IN THE MATTER OF the Resource Management Act 1991

AND

IN THE MATTER OF a Board of Inquiry appointed under section 149J of the Resource Management Act 1991 to consider The New Zealand King Co. Limited's private plan change requests to the Marlborough Sounds Resource Management Plan and resource consent applications for marine farming at nine sites located in the Marlborough Sounds.

STATEMENT OF EVIDENCE OF

ROB SCHUCKARD

AUGUST 2012

– 2 –

Contents:

1. Executive Summary

2. Background

3. Scope

4. Status of ecosystem information from Coastal Marine Environment.

5. Eutrofication, a worldwide concern and threat to coastal habitats.

6. Salmon farming and eutrofication

7. Human Population Equivalent.

8. Seabirds in the Marlborough Sounds. Impact of eutrophication on marine ecosystem of the Marlborough Sounds, a particular focus on New Zealand King Shags.

1. Executive Summary

1.1. New Zealand‘s marine habitats, the communities of marine life within them, the processes that drive marine ecosystems, the full extent of threats to marine biodiversity from human activities and broader environmental changes have not yet been identified and/or classified.

1.2. The past few decades have seen a significant global increase in coastal eutrophication, leading to widespread hypoxia and anoxia, habitat degradation, alteration of food-web structure, loss of biodiversity, and increased frequency, spatial extent, and duration of harmful algal blooms.

1.3. Nitrogen and phosphorus loading into marine waters can initiate a biological process of eutrophication that, depending on the volume and duration of nutrient loading and the assimilative capacity of the receiving waters, can culminate in a fundamental shift in the food web structure of an area and lead to ecological simplification, disrupting normal ecosystem functioning. It finally can result in a shift of phytoplankton species composition and create conditions that are favourable to nuisance and toxic algal blooms.

1.4. High and very high HAB risk in the form of shellfish and toxicity occurs near all the salmon farms, in particular during the summer months. However, similar HAB were also identified at the control sites. It is suggested that nutrient waste from existing farms are not of notable environmental significance at current levels of salmon production. – 3 –

1.5. To challenge of eutrophication and related harmful algal blooms, an integrated management framework is needed for effective nutrient management. Both changes in climate forcing and nutrient loadings are aspects of a global transformation that is expected to profoundly impact coastal hypoxia through more stratified water conditions.

1.6. Algal blooms that occur as a result of high nutrients in marine waters will reduce water clarity (and consequently sunlight availably to other organisms in the water column and benthic communities), resulting in oxygen deficit of the water column when the organisms die, sink and decompose.

1.7. Adaptive management and guidelines are being integrated in the conditions for resource consents of existing salmon farms. However, they only accommodate the industrial goal to maintain a healthy environment for the fish farm.

1.8. Intensive aquaculture has generated environmental impacts like deteriorated spawning grounds of commercially valuable fish and shellfish. The ‗‗unprized‘‘ or ‗‗hidden‘‘ support from coastal and marine ecosystems to intensively cultured species is quite substantial. The presumption to locate farms in more exposed locations to reduce the environmental impact of organic enrichment by spreading the effects may be unfounded.

1.9. Positive reinforcing feedbacks of biogeochemistry and homeostasis shift ecosystems to new stable states; such shifts can be gradual or abrupt and communities may not return to their original state once the disturbance (in this case, altered nutrient loads) is removed. The catastrophic shift cannot be reversed by a correspondingly small reversal of the parameter variable; i.e. the trajectory of recovery is very different from the pathway of decline.

1.10. About 71 % of all nitrogen in the salmon feed is released as waste in the environment. Where the deposition modelling is important, it resembles a small part of the overall waste production. At least 80% of the total losses from fish farming are available nutrients. In undisturbed temperate marine ecosystems, nutrients are abundant during winter and early spring and are gradually depleted in the surface waters during the warm season, whereas in marine culture-impacted ecosystems most of the nutrient enrichment in the water column occurs during the warm period, i.e. summer and early autumn. The proportion of in the – 4 –

plankton communities will decrease after eutrophication where Harmful Algal Blooms increased dramatically.

1.11. HABs can be initiated in offshore waters (an upwelling source) and carried inshore, where anthropogenic nutrient sources affect their dynamics, through increasing magnitude, prolonging duration, number of toxins and toxic species identified, in numbers of fisheries affected, and in economic costs.

1.12. About 30.000 tonnes of salmon produces the nitrogen waste equivalent of what is produced by 420.000 people (about 3x the people living in Marlborough, Nelson and Tasman Region together )

1.13. The average total population of King Shags estimated to be 645 birds. The species is ―VULNERABLE‖, where this ―species is facing a high risk of extinction in the wild in the medium-term future. Duffers Reef on of the most affected colonies, is also one of the biggest colonies with the highest recruitment. Birds go on average not further than 18km away from the colony with most feeding between 6 and 8 km.

1.14. King Shag is a poor flyer requiring wind assistance on return trips to the colony in particular during chick raising. It is a deep diving species of between 20-50m and dependant on clear water to maintain minimal light conditions for prey pursuit. Increase of phytoplankton biomass may impact on the light penetration to the deeper water layers and benthic communities, potentially decreasing the area suitable for King Shags to hunt.

1.15. King Shag is dependent on deep benthic prey, in particular witch flounder. This species is most common in deeper water with grained sediment. Shags in general require a high density of prey species. Small declines can already have a severe impact on the viability of the species.

1.16. In Waitata Reach the applications are all in the zone where most King Shags are feeding. Also in the Queen Charlotte, there is overlap between King Shag feeding habitat and salmon farms.

1.17. In Waitata reach, the total amount of feed-use that is proposed, consented or under appeal is the equivalent amount of nitrogen waste from about 150.000 people. Impact of toxic algae on seabirds reveal an array of responses ranging from reduced feeding activity, inability to lay eggs, and loss of motor coordination and death (Shumway et al. 2003). Bird deaths caused by HABs – 5 –

have been widely reported (in Lewitus et al. 2012). Some of the dinoflagellate produced foam destroys the waterproof layer of feathers that keeps seabirds dry, restricting flight and leading to hypothermia.

2. Background

2.1. My name is Rob Schuckard. I immigrated with my family to New Zealand in 1989. I hold a Masters Degree in Biology (University of Amsterdam), with ornithology as my main subject. In the Netherlands I worked for the Department of Conservation. I had my own consultancy for ecological studies and surveys. I operated for six years an experimental dairy farm as a model to test profitability and environmentally sound management.

2.2. In New Zealand we live in the Marlborough Sounds where I operated between 1989 and 2000 two marine farms in the Pelorus Sound. We have integrated conservation and commercial aspirations for managing our land. About 25% of our land is planted with commercial forest (pine and macrocarpa trees) and 75% is in conservation management. Since 1999, we grow annually many thousands of natives, to be planted throughout our property as a part of our restoration work. We carry out active plants and mammalians pest-control. In 2005, our property received the Marlborough Rural Environment Award for Forestry and the Supreme Award for the work we have been carrying out on our property.

2.3. In New Zealand I have dedicated considerable amounts of time to ornithological surveys and studies. Most of my studies have been carried out as projects of the Ornithological Society of New Zealand. Main study projects:

Census coordinator for shorebird surveys in Top of South Island since 1995.

Team leader for catching and marking programmes of shorebirds in South Island since 2000.

Coordinator of Gannet study at Farewell Spit since 1995.

Long term studies and monitoring of New Zealand King Shag.

2.4. I have authored or co-authored a number of publications on the above subjects. – 6 –

2.5. I have been involved and still participate in a range of community and/or conservation projects:

Secretary Marine Reserve Committee of French Pass Residents Incorporated – 1993- 1998.

Member of the Marlborough Conservation Board in 1997 and 1998.

Environmental officer of French Pass Residents Incorporated – since 1997.

Committee member for Friends of Nelson Haven and Tasman Bay – 1998-2002 and since 2011.

Member of the Sounds Advisory Group – since 2010

Scientific advisor of Sustain our Sounds Inc. - since 2012

2.6. I have read and agree to abide by the code of expert witnesses as set out in the Environment Court‘s Practice Note 2011.

3. Scope of Evidence

3.1. I have been asked by Sustain our Sounds Inc. (―SOS‖) to provide an assessment of the impact of the New Zealand King Salmon Application on the ecosystem integrity of the Marlborough Sounds.

3.2. A literature review of eutrofication of coastal marine environments and the relation with harmful algal blooms is addressed.

3.3. I will compare the environmental impact of the proposed farm expansion on the habitat of New Zealand King Shag (Phalacrocorax carunculatus).

4. Status of ecosystem information from Coastal Marine Environment

4.1. A memo from the Ministry of Fisheries to the chair of Primary Production Committee (Hon Shane Arden, 23th February 2011 makes the following assessment: “Most problems from overseas salmon farms, including those in Chile, have been the result of having poorly located farms, too many farms in close proximity to each other, or overstocked farms, so that the assimilative capacity of the environment is exceeded.‖ This – 7 –

assessment is of relevance for the application of New Zealand King Salmon for the following reasons:

Definition of poorly located farms o The Marlborough Sounds as a semi enclosed water body is marginally suitable for finfish farming due to temperature and confinement

Too many farms in close proximity to each other

o The application of New Zealand King Salmon, in combination with their existing operation and other recently granted farms are creating concentrations of farming activities in a relatively small area with ecosystem significant pulses of nitrogen waste and its environmental implications

Overstocked farms

o A number of existing farming activities are not able to stay within the conditions of their consent. This relates to measured parameters, organic seabed pollution, the spatial footprint and the heavy metal depositions of copper and zinc.

4.2. An increasing number of scientists and resource managers recognise that successful marine management cannot occur without effective monitoring and evaluation. Lack of information and knowledge to develop policies are widely regarded as main problems in planning matters of the marine environment. Due to an ever increasing demand for marine products and services, management in the marine environment is made more complex by the lack of good data and information on which to base decisions.

4.3. Worldwide, the average biodiversity index fell by 40% between 1970 and 2000, where the marine index fell by about 30% over that period (Millennium Ecosystem Assessment 2005). At present, it is estimated that between 5 and 30 million of earth‘s animal and plant species have been described. About one species is extinguished every hour. Under natural conditions, the net growth rate of a species number is 0.37% in 1 million years where now the extinction rate has increased 10.000- fold; the decrease is at least 100 times higher than the loss of species in the past 65 million years. A loss of biodiversity cannot be tolerated for ecological, ethical, religious, aesthetic, and cultural – 8 –

reasons, all the more as the destruction of biodiversity is irreversible. To maintain biodiversity, to work out theoretical principles and translate them into practical measures is one of the major tasks of the next years. The maintenance of biodiversity is closely linked to the survival of man on earth, and has thus been incorporated into the concept of ―sustainable development.‖(Kratochwill 1999).

4.4. In 1998, the director-general of the Department of Conservation (Wallace et al. 1998) provided a bleak overview of the knowledge of our marine environment:

Less than one percent of New Zealand‘s marine area has been surveyed to assess the diversity of marine species and ecosystems.

We don‘t understand fully how increasing pressures of exploitation are affecting the marine environment generally, and how those pressures should be managed.

Marine species and ecosystem knowledgebase should increase rapidly and accurately, so that we do not end up with marine kakapo scenarios.

A change in the national mind-set to a more caring attitude of the marine environment is required to increase the level of understanding and appreciation of the complexity of marine ecosystems.

The most important vision from the director-general of the Department of Conservation for the future of the marine environment was through interagency cooperation:

Assist with the identification, classification and description of New Zealand‘s seascapes (within the resources available to us) and marine habitats and biota, and the threats to them.

4.5. Concern about this lack of knowledge in the marine environment was again highlighted in the New Zealand Biodiversity Strategy 2000:

Unlike natural areas on land, only a small number of marine habitats have been fully protected due to our very limited knowledge about New Zealand‘s marine biodiversity. Evaluating the state of New Zealand‘s marine and coastal biodiversity is difficult due to our very limited information. – 9 –

We have not yet identified and classified New Zealand‘s marine habitats, the communities of marine life within them, the processes that drive marine ecosystems, or the full extent of threats to marine biodiversity from human activities and broader environmental changes.

4.6. Also, the report by the Ministry of the Environment (2001) ―Confirmed indicators for the marine environment‖ identified eight key areas of concern. One of them is the limited environmental information in the marine zone.

4.7. Locally, the recently approved draft Nelson City Biodiversity Coastal and Marine Action Plan (Approved 30 March 2009) identified as a priority joint action on Coastal Marine Research and Monitoring: ‗Review existing coverage and undertake surveys of benthic marine habitats in Tasman Bay mapping both biodiversity hot spots and risk zones.’

4.8. Marlborough District Council and Department of Conservation conducted a survey to update significant marine species and habitats in the Marlborough Sounds (Davidson et al. 2011). This approach has been in accordance with section 4.4 of the MSRMP, where Council will:‘...encourages ongoing research to define significant ecological areas.’ In a 1994, 89 ecologically important sites were identified in the Marlborough Sounds, of which 79% were unprotected (Davidson et al. 1995). Due to the enormous size of the marine component of the study area, the survey was limited:‖… and with certainty, many sites with high ecological value are yet to be discovered.‖ The latest survey found 129 ecologically significant marine sites of which many are already degraded. The report about the surveys of significant areas also refers to the same limited scope of our marine ecosystem knowledge of the Marlborough Sounds:‖….large areas of Marlborough’s marine environment have never been surveyed and the knowledge of the ecosystem is limited.”

4.9. In his evidence, Mr. Davidson is sharing his knowledge of the Marlborough Sounds as an expert for the applicant. He provides the biological values from bio geographic areas where farm applications are located and provides known ―significant‖ marine sites within and immediately adjacent to applications as a tool of the assessment. It appears that the knowledge gained for this assessment is particularly based on recent dive surveys (white arrows) and unrelated to the applied sites of King Salmon. Where the latest dive surveys are without any doubt providing a contribution to the patchy knowledge of the sounds marine habitats, it appears that previous dive surveys have not been incorporated in his assessments. For example, dive surveys in the early 1990‘s are missing from the assessment of Mr. Davidson. The Tapipi site (yellow circle) was specifically mentioned in a report – 10 –

from DOC (Duffy, 1994) as a part of the Main Tidal Channel of Pelorus Sound: ―In places (e.g. Tapipi) large colonies of bryozoans Celleporaria agglutinans are found on firm sand between 18-24 m.

Also, in 2002, NIWA (Blackwell, 2002) was commissioned to assess the abundance of blue cod in the Marlborough Sounds. Of the sub area ―Outer Pelorus Sound‖, Katera Point (yellow circle) had the highest number of blue cod.

Mr. Davidson concludes in his evidence: ―Of the five salmon farm applications located in this area, none are located near known ―significant‖ sites‖. It appears that this assessment is strictly speaking correct but was better served with:‖Most of the applied King Salmon sites have not been surveyed for the MDC/DOC 2011 report, however information from the 1990‘s warrants additional dive surveys.‖ Where old ecological information has not been integrated in the assessment and no new surveys have been conducted, it appears the assessment‘s conclusions are premature.

Where the habitat called ―Main Tidal Channel of Pelorus Sound‖ will be most affected by the NZKS application, this habitat should have been surveyed for its potentially limited distribution and uniqueness in the Marlborough Sounds and New Zealand. The only map of the distribution of this habitat was produced in 1994 for the French Pass Marine Reserve Committee (Duffy, 1994). It appears that no further progress has been made with this earlier assessment and precaution of further development in this area should prevail. – 11 –

4.10. Morrison et al. (2009) promoted research on land-based and marine stressors, designed to help uncover and address impacts important for both land and fisheries managers to address. It may be noted that proposals for these studies have been very similar since the statements by the director general of conservation were made in 1998:

fundamental and systematic inventorying of fisheries species/habitat associations for different life stages, including how changing habitat landscapes may change the relative production of different fished species; – 12 –

better knowledge of connectivity between habitats and systems at large spatial scales, where impacts at one location may have far-field cascades into distant areas through subsequent fish movements;

the role of river plumes in affecting local ecosystem processes;

the effects of land-based stressors both directly on fished species, and indirectly through impacts on nursery habitats including plants (e.g., sea grass meadows, forests, maerl beds) and animals (e.g., mussel beds, bryozoans and tubeworm mounds, sponge gardens);

a better spatially based understanding of the integrated impacts of land-based and marine- based stressors on coastal marine ecosystems;

associated spatial mapping and synthesis to provide both decision support management systems, and research tools that can help direct and interpret new research initiatives.

3.10. Despite all the concerns raised the marine environment of the Marlborough Sounds has not been subject of a more caring attitude. No protected areas have been added since the establishment of Long Island Marine Reserve in 1993 while demands for products and services from the same marine environment have increased exponentially ever since. There is a lack of science based holistic ecosystem management and there is no financial basis for required monitoring. Incremental demands for ecosystem services have resulted in a cumulative impact on the ecosystem integrity without proper monitoring regimes in place. The only precautionary principle in the plan is the allocation of areas where aquaculture is prohibited, Coastal Marine Zone I (CMZI). Allowing cage aquaculture operations specifically in these CMZI areas is not reflecting the spirit of the Marlborough Sounds Plan to accommodate both aquaculture and other occupancies.

3.11. In New Zealand, ecosystem knowledge of the marine environment which was meant to be a precursor for resource consent applications of marine ecosystem services but never came to any fruition. The Department of Conservation‘s marine conservation technical capacity has a decade long history of declining mandates to increase in science and policy. New Zealand is well behind international action on marine protected areas required by the Convention on Biodiversity Action – 13 –

Plan. The anti-science approach to marine protection and stifling involvement of responsible agencies will leave an everlasting hiatus in an already declining marine biodiversity.

3.12. Lack of ecosystem knowledge, in combination with declining diversity of life-forms, will make adaptation to changing environmental conditions in the future more challenging. To maintain bio- diversity is to maintain our future options. The future challenges to already degrading coastal habitats will be exacerbated by predicted climate change and its impact on algal blooms (Al- Ghelani, 2005). Climate-induced changes in salinity, temperature and mixing, which all influence both oxygen conditions and species‘ hypoxia tolerance will be of importance. Climate change is a rather new phenomenon and only recently, planners attempt to integrate more and more the consequences of this new reality. The impacts of eutrofication, independent on the source of the flux, will be significantly influenced by this new reality. Both changes in climate forcing and nutrient loadings are aspects of global change that is expected to profoundly impact coastal hypoxia through more stratified water conditions. Planning towards these realities is not reflected in planning matters. The effects of large-scale climate warming are causing long-term variations in oxygen content and saturation as an observed increase in temperature has led to a general decrease in oxygen solubility of water masses. Mitigation of effects should reflect the realities of an uncertain future and we should not take comfort from the poorly known assimilation capabilities of the marine environment to date.

3.13. Identification, classification and description of seascapes are not a panacea for better marine management. Despite the survey and mapping of Australia's marine environment, it is acknowledged that sensitive marine environments became degraded around the country including macro algae beds, and benthic sedimentary communities. Because of their proximity to the land, near shore habitats are also exposed to influences such as sedimentation, turbidity, and nutrient enrichment. Some impacts from such influences have been regarded as fundamental to the productivity of coastal regimes, but it is now widely recognized that eutrophication resulting from excessive nutrient inputs has severely affected the quality of Australia‘s near shore waters.

3.14. Marine fish farming can have important physical, chemical and biological effects on the pelagic environment. Although modelling of water flows is an important tool overseas, such models are only beginning to be used in New Zealand. In 2002, it was assessed that there was considerable scope for additional research regarding pelagic effects of aquaculture activities (Cole, 2002). – 14 –

However, no comprehensive assessment of the water column impact at finfish farm sites in the Marlborough Sounds has ever occurred (e.g. Taylor et al. 2010). The Marlborough Sounds system is rarely nutrient limited and supports a large phytoplankton biomass. Local rivers deliver significant inputs of nitrates, largely originating from agricultural activities, into the inner and mid- Sounds following heavy rain. Upwelling in Cook Strait results in high concentrations of inorganic nutrients, which affect the outer Sounds immediately and move gradually into the mid-Sounds with tidal inflow (Rhodes et al. 2001). However, the Pelorus (NIWA Project No.: KMP00501 Kuku Mara Partnership 2000), relative to many other coastal waters around New Zealand, has very low nitrogen concentrations, predominantly in surface waters of embayments in the summer. The nitrogen impact of the NZKS application is profound, even compared to the natural fluxes of Cook Strait upwellings.

3.15. Infaunal organisms depend upon the supply of organic material from the overlying water as a source of food. By changing the amount, the quality and the specific timing of this supply it is evident that eutrofication but also marine acidification impacts on the pelagic zone could have damaging effects in the benthos. The impact of nutrient loading in relation to the scale and setting of this relative enclosed coastal habitat of the Marlborough Sounds is not reflected in this application. The impact on fast flow habitats as a medium to assimilate and dilute the waste of this proposal is profound and likely unmatched by any other application in the past.

5. Eutrofication, a worldwide concern and threat to coastal habitats.

5.1. Eutrophication is the leading cause of water quality impairment around the world (Diaz et al 2012). Eutrophication can simply be defined as the increase in the rate of production and accumulation (over-enrichment) of water with nutrients such as nitrogen and phosphorus as a result of human activity. The sources of nutrients potentially stimulating algal blooms include sewage, atmospheric deposition, groundwater flow, as well as agricultural and aquaculture runoff and discharge. In contrast to freshwater systems, nitrogen is the key limiting nutrient of primary production in most temperate estuaries and coastal marine ecosystems (Howarth et al. 2006); thus, nitrogen levels have a greater importance to understanding eutrophication problems in marine ecosystems. However, phosphorus also contributes to eutrophication and the input of both nutrients should be managed – 15 –

(Howarth et al. 2011, Chen et al. 2012). Whether a nutrient becomes a pollutant in an aquatic system, is a function of whether it is a limiting nutrient in a given environment and the magnitude of its concentration. Nitrogen and phosphorus loading into marine waters can initiate a biological process of eutrophication that, depending on the volume and duration of nutrient loading and the assimilative capacity of the receiving waters, can culminate in a fundamental shift in the food web structure of an area and lead to ecological simplification (McClelland et al. 1998; Ingrid et al. 1997; Worm et al. 1999; Worm et al. 2000a; Worm et al. 2000b). Even small changes in nutrients can have major impacts on phytoplankton communities. Eutrofication by human activities in the coastal environment has led to changed species composition and biomass of the benthic (bottom- dwelling) communities; eventually leading to reduced species diversity and increased dominance of gelatinous organisms such as jellyfish. (Purcell 2012).

5.2. An imbalance of nutrient ratios can lead to a shift in phytoplankton species composition and create conditions that are favourable to nuisance and toxic algal blooms. Harmful algal blooms (HABs) can cause shellfish poisoning in humans and mass kills of living marine resources. Different, sometimes competing factors promote preferable growth conditions for a specific group of HAB algae. These factors encompass the natural variability of hydrodynamics, light availability, nutrient loading, individual algal physiological and behavioural patterns, trophodynamic interactions/changes in the grazing community, anthropogenic eutrophication, and global change processes (Pettersson et al. 2013)Production of toxins in algae, for example, often occurs when nutrient stoichiometry1 is different from the classic Redfield Ratio2. Redfield ratio enables scientists to study the biochemical cycles and determine which nutrient might be limiting in the system or whether the nutrients in system are well balanced. Phytoplankton controls nutrient chemistry of oceanic waters through cycling & regeneration of nutrients. Toxic species can be harmful to higher trophic levels, disrupting normal ecosystem function. The dominance of toxic algae can result in a failure of normal predator-prey interactions, which in turn enhances the transfer of nutrients that sustain toxic species at the expense of competing algal species (Glibert, 2012). An increase in Mycrocystis abundance and bloom frequency in the San Francisco Estuary started in the late 1990s and escalated over the last decade (in: Lewitus et al. 2012). Changes in nutrient loading were linked to a cascade of changes in biochemical processes and other ecosystem properties: ... increased nutrient loads (eutrophication) together with changes in nutrient ratios

1 the calculation of measurable relationships of the reactants and products in a balanced chemical equation – 16 –

(stoichiometry) has led to changes in biogeochemical conditions (e.g., high N, high N:P) and trophic cascades (e.g., abundance of an invasive macrophyte, Egeria densa) that increases pH, and the invasive clam, that removes macrozooplankton and regenerates nutrients favouring proliferation of Microcystis.

5.3. Algal blooms have been implicated in a number of marine fish kills overseas (Treasurer et al. 2003) and in New Zealand (Jones et al. 1994):

1983 – North Island – fish and shellfish died – non-toxic Cerataulina pelagic – anoxia and toxin

1987 – Akaroa South Island – Farmed Salmon died – dinoflagellate Gyrodinium aureolum – toxin

1989 – Stewart Island – Farmed Salmon – dinoflagellate Heterosigma akashiwo – toxin

1993 – Wellington Harbour – Pilchards (Sardinops sagax) – microalga Tetraselmis sp. – anoxia and suffocation. 5.4. The list of algae species causing HAB‘s in New Zealand was extended in 2001(Rhodes et al. 2001):

5.5. The evaluation of phytoplankton biomass and HAB risk in the Marlborough Sounds has been assessed with routine phytoplankton monitoring throughout the sounds as a part of Ministry of Health and Marlborough Shellfish Quality Programme. Data between 2001 and 2004 show that >90% of samples scored low for phytoplankton biomass. Moderate to high scores in the

2 when nutrients are not limiting, the molar element ratio C: N: P in most phytoplankton is 106:16:1 – 17 –

remaining samples were not different between salmon farm and control locations outside of the influence of salmon farming activities. High and very high HAB risk in the form of shellfish and fish toxicity occurs near all the salmon farms, in particular during the summer months. However, similar HAB were also identified at the control sites. In 2004, at the weight of evidence considered it is suggested that nutrient waste from existing farms are not of notable environmental significance at current levels of production.‖(Hopkins et al. 2004).

5.6. The impact of incremental increases of nitrate concentrations was assessed for the Changjiang River estuary in China (Zhou et al. 2008). By 2000, nitrogen reached a level where dinoflagellate blooms are an annual event:

– 18 –

5.7. Also in China, Sishili Bay is the most important aquaculture and tourism area for the city of Yantai. Red tides occur now frequently and have caused huge economic losses in this bay in recent years. Red tides were caused by eutrophication from terrestrial inputs and local warm weather, particularly during rainy periods (Yanju et al. 2011).

5.8. Aquatic ecosystems respond to nutrient variation in complex and dynamic ways resulting in eutrophication, hypoxia/anoxia, acidification, and changes in phytoplankton and microbial communities and jellyfish blooms. For example, warmer more acidic ocean conditions appear to be favouring smaller phytoplankton species such as dinoflagellates, increasing the occurrence of harmful algal blooms. Ecosystem disruptive algal bloom events do not simply involve a direct stimulation of growth of harmful species by increased nutrients, but rather involve complex interactions among the growth of competing algal species, differential grazing on those species, and changes in nutrient cycling that are directly linked to algal grazing (Sunda et al. 2012)

The current challenges to scientific research and management include the facts that:

the link between nutrient dynamics and ecosystem responses is poorly understood;

monitoring data to support modelling and management are scarce;

aquatic ecosystems are site-specific and/or situation-specific and are highly dynamic, giving greater complexity in research and management;

lack of regional coordination in traditional management causes trans-boundary gaps.

To address these current challenges, an integrated management framework is needed for effective nutrient management. Institutional arrangements should be developed to coordinate across multiple government agencies and other stakeholders from watershed to coast (Chen et al. 2012). Also, modelling should reflect the high diversity of biotic habitat types that are present in marine systems with separate but complementary predictive frameworks for significant and sensitive marine habitats (Smith et al. 1999).

5.9. Suspended solids in the marine environment can form plumes of discoloured water in discharge areas and increased phytoplankton populations reduce light availability below the surface, and as a result, threaten a range of benthic ecosystems leading to a loss of submerged aquatic – 19 –

communities. The optically important water quality parameters are dissolved solid and suspended solid (Khanna, 2009). The unidirectional nature of light gives rise to a vertical gradient of light intensity as a function of depth.

5.10. The maximum depth of the light zone suitable for phytoplankton is designated as the euphotic depth. Light differs from all other resources because it cannot be mixed. Restricted light penetration by turbidity affects a range of sensitive underwater habitats (Boyd 2001). Algal blooms that occur as a result of high nutrients in marine waters will reduce water clarity (and consequently sunlight availably to other organisms in the water column and benthic communities), resulting in oxygen deficit of the water column when the organisms die, sink and decompose (Wetzel 1983). In freshwater lakes, if turbidity suddenly increases (for example, an input of nutrients that generates extensive phytoplankton blooms), benthic vegetation will disappear once a critical point of turbidity is reached. Loss of vegetation reduces sediment stabilization, which increases turbidity even further and the lake may eventually shift to a turbid water state (Nyström et al. 2012). Eutrophication leads to a shift in ecological structure in coastal bays(McGlathery et al. 2007) reflected in general by the following processes:

Long-term retention of recalcitrant dissolved and particulate organic matter will decline (e.g. sea grasses are replaced by algae with less refractory material.

Benthic grazers buffer the early effects of nutrient enrichment, but consumption rates will decline as physio-chemical conditions stress consumer populations. – 20 –

Mass transport of plant bound nutrients will increase because attached perennial macrophytes will be replaced by unattached short lasting algae that move with the water.

5.11. All mentioned processes result in less removal of N from the system. Grazing of bloom-forming algae is likely to buffer the effects of nutrient enrichment only at low to moderate nutrient loading rates. Some of these responses to eutrophication are reversible, although recovery is likely to be a slow process, partly because of continuing internal loading from the sediment that would support algal populations. As light availability increases following a reduction in external nutrient loading, benthic primary producers once again become important and can accelerate the recovery process by re-oxygenating the sediment, intercepting the sediment–water column nutrient flux, and temporarily retaining nutrients in plant biomass.

5.12. The past few decades have seen a massive increase in coastal eutrophication globally, leading to widespread hypoxia and anoxia, habitat degradation, alteration of food-web structure, loss of biodiversity, and increased frequency, spatial extent, and duration of harmful algal blooms. Much of this eutrophication is due to increased inputs of anthropogenic nitrogen loading in marine waters and ultimately to coastal oceans (Howarth 1998, Levin et al. 2001). The adverse impacts of nutrient discharges into the marine environment have been demonstrated in numerous studies (Cripps et al. 1996; Hardgrave et al. 1997; Hardy 2001b; Mac Garvin 2000; Pohle et al. 2001; Sutherland et al. 2001; Wildish et al. 2004; Jessop et al. 2007) stimulating blooms of phytoplankton or microalgae populations. Anthropogenic nitrogen loading is acknowledged as the principal cause of degradation and alteration to coastal ecosystems worldwide (Bell et al. 1995; Paerl 1997; Howarth 1998; Seitzinger et al. 1999; Wu 1999). The Joint Group of Experts on the Scientific Aspects of Marine Pollution (GESAMP 1990) outlined the following biological and ecological changes that take place as eutrophication progresses:

increased primary production;

changes in plant species composition;

very dense, often toxic, algal blooms;

conditions of hypoxia (low oxygen concentration) or anoxia (no oxygen);

adverse effects on and ;

changes in the structure of benthic communities. – 21 –

5.13. Selman et al. (2008) came to a very similar assessment of the impact of eutrophication. Eutrophication can be harmful to both freshwater and marine ecosystems, and leads to a progression of symptoms that include:

Excessive phytoplankton and macro algal growth that is the source of organic carbon for accumulation. This can also reduce light penetration and lead to a loss of submerged aquatic vegetation.

An imbalance of nutrient ratios that can lead to a shift in phytoplankton species composition and creating conditions that are favourable to nuisance and toxic algal blooms. Harmful algal blooms (HABS) can cause kills of living marine resources and shellfish poisoning in humans.

Changes in species composition and biomass of the benthic (bottom-dwelling) community; eventually leading to reduced species diversity and increased dominance of gelatinous organisms such as jellyfish.

Low dissolved oxygen and formation of hypoxic or dead zones (oxygen-depleted waters). These oxygen-starved areas stress aquatic ecosystems, often leading to kills of living marine resources, altered ecosystem energy flows, and in severe cases ecosystem collapse.

5.14. Climate warming already affects phytoplankton species composition and spatial distribution, and favours species that are best adapted to the changing conditions. Shifts in phytoplankton can have far-reaching consequences for ecosystem structure and functioning. The most significant climatic effects on phytoplankton species composition will very likely be mediated through changes in thermal stratification patterns such as the extent of the growing season and vertical mixing processes (Winder et al. 2012). Both changes in climate forcing and nutrient loadings are aspects of global change that are expected to profoundly impact coastal hypoxia (Diaz et al. 2008). So in addition to anthropogenic nutrient inputs, global warming with rising temperatures may increase the release of dissolved organic matter from soil and alter vegetation cover, which may in turn elevate nutrient and organic matter export. Dissolved oxygen is a very basic requirement for aquaculture species. Dissolved oxygen is regarded to be the most critical water quality variable in aquaculture (Boyd et al. 1989). Anoxia occurs when dissolved oxygen levels in the environment – 22 –

decrease to the point where aquatic life can no longer be supported (Anonymous, 2000). Currently there are over 500 hypoxic systems covering over 240.000 km2 around the globe related to human activities (Diaz et al. 2012). Increasing nutrient concentration strongly increase the probability of hypoxic events, stressing the importance of continued extensive eutrophication management. Harmful algal blooms do now have severe impacts in the Gulf of Mexico and along the Atlantic and Pacific coast of the USA, resulting in fish kills and severe human health problems (Magaña et al., 2003, Lewitus et al. 2012).

5.15. Most of the scientific monitoring of a salmon operation involves the area close to and under the cages. The environmental monitoring is primarily dedicated to maintain a healthy environment for the fish, in contrast to providing an impact assessment of the wider health of the marine environment. Guidelines are used to monitor the suitability of the marine space for fish farming without determining the wider environmental impact of fish farm practises (Heinig, 2001).

5.16. Ecosystem, species and interspecies responses to hypoxic events can be quite different and before the threshold of hypoxia (1.41mg/L) is reached there may still be an effect on the ecosystem. Many demersal fish such as sole (Solea solea) and plaice (Pleuronectes platessa) which are important species in the northern hemisphere, are known to be affected already from concentration of 3.04mg/L (Meire, 2011). Atlantic cod (Gadus morhua) growth in the St. Laurence (USA) is reduced below about 7 mg O2/L. Shrimp and fish avoid dissolved oxygen below 2 mg O2/L in the northern Gulf of Mexico hypoxic zone, while sharks and rays emigrate from the area at oxygen concentrations below 3 mg O2/L (in: Diaz et al., 2012). Long-term decline in oxygen conditions in the Baltic Sea has had a seemingly generally negative impact on oxygen-related processes for the different life stages of eastern Baltic cod (Gadus morhua) and stock-specific processes (e.g. survival rates of eggs, settlement probability of juveniles, habitat utilization of spawning fish, age structure of successful spawners, food consumption rates of adult fish (Hinrichsen et al. 2011). Intensive aquaculture has generated environmental impacts that deteriorate spawning grounds of commercially valuable fish and shellfish and the ‗‗unprized‘‘ or ‗‗hidden‘‘ support from coastal and marine ecosystems to intensively cultured species is quite substantial (Folke et al. 1998).

5.17. Slow recovery processes from eutrophication in severely affected coastal environments have been recorded. Decreased nutrient loadings through the Baltic Sea Action Plan may not only have increased the food supply for benthivorous fish in currently hypoxic areas, but also counteract – 23 –

other negative effects of hypoxia such as decreased survival rate of eggs and settlement of juveniles. (Timmerman et al. 2012).

5.18. The impact of eutrophication on coastal biodiversity has been reported in Morrison et al. (2009):

Diverse benthic communities may disappear and be replaced by ones dominated by deposit-feeding annelids. Tracking of eutrophication effects in Scotland showed a replacement of seagrass by green algae, along with a change from a crustacean dominated assemblage…., to a benthos-poor algal matt …. . Similar processes were seen in the Baltic Sea in a number of places, leading to a loss of more than 40 macrophyte species, all replaced by a single species of brown filamentous alga. Associated with this was a drop in 33 associated in- faunal species, important as prey for fish. Fish spawning (perch and pike) grounds were also lost with the disappearance of plants. 5.19. The assimilation of nitrate by coastal phytoplankton and its conversion into organic matter is an important feature of the aquatic nitrogen cycle. Where Morrison et al. (2009) mainly deal with terrestrial sources of eutrofication, whether the nutrient source is land-based, originates directly from anthropogenic activities in the coastal marine environment or both is not relevant for its effect. Terrestrial ecosystems are becoming increasingly nitrogen-saturated due to anthropogenic activities, such as agricultural loading with artificial fertilizer. Thus, more and more reactive nitrogen is entering streams and rivers, primarily as nitrate, where it is eventually transported towards the coastal zone. As nitrate is the major plant limiting nutrient in seawater (most phytoplankton grows well at a nitrogen: phosphorus ratio of 10:1), so high nitrate levels can result in eutrophication and excessive nuisance algal and plant growth. This can have negative effects on culture species and can result in deaths due to changes in oxygen/carbon dioxide levels. Dissolved reactive nitrogen is converted into a particulate form, which eventually undergoes nitrogen removal via microbial denitrification. High and unbalanced nitrate loads to the coastal zone may alter planktonic nitrate assimilation efficiency, due to the narrow stochiometric requirements for nutrients typically shown by these organisms. This implies a cascade of changes for the cycling of other elements, such as carbon, with unknown consequences at the ecosystem level.

– 24 –

Current expert opinion is that unless urgent intervention occurs, the lagoon will almost certainly undergo a rapid change in state to an even more degraded phytoplankton dominated system (e.g. algal bloom), which would endanger the Ruppia community and change the fundamental values and character of the lagoon (Lagoon Technical Group (LTG) 2011, cited in Robertson et al. 2011). These types of system changes often represent a hysteresis function (Petraitis and Dudgeon 2004; Webster and Harris 2004). In an hysteresis, after an initial trajectory of change, only a small additional change in a parameter variable (e.g. nutrient input) can result in a catastrophic shift in a state variable (e.g. Ruppia decline, phytoplankton increase). The catastrophic shift cannot be reversed by a correspondingly small reversal of the parameter variable; i.e. the trajectory of recovery is very different from the pathway of decline (Petraitis and Dudgeon 2004, Lester and Fairweather 2008). In simple terms: if the system tips, the causal factor needs to be changed by a large amount to bring it back – this means that it is much more expensive and difficult to restore than it is to protect.

Case Study: Nutrient Loads to Protect Environmental Values in Waituna Lagoon, Southland NZ (Scanes, 2012) 5.20. The degree to which marine ecosystems may support pelagic- or benthic food chain has been shown to vary across natural and anthropogenic gradients in e.g., temperature and nutrient availability. Moreover, such external forcing may not only affect the flux of organic matter, but could trigger large and abrupt changes, i.e., trophic cascades and ecological regime shifts, which once having occurred may prove potentially irreversible. A potential regime shift from pelagic to benthic regulatory pathways has been reported in Kattegat (Denmark) as a possible first sign of recovery from eutrophication likely triggered by drastic nutrient reductions (involving both nitrogen and phosphorus), in combination with climate-driven changes in local environmental conditions (e.g., temperature and oxygen concentrations) (Lindegren et al. 2012)

5.21. Relocation of fish farms to areas with strong currents, is unlikely to prevent detrimental effects to the structure and organisation of the benthos (Lee et.al. 2006), and ‗fallowing‘ (whereby sites are left un-stocked for a period of time to allow benthic recovery) is inadvisable where slow-growing biogenic habitats such as maerl (also known as rhodoliths) are concerned, as this may expand the area impacted (Hall-Spencer et al. 2006). The latter recorded general reductions in biodiversity were crustacean communities being particularly impoverished in the vicinity of cages, and significant increases in the abundance of species tolerant of organic enrichment (e.g. Capitella spp.). Areas with diverse communities tend to have a wider range of ecological functions, including species‘ mobility and reproductive strategies, and such communities will take longer to recover from salmon farm practises than those where diversity is low and the communities are simple (Macleod et al. 2007, Thrush et al. 2001). Consequently, impacts will be more significant in areas with inherently high diversity and the presumption to locate farms in more exposed locations to reduce the environmental impact of organic enrichment by spreading the effects may in fact be – 25 –

unfounded (Macleod et al. 2006). This study shows that, under similar farming impacts, there was a greater change in the benthic in-faunal community and ecosystem function at the more exposed location compared to the more sheltered location and that the recovery response of the exposed site was slower. In addition, the overall area affected by organic deposition will be greater at exposed locations compared with more sheltered sites because the current flow and or tidal influences are greater, thus increasing the field of dispersal. The fauna at more sheltered locations where organic-rich sediments accumulate may actually have a natural resilience to organic loading, being ecologically and functionally pre-adapted to cope with an increased level of organic enrichment. The disparity between the footprints in low flow areas (most of the existing farms) compared to applied high flow areas is also shown in the NZKS application. Where zone 3 (outer edge of benthic impacted area) is not allowed to go beyond 150m from an existing farm, this has now changed to up to 900 m in this application3.

5.22. In deep water fish farms, organic waste affects the benthic community on a much larger spatial scale than at shallow water sites (Kutti et al. 2007b). The differences in the recovery time with location further reinforce the contention that managing recovery should take into account features of the receiving environment such as sediment type, organic matter content and ecological function of the resident infauna.

3 Draft Conditions – Sarah Dawson – 26 –

5.23. Corallines are distributed worldwide and are abundant in a number of diverse marine habitats, including rocky shores, sea grass meadows, tropical reefs, and rhodolith beds (or maerl) (Harvey et.al. 2005). They also have a considerable depth range, occurring from the intertidal zone down to 270 m, and are the deepest known macroscopic plant life. Corallines are important components of marine environments. Maerl beds are structurally and functionally complex habitats that support rich and diverse assemblages and host many species unique to those habitats. There is also growing evidence that maerl beds have considerable value as nursery grounds for marine species of commercial interest (Barbera et al. 2003) and can be affected by salmon farm operations (Hall- Spencer et al. 2006).

5.24. Biogeochemical pathways together with homeostatic control serve to provide the mechanism(s) whereby nutrient dynamics support trophodynamic structure. Positive reinforcing feedbacks of biogeochemistry and homeostasis shift ecosystems to new stable states; such shifts can be gradual or abrupt and communities may not return to their original state once the disturbance (in this case, altered nutrient loads) is removed (Glibert, 2012). Lags in recovery called ‗‗hysteresis‘‘ can be protracted and in some cases result in ecosystems becoming locked into degraded conditions, which can be described as being in an alternative stable state (in Nyström et al., 2012, McGlathery et al. 2007). The nitrate removal efficiency of a natural phytoplankton community decreases under high, unbalanced nitrate loads, due to the enhanced recycling of organic nitrogen and subsequent production and microbial transformation of excess ammonium. Nitrate removal efficiency was inversely correlated with the amount of nitrate present and mechanistically controlled by dissolved organic nitrogen, and organic carbon availability. These findings have important implications for the management of nutrient runoff to coastal zones (Lundau et al. 2012) and aquaculture.

6. Salmon farming and eutrofication

6.1. Mariculture in New Zealand is based mostly on filter-feeding molluscs, whereas sea cage finfish farms, which result in the introduction of ‗‗new‘‘ nutrients to the environment, are few. Since nitrogen and phosphorus are loaded from fish cages, there is always the potential for fish culture to promote eutrophic conditions; either by supplying a readily available nutrient source directly to phytoplankton; or oxygen removal, accompanied by nutrient releases, via the decomposition of waste solids. Although mussel farms do cause changes in the dynamics of nutrient – 27 –

remineralisation processes within the water column and sediments, the probable net result is the removal of nutrients from the ecosystem. To date, there is no evidence that the considerable expansion in mussel farming in the Marlborough Sounds (the major shellfish growing area) over the last decade has resulted in any increased incidence of HABs (Rhodes et al. 2001). New Zealand is not in the top or bottom ranking countries in the Mariculture Sustainable Index. New Zealand is currently not a significant mariculture producer by world standards and much of the production is molluscs, where finfish farming is only starting. If imports of fish feed were included in the assessment for the Sustainability Index, the ranking would change to a lower position (Trujillo, 2008). Production of feed makes up the single most important contributor to the greenhouse gas emissions of the finfish farming industry. For and Rainbow trout production, feed accounts for, on average, 87% of total greenhouse gas emissions (Parker, 2012).

As aquaculture continues to grow, there will be a need for a more comprehensive waste management, not only for cage aquaculture systems, but for other systems as well. The means available for reducing the potential environmental impact is highly diverse and include feeding management, site selection, alteration of site, and an active use of wastes from feeding aquaculture to produce organisms on other trophic levels, organisms that can take advantage of the wastes from cage aquaculture systems (Integrated multi-trophic aquaculture). The environmental impacts of sediments and benthic ecosystems are relatively well understood and managed, but there is a need to improve the general understanding of how pelagic ecosystems are impacted in order to establish a science based management and monitoring practices for open waters. This is important for the societal perception of aquaculture, but also for the industry itself which require pure water for its activity. Limitations of comprehensive water management (Olsen et al. 2008). 6.2. Where nitrogen and phosphorus are loaded from fish cages, there is potential for fish culture to promote eutrophic conditions; either by supplying a readily available nutrient source directly available to phytoplankton or oxygen removal via decomposition of waste solids. Considering there is now 5 times more salmon living in fish farms than have lived at any one time on earth, their waste output is not at "natural" levels. Despite better food quality and improved feeding routines which have resulted in less excess nutrients per kilo fish produced, the overall discharges have increased due to the large increase in total fish production (Skogen et al. 2009). In Europe alone, the total production of salmon in fish farms (mainly in Norway and Scotland) has increased from 70,000 t in 1990 to 148,000 t in 1996 (EEA 2002) up to 540,000 t in Norway alone in 2003 (EEA 2006a) (Airoldi et al. 2007). The global culture of salmonids (i.e. trout, salmon, charr) – 28 –

doubled from 0.94 million MT in 1995 to 1.99 million MT in 2005 (FAO 2007). World farmed salmon production first exceeded wild salmon production in 1998 (FAO 2007).

6.3. Aquaculture systems now account for roughly 35% of total fish production (FAO, 2009). Finfish accounts for about half of aquaculture production followed by mollusks and plants each 20% to 23% (Diaz et al. 2012). In 2003, aquaculture consumed 53% of fish meal and 87% of fish oil world production (Tacon, 2005). This figure has reached 63% for fish meal and 81% for fish oil in 2009 according to IFFO (International Fish meal and Fish Oil Organization). The world supply of fish meal and oil might not sustain the demand of aquaculture if the sector maintains its actual growth rate. Where it takes about 4.9kg of wild fish to grow 1 kg of salmon (in Regnier et.al. 2012, doubt has arisen about the optimism of expansion of aquaculture to meet the growing seafood demand around the world. A decline of 1.2% per year in the year-on-year growth rate in global farmed salmon production has happened since it peaked in 1966 (Liu et al. 2008). Hence the primary benefits of aquaculture are generally thought to stem from direct economic benefits rather than environmental benefits in terms of enhancing depleted fisheries (Gibbs, 2006). The capacity of marine and coastal ecosystems to produce seafood is not included in the signals that guide economic development. For example, fish farming in cages is, for its daily survival, dependent on marine ecosystem areas as large as 10 000–50 000 times the area of the cages for producing its food, and 100–200 times for processing parts of its waste. Hence, to perceive fish farming as it really is and to deal with its problems one must expand one‘s thinking and action far beyond the site of the farm (Folke et al. 1998). The assumption that the world needs not to worry about the pending demise of capture fisheries because aquaculture can replace the shortfall, may be unfounded. Also, where the rapid development of a highly industrialized production of farmed salmon has contributed to a strong belief in continued growth in aquaculture, infectious diseases represent a potent density dependent negative feedback mechanism that may limit such growth (Jansen et al., 2012). Increasing nutrient input indirectly affects the abundance of pathogens, which may lead to an increase in human and wildlife diseases (Johnson et al. 2010).

6.4. Salmon produced from aquaculture are efficient at converting feed to flesh. In 1980 to produce 1kg of salmon required almost 3kg of feed while in 1995 just over 1 kg was required (Asche et al. 2011). New feeding systems have contributed to reducing the FCR by lowering feed waste. After 1995, this figure has levelled off to a feed conversion ratio (FCR) between 1.0 and 1.5 for Atlantic salmon. With FCR levels of the moment, it seems unlikely that further gains will be developed. – 29 –

For example, 1 kilogram of salmon can be produced with as little as 1.1 to 1.8 kilograms of feed (Goldberg et al. 1997). Up to 1995, feed accounts for 35% - 60% of the cost of salmon farming (Forster, 1995; Higgs et al. 1995), by far the largest operational expense. Depending on the species, this figure now represents 40% to 70% of the total production costs (in: Regnier et al. 2012) Significant improvements have been made in the feed conversion ratio for salmon farming. Notwithstanding the improvements and independent of the efficiency to converse feed to flesh, the total amount of waste produced by fish farms is significant in relation to the natural fluxes from the marine environment.

6.5. Researchers around the world have recognized the potential harm from open net cage finfish farming and the long-term impacts on water quality, fisheries resources and sea-bed ecology. The impacts of fish cage aquaculture through the accumulation of waste products primarily of uneaten fish feed, faecal matter and other excretory wastes can be generally separated into water column and sea floor (or benthic) effects. Because of the dissolved or suspended nature of the pollutants, water column effects are far more difficult to measure compared to benthic effects. Benthic effects are measurable by changes in communities. For example, Capitella capitata, a polychaete, responds to intermediate levels of organic enrichment and reach peak abundances before conditions in the environment are strongly enriched. They are also indicator species that are increasingly more tolerant to attribution resulting from enrichment anoxia and increasingly better competitors under hyper- nitrification (Lee et.al. 2006, Tsutsumi et al., 1991). C. capitata is indicative for polluted and semi-polluted conditions and was the most dominant species at the farm stations, but was almost absent at the control stations (Neofitou et al. 2010 ). The bottom and macro faunal communities can change very rapidly with the onset of farming with significant reductions in species diversity and increases in bacterial biomass within two months of commencing operation (Pohle et al. (2001). While the cause could not be identified, the most probable explanation is that an aspect of enrichment linked to salmon farming caused changes in benthic-pelagic coupling in such a way as to exclude some species and encourage others.

6.6. Simulation and prediction of the environmental impacts of net-pen aquaculture on the water column, as well as benthic effects are discussed by Silvert (2000) and in particular Stigebrandt (2004). A model developed by the latter, estimated the holding capacity of sites for fish farming. Expressed in terms of maximum fish production per month, the holding capacity is estimated with regard to three basic environmental requirements: – 30 –

the benthic fauna at a farm site must not be allowed to disappear due to accumulation of organic material;

the water quality in the net pens must be kept high;

the water quality in the areas surrounding the farm must not deteriorate.

6.7. Considerable amounts of nutrients are released into the water column by cage farming, although there is some variability in literature about the amount of suspended and deposited waste. The largest volume of waste is fecal matter (Beveridge et al., 1991; Chen et al. 1999). Stigebrandt et al. (2004) calculated that for each kilo of fish produced, 0.3 kg of the supplied feed is not ingested by the fish. Other studies suggest a much smaller amount with a proportion of feed that is not ingested ranging from 5 to 15% (Hall-Spencer 2006, Thorpe et al., 1990; Juell, 1991; Blyth et al., 1993; Findlay et al., 1994; Wu et al., 1994; Beveridge, 1996; Beveridge et al., 1997; Cho et al., 1997). The exact amount and composition of the waste varies with different food types and feeding regimes but with an estimated release of particulate organic matter of around 10% or 25% of the dry weight of the feed consumed, discharges of particulate organic matter from one single farm may be as high as 1300 to 3250 kg per day at peak production (Kutti et al. 2007a). Isotopic examination (δ13C and δ15N) of organic matter sources and consumers was used to assess the impact and trace the dispersal of wastewater from a land based fish farm in western Mediterranean. Aquaculture waste entered the food web, altering the natural isotopic composition of organic matter sources at the base and the upper trophic levels. Waste seemed to disperse widely enough to affect the isotopic composition at the study site about 500 m from the outfall, while sites at 1 and 2 km from the outfall showed values that were similar to each other and different from those of the impacted site. The impact was detected at different ecosystem levels, although primary producers were more affected by fish farm waste, taking up aquaculture-derived nutrients (Vizzini et al., 2004). In Schotland enhancements of nutrients were detected at distances of more than 200 m and were believed to have been derived from the salmon farm (Sanderson et al. 2008). Most of the total nitrogen in the wastes is in the dissolved fraction, while the majority of the total phosphorus is in the particulate fraction (Costa-Pierce 1996). Dissolved inorganic nitrogen and dissolved inorganic phosphorus are most important drivers of change to pelagic ecosystems, while particulate organic nitrogen and particulate organic phosphorus from feces and feeds are the major drivers of benthic ecosystem change (Buschmann et al. 2007). – 31 –

6.8. About 85% of the waste will be in dissolved forms (ammonium, urea, nitrate, together called dissolved inorganic nitrogen DIN), and the rest is in particulate form (Zeldis, 2008). In Norwegian studies, the majority of the N waste is released to open waters (68% of total) whereas the majority of the P is accumulated in sediments (63%) (Olsen et al. 2008, Kutti et al. 2007a). High concentrations of P in sediments have been found to be a useful indicator that the sediment is affected by aquaculture activity (Karakassis et al. 1999). Wu (1995) shows that 23% of C, 21% of N and 53% of P of feed input is being accumulated in the bottom sediments and the significant impact is normally confined to within 1 km of the farm. Mente et al. (2006) recorded 50% of the N and 28% of the P supplied with the food is released as waste in dissolved form , very similar to yellowtail kingfish culture Seriola lalandi in South Australia where 60% of N in feed inputs is lost as soluble excretion products (Fernandes et al. 2008). The clearest impacts of the culture of yellowtail kingfish Seriola lalandi in South Australia was an 81% increase in ammonia concentrations adjacent to cages relative to controls (Tanner et al. 2010). Other studies show that deposited waste of up to 98% is re-suspended in days and moves to other areas (Milewski, 2006). At least 80% of the total losses from fish farming are plankton available nutrients and other potentially eutrophicating substances. Eutrophication is however indicated by far more signs than increases in primary production of phytoplankton. Other effects, such as changes in the energy and nutrient fluxes, changes in pelagic and benthic biomasses and community structure, changes in fish stocks, sedimentation, nutrient cycling, oxygen depletion and shifts between perennial and filamentous benthic algae, may be more sensitive and relevant indicators.

6.9. Nutrients discharged by fish farming zones into the water column are not easily detectable at the distances 2-3 nm from the fish farming zones in the eastern Mediterranean (Pitta et al. 2005). This is due to both the dispersive nature of these zones which allows rapid advection and diffusion as well as to the rapid uptake of nutrients by plankton organisms and further transfer to higher trophic levels. A study on a 2910 tonnes salmon farm in a Norwegian Fjord found an annual vertical flux of particulate organic carbon to the bottom adjacent to the farm of 365 g m−2, nine times as high as what was found 3 km (Kutti et al. 2007b). Some of the components of the organic waste were transported as far as 550 to 900 m, probably due to re-suspension of surface sediment. Despite the high sedimentation rates the content of sedimentary organic matter, total organic carbon and total organic nitrogen was not elevated in the sediment around the farm. However, phosphorus was found in higher concentrations in the sediments close to the farm, indicating that – 32 –

organic matter had settled on the sediment and been decomposed. There is a reasonable scientific consensus about the percentage of total nitrogen and phosphorus input from feed that is lost to the aquatic environment. In general term, about 60%-80% of all the nitrogen and phosphorus in feed will be released into the environment mostly dissolved.

6.10. Forrest et al. (2007) concluded that the effects on the seabed are often highly localised and largely reversible in the medium to long term. While the ecological significance of the enrichment impact on the local seabed may be relatively high, in absolute terms the broader consequences can be mitigated by appropriate site selection. However, the study does not distinguish different type of seabed habitats with different responses to enrichment (see: 4.21.). Sediments in some areas have greater resilience to organic inputs (MacLeod, 2007). Phosphate however, which does not induce stress on macro fauna during fish farming, triggered benthic algal production after the cessation of aquaculture activities. This enhancement of primary production had maximal values during summer, when light availability on the sea bed is higher, particularly after removal of the cages and the consequent elimination of shading due to the floating structures and the dust from the fish feed (Karakassis et al. 1999). In undisturbed temperate marine ecosystems, nutrients are abundant during winter and early spring and are gradually depleted in the surface waters during the warm season, whereas in marine culture-impacted ecosystems most of the nutrient enrichment in the water column occurs during the warm period, i.e. summer and early autumn (Mente et al., 2006) changing the natural cycle of nutrient loading. The highest fluxes of nitrogen from salmon farms do occur in the summer, when in natural circumstances, nitrogen is at its lowest annual level due to assimilation by the coastal marine benthic and water column ecosystems. – 33 –

6.11. Strong correlations have been demonstrated between total nitrogen input and phytoplankton production in estuarine and marine waters (Anderson et al. 2002). Elevated nutrient concentrations, along with climatic conditions, can contribute to blooms of plankton and toxic algae (MacGarvin 2000, Staniford 2002) and may also lead to changes in phytoplankton community composition via altered nutrient ratios. An increase in N: Si ratio favours the growth of flagellates rather than diatoms (Officer et al. 1980, Smayda, 1990). Also laboratory culture competition experiments with multiple marine phytoplankton species and mesocosm experiments show that diatoms dominate at high Si: N ratios, while diverse flagellates tend to dominate at low Si: N ratios (Signorini et al. 2012). Declines in silicate are widely expected in a warming world and a warmer ocean, where the vertical delivery of nutrients to the surface may be slowed due to stratification. This change in availability of silicate is a contributing factor to the restricted abundance of diatoms (in: Hallegraeff et al. 2010). Where ciliates and other diminutive heterotrophs are not grazed directly by higher order consumers like larval fishes, energy is transferred less efficiently through the food chain (Parris 2012). Dinoflagellates often dominate surface stratified waters (Winder et al. 2012); in temperate zones there may be a succession from diatoms to dinoflagellates as the relatively nutrient rich, well mixed water column of spring stabilizes to form a stratified water column with relatively warm, nutrient poor surface waters. The period from 2-13 January, 1989, a developed in Big Glory Bay, Stewart Island. Fish kills occurred and the weather was characterised by warm air and sea water temperatures, long daylight hours, light to moderate winds from an easterly quarter and an episode of intense rainfall. These conditions would result in runoff from the catchment being retained within the bay resulting in fertilization of the water and contributing towards the establishment of a stable stratified water column. These conditions were clearly suitable for the development of a – 34 –

flagellate dominated population. Conversely the same period the following year was characterised by cool air and sea water temperatures, the same day-length (but probably lower solar radiation), and moderate to strong winds from the west accompanied by intense cool rainfall. These conditions would have induced current flows that would have reduced the retention time, mixed and flushed out of the bay runoff from the catchment, promoted turbulence and prevented the establishment of a stratified water column. These conditions may have prevented the repetition of the flagellate bloom of the previous summer by inducing the development of a diatom dominated population instead (MacKenzie, 1991). The reference to runoff from the catchment of Big Glory Bay is somewhat surprising. This part of Stewart Island is very much in its original pristine state, completely surrounded by mature native forest. This in contrast to many parts of the Marlborough Sounds where land run off can be substantial.

6.12. No published nutrient data for Big Glory Bay were available before the 1989 Heterosigma bloom. However, throughout Big Glory Bay in early 1988, both the levels of inorganic nitrogen nutrients and dissolved reactive phosphate were very high; the greatest concentrations of both nutrients recorded were 48 and 14 mg/m-3, respectively (Pridmore pers. comm.). The bulk of nutrients in Big Glory Bay may have originated from Foveaux Strait through water exchange. Excretory products and nutrients leached from excess fish feed from the salmon farms might have contributed to some of this pool, but the relative significance of this source is not known (Chan et al. 1990). Under different combinations of light and N sources the maximum growth of H. carterae – 35 –

in the field is probably attained in the well-illuminated, upper water column with NO3– as the major N source (Chan et al. 1995).

6.13. Vertical mixing is one of the key variables that condition the growth performance of phytoplankton within the water column because mixing processes are usually accompanied by changes in resource availability of light and nutrients. Vertical mixing of natural waters is largely determined by meteorological variables (Winder et al. 2012). Parts of the Marlborough Sounds like Beatrix Bay are already strongly stratified, mainly caused by a difference in salinity (Sutton et al. 1997).

6.14. Although algal blooms, including those considered toxic or harmful, can be natural phenomena, the nature of the global problem of harmful algal blooms has expanded both in extent and its public perception over the last several decades. Of concern is the potential relationship between harmful algae blooms and the accelerated eutrophication of coastal waters from human activities (Anderson et al. 2002). There are two primary factors causing HABs outbreaks:

natural processes such as upwelling and relaxation,

anthropogenic loading resulting in eutrophication.

6.15. Coastal HAB monitoring and mitigation has become a management priority along the full length of the North American West coast, reflecting a common recognition that HABs and their impacts are increasing and can have a profound effect on the health and economies of their coastal communities (Lewitus et al. 2012). Lewitus links HAB‘s to coastal eutrophication but HABs can also be initiated in offshore waters (an upwelling source) and carried inshore, where anthropogenic nutrient sources affect their dynamics, through increasing magnitude and prolonging duration. Inorganic dissolved nitrogen and phosphorus supply to surface waters in the eastern tropical South Pacific is influenced by expanding oxygen minimum zones, since N loss occurs due to microbial processes under anoxic conditions while P is increasingly released from the shelf sediments. The impact of decreasing N:P supply ratios in the Peruvian Upwelling indicated that some organisms were able to benefit from low N:P fertilization ratios, especially Heterosigma sp. and Phaeocystis globosa which are notorious for forming blooms that are toxic or inadequate for mesozooplankton nutrition (Hauss et al. 2012). Overall, in many parts of the world, marine and freshwater HABs are increasing in geographic extent, in duration of occurrences, in numbers of – 36 –

toxins and toxic species identified, in numbers of fisheries affected, and in economic costs (Johnson et al. 2010).

6.16. The presence of salmon farms significantly increases density pulses of dinoflagellates (Buschmann et al. 2006). In an experiment to assess the effect of salmon farms on phytoplankton communities, tanks, filled with either effluent from salmon farms or with seawater pumped directly from the sea, were used to culture dinoflagellates. The results indicated that the density of dinoflagellates increased significantly when reared in fish farm effluent, while diatoms tended to disappear (Buschmann et al. 2006). The importance of phosphorous in the coastal marine ecosystem is evidenced by the presence of Alexandrium spp., a toxic dinoflagellate, with an evolutionary adaptation to high nutrient conditions, including phosphate (Walsh et al., 2012).

6.17. Increased nutrient loads in poorly flushed areas with high densities of farms, as well as the modification of water column nitrogen/phosphorus ratios by intensive salmon farming could enhance the risks of HABs. Nutrient loading from salmon farming together with other environmental factors (e.g. winds, depth) interact and they need to be considered in its complexity when documenting increases of HABs in the channels and fjords of southern Chile (Buschmann et al. 2007). There are differing views regarding the contribution of nutrient releases (N and P) from cage farm operations to the occurrence of harmful algal blooms in coastal waters. A number of reports document the occurrence and abundance of harmful algal blooms in the vicinity of cage farms (Wildish et al. 1990; MacKenzie et al. 2011 ), but none of these monitoring programs were experimentally or statistically designed to answer the question of whether salmon aquaculture influence blooms of harmful algal blooms. One of the difficulties in studying the impacts of N and P discharges from salmon farms is that, often, nutrients from net pens are not the only source of discharges. Models that estimate the relative contributions of nitrogen from different sources and their loading rates have been developed and applied to field conditions (Valiela et. al. 1997). In many cases, these models have been developed with a view to providing better tools for decision-making in matters of coastal zoning for whole watersheds (Valiela et. al. 1997). Many factors such as algae species presence/abundance, degree of flushing or water exchange, weather conditions, and presence and abundance of grazers contribute to the success of a given species at a given point in time. Similar nutrient loads do not have the same impact in different environments or in the same environment at different points in time. Eutrophication is one of several mechanisms by which harmful algae appear to be increasing in extent and duration – 37 –

in many locations. Although important, it is not the only explanation for blooms or toxic outbreaks. Where nutrient enrichment has been strongly linked to stimulation of some harmful species, for others it has not been an apparent contributing factor. The overall effect of nutrient over-enrichment on harmful algal species appears species specific (Anderson et al. 2002).

6.18. In January 2003, the US Environmental Protection Agency sponsored a ‗‗roundtable discussion‘‘ to develop a consensus on the relationship between eutrophication and harmful algal blooms (HABs), specifically targeting those relationships for which management actions may be appropriate (Heisler et al. 2008). Academic, federal, and state agency representatives were in attendance. The following seven statements were unanimously adopted by attendees based on review and analysis of current as well as pertinent previous data:

Degraded water quality from increased nutrient pollution promotes the development and persistence of many HABs and is one of the reasons for their expansion in the U.S. and other nations;

The composition–not just the total quantity–of the nutrient pool impacts HABs;

High-biomass blooms must have exogenous nutrients to be sustained;

Both chronic and episodic nutrient delivery promote HAB development;

Recently developed tools and techniques are already improving the detection of some HABs, and emerging technologies are rapidly advancing toward operational status for the prediction of HABs and their toxins;

Experimental studies are critical to further the understanding about the role of nutrients in HABs expression, and will strengthen prediction and mitigation of HABs; and

Management of nutrient inputs to the watershed can lead to significant reduction in HABs.

6.19. The economic impacts from HABs are diverse and large within the United States. The annual economic impacts of HABs in the United States during the 1987-92 period ranged from $34 million to $82 million per year. Perhaps more importantly, many are recurrent, and show signs of increase as the number of toxic and harmful algal species grows and as our reliance on the coastal – 38 –

zone for aquaculture, commerce and recreation expands (Anderson et al. 2000). The development of accurate predictive models of toxic dinoflagellate blooms is of great ecological and financial importance. Episodic blooms along the west coast of Florida has caused mortalities among 100 species of marine life and wreaking havoc on the valuable commercial fishing and tourism industries. Bioaccumulation of sublethal brevitoxins in fish and mollusc tissues has led to significant dolphin and seabird mortalities and neurotoxic shellfish poisoning in humans (in: Milroy et al., 2008). In the Bay of Fundy along the southwest coast of New Brunswick Canada. Commercial fishermen were interviewed about the impact of salmon farms on their business (Wiber et al. 2009):

6.20. Fishermen all reported significant environmental degradation around aquaculture sites. Within 2 years of an operation being established, fishermen reported that gravid female lobsters as well as herring avoid the area, scallop and shells become brittle, scallop meat and sea urchin roe becomes discoloured (Wiber et al. 2009). Losses to commercial fish and shellfisheries as a consequence of toxic phytoplankton blooms, macro algal blooms in shallow estuaries and anoxia has also been reported by Smith et al. (1999).

6.21. Cage aquaculture usually use diets with relative high N content (Ackefors and Enell, 1990) and, can represent the largest source of N and P release in a given area. For the L‘Etang Inlet, New Brunswick, Canada, aquaculture operations are the largest anthropogenic source of nutrient inputs of Carbon, Nitrogen and Phosphorus cycling 3.3 and 1.6 times more nitrogen and carbon compared to natural processes in the ecosystem (Strain et al. 1995). – 39 –

6.22. Chlorophyll a has often been used as a reference for N-input. Skogen (2009) established the feasibility of a ten time increase in fish production in a Norwegian fjord with a 4% increase of

chlorophyll a. However, these fjords are at about 65°N and the study couldn‘t exclude light as a limitation to the modelling of the effect of nutrients into phytoplankton: the availability of nutrients are only limiting the production in parts of the year (light being the main limitation most of the time), this puts the

effect of fish farming to primary production to only a few percentages. Other studies reported chlorphylla measurements as a doubtful tool to establish a relationship between fish farm nutrient inputs and

phytoplankton biomass (Navarro et al. 2008). The lack of a response of chlorophyll a concentration to fish farm inputs may not necessarily indicate a lack of response by the phytoplankton, as increases in cell growth may have been balanced by mortality due, for example, to predation by protozoan or metazoan zooplankton (Navarro et al. 2008). The lack of response of phytoplankton abundance to nutrient addition from fish farms is mainly due to grazing by micro zooplankton, which plays a key role in transferring nutrients up the food web (Pitta et al., 2009).

6.23. Profound changes of rocky shore communities due to eutrofication have been reported (Worm et al. 2006). Comparative experiments and observations in the Baltic and Northwest Atlantic indicate that community structure on rocky shores is controlled and maintained both by nutrient supply and consumer pressure. While grazers and predators control their prey under normal conditions, increased nutrient supply can change the interaction from predominant consumer control to predominant resource control. This can lead to the replacement of perennial algal canopies either to mussel beds or annual algal blooms or to marked declines in community diversity. Recovery of F. vesiculosus populations and associated fauna in the Baltic Sea following the reduction of nutrient loads has been recorded and wise management actions can reverse the deleterious trends even over large spatial scales (Worm et al. 2006)

6.24. Most management decisions concerning development in the coastal zone are still made on a project by project or farm by farm basis. This approach to decision-making tends to ignore synergistic interactions among multiple human influences (Worm et al. 2000a, b). Being an essentially ecologically open system, production of high volume of wastes and their release into the environment is significant. Impacts from these wastes occur over several spatial and temporal scales: internal, local and regional (Silvert, 1992; Beveridge, 1996). Unlike its land-based counterpart, cage culture relies upon natural water movement to deliver water and oxygen to – 40 –

sustain production and remove wastes (Lucas et al. 2012). On a global basis, by 2050, coastal marine systems are expected to experience, from today's levels, a 2.4-fold increase in nitrogen and 2.7-fold increase in phosphorus loading from population expansion, with serious consequences to ecosystem structure and function (Diaz et al. 2012). Nutrient input from the Changjiang River has increased about three fold between 1960-2000 (see: 4.6.). The proportion of diatoms in the plankton communities decreased from 85% in 1984 to 60% in 2000. Also, Harmful Algal Blooms increased dramatically in number and scale with 30-80 events each year since 2000. The scale of some blooms has been in access of 10.000 km2. Whole watershed management will require whole watershed information and to date this information is not available. Where e.g. salmon farming applies for fast flow habitats to dilute their waste, other sources like dairy farming and land run-off may in fact already contribute to this same potentially assimilating ―unknown‖ habitat. The use of flowing waters as convenient wastewater disposal systems should be discouraged.

Buschman et al. (1996) estimated the area of seaweeds (Gracilaria chilensis) needed to absorb released nutrients from intensive fish cultivation in an integrated coastal tank experiment. To accomplish a 65% reduction of the dissolved nitrogen released, a seaweed cultivation covering an area 4 times larger than the fish tanks is needed, while to absorb all nutrients the corresponding area would be 16 times larger. In an open system, fewer seaweeds can be grown per square meter compared to a tank system with stirring. Hence, seaweeds occupying an even larger area would be needed. To assimilate nitrogen (particulate and dissolved) released from an intensive fish tank cultivation, the natural phytoplankton production (based on a carbon production of 175 mg*m-2*d-1 along the south coast of Chile, and a Redfield C: N: P ratio by mass of 46:7:1) would have to cover an area at least 180 times the area of the fish tanks.

Case Study from Folke et al. 1998

7. Human Population Equivalent

7.1. From an ecosystem perspective, it is important to be able to quantify the wastes in measurable, comprehensible amounts of nutrients, also for this expansion of salmon farming by NZKS. A simple and direct approach is the comparison of the salmon waste with a human equivalent. Folke et al. (1994) used the human equivalent for the first time. Black et al. (1997) queried the validity of the human equivalent approach. Folke et al. (1997) argued that fish farm waste (mainly fish excretion products and faeces) and sewage (mainly human excretion products and faeces) are very similar from an ecosystem perspective, the main difference being that fish farms spread less human pathogens. Folke et al. (1997) refer to N, which is the nutrient of major concern in marine – 41 –

environments. Any net addition of nutrients either dissolved or organically bound, resulting in an increased supply of organic matter to an ecosystem should be taken into account, and both short- term and long-term effects on both spatial and temporal scales should be considered, including nutrient transports from the fish cages to other areas Folke et al. (1997).

7.2. Despite fierce critique of comparing waste of salmon farms with a human equivalent, a number of studies have made a similar comparison (Moffit 2003, Ellis and associates, 1996). Stead et al. (2002) have commented that these calculations ignore the much localised impacts of these discharges and the comparatively huge inputs of nutrients from other land-based sources such as agriculture. This comment appears more to refer to improvements of the concept instead of a rejection. As earlier indicated, an integrated overview of all anthropogenic nutrients affecting the Marlborough Sounds is long overdue. Coastal habitats as we know them can collapse under the weight of eutrophication; as indicated earlier in this evidence.

7.3. A range of data is available for this submission to make a similar comparison between salmon farming waste and a population equivalent. Forrest et.al. (2007) estimated the total nitrogen excretion of a typical salmon farm in the Marlborough Sounds 83kg per tonne4. Most of the literature on waste models is based on Atlantic salmon (Salmo salar). However, significant physiological differences between Atlantic salmon and Chinook salmon (Oncorhynchus tshawytscha) occur. Chinook salmon is not the most efficient food converter among the cultured salmon species which is reflected in a relatively high FCR. The average profile drag acting on Chinook salmon is significantly higher than on Atlantic salmon, principally because the girth of a Chinook salmon is significantly larger compared to an Atlantic salmon of similar mass. This difference in growth explains 20% of the 33% difference between typical FCR values of 1.5 for Atlantic salmon and 2.0 for Chinook salmon (Petrell et al. 2000). Chinook salmon has a higher energy cost for swimming by as high as 20% over the same production period. The sudden darts and directional changes by swimming Chinook salmon may be another cause of the consistently poorer performance in feed efficiency of Chinook salmon as compared to Atlantic salmon (Petrell et al. 2000). Also, the apparent protein digestibility coefficients of protein sources is again more efficient for Atlantic salmon compared to Chinook salmon (Hardy et al. 2001b)

4 Feeding conversion rate (FCR) of 1.5 – 42 –

7.4. As a result, the amount of nitrogen waste from a standard Chinook farm operation will be higher compared to an Atlantic salmon operation. Smolt (weighing between 20g and 300g) will take about 10 and 20 months5 to grow to a harvestable fish of about 3.5kg to 3.8kg. NZKS uses feed similar in basic composition as other parts of the world with high protein and fat and typical proportions of the feed are protein 45.2%, fat 21.5%, carbohydrate 14.2%, ash 9.8% and moisture 9.3%, with vitamins and minerals added to provide a balanced nutritious diet (Barrat-Boyes, 2007, NZKS, Te Pangu Bay 20096). Chinook salmon contains 20g of protein for every 100g of fish.

7.5. In the latest evidence, the figures for feed have now altered to 38% protein content (Wybourne, 2012). Based on the above information, the following calculations provide an assessment of the nitrogen been released into the marine environment through a standard NZKS production unit for Chinook Salmon:

Assumptions: 1800 tonne of feed produces (FCR 1.8) 1000 tonne of salmon.

Nitrogen content in protein is 16%

One tonne of salmon feed contains 380kg protein (38.0%) x 0.16 = 60.8 kg of nitrogen.

One tonne of salmon contains 200 kg protein (20%) x 0.16 = 32 kg of nitrogen

With a FCR of 1.8, the amount of nitrogen emitted is calculated as follows:

Feed usage 1800 tonne @ 60.8kgN tonne feed-1 = 109.44 tonne nitrogen

Retained in fish 1000 tonne @ 32kgN tonne fish-1 = 32.0 tonne nitrogen

Nitrogen emitted as waste = 77.440 tonne

Nitrogen emitted tonne-1 salmon = 77.5 kg

5 LEARNING RESOURCE Unit Standard 19852V2 L e v e l 2 C r e d i t 1 0. Outline the Salmon Farming Industry in New Zealand and Worldwide 6 Application for coastal permit to discharge- Te Pangu Bay, Tory Channel – Current Coastal Permit U040813 – 2009. – 43 –

6.5. To produce 1000 tonne of salmon, 1800 tonnes of feed is required with an emission of 77.44 tonne of nitrogen into the environment. About 71% of all nitrogen in the salmon feed supplied is released as waste. This amount is comparable to figures between 52% and 95% from other studies (Hall et al. 1992, Holby et al. 1991, Wu. 1995, Islam 2005, Sanderson et al. 2008).

7.6. The annual Nitrogen and Phosphorus waste produced by a person through faeces and urine is about 4000 g and 500 g respectively (Jönsson et al. 2004, Vinneras, 2001, Vinneras et al., 2002). The average time to grow a salmon to a harvestable fish is about 17 months. Nutrients produced by one person over that same period are about 5700 g nitrogen and 700 g phosphorus. People‘s nutrient release for N and P in Norway are very similar to the ones used for this study, 13 and 2g per person per day respectively (Olsen et al. 2008)

7.7. It is now possible, to analyze the nutrient waste produced by the total NZKS operation and compare these data with the population equivalent of faecal and urine waste. The discharge per tonne of salmon is 77.5 kg nitrogen with a FCR of 1.8. For phosphorus, the production of one tonne of salmon results in a discharge of 13 kg for same FCR. With a total production increase of NZKS to 15.000 tonnes, creates the same amount of nutrient waste as 210.000 people. In other words, each tonne of salmon produced by NZKS has the human population equivalent of about 14 people. – 44 –

When the production reaches the target level of 30.000 tonnes of salmon per annum, about 2500 tonnes of nitrogen will be released from the total King Salmon operation, an equivalent to 420.000 people. A population equivalent of nutrient waste of 14 people to the nutrient waste to produce 1 tonne of salmon is comparable to other studies. It needs to be reiterated that other figures relate to Atlantic salmon. A comparison requires the higher provided PE‘s due to higher FCR of Chinook compared to Atlantic salmon.

7.8. The applicant has stated: ―The nitrogen liberated from our fish is small compared to the massive amounts that come in naturally from Cook Strait‖7 In contrast to this statement, the applicants science provider has assessed the Net Oceanic exchange (DIN) between 200 and 4200 tonnes N/y8, an amount that is comparable with the total N from the upwelling from the Cook Strait. It has been assessed that the amount of nutrients released from the total King Salmon operation can on average be comparable to the average total amount of nutrients released from the Cook Strait upwelling; respectively 2500 versus 2200 tonnes of nitrogen. From here, the Cawthron‘s assessment (report 1985) becomes very important:‖ If the nutrient wastes were sufficient to significantly reduce the existing nutrient ratio in the surrounding environment below that required by diatoms, then it is possible that other species could dominate over the major bloom periods.‖ (see 4.2 of this evidence)

7.9. Whether the assimilative capacity of the receiving waters are able to absorb the volume and duration of nutrient loading of salmon farms and avoid a fundamental shift in the food web structure with

7 E.g. Grant Rosewarne – Salmon farming, Sunday Star Times, 15 July 2012. – 45 –

ecological simplification or worse, is crucial for the assessment of this application. Salmon farms now still discharge untreated wastes directly into coastal marine waters thereby using the marine environment as an open sewer. The anthropogenic addition of nutrients through aquaculture at the base of food webs can change the abundance of animals at a range of trophic levels within food webs (Gibbs, 2012).

7.10. With a proposed upper limit of 30.000 tonnes of salmon, the nitrogen waste is the equivalent of what is produced by 420.000 people. In Marlborough, Nelson and Tasman Region 130.000 people were counted during the Census of 2006. The environmental impact of this proposed application to the Outer Marlborough Sounds is the release of nitrogen waste from more than three times this amount of people without a sewage treatment plant.

7.11. At the moment, it is estimated that there are in excess of 3000 existing septic tank systems in the Sounds. It is perceived as contradictory and inconsistent that these consented septic tanks are designed to mitigate the release of nutrients and human pathogens where an industrial applicant asks for resource consent to release to waste of up to 420.000 people in the coastal marine environment without any treatment. It may be no surprise that the sites deemed suitable for salmon farming are the fast flow current areas in the outer Marlborough Sounds coastal waters with strong currents that sweep away the wastes — a strategy akin to the practice last century of building tall smokestacks, so that industrial air emissions would be carried away by the wind. We now know that the earth is not so vast that it can absorb all of mankind‘s insults, and that pollutants must be dealt with directly, not just swept away to become another community‘s - or a another ecosystem‘s - problem. In the last 20 years there has been a 790 per cent increase in nitrogen fertilizer use in New Zealand which often ends up in rivers, lakes and the sea9. Cage farms exemplify a very primitive ―dilution is the solution to pollution‖ approach to mitigating environmental discharges (Goldberg et al. 1997).

7.12. In contrast to many countries in the world (including New Zealand) some countries pay substantial amounts of money to avoid or mitigate eutrophication of coastal waters caused by the release of excess nutrients. Fish farms release excess nutrients, but the cost to society of the impact it causes is not paid by the farmer. If the polluter-pays principle were implemented, as

8 Table 7 , page 16. Gillespie, P., Knight, B., MacKenzie, L. 2011. Environmental Effects – Water Column. Report no 1985. 9 Dr Mike Joy. Hawkes Bay Today- 3rd April 2012 – 46 –

stressed in the declaration of the United Nations Conference on Environment and Development in 1992 (Agenda 21: The United Nations Programme of Action from Rio), the cost of producing the salmon (based on society‘s willingness to pay for nutrient filtering in sewage treatment plants) would become higher than the best price the industry has ever received for farmed salmon (Folke et al. 1994).

8. Alternatives to net pens 8.1. Future aquaculture development should focus on aquaculture systems other than cage so salmon sewage can be treated in a similar way as human conglomerations. A wide variety of technologies and practices are available to make aquaculture facilities environmentally friendly, and many of these are now used on commercial fish farms. These technologies and practices must be more widely adopted if aquaculture is to be widely accepted as a clean and thus desirable industry. (Goldberg and Triplett 1997). The aquaculture industry should move away from raising finfish in cage systems. Net pens are the type of aquaculture facility least amenable to control of nutrient and biological pollutants. An industry is economically sustainable when it is profitable in the long run without input from the public in the form of monetary, environmental, or other subsidies (Eagle et al. 2003). In other words, such an industry can be successful without imposing costs on the public (Hohmeyer et al. 1996). An industry is ecologically sustainable if it maintains, or is part of a management system that maintains, the natural capital upon which it and other industries depend. In the case of potentially renewable resources such as fish stocks or coastal environments, maintenance means not impairing the ability of the resource to provide services from generation to generation (Brundtland, 1987). The pollution effects of cage aquaculture represent an external cost to society, and the challenge for environmental economists has been to estimate the magnitude of these costs and to suggest ways in which they can be mitigated or ‗internalised‘ (Whitmarsh et al. 2006). Since aquaculture and in particular finfish farming is using the environment as an input, it is likely that environmental issues will arise and in fact has already arisen in New Zealand King Salmon‘s existing operations. A precursor for sustainable management is detailed knowledge of the quality, quantity and functionality of marine ecosystems. A financial basis for information gathering appears only available when applications are presented. A financial base for monitoring the impact of the existing industry on the environment is long overdue. – 47 –

8.2. A greater understanding of complex interactions between nutrients, bacteria and cultured organisms, together with advances in hydrodynamics applied to pond and tank design, have enabled the development of closed systems. These have the advantage of isolating the aquaculture systems from natural aquatic systems, thus minimizing the risk of disease and pollution on the environment. The development of new type of integrated systems (such as: Combined extensive- intensive-systems, CIE; or Integrated Multi-Trophic Aquaculture Systems, IMTA) could also contribute to the protection of the environment and the sustainable use of resources.

8.3. The promotion and the implementation of polyculture systems with the integration of aquaculture with other coastal activities potentially resulting in a synergetic reduction of the environmental impacts (Mente et al., 2006, Costa-Pierce et al. 2011). Incorporating seaweed and/or shellfish into the salmon farming system can help to solve some of the waste problems, since these organisms filter and utilise waste products. On average, 16.8 g of nitrogen is removed from estuaries for every kilogram of shellfish meats harvested (Rice, 2001). In some cases mussel productivity is enhanced due to proximity to salmon cages and translates into a measurable financial benefit which can be recognised as a genuine economy of integration‘ (Whitmarsh et al. 2006). An integrated system of, for example, salmon and seaweed or salmon and shellfish or both could reduce nutrients significantly. Moffitt (2003) also noted that nutrient enrichment from aquaculture can be captured by other species that can be harvested through polyculture schemes, within North American aquaculture (Tuten and Avault 1981; Stone 1994). Polyculture of fin-fish, shell fish and algae in coastal systems is becoming increasingly popular in Europe, Canada and the USA Here waste is treated as a resource.

8.4. Growth of seaweeds in the vicinity of fish farm cages in North West Scotland was investigated as a means of extracting, from the surrounding water, nutrients added via fish feed and excretory products. Conservative estimates of yields show that Palmaria palmate could be expected to remove up to 12% and S. latissima 5% of the waste nitrogen released during the growth of 500 tonnes of salmon in the sea over 2 years (Sanderson et al. 2012). Other studies found great differences with uptake of nitrogen compared to ammonium (Corey, 2011) in controlled conditions. 67.5% of ammonium and 12.2% of nitrate at 300 μM nitrogen was accumulated by Palmaria palmate within 12 hours of media inoculation. Under the same conditions, Chondrus crispus accumulated 68.8% of ammonium and 29.9% of nitrate. It appears that trials with New Zealand seaweeds have not occurred. – 48 –

8.5. There is some question, however, as to whether such systems could reduce significantly the overall impact of having so much organic matter concentrated in one place. For example, C and N stable isotope composition demonstrated that these bivalves did not assimilate organic wastes from the studied fish farm directly (Navarrete-Mier, 2010). Rather, such enrichment is more likely to make a significant contribution to the production of phytoplankton several kilometres away down current and polyculture of fish and bivalves need a further close analysis. However, an indirect benefit to nutrient budget of a water body from extractive aquaculture due to digestion of phytoplankton could not be excluded. A field experiment in Tasmania was set up to test if mussel performance was enhanced and organic enrichment reduced from an integrated mussel salmon culture. No difference between the mussels within 70-100 m from the cages was established compared to mussels at 500m and 1200 m from the fish cages (Cheshuk et al. 2003). Several contributing factors were:

Solid wastes from farm did not significantly increase particulate food concentrations above ambient levels

Phytoplankton production within the farm was not enhanced

Mussels were cultured too distant to intercept settling particulate wastes emanating from fish cages

Ambient seston concentrations were consistently above the pseudo faeces threshold concentration, thereby limiting ingestion of fish farm particulate wastes.

8.6. Studies in Canada (Reid et al. 2010) support the concept of culturing blue mussels in close proximity to salmon cages in IMTA systems as a means to absorb and transform solid waste organic nutrients. Simply placing mussels arbitrarily at salmon cages will not always ensure contact of the mussels with a consistent ‗stream‘ of farm particulates for ingestion. Consequently, one of the major design challenges to IMTA sites will be, understanding particulate plume dynamics existing cages so that the placement of co-cultured suspension or filter feeders will have maximum access. Closed containment technology will address numerous serious environmental problems by imposing a physical separation of farmed salmon from the marine environment. Closed systems can be on land or floating in the ocean. – 49 –

8.7. Feasibility studies of replacing open net cages with enclosed tanks equipped with proper water filtration systems for wastes were initiated by the Government of British Columbia (DFO 2008). A review of over 40 closed-containment systems from around the world found that none was producing exclusively adult Atlantic salmon and that many previous attempts to do so had failed. Reasons for failure were numerous and were often interrelated. These reasons included but were not limited to mechanical breakdown, poor fish performance, management failure, declines in market price and inadequate financing. There are continued pressures for the salmon aquaculture industry in Canada, and government departments that regulate it, to introduce technologies and practices to further reduce the risk of potentially adverse interactions between aquaculture operations and the surrounding aquatic environment. Any efforts to address these interactions must reflect local, national and global concerns pertaining to environmental impacts. New Zealand should follow according that model.

9. Assessment of Effects on the Environment

9.1. Page 4 - Additional production of seafood is required to meet the growing demand for quality fish protein, which will not be able to be met by wild fish capture.

Take note of 5.3

9.2. Page 7 – The plan change introduces...... and provide for the expansion of salmon farming at the 8 sites.

The footprint of the King Salmon operation is not the applied area but all the area that has a benthic footprint ES >210. It is estimated that the total area with the benthic footprint ES>2 is about 400ha or 0.53% of the Queen Charlotte, Tory Channel and Pelorus Sound together.

9.3. Page 10 – Management of any adverse effects from proposed structures, nets, vessels, anchoring systems and lighting systems on the following: Marine mammals, pelagic fish and seabirds:

The applicant is not able to demonstrate how the environmental impact of their activity can and will be mitigated. In an hysteresis, after an initial trajectory of change, only a small additional change in a parameter variable (e.g. nutrient input) can result in a catastrophic shift in a state variable. This catastrophic shift cannot be reversed by a correspondingly small reversal of the parameter variable. It is unclear how these ecosystem realities will be achieved by adaptive management. The trajectory of recovery is very different from the pathway of decline. The adaptive management is designed to maintain quality of the area as

10 ES 2 means minor enrichment: Low level enrichment. Can occur naturally or from other diffuse anthropogenic sources. ‘Enhanced zone’. – 50 –

a salmon growing space, it is not designed to incorporate biodiversity decline as a feedback in the response of the farm management.

9.4. Page 32 – Analysis shows that the existing Sounds are unlikely near any critical nutrient load limit that would impact on the ecological integrity of the water column environment.

The assessment only relates to phytoplankton fluctuations, not to ecosystem responses. The applicant has established a threshold for the phytoplankton to stay within the limits of the existing trophic level. The applicant fails to identify the shifts in phytoplankton communities under stress from eutrofication, fails to identify impact of eutrofication to wider ecosystem, fails to set guidelines for impact of eutrophication on habitats within the same trophic levels, fails to provide different responses from different ecosystem to enrichment.

9.5. Page 62/63 – Effects on Water Column

Effects are described as ―localized‖ without providing a spatial context of the habitats affected. Describing the impact as “minor with regard to the wider Sounds marine environment” seems presumptuous.

Nutrients can enhance the growth of phytoplankton (including harmful algal bloom (HAB) species if present). See in particular 5:14 and 5:18.

―The nutrient loading from the proposed farms is unlikely to compromise the integrity of planktonic ecosystems in the Sounds.‖ The applicant is underestimating the amount of nutrients released from the farms. The applicant has not demonstrated whether planktonic ecosystems can be used as a proxy for wider ecosystem performance.

9.6. Page 64 – ―…the mean annual chl concentrations will remain below concentrations associated with a higher ‗eutrophic‘ state than the current ‗mesotrophic‘state in the Sounds. ―

Environmental response to eutrophication is incremental, NOT stepwise from one trophic state to another. The applicant has not made clear why chlorophylla was chosen to assess the trophic state of the sounds (mesotrophic). According to the same information the sounds can arguable be oligotrophic on the basis of Total Nitrogen. The data collected for the assessment of the impact of the proposal were not the result of a program designed to answer the question of whether salmon aquaculture influence the integrity of the Marlborough Sounds through eutrofication.

9.7. Page 68 – ―...above enrichment effects on the seabed are also reversible.‖

The applicant is not recognizing the different response by different habitats to eutrophication, including benthic deposition. Areas with diverse communities tend to have a wider range of ecological functions, including species‘ mobility and reproductive strategies, and such communities will take longer to recover from salmon farm practices than those where diversity is low and the communities are simple. Consequently, impacts will be more significant in areas with inherently high diversity like ALL applied sites compared to most of the existing salmon sites in the Marlborough Sounds. The applicant continues:‘ Organic accumulation tends to be minimal at high flow sites.....which is evident – 51 –

by relatively small increases in the sediment organic content...‖ It appears that the applicant has not used the literature available, outlining the different responses of different habitats to eutrophication. Fast flow habitats can be (and certainly are on the selected sites) more delicate and sensitive to sedimentation of salmon waste. (see 4.21.). The applicant continues to describe high flow sites (page 69) as ―more resilient‖. This assessment is not supported by the literature...

The applicant has recognized that waste products from salmon farms can be detected up to 1-2km away from the farm. That assessment is supported by the literature. 9.8. In the conditions, proposed monitoring does not prescribe trigger levels or targets that should reflect on the management of the farm. Adaptive management should be designed to change the application of salmon feed towards preset desirable environmental outcomes. Where there remain significant challenges in instituting true management experiments and identifying performance indicators they are missing from the application. The ecological and institutional complexity of marine habitat management reinforces the importance of systematic decision protocols. So far, the adaptive management being integrated in the conditions for resource consents of existing salmon farms appears only to apply to the industrial goal to maintain a healthy environment for the fish. The three basic environmental requirements as proposed by Stigebrandt (2004) for fish farming enterprise are not met in the application of New Zealand King Salmon:

• the benthic fauna at a farm site must not be allowed to disappear due to accumulation of organic material;

• the water quality in the net pens must be kept high;

• the water quality in the areas surrounding the farm must not deteriorate.

9.9. The whole application should be declined on the basis of scale, impact on the environment and shortcomings in the science provided to guarantee the sustainable use of the ecosystem of the sounds.

10. New Zealand King Shag (Phalacrocorax carunculatus) 10.1. New Zealand King Shag (King Shag) is one of the rarest species of cormorants in the world, endemic to the Marlborough Sounds. Different classification systems recognise between 7 and 14 different species of these so called ―blue-eyed shags‖ (Till, 2011). New fresh material, retrieved from Duffers Reef in 2011, found the closest emerging relative was the Bounty Island shag. King – 52 –

shag should be considered a discrete species which does not share sub-species status with Stewart Island Shag and:‖...due to its small population size and general ease of disruption during nesting it is highly vulnerable...” (Till, 2011).

10.2. The average total population of King Shags estimated to be 645 birds, with 92% at four distinctive colonies; Duffers Reef, Trio Islands, Sentinel Rock, and White Rocks. Per annum, about 102-126 breeding pairs could be identified, with an annual recruitment of 40-68 birds (Schuckard 2006b). Surveys prior to 1992 may have included only about 40% of the population, because most surveys at that time seem to have been done during the middle of the day when significant numbers of shags were absent feeding. If historic counts at colonies are adjusted for birds absent feeding, numbers appear to have been stable for at least the past 50 years — and possibly over 100 years — which would suggest a long-term balance between recruitment and mortality.

10.3. The criteria of the International Union for Conservation of Nature and Natural Resources (IUCN) for threatened species has identified King Shag with 32 other New Zealand Birds as ―VULNERABLE‖, where this ―species is facing a high risk of extinction in the wild in the medium-term future”. The status of this bird is based on the latest 2000 criteria of IUCN: Area of occupancy estimated to be less than 2000 km2. They are known to exist at no more than 10 localities. Population estimated to number less than 1000 mature individuals. – 53 –

10.4. Low numbers and a very small distribution area are of a major concern for the survival of this species. In New Zealand, the conservation status of King Shag is Nationally Endangered based on its small population of between 250-1000 individuals (Miskelly et al. 2008). Duffers Reef and Trio Islands have the highest numbers of King Shags of all colonies where Duffers Reef also has the highest recruitment of all colonies.

10.5. The foraging range of birds from Duffers Reef is 8.2 ± 4.1 km (max. 24 km), very similar to the Trio Is (9.96 ± 2.78 km, max. 18 km) (Schuckard 2006b). About 75% of the Duffers Reef birds forage in two distinctive directions, Forsyth and Beatrix Bay (southeast and south) and to Waitata Reach (southwest and west)(Schuckard 1994). A slight prevalence to forage with higher numbers in the Waitata Reach is still supported by recent surveys. – 54 –

10.6. The distribution of feeding King Shags in Waitata Reach is between 2 and 10km from the colony, where the highest number is feeding between 6-8km from Duffers Reef.

10.7. Cormorants belong to the ―flapping species‖ where a high wing loading is likely related to lower energy efficiency of ―flapping flight‖ (in Spear & Ainley 1997). A lack of sufficient muscle power to fly at speed nearer to the most energy efficient air speed per distance flown has been suggested. Wing morphology and flight behaviour of Cormorants make them belong to those birds that have little leeway to speed up or slow down because they must flap at a rate near their maximum capability (i.e. they probably fly as fast as they can under any conditions) (Pennycuick 1987b and Alerstam et al. 1993). The energy use by Cormorants to reach the feeding areas is among the highest of all seabirds and may well be an evolutionary bottleneck for the species. – 55 –

10.8. Shags do not randomly use feeding areas, but target specific locations at sea. Distribution and diving behaviour of Great Cormorant (Phalacrocorax carbo) has been studied at the Chausey Islands in France (Grémillet, et al. 1999). Birds foraged within an area of approximately 1131 km² representing only 25% of the maximal potentially available area that birds may utilize considering their maximum foraging range of 35km. Individual birds remained within restricted individual foraging areas, on average 10-18% of the total utilized area. The preferences of each cormorant not only encompass the horizontal dimensions of its feeding environment, but also the maximum depth, as individuals tend to prefer a particular depth zone.

10.9. Mean foraging range for Shags Phalacrocorax aristotilis on the Island of May, Scotland was 7.0±1.9 km, with a maximum of 17km (Wanless, et al. 1991). All feeding sites were within 7km off land, >90% either within 2km from the colony or in two discrete areas 5-13 km to north and west. Use of the areas varied between years. Restricted distribution appeared to be related to water depth and bottom sediment type. Shags feed most frequently in water 21-40 m deep, with a bottom of either gravel and sand, or rock with thin patchy sediment cover.

10.10. The optimal colony location for seabirds is usually downwind of their feeding area, which provides a tail wind for the return trip by birds carrying heavy loads of prey (Pennycuick 1989, cited in Spear & Ainley 1997). About 67% of all winds in the feeding areas come from the northwest to southeast sector. Most departures from the Trio Is were to the west and southwest and birds from Duffers Reef flew into the arc from the west through to the southeast. Birds from both colonies would therefore benefit from tail winds from the west to the southeast on the homeward trip, which constitute 43.1% of all local winds. D‘Urville Island for the Trio Island population and the Bulwer Peninsula for the Duffers Reef Population provide some shelter from northwest winds (23.9% of all local winds). Like with other shag species, wind assistance for returning shags, in particular during chick rearing may be an important parameter for quality assessment of King Shag colonies.

10.11. The provided distribution of about 800 feeding King Shag provides an overview of the best known and most comprehensive knowledge of the feeding area of this species to date. Most of the birds from Duffers Reef (74%) feed in waters between 20-40m. In Admiralty Bay the shags foraged almost exclusively in waters up to 50 m deep. Departure directions, assessed near the colonies of Duffers Reef and Trio Islands and the distribution map of feeding shags resembles a – 56 –

close correlation. Areas of Port Gore and Queen Charlotte Sound have not been surveyed with the same intensity as the Pelorus Sound. However, location of all feeding areas of King Shags in relation to the different colonies appears to be similar. Important parameters for feeding King Shags are; distance to the colony, water depth and wind assistance on return flights.

– 57 –

10.12. Witch (Arnoglossus scapha), a left-eyed flatfish, dominated the diet and accounted for about 90% of prey items and 95% of wet mass. Witch is the most common species of flatfish in New Zealand, occurring from shallow waters to depths of over 400m (Stevens, 1993). Non-breeding King Shags consumed per day 655 g in November and 695 in May (Lalas and Brown 1998). Lalas and Brown (1998) identified witch as the main prey species for King Shags in the Pelorus Sound where: ”…..King Shags are actively targeting witch over other available prey items in the area.” Whether the high density of witch in New Zealand and the Marlborough Sounds waters makes the species a likely prey or that King Shag specifically target witch remains speculative.

Witch Flounder or Megrim (Arnoglossus scapha),

10.13. In Wellington Harbour the distribution of five flatfish species including witch was studied (Livingstone 1987). Witch predominated in clear deeper water with a greater influx of oceanic water from the Cook Strait and coarser grained sediment. Other flatfish species were more common in shallow areas underlain by fine sediments; ―The non-random distribution of flatfish species in the harbor may be related to sediment types and water depth or associated with distribution of prey in different sediment types‖. Below 18m, the main tidal channel of Pelorus Sound, between West Entry Point and Maud Island, generally grades into coarse, silty sand (C. Duffy, 1994). This habitat resembles similarities with the witch habitat of Wellington Harbor, as described by Livingstone (1987). Witch is very distinctive from all the other species of flatfish, feeding on pelagic and epibentic active prey. These preys are most common in deeper water with coarser grained sediments. – 58 –

10.14. All the King Shag colonies were visited to collect pellets with otolith samples during the summer of 2011 (Schuckard et al. in prep.). The results provide a first ever insight in prey selection from the birds from the main colonies in the non breeding period.

10.15. Results of this study are in a preliminary stage. The preys have been analyzed but the relative weight contribution per species to the overall diet is still pending. Witch is still a very important prey species in the overall King Shag diet. However, in numerical sense, the diet from Duffers Reef and Sentinel Rock showed more diversity in prey with an even occurrence of Witch, Lemon sole (Pelotretis flavilatus) and Opalfish (Hemerocoetes monopterygius). All the preferred prey items are predominantly benthic and epibenthic species, highlighting the deep diving capabilities and the reliance and dependency on a healthy benthic and epibenthic environment of the deeper water of the Marlborough Sounds. In earlier studies, lemon sole, flounder, sole and opalfish have also been mentioned as prey for King Shag (Nelson 1971, Lalas & Brown 1998, Lalas 2001, 4.4) However, their contribution to the overall diet of King Shag was relatively small compared to the recent study. Other incidental preys are; blue cod, sea perch, leatherjacket, scorpionfish, and red rock lobster. All these species are linked to communities of rocky sea-bottoms and/or bottoms with coarser substrates. Opalfish has also been linked to a demersal and benthic habitat as a prey species for Snapper (Coleman, 1972), New Zealand Sea Lion (Phocarctos hookeri) (Meynier, et al. 2009) and selectively provisioned to chicks of Yellow-eyed Penguin (Megadyptes antipodes) as a smaller prey size (Browne et al. 2011).

10.16. Water temperature and dive depth influence the cost of diving but foraging parameters of shags are most strongly influenced by the availability of prey. Even a small reduction in prey density will prevent birds meeting their energy requirements (Grémillet et al. 1999). A reduction of prey density of only 25% results in search time increase of 50%-100%. If prey density decreases to 50%, females will fail to reach the foraging efficiency of 1.0, irrespective of temperature or diving depth. – 59 –

Models of the effects of environmental conditions and energy requirements on the feeding performance and distribution of shags (Phalacrocorax aristotelis) predicted that bird numbers would decline where predicted daily feeding times were high (Wanless et al. 1997).

10.17. Factors affecting the distribution of other members from the ‗blue-eyed‘ shags also provide a comparison for King Shags The mean diving depth of the Antarctic shag (Leucocarbo bransfieldensis) was 37.8 m (Casaux et al. 2001), of the Macquarie shag (L. purpurascens) 27.1-39.3 m (Kato et al. 1996), and studies of the South Georgian shag (L. georgianus) found the birds foraging in depths of 61.4-83.9 m (Kato et al. 1992) and 56 and 91 m (Wanless & Harris 1993). Imperial Shag (L. atriceps) was diving average depths of 80-90m (max 116m) twice more common compared to shallow dives (Croxall et al. 1991). The mean diving depths of shags are thought to reflect the availability of food or the physical characteristics of the feeding area for both, rather than the diving ability of the bird (Casaux et al. 2001; Wilson & Wilson 1988). That the diving ability may not be the main control is shown by the maximum diving depths being sometimes much greater than mean diving depths, for example 125 m for South Georgian shag (Wanless & Harris 1993), 116 m for Imperial shag (L. atriceps) (Croxall et al. 1991), and 112.6 m for the Antarctic shag (Casaux et al. 2001). A close relative of King Shag, the Kerguelen Shag (L. verrucosus) provided a combined set of data of diving depth, GPS, air speed and under water speed (Watanabe et al. 2011). The average distance from the colony was 8.1 km with a maximum of 26km with an average diving depth of 23.4 m, an average maximum of 45.6 m and absolute maximum of 94.2m. These birds regularly rested at sea during both outbound and inbound flights without any diving, which were interpreted by the authors as necessary recuperation for the high flight costs. The implication of deep diving at the cost of flight performance was an important outcome of this study.

10.18. King Shag is like most blue eyed shag a deep diver and subsequently the energy to fly to the feeding area is in budgetary terms very significant. King Shags are feeding 2-10km in mostly south western direction from the colony in waters between 20-50m deep. Duffer‘s reef is one of the two bigger colonies of New Zealand Kings shags with the highest recruitment. The applications for the new salmon farms are all in the middle of the main feeding areas of Duffers Reef and White Rock – 60 –

colonies, potentially affecting 53% of the total population.

10.19. The proposed and also existing farms in Waitata Reach are located between 2 and 10km from Duffers Reef. The highest density of farms, but also the highest application of salmon feed is also located between 6 and 8 km from the colony. This zone of high salmon impact overlaps with the area where most King Shags from Duffers Reef are feeding (see 9.6.). Between 16300 and 30600 tonnes of feed are potentially released if the NZKS applications and others under appeal do go ahead. An assumed average feed use of 23.500 tonnes per annum will result in a monthly nitrogen releases of between 40 and 130 tonnes, where the spike of nitrogen release through feces will take place during the warmest temperatures of the year (between December and April).

– 61 –

10.20. The population equivalent in nitrogen terms of using 23.500 tonnes of feed per annum in Waitata Reach is about 150.000 people. The waste will be released between 2 and 10km from Duffers Reef. I would regard the impact of eutrofication in this main feeding area as an unacceptable experiment, threatening the survival of a very significant portion of the total King Shag population. Impacts of toxic algae on seabirds reveal an array of responses ranging from reduced feeding activity, inability to lay eggs, and loss of motor coordination and death (Shumway et al. 2003). Bird deaths caused by HABs have been widely reported (in Lewitus et al. 2012). Some of the dinoflagellate produced foam destroys the waterproof layer of feathers that keeps seabirds dry, restricting flight and leading to hypothermia. – 62 –

10.21. The Queen Charlotte Sound is also an important feeding area for King Shags where the Tory Channel meets the Queen Charlotte. Where the numbers recorded are less compared to the Pelorus Sound (less surveys and smaller colony), the overlap of feeding areas and the waste from nearby salmon farm proposals may also affect King Shags from White Rock colony. Kaitapeha and Ruaomoko combined have applied for 5000 to 10000 tonnes of feed per annum on top existing farms.

10.22. In the underwater environment, illumination is primarily influenced by water depth. Depth utilization of diving shags is deeper around midday compared to earlier or later in the day. Most seabirds are visual hunters and are thus strongly affected by light levels. Light limited foraging patterns have been recorded by blue-eyed shags (Phalacrocorax atriceps) to be between -0.3 and 2 log10 lx. (0.5 and 100 lux) (Wanless et al. 1999). NIWA made an assessment of sustainable production levels related to a proposed marine mussel farm development in Port Ligar for Kuku Mara Partnership in 2000 (Project No.: KMP00501). The reports comments that relative to many other coastal waters around New Zealand, Pelorus can exhibit very low nitrogen concentrations, predominantly in surface waters of embayments in summer. Small increases in phytoplankton in an oligotrophic system (like the sounds based on total nitrogen) can greatly compress the range available for deep diving blue eyed shags. For example, Chlorophyll a levels measured in the Sounds average about 1 mg/m3. If that was to increase to 2 mg/m3 the available depth range would shrink from 52 to 37m.

– 63 –

10.23. The effect of water clarity on the distribution of marine birds in near-shore waters of Monterey Bay in California was studied (Henkel, 2006). Brandt‘s Cormorant, a generalist predator but typically feeding on epibenthic fishes, occurred most often in the clearest water available (>5m Sechi depth). However, some plasticity towards water clarity was observed by this species.

10.24. The aquatic visual acuity of great cormorants was measured under a range of viewing conditions and found to be unexpectedly poor (White et al. 2007, Martin, 2011) and comparable to unaided humans under water. The efficient hunting of cormorants involves the use of specialized foraging techniques which employ brief short-distance pursuit and/or rapid neck extension to capture prey that is visually detected or flushed only at short range (less than 1m). Prey detestability is constrained by their poor aquatic visual acuity.

11. NIWA Client Report No: CHC2011-058 – Assessment of potential environmental effects of the proposed NZ King Salmon expansion on seabirds, with particular reference to the NZ King Shag – Dr. Paul Sagar

11.1. Executive summary: page 5 and page 6 – Dr. Sagar assessed that the information status of NZ King Shag is restricted to: ...‖anecdotal observations or short term studies.‖ He continues:‖...gaps in knowledge of the biology of the species may reduce the confidence with which some environmental effects may be predicted‖. However, no additional information is provided to alleviate these concerns through a measured precautionary approach:

Dr. Sagar does not address the limited adaptability for this deep diving species with poor flight capabilities.

The impact of the nutrient composition comparable with 150.000 people in the important feeding zone from the Duffers Reef population is not addressed.

How the impact and magnitude of the nutrient waste reflects on phytoplankton shading and light conditions near benthic habitats and King Shag feeding area is not addressed.

The multiple risks of HABs for birdlife but also for marine mammals is not addressed

The specific important role of Duffers Reef as one of the main colonies and the colony with the highest recruitment is not addressed. – 64 –

In general, Dr. Sagar has not made any attempt to provide the board with an assessment of the limited spatial occupancy of this species in the Marlborough Sounds.

I regard the assessment of Dr. Sagar as deficient in providing the board with an assessment of the status of this endemic seabird of the Marlborough Sounds.

11.2. Table 3-1 – A number of species in this table are missing from the list of breeders in the Marlborough Sounds:

Sooty Shearwater – breeds in the Marlborough Sounds

Flesh-footed Shearwater – breeds in the Marlborough Sounds

White fronted Tern – breeds in the Marlborough Sounds. About 400-500 pair are breeding, about 3% of the national population (Schuckard, 2005)

Diving Petrel – breeds in the Marlborough Sounds.

Little Shag – breeds in the Marlborough Sounds

11.3. Dr. Sagar refers to a number of publications to assess the impact of aquaculture structures on seabirds (table 4-1). Where the impact of a mussel farm and a finfish farm is profoundly different, it is not clear if table 4-1 is related to mussel farms, finfish farms or both. Therefore, the relevance of the statements (4.2.2): ―Kings Shags use mussel floats for roosting and have been observed feeding within mussel farms, although there are no reports of them using salmon farm structures‖ is unclear. Also the relevance to salmon farming of the following sentence is also unclear: ―However, as with other seabirds, use of such new roosting sites may reduce the energy expenditure of the birds because they do not have to fly to and from their natural roosting sites which may be some distance from their foraging area‖.

11.4. Dr. Sagar states that the ―extent of the proposed salmon farms is insignificant compared to the overall area of such depths within the foraging range of NZ King Shags”. Consequently, the potential effects of habitat exclusion are considered to be ―insignificant.‖ No figures or maps are provided of the suitable habitat and how much of this habitat may or may not become affected by the application. The assessment seems ad hoc. – 65 –

11.5. Dr. Sagar states in 4.2.5: ―currently, it is assumed that they are diurnal feeders...” In the last 20 years, there is no evidence that King Shag feeds at night, there is however sufficient evidence that daylight is an important factor to depart the colony. In winter time, departures are very strongly correlated to sunrise.

11.6. The draft conditions only provide a somewhat meager reference to maintain the well-being of New Zealand King Shag:

11.7. I firmly disagree with Dr. Sagar‘s selection parameters that he selected for his assessment. Instead, deep diving, poor flying, depending on clear water, low adaptibility for changing habits should have been the guiding principles for the assessment. I perceive this application as a significant risk for the species.

12. References

– Ackefors, H., Enell, M., 1990. Discharge of nutrients from Swedish fish farming to adjacent sea areas. Ambio 19 (1), 28–35.

– Airoldi, L. and Beck, M.W. 2007. Loss, status and trends for coastal marine habitats in Europe. Oceanography and Marine Biology: An Annual Review, 2007, 45, 345-405.

– Al-Ghelani, H.M., AlKindi, A.Y.A., Amer, S., and Al-Akhzami, Y.K. 2005. Harmful Algal Blooms: Physiology, Behavior, Population Dynamics and Global Impacts – A Review. SQU Journal For Science, 10: 1-30.

– Alerstam, T. & Gudmundsson, G. A. & Larsson, B. 1993. Flight tracks and speeds of Antarctic and Atlantic seabirds: Radar and optical instruments. Phil. Trans. R. Soc. Lond. B.:55-67.

– Anderson, D.M., Glibert, P.M. and Burkholder, J.M. 2002. Harmful Algal Blooms and Eutrophication: Nutrient Sources, Composition, and Consequences. Estuaries, Vol 25, No 4b, p 704-726. – 66 –

– Anderson, D.M., Hoagland, P.H., Kaoru, Y. and White, A.W. 2000. Estimated Annual Economic Impact from Harmful Algal Blooms (HABs) in the United States. Woods Hole Oceanographic Institution., Techinical Report WHOI- 2000-11.

– Anonymous, 2000. Relationship between water quality, aquaculture production and human food safety.

– Asche, F. and Bjørndall, T. 2011. The economics of Salmon Aquaculture. Wiley-Blackwell Publication.

– Ashe, F., Roll, K.H., Tveteras, R. 2009. Economic inefficiency and environmental impact: An application to aquaculture production. Journal of Environmental Economics and Management 58: 93-105.

– Barbera, C., Bordehore, C., Borg, J.A., Glémarec, M., Grall, J., Hall-Spencer, J.M., de la Huz, Ch., Lanfranco, E., Lastra, M., Moore, P.G., Mora, J., Pita, M.E., Ramos-Esplá, A.A., Rizzo, M., Sánchez-Mata, A., Seva, A., Schembri, P.J. & Valle, C. 2003. Conservation and management of northeast Atlantic and Mediterranean maerl beds. Aquatic Conservation: Marine and Freshwater Ecosystems 13 (Suppl.), 65–76.

– Barratt-Boyes, M. 2007. New Zealand Salmon Reigns Supreme. New Zealand Aquaculture, issue 19: 5-7.

– Bell, P. R. F., and I. Elmetri. 1995. Ecological indicators of large-scale eutrophication in the Great Barrier Reef lagoon. Ambio 24: 208-215.

– Beveridge, M.C.M. 1996. Cage Aquaculture. 2nd edn 346 pp. Fishing News Books: Oxford.

– Beveridge, M.C.M., Phillips, M.J. and Clarke, R.M. 1991. A quantitative and qualitative assessment of wastes from aquatic animal production. In: Advances in World Aquaculture Volume 3, D.E. Brune and J.R. Tomasso (eds), pp. 506–533. World Aquaculture Society, Baton-Rouge, USA.

– Beveridge, M.C.M., Phillips, M.J. and Macintosh, D.C. 1997. Aquaculture and the environment: the supply and demand for environmental goods and services by Asian aquaculture and the implications for sustainability. Aquaculture Research 28: 101–111.

– Black, E., Gowen, R., Rosenthal, H., Roth, E., Stechy, D., and Taylor, F.J.R.. 1997. The Costs of Eutrophicationh from Salmon Farming: Implications for Policy – A comment. Journal of Environmental Management (1997) 50, 105–109

– Blyth, P.J., Purser, G.J. and Russell, J.F. 1993. Detection of feeding rhythms in sea caged Atlantic salmon using new technology. In: Fish Farming Technology, H. Reinerstsen, L.A. Dahle, L. Joergensen and K. Tvinnereim (eds), pp. 209–216. A.A. Balkema, Rotterdam. – 67 –

– Boyd, C.E. and B.J. Watten. 1989. Aeration systems in aquaculture. Reviews of Aquatic Science 1:425-472.

– Boyd, C.E. 2001. Water quality standards: total suspended solids. Global Aquaculture Advocate. 4(1), 70-71.

– Browne, T. Lalas, C., Mattern, T., Heezik, Y.van. 2011. Chick starvation in yellow-eyed penguins: Evidence for poor diet quality and selective provisioning of chicks from conventional diet analyses and stable isotopes. Austral Ecology Vol 36, issue 1:99-108.

– Brundtland H. Our common future (for the World Commission on Environment and Development. Oxford: Oxford University Press; 1987. p. 45–65.

– Buschmann, A., Costa-Pierce, B. A., Cross, S., Iriarte, J.L., Olsen, Y. and Reid, G.. 2007. Nutrient Impacts of Farmed Atlantic Salmon (Salmo salar) on Pelagic Ecosystems and Implications for Carrying Capacity. Report of the technical working group (TWG) on nutrients and carrying capacity of the salmon aquaculture dialogue. 30 August 2007.

– Buschmann, A.H., Cabello, F., Kyle, Y., Carvajal, J., Varela, D.A. and Henriquez, L. 2009. Salmon aquaculture and coastal ecosystem health in Chile: Analysis of regulations, environmental impacts and bioremediation systems. Ocean & Coastal Management (2009), doi:10.1016/j.ocecoaman.2009.03.002

– Buschmann, A.H., Riquelme, V.A., Hernandez-Gonzalez, M.C., Varela, D., Jimenez, J.E., Henriquez, L.A., Vergara, P.A., Guinez, R. and Fiun, L. 2006 A review of the impacts of salmonid farming on marine coastal ecosystems in Southeast Pacific. ICES Journal of Marine Science, 63: 1338-1345.

– Casaux, R.; Favero, M.; Silva, P.; Baroni, A. 2001. Sex differences in diving depths and diet of Antarctic shags at South Shetland Islands. Journal of field ornithology 72: 22-29.

– Chan, F.H., Anderson, C., Boustead, N.C., 1990. First record of a Heterosigma (Raphidophyceae) bloom with associated mortality of cage reared salmon in Big Glory Bay New Zealand. New Zealand Journal of Marine and Freshwater Research, 1990, Vol.24: 461-469.

– Chan, F.H. and Page, M. 1995. Influence of light and nitrogen sources on growth of Hetrosigma carterae (Raphidophyceae). New Zealand Journal of Marine and Freshwater Research, Vol.29:299-304.

– Chen, N. and Hong, H. 2012. Integrated managementof nutrients from the watershed to coast in the sybtropical region. Environmental Sustainability. http://dx.doi.org/10.1016/j.cosust.2012.03.007 – 68 –

– Chen, N., Peng, B., Hong, H., Turyaheebwa, Cui, S., Mo, X. 2012. Nutrient enrichment and N:P ratio decline in a coastal bay-river system in southeast China: The need for a dual nutrient (N and P) management strategy. Ocean & Coastal Management. http://dx.doi.org/10.1016/j.ocecoaman.2012.07.013

– Chen, Y-S, Beveridge, M.C. and telver, T.C. 1999. Physical characteristics of commercial pelleted Atlantic salmon feeds and consideration of implications for modelling of waste dispersion through sedimentation. Aquaculture International 7: 89-100, 1999.

– Cheshuk, B.W., Purser, G.J., Quintana, R. 2002. Integrated open-water mussel (Mytilus planulatus ) and Atlantic salmon (Salmo salar) culture in Tasmania, Australia. Aquaculture 218: 357-378.

– Cho, C.Y. and Bureau, D.P. 1997. Reduction of waste output from salmonid aquaculture through feeds and feeding. Progressive Fish-Culturist 59: 155–160.

– Cole, R. 2002. Impact of marine farming on wild fish populations. NIWA Final Research Report for Ministry of Fisheries Research Project ENV2000/08.

– Colman, J.A. 1972. Food of Snapper (Chrysophyrys auratus), in the Hauraki Gulf, New Zealand. New Zealand Journal of Marine and Freshwater Research, Vol. 6, Number 3: 221-239

– Corey, P. 2011. Nutrient Removal by Palmaria palmata and Chondrus crispus in Bioremediation of Aquaculture Effluent. Submitted in partial fulfillment of the requirements for the degree of Master of Science at Dalhousie University Halifax, Nova Scotia in co-operation with Nova Scotia Agricultural College Truro, Nova Scotia

– Costa-Pierce, B. A. 1996. Environmental Impacts of nutrients from aquaculture: Towards the evolution of sustainable aquaculture. In: Aquaculture and Water Resource Management, ed. 188

– Costa-Pierce, B.A., Bartley, D.M., Hasan, M., Yusoff, F., Kaushik,S.J., Rana, K., Lemos, D., Bueno, P. and Yakupitiyage, A. 2011, in press. Responsible use of resources for sustainable aquaculture. Global Conference on Aquaculture 2010, Sept. 22-25, 2010, Phuket, Thailand. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy.

– Cripps, S. J. and L. A. Kelly. 1996. ―Reductions in wastes from aquaculture,‖ In D. J. Baird et al. (eds.), Aquaculture and water resource management, Blackwell Science, Oxford.

– Croxall, J.P.; Naito, Y.; Kato, A.; Rothery, P.; Briggs, D.R. 1991. Diving patterns and performance in the Antarctic blue-eyed shag Phalacrocorax atriceps. Journal of zoology, London 225: 177-199. – 69 –

– Davidson, R.J., Courtney, S.P., Millar, I.R., Brown, D.A., Deans, N.A., Clerke, P.R., Dix, J.C., Lawless, P.F., Mavor, S.J., McRae, S.M. 1995. Ecologically Important Marine, Freshwater, Island and Mainland Areas From Cape Soucis To Ure River, Marlborough, New Zealand: Recommendations For Protection. Nelson/Marlborough Conservancy Occasional Publication Number 16.

– Davidson, R., Duffy, C., Gaze, P., Baxter, A., DuFresne, S., Courtney, S., Hamill, P. 2011. Ecologically Significant Marine Sites in Marlborough, New Zealand. Marlborough District Council and Department of Conservation.

– DFO. 2008. Potential Technologies for Closed-containment Saltwater Salmon Aquaculture. DFO Can. Sci. Advis. Sec. Sci. Advis. Rep. 2008/001.

– Diaz, R., Rabalais, N.N. and Breitburg, D.L. 2012. Agriculture‘s Impact on Aquaculture: Hypoxia and Eutrofication in Marine Waters. OECD Publishing 2012.

– Diaz, R.J. and Rosenberg, R. 2008. Spreading Dead Zones and Consequences for Marine Ecosystems. Science 321, 926-929.

– Duffy, L. 1994. Letter and map to French Pass Residence marine Reserve Committee.

– Eagle, J., Naylor, R. and Smith, W. 2003. Why farm salmon outcompete fishery salmon. Marine Policy doi:10.1016/j.marpol.2003.08.001

– EEA. 2002. Europe‘s Biodiversity — Biogeographical Regions and Seas. Online. Available HTTP: http://reports.eea.eu.int/report_2002_0524_154909/en (accessed 4 August 2006).

– EEA. 2006a. The Changing Faces of Europe‘s Coastal Areas. EEA Report 6/2006. Luxembourg: OPOCE. Online. Available HTTP: http://reports.eea.europa.eu/eea_report_2006_6/en/eea_report_6_2006.pdf (accessed 7 August 2006).

– Ellis, D. and Associates. 1996. Net loss: The salmon netcage industry in British Columbia: Vancouver, B.C.: The David Suzuki Foundation.

– FAO (2007) FAO Fisheries Department, Fishery Information, Data and Statistics Unit. Fishstat Plus: Universal software for fishery statistical time series (FAO Fishstat Plus database version 2.30). Available at: http://www.fao.org [Accessed 30 March 2007].

– Fernandes M, Tanner J (2008) Modelling of nitrogen loads from the farming of yellowtail kingfish Seriola lalandi (Valenciennes, 1833). Aquacult Res 39:1328–1338 – 70 –

– Findlay, R.H. and Watling, L. 1994. Towards a process level model to predict the effects of salmon netpen aquaculture on the benthos. In: Modelling Benthic Impacts of Organic Enrichment from Marine Aquaculture (Hargrave, B.T., ed.)Canadian Technical Report of Fisheries and Aquatic Sciences, Canada, pp. 47-48.

– Folke C. Kautsky N. Troell M 1994. The costs of eutrophication from salmon farming: Implications for policy. Journal of Environmental Management 40: 173-182.

– Folke, C. , Kautsky, N., Berg, H., Jansson, A., Troell, M.. 1998. The ecological footprint concept for sustainable seafood production: A review. Ecological Applications, 8(1) Supplement, 1998, pp. S63–S71

– Folke, C., Kautsky, N. and Troell, M. 1997. Salmon Farming in Context: Response to Black et al.. Journal of Environmental Management (1997) 50, 95–103

– Forrest, B.; Keeley, N.; Gillespie, P.; Hopkins, G.; Knight, B.; Govier, D. (2007). Review of the ecological effects of marine finfish aquaculture: Final report. Prepared for the Ministry of Fisheries. Cawthron Report No. 1285.

– Forster, J. 1995. Cost trends in farmed salmon. The Alaska Department of Commerse and Economic Development, Division of Economic Development, Juneau, Alaska. 40pp.

– Gibbs, M. 2006. Aquaculture in New Zealand with Respect to Sustainable Development Principles. Prepared for Parliamentary Commisioner for the Environment. Cawthron Report No.1211.

– Gibbs, M.T. 2012. The changing focus of the ecological risk assessment of aquaculture operations: From local to cumulative risk assessment. Human and Ecological Risk Assessment: An International Journal, 18: 488-500.

– Glibert, P.M. 2012. Ecological stoichiometry and its implications for aquatic ecosystem sustainability. Environmental Sustainability 2012, 4:1–6

– Goldburg, R. and Triplett, T. 1997. Murky Waters: Environmental Effects of Aquaculture in the US. Environmnetal Defence Fund.

– Gowen, R.J. and Bradbury NB, 1987. The ecological impact of salmonid farming in coastal waters: A review. Oceanogr. Mar. Biol. Ann. Rev. 25:563-575.

– Grémillet, D.; Wilson, R.P. 1999. A life in the fast lane: energetics and foraging strategies of the great cormorant. Behavioral ecology 10: 516-524. – 71 –

– Hall, P.O.J., Holby, O., Kolberg, S., Samuelsson, M.O.. 1992. Chemical fluxes and mass balances in a marine cage farm. IV Nitrogen. Marine Ecology Progress Series, 61: 61-73.

– Hall, POJ, Holby, O, Kollberg S., Samuelsson, M-O. 1992. Chemical fluxes and mass balances in a marine fish cage farm. I. Carbon. Mar. Ecol. Prog. Ser. 89:81-91.

– Hallegraeff G., Beardall, J., Brett, S., Doblin, M., Hosja, W., de Salas, M. and Thompson, P. (2009) Phytoplankton In A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 (Eds. E.S. Poloczanska, A.J. Hobday and A.J. Richardson), NCCARF Publication 05/09, ISBN 978-1-921609-03-9.

– Hall-Spencer J., White, N., Gillespie, E., Gillham, K., Foggo, A.. (2006). Impact of fish farms on maerl beds in strongly tidal areas. Mar Ecol Prog Ser: Vol.326: 1-9, 2006.

– Hardy, R. W. 2001a. Fish feeds and nutrition. Aquaculture Magazine 27 (2): 51-54.

– Hardy, R. W. 2001b. Urban legends and fish nutrition part 2. Aquaculture Magazine 27(2):57-60.

– Hargrave, B. T., G. A. Phillips, L. I. Doucette, M. J. White, T. G. Milligan, D. J. Wildish, and R.E. Cranston. 1997. Assessing benthic impacts of organic enrichment from marine aquaculture. Water, Air, and Soil Pollution 99:641-650.

– Harvey, A., Woelkerling, W., Farr, T., Neill, K. and Nelson, W.. 2005. Coralline algae of central New Zealand. An identification guide to common ‗crustose‘ species. NIWA Information Series No. 57.

– Hauss, H., Franz, J.M.S. and Sommer, U. 2012. Changes in N:P stoichiometry influence taxonomic composition and nutritional quality of phytoplankton in the Peruvian upwelling. Journal of Sea Research. http://dx.doi.org/10.1016/j.seares.2012.06.010

– Heinig, C.M. 2001. The impacts of Salmon Aquaculture: The Difficulties of Establishing Acceptability Limits and Standards in: Tlusty, M.F., D.A. Bengston, H.O. Halvorson, S.D. Oktay, J.B. Pearce, and R.B. Rheault, Jr. (eds.) 2001. Marine Aquaculture and the Environment: A Meeting for Stakeholders in the Northeast. Cape Cod Press, Falmouth, Massachusetts.

– Heisler, J., Glibert, P.M., Burkholder, J.M., Anderson, D.M., Cochlan, W., Dennsion, W.C., Dortch, Q., Gobler, C.J., Heil, C.A., Humphries, E., Lewitus, A., Magnien, R., Marshall, H.G., Sellner, K., Stockwell, D.A., Stoecker, D.K., and Suddleson, M. 2008. Eutrophication and harmful algal blooms: A scientific consensus. Harmful Algae 8: 3-13. – 72 –

– Henkel, L.A. 2006. Effect of water clarity on distribution of marine birds in nearshore waters of Monterey Bay, California. J. Field Ornithol. 77(2): 151-156.

– Higgs, D.A., Dosanjh, B.S., Prendergats, A.F., Beames, R.M., Hardy, R.W., Riley, W. and Deacon, G. In: Nutrintion and utilization technology in aquaculture, C. Lim and D. Sessa (Eds.), AOCS Press, Champaign, IL. Pp. 130 – 156. 1995.

– Hinrichsen, H.H., Huwer, B., Makarchouk, A., Petereit, C., Schaber, M. And Voss, R. 2011. Climate-driven long term trends in Baltic Sea oxygen concentrations and the potential consequences for eastern Baltic cod (Gadus morhua). ICES Journal of Marine Science, 68: 2019-2028.

– Hohmeyer O, Ottinger, R.L. and Rennings, K.(eds), 1996. Social costs and Sustainability: Valuation and Implementation in the Energy and Transport Sector. . Berlin: Springer; 1996.

– Holby, O, Hal POJ 1991. Chemical fluxes and mass balances in a marine fish cage farm. II. Phosphorus. Mar. Ecol. Prog. Ser. 70: 263-272.

– Hopkins GA 2004. Seabed impacts of Marlborough Sounds salmon farms: Te Pangu Bay monitoring 2003. Cawthron Report No. 847b. 25p plus appendices.

– Howarth RW and Marino R. 2006. Nitrogen as the limiting nutrient for eutrophication in coastal marine ecosystems: Evolving views over three decades. Limnol. Oceanogr. 5:364–376.

– Howarth, R. W. 1998. An assessment of human influences on fluxes of nitrogen from the terrestrial landscape to the estuaries and continental shelves of the North Atlantic Ocean. Nutrient Cycling in Agroecosystems 52:213-223.

– Howarth, R.W., Chan F., Conley, D.J., Garnier J., Doney S.C., Marino R., and Billen, G. 2011. Coupled biogeochemical cycles: eutrophication and hypoxia in temperate estuaries and coastal marine ecosystems. Front Ecol Environ 2011; 9(1): 18–26, doi:10.1890/100008

– Ingrid, G., T. Andersen, and O. Vadstein. 1997. Pelagic food webs and eutrophication of 191 coastal waters: impact of grazers on algal communities. Mar. Pollut. Bull. 33(1-6): 22-35.

– Islam, M.S. 2005. Nitrogen and phosphorus budget in coastal cage aquaculture and impacts of effluent loading on ecosystem: review and analysis towards model development. Marine Poillution Bulletin 50: 48- 61. – 73 –

– Jansen, P.A., Kristoffersen, A.B., Viljugrein, H., Jimenez, D., Aldrin, M, and Stein, A. 2012. Sea lice as a density-dependent constraint to salmonid farming. Proc. R. Soc. B (2012) 279, 2330–2338

– Jessopp, M., MvAllen, R., O‘Halloran, J., and Kelly T. 2007. Nutrient and Ecosystem Dynamics in Ireland‘s only Marine Nature Reserve. Prepared for the Environmental Protection Agency by School of Biological, Earth and Environmental Sciences, University College Cork

– Johnson PTJ, Townsend AR, Cleveland CC, Glibert PM, Howarth RW, McKenzie VJ, Rejmankova E, Ward MH: Linking environmental nutrient enrichment and disease emergence in humans and wildlife. Ecol Appl 2010, 20:16-29.

– Jönsson, H. and Vinnerås, B. 2004. Adapting the nutrient content of urine and faeces in different countries using FAO and Swedish data. In: Ecosan – Closing the loop. Proc. 2nd Intern. Symp. Ecological Sanitation, April 2003, Lübeck, Germany. p 623-626. (www2.gtz.de/ecosan/ download/ecosan-Symposium-Luebeck- session-f.pdf)

– Jones, J.B. and Rhodes, L.L. 1994. Suffocation of pilchards (Sardinops sagax) by a green microalgal bloom in Wellington Harbour. New Zealand Journal of Marine and Freshwater Research Vol.28: 379-383.

– Juell, J.E. 1991. Hydroacoustic detection of food waste – a method to estimate maximum food intake of fish populations in sea cages. Aquaculture Engineering 10: 207–217.

– Karakassis, I., Hatziyanni, E., Tsapakis, M., and Plaiti, W. 1999. Benthic recovery following cessation of fish farming: a series of successes and catastrophes. Mar. Ecol. Prog. Ser. 184: 205-218.

– Kato, A.; Croxall, J.P.; Watanuki, Y.; Naito, Y. 1992. Diving patterns and performance in male and female blue-eyed cormorants Phalacrocorax atriceps at South Georgia. Marine ornithology 19: 117-129.

– Kato, A.; Naito, Y.; Watanuki, Y.; Shaughnessy, T.D. 1996. Diving pattern and stomach temperatures of foraging king cormorants at subantarctic Macquarie Island. Condor 98: 844-848.

– Khanna, D.R., Bhutiani, R., Chandra, K.S.. 2009. Effect of the Euphotic Depth and mixing depth on phtolanktonic growth mechanism. Int. J. Environ. Res., 3(2):223-228,

– Kratochwill, A. 1999. Biodiversity in ecosystems: principles and case studies of different complexity levels. Kluwer Acedemic Publishers.

– Kutti, T., Ervik, A. and Hansen, P.K.. 2007a. Effects of organic effluents from a salmon farm on a fjord system.I. Vertical export and dispersal proceses. Aquaculture 262: 367-381. – 74 –

– Kutti, T., Hansen, P.K., Ervik, A., Høisæter, T. and Johannessen, P. 2007b. Effects ot organic effluents from a salmon farm on a fjord system. II. Temporal and sp[atial patterns in infauna community composition. Aquaculture 262 (2007): 355-366.

– Lalas C.; Brown, D. 1998. The diet of New Zealand king shags (Leucocarbo carunculatus) in Pelorus Sound. Notornis 45: 129-139.

– Lee H.W., Bailey-Brock JH, McGurr MM (2006) Temporal changes in the polychaete infaunal community surrounding a Hawaiian mariculture operation. Mar Ecol Prog Ser 307:175–185

– Levin, L.A.; Boesch, D.F.; Covich, A.; Dahm, C.; Ereus, C.; Ewel, K.C.; Kneib, R.T.; Moldenke, A.; Palmer, M.A.; Snelgrove, P.; Strayer, D.; Marcin Weslawski, J. (2001). The function of marine critical transition zones and the importance of sediment biodiversity. Ecosystems 4: 430–451.

– Lewitus, A.J., Horner, R.A., Caron, D.A., Garcia-Mendoza, E., Hickey, B.M., Hunter, M., Huppart, D.D., Kudela, R.M., Langlois, G.W., Largier, J.L., Lessard, E.J., RaLonde, R., Rensel, J.E.J., Strutton, P.G., Trainer, V.L., Tweddle, J.F.. 2012. Harmful algal blooms along the North American west coast region: History, tgrends, causes and impacts. Harmful Algae 820:1-27.

– Lindegren, M., Blenckner, T. and Stenseth, N.C. 2012. Nutrient reduction and climate change cause a potential shift from Pelagic to Benthic pathways in a Eutrophic Marine Ecosystem. Global Change Biology. DOI: 10.1111/j.1365-2486.2012.02799.x

– Liu, Y., Sumaila, U.R. 2008. Can farmed salmon production keep growing? Marine Policy 32 (2008) 497-501.

– Livingstone M.E. 1987. Food resource use among five flatfish species (Pleuronectiformis) in Wellington Harbour, New Zealand. N.Z.J.Mar.Freshw.Res.21:281-293.

– Lucas, J.S. & Southgate, P.C. (eds). 2012. Aquaculture, Farming Aquatic Animals and Plants. Second Edition. Chapter 4:Environmental Aspects: Martin Kumar and Simon Cripps: pp. 84-106. Blackwell Publishing Ltd.

– Lundau, M., Voss, M., Erickson, M., Dziallas, C., Casciotti, K. and Duckjlow. H. 2012. Excess nitrate loads to coastal waters reduces nitrate removal efficiency: mechanism and implications for coastal eutrophication. Environmental Microbiology, DOI: 10.1111/j.1462-2920.2012.02773.x

– MacGarvin, M. 2000. Scotland‘s secret? Aquaculture, nutrient pollution eutrophication, and toxic blooms. Modus Vivendi. WWF Scotland. Available at http://www.wwf.no/english/aquaculture/scotlands_secret.pdf. – 75 –

– MacKenzie, L. 1991. Toxic and noxious phytoplankton in Big Glory Bay, Stewart Island, New Zealand. Journal of Applied Phycology 3: 19-34.

– MacKenzie, L.A., Smith, K.F., Rhodes, L.L., Brown, A., Langi, V., Edgar, M., Lovell, G., Preece, M. 2011. Mortalities of sea-cage salmon (Oncorhynchus tshawytscha) due to a bloom of Pseudochattonella verruculosa (Dictyochophyceae) in Queen Charlotte Sound, New Zealand. Harmful Algae (2011), doi:10.1016/

j.hal.2011.07.003

– MacLeod, C.K., Motschaniwskyi, N.A., Crawford, C.M. 2006. Evaluation of short term fallowing as a strategy for the management of recurring organic enrichment under salmon cages. Marine Pollution Bulletin 52 (2006) 1458–1466

– Magaña, H. A.; Contreras, C; & Villareal, T. A. (2003) A historical assessment of Karenia brevis in the western Gulf of Mexico. Harmful Algae, vol.2, No.3, pp.163-171, ISSN 1568-9883

– Martin, G.R. 2011. Through birds eyes: insight into avain sensory ecology. J.Ornithology DOI 10.1007/s10336-011-0771-5.

– McClelland, J. W., and I. Valiela. 1998. Changes in food web structure under the influence of increased anthropogenic nitrogen inputs to estuaries. Mar. Ecol. Prog. Ser. 168:259-271.

– McGlathery, K.J., Sundbäck, K., Anderson, I.C. 2007. Eutrophication in shallow coastal bays and lagoons: the role of plants in the coastal filter. Mar. Ecol. Prog. Ser. 348: 1–18.

– Meire, L. 2011. Impact of global change on coastal hypoxia. Thesis Master of Science in Marine and Lacustrine Sciences. University of Gent 2011.

– Mente, E., Pierce, G.J., Santos, M.B. and Neofitou, C. 2006. Effect of feed and feeding in culture of salmon on the marine aquatic environment: a synthesis for European aquaculture. Aquaculture International (2006) 14: 499-522.

– Meynier, L., Mackenzie, D.D.S., Duigan, P.J., Chilvers, L., Morel, P.C.H. 2009. Variabilityin the diet of New Zealand sea lion (Phocarctos hookeri) at the Auckland Islands, New Zealand. Marine Mammal Science, Vol 25, issue 2:302-326.

– Milewski, E. 2006. Salmon Aquaculture: Lessons from New Brunswick. PowerPoint from Science Advisor Conservation Council of New Brunswick, Port Mouton, Nova Scotia, November 15, 2006. – 76 –

– Millenium Ecosystem Assessment. Ecosystems and Human Well Being: Current State and Trends. Island Press 2005

– Milroy, S.P., Dieterle, D.A., He, R., Kirkpatrick, G.J., Lester, K.M., Steidinger, K.A., Vargo, G.A., Walsh, J.J. and Weisberg, R.H.. 2008. A three-dimensional biophysical model of Karenia brevis dynamics on the west Florida shelf: A look at physical transport and potential zooplankton grazing controls. Continental Shelf Research 28 (2008) 112–136

– Miskelly, C.M.; Dowding, J.E.; Elliott, G.P.; Hitchmough, R.A.; Powlesland, R.G.; Robertson, H.A.; Sagar, P.M.; Scofield, R.P.; Taylor, G.A. (2008). Conservation status of New Zealand birds, 2008. Notornis 55: 117-135.

– Moffitt, C.M. 2003. The implications of Aquaculture Production and Development on Sustainable Fisheries. Manuscript in Press, Symposium on Sustainable Fisheries. American Fisheries Society Symposium.

– Morrison, M.A.; Lowe, M.L.; Parsons, D.M.; Usmar, N.R.; McLeod, I.M. 2009. A review of land-based effects on coastal fisheries and supporting biodiversity in New Zealand. New Zealand Aquatic Environment and Biodiversity Report No. 37. 100 p.

– Navarrete-Mier, F., Sanz-Lázaro, C., Marín, A. 2010. Does bivalve mollusc polyculture reduce marine fin fish farming environmental impact? Aquaculture 306: 101-107.

– Navarro, N., Leakey, R. J. G. and Black, K.D. 2008. Effects of salmon cage aquaculture on the pelagic environment of temperate coastal waters: seasonal changes in nutrients and microbial community. Mar. Ecol. Prog. Ser. Vol. 361: 47-58.

– Nelson, A. 1971. King shags in the Marlborough Sounds. Notornis 18: 30-37.

– Neofitou, N., Vafidid, D., Klaoudatos, S. 2010. Spatial and temporal effects of fish farming on benthic community structure in a semi-enclosed gulf of the Eastern Mediterranean. Aquacult Environ Interact 1: 95–105, 2010

– Nyström, M., Norström, A.V., Blenckner, T., de la Torre-Castro, M., Eklöf, J.S., Folke, C., Österblom, H., Steneck, R.S., Thyresson, M., and Troell, M..Confronting Feedbacks of degraded Marine Ecosystems. Ecosystems DOI: 10.1007/s10021-012-9530-6

– Officer CB, Ryther J.H. (1980) The possible importance of silicon in marine eutrophication. Mar Ecol Prog Ser 3:83–91 – 77 –

– Olsen Y. and Olsen L.M. 2008. Environmental Impact of Aquaculture on Coastal Planktonic Ecosystems. K. Tsukamoto, T. Kawamura, T. Takeuchi, T. D. Beard, Jr. and M. J. Kaiser, eds. Fisheries for Global Welfare and Environment, 5th World Fisheries Congress 2008, pp. 181–196.

– Paerl, H. W. 1997. Coastal eutrophication and harmful algal blooms: importance of atmospheric deposition and groundwater as ―new‖ nitrogen and other nutrient sources. Limnol. Oceanogr. 42(2): 1154-1165.

– Parker, R. 2012. Review of life cycle assessment research on products derived from fisheries and aquaculture: A report for seafish as part of collective action to address greenhouse gas emissions in seafood. Final report prepared for Sea Fish Industry Authority, Edinburgh, UK, 2012.

– Parris, D. 2011. Phytoplankton Ecology in the Skidaway River Estuary. A Thesis Submitted to the Graduate Faculty of Georgia Southern University University in Partial Fulfillment of the requirements for the degree Masters of Biology.

– Pennycuick, C.J. 1987a. Flight mechanics in seabirds. In: Croxall, J.P. (ed) Seabirds: Feeding ecology and role in marine ecosystems: 43-62. Cambridge: Canbridge University Press.

– Pennycuick, C.J. 1987b. Flight of aucks (Alcidae) and other northern sea birds compared with southern Procellariformes: J. Exp. Biol. 128:335-347.

– Petrell, R.J. and Jone, R.E. 2000. Power requirement of swimming in Chinook salmon and Atlantic salmon and implications for food conversion and growth performance. Aquacultural Engineering 22 (2000) 225– 239

– Petterson, L.H. and Pozdnyakov, D. 2013. Monitoring of Algal Blooms. Springer Praxis Books, 2013, 25-47, DOI: 10.1007/978-3-540-68209-7_2

– Pitta, P. Tsapakis, M., Apostolaki, E. T., Tsagaraki, T., Holmer, M.. Karakassis, I. 2009. ‗Gost nutrients‘ from fish farms are treansferred up the food web by phytoplankton grazers. Mar Ecol Prog Ser 374: 1–6, 2009

– Pitta, P., Apostolaki, E.T., Giannoulaki, M., Karaakassis, I. 2005. Mesoscale changes in the water column in reponse to fish farming zones in three coastal areas in the Eastern Mediterranean Sea. Estuarine, Coastal and Shelf Science 65 (2005) 501-512

– Pohle, G., B. Frost, and R. Findlay. 2001. Assessment of regional benthic impact of salmon mariculture within the Letang Inlet, Bay of Fundy. ICES Journal of Marine Science 58:417-426. – 78 –

– Purcell, J.E. 2012. Jellyfish and Ctenophore Blooms with Human Proliferations and Environmental Perturbations. Annual Review of Marine Science. Vol. 4: 209-235.

– Regnier, E. and Schubert, K. 2012. Modeling of Interactions between aquaculture and capture fisheries.Paris School of Economics, University Paris 1 Pantheon Sorbonne.

– Reid, G.K., Liutkus, M., Bennett, A., Robinson, S.M.C., MacDonald, B. and Page, F. 2010. Absorption efficiency of blue mussels (Mytilus edulis and M. trossulus) feeding on Atlantic salmon (Salmo salar) feed and faecal particulates: Implications for integrated multi-trophic aquaculture.. Aquaculture 299:165-169.

– Rhodes, L.L., Mackenzie, A.L., Kaspar, H.F. and Todd, K.E. 2001. Harmful Algae and mariculture in New Zealand. ICES Journal of Marine Science, 58: 398-403

– Sagar, P. 2008. Assessment of the potential impacts on waders and seabirds of finfish farming in the Firth of Thames. NIWA Client Report: CHC2008-094 October 2008 NIWA Project: EVW009501

– Sanderson, J.C., Cromney, C.J., Dring, M.J. and Kelly, M.S.. 2008. Distribution of nutrients for seaweed cultivation around salmon cages at farm sites in north-west Scotland. Aquaculture 278 (2008) 60–68

– Sanderson, J.C., Dring, M.J., Davidson, K. and Kelly, M.S. 2012. Culture, yield and bioremediation of Palmaria palmate and Saccharina latissima adjacent to fish farm cages in north west Scotland. Aquaculture in press 4th April 2012.

– Scanes, P. 2012. Nutrient Loads to Protect Environmental Values in Waituna Lagoon, Southland NZ.

– Schuckard, R. 1994. New Zealand king shag (Leucocarbo carunculatus) on Duffers Reef, Marlborough Sounds. Notornis 41: 93-108.

– Schuckard, R. 2005. Marlborough Sounds, an important nesting and wintering area for white-fronted terns. Notornis 2005, Vol 52:165-167.

– Schuckard, R. 2006a. Distribution of New Zealand king shag (Leucocarbo carunculatus) foraging from the Trio Is and Stewart I colonies, Marlborough Sounds, New Zealand. Notornis 53(3): 291-296.

– Schuckard, R. 2006b. Population status of the New Zealand king shag (Leucocarbo carunculatus). Notornis 53(3): 297-307.

– Seitzinger S. P., and R. W. Sanders. 1999. Atmospheric inputs of dissolved organic nitrogen stimulate estuarine bacteria and phytoplankton. Limnol. Oceanogr. 44(3) 721-730. – 79 –

– Selman, M., Greenhalgh, S., Díaz, R. and Sugg, Z. (2008), Eutrophication and hypoxia in coastal areas: A global assessment of the state of knowledge. World Resources Institute Policy Note, No, 1. Washington D.C. 6 p.

– Shumway, S.E., Allen, S.M., Boersma, P.D. 2003. Marine birds and harmful algal blooms:sporadic victims or under-repoted events.? Harmful Algae Vol. 2, issue 1:1-17.

– Signorini, S.R., Häkkinen, S., Gudmundsson, K., Olsne, A., Olafsson, J., Reverdin, G., Henson, S.A., McClain, C.R. and Worthen, D.L. 2012. The role of phytoplankton dynamics in the seasonal and interannual variability of carbon in the subpolar North Atlantic – a modelling study. Geosci. Model Dev., 5, 683–707, 2012

– Silvert, W., 1992. Assessing environmental impacts of finfish aquaculture in marine waters. Aquaculture 107, 67–79.

– Skogen, M.D., Eknes, M., Asplin, L.C. and Sandvik, A.D. 2009. Modelling the environmental effects of fish farming in a Norwegian fjord. Aquaculture 298 (2009) 70 – 75.

– Smayda, T.J. 1990. Novel and nuisance phytoplankton bloom in the sea: evidence for a global epidemic. In: Toxic Marine Phytoplankton. (Graneli, E., Sundstrom, B., Edler, L. and Anderson, D.M., eds.) pp. 29-40. Elsevier Science Publishing Co., New York.

– Smith, V.H., Tilman, G.D. and Nekola, J.C. 1999. Eutrophication: impacts of excess nutrient inputs on freshwater, marine, and terrestrial ecosystems. Environmental Pollution 100 (1999): 179-196.

– Spear, L.B.; Ainley, D.G. 1997. Flight behaviour of seabirds in relation to wind direction and wing morphology. Ibis 139: 221-233.

– Spear, L.B.; Ainley, D.G. 1997. Flight speed of seabirds in relation to wind speed and direction. Ibis 139: 234- 251.

– Staniford, D. 2002. Sea cage fish farming: an evaluation of environmental and public health aspects (the five fundamental flaws of sea cage fish farming). Paper presented by Don Staniford at the European Parliament‘s Committee on Fisheries public hearing on ‗Aquaculture in the European Union: Present Situation and Future Prospects‘, 1st October 2002: http://www.europarl.eu.int/hearings/20021001/pech/programme_en.pdf,

– Stead, S.M. and Laird, L. 2002. The Handbook of Salmon Farming. Spriner Praxis Books/Food Sciences. – 80 –

– Stevens, D.W. 1993. A comparative study of otoliths and associated labyrinthine apparatus of the New Zealand flatfish (order Pleuronectiformes). Part one, Introduction and comparative morphology. Thesis Victoria University of Welington.

– Stigebrandt, A., Aure, J., Ervik, A. and Hansen, P.K.. 2004. Regulating the local environmental impact of intensive marine fish farming III. A model forestimation of holding capacity in the modelling – Ongrowing fish farm – Monitoring system. Aquaculture 234 (2004) 239 – 261.

– Stone, N. 1994. Mixing bighead carp culture with catfish farming. The catfish journal 9(2):10-23.

– Strain, P.M. and B.T. Hargrave. 2005. Salmon aquaculture, nutrient fluxes and ecosystem processes in southwestern, New Brunswick. Chapter 2 in: The Handbook of Environmental Chemistry, Vol. 5: Water Pollution. Environmental Effects of Marine Finfish Aquaculture. Springer, Berlin Heidelburg.

– Sunda, W.G. and Shertzer. K.W. 2012. Modeling ecosystem disruptive algal blooms: positive feedback mechanisms. Mar.Ecol.Prog.Ser.. Vol.447:31-47.

– Sutherland, T., A. J. Martin, and C. D. Levings. 2001. Characterization of suspended particulate matter surrounding a salmonid net-pen in the Broughton Archipelago, British Columbia. ICES Journal of Marine Science 58:404-410.

– Sutton, P.J.H. and Hadfield, M.G. 1998. Aspects of the hydrodynamics of Beatrix Bay and Pelorus Sound, New Zealand. New Zealand Journal of Marine and Freshwater Research, 1997, Vol. 31: 271-279

– Tacon A. G.J. and Metian, M. 2008. Global overview on the use of fish meal and fish oil in industrially compounded aquafeeds: Trends and future prospects. Aquaculture, 285: 146-158.

– Tanner, J.E. and Fernandes, M. 2010. Environmental effects of yellowtail kingfish aquaculture in South Australia. Aquaculture Environment Interactions, Vol.1: 155-165.

– Taylor, D., Keeley, N., Forrest, R., Knight, B., Dunmore, R.. 2010. Assessment of the environmental effects of converting a mussel farm to salmon farm, MF 8080, Port Ligar. Report No. 1883.

– Thorpe, J.E., Talbot, C., Miles, M.S., Rawling, C. and Keay, D.S. 1990. Food consumption in 24 hours by Atlantic salmon (Salmo salar L.) in a sea cage. Aquaculture 90: 41–47.

– Thrush SF, Whitlach RB (2001) Recovery dynamics in benthic communities: balancing detail with simplification. In: Reise K (ed) Ecological comparisons of sedimentary shores. Springer-Verlag, Berlin, p 297–311 – 81 –

– Till, C, E. 2011. Species, sub Species or New Species? A Phylogenetic Analyses of the Blue Eyed Shag Complex. University of Otago, Dunedin, Genetics.

– Timmerman, K., Norkko, J., Janas, U., Norkko, A., Gustafsson, Bo.G., Bonsdorff, E. 2012, Modelling macrofaunal biomass in relation to hypoxia and nutrient loading. Journal of Marine Systems (2012), doi: 10.1016/j.jmarsys.2012.06.001

– Treasurer, J.W., Hannah, F. and Cox, D. 2003. Impact of a phytoplankton bloom on mortalities and feeding response of farmed Atlantic salmon, Salmo salar, in west Scotland. Aquaculture 218: 103-113.

– Trujillo, P. 2008. Using a Mariculture Sustainability Index to Rank the Performances of 53 Countries. Fiesheries Centre, University of British Colombia, Vancouver, Canada.

– Tsutsumi, H., Kikuchi, T., Tanaka, M., Higashi, T., Imasaka, K., Miyazaki,M., 1991. Benthic faunal succession in a cove organically polluted by fish farming. Mar. Pollut. Bull. 23, 233–238.

– Tuten, J. S., and J. W. Avault. Jr. 1981. Growing red swamp crawfish (Procambarus clarkii) and several North American fish species together. Progressive Fish-Culturist 43:97-99.

– Valiela I, McClelland J, Hauxwell J, Behr PJ, Hersh D, Foreman K (1997) Macroalgal blooms in shallow estuaries: controls and ecophysiological and ecosystem consequences.

– Vinneras, B., 2001. Faecal separation and urine diversion for nutrient management of household biodegradable waste and waste water. Department of Agricultural Engineering, Report 245, SLU, Uppsala.

– Vinnerås, B., Jönsson, 2002. The performance and potential of faecal separation anhd urine diversion to recycle plant nutrients in household wastewater. Bioresource Technology 84:275-282.

– Vizzini, S. and Mazzola, A. 2004. Stable isotope evidence for the environmental impact of a land based fish farm in the western Mediterranean. Marine Pollution Bulletin, Volume 49, issues 1-2, pp 61-70.

– Wallace, C., Weeber, B. and Buchanan, S. (eds) 1998. Seaviews, Marine Ecosystem Management: Obligations and Oppertunities. Proceedings of the conference held in Wellington, 11 – 14th of February 1998.

– Wanless S., Harris M.P. and Morris J.A. 1991. Foraging range and feeding locations of Shags Phalacrocorax aristotelis during chick rearing. Ibis 133: 30-36 – 82 –

– Wanless, S., Finney, S.K., Harris, M.P. and McCafferty, D.J. 1999. Effect of the diel light cyscle on the diving behaviour of two bottom feding marine birds: the blue-eyed shag Phalacrocorax atriceps and the European shag P. aristotilis. Mar. Ecol. Prog. Ser. 188:219-224.

– Wanless, S.; Bacon, P.J.; Harris, M.P.; Webb, A.D.; Greenstreet, S.P.R.; Webb, A. 1997. Modeling environmental and energetic effects on feeding performance and distribution of shags (Phalacrocorax aristotelis): integrating telemetry, geographical information systems, and modeling techniques. ICES journal of marine science 54: 524-544.

– Wanless, S.; Harris, M.P. 1993. Use of mutually exclusive foraging areas by adjacent colonies of blue-eyed shags (Phalacrocorax atriceps) at South Georgia. Colonial waterbirds 16: 176-182.

– Weber, M.L. 1997. Farming Salmon: A Briefing Book. Consulatative Group on Biological Diversity.

– Wetzel R. 1983. Limnology (2nd edition). Saunders College Publishing, Toronto. 767 pp.

– White, C.R., Day, N., Butler, P.J., Martin, G.R. 2007 Vision and Foraging in Cormorants: More like Herons than ONE 2(7): 1-6.Hawks? PLoS

– Whitmarsh, D.J., Cook, A.J., Black, K.D. 2006 Searching for sustainability in aquaculture: An investigation into economic prospects for an integrated salmon-mussel production system. Marine Policy 30: 293-298.

– Wiber, M.G., Young, S. and Wilson, L. 2009. Impact of Aquacultue on Commercial Fisheries: Fishermen‘s Local Ecological Knowledge. Human Ecology, DOI 10.1007/s10745-011-9450-7

– Wildish, D. J. et al. 2004. Acoustic detection of organic enrichment in sediments at a salmon farm is confirmed by independent groundtruthing methods. Marine Ecology Progress Series 267: 99-105.

– Wildish, D.J., Martin, J.L., Wilson, A.J., Saulnier, A.M. 1990. Environmental monitoring of the Bay of Fundy salmonid mariculture industry during 1988 and 1989. Can.Tech.Rep.Fish.Aquat.Sci.

– Wilson, R.P. and Wilson, M.P. 1988. Foraging behaviour in four sympatric cormorants. Journal of animal ecology, 57:943-955.

– Wilson, R.P. and Wilson, M.P. 1988. Foraging behaviour in four sympatric cormorants. Journal of animal ecology, 57:943-955.

– Winder, M. and Sommer, U. 2012. Phytoplankton response to a changing climate. Hydrobiologia DOI 10.1007/s10750-012-1149-2 – 83 –

– Worm B., and H. K. Lotze. 2000a. Nutrient pollution, low-trophic level harvesting and cumulative impacts on coastal ecosystems. In: Proceedings of the symposium ―Rockweed: management in the face of scientific uncertainty‖,ed. R.W. Rangeley. St. Andrews, New Brunswick, Canada. Huntsman Marine Science Centre.

– Worm B., and H. K. Lotze. 2006. Effects of eutrofication, grazing, and algal blooms on rocky shores. Limnol. Oceanogr., 51: 569-579. 2006

– Worm, B., H. K. Lotze, and U. Sommer. 2000b. Coastal food web structure, carbon storage, and nitrogen retention regulated by consumer pressure and nutrient loading. Limnolology and Oceanography 45(2):339- 349.

– Worm, B., H. K. Lotze, C. Boström, R. Engkvist, V. Labanauskas, and U. Sommer. 1999. Marine diversity shift linked to interactions among grazers, nutrients and dormant propagules. Mar Ecol. Prog. Ser. 185: 309-314.

– Wu, R.S.S., Lam, K.S., MacKay, D.W., Lau, T.C. and Yam, V. 1994. Impact of marine fish farming on water quality and bottom sediment: A case study in the sub-tropical environment. Marine Environmental Research 38: 115–145.

– Wu, RSS, 1995. The environmental impact of marine fish culture: Towards a sustainable future. Mar. Poll. Bull. 31: 159- 166.

– Wybourne, B.A. 2012. Statement of evidence of Ben Wybourne in relation to feed discharge for New Zealand King Salmon Co. Ltd.

– Yanju, H., Danling, T., Long, Y. And Qianguo, X. 2011. Nutrient and chlorophyll a anomaly in red tide periods of 2003- 2008 in Sishili Bay, China. Chinese Journal of Oceanology and Limnology. Vol.29 No.3, p 664-673.

– Zeldis, J. 2008. Exploring the carrying capacity of the Firth of Thames for finfish farming: a nitrogen mass- balance approach. Prepared for Environment Waikato. NIWA Client Report: CHC2008-02. June 2008. NIWA Project: EVW08501

– Zhou, M-j., Shen, Z-l., Yu, R-c. 2008. Responses of a coastal phytoplankton community to increased nutrient inputs from the Changjiang (Yangtze) River. Continental Shelf Research 28 (2008) 1483– 1489

– 84 –

Rob Schuckard

PO Box 98

Rai Valley 7145

03 5765371 [email protected]