduce effects on fisheries can include exclusion<br />

zones to avoid important fishery sites. Other mitigation actions also include activities to facilitate<br />

66


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

communication and cooperation among potentially conflicting users such as appointing fishery<br />

industry liaisons to facilitate communications and providing advance notice to other users <strong>of</strong><br />

changing dredging.<br />

� ��������������������<br />

The potential effects <strong>of</strong> dredging may have an increased environmental impact during particular<br />

times <strong>of</strong> the year e.g. during spawning periods. The English aggregate dredging industry<br />

mitigates against these effects by adopting temporal mitigation measures i.e. through minimizing<br />

or avoiding screening (or prohibiting dredging entirely) on certain license areas during particular<br />

environmental windows. Temporal mitigation may take place during certain phases <strong>of</strong> flood or<br />

ebb; during the Spring-Neap tidal cycle, or over particular seasons. As an example, mitigation <strong>of</strong><br />

the potential effects <strong>of</strong> benthic boundary layer plumes on breeding areas for crab has been<br />

achieved by dredging only when the tidal stream transports sediments away from sensitive<br />

areas.<br />

� ����������������������������<br />

The marine aggregate industry in the UK has significant sums <strong>of</strong> money in increasing knowledge<br />

<strong>of</strong> the marine environment and in enhancing the understanding <strong>of</strong> the impacts <strong>of</strong> aggregate<br />

dredging. It has done this through internal research projects, the Environmental Impact<br />

Assessment process and most significantly through the <strong>Aggregate</strong> Levy Sustainability Fund<br />

(ALSF). In 2002 the Government imposed a levy on all primary aggregates production (including<br />

marine aggregates) to reflect the environmental costs <strong>of</strong> winning these materials. A proportion <strong>of</strong><br />

the revenue generated was used to provide a source <strong>of</strong> funding for research aimed at minimising<br />

the effects <strong>of</strong> aggregate production. This fund, delivered through <strong>Defra</strong>, is known as the<br />

<strong>Aggregate</strong> Levy Sustainability Fund (ALSF). The <strong>Marine</strong> ALSF (MALSF) is currently administered<br />

by the Centre for Environment, Fisheries and Aquaculture Science (<strong>Cefas</strong>) through the <strong>Marine</strong><br />

Environmental Protection Fund (MEPF) and English Heritage.<br />

The objectives <strong>of</strong> the ALSF in the marine environment are to reduce the environmental footprints<br />

<strong>of</strong> marine extraction. Priority areas for funding have been identified in line with five main aims.<br />

� To develop and use seabed mapping techniques to improve the evidence base <strong>of</strong><br />

nature, distribution and sensitivity <strong>of</strong> marine environmental and archaeological<br />

resources relevant to marine aggregate activities.<br />

� To increase understanding <strong>of</strong> the effects <strong>of</strong> aggregate dredging activities, including<br />

noise, and their significance.<br />

� To develop monitoring, mitigation and management techniques, underpinned by<br />

scientific research.<br />

� To research and understand the socio-economic issues associated with aggregate<br />

dredging activities.<br />

� To promote co-ordination and establishment <strong>of</strong> sustainable archives for the<br />

dissemination <strong>of</strong> research related to these aims to a wide range <strong>of</strong> stakeholders.<br />

67


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

The <strong>Marine</strong> ALSF supports a wide range <strong>of</strong> projects that have addressed a range <strong>of</strong> scientific and<br />

socio-economic fields including exploring the ecology, geology and heritage <strong>of</strong> the seabed<br />

around the UK; environmental protection and impact assessment, protection <strong>of</strong> the historic<br />

environment, recoverability, mitigation, technological advances and the raising <strong>of</strong> public<br />

awareness. Since 2002, over £20 million has been invested in projects increasing the knowledge<br />

and understanding <strong>of</strong> marine environmental resources through applied, science led research.<br />

This directly supports marine planning and decision making through the provision <strong>of</strong> robust state<strong>of</strong>-the-art<br />

evidence. The MALSF has commissioned projects that have addressed a range <strong>of</strong><br />

scientific and socio-economic themes.<br />

The MALSF seeks to apply the “collect once, use many times” principle to the data collected and<br />

recognizes that data collected for a project specific purpose may have wider applicability, if<br />

accessed by other scientific disciplines and be <strong>of</strong> use beyond the requirements <strong>of</strong> any single<br />

project. Accordingly, <strong>Cefas</strong> (2008) produced a summary report for the MALSF which examined<br />

current practices in MALSF research and provided recommendations on how they may be<br />

adapted to maximise the value <strong>of</strong> the data collected in future, for the benefit <strong>of</strong> all stakeholders.<br />

For transparent planning and management <strong>of</strong> the marine environment, the information and<br />

knowledge gained should be readily available to all ‘stakeholders’ – from Government through to<br />

the general public. The MALSF aims to make the results <strong>of</strong> the projects accessible through a<br />

range <strong>of</strong> communication and dissemination materials and activities.<br />

The increasing amount <strong>of</strong> knowledge gained as a result <strong>of</strong> the MALSF has allowed development<br />

to commence <strong>of</strong> a <strong>Marine</strong> <strong>Aggregate</strong> Extraction Risk Assessment (MARA) Framework to explore<br />

the potential <strong>of</strong> an approach for assessing hazard probability and consequences for receptors<br />

and to present these in the context <strong>of</strong> the overall risk arising from dredging (HR Wallingford,<br />

2007).<br />

������ ���������������������<br />

The English marine aggregate industry is licensed commercially by The Crown Estate, which<br />

owns the seabed to the 12-mile territorial limit and holds the non-energy mineral rights out to 200<br />

miles as part <strong>of</strong> the hereditary possessions <strong>of</strong> the Sovereign. Under The Crown Estate Act 1961,<br />

The Crown Estate Commissioners have a duty to maintain and enhance the value <strong>of</strong> the estate’s<br />

assets and to secure revenue from them. It receives a royalty for every tonne <strong>of</strong> aggregate<br />

extracted from licensed areas. Companies seeking to dredge are obliged to obtain authorization<br />

from the Government and a license from The Crown Estate, before they can legally operate<br />

(Gubbay, 2005).<br />

Before dredging can take place on an English licence area the dredging company must obtain<br />

consent, termed a dredging permission. The dredging permission process in England is<br />

administered by <strong>Defra</strong>’s <strong>Marine</strong> and Fisheries Agency. In 2007, the Environmental Impact<br />

Assessment and Natural Habitats (Extraction <strong>of</strong> Minerals by <strong>Marine</strong> <strong>Dredging</strong>) (England and<br />

Northern Ireland) Regulations 2007 (the <strong>Marine</strong> Minerals Regulations) were introduced which<br />

provided a statutory framework for the environmental impact assessments required for proposed<br />

68


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

marine minerals extraction. The <strong>Marine</strong> & Fisheries Agency carry out the licensing and<br />

enforcement duties resulting from the introduction <strong>of</strong> the regulations. The procedure to follow is<br />

summarized in Figure 3-33 below.<br />

The English regulatory process is transparent, puts the environment first, and involves significant<br />

consultation to issues that need addressing in the environmental impact assessment. The<br />

process is supported by Government Guidance – primarily <strong>Marine</strong> Mineral Guidance Note 1:<br />

Guidance on the Extraction by <strong>Dredging</strong> <strong>of</strong> Sand, Gravel and Other Minerals from the English<br />

Seabed (MMG1) (ODPM, 2002) which describes the policies and procedures for marine minerals<br />

dredging in English waters and provides guidance to regulators, scientific advisors, industry and<br />

wider stakeholders as to how marine minerals extraction should be undertaken so that it is<br />

consistent with the principles <strong>of</strong> sustainable development.<br />

An environmental statement and supporting studies are widely circulated to stakeholders, with<br />

the applicant then tasked with overcoming concerns. Unless the applicant can demonstrate<br />

that the environmental impacts are acceptable at the end <strong>of</strong> the process, then the administering<br />

government department will not award a dredging permission. Consent decisions are<br />

accompanied by a schedule <strong>of</strong> legally enforceable conditions that form an integral part <strong>of</strong><br />

any Crown Estate production licence to dredge and must be adhered to by the licensee<br />

(BMAPA, 2009).<br />

69


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

���������������������<br />

Applicant supplies<br />

necessary information<br />

to MFA<br />

��������<br />

������������<br />

MFA consult widely<br />

statutory and other<br />

stakeholders<br />

EIA or AA<br />

required?<br />

Applicant applies to MFA for<br />

pre-application advice from<br />

<strong>Cefas</strong> (recommended)<br />

Applicant to pay fee to MFA<br />

MFA to instruct <strong>Cefas</strong> to<br />

engage with applicant and<br />

can supply formal scoping<br />

opinion if required<br />

Applicant to complete fully<br />

worked up ES<br />

Applicant provides application<br />

form, fees, ES and all other<br />

necessary supporting<br />

information to MFA<br />

MFA make decision<br />

based on consultation<br />

responses<br />

�<br />

�<br />

������������������������������������������������������������������������������<br />

�������������������������������������<br />

70<br />

Screening determination<br />

sent to applicant and<br />

published<br />

Decision sent to<br />

applicant and<br />

published


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

71


Benchmarking Equipment, Practices and Technologies<br />

��� ��������� ������� ���� ������ �� �������������� ������������� ���� ����������<br />

�������������������������������<br />

���� �����������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

This Section <strong>of</strong> the report will describe the characteristics <strong>of</strong> the current world fleet <strong>of</strong> dredgers<br />

and compare these with the situation in the English fleet. In order to make comparison easier it<br />

will mirror the structure <strong>of</strong> Chapter 3.1 and focus on:<br />

� Vessel age;<br />

� Design;<br />

� Crew;<br />

� Vessel size; and<br />

������ �������������<br />

� Power and Productivity Rates<br />

The most recent data available estimates that there are in excess <strong>of</strong> 1300 vessels in the current<br />

world dredging fleet (Clarkson Research Services Ltd, 2009). These vessels are divided into a<br />

number <strong>of</strong> different classes <strong>of</strong> dredging vessel as shown in below.<br />

������������� �������<br />

Backhoe / Dipper / Grab Dredger 243<br />

Barge Unloading Dredger 19<br />

Bucket Ladder Dredger 61<br />

Cutter Suction / Bucket Wheel Dredger 420<br />

Special Equipment Dredger 32<br />

Suction Dredger 40<br />

Trailer Suction Hopper Dredger 494<br />

TOTAL 1309<br />

�����������������������������������������������������������������������������<br />

Table 4-1 shows that TSHDs make up 38% <strong>of</strong> the world’s dredging fleet, and the remainder <strong>of</strong><br />

Chapter 4.1 will concentrate on these 494 vessels.<br />

������ �����������<br />

The age <strong>of</strong> TSHDs in the world fleet varies greatly. The oldest TSHD in the world fleet is the<br />

TSHD “Westerschelde”, owned by Dutch <strong>Dredging</strong>, which was built in 1935 and is 74 years old.<br />

The mean age <strong>of</strong> vessels in the world TSHD fleet is 24 years. The data are skewed, however, by<br />

the number <strong>of</strong> older vessels working in the developing world, since vessels that have reached the<br />

end <strong>of</strong> their working life in Europe or North America are <strong>of</strong>ten sold to operators in emerging<br />

72


Benchmarking Equipment, Practices and Technologies<br />

������ �������<br />

Report No: 10/J/1/06/1309/0996<br />

markets to continue working. A better comparison with the English dredging fleet would therefore<br />

be to examine the fleets <strong>of</strong> the major European dredging companies i.e. Van Oord, Jan de Nul<br />

N.V., <strong>Dredging</strong> International and Boskalis. The oldest vessel is Van Oord’s vessel “HAM 350”<br />

which is 35 years old, and the mean age <strong>of</strong> the vessels in the fleets <strong>of</strong> these four companies is<br />

11.1 years. This is considerably younger than the mean age <strong>of</strong> English TSHDs (21.9 years). 6<br />

further vessels are due to be launched in 2010 – 2011 which is likely to reduce the mean age<br />

further.<br />

Data on configurations are available for 413 <strong>of</strong> the world fleet (Bert Visser’s Directory <strong>of</strong><br />

Dredgers, 2009) and are summarized in Figure 4-1 below.<br />

��������������������������������������������������������������<br />

It can be seen that for those vessels where data are available the majority have an aft design<br />

configuration, while approximately 20% <strong>of</strong> the vessels have a forward configuration. There is also<br />

a small percentage (2% - 10 vessels) that have a midships bridge and accommodation block.<br />

Presumably this is to reduce movement and increase crew comfort but places an obvious<br />

restriction on the capacity <strong>of</strong> the dredge hopper.<br />

Modern ship design can also play a part in altering the sailing resistance (and hence fuel use) <strong>of</strong><br />

a ship. The resistance to a ship’s sailing depends on its speed, the way water flows around the<br />

hull and the waves the ship makes as it moves. Different hull designs can significantly alter the<br />

sailing resistance with the bow design being <strong>of</strong> particular importance. Many modern TSHDs have<br />

a "bulb" (i.e. artificial nose) mounted on the hull at the waterline which alters the flow <strong>of</strong> water<br />

around the hull, and resistance is reduced because <strong>of</strong> the reduction <strong>of</strong> induced waves. Fuel<br />

savings <strong>of</strong> up to 10% can be made – but this is only optimized for one draught. Figure 4-2 shows<br />

two TSHDs – one with a conventional bow and one with a bulb. It can be seen that the waves<br />

73


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

induced by the ship with a bulb are significantly smaller. The design <strong>of</strong> the ship’s stern can also<br />

have an effect on sailing efficiency with carefully designed sterns resulting in an ideal approach<br />

<strong>of</strong> flow to the propellers resulting in less vibration and higher propulsion efficiency.<br />

�<br />

�����������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������<br />

Ship propellers are reasonably efficient in transforming mechanical engine power into thrust<br />

power with efficiencies mostly in the range <strong>of</strong> around 70%. Propeller efficiency can be limited by<br />

conditions imposed by the hull and the ports the vessel serves, for example propeller diameter is<br />

restricted by the operational draught restrictions. Other modern techniques for reducing hull drag<br />

include anti-fouling paints and gel coats; high quality hull welding, retro-fitted propeller pods etc.<br />

A detailed study <strong>of</strong> vessel design is beyond the scope <strong>of</strong> this project, however a recently<br />

commissioned ALSF study (MEPF REF 09/P133) will review vessel design and identify key<br />

issues including optimum operation (e.g. optimum trim, optimum power setting in restricted<br />

waters), maintenance (engine room tuning, polishing <strong>of</strong> propellers) and refits (application <strong>of</strong><br />

Energy Savings Devices, improved propeller design). It will also quantify potential reductions in<br />

environmental impacts for any future new build vessels.<br />

74


Benchmarking Equipment, Practices and Technologies<br />

������ �����<br />

������ ������<br />

Report No: 10/J/1/06/1309/0996<br />

Data on crew numbers are more restricted than vessel age data and the accuracy <strong>of</strong> the data are<br />

unknown, however the mean crew on TSHDs worldwide is 27. The largest reported crew<br />

numbers are 120 crew members on the US Army Corps <strong>of</strong> Engineers operated TSHD “Esayons”,<br />

and 92 crew on the Iraqi Ports operated TSHDs “Al Najaf” and “Alkhalij Alarabi”. These contrast<br />

with the 3 crew members <strong>of</strong> the TSHD “Toste” operated by the Danish Coastal Authority, and<br />

which is used for beach recharge operations. Again these numbers are skewed by practices on<br />

vessels worldwide being radically different from those in the English industry.<br />

A trend within European dredging companies, particularly DEME, is the development <strong>of</strong> oneman-operated<br />

dredgers which have been claimed to be safer and less prone to<br />

misunderstandings between operators (IHC, 2007). Typically dredging vessels, including all<br />

English dredging vessels have had at least two crew members on duty on the bridge - one<br />

handles the ship, navigation and lookout while the other controls the dredging equipment.<br />

The most recent DEME vessels including the TSHDs "Marieke", "Brabo" and "Breydel” have<br />

been built so that a single crew member can operate the whole process. The wheelhouse is<br />

ergonomically designed for a one-man-operation with information supplied through 6 displays<br />

and alarm systems (Figure 4-3). Since the wheelhouse is positioned in front <strong>of</strong> the vessel, and<br />

the dredging equipment aft, the lone operator needs CCTV-systems to check the equipment aft.<br />

�����������������������������������������������������������������������<br />

As with crew numbers, there is a wide variety <strong>of</strong> TSHD size operating worldwide - from the<br />

smallest vessels up to the megadreders, which are discussed as a case study below. Table 4-2<br />

below summarizes the available data (Clarkson Research Services Ltd, 2009).<br />

75


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

� ����������� ������������ ���������������������<br />

Mean 96 17 6.1<br />

Maximum 231<br />

Minimum<br />

TSHD “Queen <strong>of</strong> the<br />

Netherlands”<br />

35<br />

TSHD“Ria de Navia”<br />

76<br />

41<br />

TSHD “Christobal<br />

Colon”<br />

7.4<br />

TSHD “Oosterschelde”<br />

15.2 TSHD “Christobal<br />

Colon”<br />

1.4<br />

TSHD “New Era”<br />

������ ����� ����� ������ ��� ���� ������ ������ ���� ���� �������� ���� �������� ��������� ���������� ���������<br />

���������������������<br />

The size <strong>of</strong> the vessel is partially governed by the type <strong>of</strong> dredging activity undertaken i.e. the<br />

capital and maintenance dredging industries have very different drivers from the marine<br />

aggregate industry. Capital dredging involves the creation <strong>of</strong> new or improved facilities e.g.<br />

harbour basins, navigational channels etc. while maintenance dredging is the removal <strong>of</strong> siltation<br />

from channels in order to maintain the design depths <strong>of</strong> navigation channels and ports (Van<br />

Raalte, 2006). Capital dredging typically requires the relocation <strong>of</strong> large quantities <strong>of</strong> sediment as<br />

quickly as possible, since additional dredging time equates to additional development costs while<br />

maintenance dredging in artificially deepened navigation areas needs to occur quickly so as to<br />

not compromise the operation <strong>of</strong> a facility. As a consequence this has driven the development <strong>of</strong><br />

larger vessels, able to rapidly dredge large volumes <strong>of</strong> sediment at a time, and with high loading<br />

and transit speeds. Chapter 3.1.5 showed that the driver on vessel size for the English fleet was<br />

their ability to enter the ports they supply and thus the vessel size is constrained by these<br />

parameters<br />

� �����������������������������������������������������<br />

Due to its small size, high population density and fast industrial development Singapore has<br />

constantly reclaimed land for development purposes. Planned construction and land reclamation<br />

led Moch and Tiarma (2002) to estimate that 1.8 billion m 3 <strong>of</strong> sand would be needed between<br />

2001 and 2010. To serve this market a range <strong>of</strong> larger Jumbo dredging vessels with hopper<br />

capacities between 17,000 and 25,000m 3 were built. The first example <strong>of</strong> a jumbo dredger was<br />

TSHD “Pearl River” (17,000m 3 ) followed later by vessels such as “Queen <strong>of</strong> the Netherlands”,<br />

“Volvox Terranova” and others.<br />

A new phase <strong>of</strong> major land reclamation projects in the Persian Gulf (e.g. Palm Island, The World)<br />

led the major world dredging companies to plan and commission even larger dredging vessels –<br />

the so called Megadredgers. The first <strong>of</strong> these to be commissioned was the Jan de Nul dredger<br />

“Vasco Da Gama” (33,000 m 3 ) in 2000. The vessel has twin 1400mm diameter dredge pipes and<br />

can dredge to a maximum depth <strong>of</strong> 125m with a loading time for the hopper <strong>of</strong> just one hour.<br />

“Vasco Da Gama” has mainly been used for land reclamation. It has also been used for diamond


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

mining in Southern Africa and deep <strong>of</strong>fshore dredging <strong>of</strong> iceberg protection wells for sub-sea<br />

production systems <strong>of</strong>f Newfoundland.<br />

A number <strong>of</strong> further Megadredgers are in the building or procurement phases. “Queen <strong>of</strong> the<br />

Netherlands is scheduled to be lengthened in 2009, while DEME ordered a vessel from IHC in<br />

early 2009 for delivery in 2011. Jan de Nul has two sisterships – the “Cristobal Colon” and the<br />

“Leiv Eriksson” – being built and these will both have 46,000m 3 hoppers. “Vox Maxima” is Van<br />

Oord’s new megatrailer with a 31,000m 3 capacity and was launched in 2009. Table 4-3 below<br />

gives comparisons for all the Newbuild (N) or Lengthened (L) mega dredgers.<br />

����������� ������� ������ ������ ���������� ���� ����� ����������<br />

���������<br />

���<br />

�����<br />

������� �������� ������� �����<br />

Owner VOA JDN BOKA JDN VOA DEME JDN<br />

N or L L N L N N N N<br />

Hoper (m³) 37,293 33,000 35,500 46,000 31,136 30,000 30,500<br />

Installed<br />

power (MW)<br />

Pump ashore<br />

(MW)<br />

Dredge power<br />

(MW)<br />

<strong>Dredging</strong><br />

depth (m)<br />

8.6 37.1 27.6 41.5 31.3 ??? 23.6<br />

11 16 12 16 13.3 ??? 15<br />

2 x 2.5 2 x 4.5 2 x 6 2 x 6.5 1 x 6 ??? 2 x 3.4<br />

101 140 115 2 x 155 125 ??? 93.5<br />

Draught (m) 13.37 14.6 11.49 15.15 14.5 max. 12<br />

Speed (kn) 15.5 16.3 ??? 18.0 17.0 ??? 16.0<br />

Commissioned 2008 2000 2009 2009 2009 2011 2010<br />

����������������������������������������������������������������������������������<br />

The global financial crisis and the curtailment or cancellation <strong>of</strong> many <strong>of</strong> the building projects in<br />

the Persian Gulf means the markets that these Megadredgers were designed to serve are now<br />

uncertain. Despite this, the major dredging companies are completing the building programmes<br />

for these ships. It may be that projections <strong>of</strong> global recovery indicate that major building projects<br />

are likely to recommence before delivery <strong>of</strong> the ships – or that the ships themselves can create<br />

their own markets by <strong>of</strong>fering previously impossible technical solutions.<br />

77<br />

m<br />

11.0


Benchmarking Equipment, Practices and Technologies<br />

������ �����������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

As with crew numbers and size there is an extremely wide variety <strong>of</strong> engine types, engine power,<br />

generators, dredge pumps and dredge powers on vessels <strong>of</strong> the world fleet. Data are not<br />

available for all vessels but Table 4-4 below summarizes the available data (Clarkson Research<br />

Services Ltd, 2009).<br />

� ����������������������<br />

������<br />

Maximum<br />

reported<br />

Minimum<br />

reported<br />

38,600 (i.e. 2 x 19,300)<br />

TSHD “Christobal<br />

Colon”<br />

320<br />

TSHD “Westerschelde”<br />

78<br />

�������������������<br />

������<br />

10,000<br />

TSHD “Vasco de<br />

Gama”<br />

100<br />

TSHD “Bir Anzarane”<br />

������������������<br />

�����<br />

13,100 (i.e. 2 x 6550)<br />

TSHD “Xin Hai Long”<br />

119<br />

TSHD “Currituck”<br />

������ ����� ��������� �������� ���� �������� �������� ���������� ���� ������� ����� ������� ����������<br />

�����������������������������<br />

There is very little data available on general productivity rates for TSHDs. Some examples <strong>of</strong><br />

specialized vessels with high productivities can, however, be derived from literature (Table 4-5)<br />

������������ ������������������ � �� ���������������� ������� � ��������<br />

“Brisbane” 2900 25 116<br />

“Wheeler” 6120 11 556<br />

�����������������������������������������������<br />

���� �����������������<br />

������ ����������<br />

At the lower end <strong>of</strong> the dredge pipe a draghead is installed - the function <strong>of</strong> which is to excavate<br />

the sediment and mix it with water for hydraulic transport. It has been shown that dragheads<br />

used by the UK dredging fleet are older-style dragheads (like the California type dragheads)<br />

where the primary source <strong>of</strong> production is erosion <strong>of</strong> sediment around the edges <strong>of</strong> the draghead<br />

caused purely by the water being sucked into the system by the dredge pump.<br />

One <strong>of</strong> the keys to maximizing the efficiency <strong>of</strong> dredging is to minimize the dredging cycle time<br />

and hence the time taken to load the hopper. This is particularly important for the maintenance<br />

and capital dredging sectors where there is a key financial driver to minimise dredging time. This<br />

is <strong>of</strong> less importance for the English aggregate dredging industry where tidal constraints limit<br />

many <strong>of</strong> the cycle times.<br />

Hopper loading time is strongly controlled by the performance <strong>of</strong> the draghead. Much research<br />

has taken place, mainly by the Dutch dredging industry, to maximize the performance <strong>of</strong> the<br />

draghead, and many dredging applications around the world now use more active draghead


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

systems. In many situations around the world, modern dragheads (Figure 4-4) use cutting teeth<br />

or knives, jetting (loosening the soil by the use <strong>of</strong> waterjets), or a combination <strong>of</strong> both to mobilise<br />

the dredged sediment.<br />

� ��������������<br />

���������������������������������<br />

One simple method <strong>of</strong> loosening the seabed sediment and increasing the erosion rate is the use<br />

<strong>of</strong> cutting teeth or knives on the draghead <strong>of</strong> the TSHD. The teeth act to cut the sediment and<br />

move the sediment particles into the water flow to form a slurry, which is then sucked up the<br />

dredge pipe. It has been reported (Brogdon Jr et al., 1994) that knives tend to increase<br />

production at high travel speeds, since they act as sediment displacement devices.<br />

Despite the fact that adding cutting teeth to a draghead seems a simple solution to increasing the<br />

erodibility <strong>of</strong> the sediment, it may have other operational consequences that are not immediately<br />

apparent, for instance Herbich (2000) reports that teeth can impart high and variable stresses<br />

onto the dredge pipe, particularly in the transition <strong>of</strong> the draghead from unconsolidated to<br />

cohesive sediments. There is therefore a risk <strong>of</strong> damage to the dredge pipe. In addition the teeth<br />

need to be sharp for optimum cutting <strong>of</strong> bed sediment and therefore need to be replaced on a<br />

regular basis and this may result in an increased down time in port.<br />

TSHDs with cutting teeth on the draghead are being commonly used in major projects worldwide<br />

– examples include “Queen <strong>of</strong> the Netherlands” which has been used to deepen the shipping<br />

channel for Melbourne Port; and the “Vasco de Gama” which was fitted with a specially modified<br />

draghead with an adaptable number <strong>of</strong> cutting teeth for dredging in Port Sudan (Malherbe and<br />

De Pooter, 2008).<br />

Figure 4-5 shows a schematic overview <strong>of</strong> the cutting process in saturated granular sediment. A<br />

blade is moving from left to right and cuts a layer with thickness ��. A shear zone develops in<br />

front <strong>of</strong> the blade and within the shear zone the soil is dilating and the pore volume increases due<br />

to shear which is a typical behaviour for densely packed sands.<br />

79


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

������������������������������������������������������<br />

The phenomenon <strong>of</strong> dilation during cutting causes the development <strong>of</strong> negative pore pressure<br />

(under pressure) in the soil, because water needs to flow to the area where dilation occurs. The<br />

negative pore pressure increases the effective pressure which can result in very high cutting<br />

forces under water. The magnitude <strong>of</strong> the decrease in pore pressure is limited since the pressure<br />

cannot be lower than zero (vacuum) and in that case cavitation <strong>of</strong> the pore water will occur (Van<br />

Os & Van Leusen, 1987).<br />

In calculating the cutting forces under water, two different regimes can be distinguished – the<br />

drained regime and the cavitating regime. Miedema (1994) defines the cutting force in the<br />

drained regime as:<br />

� ����� ��� �<br />

�<br />

� � � �<br />

� � � � � � � ������<br />

In which:<br />

� � Constant [-]<br />

� � Horizontal cutting force [N]<br />

� Width <strong>of</strong> the cutting tool [m]<br />

� Cutting depth [m]<br />

� � Relative increase in pore volume, defined as:<br />

80


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

� � �<br />

��� ��<br />

�<br />

��� �<br />

���<br />

�� � � � � � � ������<br />

In the cavitating regime the pore pressure in the shear zone is constant and equal to the vapour<br />

pressure. The pressure difference, which determines the effective pressure, is therefore<br />

proportional to the water pressure outside <strong>of</strong> the sediment, hence the water depth.<br />

�� �������������� � � � � � � ������<br />

� �<br />

In which:<br />

� � Constant [-]<br />

� water depth [m]<br />

The specific energy is therefore the amount <strong>of</strong> energy needed to excavate a certain volume <strong>of</strong><br />

sediment and is equal to the ratio between cutting power and production. In the cavitating regime<br />

this quantity is independent <strong>of</strong> the operational conditions, except the water depth.<br />

� �<br />

����� ��<br />

� � � �<br />

���� ��<br />

81<br />

� ���<br />

�<br />

��� � � � ���<br />

� � � � � � ������<br />

The coefficient �� depends on the cutting angle � and internal friction angle �� <strong>of</strong> the soil<br />

(Miedema, 1994):<br />

�� ������ ��������� �������� � � � � � � ������<br />

� �<br />

� � ��������������������������������������<br />

� ��������<br />

During dredging the weight <strong>of</strong> the draghead and dredge pipe penetrates the upper layer <strong>of</strong> the<br />

sediment. However when the sediment is compacted or armoured the weight may not be<br />

sufficient to penetrate the sediment and erosion <strong>of</strong> the sediments will be reduced. Reduced<br />

erosion results in low mixture densities and a reduced production speed.<br />

The theory behind incorporating jet nozzles is that by combining sufficient surface contact<br />

pressure with the sea bed and injection <strong>of</strong> jetwater the draghead can liquefy densely compacted<br />

sediments and increase loading speeds. To maximize efficiency <strong>of</strong> the jetting Vandyke (2002)<br />

suggests that the nozzles have to be integrated into the points <strong>of</strong> the draghead so that the<br />

waterjets cut the sediment only moments before the draghead penetrates. Brogdon Jr et al.<br />

(1994) report that water jet erosion benefits tend to increase as the travel speed <strong>of</strong> the dredging<br />

vessel decreases, since the jetting forces have a longer time to attack individual particles. The jet<br />

momentum available for the erosion process can be defined as:


Benchmarking Equipment, Practices and Technologies<br />

� �<br />

Where �<br />

Report No: 10/J/1/06/1309/0996<br />

� � �<br />

� ��� � ������� �����<br />

� � � � � ������<br />

�<br />

�<br />

� is the total jet discharge and � � the jet pressure drop over the nozzles in the<br />

draghead. So jet production can be increased by increasing jet discharge and/or jet pressure.<br />

The amount <strong>of</strong> hydraulic power � � (and likewise fuel consumption) used for the jetting system is<br />

equal to:<br />

� �<br />

�� � �� � � � � � � � � � � ������<br />

Combination the above two equations gives:<br />

� �<br />

� �<br />

�<br />

� �<br />

��� � ��<br />

� � � � � � � � ������<br />

� �<br />

From this equation it can be concluded that for a certain level <strong>of</strong> installed jet power, the<br />

momentum and therefore jet production will decrease with the jet pressure. In other words a<br />

combination <strong>of</strong> low jet pressure and high jet discharge is more efficient than the other way<br />

around. This however is not entirely true since the jet discharge will increase with decreasing<br />

pressure for a certain value <strong>of</strong> jet power. This will dilute the sand water mixture created on the<br />

seabed. The diluting effect can be shown using a simplified approach where we assume that the<br />

original jet water volume mixes with the excavated soil volume. The volume <strong>of</strong> excavated<br />

sediment (including the pore volume) � � is:<br />

�<br />

�<br />

�<br />

� �<br />

�<br />

����� � �<br />

�<br />

�� � � � � � � ������<br />

The sediment concentration � � in a mixture <strong>of</strong> sediment and water created by the jets is:<br />

� �<br />

�<br />

�<br />

�<br />

� � �<br />

� ���� � � � � � � ����������������������<br />

�� � ��<br />

The density � � <strong>of</strong> the jet mixture follows from:<br />

� ��� �<br />

� �<br />

� � � � � � �� � � � � ����������������������<br />

� � � � �<br />

This effect is illustrated in Figure 4-6 where jet production and the density <strong>of</strong> the mixture created<br />

by the jet is shown as a function <strong>of</strong> the jet pressure for a certain value <strong>of</strong> jet power. In the lower<br />

82


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

panel we see the jet production decreases with applied pressure. In the upper panel it is shown<br />

that density increases. This simple approach yields the upper limit <strong>of</strong> the jet density. In reality<br />

density will be lower because a water jet entrains surrounding water and therefore the excavated<br />

volume <strong>of</strong> sediment will be diluted with a larger volume <strong>of</strong> water than leaving the nozzle.<br />

�������������������������������������������������������������������������<br />

The sand – water mixture created by the jets is mixed with the water entering the draghead and<br />

sucked into the suction pipe. Therefore the mixture density transported to the hopper will always<br />

be lower than created by the jets. The combination <strong>of</strong> jet pressure and discharge determines the<br />

upper limit <strong>of</strong> the mixture density. This explains the tendency to increase jet pressure in modern<br />

dragheads. The specific energy <strong>of</strong> mechanical excavation (cutting) can be compared with the<br />

specific energy for hydraulic excavation.<br />

� �<br />

�<br />

���<br />

�� �<br />

� �<br />

� � � � � � ����������������������<br />

�� � �<br />

� � � �<br />

�<br />

�<br />

� ��� � �<br />

�� � �<br />

� � �<br />

��<br />

�<br />

�<br />

�<br />

The value �� is the excavation constant specific for jetting. The specific energy for cutting is<br />

compared with the specific energy for jetting in Figure 4-7 as a function <strong>of</strong> the jet pressure and<br />

different water depths. The figure shows that in the jetting process the specific energy increases<br />

with jet pressure, which is in agreement with Figure 4-6 where jet production decreases with<br />

pressure. The specific energy for cutting is higher than for jetting and the difference increases<br />

83


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

with water depth. The actual difference is even larger since here the hydraulic and mechanical<br />

power consumptions are compared. The cutting force is a result <strong>of</strong> the ships propulsion system<br />

and the jet power originates from a centrifugal pump. The efficiency <strong>of</strong> a jet pump is higher than<br />

the efficiency <strong>of</strong> a ships propeller, so when the comparison is made taking these efficiencies into<br />

account the difference between the two specific energies becomes larger.<br />

�����������������������������������������������������������������������<br />

So for dredging <strong>of</strong> granular sediment jetting is more efficient compared with mechanical<br />

excavation. This explains the fact that many modern TSHDs are nowadays equipped with a<br />

considerable amount <strong>of</strong> jet power as standard.<br />

84


Benchmarking Equipment, Practices and Technologies<br />

� ���������������������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

One system that incorporates high pressure jetting nozzles in the draghead <strong>of</strong> a conventional<br />

TSHD is the DEME patented DRACULA draghead system (DRedging And Cutting Using Liquid<br />

Action) (Vandycke, 2002) and was initially designed for dredging stiff clays. The DRACULA<br />

system was successfully used in a dredging operation to deepen the access channel to the port<br />

<strong>of</strong> Antwerp.<br />

Model tests indicated that in order to dredge stiff clays water pressures <strong>of</strong> between 300-350 bar<br />

were needed and therefore the DRACULA water jets generated 380 bar <strong>of</strong> pressure. This was<br />

deemed sufficient to erode the seabed sediment at trailing speeds <strong>of</strong> 2-2.5 knots.<br />

In order to drive water through draghead nozzles high pressure pumps must be installed. For the<br />

DRACULA system the pumps were driven by diesel engines, and any pumping system would<br />

therefore have an additional fuel implication if used on the English fleet. Vandyke (2002)<br />

suggests, as a rule-<strong>of</strong>-thumb, that one pump can feed ca. 20 nozzles <strong>of</strong> 2.0 mm diameter.<br />

Although the sediment type dredged was different from that dredged by the English marine<br />

aggregate fleet it is still useful to consider the results <strong>of</strong> using the jetting nozzles. It was found<br />

that production <strong>of</strong> the dredger increased between 15% and 27% and that when the system was<br />

active, the average fuel rack was 5% less compared to dredging without DRACULA. It was also<br />

observed that the maximum increase in production (27%) occurred when the jets in the visor<br />

together with the jets on the heel were active (Vandyke, 2002).<br />

� ��������������������������������<br />

The DRACULA case study showed how introducing water jets to the draghead can increase the<br />

efficiency <strong>of</strong> production, while chapter 4.1.1 described how the addition <strong>of</strong> knives or teeth to a<br />

draghead can also improve the rate <strong>of</strong> sediment loading. An obvious further experimental step<br />

was therefore to test a combination <strong>of</strong> knives and jetting as a means to improve dredging<br />

efficiency. The US Army Corps <strong>of</strong> Engineers first investigated how combinations <strong>of</strong> knives and<br />

jets affected dredging performance as long ago as 1994. These tests showed that the addition <strong>of</strong><br />

knives, placed in front <strong>of</strong> water jets increased the efficiency <strong>of</strong> dredging (Brogdon Jr et al., 1994).<br />

� �������������������������<br />

IHC Parts & Services from the Netherlands designed and patented the Wild Dragon draghead,<br />

which is characterized by a double row <strong>of</strong> cutting teeth with water jets contained within the teeth<br />

themselves. IHC claims that “the dredger production is increased, the dredging process produces<br />

less wear on components, the hopper mixture’s velocity is much lower, spoil settles sooner and<br />

the hopper is loaded faster. Lower emissions and less turbidity thanks to lower overflow from the<br />

hopper make the ‘Wild Dragon’, much better for the environment” (IHC, 2005). The draghead has<br />

been trialled on the TSHD “Xin Hai Long” in China (Figure 4-8).<br />

85


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

�����������������������������������������������������������������������<br />

����������������������������������������������������������<br />

The area where the “Xin Hai Long” normally operates, the Shanghai estuary, is known for its<br />

layers <strong>of</strong> extremely hard packed sand and has proved to be impossible to dredge by TSHD’s with<br />

acceptable production rates. Cutter suction dredgers could do the job but their deployment is<br />

impossible because <strong>of</strong> the intensive vessel traffic and wave conditions. IHC (2005) reports that<br />

the use <strong>of</strong> the Wild Dragon draghead on the “Xin Hai Long” was very successful. During the<br />

trials, dredging operations were carried out with two types <strong>of</strong> dragheads simultaneously; a normal<br />

one at port side and a Wild Dragon draghead at starboard side and the claimed improvement in<br />

production was up to 100% (Figure 4-9).<br />

���������������������������������������������������������������������������������������������������������<br />

��������������������������������������������<br />

������������������������������������������������������������������������������������������<br />

���������������������������������������������<br />

86


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Similar combinations <strong>of</strong> teeth and water jets are also in operation in dredgers operated in<br />

different industrial sectors e.g. in diamond mining <strong>of</strong>f the coast <strong>of</strong> Namibia dragheads on the<br />

TSHD “Vasco da Gama” are fitted with teeth and water jets to break up cohesive muds and clays<br />

and allow diamondiferous sediments to be loaded into the vessel for subsequent processing<br />

(Pisces Environmental Services (Pty) Ltd, 2008).<br />

Numerical tests on draghead efficiency with the addition <strong>of</strong> cutting knives and/or water jets were<br />

conducted by Brogdon Jr et al. (1994). Table 4-6 (from Brogdon Jr et al., 1994) summarizes the<br />

results, and it was found that water jets could significantly increase dredging production when a<br />

dredge was operating at slow speeds (less than 0.9 ms-1 in the tests conducted by Brogdon Jr et<br />

al. (1994)).<br />

�������������������<br />

�����<br />

�������������������<br />

������<br />

87<br />

��������������������<br />

Without knife With knife<br />

No jet n/a 0.03 0.05<br />

19 275 0.05 0.07<br />

19 483 0.1 0.07<br />

25 138 0.04 0.06<br />

25 275 0.05 0.07<br />

����������������������������������������������������������������������������������������������������


Benchmarking Equipment, Practices and Technologies<br />

� ����������������������������<br />

� �����������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

A technical modification that allows an increased area <strong>of</strong> seabed to be dredged on each pass<br />

(and hence reduce dredging time) is to increase the width <strong>of</strong> the draghead. The Japanese TSHD<br />

“Seiryu-maru" claims to have the world's largest-class wide span drag head which permits high<br />

accuracy levelled dredging without leaving excessive ridges (Yano et al., 2006).<br />

The ship has a 7.2m wide monobloc draghead, which is divided into four sections (Figure 4-10).<br />

These sections increase the ability <strong>of</strong> the draghead to follow undulations in the seabed and help<br />

prevent water suction caused by undulations which is a drawback with rigid monobloc<br />

dragheads. This draghead also incorporates a recycling system that returns the hold water to the<br />

drag head.<br />

Adoption <strong>of</strong> the wide span drag head has however resulted in a hull layout that has an aft centre<br />

dredge pipe (which is discussed in more detail in section 4.2).<br />

������������������������������������������������������������������������������������<br />

Dredgers with wide dragheads have also been used in other sectors e.g. the “Vasco de Gama”<br />

was fitted with a 7m wide draghead, capable <strong>of</strong> excavating 0.5m depth at a pass, for diamond<br />

mining <strong>of</strong>f the coast <strong>of</strong> Namibia (Pisces Environmental Services (Pty) Ltd, 2008).<br />

88


Benchmarking Equipment, Practices and Technologies<br />

�<br />

� �����������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

The Vic Vac dredge head was designed by a US company (J.F. Brennan Co., Inc.) to target<br />

removal <strong>of</strong> thin layers <strong>of</strong> s<strong>of</strong>t sediment above a firm substrate. The Vic Vac dredge does not use<br />

a conventional draghead, but instead consists <strong>of</strong> a unit that focuses suction directly below the<br />

assembly. Figure 4-11 shows the Vic Vac dredge head while Figure 4-12 show an isometric view<br />

<strong>of</strong> the bottom <strong>of</strong> the assembly (Green, ������������).<br />

���������������������������������������������������������������������������������������������<br />

���������������������������������������������������������������������������������<br />

��������������������������������������������<br />

The Vic Vac was designed to allow precise removal <strong>of</strong> sediment without resuspending large<br />

amounts <strong>of</strong> sediment and hence minimizing the plume caused by dredging. The Vic Vac has<br />

89


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

been successfully used to dredge contaminated sediments with minimal resuspension in the<br />

Ashtabula River, near Lake Erie, Ohio, USA (USEPA, 2008).<br />

A second version <strong>of</strong> a focussed suction system, also developed in the United States, is the<br />

Tornado Motion dredging system. This is based on an eddy pump (Figure 4-13) which consists <strong>of</strong><br />

an energy generating rotor attached to the end <strong>of</strong> a drive shaft. As the rotor begins to spin, it sets<br />

into motion the ambient fluid present and this spinning fluid is forced down into the hollow centre<br />

<strong>of</strong> the intake chamber. Within the intake chamber it creates a high speed, swirling synchronized<br />

column <strong>of</strong> fluid which agitates the sediment to be pumped and the agitated material travels by<br />

reverse flow along the sides <strong>of</strong> the intake chamber, into a volute. Here the sediment, under<br />

pressure from below, is forced into the discharge pipe (Weinrib, pers comm.).<br />

��������������������������������������������������������<br />

�������������������������������������������������������������<br />

The eddy pump was first used in a demonstration project in 1994 and it was found that there was<br />

little resuspension <strong>of</strong> sediment at the suction head. Creek and Sagraves (1995) reported that,<br />

average turbidity around the suction head exceeded background levels by 1.2 Nephelometric<br />

Turbidity Units (NTU). The maximum turbidity increase at the head was 1.9 NTU and maximum<br />

suspended solids measurement was 9.0 mgl-1. The maximum turbidity <strong>of</strong> the discharge plume<br />

was only 12 NTU over ambient. Figure 4-14 shows the Tornado Motion suction head.<br />

������������������������������������������<br />

�������������������������������������������������������������<br />

90


Benchmarking Equipment, Practices and Technologies<br />

�<br />

� �����������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

91<br />

�<br />

An extreme way <strong>of</strong> controlling the draghead is to make it semi-autonomous, and the Penta-<br />

Ocean Construction Co. Ltd. in Japan has done that in developing a Submersible <strong>Dredging</strong><br />

Robot for highly accurate seabed dredging. This has been named the “Futaba” and is an 8legged<br />

robot, which walks on the seabed with automatic movement and posture control, and can<br />

be controlled by a single operator. The dredge sediment is pumped back along a pipe to shore or<br />

hopper. “Futaba” has a 320kW dredge pump, however its rate <strong>of</strong> production is low at only<br />

70m 3 /hr (Hisao and Sadayoshi, 1999). It is claimed, however, that it is capable <strong>of</strong> generating<br />

accurate dredging pr<strong>of</strong>iles since it transmits real-time positions and status back to the operator.<br />

In addition since it operates on the seabed it is relatively insensitive to wave action even under<br />

severe sea conditions (Penta-Ocean Construction Co., 2009).<br />

� ������������������������������������<br />

The low turbidity draghead is a modified draghead <strong>of</strong> a TSHD with a specially designed suction<br />

box. A moveable visor or valve makes it possible to use the draghead in a sweeping motion in<br />

two different directions. The cutting height can be adjusted to dredge layers <strong>of</strong> different<br />

thicknesses with the lower cutting edge shaving the sediment layer and the upper visor adjusting<br />

to the bottom pr<strong>of</strong>ile in order to prevent excess water flowing into the draghead. The system<br />

includes a degassing system which prevents cavitation in the suction pump (Bray, 2008).<br />

������ ����������������������������������<br />

The efficiency <strong>of</strong> the dredge pump and impeller to move sediment from the seabed within the<br />

dredging area and into the hopper <strong>of</strong> a dredger is fundamental to controlling the loading time <strong>of</strong><br />

the vessel. Therefore, the better the efficiency <strong>of</strong> the pump and impeller the less time the vessel<br />

is directly impacting the environment by loading.<br />

The pressure difference generated between the suction and discharge sides <strong>of</strong> the dredge pump<br />

is called the manometric head. The pump must energize the sediment/water mixture, move it<br />

through the dredge pipe, and overcome the effects <strong>of</strong> friction and gravity. Conventional dredge<br />

pumps (Figure 4-15) have an efficiency <strong>of</strong> approximately 80% at their Best Efficiency<br />

Point however research by Dutch dredging scientists has led to the development <strong>of</strong> high<br />

efficiency pumps (Figure 4-16) which, it is claimed, have an efficiency <strong>of</strong> about 90% at the BEP<br />

(IHC, 2004).


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

��������������������������������������������������������������������������������<br />

����������������������������������������������������������������������������<br />

Van den Berg (2001) reports that conventional dredge pumps have impellers with single curved<br />

blades, a suction shield with a small radius at the entrance and a casing with a large throat area<br />

while high efficiency pumps have blades that are curved in three dimensions, an increased radius<br />

<strong>of</strong> the suction shields, a decrease <strong>of</strong> the inlet angles <strong>of</strong> the blades, a change in the design <strong>of</strong> the<br />

inlet edge <strong>of</strong> the blades and optimisation <strong>of</strong> the blade angle. This has resulted in pumps which<br />

generate less turbulence, fewer impact losses and less friction. Figure 4-17 shows the increased<br />

efficiency when using high efficiency pumps and impellers in coarse sand and gravel and<br />

medium fine sand substrates.<br />

92


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

���������������������������������������������������������������������������������������������������������<br />

������������������<br />

These dredge pumps are designed to allow them to be retro-fitted to existing dredgers and not<br />

just new builds.<br />

There is a wide range <strong>of</strong> dredge pipe diameters on the TSHDs <strong>of</strong> the world fleet. The largest<br />

dredge pipes currently in service (1400 mm diameter) are on the TSHDs “Vasco de Gama” and<br />

“VolvoxTerranova”, however it is reported that a currently unnamed dredger being built by<br />

Guangzhou Wenchong in China will have a 9000 mm diameter pipe. These large pipes can be<br />

contrasted with the USACE TSHD “Currituck” which has a pipe <strong>of</strong> only 254 mm diameter.<br />

Another development in dredge pipe technology is IHC’s patented TelePipe design which<br />

consists <strong>of</strong> an extendable, synthetic suction tube enclosed in a reinforced synthetic lattice system<br />

with special gliders that are operated by a remotely controlled winch. The system is still being<br />

developed by IHC but they report that feasibility studies are promising (IHC, 2009b). The goal is<br />

to make deeper dredging possible using existing TSHDs, without costly ship extensions or major<br />

alterations to the gantry arrangement. The system will replace the lower part <strong>of</strong> conventional<br />

suction pipes. A typical TelePipe measuring 30 metres extends to about 50 metres and IHC<br />

claims that the weight, strength and reliability will match the replaced pipe.<br />

93


Benchmarking Equipment, Practices and Technologies<br />

� �����������������������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

High efficiency pumps have been installed by Warman International on the TSHD “Brisbane”<br />

which is one <strong>of</strong> Austalia’s largest dredging vessels at 85-metres long and 5000 tonnes gross<br />

weight (Ferret, 2001). The pumps weigh 36 tonnes each and can pump directly into the ship’s<br />

hopper or directly ashore. They are built with specially designed frames and cover plates to<br />

withstand the high pressures generated by the pumps which can reach up to 1500 kPa. On<br />

conventional pumps the frame and coverplates are generally cast however in these high<br />

efficiency pumps they are fabricated to meet the higher pressure requirements. The vessel is<br />

used to dredge the Brisbane River estuary at depths <strong>of</strong> up to 25 metres, and the pumps have the<br />

capacity to pump at 2500 m 3 /s. Complete loading <strong>of</strong> the 2900 m 3 hopper can be completed in 25<br />

minutes.<br />

� ��������������������������������<br />

The Dutch dredging company Pinpoint <strong>Dredging</strong> has constructed and successfully tested and<br />

operated the Punaise (Dutch for "thumbtack") system which is a remotely operated pumping<br />

system. The Punaise sits on the seabed, is watertight and contains the dredge pump, motor and<br />

ballast tanks (Figure 4-18). It is connected to the shore via an umbilical that consists <strong>of</strong> control<br />

connections, the power supply and the sediment discharge pipe (Van Rijn et al., 2005).<br />

��������������������������������������������������<br />

Since it sits on the seabed it can remove bottom sediments without impacting navigation and<br />

without being affected by storms.<br />

The first stage <strong>of</strong> operation consists <strong>of</strong> positioning the Punaise in an appropriate location and<br />

sinking the housing by filling the ballast tanks. Fluidizers are activated, which allows the supports<br />

to settle into the seabed and dredging can begin. As sediment is dredged and pumped ashore a<br />

94


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

pit is formed with the Punaise at the lowest point. It dredges by means <strong>of</strong> hydraulic erosion where<br />

sediment flows down the slopes <strong>of</strong> the pit and into the suction entrance. It is only suitable,<br />

therefore, for dredging thick layers <strong>of</strong> sediments that are non-cohesive and can flow down slope.<br />

The Punaise system is not new, and the first prototype was developed in 1990. It was used for<br />

two years to dredge 1.2 million m 3 <strong>of</strong> silt from Flushing Harbour in The Netherlands. It was sited<br />

in the centre <strong>of</strong> the turning basin and pumped the silt through a submerged pipeline directly to<br />

the Schelde estuary which ends in the North Sea (Williams and Visser, 1997).<br />

� ���������������������������<br />

A technical modification that allows an increased area <strong>of</strong> seabed to be dredged on each pass<br />

(and hence reduce dredging time) is to increase the width <strong>of</strong> the draghead and the use <strong>of</strong> this<br />

technique on the Japanese TSHD “Seiryu-maru" has been discussed in Chapter 4.1.4 above. A<br />

further modification necessitated by the use <strong>of</strong> the wide draghead was the hull layout <strong>of</strong> the<br />

“Seiryu-maru” that has an aft centre dredge pipe (Figure 4-19).<br />

��������������������������������������������������������������������������������������������<br />

An environmental benefit that may be afforded by an aft centre system is that it allows the<br />

dredging line to be shown by the ship’s wake. This may improve dredging accuracy and<br />

efficiency without the need for a Dredge Track Presentation System – if there are other visual<br />

reference points available. The presence <strong>of</strong> an aft centre pipe does, however, potentially<br />

compromise the volume <strong>of</strong> the hold – the 104m-long “Seriyu-maru” has a hopper capacity <strong>of</strong> only<br />

1700m 3 (although it also contains 1500m 3 <strong>of</strong> oil recovery tanks) (Yano et al., 2006).<br />

Another vessel with a central pipe is the US Army Corps <strong>of</strong> Engineers vessel TSHD “Wheeler”,<br />

built as long ago as 1983. Unlike the “Seriyu-maru” which only has a central pipe the “Wheeler”<br />

has a central 1000mm diameter dredge pipe in addition to two 660mm diameter side pipes. The<br />

vessel is solely a maintenance dredger and can dredge at depths <strong>of</strong> up to 29m for the side pipes<br />

but the central pipe can only reach a depth <strong>of</strong> 16.8m. When all three pipes are able to be<br />

deployed the vessel loads rapidly – the 6120m 3 hopper can be filled in as little as 11 minutes<br />

(USACE, 2006).<br />

95


Benchmarking Equipment, Practices and Technologies<br />

������ ����������������������<br />

Report No: 10/J/1/06/1309/0996<br />

Similar to the UK dredging industry, two screening systems are commonly found in dredgers<br />

worldwide – these are the central screening tower or the revolving system. The central screening<br />

tower with midships loading channel is the simplest but must be designed to minimise<br />

interference with the loading and unloading process. The revolving system involves the<br />

installation <strong>of</strong> a rotatable screening tower to the side <strong>of</strong> the hopper. Because it can turn almost<br />

300º it can evenly spread material over the hopper thus improving the hopper fill ratio. Existing<br />

hopper dredgers can also be effectively retr<strong>of</strong>itted with screening hardware.<br />

������ �����������������<br />

During dredging a mixture <strong>of</strong> sediment is discharged into the hopper <strong>of</strong> the dredger. The intention<br />

is that the sediment settles in the hopper and the excess water flows overboard. Generally,<br />

however, a proportion <strong>of</strong> the incoming sediment does not settle in the hopper but instead flows<br />

overboard with the excess water. Depending on the particle size distribution (PSD) <strong>of</strong> the<br />

sediment, the hopper geometry and other process parameters this overflow loss can reach<br />

values <strong>of</strong> up to 30-40 % <strong>of</strong> the total volume <strong>of</strong> sediment pumped into the hopper. It is important to<br />

quantify the overflow loss for three reasons:<br />

� The loading time, and therefore the amount <strong>of</strong> time the environment is directly affected<br />

by dredging (as well as the economic cost price <strong>of</strong> the sediment dredged) increases due<br />

to the overflow losses.<br />

� In particular the finer fractions <strong>of</strong> the particle size distribution (PSD) are present in the<br />

overflow mixture, implying that sediment contained in the overflow will settle slowly and<br />

can create a turbid plume around the dredger.<br />

� For environmental modelling reasons it is important to quantify the quantity and the PSD<br />

<strong>of</strong> the overflow sediments since these are needed as inputs to the dispersion models.<br />

The function <strong>of</strong> the overflow system is to discharge the excess water and to control the water<br />

(and therefore sediment) level in the hopper. The excess water is released overboard through the<br />

overflow structure, which acts as a weir. In modern aggregate dredgers these structures are<br />

mostly adjustable in vertical position to regulate the overflow level in the hopper. A telescopic<br />

overflow structure in the cross section <strong>of</strong> a hopper is shown in Figure 4-20.<br />

96


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

������������������������������������������������<br />

The operational conditions which have an influence on loading production and hence overflow<br />

losses are:<br />

� Discharge<br />

� Loading Concentration<br />

� Loading Time<br />

� Loading procedure<br />

� Water Temperature<br />

The discharge Q, together with the hopper area has the largest influence on the overflow losses.<br />

The ratio between the discharge and hopper area is called the overflow rate and can be regarded<br />

as an average vertical flow velocity in the hopper. The sediment concentration <strong>of</strong> the mixture<br />

discharged in the hopper has also an important influence on loading and losses.<br />

The sediment production sucked from the seabed is the product <strong>of</strong> discharge and concentration<br />

and a larger concentration means a larger suction and loading production.<br />

The settling velocity <strong>of</strong> the sediment particles decreases with sediment concentration - this is<br />

called hindered settling (Richardson & Zaki, 1954). The hindered settling effect will increase the<br />

cumulative overflow losses. The gradient <strong>of</strong> the loading production versus concentration will<br />

therefore decrease with concentration and Figure 4-21 shows the loading production as a<br />

function <strong>of</strong> the inflow concentration for three different grain sizes while inflow discharge is<br />

constant. In this example overflow losses for the coarsest sediment type (250 micron) are low<br />

and loading production increases linearly with the inflow concentration. For the finer sands<br />

hindered settling plays a role in this example. This effect increases with inflow concentration and<br />

loading production decreases relatively. The difference between the straight line and the curved<br />

97


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

lines is the overflow loss. For example, with an inflow concentration <strong>of</strong> 0.3 and particle size <strong>of</strong><br />

150 microns loading production is 6 ton/s. The suction production is 8 ton/s. The difference is the<br />

overflow loss (2 ton/s). The overflow losses will increase normally with the time elapsed since the<br />

start <strong>of</strong> loading. This is mainly caused by the increasing sediment concentration above the settled<br />

sediment bed in the hopper.<br />

��������������������������������������������������������������������������<br />

The water temperature influences the viscosity <strong>of</strong> water. For the finer fractions <strong>of</strong> the particles<br />

size distribution the flow around the particles is in the laminar <strong>of</strong> transitional regime where<br />

viscosity <strong>of</strong> the fluid influences settling velocity. This is shown in Figure 4-22 where the settling<br />

velocity for two different values (10 and 40 o C) <strong>of</strong> water temperature is shown. For coarser<br />

sediment the flow pattern around the particles becomes turbulent and so viscosity will play not an<br />

important role anymore.<br />

������������ ����������������������������������������������������<br />

98


Benchmarking Equipment, Practices and Technologies<br />

� ���������������������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

The rationale behind anti-turbidity systems is to prevent air entrainment in the overflow mixture. If<br />

air is entrained then bubbles are dragged down with the flow through the overflow pipe and are<br />

released below the ships hull. The bubbles then rise to the water surface and create an upwards<br />

flow (air-lift effect) from the overflow outflow location towards the water surface. Sediment<br />

particles are transported to the surface with this flow and form a plume. Depending on the<br />

location <strong>of</strong> the overflow and the flow along the ships hull the air-water flow can even guide the<br />

sediment particles towards the propeller wake behind the ship. The high turbulence level at this<br />

position will effectively mix the sediment through the water column. Hence the presence <strong>of</strong> air<br />

bubbles will stimulate the forming <strong>of</strong> a passive plume, especially at the water surface, where very<br />

limited suspended sediment concentrations are clearly visible. Due to the very low settling<br />

velocity <strong>of</strong> the particles in this passive plume, measured concentrations will be larger around the<br />

dredger compared with situations where air entrainment is prohibited and a dynamic plume is<br />

present. In that case the trajectory <strong>of</strong> the plume will be directed straight towards the seabed and<br />

<strong>of</strong>ten no turbidity at the water surface will be visible.<br />

Designs for overflow svstems that will reduce the surface plume have been proposed for over 30<br />

years. Ofuji and Ishimatsu (1976) reported on an anti-turbidity overflow system (Figure 4-23) that<br />

had proven effective in reducing the surface turbidity generated by water overflowing from a<br />

TSHD. Their system could be incorporated in the existing dredgers through simple modifications<br />

<strong>of</strong> existing overflow systems and suppressed the generation <strong>of</strong> air bubbles in the overflow water<br />

by inserting an inclined baffle plate in the overflow chute. The chute was modified so the overflow<br />

water was retained long enough for any contained air bubbles to rise and vanish and the<br />

overflow water was subsequently discharged through a submerged outlet.<br />

������������������������������������������������������������������������������<br />

99


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

The combination <strong>of</strong> overflow ports on the bottom <strong>of</strong> the hull and reduced air bubbles in the water<br />

leads to the overflow descending rapidly to the seabed with a minimum amount <strong>of</strong> dispersion in<br />

the water column. Herbich and Brahme (1991) reported that the Ishikawajima-Harima Heavy<br />

Industries Company Ltd., Japan, further developed the system and incorporated it in three<br />

Japanese TSHDs with capacities ranging from 2,000 to 3,975 m 3 . Herbich and Brahme (1991)<br />

also report that tests were carried out on the dredger “Kairyu Maru” and showed considerable<br />

reductions in the turbidity at the surface and at 1 m below the surface, both by the side and aft <strong>of</strong><br />

the ship. These data are shown in Table 4-7 below:<br />

� ����������������������������� �����������������������������<br />

Average<br />

At surface 1m below At surface 1m below<br />

Concentration <strong>of</strong><br />

Suspended<br />

Solids (ppm)<br />

surface<br />

surface<br />

At side <strong>of</strong> ship 627 272 6.0 8.2<br />

Aft <strong>of</strong> the ship 119 110 6.5 8.9<br />

������������������������������������������������������������������������������������������������������<br />

����������������������������������������������������<br />

No data are available on sediment concentrations at the bottom <strong>of</strong> the water column by the side<br />

and behind the dredger.<br />

� ����������������������������������������������<br />

In a standard overflow system a large volume <strong>of</strong> air is mixed with the water and sediment due to<br />

the high fall height and this causes turbulence and increased spreading <strong>of</strong> the dredge plume<br />

(Minerals Management Service, 2004). Jan de Nul (2003) has developed a low turbidity valve<br />

which is an adjustable valve in the overflow funnel which chokes the flow such that no air is taken<br />

down with the suspension leaving the hopper (Figure 4-24).<br />

�����������������������������������������������������������������������<br />

100


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

Minerals Management Service (2004) notes that data supplied by the manufacturer indicate that<br />

the size <strong>of</strong> the plume and the total amount <strong>of</strong> sediment in the plume are reduced through using<br />

the low turbidity valve (Figure 4-27). From 1995-2000 a monitoring project for the maintenance<br />

dredging operations in Belgian coastal harbours and waterways was carried out under the name<br />

MOBAG 2000 (Van Parys et al., 2001). Turbidity was measured during dredging with a TSHD<br />

using an environmental valve in the overflow and a standard overflow without environmental<br />

valve. It was found that dredging with the standard method gave the highest turbidities and the<br />

natural background situation returned after 30-45 minutes. The environmental valve reduced the<br />

turbidity to 40% <strong>of</strong> the standard method and background situation returned after 20-30 minutes.<br />

The overflow mixture leaves the hopper through the overflow arrangement, which is normally a<br />

telescopic pipe. Depending on:<br />

� The draught <strong>of</strong> the vessel,<br />

� The mixture level in the hopper,<br />

� The overflow discharge,<br />

� The diameter <strong>of</strong> the overflow pipe and overflow crest, and<br />

� The relative opening <strong>of</strong> the green valve,<br />

The following two situations can occur as shown in Figure 4-25. In the left panel <strong>of</strong> the figure the<br />

water level in the overflow pipe is lover than in the hopper. The mixture flows over the crest <strong>of</strong> the<br />

overflow weir and plunges into the overflow pipe. Hence the mixture enters the pipe as a<br />

plunging jet. It is known that above a certain critical speed a plunging jet entrains air at the<br />

interface between the impinging jet and the stagnant water (Ervine (1976); Bins (1988)). The<br />

entrained air bubbles will be transported downward and leave the vessel with the sediment<br />

mixture at the bottom <strong>of</strong> the ship. The only method to avoid entrainment <strong>of</strong> air is taking care that<br />

the plunging jet situation does not occur. This can be achieved by increasing the water level in<br />

the overflow pipe by increasing the hydraulic resistance <strong>of</strong> the overflow pipe. In that case the<br />

situation as shown in the right panel <strong>of</strong> Figure 4-25 will occur. The water level must be high<br />

enough relative to the overflow level to avoid air entrainment, but should not be too high<br />

otherwise the maximum draught <strong>of</strong> the vessel will be exceeded. The mixture level should stay<br />

between these two limits. This optimal level will depend on the influences mentioned above,<br />

therefore the (extra) hydraulic resistance due to the green valve must be variable. The variable<br />

resistance is mostly created by placement <strong>of</strong> a butterfly valve in the overflow pipe. A control unit<br />

must be present to adjust the valve position to keep the level between the limits without<br />

exceeding maximum draught.<br />

�<br />

������������ ������������������������������������������������������������������������������������������<br />

101


Benchmarking Equipment, Practices and Technologies<br />

� �������������������������������������<br />

A system, sometimes called the “Green Pipe”, has been developed by both IHC, Holland and<br />

Krupp Foerdertechnik <strong>of</strong> Germany to reduce plumes and increase dredging efficiency. This<br />

system recirculates the overflow waters back to the draghead in a closed system and reduces<br />

the surface turbidity plumes. Instead <strong>of</strong> allowing the overflow (composed <strong>of</strong> fine organics and<br />

transport water) to go directly back into the water, the new system pumps the overflow along the<br />

dragarm to the draghead to assist in suction operations (McLellan and Hopman, 2000). Reusing<br />

overflow water in this manner maintains a closed system so that only a minimal turbidity plume is<br />

produced at the sea surface and potential contaminated sediment is not discharged. This is<br />

sometimes called the “Green Pipe” system.<br />

Report No: 10/J/1/06/1309/0996<br />

�<br />

�<br />

�����������������������������������������������������������������������<br />

This technique has three main benefits: (a) minimization <strong>of</strong> sediment/turbidity plumes at the sea<br />

surface by avoiding transport water overflow and a possible related water pollution problem<br />

(Figure 4-27); (b) increased material flow in relation to transport water, which in turn increases<br />

dredge efficiency by reducing loading time, and (c) decreased pressure drop inside the<br />

draghead, which reduces dredge pulling force. This reduces energy needed for propulsion and<br />

lowers fuel consumption. Some manufacturers claim a 20% increase in efficiency for dredging<br />

silty sand with the recirculating system (Francingues et al., 2000).<br />

102


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

����������������������������������������������������������������������<br />

Minerals Management Service (2004) states that sediment balance considerations suggest that<br />

the same amount <strong>of</strong> sediment would be “overflowed” with the “Green Pipe” - the only difference<br />

being its release close to the bed. Minerals Management Service (2004) also states that there is<br />

a concern that this could promote the development <strong>of</strong> a turbidity current near the bed (i.e., due to<br />

the high and concentrated sediment loading at the drag head with this approach) and this if such<br />

a turbidity current were to develop at the bed, the sedimentation footprint may extend much<br />

further from the dredge site than normally expected.<br />

������ ������������������<br />

Discharging sediment from the dredger can have potential environmental impacts – particularly<br />

where the sediment is contaminated with pollutants or fine grained material. Indeed many<br />

contaminated cargos are simply dumped, either at licensed dumping areas at sea, or in landfill<br />

sites.<br />

One answer to the problem <strong>of</strong> discharging contaminated sediment is the Mobile Soil Washing<br />

Plant (MSWP) which was developed by Boskalis Dolman and has been used on dredging<br />

projects on the Miami River (Taylor et al., 2006). The MSWP is “portable” – prior to its use in<br />

Miami it was in use in south-eastern England – although it did take three weeks to disassemble,<br />

transport and reassemble at site. The MSWP is a high capacity plant, processing up to 150<br />

tonnes <strong>of</strong> sediment per hour and combines coarse fraction separation via a stationary grizzly<br />

screen, rotating sieve drums, vibrating shaker screens and hydrocyclones, and mechanical dewatering<br />

<strong>of</strong> the fine fraction through a pre-thickener tank and a series <strong>of</strong> belt filter presses.<br />

Taylor et al. (2006) describe the operation <strong>of</strong> the MSWP - coarse material with a particle size <strong>of</strong><br />

from 1 m to 3 cm pass the grizzly screen and are placed in a rotating wash and sieve drum. The<br />

finer sediment is placed through vibrating shaker screens before the sand separation process<br />

103


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

begins using hydrocyclones and countercurrent washing. The de-sanded sediment is then put<br />

through a mechanical de-watering process where flow and density measurements are conducted<br />

and automated polymer dosing occurs. This results in the pre-conditioning <strong>of</strong> fines <strong>of</strong> 20 percent<br />

dry solids in a pre-thickener tank. The mechanical de-watering table uses belt filter presses which<br />

reduces the remaining dry solids content by about 50 percent. More than 50 percent <strong>of</strong> the<br />

dredged sediment was processed clean (coarse) fractions which could be beneficially used in the<br />

community at little to no cost to the project; and around one quarter was dewatered filtered cake<br />

which was transported by truck for ultimate disposal at one <strong>of</strong> three licenced landfills.<br />

���� ���������������������������������������������������������<br />

������ �������������������<br />

� ������������<br />

While it is not possible to state definitively, it is likely that navigation on the majority (if not all <strong>of</strong><br />

the world’s TSHDs) utilizes Differential GPS.<br />

� ���������������������������������������<br />

In addition the International Maritime Organization (IMO) requires all vessels over 299 Gross<br />

Tonnes to be fitted with an Automatic Identification System (AIS). Only one TSHD has a weight<br />

less than this – the Spanish vessel “Ria de Navia” (reported 298 Gross Tonnes (Clarkson<br />

Research Services Ltd, 2009)). The AIS transponder transmits the vessel name, position, speed<br />

and course is intended to help ships avoid collisions, as well as assisting port authorities to better<br />

control sea traffic.<br />

A refinement <strong>of</strong> navigation technology and electronic charting is the development <strong>of</strong> draghead<br />

steering for dredging management. This has a number <strong>of</strong> advantages both for production and for<br />

environmental efficiency. The ability to accurately steer the draghead means that the <strong>of</strong>ftake <strong>of</strong><br />

resource within a dredging area can be maximised. Ridges <strong>of</strong> sediment between dredging lanes<br />

that were previously left undredged can be dredges, thus helping to minimise the overall area <strong>of</strong><br />

seabed dredged.<br />

IHC has developed a system that can be used on TSHDs called the Dredge Track Presentation<br />

System (DTPS) (Mallee, 2000). This is a s<strong>of</strong>tware package, written in C++ and running on a<br />

standard Windows PC, that allows the dredge operator to see in real-time the position <strong>of</strong> the<br />

draghead and an updated bathymetry (Figure 4-28). The DTPS has been designed to operate as<br />

either a stand-alone system or be integrated with other electronic systems on a dredger, such as<br />

the ECDIS, any dredge pipe position monitoring system and dynamic positioning and dynamic<br />

tracking (DP/DT) systems.<br />

104


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

���������������������������������������������������������������������������������������������<br />

The DTPS was developed by IHC Systems in close co-operation with DEME and the system<br />

underwent extensive testing on DEME vessels. The system has been in use since 1999 when it<br />

was trialled on the DEME vessel “Lange Wapper” and used for dredging in the Western Schelde<br />

(IHC, 2001).<br />

� ���������������������������������������������������������<br />

A number <strong>of</strong> recent dredging operations include almost real time management <strong>of</strong> dredging<br />

efficiency by using multi-beam sonar to provide instantaneous data to the bridge <strong>of</strong> the TSHD<br />

and interfacing these data with the navigation systems. One example is a trenching and<br />

backfilling operation for a pipeline running from Yung-An to Tung-Hsiao in Taiwan and reported<br />

by Van Melkebeek (2002). During this campaign the TSHD “Gerardus Mercator” installed a multibeam<br />

echosounding device in the moon pool <strong>of</strong> the dredger. This provided on-line, day-to-day,<br />

survey information for the project team and client representatives on board the vessel. Similar<br />

multi-beam survey equipment was installed on the TSHD “Oranje” for use in a dredging<br />

campaign for the Balgzand-Bacton pipeline. This allowed the vessel to survey the efficiency <strong>of</strong> its<br />

operations in dredging 381 sand ridges in water depths <strong>of</strong> up to 50 m (Boskalis Offshore bv,<br />

2009).<br />

In 2003 the TSHD “Vasco da Gama” was employed by Husky Energy, a Canadian oil company,<br />

to excavate a Glory Hole approximately 200 nautical miles south-east <strong>of</strong> Newfoundland on the<br />

Grand Banks in the Atlantic Ocean. A Glory Hole is a depression, up to 9 m below seabed level,<br />

which protects wellheads against iceberg scouring. During the dredging process, progress was<br />

monitored in real-time on board the dredger by means <strong>of</strong> the dredging control systems. The<br />

105


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

actual draghead depth and the target depth per layer were compared online and any difference<br />

displayed. The position <strong>of</strong> the draghead was visualised on screen on a background <strong>of</strong><br />

bathymetric data by means <strong>of</strong> a plan view with a differential colour chart showing the amount still<br />

to be dredged together with a longitudinal and cross pr<strong>of</strong>ile <strong>of</strong> the Glory Hole, marking seabed<br />

level and target level. Progress was checked using a multibeam system installed in the moon<br />

pool and the system continuously updated the dredge levels achieved (Jan de Nul, 2009).<br />

In Belgium, the Maritime Schelde Department has standardized a management system that<br />

records real-time data on dredging operations including: location, depth <strong>of</strong> cut, sediment mixture<br />

concentration, and several other parameters that help determine the performance <strong>of</strong> the<br />

operation (Francingues et al., 2000).<br />

� �����������������������������������������������������<br />

A further refinement <strong>of</strong> on-board real-time management is a system that extends to enabling<br />

shore-based <strong>of</strong>fices to access real-time data from a dredging vessel. In 2005, IHC Systems was<br />

awarded a contract by the Sri Lanka Port Authority to develop, build and install such a system on<br />

the TSHD “Hansakawa” (IHC, 2006). The system consisted <strong>of</strong>:<br />

� Real Time Kinematic (RTK) DGPS positioning system<br />

� Suction Tube Position Monitor (STPM)<br />

� Dredge Track Presentation System (DTPS) for the vessel and a second for an onshore<br />

<strong>of</strong>fice<br />

� Interfaces for existing instrumentation systems.<br />

The RTK system allowed the position <strong>of</strong> the ship to be determined with an accuracy <strong>of</strong> a few<br />

centimetres and the set up included a dual RTK DGPS receiver on board the ship and an RTK<br />

DGPS base to serve as a shore reference station. The shore-based DTPS system interfaced with<br />

the “Hansakawa” using a full-duplex radio link. An interface was created between the DTPS and<br />

the dredger permitting online presentation <strong>of</strong> production both on board and at the SLPA <strong>of</strong>fices<br />

and the system was fully operational by July 2006 (IHC, 2006).<br />

106


Benchmarking Equipment, Practices and Technologies<br />

� �����������������������������������<br />

� ������������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

Chapter 3.2.5 described how the UK aggregates industry uses an Electronic Monitoring System<br />

(EMS) to monitoring the activities <strong>of</strong> dredgers on English licence areas. The use <strong>of</strong> similar<br />

systems on board aggregate dredging vessels is now common practice among vessels from<br />

other European countries including the Netherlands (Figure 4-29), Denmark, Germany, Spain<br />

and Belgium.<br />

������������������������������������������������������������������������������������������������<br />

�����������������������������������������������������������������������������������������������<br />

���� �������� �������� ��� ������� ������� �������� ���� ��������� ������� ���� ����� ������� ������� ���<br />

��������� ��� ���� ����� ����� ������� ��������� �������� �������� ������� ���� ��������� ����� ����� ��<br />

��������������������������������������������������������������������������������������������<br />

������������������������������������������������������������������������������������������������<br />

Report: ����������������������������������/����������������������������������108<br />

107


Benchmarking Equipment, Practices and Technologies<br />

� ��������������������������������<br />

Minerals Management Service (2004) report that the US Army Corps <strong>of</strong> Engineers uses an<br />

automated monitoring system referred to as “Silent Inspector” (SI). SI consists <strong>of</strong> a Dredge<br />

Specific System (DSS), a Ship Server, and a Shore Server. The DSS collects and displays<br />

standard information on dredging operations that is then transmitted to the Ship Server. Figure<br />

4-31 shows the DSS dredge display which gives real time sensor data including production rate,<br />

mixture velocity and density, pumping effort etc.<br />

Report No: 10/J/1/06/1309/0996<br />

�������������������������������������������������������������������������������������������������������<br />

������<br />

The Ship Server acts as the dredge based data archive and report creation centre as well as<br />

performs automated reviews <strong>of</strong> the data. The Shore Server is a larger system operated and<br />

maintained by the US Army Corps <strong>of</strong> Engineers. “Silent Inspector” provides information on<br />

dredge location history, quantity history, and status <strong>of</strong> a given project. It monitors all aspects <strong>of</strong><br />

operations from contract compliance to assurance that the operation is being performed in an<br />

environmentally safe manner (Rosati, 1999).<br />

108


Benchmarking Equipment, Practices and Technologies<br />

������ ����������������������������������<br />

� �������������������<br />

Spatial mitigation techniques are common in dredging projects worldwide, and work on similar<br />

principles to spatial mitigation in English areas – i.e. potential impacts are mitigated by reducing<br />

the total area <strong>of</strong> seabed dredged. Reducing the area dredged potentially mitigates the effects <strong>of</strong><br />

the dredging in a number <strong>of</strong> ways including:<br />

Report No: 10/J/1/06/1309/0996<br />

� reducing the area <strong>of</strong> direct removal <strong>of</strong> biomass<br />

� allows the majority <strong>of</strong> an area to be undredged, or undergoing recolonisation, at any<br />

particular time<br />

� reduces the area over which sediments are returned to the seabed<br />

� targeting sediments with low silt and clay reduces plumes<br />

� reduces conflict with other users <strong>of</strong> an area<br />

It is beyond the scope <strong>of</strong> this report to discuss all dredging projects worldwide where spatial<br />

mitigation has been used, however four recent case studies where spatial mitigation techniques<br />

were successfully used will be described.<br />

� �<br />

� �������������������������������������������������<br />

The Port <strong>of</strong> Melbourne is Australia’s largest container and general cargo port, and aims to<br />

expand its operations. The port’s expansion plan depends upon deepening the entrance to Port<br />

Phillip Bay, which also acts as a social, cultural and recreational asset (Bradford and Siebinga,<br />

2009). Port Phillip Bay includes two <strong>Marine</strong> National Parks, a Ramsar wetland and is a habitat for<br />

multiple fish species, little penguins, whales, dolphins, seals, various coldwater coral species and<br />

natural seagrass habitats and is used extensively for swimming, diving and boating.<br />

Bradford and Siebinga (2009) report that the main environmental impacts associated with<br />

dredging the 400,000 m 3 required related to turbidity and contaminated material. The concern<br />

was that dredging induced plumes within the Bay could potentially harm benthic organisms,<br />

seagrasses and fauna that depend on these habitats for food and protection. Trial dredging was<br />

conducted using the TSHD “Queen <strong>of</strong> the Netherlands” and showed that although the draghead<br />

was dredging efficiently, large amounts <strong>of</strong> sediment and rubble were being left on the seabed<br />

and were subsequently remobilized by currents and waves and deposited on flora and fauna<br />

living on a nearby deep canyon reef.<br />

To mitigate against these impacts a number <strong>of</strong> measures were put in place to minimise the<br />

possibility <strong>of</strong> sediments and rubble being transported into the canyon and impacting the marine<br />

life. These included spatial mitigations including:<br />

� When dredging towards the canyon the draghead had to be lifted so that no dredging<br />

occurred within 5 metres <strong>of</strong> the canyon edge.<br />

� When dredging the canyon edge itself only dredging towards the plateau was allowed.<br />

109


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

� When dredging the area closest closest to the Port Phillip Heads <strong>Marine</strong> National Park,<br />

a dredge zone was implemented so that a ridge was left in place until the remaining area<br />

had been dredged to design level. This ridge was then removed separately, after<br />

additional clean up <strong>of</strong> the area behind the ridge.<br />

Post-dredging video surveys demonstrated that the spatial mitigation measures put in place were<br />

effective in reducing the amounts <strong>of</strong> sediment and rubble deposited on the deep water reef<br />

(Bradford and Siebinga, 2009).<br />

� �������������������������������������������������<br />

Sand dredging was proposed for the Gulf <strong>of</strong> Mexico and the potential impact <strong>of</strong> operations on<br />

sea turtles was a significant concern. The US National <strong>Marine</strong> Fisheries Service published a<br />

Regional Biological Opinion for dredging in the Gulf <strong>of</strong> Mexico (NMFS, 2004) which designated a<br />

number <strong>of</strong> Hardground Buffer Zones. This stipulated that all dredging in potential aggregate<br />

areas should be designed to ensure that dredging did not occur within a minimum <strong>of</strong> 400 feet<br />

from any significant hardground areas or bottom structures that served as attractants to sea<br />

turtles for foraging or shelter (NMFS, 2004). Minerals Management Service (2004) reports that<br />

for the purposes <strong>of</strong> this spatial mitigation strategy only, significant hardgrounds in dredging areas<br />

were defined as areas that over a horizontal distance <strong>of</strong> 150 feet had an average elevation above<br />

the sand <strong>of</strong> 1.5 feet or greater, and had algae growing on them. It also required that pre-dredge<br />

seabed surveys were carried out at a sufficient scale to allow the hardgrounds to be mapped.<br />

� �������������������������������������������������<br />

The island <strong>of</strong> Vilufushi in the Maldives was partially destroyed by the tsunami <strong>of</strong> 2004, however a<br />

decision was taken by the government <strong>of</strong> the Maldives to reconstruct and increase the size <strong>of</strong> the<br />

island. This would be done through dredging coral sand from an area <strong>of</strong> the northern reef edge <strong>of</strong><br />

Vilufushi and pumping it directly into the area to be filled. This involved dredging and reclamation<br />

<strong>of</strong> 1,000,000 m 3 <strong>of</strong> coral sand from the reef to increase the surface and raise the height <strong>of</strong> the<br />

island (Bosschieter, 2007).<br />

The very rich biodiversity required protecting through monitoring and spatial mitigation<br />

procedures. To ensure that adverse effects were minimized environmental monitoring measured<br />

a range <strong>of</strong> parameters including turbidity, suspended solids concentrations (SSC) and dissolved<br />

oxygen levels in the water at 8 coral reef locations; sedimentation rates at two seagrass locations<br />

and potential erosion.<br />

The spatial mitigation involved the pre-dredging construction <strong>of</strong> bunds around the active dredging<br />

area, isolating it from the ocean. The dredge plume was therefore confined only to a minimised<br />

spatial area. The active dredge zone was subdivided into four compartments and the plume<br />

waters were made to flow through the four compartments before being discharged to the ocean.<br />

This allowed the suspended sediment to settle as much as possible, thus mitigating the effect <strong>of</strong><br />

releasing the plume into the ocean.<br />

110


Benchmarking Equipment, Practices and Technologies<br />

� ���������������������������<br />

The previous three case studies have described spatial mitigation through zoning (similar to the<br />

English practice) and bunding (where physical barriers are built). A further method <strong>of</strong> spatially<br />

mitigating the impact <strong>of</strong> overflow on the environment is to use a temporary physical barrier within<br />

the water column e.g. the use <strong>of</strong> silt curtains, silt screens and bubble curtains. Silt curtains and<br />

silt screens are flexible barriers that hang down from the water surface and both systems use a<br />

series <strong>of</strong> floats on the surface and a ballast chain or anchors along the bottom. Their use is<br />

generally limited to areas where maximum wave height is less than 1.5m and current speeds are<br />

less than 0.5 ms -1 .<br />

Report No: 10/J/1/06/1309/0996<br />

Silt curtains are made <strong>of</strong> impervious materials, such as coated nylon, while screens are made<br />

from synthetic geotextile fabrics, which allow water to flow through, but retain a large fraction <strong>of</strong><br />

the suspended solids inside the screened area (Averett et al., 1990) (Figure 4-32). Both systems<br />

can be kilometres in length and can be used to screen environmentally sensitive areas from the<br />

effects <strong>of</strong> plumes (USEPA, 1994).<br />

����������������������������������������������������������������������������������������������������������<br />

��������������������������<br />

Francingues and Palermo (2005) suggest that at depths greater than approximately 4-5 m loads<br />

and pressures on curtains and mooring systems become excessive and can result in failure <strong>of</strong><br />

standard construction materials. Because these barriers contain the sediment plume, they may<br />

temporarily increase sediment concentrations within the barrier compared with the concentrations<br />

that would occur if the barrier was absent (USEPA, 2005; Francingues and Thompson, 2006).<br />

Silt curtains have been used in dredging projects worldwide to protect the environment, with<br />

varying degrees <strong>of</strong> success. For example, silt curtains were ineffective in limiting sediment plume<br />

migration during dredging operations at New Bedford Harbour, USA, primarily because <strong>of</strong> tidal<br />

fluctuations and wind (Averett et al. 1990). Conversely a silt curtain was found to reduce<br />

suspended solids from approximately 400 mg/l (inside) to 5 mg/l (outside) during dredging in<br />

Halifax Harbour, Canada (USEPA 1994). Sydney Ports Corporation, Australia, has planned the<br />

construction and operation <strong>of</strong> a new container terminal, the Port Botany Expansion, where the<br />

dredging will take place both within, and outside, a range <strong>of</strong> silt curtains (Sydney Ports, 2009).<br />

111


Benchmarking Equipment, Practices and Technologies<br />

� ��������������������<br />

Temporal mitigation techniques have been used in dredging projects worldwide, and are based<br />

on similar principles to temporal mitigation in English areas – i.e. potential impacts are mitigated<br />

by restricting dredging or minimising sediment entering the water column during specific<br />

seasons, time frames or tidal states. Three case studies describing temporal mitigation measures<br />

(<strong>of</strong>ten described as Environmental Windows) will be described below.<br />

� ������������������������������������<br />

The Øresund Fixed Link is a bridge linking Denmark and Sweden, and was one <strong>of</strong> the largest<br />

infrastructure projects built in Europe and involved dredging in excess <strong>of</strong> 7 million m 3 <strong>of</strong> flint,<br />

limestone and clay till. The water in the Øresund has a low natural turbidity, resulting in good<br />

underwater visibility and allows species such as eelgrass to flourish (Jensen, 2006). Introduction<br />

<strong>of</strong> turbid plumes to the water column would therefore have a potentially major environmental<br />

impact.<br />

Report No: 10/J/1/06/1309/0996<br />

The overspill sediment during the project was limited to a maximum <strong>of</strong> 5% <strong>of</strong> the dredged<br />

quantity. In addition there were additional temporal limitations on the daily and weekly tonnage<br />

allowed as overspill in defined environmental areas during prescribed seasonal windows which<br />

were linked to the eelgrass growing season (March to November).<br />

� �������������������������������������������������������������������<br />

The idea <strong>of</strong> dredging during Environmental Windows was introduced in the USA in the National<br />

Environmental Policy Act in 1969 (Burt, 2002; Randall, 2006). This was reinforced by a US<br />

National Academy <strong>of</strong> Sciences (NAS) workshop in 2001, which resulted in a guidance document<br />

- “A Process for Setting, Managing and Monitoring Environmental Windows for <strong>Dredging</strong><br />

Projects” (NAS, 2001). The environmental window is a period set aside when no dredging is<br />

permitted to protect biological resources and habitats from the potential harmful effects <strong>of</strong> the<br />

dredging.<br />

The environmental window concept has been widely applied to a number <strong>of</strong> dredging projects in<br />

the United States, with about 90% <strong>of</strong> civil and maintenance dredging confined to specific times <strong>of</strong><br />

the year (Burt and Hayes, 2005). Randall (2006) reports that physical disturbance to habitat and<br />

nesting is the justification for more than 75% <strong>of</strong> the environmental windows in the US. Reine et<br />

al. (1998) suggested that potential detrimental impact to either individual or groups <strong>of</strong> sport and<br />

anadromous fishes and the protection <strong>of</strong> threatened and/or endangered species (e.g. sea turtles,<br />

marine mammals) were the justification for the majority <strong>of</strong> environmental windows affecting the<br />

USACE.<br />

Randall (2006) also indicated that the use <strong>of</strong> environmental windows increases the costs <strong>of</strong><br />

dredging and since the majority <strong>of</strong> environmental windows constrain dredging operations during<br />

spring and summer months (March-September) to avoid potential conflicts with biological<br />

activities such as migration, spawning, and nesting, <strong>of</strong>ten restrict dredging to winter months when<br />

weather conditions are most hazardous for dredging operations.<br />

112


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Concerns were raised by the USACE (Reine at al., 1998) about the use <strong>of</strong> environmental<br />

windows in the US suggesting that the technical justification for a window was <strong>of</strong>ten anecdotal. In<br />

addition, in some districts, environmental windows are in place during all seasons making it<br />

difficult to fulfil dredging requirements. An example is given in Figure 4-33 which depicts a<br />

“typical” pr<strong>of</strong>ile taken from a dredging project file in the New England District. In this example the<br />

dredging operation could not be completed within the remaining unrestricted period,<br />

necessitating an “exemption” during September through mid-November and the latter half <strong>of</strong><br />

January (Reine et al., 1998).<br />

�������������������������������������������������������������������������������������������������������<br />

��������������������������������������������<br />

Burt (2002) concludes therefore that the severe concept <strong>of</strong> environmental windows as applied in<br />

the United States would unreasonably restrict dredging operations if applied to Europe, where<br />

the majority <strong>of</strong> dredging operations take place all year round.<br />

113


Benchmarking Equipment, Practices and Technologies<br />

� �������������������<br />

� �����������������������������������<br />

One means <strong>of</strong> preventing the sediment plume associated with overflow is to prevent any overflow<br />

water from the dredging vessel entering the environment. One way <strong>of</strong> doing this is to contain all<br />

the dredged sand-water mixture within a vessel’s hopper. An example <strong>of</strong> this in practice is a<br />

dredging programme that took place at the Ketelmeer, in the Netherlands, in 2005 (Van der<br />

Linde et al., 2005). During this project overflowing was prohibited, therefore all the water-sand<br />

mixtures dredged were discharged into hopper barges with a capacity <strong>of</strong> approximately 800 m 3 .<br />

To ensure that no overflow occurred the barges were actually underfilled – containing<br />

approximately 600 m 3 (150 m 3 sand and 450 m 3 water) – before being transported to the<br />

discharging site in the mouth <strong>of</strong> the River Ijssel where the mixtures were pumped into a settling<br />

basin (Van der Linde et al., 2005). A similar method <strong>of</strong> preventing overflow from the TSHD<br />

entering the environment is to decant the overflow water to barges alongside the dredging<br />

vessel. These overflow waters can then be conducted to settling ponds and floculant added to<br />

remove the fine grained sediment.<br />

� ���������������������������������������������������������<br />

The Øresund Fixed Link is a road crossing between Denmark and Sweden, and the<br />

environmental concerns associated with its construction focused on the effects <strong>of</strong> blocking water<br />

flow between the Baltic and the North Sea, and the effect <strong>of</strong> the dredging and landfill activities on<br />

the local environment (Jensen, 2006). The EIA process indicated that the largest environmental<br />

impacts would be associated with dispersal <strong>of</strong> overspill sediment and a feedback monitoring<br />

programme was implemented as part <strong>of</strong> the environmental management system. Feedback<br />

monitoring is a form <strong>of</strong> surveillance monitoring where a few fast reacting environmental variables<br />

are forecast by modelling and then monitored continuously while the dredging is occurring<br />

(Netzband and Adnitt, 2009). The principles <strong>of</strong> feedback monitoring are illustrated in Figure 4-34.<br />

Report No: 10/J/1/06/1309/0996<br />

114


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

NO<br />

NO<br />

New work plan from the contractor<br />

including a spill scenario<br />

Environmental impact assessment <strong>of</strong> the<br />

work plan based on numerical modelling <strong>of</strong><br />

the spill scenario<br />

Are the operational criteria fulfilled?<br />

The dredging work starts<br />

Monitoring <strong>of</strong> selected variables<br />

Are the operational criteria exceeded?<br />

Intensified monitoring<br />

Numerical modelling<br />

Are the authorities criteria exceeded?<br />

�������������������������������������������������������������������<br />

The purpose <strong>of</strong> the feedback monitoring is to ensure that possible exceedence <strong>of</strong> environmental<br />

criteria can be forecast in such good time that dredging plans can be altered accordingly and<br />

costly down-time avoided (Netzband and Adnitt, 2009).<br />

In the Øresund example, integrated into the feedback monitoring programme were the outputs <strong>of</strong><br />

computer models which forecasted sediment dispersion, sedimentation, impact on water quality<br />

and impact on key habitats (such as eelgrass beds). Jensen (2006) states that for feedback<br />

monitoring it is essential that the variables measured meet certain demands:<br />

115<br />

YES<br />

NO<br />

YES


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

� Must have an unambiguous, easily measurable relationship to the organisms<br />

representing the ecosystem concerned;<br />

� Measurement result must be available in a short timescale (few days maximum);<br />

� Background data for determination <strong>of</strong> statistically relaiable limit values<br />

� Impact <strong>of</strong> various conditions should be calculable in advance.<br />

Therefore for monitoring <strong>of</strong> eelgrass in the Øresund case, shoot density, leaf and root biomass<br />

and carbohydrates in the rhizomes were the monitored variables and sediment concentration and<br />

sedimentation rates and total were measured weekly to give early warning data and validation for<br />

the computer model (Jensen, 2006). The use <strong>of</strong> indicators is a growing element <strong>of</strong> marine<br />

management and regulation (Gubbay, 2004). Bayer et al. (2008) report that their use is <strong>of</strong>ten<br />

masked by a range <strong>of</strong> synonymous terms e.g. standards, targets, thresholds, criteria—against<br />

which specific variables (i.e. indicators) are assessed .<br />

������ ����������������������<br />

A complete review <strong>of</strong> how dredging is regulated worldwide is beyond the scope <strong>of</strong> this report,<br />

however a comprehensive summary can be found in Sutton and Boyd (2009). Instead this report<br />

will examine four case studies, looking at transnational regulation (e.g. across Europe) and<br />

national regulations in the Netherlands, the USA and Australia.<br />

� ���������������������������������������������<br />

In the case <strong>of</strong> environmental law, countries <strong>of</strong> the European Union have agreed to transfer<br />

legislative and executive competancies to a transnational level. EU laws have three levels:<br />

� Framework Directives (Directive is equivalent to a law in national legislation) define a<br />

general approach which sets a number <strong>of</strong> boundary conditions and constraints and have<br />

to be implemented by each member state in accordance with its specific circumstances.<br />

� Directives are equivalent to laws and are binding on the member states, except for the<br />

fact that they first have to be “transposed” into national law.<br />

� Regulations are legal decisions taken at EU level that are binding as such for the member<br />

states and do not need transposition.<br />

(Mink et al., 2006)<br />

In addition to EU law there are also International Conventions and Treaties which may influence<br />

regulation and take priority <strong>of</strong> EU laws e.g. the London Convention and the Oslo-Paris (OSPAR)<br />

Convention for the Atlantic and the North Sea. The London Convention is a global convention<br />

and member countries agree to introduce legislation in their own countries to implement the<br />

Convention. Its aim is to protect the marine environment and although it is limited to nonterritorial<br />

waters many signatories choose to apply it to their territorial waters, including estuaries<br />

(Burt and Hayes, 2005). It defines the nature, temporal and spatial scales and duration <strong>of</strong><br />

expected impacts based on reasonably conservative assumptions. OSPAR functions in a similar<br />

116


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

way to the London Convention but only applies to countries bordering the North Sea and North<br />

East Atlantic.<br />

EU law does not deal specifically with dredged material, however a number <strong>of</strong> Directives have an<br />

impact on dredging. For maintenance and capital dredging the Waste Framework Directive is<br />

highly important since the dredged sediment may potentially be defined as waste. “Waste” is<br />

defined as “any substance or object which the holder discards or intends to discard” and the EC<br />

argues that dredged sediment is “waste” since the holder needs to dispose <strong>of</strong> it in some way.<br />

This obviously has less impact on the marine aggregate dredging industry where the dredged<br />

sediment is a commercially valuable resource. The EuDA Environment Committee concludes<br />

that for dredging in marine waters current EU law may only be relevant, and applicable to marine<br />

aggregate dredging where there is a requirement to dispose <strong>of</strong> dredged material on land when<br />

the Landfill Directive may come into effect.<br />

An EU Directive that may eventually impact all dredging operations, including marine aggregate<br />

dredging, is the Water Framework Directive. Its goal is to gradually improve the quality <strong>of</strong><br />

European waters to some standard which may be called “good”. This is a long-term goal, and it is<br />

recognised that water quality varies considerably over time and as a result <strong>of</strong> many factors. How<br />

severely the Directive will impact dredging will depend on the consideration <strong>of</strong> this temporal<br />

variability – for instance will short term increases in suspended sediment due to dredging matter<br />

if average values are within defined limits?<br />

The Habitats and Birds Directives aim to protect biodiversity and rare biotopes and species and<br />

their implementation process has led to the establishment <strong>of</strong> Natura 2000 sites across Europe.<br />

These consist <strong>of</strong> Special Areas <strong>of</strong> Conservation (SACs) under the Habitats Directive and Special<br />

Protection Areas (SPAs) under the Birds Directive. These may impact marine aggregate<br />

dredging since dredging areas may be located close to or in these Natura 2000 sites which<br />

imposes restrictions on operation or obstacles to permitting <strong>of</strong> new Licences.<br />

Finally the EC-published Thematic Strategy on the Protection and Conservation <strong>of</strong> the <strong>Marine</strong><br />

Environment (see http://ec.europa.eu/environment/water/marine.htm) may have implications for<br />

dredging in the future since it aims to achieve “good environmental status” <strong>of</strong> European marine<br />

waters by 2021 and also claims competence to regulate the status <strong>of</strong> the seabed and its subsoils.<br />

Mink et al. (2006) suggest that by analogy with the Water Framework Directive “good<br />

environmental status” is likely to be defined on the basis <strong>of</strong> parameters including physical and<br />

chemical conditions, biological and ecological processes and physiographic and geographic<br />

factors.<br />

� ����������������<br />

<strong>Marine</strong> aggregate dredging in the Netherlands is regulated according to the Sediment Extraction<br />

Law, amended in 1997, and Licences are granted by Ministry <strong>of</strong> Transport, Public Works, and<br />

Water Management; the Directorate�General <strong>of</strong> Public Works and Water Management, and the<br />

North Sea Directorate. Sutton and Boyd (2009) indicate that from 2004 a distinction was made by<br />

the Dutch Government between licences that allowed extraction <strong>of</strong> < 10 million m 3 per licence<br />

and those that allowed extraction <strong>of</strong> > 10 million m 3 . For the smaller-scale extractions maximum<br />

117


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

extraction depth was set at 2 m while larger�scale extractions allowed an extraction depth <strong>of</strong><br />

more than 2 m if this was a preferred option in an EIA.<br />

Not all dredging applications in require an EIA – an EIA only has to be produced when an<br />

extraction exceeds an area <strong>of</strong> 500 ha (EEZ) or 100 ha (territorial sea) and/or exceeds a volume<br />

<strong>of</strong> 10 million m 3 . In addition the landward limit for extraction <strong>of</strong> marine sediments is the<br />

established NAP 20 m depth contour, defined by the NAP (Dutch Ordnance Level ~ Mean Sea<br />

Level). Seaward <strong>of</strong> this contour, extraction is allowed in principle (Sutton and Boyd, 2009).<br />

� ��������<br />

The major US statutes and laws that relate to dredging activities are mainly concerned with<br />

disposal <strong>of</strong> dredged material within United States waters, rather than dredging for marine<br />

aggregates (Randall, 2006) and regulation <strong>of</strong> disposals is a shared responsibility between the US<br />

Army Corps <strong>of</strong> Engineers (USACE) and the Environmental Protection Agency. The Outer<br />

Continental Shelf Act 1983 (amended in 1994), however, allows for the leasing <strong>of</strong> areas <strong>of</strong> the<br />

shelf for sand and gravel extraction. The agencies responsible for these activities would be the<br />

USACE and the Minerals Management Service. Any extraction <strong>of</strong> marine sand and gravel would<br />

need to take into account the Sustainable Fisheries Act (1996), which requires the National<br />

<strong>Marine</strong> Fisheries Service to define essential habitats for various commercial species and all<br />

federal agencies must consult the National <strong>Marine</strong> Fisheries Service on any action that may<br />

adversely affect essential fish habitats.<br />

<strong>Dredging</strong> <strong>of</strong> <strong>of</strong>fshore sediment for beach nourishment is relatively common in the United States<br />

however the Minerals Management Service has decided not to proceed with the designation and<br />

leasing <strong>of</strong> <strong>of</strong>fshore areas for marine aggregate mining. Thus the US currently has only two<br />

commercial marine aggregate dredging operations, neither <strong>of</strong> which is <strong>of</strong>fshore. Since 1985 there<br />

has been a licence to dredge aggregates from the entrance to New York Harbour, producing<br />

between 1.2 and 1.9 million m 3 <strong>of</strong> sand and gravel per year. The company responsible has also<br />

attempted to gain a licence to dredge a large sand and gravel deposit located in Federal waters<br />

<strong>of</strong>fshore <strong>of</strong> northern New Jersey, however public opposition to the project has prevented the<br />

Minerals Management Service from allowing this. The only other aggregates mining is on Lake<br />

Erie and produces about 306,000 m 3 <strong>of</strong> aggregates annually, using a TSHD. These small<br />

volumes <strong>of</strong> aggregate compare with a total volume dredged in the US <strong>of</strong> between 200 and 250<br />

million m 3 a year.<br />

A further legislative influence on dredging in the US is the “Jones Act” (the 1920 Merchant <strong>Marine</strong><br />

Act). Dredgers working in the US must comply with the “Jones Act” which is a cabotage law<br />

(cabotage is the transport <strong>of</strong> goods or passengers between two points in the same country). The<br />

“Jones Act” requires that all goods transported by water between US ports be carried in USflagged<br />

ships, constructed in the US, owned by US citizens, and at least 75% <strong>of</strong> the crew must<br />

be US citizens. Operating under the “Jones Act” increases the costs associated with shipping the<br />

product carried.<br />

Every dredger operating within the US is therefore subject to the Jones Act and this has led to<br />

the US being a closed market. It is not subject to foreign competition, and approximately 300<br />

118


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

dredging companies operate with an outdated fleet <strong>of</strong> mainly small vessels (the largest TSHD is<br />

the USACE’s vessel "Wheeler", which has a capacity <strong>of</strong> 8000 m 3 ).<br />

� ����������<br />

Unlike the English, Dutch and (to a lesser extent) American examples, the Australian marine<br />

aggregate dredging industry is in its infancy. However the sector is likely to expand in the near<br />

future due to diminishing land-based aggregate resources close to the major conurbations<br />

(Littleboy and Boughen, 2007; Johns, 2008). Because <strong>of</strong> this anticipated expansion <strong>of</strong> the sector<br />

CSIRO has undertaken a review <strong>of</strong> potential operations as part <strong>of</strong> National Research Flagships<br />

Wealth From the Oceans program (Johns, 2008).<br />

A number <strong>of</strong> different acts, guidelines and policies comprise Australian legislation affecting<br />

mining <strong>of</strong> <strong>of</strong>fshore minerals:<br />

� Seas and Submerged Lands Act 1973, as amended by the Maritime Legislation<br />

Amendment Act 1994<br />

� Commonwealth Offshore Minerals Act 1994<br />

� State/Northern Territory Offshore Minerals Acts 1998, 1999, 2000, and 2003<br />

� Native Title Act 1993<br />

� Environmental Protection and Biodiversity Conservation Act 1999<br />

� Guidelines on the Application <strong>of</strong> the Environment Protection and Biodiversity Conservation<br />

Act 1999 to Interactions Between Offshore Seismic Operations and Larger Cetaceans<br />

October 2001<br />

The Seas and Submerged Lands Act 1973 (SSLA) provides legislation which relates to<br />

Australia’s sovereignty over defined waters <strong>of</strong> the sea, airspace over, and the seabed and subsoil<br />

beneath those waters, and to sovereign rights relating to the continental shelf and the recovery <strong>of</strong><br />

minerals (other than petroleum) from the above listed areas. Defined waters <strong>of</strong> the sea involve<br />

those <strong>of</strong> the state/territory coastal waters, territorial sea, contiguous zone, exclusive economic<br />

zone, and continental shelf.<br />

The Commonwealth <strong>of</strong> Australia Offshore Minerals Act 1994 (OMA), along with six other<br />

associated Acts provide the legal framework for the exploration and recovery <strong>of</strong> minerals, other<br />

than petroleum, on Australia’s continental shelf, including the payment <strong>of</strong> royalties, annual fees<br />

and user charges for licences. The OMA establishes a code for the development <strong>of</strong> <strong>of</strong>fshore<br />

minerals in the area under the Australian Government (Commonwealth) jurisdiction beyond the<br />

outer limit <strong>of</strong> the States and Northern Territory coastal waters (i.e. > 3 nm). There is no<br />

Commonwealth involvement in the administration <strong>of</strong> <strong>of</strong>fshore mineral activities within the State or<br />

Northern Territory jurisdiction (i.e. < 3 nm). Under the OMA, a mineral is defined as a naturally<br />

occurring substance or mixture <strong>of</strong> substances, which may be in the form <strong>of</strong> sand, gravel, clay,<br />

limestone, rock, evaporates, shale, oil-shale, or coal; but does not include petroleum.<br />

The Act covers 5 types <strong>of</strong> licences: Exploration, Retention, Mining, Works, and Special Purpose<br />

Consent Licences. An Exploration Licence (4 years) enables the holder to explore for all minerals<br />

119


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

and may be renewed for up to 10 years, but the licence area must be reduced by 50% on each<br />

renewal. A Retention Licence (up to 5 years) allows an Exploration Licence holder to retain rights<br />

where a significant mineral deposit exists but is not currently commercially viable. A Mining<br />

Licence (up to 21 years) enables the holder to explore for and recover minerals over an area<br />

where a significant mineral deposit has been identified and evaluated. Before a Mining Licence is<br />

granted, the applicant must lodge a mining development plan and environmental impact<br />

statement for assessment and approval. A Works Licence (up to 5 years) enables a licence<br />

holder to carry out major licence-related operations outside the licence area <strong>of</strong> his exploration,<br />

retention or mining licence. A Special Purpose Consent title (up to 1 year) allows the holder to<br />

conduct scientific investigation, reconnaissance surveys, or small mineral sample collection.<br />

The State/Northern Territory Offshore Minerals Acts 1998, 1999, 2000, and 2003 allow each<br />

State and the Northern Terrirtory to have its own Offshore Minerals Act which provide legislation<br />

for the exploration and mining <strong>of</strong> minerals (other than petroleum) within the first 3 nm <strong>of</strong> the<br />

territorial sea.<br />

The Native Title Act 1993 (NTA) provides the legal basis for the recognition and protection <strong>of</strong><br />

native title. Native title represents the rights and interests (i.e. hunting, gathering, or fishing) <strong>of</strong><br />

Aboriginal and Torres Strait Islander people in relation to land and waters according to their<br />

traditional laws and customs, which are recognised under Australian law. Under the NTA,<br />

exploration and mining companies are required to negotiate access arrangements for land on<br />

which native title has been granted or is subject to claim. During this process many Indigenous<br />

communities have become important stakeholders in mineral resource evaluation and<br />

development. The NTA requires that when the Government intends to grant an Exploration or<br />

Mining Licence, it must provide proper notification including advertising its intention to grant the<br />

licence in a major newspaper that circulates in the area to which the notice relates and an<br />

Aboriginal newspaper.<br />

The Environmental Protection and Biodiversity Act 1999 (EPBC) provides the legal basis for the<br />

protection <strong>of</strong> the environment and conservation <strong>of</strong> heritage and aspects <strong>of</strong> national environmental<br />

significance. The EPBC specifies the process for environmental assessment and approvals,<br />

environmental conservation and management. The EPBC also promotes ecologically sustainable<br />

development; conservation <strong>of</strong> biodiversity; cooperation between governments, communities,<br />

land-holders, and indigenous peoples; and recognition <strong>of</strong> the conservation role and knowledge <strong>of</strong><br />

indigenous peoples to the sustainable use <strong>of</strong> Australia’s biodiversity.<br />

In all Australian waters, including state and territory waters, the EPBC regulates actions that will<br />

have, or are likely to have, a significant impact on the environment <strong>of</strong> any listed threatened and<br />

migratory species (including cetaceans and other marine species) in a Commonwealth marine<br />

area. <strong>Marine</strong> species, not including whales and other cetaceans, include: sea-snakes, seals,<br />

crocodiles, dugong, turtles, seahorses, sea-dragons and pipefish, and birds which occur naturally<br />

in Commonwealth marine areas.<br />

Guidelines on the Application <strong>of</strong> the Environment Protection and Biodiversity Conservation Act<br />

1999 to Interactions Between Offshore Seismic Operations and Larger Cetaceans October 2001<br />

were drawn up to ensure potential impacts on whales from industry-related activities (particularly<br />

120


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

seismic surveys) are avoided. A proposed seismic operation would therefore require the approval<br />

<strong>of</strong> the Minister for Environment and Water Resources if it is regarded as being likely to have a<br />

significant impact on cetacean/whale species (including threatened and migratory cetacean<br />

species).<br />

121


Benchmarking Equipment, Practices and Technologies<br />

��� �����������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

Benchmarking is a qualitative process looking at an objective and comparing alternative ways <strong>of</strong><br />

achieving it measured in differing ways e.g. cost, efficiency, environmental impact, practicality,<br />

etc. The following chapter will benchmark the technologies and practices identified in Chapter 4<br />

against the English example. It will compare the technologies and practices in use in English<br />

aggregate dredging to those used elsewhere in the world. It will rate each option under the<br />

categories ‘Best Practice’, ‘Fit for Purpose’ and ‘Sub-optimal’ as they apply to the English<br />

situation.<br />

122


Report No: 10/J/1/06/1309/0996<br />

123<br />

������ �����������<br />

���������� �������� ������� ������� ����������������<br />

Dragheads � Single visor<br />

Dredge<br />

pumps and<br />

impellers<br />

� California<br />

� No jetting or<br />

cutting<br />

� Standard<br />

pump /<br />

impeller<br />

� Cutting teeth<br />

and jetting<br />

common<br />

� High<br />

efficiency<br />

pump /<br />

impeller<br />

� Punaise<br />

system<br />

Fit for<br />

purpose<br />

Fit for<br />

purpose<br />

�������������<br />

�����������<br />

� � English draghead technology is unadvanced compared with<br />

worldwide technologies, however it is rated as being fit for<br />

purpose.<br />

� Cutting teeth and jetting are common in dragheads on most new<br />

build TSHDs however many <strong>of</strong> these vessels are used in capital<br />

or maintenance dredging where soils are <strong>of</strong>ten consolidated,<br />

unlike English aggregate deposits.<br />

� Other speciality dragheads identified as part <strong>of</strong> this project (wide<br />

dragheads, focussed suction, semi autonomous) are rated as<br />

sub-optimal and inapplicable to the English aggregate dredging<br />

industry.<br />

� � Many TSHDs worldwide use high efficiency pumps and<br />

impellers compared with standard pumps and impellers at use in<br />

the English fleet.<br />

� High efficiency pumps and impellers could improve loading rates<br />

by 10% or more, however since much <strong>of</strong> the English industry is<br />

governed by tidal turnaround the time saved would not allow<br />

increased cargo collection.<br />

� Faster loading could however reduce the environmental impacts<br />

<strong>of</strong> dredging by reducing the amount <strong>of</strong> time the vessel was<br />

actively impacting the dredging area<br />

� The Punaise system is rated sub-optimal and inapplicable to the<br />

Benchmarking Equipment, Practices and Technologies


Report No: 10/J/1/06/1309/0996<br />

124<br />

Dredge pipe � Single pipe � Single pipe<br />

Screening � Fixed screens<br />

Overflow<br />

control<br />

� Screening<br />

towers<br />

� Generally<br />

none<br />

� Moveable<br />

spillways<br />

� Twin pipe<br />

� Central pipe<br />

� Fixed<br />

screens<br />

� Screening<br />

towers<br />

� Moveable<br />

spillways<br />

� Anti-turbidity<br />

valve<br />

Best<br />

practice<br />

Best<br />

practice<br />

Sub<br />

optimal /<br />

fit for<br />

purpose<br />

English aggregate industry.<br />

X � While single pipes are the most common arrangement in the<br />

world fleet a number <strong>of</strong> vessels have twin pipes. This allows<br />

faster loading <strong>of</strong> the hopper.<br />

� Since much <strong>of</strong> the English industry is governed by tidal<br />

turnaround the time saved would not allow increased cargo<br />

collection.<br />

� Faster loading could however reduce the environmental impacts<br />

<strong>of</strong> dredging by reducing the amount <strong>of</strong> time the vessel was<br />

actively impacting the dredging area<br />

� Central pipes would be sub-optimal for the English industry<br />

since they impact on hopper capacity<br />

� � Where screening is carried out the screen arrangement <strong>of</strong> fixed<br />

screens or screening towers are used. No novel technologies<br />

with regards screening were identified.<br />

� The use <strong>of</strong> fixed screens is fit for purpose while rotating screen<br />

towers are considered best practice.<br />

� Screening technology is in excess <strong>of</strong> 20 years old –<br />

investigation <strong>of</strong> latest land based screening techniques is<br />

recommended<br />

� � The majority <strong>of</strong> English aggregate vessels have no overflow<br />

control while a small number <strong>of</strong> vessels have moveable<br />

spillways<br />

� Anti-turbidity valve technology has been available for a number<br />

Benchmarking Equipment, Practices and Technologies


Report No: 10/J/1/06/1309/0996<br />

125<br />

Vessel<br />

design<br />

� Aft or forward<br />

config.<br />

� Typically no<br />

bulb on bow.<br />

� Relatively<br />

large crew<br />

numbers<br />

� Green pipe<br />

recirc.<br />

system<br />

� Aft or<br />

forward<br />

config.<br />

� New hulls<br />

have bulb<br />

on bow<br />

� Move<br />

towards<br />

smaller crew<br />

numbers<br />

� New<br />

designs<br />

include<br />

reduced<br />

drag<br />

measures<br />

including<br />

Sub<br />

optimal /<br />

Fit for<br />

purpose<br />

<strong>of</strong> years and is efficient, having been shown to reduce turbidity<br />

to 40% <strong>of</strong> the standard method. Anti-turbidity valves are rated as<br />

Best Practice for overflow control.<br />

� The green pipe recirculating system has also been shown to<br />

reduce surface turbidity and hence the sediment plume.<br />

However the same amount <strong>of</strong> sediment is returned to the<br />

environment and concerns have been expressed that it may<br />

increase bed turbidity currents and the size <strong>of</strong> the sediment<br />

footprint and is hence not applicable to the English industry.<br />

� � Vessels in the English aggregate fleet are old compared with the<br />

fleets <strong>of</strong> the major European dredging contractors.<br />

� Designs <strong>of</strong> English ships do not incorporate the latest<br />

technologies to reduce drag and improve hull efficiency and fuel<br />

comsumption.<br />

� There is a move on the newest European dredging vessels to<br />

reduce crew numbers<br />

� The design <strong>of</strong> vessels in the English fleet is considered sub-<br />

optimal when compared with the design <strong>of</strong> the newest TSHDs<br />

being built, but is fit for purpose for dredging in English licence<br />

areas.<br />

� There is, however, major scope for development in the design <strong>of</strong><br />

English aggregate vessels when new builds are commissioned.<br />

A recently commissioned ALSF study (MEPF REF 09/P133) will<br />

review vessel design and identify key issues including optimum<br />

operation and quantify potential reductions in environmental<br />

Benchmarking Equipment, Practices and Technologies


Report No: 10/J/1/06/1309/0996<br />

126<br />

Dry<br />

Discharge<br />

� Shore<br />

discharge<br />

� Drag scrapers<br />

� Bucket wheel<br />

system<br />

� Grab<br />

discharge<br />

anti-fouling<br />

paints; gel<br />

coats; high<br />

quality hull<br />

welding,<br />

propeller<br />

pods etc.<br />

� Pump out Best<br />

practice<br />

������ ��������������������������������������������������������<br />

� �������� ������� ������� ����������������<br />

�������������<br />

Positioning � DGPS<br />

� DGPS<br />

Best<br />

practice<br />

impacts for any future new build vessels.<br />

� � No novel technologies with regards to dry discharging were<br />

identified.<br />

� Shore discharge is considered sub-optimal, drag scrapers are fit<br />

for purpose while bucket wheel systems / grab discharge are<br />

considered best practice<br />

� There may be further opportunities for development by<br />

examining land-based technologies and adapting suitable ones<br />

for TSHDs<br />

�����������<br />

X � DGPS navigation systems are common to most modern<br />

dredging projects worldwide.<br />

Benchmarking Equipment, Practices and Technologies


Report No: 10/J/1/06/1309/0996<br />

127<br />

Navigation<br />

and<br />

electronic<br />

sensors<br />

Electronic<br />

Monitoring<br />

Spatial<br />

mitigation<br />

� AIS / ECDIS<br />

� Draghead<br />

<strong>of</strong>fset from<br />

DGPS<br />

position<br />

� Pipe angle<br />

indicator<br />

� AIS / ECDIS<br />

� Draghead<br />

position<br />

� Draghead<br />

angle and<br />

cutting<br />

depth<br />

� Mixture<br />

density<br />

� Flow<br />

velocity<br />

� Vacuum<br />

� Production<br />

rate<br />

� Draghead<br />

steering<br />

� Real time<br />

bathy<br />

Sub<br />

optimal /<br />

fit for<br />

purpose<br />

� EMS � EMS Best<br />

practice<br />

Best<br />

practice<br />

� � Linkage <strong>of</strong> navigation data to an electronic display<br />

information system is common to most modern dredging<br />

projects worldwide<br />

� The English dredging industry is sub-optimal with regards<br />

to dredging sensor information and draghead steering<br />

� Many modern TSHDs are fitted with a much wider range <strong>of</strong><br />

electronic sensors and information displays than vessels in<br />

the English fleet<br />

X � The use <strong>of</strong> a tamper pro<strong>of</strong> electronic monitoring system is<br />

common, particularly in Europe and the United States.<br />

� � There is a commitment by English dredging companies to<br />

relinquish unproductive license areas.<br />

� <strong>Dredging</strong> within English licence areas is controlled through<br />

voluntary zoning to minimize the area dredged at any one<br />

Benchmarking Equipment, Practices and Technologies


Report No: 10/J/1/06/1309/0996<br />

128<br />

Temporal<br />

mitigation<br />

Best<br />

practice<br />

time.<br />

� Communication with other users and voluntary exclusion<br />

zones to minimize conflict with fishermen are routine within<br />

the English dredging industry.<br />

� Similar means <strong>of</strong> spatial mitigation are also used on<br />

dredging projects worldwide.<br />

� The use <strong>of</strong> spatial mitigation measures by the English<br />

industry is considered best practice when compared with<br />

world wide practices, however there is continued<br />

opportunity for development if knowledge and<br />

understanding continues to grow.<br />

� � English dredging industry mitigates against environmental<br />

impacts on vulnerable species by reducing or avoiding<br />

screening; or prohibiting dredging entirely, on certain<br />

licence areas during particular time frames.<br />

� Similar temporal mitigation has been used worldwide to<br />

protect species or habitats vulnerable at certain times <strong>of</strong><br />

the year.<br />

� Environmental Windows as widely used in the USA<br />

severely restrict dredging operations and would<br />

unreasonably restrict aggregate dredging operations if<br />

applied in English waters. US style Environmental<br />

Windows are therefore considered sub-optimal in the<br />

English setting.<br />

� The use <strong>of</strong> temporal mitigation measures by the English<br />

Benchmarking Equipment, Practices and Technologies


Report No: 10/J/1/06/1309/0996<br />

129<br />

Knowledge<br />

and<br />

understandi<br />

ng<br />

Best<br />

practice<br />

Permitting Best<br />

practice<br />

industry is considered best practice when compared with<br />

world wide practices, however there is continued<br />

opportunity for development if knowledge and<br />

understanding continues to grow.<br />

� � <strong>Marine</strong> aggregate industry in the UK has invested<br />

significant sums <strong>of</strong> money in increasing knowledge and <strong>of</strong><br />

the marine environment and enhancing understanding <strong>of</strong><br />

the impacts <strong>of</strong> aggregate dredging.<br />

� It has done this through internal research projects, the<br />

Environmental Impact Assessment process and most<br />

significantly through the <strong>Aggregate</strong> Levy Sustainability<br />

Fund (ALSF).<br />

� Worldwide knowledge is not so advanced, although other<br />

research programmes exist such as the DOER programme<br />

in the US and Australian CSIRO research projects<br />

� It is considered that the knowledge and understanding<br />

developed within the English system is world-leading,<br />

although further opportunities to develop and maintain this<br />

position do exist.<br />

X � The English regulatory process is transparent, puts the<br />

environment first, and involves significant consultation.<br />

� An environmental statement and supporting studies are<br />

required in al cases and unless the applicant can<br />

demonstrate that the environmental impacts are<br />

acceptable then a dredging permission will not be granted.<br />

Benchmarking Equipment, Practices and Technologies


Report No: 10/J/1/06/1309/0996<br />

130<br />

� Consents are accompanied by a schedule <strong>of</strong> legally<br />

enforceable conditions<br />

� Regulation worldwide differs between countries, however<br />

examples from Europe (Netherlands) showed that, unlike<br />

the UK, not all dredging applications required an EIA<br />

� US regulations are concerned mainly with disposal <strong>of</strong><br />

dredged material rather than aggregate dredging.<br />

Responsibilities for regulation are split between the<br />

USACE and the Environmental Protection Agency.<br />

� Australian industry is evolving and the regulatory<br />

framework is similarly evolving. Currently regulation is<br />

complex with 6 Acts, Guidelines and Policies comprising<br />

the Australian legislation.<br />

� The English regulatory process is considered best practice<br />

when compared with examples worldwide.<br />

Benchmarking Equipment, Practices and Technologies


Benchmarking Equipment, Practices and Technologies<br />

����� ������������<br />

Report No: 10/J/1/06/1309/0996<br />

Within the English fleet there are a number <strong>of</strong> old vessels which are reaching the end <strong>of</strong> their<br />

serviceable lives, which is determined by a combination <strong>of</strong> the ability to maintain the machinery in<br />

a reliable condition and the cost <strong>of</strong> maintaining the vessel’s structure which escalates significantly<br />

beyond 30 years old. Operators are therefore going to be faced the dilemma <strong>of</strong> developing cost<br />

effective vessel designs that can continue to serve the existing wharves or alternatively find new<br />

ways <strong>of</strong> supplying the markets that the wharves served. There is may be pressure to move<br />

operations into deeper, more exposed waters, particularly in the North East and the West Coast.<br />

There will also be continued environmental demands imposed on the industry, some <strong>of</strong> which will<br />

provide economic benefits.<br />

From the benchmarking process it is concluded that in terms <strong>of</strong> technology the English fleet is<br />

generally fit for purpose when compared with the world fleet:<br />

� Dragheads – fit for purpose<br />

� Dredge pumps and impellers – fit for purpose<br />

� Dredge pipes – best practice<br />

� Screening – best practice<br />

� Overflow control – sub-optimal / fit for purpose<br />

� Vessel design – sub-optimal / fit for purpose<br />

� Dry discharge – best practice.<br />

The English fleet is older than comparable fleets operated by the major European dredging<br />

contractors and, therefore, it is concluded that while many <strong>of</strong> the technologies are fit for purpose,<br />

or even current best practice, there are opportunities for development in the fields <strong>of</strong> dragheads;<br />

dredge pumps and impellers; screening, overflow control, vessel design and discharge.<br />

Compared with the technological side the benchmarking process has shown that in terms <strong>of</strong><br />

dredging practices, mitigation and Regulatory Framework the English example is generally worldleading:<br />

� Positioning – best practice<br />

� Navigation and electronic sensors – sub-optimal / fit for purpose<br />

� Electronic monitoring – best practice<br />

� Spatial mitigation – best practice<br />

� Temporal mitigation – best practice<br />

� Knowledge and understanding – best practice<br />

� Permitting – best practice<br />

There is a distinct difference between the navigation information relating to drag head position,<br />

steering and electronic dredging sensors employed in the English fleet and those used on the<br />

131


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

newest build comparable TSHDs. However in all other regards the benchmarking process<br />

considers the English examples to be best practice. There are, however, further opportunities for<br />

development in the fields <strong>of</strong> navigation and electronic sensors and knowledge and<br />

understanding; and should knowledge and understanding <strong>of</strong> impacts continue to grow then there<br />

would be opportunity to further enhance spatial and temporal mitigation practices.<br />

The overall conclusion <strong>of</strong> the benchmarking exercise is that having surveyed the world fleet,<br />

technology and practices it is not possible to identify major improvements that are applicable to<br />

the English industry and that would significantly enhance the environmental performance <strong>of</strong> the<br />

English fleet. The English industry is, however, open with regards to its environmental reporting<br />

and willing to develop knowledge on enhancing environmental performance.<br />

132


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

133


Benchmarking Equipment, Practices and Technologies<br />

�����������<br />

Report No: 10/J/1/06/1309/0996<br />

Austen MC, Hattam C, Lowe S, Mangi SC, and Richardson K, (2009). Quantifying and Valuing<br />

the <strong>Impacts</strong> <strong>of</strong> <strong>Marine</strong> <strong>Aggregate</strong> Extraction on Ecosystem Goods and Services. MALSF Report,<br />

MEPF 08-77, 72pp.<br />

Averett DE, Perry BD, Torrey EJ, and Miller JA, (1990). Review <strong>of</strong> removal, containment and<br />

treatment technologies for remediation <strong>of</strong> contaminated sediment in the Great Lakes.<br />

Miscellaneous Paper EL-90-25. US Army Engineer Waterways Experiment Station, Vicksburg,<br />

MS.<br />

Bayer E, Barnes RA, and Rees HL, (2008). The regulatory framework for marine dredging<br />

indicators and their operational efficiency within the UK: a possible model for other nations? ICES<br />

Journal <strong>of</strong> <strong>Marine</strong> Science, 65, 1402-1406.<br />

Bernard W, (1978). Prediction and control <strong>of</strong> dredged material dispersion around dredging and<br />

openwater pipeline disposal operations. Technical report, Dredged Material Research Program.<br />

USWES Environmental Laboratory Vicksburg.<br />

Bert Visser’s Directory <strong>of</strong> Dredgers (2009). http://www.dredgers.nl/. Accessed 20/05/09.<br />

Blake NJ, Doyle LJ, and Culter JJ, (1996). <strong>Impacts</strong> and direct effects <strong>of</strong> sand dredging for beach<br />

renourishment on the benthic organisms and geology <strong>of</strong> the West Florida Shelf, Final Report. US<br />

Department <strong>of</strong> the Interior, Minerals Management Service, Office <strong>of</strong> International Activities and<br />

<strong>Marine</strong> Minerals, Herndon, VA. OCS Report MMS 95-0005, 109 pp.<br />

Bins AK, (1988). Minimum air entrainment velocity <strong>of</strong> a vertical plunging liquid jets. Chem. Eng.<br />

Sc., 43, (2), pp 379-543.<br />

BMAPA (2006). Strength From The Depths: A Sustainable Development Strategy For The British<br />

<strong>Marine</strong> <strong>Aggregate</strong> Industry. BMAPA, London.<br />

BMAPA (2009). Strength From The Depths: Third Sustainable Development Report For The<br />

British <strong>Marine</strong> <strong>Aggregate</strong> Industry. BMAPA, London.<br />

Boskalis Offshore bv, (2009). Balgzand - Bacton Pipeline (BBL). A gas pipeline from Balgzand<br />

(The Netherlands) to Bacton (UK).<br />

http://www.boskalis.com/media/downloads/en_US/Projects/Europe/Netherlands%20-<br />

%20Balgzand%20Bacton.pdf. Accessed 11/11/09.<br />

Bosschieter C, (2007). Environmental monitoring for the reconstruction <strong>of</strong> Vilufushi, Maldives.<br />

Terra et Aqua, 109, 14-22.<br />

Boyd SE, and Rees HL, (2003). An examination <strong>of</strong> the spatial scale <strong>of</strong> impact on the marine<br />

benthos arising from marine aggregate extraction in the Central English Channel. Estuarine,<br />

Coastal and Shelf Science, 57: 1 – 16.<br />

Boyd SE, Rees HL, Vivian CMG, and Limpenny DS (2003a). Review <strong>of</strong> current state <strong>of</strong><br />

knowledge <strong>of</strong> the impacts <strong>of</strong> marine aggregate extraction – a UK perspective. In European<br />

<strong>Marine</strong> Sand and Gravel – Shaping the Future, EMSAGG Conference, 21–22 February 2003,<br />

Delft University, the Netherlands. 5 pp.<br />

134


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

Boyd SE, Limpenny DS, Rees HL, Cooper KM, and Campbell S, (2003b). Preliminary<br />

observations <strong>of</strong> the effects <strong>of</strong> dredging intensity on the recolonization <strong>of</strong> dredged sediments <strong>of</strong>f<br />

the south�east coast <strong>of</strong> England (Area 222). Estuarine, Coastal and Shelf Science, 57: 209 –<br />

223.<br />

Boyd SE, Cooper KM, Limpenny DS, Kilbride R, Rees HL, Dearnaley MP, Stevenson J,<br />

Meadows WJ and Morris CD, (2004). Assessment <strong>of</strong> the re-habilitation <strong>of</strong> the seabed following<br />

marine aggregate dredging. Sci. Ser. Tech. Rep., CEFAS Lowest<strong>of</strong>t, 121: 154pp.<br />

Boyd SE, Limpenny DS, Rees HL and Cooper KM, (2005). The effects <strong>of</strong> marine sand and gravel<br />

extraction on the macrobenthos at a commercial dredging site (results 6 years post-dredging).<br />

ICES Journal <strong>of</strong> <strong>Marine</strong> Science, 62,145-162.<br />

Bradford S, and Siebinga M, (2009). Communicating about dredging in a precious environment:<br />

port <strong>of</strong> Melbourne channel deepening project. Terra et Aqua, 116, 12-20.<br />

Bray, RN, (Ed.), (2008). Environmental Aspects <strong>of</strong> <strong>Dredging</strong>. Taylor and Francis, Leiden, The<br />

Netherlands, 386pp.<br />

Brogdon Jr NJ, Banks GE, and Ashley JA, (1994). Water Jets As A Draghead Enhancement<br />

Device. US Army Corps <strong>of</strong> Engineers Technical Report DRP-94-2. USACE, Washington DC,<br />

USA, 60pp.<br />

Burt TN, (2002). Environmental Windows as emerging issues in Europe. Presented at<br />

Environmental Windows Workshop: Achieving dredging decisions that balance economic and<br />

environmental concerns. PIANC US Section 100th Anniversary, Vicksburg, Mississippi, 16 April<br />

2002.<br />

Burt TN, and Hayes DF, (2005). Framework for research leading to improved assessment <strong>of</strong><br />

dredge generated plumes. Terra et Aqua, 98, 20-31.<br />

Camp TR, (1946). Sedimentation and the Design <strong>of</strong> Settling Tanks. Trans. ASCE, 111(2285), pp.<br />

895-936.<br />

<strong>Cefas</strong>, (2008). ‘Best Use’– Maximising the value <strong>of</strong> data collected under the ALSF projects and<br />

the <strong>Aggregate</strong> <strong>Dredging</strong> Industry. MALSF Report, MEPF 07/08, 36pp.<br />

Clarke D, Dickerson C, and Reine K, (2002). Characterization <strong>of</strong> Underwater Sounds Produced<br />

by Dredges. <strong>Dredging</strong> 2002. ASCE, Orlando, Florida, USA, 64-65.<br />

Clarkson Research Services, (2009). Dredgers Of The World 7 th Edition 2009 / 2010 Access<br />

Database.<br />

Cooper KM, Eggleton JD, Vie SJ, Vanstone K, Smith R, Boyd SE, and Ware S, (2005).<br />

Assessment <strong>of</strong> the rehabilitation <strong>of</strong> the seabed following marine aggregate dredging: Part II.<br />

CEFAS Science Series Technical Report 130. 86 pp.<br />

135


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Creek KD, and Sagraves TH, (1995). Eddy Purnp <strong>Dredging</strong>: Does It Produce Water Quality<br />

<strong>Impacts</strong>? ASCE Waterpower '95 Conference Proceedings, 2246-2252.<br />

Dearnley M, Feates N, and Benson T, (2009). The development <strong>of</strong> an instrument array to<br />

measure the concentration <strong>of</strong> silt and sand in the overflow from aggregate dredgers. MALSF<br />

Report, MEPF/08/P70, 134pp.<br />

<strong>Defra</strong> (2003). Preliminary investigation <strong>of</strong> the sensitivity <strong>of</strong> fish to sound generated by aggregate<br />

dredging and marine construction. Project AE0914 Final Report. <strong>Defra</strong>, London, UK.<br />

Desprez M, (2000). Physical and biological impact <strong>of</strong> marine aggregate extraction along the<br />

French coast <strong>of</strong> the Eastern English Channel. Short and long�term post�dredging restoration.<br />

ICES Journal <strong>of</strong> <strong>Marine</strong> Science, 57: 1428 – 1438.<br />

De Wit L, (2009). Modelling TSHD overflow plumes. Internal report, <strong>Dredging</strong> Engineering, Delft<br />

University <strong>of</strong> Technology.<br />

Diesing M, Schwarzer K, Zeiler M, and Klein H (2006). Comparison <strong>of</strong> marine sediment<br />

extraction sites by means <strong>of</strong> shoreface zonation. Journal <strong>of</strong> Coastal Research, Special Issue,39:<br />

783 – 788.<br />

Ervine DA, (1976). The entrainment <strong>of</strong> air in water. International Water Power and Dam<br />

Construction, 28, (12), pp 142-153.<br />

Ferret (2001). Big pumps for biggest dredge. http://www.ferret.com.au/n/Big-pumps-for-biggest-<br />

dredge-n732153. Accessed 14/06/09.<br />

Francingues NR, and Palermo MR, (2005). Silt curtains as a dredging project management<br />

practice. DOER Technical Notes Collection. ERDC TN-DOER-E21. Vicksburg, MS: U.S. Army<br />

Engineer Research and Development Center.<br />

Francingues NR, and Thompson D, (2006). Control <strong>of</strong> resuspended sediments in dredging<br />

projects. Proceedings <strong>of</strong> WEDA XXVI Annual Meeting and 38th TAMU <strong>Dredging</strong> Seminar, June<br />

25-28, San Diego, CA.<br />

Francingues NR, Hopman RJ, and Alexander MP, (2000). Innovations in <strong>Dredging</strong> Technology:<br />

Equipment, Operations, and Management. Proceedings Western <strong>Dredging</strong> Association,<br />

Twentieth Technical Conference. Texas A&M <strong>Dredging</strong> Seminar, pp. 3- 12.<br />

�������������������������������������������������������������������������������������<br />

�������������������������������������������������������������������<br />

Greene CR, (1987). Characteristics <strong>of</strong> Oil Industry, Dredge and Drilling Sounds in the Beaufort<br />

Sea. Journal <strong>of</strong> the Acoustical Society <strong>of</strong> America, 82,1315-1324.<br />

Gubbay S, (2004). A Review <strong>of</strong> <strong>Marine</strong> Environmental Indicators Reporting on Biodiversity<br />

Aspects <strong>of</strong> Ecosystem Health. The Royal Society for the Protection <strong>of</strong> Birds, Sandy. 73 pp.<br />

Gubbay S, (2005). A Review <strong>of</strong> <strong>Marine</strong> <strong>Aggregate</strong> Extraction in England and Wales, 1970–2005.<br />

The Crown Estate, London. 36 pp.<br />

136


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Hayes, D., G. Raymond, and T. Mc Lellan, (1984), Sediment resuspension from dredging<br />

activities. <strong>Dredging</strong> and dredged material disposal, 1, Proc. <strong>of</strong> Conf. <strong>Dredging</strong>, 72-82, Florida,<br />

USA.<br />

Herbich JB, (2000). Handbook <strong>of</strong> dredging engineering. McGraw-Hill. 992pp.<br />

Herbich JB, and Brahme SB, (1991). Literature review and technical evaluation <strong>of</strong> sediment<br />

resuspension during dredging: Contract Report HL-91-1 for U.S. Army Corps <strong>of</strong> Engineers<br />

Waterways Experiment Station, Vicksburg, MS, 87 p.<br />

Hisao T, and Sadayoshi Y, (1999). <strong>Dredging</strong> robot “No. 2 Futaba”. Workvessel, 246, 4-11.<br />

Hitchcock DR, and Bell S, (2004). Physical impacts <strong>of</strong> marine aggregate dredging on seabed<br />

resources in coastal deposits. Journal <strong>of</strong> Coastal Research, 20, 101–114.<br />

Hitchcock DR, and Drucker BR, (1996). Investigation <strong>of</strong> benthic and surface plumes associated<br />

with marine aggregates mining in the UK. In Oceanology International ’96: The Global Ocean –<br />

Towards Operational Oceanography, 5 – 8 March 1996, Brighton, UK, pp.221 – 234. Conference<br />

Proceedings, Spearhead Exhibitions Ltd, Surrey, Vol. 2.<br />

HR Wallingford, (2007). Development <strong>of</strong> a risk assessment framework: MARA (<strong>Marine</strong> <strong>Aggregate</strong><br />

Extraction Risk Assessment). MALSF Report, MEPF 04/03, 62pp.<br />

IHC, (2001). IHC’s new Dredge Track Presentation System (DTPS). Ports and <strong>Dredging</strong>, 156, 7-<br />

10.<br />

IHC, (2004). Upgrade units for conventional dredge pumps: Possibility <strong>of</strong> upgrade. Ports and<br />

<strong>Dredging</strong>, 162, 24-26.<br />

IHC (2006). Advanced dredging system for Sri Lankan hopper dredger HANSAKAWA. Ports and<br />

<strong>Dredging</strong>, 166, 12-13.<br />

IHC, (2007). One man-operated bridge. How dedication succeeded in starting a new era in<br />

dredging. Ports and <strong>Dredging</strong>, 167, 14-21.<br />

IHC, (2009a). Swell compensators. Achieving optimal production.<br />

http://www.ihcholland.com/fileadmin/documents/Company/swell_compensators.PDF. Accessed<br />

07/01/10.<br />

IHC, (2009b). IHC Merwede sets the standard with smart configuration management. Ports and<br />

<strong>Dredging</strong>, 173, 6-13.<br />

International Maritime Organisation (IMO), (2008). Prevention <strong>of</strong> air pollution from ships. Report<br />

<strong>of</strong> the working group on greenhouse gas emissions from ships. IMP circular MEPC/58 WP.8.<br />

Jan de Nul, (2003). Specialised alternative dredging methods. 3er Congreso Argentino de<br />

Ingenieria Portuaria 2003. 11pp.<br />

Jan de Nul, (2009). <strong>Dredging</strong> as protection against icebergs.<br />

http://www.jandenul.com/data/movies/Press/media/pdf/NF2003_icebergs.pdf. Accessed<br />

07/10/09.<br />

137


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Jensen A, (2006). Environmental investigation and monitoring <strong>of</strong> the fixed links across the Danish<br />

Straits. In: Eisma D (Ed.). <strong>Dredging</strong> in Coastal Waters. Taylor and Francis, London, UK, 41-49.<br />

Johns S, (2008). Exploring the social dimensions <strong>of</strong> an expansion to the seafloor exploration and<br />

mining industry in Australia. Desktop study: Australian context review. CSIRO Report P2007/661,<br />

Wealth From The Oceans Progamme, 54pp.<br />

Kemp R. (2008). Energy consumption <strong>of</strong> marine aggregate extraction. The Crown Estate,<br />

London, 21 pp.<br />

Kenny AJ, and Rees HL, (1994). The effects <strong>of</strong> marine gravel extraction on the macrobenthos:<br />

early post dredging recolonization. <strong>Marine</strong> Pollution Bulletin, 28: 442 – 447.<br />

Kenny AJ, and Rees HL, (1996). The effects <strong>of</strong> marine gravel extraction on the macrobenthos:<br />

Results 2 years post�dredging. <strong>Marine</strong> Pollution Bulletin, 32: 615 – 622.<br />

Kenny AJ, Rees HL, Greening J, and Campbell S, (1998). The effects <strong>of</strong> marine gravel extraction<br />

on the macrobenthos at an experimental dredge site <strong>of</strong>f North Norfolk, UK (results 3 years<br />

post�dredging). ICES Document CM 1998/V:14. 7 pp.<br />

Kirby R. and Land J, (1991). The impact <strong>of</strong> dredging: A comparison <strong>of</strong> annual and man-made<br />

disturbances to cohesive sedimentary regimes. Paper B3. Proc. CEDA-PIANC Conference,<br />

Amsterdam, The Netherlands.<br />

Land J, Burt JN, and Otten H, (2004). Application <strong>of</strong> a new international protocol to measurement<br />

<strong>of</strong> sediment release from dredgers. WODCON XVII .<br />

Limpenny DS, Boyd SE, Meadows WJ, Rees HL, and Hewer AJ, (2002). The utility <strong>of</strong> sidescan<br />

sonar techniques in the assessment <strong>of</strong> anthropogenic disturbance at aggregate extraction sites.<br />

ICES Document CM 2002/K:04. 20 pp.<br />

Littleboy A, and Boughen N, (2007). Exploring the social dimensions <strong>of</strong> an expansion to the<br />

seafloor exploration and mining industry in Australia: Synthesis Report. CSIRO Report<br />

P2007/917, Wealth From The Oceans Progamme, 24pp.<br />

Malherbe B, and De Pooter P, (2008). New possibilities for ripper dredging <strong>of</strong> rock. Terra et<br />

Aqua, 110, 9-12.<br />

Mallee GT, (2000). A new innovative Dredge Track Presentation System (DTPS). IHC Report.<br />

12pp.<br />

<strong>Marine</strong> and Fisheries Agency, (2009). Procedure for considering marine aggregate permissions.<br />

MFA, London. 18pp.<br />

McLellan NT, and Hopman RJ, (2000). Inovations in <strong>Dredging</strong> Technology: Equipment,<br />

Operations and Management. USACE ERDC, Report No. TR-DOER-5.Vicksburg, MS.<br />

Miedema SA, (1994). On the snow-plough effect when cutting water saturated sand with inclined<br />

straight blades. ASCE Proc. <strong>Dredging</strong> 94’, Orlando, Florida, USA.<br />

Miedema SA, and Van Rhee C, (2007). A sensitivity analysis on the effects <strong>of</strong> dimensions and<br />

geometry <strong>of</strong> trailing suction hopper dredges. WODCON, Orlando, USA.<br />

138


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Millner RS, Dickson RR, and Rolfe MS, (1977). Physical and biological studies <strong>of</strong> a dredging<br />

ground <strong>of</strong>f the east coast <strong>of</strong> England. ICES Document CM 1977/E:48. 11 pp.<br />

Minerals Management Service, (2004). Review <strong>of</strong> existing and emerging environmentally friendly<br />

<strong>of</strong>fshore dredging technologies. OCS Report MMS 2004-076, 441pp.<br />

Mink F, Dirks W, Van Raalte G, De Vlieger H, and Russell M, (2006). Impact <strong>of</strong> European Union<br />

environmental law on dredging. Terra et Aqua, 104, 3-10.<br />

Moch NK, and Tiarma S, (2002). House calls for sand-mining ban as marine damage looms. The<br />

Jakarta Post September 10 2002. Jakarta, Indonesia.<br />

National Academy <strong>of</strong> Sciences (NAS), (2001). A Process for Setting, Managing,and Monitoring<br />

Environmental Windows for <strong>Dredging</strong> Projects. Transportation Research Board Special Report,<br />

262, 96pp.<br />

National <strong>Marine</strong> Fisheries Service (NMFS), (2004). Endangered Species Act Section 7<br />

Consultation on <strong>Dredging</strong> <strong>of</strong> Gulf <strong>of</strong> Mexico Navigation Channels and Sand Mining (“Borrow”)<br />

Areas Using Hopper Dredges by COE Galveston, New Orleans, Mobile, and Jacksonville<br />

Districts (Consultation Number F/SER/2000/01287), 119 pp.<br />

Netzband A, and Adnitt C, (2009). <strong>Dredging</strong> management practices for the environment: A<br />

structured selection approach. Terra et Aqua, 114, 3-8.<br />

Newell RC, and Seiderer LJ, (2003). Literature Review On Ecological <strong>Impacts</strong> Of <strong>Dredging</strong>. In:<br />

Minerals Management Service, (2004). Review <strong>of</strong> existing and emerging environmentally friendly<br />

<strong>of</strong>fshore dredging technologies. OCS Report MMS 2004-076, 441pp.<br />

Newell RC, Seiderer LJ, and Hitchcock DR, (1998). The impact <strong>of</strong> dredging works in coastal<br />

waters: a review <strong>of</strong> the sensitivity to disturbance and subsequent recovery <strong>of</strong> biological resources<br />

on the seabed. Oceanography and <strong>Marine</strong> Biology, 36: 127 – 178.<br />

Newell RC, Seiderer LJ, Simpson NM, and Robinson JE, (2002). Impact <strong>of</strong> marine aggregate<br />

dredging and overboard screening on benthic biological resources in the central North Sea:<br />

Production Licence Area 408. Coal Pit. <strong>Marine</strong> Ecological Surveys Limited. Technical Report No.<br />

ER1/4/02 to the British <strong>Marine</strong> <strong>Aggregate</strong> Producers Association (BMAPA). 72 pp.<br />

Newell RC, Seiderer LJ, Simpson NM, and Robinson JE, (2004). <strong>Impacts</strong> <strong>of</strong> marine aggregate<br />

dredging on benthic macr<strong>of</strong>auna <strong>of</strong>f the south coast <strong>of</strong> the United Kingdom. Journal <strong>of</strong> Coastal<br />

Research, 20: 115 – 125.<br />

Office <strong>of</strong> the Deputy Prime Minister (ODPM), (2002). <strong>Marine</strong> Mineral Guidance 1: Extraction by<br />

dredging from the English seabed. 32pp.<br />

Ofuji I, and Ishimatsu N, (1976). Anti-turbidity overflow system for hopper dredge. World<br />

<strong>Dredging</strong> Congress (WODCON VII), 207-233.<br />

OSPAR, (2009). Overview <strong>of</strong> the impacts <strong>of</strong> anthropogenic underwater sound in the marine<br />

environment. OSPAR Commission, London, 134pp.<br />

139


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Parvin SJ, Nedwell JR, Kynoch J, Lovell J and Brooker AG (2008). Assessment <strong>of</strong> underwater<br />

noise from dredging operations on the Hastings shingle bank. Report No. Subacoustech<br />

758R0137<br />

Pennekamp JEA, (1996). Turbidity caused by dredging viewed in perspective. Terra et Aqua No.<br />

64 .<br />

Penta-Ocean Construction Co. (2009). Submersible <strong>Dredging</strong> Robot [Technologies and<br />

Innovation] - PentaOcean. http://www.penta-ocean.co.jp/english/business/civil/sdr.html.<br />

Accessed 24/08/09.<br />

Pisces Environmental Services (Pty) Ltd, (2008). Data Gathering and Gap Analysis for<br />

Assessment <strong>of</strong> Cumulative Effects <strong>of</strong> <strong>Marine</strong> Diamond Mining Activities on the BCLME Region:<br />

Chapter 4: Mining Methods. Report Project: BEHP/CEA/03/02. 52pp.<br />

Pointer IR, and Kennedy R, (1984). Complex patterns <strong>of</strong> change in the macrobenthos <strong>of</strong> a large<br />

sandbank following dredging. Journal <strong>of</strong> <strong>Marine</strong> Biology, 78: 335 – 352.<br />

Randall RE, (2006). <strong>Dredging</strong> in the United States. In: Eisma D (Ed.). <strong>Dredging</strong> in Coastal<br />

Waters. Taylor and Francis, London, UK, 139-166.<br />

Reine KJ, Dickerson DD, and Clarke DG, (1998).Environmental windows associated with<br />

dredging operations. DOER Technical Notes Collection (TNDOER-E2), US Army Engineer<br />

Research and Development Center, Vicksburg, MS.<br />

Richardson JF, and Zaki WN, (1954). Sedimentation and Fluidization I. Transactions <strong>of</strong> the<br />

Institute <strong>of</strong> Chemical Engineers, 32, 35-53.<br />

Richardson WJ, Fraker MA, Wuersig B, and Wells RS, (1985). Behaviour <strong>of</strong> bowhead whales,<br />

Balaena mysticetus summering in the Beaufort sea: reactions to industrial activities. Biological<br />

Conservation, 32, 195-230.<br />

RIZA, (2005).Vertroebeling tijdens en na baggeren met sleephopperzuiger in het<br />

Noordzeekanaal (In Dutch). Technical report, RIZA rapport 2005.006.<br />

Robinson JE, Newell RC, Seiderer LJ, and Simpson NM, (2005). <strong>Impacts</strong> <strong>of</strong> aggregate dredging<br />

on sediment composition and associated benthic fauna at an <strong>of</strong>fshore dredge site in the southern<br />

North Sea. <strong>Marine</strong> Environmental Research, 60: 51 – 68.<br />

Rodi W, (1993). Turbulence Models and their Application in Hydraulics, A state <strong>of</strong> the art review.<br />

IAHR, Third Edition.<br />

Rosati J, (1999). Automated Inspection Tool "Silent Inspector" Undergoes Field Testing.<br />

<strong>Dredging</strong> Research, 2, 1-3.<br />

Spearman J, Bray RN, and Burt TN, (2004). Dynamic representation <strong>of</strong> trailer dredger source<br />

terms in plume dispersion modeling. Wodcon XVI, Hamburg, Germany.<br />

Sutton G, and Boyd S, (Eds), (2009). Effects <strong>of</strong> Extraction <strong>of</strong> <strong>Marine</strong> Sediments on the <strong>Marine</strong><br />

Environment 1998 – 2004. ICES Cooperative Research Report No. 297. 180 pp.<br />

140


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Sydney Ports, (2009). Port Botany Expansion. Ship Turning Area <strong>Dredging</strong> Section 96 ( 1a )<br />

Modification. Sydney Ports Corporation Report. 44pp.<br />

Taylor A, McWilliams J, Van Dam H, and Lammers B, (2006). Deepening, cleaning and<br />

processing sediment from the Miami river. Terra et Aqua, 103, 23-29.<br />

The Crown Estate, (2008). <strong>Marine</strong> aggregate dredging 1998 -2007: A ten year review. The<br />

Crown Estate, London, 28pp.<br />

The Crown Estate, (2009). The Crown Estate Electronic Monitoring System 2008 Report. The<br />

Crown Estate, London. 8pp.<br />

Thomsen F, McCully S, Wood D, Pace F, and White P (2009). A generic investigation into noise<br />

pr<strong>of</strong>iles <strong>of</strong> marine dredging in relation to the acoustic sensitivity <strong>of</strong> the marine fauna in UK waters<br />

with particular emphasis on aggregate dredging: PHASE 1 Scoping and review <strong>of</strong> key issues.<br />

MEPF Report, Ref No. MEPF/08/P21.<br />

US Army Corps <strong>of</strong> Engineers (USACE), (2006). The Corps Hopper Dredge Wheeler.<br />

http://www.mvn.usace.army.mil/pao/bro/wheeler.htm. Accessed 12/12/09.<br />

US Environmental Protection Agency (USEPA), (1994). Remediation guidance document. EPA<br />

905-R94-003. Chicago, IL: Assessment and Remediation <strong>of</strong> Contaminated Sediments Program,<br />

Great Lakes National Program Office.<br />

US Environmental Protection Agency (USEPA), (2005). Contaminated sediment remediation<br />

guidance for hazardous waste sites. OSWER 9355.0-85. Washington, DC: Office <strong>of</strong> Solid Waste<br />

and Emergency Response, US Environmental Protection Agency.<br />

US Environmental Protection Agency (USEPA), (2008). Innovative <strong>Dredging</strong> Technology<br />

Accelerates Removal <strong>of</strong> Residual Contamination in Ashtabula River. Technology News and<br />

Trends (USEPA Newsletter), 38, 1-3.<br />

Van de Ketterij RG, Stapersma D, Kramers CHM, and Verheijen LTG, (2009). CO2 index:<br />

Matching the dredging industries needs with IMO legislation. Proceedings Ceda <strong>Dredging</strong> Days,<br />

Rotterdam 2009.<br />

Van den Berg CH, (1986). The Use <strong>of</strong> Underwater Pumps On Trailing Suction Hopper Dredgers.<br />

Ports and <strong>Dredging</strong>, 125, 1-14.<br />

Van den Berg CH, (2001). The need for different pump types. Ports and <strong>Dredging</strong>, 154, 10-13.<br />

Van der Linde P, Bodegom D, and Janssen-Stelder B, (2005). Sub-Surface <strong>Dredging</strong>: An<br />

Economic and Environmentally Friendly Technique for Lowering Ground Levels. Terra et Aqua,<br />

99, 11-14.<br />

Van Melkebeek E, (2002). Pre-Trenching, Pre-Sweeping and Backfilling for the 36” Offshore<br />

Pipeline Project in Taiwan. Terra et Aqua, 87, 19-25.<br />

141


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Van Os AG, and Van Leussen W, (1987). Basic research on cutting forces in saturated sand,<br />

Journal <strong>of</strong> Geotechnical Engineering. Vol 113, No. 12, pp 1501-1516.<br />

Van Parys M, Dumon PA, Claeys S, Lanckneus J, Lancker VV, Vangheluwe M, Sprang PV,<br />

Speleers LV, and Janssen C, (2001). Environmental monitoring <strong>of</strong> the dredging and relocation<br />

operations in the coastal harbours in Belgium, Mobag 2000, WODCON XVI .<br />

Van Raalte GH, (2006). <strong>Dredging</strong> techniques; adaptations to reduce environmental impact. In:<br />

Eisma D (Ed.). <strong>Dredging</strong> in Coastal Waters. Taylor and Francis, London, UK, 1-40.<br />

Van Rhee C, (2001). Numerical Simulation <strong>of</strong> the Sedimentation Process in a Trailing Suction<br />

Hopper Dredger. 16th World <strong>Dredging</strong> Congress and Exhibition (WODCON), Kuala Lumpur.<br />

Van Rhee C, (2002a). On the sedimentation process in a trailing suction hopper dredger.<br />

Unpublished PhD Thesis, Delft University <strong>of</strong> Technology.<br />

Van Rhee, C. (2002b). Modelling the sedimentation process in a trailing suction hopper dredger.<br />

Terra et Aqua, 18-27.<br />

Van Rijn LC, (2005). Principles <strong>of</strong> sedimentation and erosion engineering in rivers, estuaries and<br />

coastal seas. Aqua Publications.<br />

Van Rijn LC, Soulsby RL, Hoekstra P, and Davies AG (Eds), (2005). SANDPIT: Sand Transport<br />

and Morphology <strong>of</strong> Offshore Sand Mining Pits. Process Knowledge and Guidelines for Coastal<br />

Management. End document May 2005, EC Framework V Project No. EVK3 �2001� 00056.<br />

Aqua Publications, the Netherlands. 716 pp.<br />

VBKO (2003). Protocol for the field measurement <strong>of</strong> sediment release from dredgers, issue 1,<br />

August 2003, Technical Report, VBKO TASS Project.<br />

Vlasblom WJ, and Miedema SA, (1995). A theory for determining sedimentation and overflow<br />

losses in hoppers. Proceedings <strong>of</strong> the 14th World <strong>Dredging</strong> Congress, Amsterdam.<br />

Wakeman T, Sustar J, and Dickson W, (1975). Impact <strong>of</strong> three dredge types compared in San<br />

Francisco district. World <strong>Dredging</strong> and <strong>Marine</strong> Construction, 9-14.<br />

Williams GL, and Visser KG, (1997). The Punaise: a remotely operated submerged dredging<br />

system. Terra et Aqua, 69, 20-28.<br />

Willoughby M, and Crabb D, (1983). The behaviour <strong>of</strong> dredge generated sediment plumes in<br />

Moreton Bay. Sixth Australian Conference on Coastal and Ocean Engineering, 182-186.<br />

Winterwerp, J. (2002). Near field behavior <strong>of</strong> dredging spill in shallow water. Journal <strong>of</strong><br />

Waterway, Port, Coastal and Ocean Engineering, 128 (2), 96-98<br />

Yano S, Kumamoto H, Watanabe T, Ogawa T, and Miyake N, (2006). "Seiryu-maru" – Trailing<br />

Suction Hopper Dredger With Oil Recovery System. Mitsubishi Heavy Industries Ltd. Technical<br />

Review, 43, 1, 1-2.<br />

142


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

143


Benchmarking Equipment, Practices and Technologies<br />

�����������<br />

������������������������������������<br />

�����<br />

Report No: 10/J/1/06/1309/0996<br />

The sedimentation process in a hopper has a strong resemblance to the sedimentation process<br />

in secondary sedimentation tanks used in the field <strong>of</strong> sewage and water treatment. It is therefore<br />

no coincidence that some models used in the dredging industry are based on work done in<br />

wastewater treatment. In this respect the work <strong>of</strong> Camp (1946) must be mentioned since this<br />

forms the basis <strong>of</strong> many subsequent models.<br />

The method <strong>of</strong> Camp is based on the supposed trajectories <strong>of</strong> particles during settling (see<br />

Figure A-1). The mixture enters the hopper in the inlet zone and is transported towards the outlet<br />

zone. The trajectories <strong>of</strong> particles are observed. A particle with a settling velocity �� reaches the<br />

bed regardless <strong>of</strong> the vertical position leaving the inlet. Particles with lower settling velocity only<br />

reach the bed in cases where the vertical starting position is not too high.<br />

�<br />

����������������������������������������������������������<br />

144


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

This leads to the settling efficiency for a certain fraction <strong>of</strong> the particle size distribution. The<br />

settling efficiency (or removal ratio) for a certain fraction is:<br />

�<br />

�<br />

�<br />

� ������<br />

�<br />

�<br />

The removal ratio is the ratio <strong>of</strong> the incoming particles <strong>of</strong> a certain fraction that settle in the<br />

hopper. With v as settling velocity <strong>of</strong> that fraction and �� defined as overflow rate:<br />

Where:<br />

�<br />

�<br />

�<br />

� ������<br />

��<br />

� Mixture discharge into hopper [m 3 /s]<br />

� Length <strong>of</strong> the hopper [m]<br />

� Width <strong>of</strong> the hopper [m]<br />

In Figure A-1 the particles follow a straight line during settling. In reality this will not be the case<br />

since flow will be turbulent. To take the influence <strong>of</strong> turbulence into account Camp solved an<br />

advection / diffusion equation with a constant diffusion coefficient. The result <strong>of</strong> this analysis is<br />

shown in Figure A-2.<br />

����������� �����������������������������������������������������������������������������<br />

145


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

Lines are drawn for different values <strong>of</strong>������ and the position on the horizontal axis includes the<br />

influence <strong>of</strong> turbulence which creates mixing. The mixing, or diffusion coefficient is � hence high<br />

values <strong>of</strong> turbulence are shown on the left side <strong>of</strong> the graph.<br />

�������������������<br />

The Vlasblom and Miedema (1995) model is an refinement <strong>of</strong> the Camp model. The Camp model<br />

uses a constant mixture level in the basin. The sludge that settles on the bottom is scraped to a<br />

central location and sucked away. In a hopper <strong>of</strong> a TSHD the mixture level decreases in time due<br />

to the rising bed level and this effect is taken into account in the Vlasblom & Miedema model.<br />

Furthermore the effect <strong>of</strong> hindered settling is also taken into account. Hindered settling is the<br />

decrease <strong>of</strong> settling velocity with increasing concentration.<br />

������������������<br />

Van Rhee (2001) published a one dimensional model based on the observation that the flow in<br />

the hopper is dominated by density effects. In the inlet zone the flow is directed vertically towards<br />

the bed due to buoyancy. At this location settling occurs and on average the flow is vertical in the<br />

area above the density current. This vertical flow is modelled using an advection / diffusion<br />

equation for different fractions <strong>of</strong> the PSD.<br />

�������������������<br />

The flow in a hopper can be regarded as two dimensional in the horizontal (length) and vertical<br />

direction over the largest part in a hopper. Based on this observation Van Rhee (2002a, 2002b)<br />

developed a 2DV model based on the Reynolds Averaged Navier Stokes (RANS) equations with<br />

a �-� turbulence model (Rodi (1993)). In the model the momentum equations and sediment<br />

transport equations are coupled. The model has a movable bed and water level (the latter for<br />

changing overflow levels during time). Inflow and overflow location, inflow concentration and<br />

velocity and the inflow PSD can all be varied in time.<br />

This model is the most sophisticated model available. In Figure A-3 a result <strong>of</strong> this model is<br />

shown. In the upper panel the flow velocity is shown while the lower panel shows the<br />

concentration. The inflow is at the left side <strong>of</strong> the hopper, the overflow at the right side. A settled<br />

bed can be seen in both panels. Where the mixture enters the hopper at the left side an erosion<br />

crater develops. The mixture flows out <strong>of</strong> this crater as a heavy fluid – a density current. Due to<br />

this phenomenon the velocity is not uniformly distributed over height.<br />

146


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

����������� ��������������������������������������������������������������������<br />

���������������������<br />

For a first quick estimate <strong>of</strong> the cumulative overflow loss one might expect during loading <strong>of</strong><br />

sediment with a certain PSD when the hopper size and discharge is known Camp’s (1946)<br />

method provides good results.<br />

To get more insight on the development during (loading) time <strong>of</strong> losses the Vlasblom and<br />

Miedema (1995) model can be used. This model predicts the particle size distribution <strong>of</strong> the<br />

overflow mixture during time. Calculation effort is limited therefore a large number <strong>of</strong> calculations<br />

can be performed in a restricted time. If more results are needed, for instance to investigate<br />

loading and overflow strategy, the effect <strong>of</strong> the water level in the hopper at the start <strong>of</strong> the loading<br />

process and the influence <strong>of</strong> the hopper shape, the<br />

.<br />

Vlasblom and Miedema (1995) model is too<br />

restricted and a more sophisticated model is needed<br />

147


Benchmarking Equipment, Practices and Technologies<br />

�����������<br />

�������������������������<br />

Report No: 10/J/1/06/1309/0996<br />

������������<br />

IMO ship pollution rules are contained in the “International Convention on the Prevention <strong>of</strong><br />

th<br />

Pollution from Ships”, known as MARPOL 73/78. On 27 September 1997, the MARPOL<br />

Convention<br />

was been amended by the “1997 Protocol”, which includes Annex VI,<br />

“Regulations for the Prevention <strong>of</strong> Air Pollution from Ships”. MARPOL Annex VI sets limits on<br />

NOx and SOx emissions from ship exhausts, and prohibits deliberate emissions <strong>of</strong> ozone<br />

depleting substances. The IMO emission standards are commonly referred to as Tier I - III<br />

standards. The Tier I standards were defined in the 1997 version <strong>of</strong> Annex VI, while the Tier II/III<br />

standards were introduced by Annex VI amendments adopted in 2008.<br />

�����������������������<br />

The 1997 Protocol to MARPOL, which includes Annex VI, became effective 12 months after<br />

being accepted by 15 States with not less than 50% <strong>of</strong> world merchant shipping tonnage. On 18 th<br />

May 2004, Samoa deposited its ratification as the 15th State (joining Bahamas, Bangladesh,<br />

Barbados, Denmark, Germany, Greece, Liberia, Marshal Islands, Norway, Panama, Singapore,<br />

Spain, Sweden, and Vanuatu). At that date, Annex VI was ratified by States with 54.57% <strong>of</strong> world<br />

merchant shipping tonnage.<br />

Accordingly, Annex VI entered into force on 19 th May 2005. It applies retroactively to new<br />

engines greater than 130 kW installed on vessels constructed on or after January 1 st , 2000, or<br />

which undergo a major conversion after that date. The regulation also applies to fixed and<br />

floating rigs and to drilling platforms (except for emissions associated directly with exploration<br />

and/or handling <strong>of</strong> sea-bed minerals). In anticipation <strong>of</strong> the Annex VI ratification, most marine<br />

engine manufacturers have been building engines compliant with the above standards<br />

since 2000.<br />

������������������������������<br />

Annex VI amendments adopted in October 2008 introduced:<br />

� New fuel quality requirements beginning from July 2010.<br />

� Tier II and III NOx emission standards for new engines.<br />

� Tier I NOx requirements for existing pre-2000 engines.<br />

The revised Annex VI enters into force on 1 st July 2010. By October 2008, Annex VI was ratified<br />

by 53 countries (including the Unites States), representing 81.88% <strong>of</strong> tonnage.<br />

Emission Control Areas<br />

Two sets <strong>of</strong> emission and fuel quality requirements are defined by Annex VI:<br />

� Global requirements<br />

148


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

� More stringent requirements applicable to ships in Emission Control Areas (ECA).<br />

An Emission Control Area can be designated for SOx and PM, or NOx, or all three types<br />

<strong>of</strong> emissions from ships, subject to a proposal from a Party to Annex VI. Existing SOx<br />

Emission Control Areas include the Baltic Sea (adopted: 1997 / entered into force: 2005)<br />

and the North Sea (2005/2006). Future Emission Control Areas could also include zones<br />

around pollution sensitive ports.<br />

�����������������������<br />

NOx emission limits are set for diesel engines depending on the engine maximum operating<br />

speed (n, rpm) as indicated in Figure B-1. Tier I and Tier II limits are global, while the Tier III<br />

standards apply only in NOx Emission Control Areas.<br />

����������� �������������������������������������<br />

Tier II standards are expected to be met by combustion process optimization. The parameters<br />

examined by engine manufacturers include fuel injection timing, pressure, and rate (rate<br />

shaping), fuel nozzle flow area, exhaust valve timing, and cylinder compression volume.<br />

Tier III standards are expected to require dedicated NOx emission control technologies such as<br />

various forms <strong>of</strong> water induction into the combustion process (with fuel, scavenging air, or incylinder),<br />

exhaust gas recirculation, or selective catalytic reduction.<br />

149


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

The main changes to MARPOL Annex VI will see a progressive reduction in sulphur oxide (SOx)<br />

emissions from ships, with the global sulphur cap reduced initially to 3.50% (from the current<br />

4.50%), effective from 1 st January 2012; then progressively to 0.50 %, effective from 1 st January<br />

2020, subject to a feasibility review to be completed no later than 2018.<br />

The limits applicable in Sulphur Emission Control Areas (SECAs) will be reduced to 1.00%,<br />

beginning on 1 st July 2010 (from the current 1.50 %); being further reduced to 0.10 %, effective<br />

from 1 st January 2015.<br />

Progressive reductions in nitrogen oxide (NOx) emissions from marine engines were also<br />

agreed, with the most stringent controls on so-called “ Tier III ”<br />

engines, i.e. those installed on<br />

st<br />

ships constructed on or after 1 January 2016, operating in Emission Control Areas.<br />

(IMO, 2008). The consequence <strong>of</strong> the new IMO legislation on fuel oil is that from 2015 only light<br />

distillates can be used in SECA’s and from 2020 only light distillates can be used worldwide.<br />

�������������<br />

The Kyoto Protocol is an international agreement linked to the United Nations Framework<br />

Convention on Climate Change. The major feature <strong>of</strong> the Kyoto Protocol is that it sets binding<br />

targets for 37 industrialized countries and the European community for reducing greenhouse gas<br />

(GHG) emissions.These amount to an average <strong>of</strong> five per cent against 1990 levels over the fiveyear<br />

period 2008-2012. Shipping and aircraft were however excluded from these agreements. At<br />

the Bali summit <strong>of</strong> 2007 governments agreed to include both shipping and aircraft during the<br />

Copenhagen summit in December 2009 (COP15 conference). This forced the IMO into action<br />

and it started procedures to come up with an Energy Efficiency Design Index (CO 2 index) for<br />

ships. Such an index will probably be used by (port) authorities in the tender process for<br />

maintenance and land reclamation. Irrespective <strong>of</strong> such an index and the consequences it may<br />

have for ship’s operations, the upcoming agreements from the COP15 conference will result in<br />

limited CO2 emission rights for ship owners, and significant costs for excess <strong>of</strong> CO 2 emission<br />

limits. These emission rights will be limited progressively in time. CO 2 emissions must be<br />

reduced by 20% in 2020 with respect to the emissions in 1990. It is expected that emissions<br />

reduction targets for 2050 will amount to 50 %. It is therefore <strong>of</strong> the utmost importance that the<br />

CO2 index calculated according to IMO norms, correctly reflect the energy consumption and<br />

energy efficiency <strong>of</strong> dredging equipment (Van de Ketterij et al. 2009). .<br />

It is to be expected that the agreements on CO2 reduction in the future will lead to increased<br />

costs for emission and limited emission rights for ship owners. Apart from this cost factor it could<br />

lead to the decision <strong>of</strong> authorities not to allow the use <strong>of</strong> ships (dredgers) with insufficient<br />

energy efficiency.<br />

150


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

To compare the energy efficiency <strong>of</strong> vessels IMO proposed an energy index during the MEPC58<br />

conference (IMO, 2008). The essence <strong>of</strong> the index is that the ratio between the CO2 emission<br />

from the installed power sources and the product <strong>of</strong> tonnage and speed is computed. The<br />

product <strong>of</strong> tonnage and speed can be regarded as a benefit for society, while the CO2 emission is<br />

the price that has to be paid. The index in simplified shape reads:<br />

Where :<br />

���� �<br />

���<br />

�<br />

���<br />

� �<br />

� ��� � � � ��� �<br />

���� ��� ��� �� �� ��<br />

�� ��<br />

���� ��<br />

�� Conversion factor between fuel consumption and CO2 emission. [-]<br />

��� ��� Specific Fuel consumption <strong>of</strong> a main engine<br />

(at 75 % <strong>of</strong> the rated installed power (MCR)) [g/kWh]<br />

��� �� Specific Fuel consumption <strong>of</strong> the auxiliary engines [kW]<br />

� ��<br />

75 % <strong>of</strong> the MCR <strong>of</strong> the main engine nr. i [kW]<br />

� ��<br />

Power <strong>of</strong> the auxiliary engines [kW<br />

��� Number <strong>of</strong> main engines [-]<br />

�� Deadweight [ton]<br />

� ���<br />

Ships speed on deep water at max load [kn]<br />

151<br />

���<br />

(B-1)<br />

The CO2 index for ships generally is based on the energy index. This index has been introduced<br />

by Gabrielli and Von Karman (1950) in a famous paper. In this paper they tried to compare<br />

energy consumption <strong>of</strong> various means <strong>of</strong> transportation, making them dimensionless by division<br />

by a transport capacity.<br />

�<br />

� � (B-2)<br />

��


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

����������� �����������������������������������������������������������������������������������<br />

An updated version, with modern transport means, <strong>of</strong> this figure is shown in Figure B-3. Here it is<br />

shown that transport by shipping is very energy efficient compared with the other transport<br />

methods.<br />

For a Trailing Suction Hopper Dredge the energy index cannot be simply captured with the<br />

definition <strong>of</strong> equation B-2,<br />

and likewise the definition <strong>of</strong> a CO2 index in the shape <strong>of</strong> equation<br />

B-1 cannot be used because the power installed on board <strong>of</strong> a TSHD is not primarily used for<br />

horizontal transportation <strong>of</strong> sediment. For the loading phase and the discharge phase a<br />

considerable amount <strong>of</strong> energy is needed as well. Therefore Van de Ketterij et al. (2009)<br />

proposed an alternative method to compute the energy efficiency <strong>of</strong> a TSHD taken the complete<br />

dredge cycle into account.<br />

152


Benchmarking Equipment, Practices and Technologies<br />

�<br />

Report No: 10/J/1/06/1309/0996<br />

����������� ������������������������������������������������������������������<br />

The total amount <strong>of</strong> energy needed for the complete dredge cycle reads:<br />

���� � �� �� �� ��� � � ��� �� �� � �������� B-3<br />

Where:<br />

� ��� = total energy needed for the dredge cycle, �� � = the power used during sailing empty, � �� =<br />

the power used for sailing fully loaded, � � = the power used during dredging, � ��� = the power<br />

used for discharging the load, ��������� � and � ��� the time needed for sailing loaded, sailing full,<br />

dredging and discharging respectively.<br />

By taking the average energy consumption over the total cycle the following index is found:<br />

� � � � �<br />

� �� �� ��<br />

� � � � �<br />

� � �<br />

�� ���� �� ���� ��� �� ��<br />

� ���<br />

�� �� � ���<br />

����� ����� ����� ����� �����<br />

153<br />

B-4


Benchmarking Equipment, Practices and Technologies<br />

Report No: 10/J/1/06/1309/0996<br />

The gravitational acceleration was introduced in the denominator to get a dimensionless index.<br />

Velocity � is defined as the sailing distance full divided by the cycle time:<br />

B-5<br />

�<br />

� �<br />

�<br />

����<br />

�����<br />

This energy index takes all phases <strong>of</strong> the dredge cycle into account and is therefore a much<br />

more objective measurement to compare the energy efficiency (and CO2 efficiency since CO2 is<br />

directly coupled to the energy efficiency). Using this method the expression in eq. B-1 can be<br />

extended to an EEDI index for a TSHD by splitting up the first term in the nominator in the power<br />

contributions over the different phases and using definition B-5 for the sailing speed.<br />

A complication <strong>of</strong> the method is that a standard dredging cycle must be defined to compare<br />

different dredgers. This involves a certain selection <strong>of</strong> sediment properties, sailing distance,<br />

dredging depth and discharge method. The selection itself can have a large influence. If for<br />

instance the particle size is very coarse efficiency improvement <strong>of</strong> the loading system will not be<br />

rewarded. If a long sailing distance is chosen the sailing phase dominates over the other phases<br />

and improvements <strong>of</strong> the dredging equipment will not be measured. In Van de Ketterij et al. (2009)<br />

a certain standard cycle was proposed, but discussions on this topic are ongoing.<br />

154


<strong>Marine</strong><br />

<strong>Aggregate</strong> Levy<br />

Sustainability Fund<br />

MALSF<br />

© Crown Copyright 2010<br />

Published by the <strong>Marine</strong> <strong>Aggregate</strong> Levy Sustainability Fund (MALSF).<br />

For more information, please visit:<br />

http://www.alsf-mepf.org.uk/<br />

http://www.english-heritage.org.uk/ALSF<br />

http://alsf.defra.gov.uk<br />

Cover Image Credits<br />

Top Row: Left to Right Bottom Row: Left to Right<br />

BMAPA Crown Copyright, Courtesy <strong>of</strong> <strong>Cefas</strong><br />

Crown Copyright, Courtesy <strong>of</strong> <strong>Cefas</strong> BMAPA<br />

St Andrews University & English Heritage Wessex Archaeology & English Heritage

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!