Download The Implementation of UNGA Resolutions 61/105 and 64/72 in the

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts
no text concepts found
Transcript
ocean
The Implementation of UNGA Resolutions
61/105 and 64/72 in the Management
of Deep-Sea Fisheries on the High Seas
A report from the International Programme on the State of the Ocean
Dr Alex D. Rogers
Matthew Gianni
MAY 2010
The International Programme on the State of the
Ocean (IPSO) brings together world experts in the
science, socioeconomics and governance of marine
ecosystems to identify how humankind is changing
the capacity of the Global Ocean to support life and
human societies on Earth.
IPSO will use this knowledge to identify solutions
to restore the health of the Ocean, so as to sustain
environmental security and benefits for the present
and future generations. The programme will
communicate its findings to the public, industry and
policymakers in order to impel the required changes in
human behaviour needed to achieve these solutions.
www.stateoftheocean.org
D
The Deep Sea Conservation Coalition (DSCC) is
a coalition of over 60 organizations worldwide
promoting fisheries conservation and the
protection of biodiversity on the high seas.
The DSCC has been actively involved in the
international debate and negotiations
concerning the adverse impacts on deep-sea
biodiversity in areas beyond national
jurisdiction from bottom trawling and other
methods of bottom fishing on the high seas
since 2003/2004.
www.savethehighseas.org
The Implementation of UNGA Resolutions 61/105 and 64/72 in
the Management of Deep-Sea Fisheries on the High Seas
Dr Alex David Rogers
Scientific Director,
International Programme on the State of the Ocean,
Institute of Zoology,
Zoological Society of London,
Regent’s Park,
London,
NW1 4RY
Matthew Gianni
High Seas Fisheries Consultant,
Political and Policy Advisor, Deep Sea Conservation Coalition
Amsterdam,
The Netherlands
Contents
SUMMARY
2
SUMMARY TABLE
6
RECOMMENDATIONS
8
INTRODUCTION
10
METHODS
13
Northeast Atlantic Ocean
15
Northwest Atlantic Ocean
35
Mediterranean Sea
45
Southwest Atlantic Ocean
51
North Pacific Ocean
56
South Pacific Ocean
62
Southwest Indian Ocean
68
Cover photograph:
Southern Ocean
72
Mediterranean roughy (Hoplostethus mediterraneus), over coral garden
habitat mainly comprising Acanthogorgia hirsuta, Faial Island, Azores, North
Atlantic, 350m depth. © A.D. Rogers and Rebikoff Foundation.
REFERENCES
82
ANNEXES
93
Reviewed by Dr Richard Haedrich
Professor of Fisheries Biology emeritus,
Department of Biology,
Memorial University
Newfoundland
Citation:
Rogers, A.D., Gianni, M. (2010) The Implementation of UNGA Resolutions
61/105 and 64/72 in the Management of Deep-Sea Fisheries on the
High Seas. Report prepared for the Deep-Sea Conservation Coalition.
International Programme on the State of the Ocean, London, United
Kingdom, 97pp.
About this report:
This report was prepared for the Deep-Sea Conservation Coalition by the
International Programme on the State of the Ocean.
The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 1
Summary
details of fishing history, intended fishing
operations, gear to be used, a full definition
of VMEs likely to be encountered, and a full
ecological risk assessment in consultation with
scientists, managers and industry to assess
the potential impacts of the proposed fishing
operations. Other impact assessments lacked
sufficient information to assess the impacts of
proposed fishing operations or were based on
incorrect assumptions about the presence or
lack of presence of VMEs. In addition, several
RFMOs have not required impact assessments
for exploratory fisheries in new areas and/
or existing fishing areas, despite the UNGA
resolutions and FAO Guidelines (FAO, 2009a)
that call for all deep-sea bottom fisheries to be
assessed.
For the past eight years, the issue of protecting biodiversity in the deep sea in
areas beyond national jurisdiction – the high seas – has been extensively debated
by the United Nations General Assembly (UNGA) and in other international
fora. The UNGA adopted a series of resolutions, beginning with Resolution
59/25 in 2004, which called on high seas fishing nations and regional fisheries
management organisations (RFMOs) to take urgent action to protect vulnerable
marine ecosystems (VMEs) from destructive fishing practices, including bottom
trawl fishing, in areas beyond national jurisdiction (UNGA, 2004).
A report from the United Nations (UN)
Secretary General in 2006 on progress on
the implementation of the 2004 resolution
concluded that little action had been taken to
protect deep-sea ecosystems on the high seas
from the adverse impacts of bottom fisheries
despite the fact that “deep-sea habitats in these
areas are extremely vulnerable and require
protection”. (UNSG, 2006)1
As a result of a review by the UNGA regarding
the effectiveness of the measures called for in
Resolution 59/25, the UNGA called for a series
of specific actions to be taken by states and
RFMOs in UNGA Resolution 61/105, adopted by
consensus in December 2006 (UNGA, 2007).
Resolution 61/105 committed nations that
authorise their vessels to engage in bottom
fisheries on the high seas to take a series
of actions, outlined in Paragraph 83 of the
resolution (see Annex I of this report). The four
main action points are summarised as follows.
1. Paragraph 204: “Some
States have undertaken,
or are in the process of
undertaking, extensive
efforts to protect some
fishery habitat areas
within their national
jurisdiction, in particular
through the establishment
of protected areas.
However, this is not the
case on the high seas,
though deep-sea habitats
in these areas are
extremely vulnerable and
require protection.”
•Conduct assessments of whether bottom
fishing activities have significant adverse
impacts (SAIs) on VMEs.
• To ensure that if fishing activities have
significant adverse impacts they are managed
to prevent such impacts, including through
closing areas to bottom fishing where VMEs
are known or likely to occur, or they are not
authorised to proceed.
•To establish and implement protocols to cease
fishing where an encounter with VMEs occurs
during fishing activities, and to report such
encounters so that appropriate measures can
be adopted with respect to that site.
•To implement measures in accordance with
the precautionary approach, ecosystems
approaches and international law, and to
sustainably manage deep-sea fish stocks.
A set of International Guidelines for the
Management of Deep-Sea Fisheries in the High
Seas (FAO Guidelines) were then negotiated
under the auspices of the United Nations Food
and Agriculture Organization (UN FAO) to, inter
alia, further define and agree to criteria for
the conduct of impact assessments of high
seas bottom fisheries; identify VMEs; and then
assess whether deep-sea fisheries would have
“significant adverse impacts” on VMEs. The
FAO Guidelines were adopted in August 2008.
Key elements of the Guidelines are contained in
Annex II of this report (FAO, 2009a).
In 2009, the UNGA determined that Resolution
61/105 had not been implemented sufficiently.
As a result the General Assembly adopted
additional provisions in Resolution 64/72
(UNGA, 2009). This resolution reaffirmed the
2006 resolution and made it clear that the
measures called for in Resolution 61/105
should be implemented, consistent with the
FAO Guidelines, by flag states and RFMOs prior
to allowing, or authorising, bottom fishing on
the high seas to proceed. Resolution 64/72
placed particular emphasis on conducting
impact assessments of bottom fisheries on the
high seas and called on states and RFMOs to
“ensure that vessels do not engage in bottom
fishing until such assessments have been
carried out”. Resolution 64/72 further called for
stock assessments and conservation measures
to ensure the long-term sustainability of deepsea fish stocks and non-target species, and
the rebuilding of depleted fish stocks (UNGA,
2009: Paras 119–120). Key paragraphs of both
resolutions are contained in Annexes I and III of
this report.
A comprehensive review of the extent to which
RFMOs and states have been implementing the
2 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Argos Georgia in Port
Stanley, the Falkland
Islands. U.K. vessel
involved in fishing for
toothfish (Dissostichus
spp.) in the Ross Sea,
2008/2009. © A.D. Rogers
relevant UNGA resolutions has not previously
been conducted. This report assesses the
measures and regulations adopted with regards
to the four key actions in the 2006 UNGA
Resolution 61/105 and reinforced by Resolution
64/72 by the following RFMOs: North East
Atlantic Fisheries Commission (NEAFC);
Northwest Atlantic Fisheries Organization
(NAFO); General Fisheries Commission for
the Mediterranean (GFCM); South East
Atlantic Fisheries Organisation (SEAFO); and
Commission for the Conservation of Antarctic
Marine Living Resources (CCAMLR). The report
also reviews the interim measures adopted by
the states participating in the negotiation of
the new North Pacific Fisheries Commission
(NPFC), the South Pacific Regional Fisheries
Management Organisation (SPRFMO), and in the
Southern Indian Ocean. The review covers the
measures adopted both prior to and in response
to the 2006 UNGA resolution. The key findings
of the report include the following.
Conducting impact
assessments of individual
bottom fishing activities
The degree to which nations conducted impact
assessments varied widely. Despite the call
from the UNGA for impact assessments for all
bottom fisheries in the high seas, some RFMOs
have had no Contracting Parties conduct impact
assessments (e.g. NEAFC, NAFO), while in other
areas all Contracting Parties have submitted
impact assessments (e.g. CCAMLR, NPFC),
or some Contracting Parties have conducted
impact assessments (e.g. SPRFMO). The impact
assessments undertaken also varied in their
scope. In some cases, Contracting Parties
conducted full risk assessments that included
Preventing impacts
on vulnerable marine
ecosystems
RFMOs have undertaken a variety of measures
to protect known or suspected VMEs within their
Regulatory Areas. In some cases, technical
measures were adopted, such as the banning of
gillnets below a certain depth or from the entire
region because of the high risk of by-catch and
ghost fishing (e.g. NEAFC, SEAFO, SPRFMO) or
prohibiting of bottom trawling (CCAMLR). Most
RFMOs have adopted spatial conservation
measures to protect VMEs, although the extent
and type of closures implemented by the
RFMOs varied (e.g. NEAFC, NAFO, SEAFO, GFCM
and, most recently, CCAMLR). Some have not
closed all areas despite strong evidence of the
presence of VMEs (e.g. NEAFC) and some have
closed very few areas despite evidence of wideranging destruction of VMEs by bottom fishing
and potential ecological consequences, not only
in terms of ecosystem function but also in terms
of loss of essential habitat for species targeted
by fisheries (e.g. GFCM). In most cases,
closures have not been implemented because
the lack of information on deep-sea ecosystems
has prevented RFMOs from identifying where
VMEs exist and scientific information on where
some VME types (e.g. stony corals) are likely
to occur has not been used. There is also
evidence that some RFMOs have limited their
interpretation of which species can form VMEs
(e.g. only corals or sponges; NEAFC, NPFC) or
what structurally constitutes a VME (e.g. only
areas where a very high density of individuals
on the seabed are recognised as VMEs; NPFC).
In most cases, this likely reflects the use of
the few example VMEs referred to in the UNGA
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 3
resolutions and FAO Guidelines rather than
being based on a scientific assessment of the
full range of types of VMEs that may be found
within a specific geographic area (FAO, 2009a).
Sustainably managing
deep-sea fish stocks
For most of the target and by-catch species
taken in deep-sea bottom fisheries on the
high seas, there is insufficient information on
the biology, life history, fishing mortality and
geographic range of stocks of these species.
This information is crucial for evaluating stock
status, sustainable harvest levels and biological
reference points for each population. In the
absence of such data it is important that
the precautionary principle is applied in the
management of deep-sea fish stocks. Instead,
the report found evidence that many deep-sea
fish stocks were not subject to assessment
or long-term management plans. Furthermore,
where specific management advice was
provided by scientists or scientific bodies (e.g.
the International Council for Exploration of the
Sea [ICES]), total allowable catches (TACs) set
by RFMOs or states often exceeded advice,
even where there was a significant possibility
of overfishing or collapse of a fish stock. The
high biodiversity of high seas fish communities
means that by-catch in many high seas fisheries
forms a significant proportion of overall catch.
In some cases, populations of by-catch species
have collapsed to the point where they have
become threatened with local extirpation or
extinction under IUCN Red List criteria. In many
cases, little action has been taken to manage
by-catch species with a low productivity, although
exceptions include skates, rays and grenadiers
in Antarctica and the banning of gillnets by
several RFMOs, which are associated with high
by-catch of species like sharks. For several
of the RFMOs reviewed, there was evidence
from observer information and catch data
from scientific advisory bodies to RFMOs, of
significant levels of misreporting, under-reporting
or non-reporting of catch, particularly of by-catch
species, in the deep-sea fisheries. For the other
RFMOs the extent of reporting of catches is
unknown. Accurate reporting of catches of target
and by-catch species is required to assess
fishing mortality on populations and, without
such data, formulation of management plans
that ensure sustainable levels of exploitation
are extremely difficult.
For most areas, with the possible exception of
the Southern Ocean, most of these deep-sea
fisheries are also not regulated sufficiently to
ensure sustainable levels of exploitation of
target species or mortality of non-target
by-catch species. Fishery management plans
for deep-sea fisheries in high seas areas and
the establishment of biological reference points
aimed at ensuring the long-term sustainability of
deep-sea fisheries are rare.
Mediterranean
roughy (Hoplostethus
mediterraneus), over coral
garden habitat mainly
comprising Acanthogorgia
hirsuta, Faial Island,
Azores, North Atlantic,
350m depth
(© A.D. Rogers and
Rebikoff Foundation.)
4 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Encounter rules
The requirement to establish rules to ensure
that fishing ceases when potential VMEs are
encountered is a complex area of the UNGA
resolutions. Implementation of these rules
is particularly problematic for deep-water
regions of the high seas where there are few
data available on benthic ecosystems and the
interactions between bottom fishing gear and
VMEs. Encounter protocols have been generally
implemented as move-on rules, whereby, at a
threshold weight of by-catch of VME-associated
species in a single trawl tow or set of static
fishing gear, a vessel moves away from the area
and reports the encounter. In some cases, the
diversity of VME-associated species is also
taken into account.
A number of significant problems with move-on
rules were identified in the present report. For
many RFMOs, move-on rules for VME encounters
apply to only a limited number of VME-related
species, despite scientific evidence of and,
sometimes, specific advice by scientific bodies
on the presence of, various types of VMEs within
RFMO Regulatory Areas. This has resulted from
RFMOs using only the example VMEs mentioned
in UNGA Resolution 61/105 and the FAO
Guidelines or from simply using move-on rules
developed by other RFMOs without considering
the specific biogeography or biodiversity within
a region. Further, the threshold by-catch weights
that trigger move-on rules are set at such a high
level by many RFMOs that they are unlikely to
result in triggering the action to cease fishing in
the vicinity of a VME, nor to report the presence
of a VME to the responsible management
authority. Many RFMOs are also using the same
threshold levels for different kinds of fishing
gear and for different kinds of organisms. These
practices fail to take into account the different
impacts of active and passive fishing gear,
nor the different vulnerability and likelihood
of retention of different VME species when
impacted by fishing gear. In most cases, this
is likely to lead to underestimation of VME
encounters. Many RFMO encounter rules require
a vessel to move two nautical miles (nm) when
a threshold weight of VME organisms is caught
as by-catch. This is likely to be ineffective as
a conservation measure for mobile fishing
gears with long tow times as it is impossible to
identify where a VME encounter occurs along
a tow using commercial bottom trawl gear
(commercial trawl tows are up to 20nm long). In
this case the mid-point of the tow, usually the
point used as the centre of the 2nm temporary
exclusion zone for the fishing vessel, could be
as far as 10nm from the actual VME. It is also
questionable whether a 2nm move-on rule is
effective for passive fishing gears, such as longlines, where the gear may be up to 20km long,
although a better idea of encounter position
can be attained by recording which VME species
were caught on which segments of the gear and
then estimating the area of encounter on the
seabed from the position of deployment. Several
RFMOs (e.g. NEAFC, NAFO) also use move-on
rules that differentiate between fished and nonfished areas. This is inconsistent with Paragraph
23 of the FAO Guidelines, which requires
that deep-sea fisheries should be rigorously
managed throughout all stages of their
development, including experimental, exploratory
and established phases.
The following table (see page 6) provides
an overview of actions taken, or not taken,
by existing and incipient RFMOs in relation
to the key actions called for in the 2006
UNGA Resolution 61/105 and reinforced
by Resolution 64/72.
Summary of the findings
of the report on the
implementation of UNGA
resolutions and FAO
Guidelines by RFMOs
The table shows a selection of managed and
unmanaged stocks of target and by-catch
species (the list is not necessarily complete
for unmanaged stocks). It also shows whether
scientific recommendations have been followed
by RFMOs in setting sustainable harvest levels
of target and by-catch species, whether or not
there is evidence of non-reporting or misreporting
of catches and information on closed areas, and
the application of the move-on rule. Whether
or not Contracting Parties have submitted
environmental impact assessments for fisheries
has also been included. Note: A managed
fishery is one where reliable catch data have
been collected in recent years (last five years)
and preferably where there has been fisheriesindependent scientific assessment of stock
status. On the basis of these data, harvest and
management plans have been developed by
fisheries scientists to determine TACs appropriate
to maintain the long-term sustainability of the
fishery. In some cases, where data are not
available, precautionary TACs have been set.
These cases are denoted with a ‘*’ symbol.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 5
Deep-water species
managed
Deep-water species not managed
TACs fall within
scientific
recommendations
Evidence of
misreporting
of catches
or catches
being
unreported
Closure of areas to
protect VMEs
Closures for
other reasons
Move-on rule
Threshold
Coral (kg)
Sponge (kg)
Yes >10
Yes
Yes
60
800
No
Environmental
impact
assessments of
fisheries
NEAFC
Hoplostethus atlanticus;
Micromesistius poutassou;
Sebastes mentella
Alepocephalus bairdii; A. rostratus; Argentina
silus; Beryx spp.; Brosme brosme; Chaceon
affinis; Chimaera monstrosa; Coryphaenoides
rupestris; Epigonus telescopes; Helicolenus
dactylopterus; Hydrolagus spp.; Lepidopus
caudatus; Macrourus berglax; Molva molva;
Phycis blennoides; Polyprion americanus;
sharks 2
No. Examples:
Hoplostethus
atlanticus; Sebastes
mentella
Yes
NEAFC
NAFO
Pandalus spp.; Penaeus
spp.; Rajidae; Reinhardtius
hippoglossoides; Sebastes
spp.; Urophycis tenuis
Anarhichas lupus; Anarhichas minor; Anarhichas
denticulatus; Antimora rostrata; Chimaeridae;
Coryphaenoides rupestris; Macrourus berglax;
sharks
No. Examples:
Sebastes spp. and
skates
Yes
NAFO
Yes >10
No
Yes
60
800
No
GFCM
Aristeus antennatus;
Merluccius merluccius;
Nephrops norvegicus;
Parapenaeus longirostris
Sharks; others not known
Not identified in
current report
Unknown
GFCM
Yes
<5 but fishing banned
below 1000m depth
No
No
N/A
N/A
No
SEAFO
Beryx spp.*; Chaceon
spp.*; Dissostichus
eleginoides*;
Hoplostethus atlanticus*;
Pseudopentaceros
richardsoni*; sharks*
Not Known
Unknown
Unknown
SEAFO
Yes >10
No
Yes
60
800
No
NPFC
None
Allocyttus verrucosus; Beryx decadactylus; Beryx
splendens; Chaceon spp.; Chioniocetes tanneri;
Corallium spp.; Coryphaenoides spp.; Epigonus
denticulatus; Erilepis zonifera; Helicolenus spp.,
Lepidocybium flavobrunneum; Paralomis spp.;
Physiculus spp.; Pseudopentaceros wheeleri;
sharks; Zenopsis nebulosa; and many other
species
No. Example: Beryx
splendens
Yes
NPFC
No
Yes. (Spatial
measures for
alfonsino)
Yes
50
N/A
Yes
SPRFMO
None
Allocyttus niger; Allocyttus verrucosus; Beryx
spp.; Caprodon longimanus; Centroberyx
affinis; Dissostichus eleginoides; Epigonus
spp.; ; Etelis carbunculus; Etelis coruscans;
Helicolenus spp.; Hoplostethus atlanticus;
Jasus spp.; Macrouridae; Micromesistius
australis; Mora moro; Nemadactylus spp.;
Neocyttus rhomboidalis; Paristiopterus labiosus;
Pentaceros richardsoni; Pentaceros japonica
Polyprion oxygeneios, Polyprion americanus;
Pseudocyttus maculatus;Rexea spp.; Seriola
lalandi; sharks
Not identified in
current report
Unknown
SPRFMO
No, but note that New
Zealand has closed
areas to its vessels
No
Yes
30 (New
Zealand)
100 (Spain)
50 (New
Zealand)
1000 (Spain)
Yes
SIOFA
(South
Indian
Ocean
Fisheries
Agreement)
None
Beryx decadactylus; Beryx splendens;
Epigonus spp.; Hoplostethus atlanticus;
Pseudopentaceros richardsoni; Plagiogeneion
rubiginosum.
All other low-productivity deep-sea species
taken as catch or by-catch
N/A
Unknown
SWIO
Yes
Yes
No
N/A
N/A
No
CCAMLR
Champsocephalus gunnari;
Dissostichus eleginoides;
Dissostichus mawsoni;
Macrouridae; Rajiformes
Antimora rostrata and other species
Yes
No
CCAMLR
Yes
Yes
Yes
10
10
Yes
2 Note that many of these species are listed as “regulated” by NEAFC but are only covered by general measures to reduce effort on deep-sea
fisheries (NEAFC, 2010a). These measures have not been effective at reducing catches of deep-sea species collectively and do not represent effective
management of individual species (see main report).
6 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 7
Recommendations
The following are a set of recommendations for improving the
implementation of UNGA Resolutions 61/105 and 64/72 by RFMOs and
flag states, including in regions where RFMOs are under negotiation or
have not yet been established. These are organised to reflect the four key
requirements of the resolutions.
Conduct assessments of whether
bottom fishing activities have
significant adverse impacts on VMEs
•A standard for assessments of deep-sea
bottom fisheries on the high seas should
be developed with participation of fisheries
managers, the industry and scientists.
Examples of comprehensive assessments
exist (e.g. New Zealand assessments for
CCAMLR and SPRFMO) and can be built upon.
•Part of any assessment should include
consideration of which VMEs are present
within the geographic region in which fishing
activities occur or will occur in accordance
with the FAO Guidelines. These should include
fragile habitats with a low resilience to fishing
impacts and biologically significant areas,
such as spawning grounds and threatened
or endangered species. Such data are often
unavailable for deep-sea ecosystems so this
may require investment in new research and/
or synthesis of existing data.
•States whose vessels engage in bottom
fisheries on the high seas should perform
impact assessments consistent with
the criteria agreed in the FAO Guidelines
(paragraphs 47, 42, 17–20) as a precondition
to further authorising bottom fishing in areas
that have been historically fished as well as
those where exploratory fishing activities are
proposed.
To ensure that if fishing activities
have significant adverse impacts
they are managed to prevent such
impacts, including through closing
areas to bottom fishing where VMEs
are known or likely to occur, or they
are not authorised to proceed
•Where impact assessments cannot make
a clear determination that bottom fishing
will not produce SAIs on VMEs, fishing
should be prohibited, particularly in respect
of bottom trawl fisheries, in accordance
with the precautionary approach, especially
where knowledge of deep-sea ecosystems is
deficient.
•All areas where VMEs are known or likely
to occur should be closed to bottom
fishing with immediate effect, unless or
until an assessment has determined that
management measures for fisheries in these
areas would not result in SAIs to VMEs.
•States should implement measures sufficient
to protect VMEs, even where an RFMO fails to
adopt sufficient measures, e.g. if the decisionmaking structure of an RFMO has allowed
one or more Contracting Parties to block the
adoption of measures necessary to effectively
implement UNGA Resolutions 61/105 and
64/72, the other Contracting Parties should
nonetheless establish measures to regulate
their high seas fleets to ensure the full
and effective implementation of the UNGA
resolutions.
•The widespread deep-sea bottom fisheries
on the high seas in the 1960s to 1990s
have impacted on a large area of the seabed
likely to be suitable for the occurrence of
VMEs. The species diversity of many such
ecosystems is unknown, as is the capacity for
recovery. Where there is a history of bottom
fishing on the high seas then, at a minimum,
states and RFMOs should establish closures
of representative sites in historically fished
areas where VMEs are likely to have previously
occurred, to allow for recovery or regeneration
of degraded areas.
•All closures of areas of seabed to bottom
fishing should be considered within the
framework of a network of protected areas,
with clear objectives in terms of conservation
and/or fisheries management.
8 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
To establish and implement protocols
to cease fishing where an encounter
with VMEs occurs during fishing
activities, and to report such
encounters so that appropriate
measures can be adopted with
respect to that site
•The trigger thresholds for encounter rules
should be based on rigorous scientific
analyses of relationships between
by-catch and the presence of VMEs within
the geographic region in which bottom fishing
activites take place. Such analyses can be
undertaken on fisheries-independent catch
data or on fisheries data in combination
with scientific surveys or other information.
Thresholds should be specific to particular
groups or size-classes of organisms and to
the fishing gear and methods used.
•Evidence of by-catches of VME indicator
species at levels indicated by scientists to
represent a likely encounter with a VME should
trigger an immediate (and at least temporary)
cessation of fishing and closure of the area
until an assessment of the area has been
conducted and a determination has been
made as to whether fishing can be resumed in
the area without SAIs on VMEs.
•Move-on rules should ensure that subsequent
to an encounter there is no risk of SAIs
occurring on identified VMEs as a result of
continuing fishing activities. Move-on distances
should reflect the accuracy with which the
location of a VME has been identified.
To implement measures in
accordance with the precautionary
approach, ecosystems approaches
and international law, and to
sustainably manage deep-sea fish
stocks
•The fish stocks targeted by deep-sea bottom
fisheries should be subject to scientific
assessment of status at a minimum of every
five years or more frequently where scientists
and managers consider it appropriate.
Based on such assessments, TACs should
be determined that ensure long-term
sustainability.
•The impact of fishing mortality on by-catch
species should be assessed to determine
whether there are SAIs on population viability.
Where such impacts take place, management
measures should be applied to ensure the
long-term sustainability of populations of nontarget species.
•Scientific recommendations on annual
catches and other measures to ensure the
sustainability of target and by-catch species
should be adopted by RFMOs and states
unless a clear case can be made that the
information on which such decisions were
based is inaccurate. This is likely to occur
when new information becomes available.
In situations where there is a dispute
over scientific information or advice, the
precautionary approach should be adopted
when making management recommendations
for a stock.
•High seas fisheries taking low-productivity
species (either as targeted catch or as
by-catch), where the long-term sustainability of
the target species or viability of populations
of non-target species cannot be ensured
through management plans based on sound
scientific assessment of the state of stocks or
populations, should be closed. Such fisheries
should remain closed until management
plans are in place and can ensure, with a
high degree of confidence and taking into
account any uncertainties with regard to data
or other information, that such fisheries are
sustainable and consistent with ecosystembased and precautionary approaches.
•All deep-sea bottom fisheries operating on the
high seas should ensure that data on catches,
utilised by-catch and discards are collected
accurately and to the species level. Where
there are issues of species identification
then by-catch should be retained for expert
identification on land or observers with the
expertise to accurately evaluate catch should
be carried.
•Where misreporting is suspected, systems
that ensure correct reporting of catches
should be implemented.
•In regions where there are few data,
collaborative programmes between managers,
scientists and industry should be established
to help with identification of catch and by-catch
species.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 9
Introduction
The protection of biodiversity in the deep-sea in areas beyond national jurisdiction
has been extensively debated by the United Nations General Assembly (UNGA)
and other international fora. The United Nations Food and Agriculture Organization
(UN FAO), in 2008, published a report that estimated the total global catch in
high seas bottom fisheries in 2006 was some 250,000 tonnes, representing 0.3%
of the marine capture fisheries worldwide (Bensch et al., 2008). The value of the
high seas bottom catch in 2006 was estimated at approximately $450 million US
dollars or €360 million euros. Of this amount, approximately 105,000t were caught
in high seas bottom fisheries in the North Atlantic in 2006.
The UN FAO report (Bensch et al., 2008) also
estimated that 285 vessels flagged to 27
countries engaged in high seas bottom fisheries
in 2006, though many of the vessels were only
involved in bottom fishing on the high seas on
a part-time basis. Of this number, 80 percent
were flagged to only 10 states: Spain, Republic
of Korea, New Zealand, Russia, Australia, Japan,
France, Portugal, Belize and Estonia. Over
one-third were flagged to European Union (EU)
countries and the EU fleet took one-half or more
of the total high seas catch, mostly through
bottom trawling. The EU plays an even more
significant role in high seas bottom fishing within
some geographic regions. For example, the EU
fleet is responsible for approximately 80 percent
of the catch in high seas bottom fisheries in the
northwest Atlantic and 95 percent of the catch
in the northeast Atlantic. Most high seas bottom
fishing is done by bottom trawl vessels. The
conclusions of the UN FAO report were similar to
the findings of an earlier study published by the
International Union for Conservation of Nature
(IUCN) in 2004.
Deep-sea bottom fisheries have significant
impacts on deep-sea communities formed by
emergent epifaunal animals such as corals
and sponges. Relevant studies have included
comparisons between adjacent-fished versus
unfished areas using seafloor observations with
towed cameras; observations of fishing impacts
with submersibles or remotely operated vehicles
(ROVs); acoustic imaging of the seafloor;
sampling of seabed communities in impacted
versus non-impacted areas; and documenting
by-catch of benthic invertebrates in fishing gear
(Koslow & Gowlett-Holmes, 1998; Rogers, 1999;
Roberts et al., 2000, 2009; Koslow et al., 2001;
ecosystems in general, potentially including
coral, sponge and other communities that form
VMEs, and individual species, especially if they
influence food webs within such habitats (Clark
& Koslow, 2007; DeVries et al., 2007). Offal
from hoki fisheries off New Zealand has been
observed to alter oxygen concentrations at
depths as great as 800 metres (m) and change
the composition of the benthic community there
(Clark & Koslow, 2007). The removal of Antarctic
toothfish (Dissostichus mawsoni) from the Ross
Sea by deep-sea fishing has been implicated
in local/regional declines in the abundance of
predators such as killer whales (DeVries et al.,
2007).
Hall-Spencer et al., 2002; Fosså et al., 2002;
Anderson & Clark, 2003; Clark & O’Driscoll,
2003; Freiwald et al., 2004; Ardron, 2005;
Gass & Willison, 2005; Mortensen et al., 2005;
Shester & Ayers, 2005; Stone, 2006; Clark &
Koslow, 2007; Edinger et al., 2007a; Althaus et
al., 2009; Clark & Rowden, 2009). Such work
has demonstrated that bottom fishing damages
or destroys long-lived epifaunal animals such
as corals, reducing the three-dimensional
complexity of the seabed and leading to
decreased species diversity and faunal biomass
(Koslow et al., 2001; Reed et al., 2005; Stone,
2006; Althaus et al., 2009; Clark & Rowden,
2009). Bottom trawling is likely to have the most
serious adverse impacts on vulnerable deep-sea
benthic species, given the size and weight of
bottom trawl gear, the scale of the seabed area
impacted by bottom trawl tows, and the fact
that it is the dominant method of bottom fishing
for deep-sea species on the high seas (Gianni,
2004; Friewald et al., 2004; Davies et al., 2007;
WGDEC, 2008). However, such effects arise not
only from bottom trawling but from all bottomcontact fishing methods, including benthic
longlines, gillnets and pots (e.g. Stone, 2006;
Edinger et al., 2007a; FAO, 2008). The intensity
of impact differs between gears and can be
influenced by fishing practices (WGDEC, 2006;
FAO, 2008).
Observations of significant adverse impacts
(SAIs) of fishing on deep-water coral
communities have been widely reported in all
oceans, including especially the northeastern
(Hall-Spencer et al., 2002; Fosså et al., 2002;
Wheeler et al., 2005) and northwestern
Atlantic (Mortensen et al., 2005; Edinger et al.,
2007a), and the northeastern (Stone, 2006;
Krieger, 1998, 2001; Stone et al., 2005) and
southwestern Pacific (Koslow & Gowlett-Holmes,
1998; Koslow et al., 2001; Clark & O’Driscoll,
2003; Rowden et al., 2004; Althaus et al., 2009;
Clark & Rowden, 2009), but recovery of deepsea ecosystems from the mechanical impacts
of bottom fishing has been poorly studied.
However, it is likely that such ecosystems will
only recover very slowly, if at all, as habitatforming corals have slow growth rates, especially
some Antipatharia and Octocorallia (Roark et
al., 2006, 2009; Sherwood & Edinger, 2009),
and the coral habitat itself may have taken
thousands of years to develop (Hall-Spencer et
al., 2002). In some areas impacted by trawling,
there have been observations of the occurrence
on the seabed of stylasterid corals, potentially
indicating that these are capable either of
surviving trawling impacts or of colonising areas
of rock relatively quickly after disturbance (Clark
& Rowden, 2009). However, in many cases
no recovery of seabed ecosystems has been
observed even many years after fishing impacts
(Waller et al., 2007; Althaus et al., 2009).
Vulnerable marine ecosystems (VMEs) on the
sea bottom may also be susceptible to the
direct and indirect effects of increased sediment
load in the water overlying the seabed that can
smother live colonies or bury hard substrata
required for settlement of larvae. Removal of
target fish species and the dumping of
by-catch or offal from fish processing can impact
The UNGA adopted Resolutions 59/25 and
61/105 in 2004 and 2006, respectively, to
address international concerns regarding the
adverse impacts of deep-sea fisheries on VMEs
and individual species, including targeted and
non-targeted fish, in the deep sea (UNGA, 2004,
2007). The latter of the two resolutions called
on states and regional fisheries management
10 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
organisations (RFMOs) to regulate high seas
bottom fisheries through conducting impact
assessments to determine whether SAIs on
VMEs would occur. Furthermore, the resolution
calls on high seas fishing nations to close areas
of the high seas to bottom fishing where VMEs
were known or likely to occur unless such fishing
could be managed to prevent SAIs on VMEs.
The resolution also called for high seas bottom
fisheries to be managed to ensure the long-term
sustainability of deep-sea fish stocks targeted
or otherwise impacted, e.g. caught as by-catch
(UNGA, 2007: Paras 80–91). Since 2006, a
number of states and RFMOs, including those
involved in high seas bottom fisheries in the
North Atlantic, northwest Pacific, South Pacific
and Southern Ocean have adopted framework
agreements to implement UNGA Resolution
61/105.
Subsequent to the adoption of the UNGA
resolution, the UN FAO hosted a series of
consultations and negotiations to draft a set
of guidelines for the implementation of UNGA
Resolution 61/105. The UN FAO International
Guidelines for the Management of Deep-Sea
Fisheries in the High Seas (FAO Guidelines),
adopted by member countries of the UN FAO in
2008, sought to elaborate science-based criteria
to identify VMEs, conduct impact assessments
of bottom fisheries on the high seas, and
determine whether “significant adverse impacts”
to such ecosystems would occur (FAO, 2009a).
In 2009, the UNGA reviewed the implementation
of Resolution 61/105, adopted in 2006.
Recognising that the implementation of
the resolution was insufficient, the General
Assembly adopted Resolution 64/72 (UNGA,
2009). This resolution reaffirmed the 2006
resolution and made it clear that the measures
called for in Resolution 61/105 should be
implemented consistent with the FAO Guidelines
by flag states and RFMOs prior to allowing,
or authorising, bottom fishing on the high
seas. Resolution 64/72 places particular
emphasis on conducting impact assessments
of bottom fisheries on the high seas and calls
on states and RFMOs to “ensure that vessels
do not engage in bottom fishing until such
assessments have been carried out”. Resolution
64/72 further calls for stock assessments and
conservation measures to ensure the long-term
sustainability of deep-sea fish stocks and nontarget species, and the rebuilding of depleted
fish stocks (UNGA, 2009: Paras 119–120).
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 11
The present report reviews the regulations
adopted by the RFMOs with responsibility
for the management of deep-sea bottom
fisheries on the high seas in the North Atlantic,
Mediterranean, southeast Atlantic, North Pacific,
South Pacific and the Southern Ocean. The
southwest Indian Ocean has also been included
as an agreement to establish an RFMO has
been adopted but is not yet in force, while
several high seas deep-water fishing companies
have undertaken voluntary conservation
measures within the region to protect deepsea ecosystems. The report includes a set of
recommendations on further actions needed by
both RFMOs and states to ensure the effective
management of bottom fisheries on the high
seas so as to protect deep-sea ecosystems and
ensure the long-term sustainability of deep-sea
fish stocks and species.
Note that the European Commission has
also implemented UNGA Resolution 61/105
through Regulation 734/2008, which requires
all European states operating deep-sea bottom
fisheries outside areas where competent RFMOs
exist to undertake impact assessments of those
fisheries or not to issue licences for such fishing
activities. An assessment has been undertaken
of deep-sea bottom fisheries by Spain for the
southwest Atlantic but this was not available in
a form suitable for analysis by the authors in
the present report. Both Spain and New Zealand
undertook assessments for fisheries in the
South Pacific and these have been analysed
under the section of this report on the South
Pacific Ocean.
Methods
This report reviews the implementation of UNGA Resolutions 59/25, 61/105
and 64/72 (UNGA, 2004, 2007, 2009) with respect to deep-sea bottom fisheries
in the high seas, mainly by RFMOs but also by individual flag states. Here,
interpretation of the requirements of Resolutions 61/105 and 64/72 has been
largely based on the UN FAO International Guidelines for the Management of DeepSea Fisheries in the High Seas (FAO, 2009a). The FAO Guidelines were developed
as practical guidance on what was required to enable fisheries managers to develop
sustainable ecosystem-based deep-sea fisheries on the high seas in accordance
with international law and agreements related to fisheries, as directed by the UNGA
resolutions. The FAO Guidelines were produced through a series of international
workshops of experts, including scientists, fisheries managers and representatives
of the fishing industry, and were agreed by Member States of the UN FAO Committee
on Fisheries in 2008/09 and endorsed by the UNGA in 2009.
To review the implementation of UNGA
resolutions by RFMOs and flag states in respect
of management of deep-sea bottom fisheries on
the high seas, this report focused on five main
areas.
Stylasterid or hydrocoral,
Faial Island, Azores, 350500m depth (© A.D.
Rogers and the Rebikoff
Foundation).
12 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
1.Development of assessment and
management regimes which ensure that
harvest levels for target species of deepsea fish characterised by low productivity
are sustainable in the long term. Here, we
define a managed fishery as one where
reliable catch data have been collected in
recent years (last five years) and where there
has been a scientific assessment of stock
status. Ideally, such assessments should
be undertaken with fisheries-independent
data. On the basis of these data, harvest
and management plans have been developed
to determine Total Allowable Catches (TACs)
and biological reference points appropriate to
maintaining the long-term sustainability
of fished populations.
2.Development of assessment and
management regimes that ensure that
by-catch of non-target species of fish and
other marine organisms are not a threat to
the viability of populations, nor to biodiversity.
Thus, harvest and management plans for
deep-sea high seas fisheries must ensure the
long-term survival of populations and species
within the regulated area.
3.Protection of known or recently identified
VMEs, and areas where VMEs are likely to
occur.
4.The development of management tools and
practices to detect and subsequently protect
previously unidentified VMEs encountered
during the course of fishing.
5.The development of impact assessments by
states for their deep-water high seas fisheries
in order to achieve sustainable ecosystembased management.
Thus the present report reviews progress in
the implementation of all aspects of the UNGA
resolutions.
To achieve this review it was necessary to
identify which species of fish were categorised
as having a low productivity and which were
vulnerable to overfishing in the areas of the
high seas regulated by RFMOs. Typically, these
were species that exhibit low rates of growth,
high longevity and low levels of fecundity and
which may aggregate during periods of spawning,
rendering them vulnerable to high levels of
fishing pressure. It was also necessary to
identify which by-catch species showed a high
vulnerability to overfishing or damage by bottom
fisheries as well as a low capacity for recovery
following impact (low resilience). This required
study of peer-reviewed scientific literature,
the work of scientific advisory bodies to state
governments and RFMOs, and analyses of data
presented by RFMOs themselves in annual
reports or the reports of RFMO committees.
Once low-productivity or vulnerable species
were identified, all the information pertaining to
the management, catch levels and population
size of each species was reviewed. Data were
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 13
obtained through RFMO annual and committee
reports, reports from other sources (e.g.
non-peer-reviewed scientific reports, reports
from governmental and non-governmental
organisations) and peer-reviewed scientific
literature. Based on a synthesis of all literature
sources, it was determined whether fisheries
were or were not being managed to ensure longterm sustainability in populations of target or
non-target species taken in individual fisheries.
To identify whether or not action had been taken
to protect known VMEs, all literature relating to
the distribution of VMEs, particularly cold-water
coral reefs, coral gardens and sponge grounds,
was reviewed and the occurrence of these
ecosystems identified in the high seas. This was
possible for the northeastern and northwestern
Atlantic but was difficult for many other regions
of the world’s oceans and seas, where levels of
scientific knowledge on the seabed fauna are
poor. All sources of literature were reviewed,
especially peer-reviewed scientific literature and
reports of scientific advisory bodies to state
governments and RFMOs, but also RFMO reports
and reports from other sources. Progress of
individual RFMOs in the protection of known VME
localities using site-specific (e.g. establishment
of protected areas) or activity-specific (e.g.
banning of gillnets) measures were also
reviewed.
Measures to identify and report previously
unidentified VMEs during fishing operations and
to subsequently act to protect such VMEs from
fisheries impacts were also reviewed for each
RFMO. In particular, attention was paid to the
design of move-on rules and the thresholds of
by-catch of VME species. Here, the aim was to
determine whether or not the threshold levels
of by-catch were based on sound scientific
analyses or previous experience within fisheries
and to determine whether they would be likely
to trigger management responses to protect
VMEs, if present. This represents one of the
complex areas of the UNGA resolutions and FAO
Guidelines and it required detailed analyses not
only of individual RFMOs and the ecosystems
within regions but also the comparison of
practices across RFMOs or what is known about
VMEs across different geographic regions.
Knowledge of the distribution and abundance of
species that comprise VMEs was gathered from
the scientific peer-reviewed literature and from
the reports of the scientific advisory bodies and
committees of RFMOs and states. RFMO annual
and committee reports, recommendations and
NorthEast Atlantic Ocean
regulations were scrutinised for details of moveon rules and threshold levels.
The Northeast Atlantic Ocean (Fig. 1) is
important in terms of fisheries production,
with catches of 9.1 million tonnes of fish
recorded for 2006 (FAO, 2009b). The North
East Atlantic Fisheries Commission (NEAFC) is
the competent RFMO in this area.
Finally, the UNGA resolutions and FAO Guidelines
require the performance of impact assessments
for deep-sea fisheries. In the present report,
it was identified whether or not states had
carried out such impact assessments. If such
assessments had been carried out then their
content was compared to the requirements
of RFMOs, as several have requested
specific information to be contained in such
assessments, as well as the FAO Guidelines.
In compiling the present report, we encountered
limitations to our assessments. The lack of
availability of data on the catch and by-catch
of deep-sea fisheries at high spatial resolution
was a significant barrier to understanding
the impacts of fishing on target and nontarget populations of fish and other species.
In addition, data for by-catch were often not
available for individual species but rather were
aggregated by genus (e.g. Sebastes spp.) or
taxonomic group such as ‘skates’ or ‘sharks.
In some instances there was also evidence
of non-reporting or misreporting of catches of
deep sea species, which is a significant problem
when assessing the effects of harvesting on
exploited fish stocks. For many regions of the
deep-sea, fisheries are exploiting stocks in areas
where there is limited or no knowledge of the
occurrence or extent of distribution of VMEs.
This is particularly the case for waters that are
a long distance from developed countries, such
as the southern Indian Ocean or large parts of
the Pacific Ocean. In such cases, it is almost
impossible, given the current state of knowledge,
to identify where fisheries may impact VMEs or
populations of by-catch species. The problem
of a lack of baseline knowledge on the diversity
and distribution of species across large parts of
the world’s oceans cannot be over-emphasised
in the context of sustainable ecosystem-based
management of fisheries.
The four main fisheries under regulation by
NEAFC are spring-spawning herring (Clupea
harengus), mackerel (Scomber scombrus),
blue whiting (Micromesistius poutassou) and
pelagic redfish (Sebastes mentella). Of these
species, blue whiting and pelagic redfish are
found in deep water, with the former classed as
mesopelagic at depths between 30–400m and
the latter as benthic or bathypelagic at depths
between 300 and >1,440m (Whitehead et al.,
1986; http://www.fishbase.org). These fisheries
make up a catch of < 4 million tonnes in the
northeast Atlantic of which approximately one
million tonnes, worth in the region of US$236
million or some €190 million, were reported as
caught in the Regulatory Area (NEAFC high seas
areas) in 2005 (Arbuckle et al., 2008; NEAFC
press release, 2009). These fisheries are mainly
pelagic and NEAFC states that they are relatively
clean (i.e. catches are close to 100 percent
target species) and do not impact on the seabed.
This assertion has been questioned in a recent
review of management by NEAFC, which called
for verification of low levels of by-catch through
scientific studies, so that a full understanding of
the ecosystem-level impacts of these fisheries
can be reached (Arbuckle et al., 2008). In
addition, there are fisheries for a variety of
demersal species, including cod, haddock, hake
and others that take place both within EU waters
and outside waters under national jurisdiction.
While all are high-productivity fish species, they
are fished using bottom-contact fishing gear that
has the potential to impact on VMEs associated
with the seabed. In particular, the haddock
fishery on the Rockall Bank has been identified
as having the potential to impact populations
of by-catch fish species as well as benthic
communities (Arbuckle et al., 2008). Finally,
there are a variety of deep-sea species of fish
that are exploited in the region using a variety of
different types of gear, including bottom trawls,
longlines, gillnets and pots.
Management of fisheries for deep-sea
species of low productivity
Pelagic redfish (Sebastes mentella)
Whether or not pelagic redfish fall within the
remit of the FAO Guidelines has not been
Figure 1.
The NEAFC
Regulatory
Area (NEAFC,
2009).
Finally, the reader should be reminded that the
report only covers geographic areas in which
high seas deep-sea fisheries are regulated by
RFMOs or areas where agreements (treaties) to
establish RFMOs are either under negotiation
or have been adopted but have not yet entered
into force. Deep-sea fisheries on the high
seas in many parts of the world’s oceans
that fall outside these geographic regions are
unmanaged, unless there is state regulation of
the activities of their flagged vessels.
14 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 15
addressed by NEAFC or states in the region,
perhaps because it is assumed this is not a
low-productivity species. However, S. mentella
exhibits several features that suggest that it may
not be able to sustain the levels of exploitation
possible for high-productivity species. These
features include ovoviviparity (live bearing) and
a high longevity (ageing studies have indicated
that adults were generally 65 years old, but
ages up to 75 years were found; Campana et
al., 1990). The International Council for the
Exploration of the Sea (ICES), a body of more
than 1,600 scientists from 200 institutes
from around the North Atlantic that is the
prime source of scientific advice on marine
ecosystems and marine living resources in
the region, ranked S. mentella highly on a
scale of vulnerability for deep-sea species,
lying below roundnose grenadier and orange
roughy (the former species having a similar
vulnerability score to S. mentella, the latter
having a considerably higher vulnerability score;
ICES, 2001; see also Koslow et al., 2000;
WGEAFM, 2008a). ICES has advised that this
species is vulnerable to overfishing in the
NEAFC Regulatory Area and the stock size at the
present time is estimated to be low compared
to the early 1990s, although clear trends have
not been apparent since 1999 (Arbuckle et al.,
2008).
Pelagic redfish are fished in two regions of
the NEAFC Regulatory Area: the Irminger Sea
between east Greenland and Iceland, and the
Norwegian Sea. These fisheries occur both
within Exclusive Economic Zones (EEZs) and
on the high seas. There has been substantial
disagreement on whether pelagic redfish form
a single or multiple stocks in the Irminger Sea
region, with objections to NEAFC’s management
proposals arising mainly from Iceland and
Russia (Arbuckle et al., 2008), and these
continued through 2009 (NEAFC, 2010a). The
result has been that no management objectives
or harvest controls (biological reference points)
have been set for the species (Arbuckle et al.,
2008; NEAFC, 2009a). NEAFC has set overall
TACs for pelagic redfish but they have been
exceeded because of illegal, unreported and
unregulated (IUU) fishing and as a result of
objections to management measures by coastal
states (Arbuckle et al., 2008). Since 2002, this
has meant that specific advice from ICES on
TACs for pelagic redfish has not been agreed
upon and catches have exceeded recommended
limits (NEAFC, 2009a). The independent
review of NEAFC concluded that there was an
urgent requirement to resolve outstanding
issues regarding stock structure by means
consistent with the precautionary principal so
that management for the species, regarded as
vulnerable, could progress in a manner that
ensured sustainability (Arbuckle et al., 2008).
In addition, the report indicated significant
problems with catch data for pelagic redfish,
with catches not always being reported, as well
as inconsistencies in acoustic survey data for
the species resulting in uncertain estimates
of stock size (Arbuckle et al., 2008). Problems
of reporting of catches of redfish were again
discussed during the NEAFC Annual Meeting
in 2009 (NEAFC, 2010a).
they should be closed to fishing. Currently, in
Icelandic waters, spawning areas are closed but
there is no TAC for blue ling and the increased
effort from longline fishing targeted at this
species is contrary to ICES advice (WGDEEP,
2009). Other spawning areas have also been
identified in northern regions off the Norwegian
continental slope (Storegga) and on the
Reykjanes Ridge near the Icelandic EEZ (ICES,
2007a: Fig 2). In November 2009 NEAFC agreed
to close an identified area during the spawning
season where another spawning aggregation
exists for blue ling in the northern part of the
Mid-Atlantic Ridge, just south of the Icelandic
EEZ (NEAFC, 2010a). This area is bounded by
the following coordinates:
Directed deep-water bottom fisheries
60°58’76
60°56’02
60°59’76
61°03’00
Ling (Molva molva)
Ling are fished in deep water but are regarded
as one of the less vulnerable species (ICES,
2008). Ling are mainly targeted by longline
fisheries within EEZs but some catches are from
high seas areas, such as the western Rockall
Bank. The species are also taken by trawl
fisheries within the northeast Atlantic, mainly
as by-catch. Catch per unit effort (CPUE) data
from the fishery indicate a decline in catches
from the 1970s to 1990s and stocks remain at
a reduced level (ICES, 2008). There has been
limited provision of data from some of the major
fisheries for this species in the NEAFC region. At
present, length and age data are inadequate for
reliable age-based assessments of ling (WGDEEP,
2009) and there are no data on discards of this
species in the northeast Atlantic.
16 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
–
–
–
–
27°27’32
27°31’16
27°43’48
27°39’41
W
W
W
W
Blue ling has been important in by-catch from
mixed deep-water trawl fisheries on the Hatton
Bank (ICES, 2008). In other high seas areas,
such as in the Norwegian Sea and north of the
Azores, it is taken in small quantities. In the
Hatton Bank area, CPUEs have been variable
but a dramatic reduction in catches occurred
between 2002 and 2006. No catches were
reported from this region in 2007 and 2008.
The main states involved in this fishery were
Spain and France. Recently, results from the
EU POORFISH, a project aimed at developing
methods to provide advice on fisheries where
data are poor, together with other information,
enabled ICES to advise NEAFC and the EU on
the likely presence of spawning aggregations of
blue ling in the Hatton and Rockall Bank areas.
Blue ling (Molva dypterygia)
Fisheries for this species began around Iceland,
targeting spawning aggregations, but these were
depleted relatively quickly. The species is now
taken mainly as by-catch in redfish and other
fisheries, mostly within the Icelandic EEZ. Some
catches are from the high seas portion of the
Irminger Sea and Rockall and Hatton Banks.
In northern areas fishing has been mainly by
bottom trawls, although the species is now
also targeted by longlines. Landings from high
seas areas have been very variable and little
information is available on these (e.g. Spanish
landings in 2003). ICES advised that there
should be no directed fisheries towards blue ling
because the species is vulnerable to overfishing,
especially when spawning aggregations are
targeted (WGDEEP, 2009). It also advises that
areas where spawning aggregations are present
N
N
N
N
These studies indicated spawning grounds for
blue ling on the continental slope off western
Scotland and around the Hatton, Rockall and
Rosemary Banks (ICES, 2007a; Fig. 3). In EU
waters these spawning grounds have been
protected from fishing during the spawning
period by specific fisheries measures. As yet, no
specific measures on the basis of this advice
have come from NEAFC, despite the information
having been available since 2007/08. However,
the measures that extended the closures to
bottom trawling on Hatton Bank in 2007 and
2008 include at least part of the putative
spawning area for blue ling (NEAFC, 2006a,
2007a). The presence of a spawning ground is
viewed as a feature that identifies a geographic
area as a VME (FAO, 2009a: Para 42 [i]).
Fishing on spawning aggregations may result
in underestimation of stock trends and so any
such analyses for blue ling require cautious
interpretation (Arbuckle et al., 2008). Ageing is
difficult in this species, making stock analysis
for management purposes problematic (WGDEEP,
2009).
Tusk (Brosme brosme)
Genetic evidence suggests that it is likely
that tusk form several distinct stocks in the
northeast Atlantic region and that stocks from
the northwest Atlantic, Mid-Atlantic Ridge and
Rockall should be treated as distinct (ICES,
2008; WGDEEP, 2009). Where there is a
directed fishery, management is in place (e.g.
Icelandic EEZ). However, in other areas there
is no species-specific management (e.g. ICES
Areas I and II, including the high seas areas
of the Banana Hole and Barents Sea). On the
Mid-Atlantic Ridge tusk is taken as by-catch
Used with the kind permission of the International Council for the Exploration of the Sea
64°
63°
Known area
62°
Known area
61°
28°
26°
24°
22°
20°
Figure 2. Blue ling: Spawning areas in the Icelandic EEZ
(WGDEEP, 2009).
18°
16°
Figure 3. Blue ling: Spawning areas identified in ICES
areas XII, VIb, VIb and V (ICES, 2007a).
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 17
from gillnet and longline fisheries. There does
not seem to be any specific management of the
species in this region of the high seas and there
are no clear trends in available catch data. The
same is true for Rockall, where tusk is also a
by-catch species landed mainly by Norwegian
vessels. Length and age data are inadequate
for aged-based population analyses (WGDEEP,
2009). The species is considered Threatened in
the northwest Atlantic (COSEWIC, 2003).
Greater silver smelt (Argentina silus)
Silver smelt were taken as by-catch and
generally discarded up to 1996. Since 1997,
a directed bottom trawl fishery has existed
for this species, although it is also taken as
by-catch in the redfish fisheries in the region.
ICES advised that greater silver smelt is a lowproductivity species and that it can sustain
only low levels of exploitation (ICES, 2008). It
also shows aggregating behavior increasing its
vulnerability to overfishing (ICES, 2008). Catches
have been very variable for this species along
the continental slope west of the British Isles
and Ireland, in the Norwegian Sea and around
the Faroes and Iceland. In some cases, there
is evidence of overexploitation but in others
no evidence of significant decline is apparent;
catches seem to reflect market demand. Stock
structure in greater silver smelt has not been
resolved and there is an urgent requirement for
genetic studies to identify biologically relevant
management units. The greater silver smelt is
often caught as by-catch in other fisheries and
data on discards are lacking (ICES, 2008). Like
redfish, species of Argentina appear very similar
and confusion exists in regard to identification of
smelts in the region. There have been no recent
stock assessments.
Orange roughy (Hoplostethus atlanticus)
Orange roughy are recognised as a lowproductivity species that aggregate at
seamounts to spawn. Throughout the
northeastern Atlantic this species has been
targeted by fishing vessels when aggregations
have been located and, in almost all cases,
depletion of stocks has occurred (WGDEEP,
2009). ICES recommends that no fisheries
be directed at this species as a result of
its vulnerability (ICES, 2008). There are no
stock assessments for orange roughy, but
stock status is based on CPUEs that fluctuate
strongly, perhaps as a result of fishing being
targeted at spawning aggregations. It is likely
that management should be directed at the
level of individual aggregations around specific
topographic features but spatial fisheries
statistics are not available to enable such an
approach. Despite continued advice that there
should be no fisheries directed at orange roughy,
NEAFC agreed both in 2008 and 2009 to allow
a TAC of 150t for each Contracting Party (a
maximum of 750t) with no directed fishing in
ICES Sub-Areas V, VI and VII (NEAFC, 2010a).
Roundnose grenadier (Coryphaenoides
rupestris)
In 2001 ICES ranked roundnose grenadier high
on a scale of vulnerability for deep-sea species,
lying below orange roughy (ICES, 2001). The
females mature at age 9–11 years and produce
a small number of large eggs (Kelly et al.,
1996), with large individuals making the largest
contribution to egg production (Alekseyev et al.,
1992). This species is subject to a directed
trawl fishery west of the British Isles including
around the Rockall and Hatton Banks and
has been fished in the Skagerrak and on the
northern Mid-Atlantic Ridge. The Ridge fishery
was closed to bottom trawling and static gear
in 2007 (Shibanov & Vinnichenko, 2008). In
the western part of the northeast Atlantic it is
suspected that landings in international waters
are unreported, especially in areas to the west
of the Hatton Bank where Spanish vessels have
been operating. This is a concern as catches
in this area have been high. In almost all these
areas there have been indications of declining
stocks with major reductions in catches
(WGDEEP, 2009). Virtual population analysis
of stocks, to the west of the UK and around
the Faroes, indicates a significant decline in
roundnose grenadier (WGDEEP, 2009). Statistics
from the area to the west of the Hatton Bank
were not included in this analysis because
of their unreliability. It is also reported that
numbers of large fish are declining, a concern
because of the large contribution to spawning
capacity in this species (Arbuckle et al., 2008;
see also Kelly et al., 1997, for the Rockall
Trough).
Fisheries on the Mid-Atlantic Ridge for
roundnose grenadier commenced in the 1970s
and were targeted at seamounts, using both
pelagic and bottom trawls. From a peak catch of
29,900t in 1975, the fishery has since declined
with the fall of the Soviet Union, although
sporadic fishing by various nations has taken
place since then. Roundnose grenadier is also
taken as by-catch in orange roughy and blue ling
fisheries in this area. There is insufficient data
on current exploitation of this species on the
18 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Mid-Atlantic Ridge, which in 2005 amounted to
about 2,000t (Shibanov & Vinnichenko, 2008).
ICES identifies roundnose grenadier as a lowproductivity species and has recommended
restricted fishing with no further fisheries to
be developed until it is shown that they are
sustainable. NEAFC has initiated regulations to
reduce effort in deep-water fisheries although
these have been ineffective (see below); the
EU has established species-specific TACs for
this species in the northeastern Atlantic for its
vessels; NAFO has banned a directed fishery.
There are age-based stock assessments for this
species for some areas. Roundnose grenadier is
considered Endangered in the northwest Atlantic
(COSEWIC, 2008).
Black scabbardfish (Aphanopus carbo)
In the northern parts of the northeast Atlantic
this species has been targeted along the
continental slope and off the Rockall and Hatton
Banks by bottom trawl fisheries. Further south,
in areas such as the Azores, it is targeted by
longline fisheries. In most areas, in the northern
part of the northeast Atlantic, stocks have
shown significant declines in CPUE (to ~20
percent of original CPUE estimates; WGDEEP,
2009). ICES has recommended that catches be
restricted to 50 percent of the level prior to the
fishery expansion in 1992/93 in the northern
area and that no further fisheries be developed
unless they can be proven sustainable (ICES,
2008).
It is suspected that this species undergoes
significant ontogenetic migrations over
considerable distances but further information
is required on the reproductive and population
biology of the species before this can be
confirmed.
Goldeneye perch (Beryx splendens and Beryx
decadactylus)
These species are generally taken as by-catch in
the northeast Atlantic region, particularly along
the Mid-Atlantic Ridge, on the high seas to the
north of the Azores EEZ. General trends are
for a decline in catches. No assessments are
available but there is concern about sequential
depletion of stocks and under-reporting of
catches from high seas areas (WGDEEP, 2009).
ICES has identified these species as being
highly vulnerable to trawl fisheries as a result of
their aggregating behaviour around seamounts,
the result of which may be that Beryx can only
sustain low levels of exploitation. There are
insufficient data for assessment of the status
of populations. The stock structure of this
species is poorly understood and aspects of
reproductive biology are poorly known. It has
been recommended that no new stocks of Beryx
spp. are exploited prior to assessments being
undertaken to determine sustainable levels of
fishing (ICES, 2008).
Blackspot sea bream (Pagellus bogaraveo)
The species has been fished on the continental
slope off the UK, France, Portugal and Spain, as
well as the Azores. Stocks along the northern
European continental shelf collapsed in the
1980s following overfishing over at least
10 years when the fishery was unmanaged.
Fisheries continue in other areas. It is
speculated that the life history of this species (a
protandrous hermaphrodite) makes it vulnerable
to overfishing as all large fish are female. There
is evidence of population differentiation between
the European continental slope/shelf and the
Mid-Atlantic Ridge (Stockley et al., 2005).
Greater forkbeard (Phycis blennoides)
This species is mainly taken as by-catch in
bottom trawl and longline fisheries throughout
the northeast Atlantic region, along the
European continental slope, offshore banks
and oceanic islands and the Mid-Atlantic Ridge.
Trends in catches are unclear and vary markedly
from area to area. In general, data on catches of
this species are not reliable as it is a by-catch
species and not always recorded and is also
confused in landing statistics with other species
(hakes and morids), which it resembles.
Deep-sea sharks
Sharks are low-productivity species as a result
of life histories exhibiting low fecundity, slow
rates of growth and a long time to maturity
(ICES WGEF, 2007; Gibson et al., 2008). In the
northeast Atlantic region, sharks are captured
along the European continental shelf, including
the Rockall and Hatton Banks, and on the
Mid-Atlantic Ridge. They have been targeted
directly by gillnet and longline fisheries, make
up an important component of mixed deepwater species fisheries (Hareide et al., 2004)
and are taken as by-catch, especially in hake
and monkfish fisheries but also others. At
present, TACs are set by the EU for ‘deepwater sharks’ including Portuguese dogfish
(Centroscymnus coelolepis), leafscale gulper
shark (Centrophorus squamosus), birdbeak
dogfish (Deania calceus), kitefin shark (Dalatias
licha), greater lanternshark (Etmopterus
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 19
princeps), velvet belly (Etmopterus spinax), black
dogfish (Centroscyllium fabricii), gulper shark
(Centrophorus granulosus), blackmouth dogfish
(Galeus melastomus), mouse catshark (Galeus
murinus), Iceland catshark (Apristurus spp.),
rough longnose dogfish (Deania hystricosum)
and arrowhead dogfish (Deania profundorum).
For the majority of these species, fisheries
are essentially unmanaged and reporting on
catches, by-catches and discards is unreliable.
This is exacerbated by confusion over the
identification of shark species or the placing of
many species into a single category of ‘deepwater sharks’ (ICES, 2008).
Siki sharks (Centroscymnus coelolepis and
Centrophorus spp.)
These sharks are widely distributed in the
northeast Atlantic but many aspects of their
biology are poorly understood. Fisheries
directed towards these sharks commenced
in the late 1980s, while in the last decade
CPUEs have declined substantially, indicating
that stocks are depleted (ICES WGEF, 2007).
In 2006 ICES recommended that no fisheries
should be targeted at these species unless
sufficient information was available to determine
sustainable levels of exploitation, and that
efforts to limit by-catch of these species
should also be taken. All three appear on the
IUCN Red List as being at risk of extinction:
Near Threatened – Centroscymnus coelolepis
(Endangered in the northeast Atlantic);
Vulnerable – Centrophorus granulosus
(Critically Endangered in the northeast Atlantic);
Centrophorus squamosus (Endangered in the
northeast Atlantic).
In the northeast Atlantic, gillnets have now been
banned from waters deeper than 200m on the
high seas and in the areas around the Azores,
Madeira and Canary Islands, and deeper than
600m elsewhere.
Other deep-water species
A variety of other deep-sea fish species are
fished in the northeast Atlantic, including high
seas areas such as the Hatton Bank. These
include roughhead grenadier (Macrourus
berglax), common mora (Mora moro) and other
Moridae, rabbit fish (Chimaera monstrosa
and Hydrolagus spp.), Baird’s smoothhead
(Alepocephalus bairdii), Risso’s smoothhead
(A. rostratus), wreckfish (Polyprion americanus),
bluemouth (Helicolenus dactylopterus), silver
scabbard fish (Lepidopus caudatus), deep-water
cardinalfish (Epigonus telescopus) and deepwater red crab (Chaceon affinis). Some of these
species appear to be showing marked declines
in catch. All of these species are now listed as
Regulated Species by NEAFC (NEAFC, 2009b)
but there appears to be no specific management
in place for them on the high seas other than
the requirement for a general reduction in
effort for deep-sea fisheries. Furthermore,
because data on catches are likely to be
unreliable as a result of misreporting or underreporting of catches (Arbuckle et al. 2008), and
because some species are discarded, realistic
assessments of population status and trends
are not feasible under the present management
regime. Catches of Alepocephalus bairdii, in
particular, increased markedly in recent years
(NEAFC, 2004, 2006b, 2007b), perhaps a
reflection of increased effort in the deeper
waters where this species occurs. The latest
figures indicate a major decrease in catches of
this species in the NEAFC Regulatory Area from
2007 to 2008 (NEAFC, 2010b).
NEAFC has made several requests to ICES with
respect to improving information on the spatial
and temporal patterns of deep-sea fishing in its
area of competence. ICES reported (WGDEEP,
2009) that the Vessel Monitoring System (VMS)
and reported catch data provided by NEAFC
was insufficient for these purposes. This was
because of the low frequency of VMS reporting
on vessel positions (once every two hours,
although it has recently increased in frequency
to every hour; NEAFC, 2010a) and because
fishing operations were not generally logged to
tally with VMS data. ICES also judged it likely
that there were significant amounts of missing
data or misreporting of catches. In addition, 70
percent of vessels reporting demersal catches
only reported catches of a single species. This
is highly unlikely given that demersal deep-sea
fish communities are of relatively high diversity
and thus almost all deep-sea fisheries are de
facto multispecies fisheries. The conclusion
must be that only the target species or most
abundant species in catches are being reported
and that data on total catches are incomplete
or misreported. For similar reasons, ICES was
unable to help NEAFC classify deep-sea fishing
activities into management categories (e.g.
targeted fishery, by-catch fishery, etc.; WGDEEP,
2009).
The review of management of fisheries by NEAFC
(Arbuckle et al., 2008) identified that deep-sea
fisheries in the area were not subject to longterm management objectives and therefore
long-term management plans were not in place.
Some unilateral and multilateral agreements
had been initiated for some species/fisheries.
It was also concluded that the status of many
of the deep-sea species targeted in the NEAFC
Regulatory Area was unknown. The review panel
summed up its concerns:
“The Panel nonetheless considers the situation
for deep-sea species to be inadequate, in
particular as regards knowledge of the species,
nature of the fisheries, status of the resources
and management planning. This is a critical
issue that NEAFC needs to address as a priority
and every effort should be made to develop
the necessary fisheries database to begin this
process.” (Arbuckle et al., 2008).
In 2004 NEAFC established a cap on fishing
effort (no more than the highest level in previous
years) for deep-sea species in its Regulatory
Area – the first measure to regulate fisheries
for deep-sea species on the high seas of the
northeast Atlantic. In 2006, NEAFC Contracting
Parties agreed to further reduce fishing effort
by 35 percent in fisheries for deep-sea species
(Bjorndal, 2009). Over the duration of these
regulations, the reported catch of deep-sea
species in bottom fisheries in the NEAFC area
has varied several fold but has risen from just
over 25,000t in 2004 to more than 45,000t
in 2008, with 2007 representing the year of
highest catches at over 90,000t (Table 1).
EU fleets are responsible for 95 percent of
the reported catch of deep-sea species in
the NEAFC area. While these catches include
species that are of high or medium productivity,
catches of species around which there is
considerable concern, including blue ling, sharks
and argentines, have increased dramatically
(catches of orange roughy declined). Information
on whether these catches are from high seas
areas or not is difficult to ascertain, as are the
accuracy of the catch data presented, but it is
certain that there has been a dramatic increase
in reported catches of deep-sea species over the
last five years for which statistics are available
in the NEAFC Regulatory Area.
The NEAFC Commission agreed to maintain the
35 percent reduction in fishing effort on deepsea species until 2012 (NEAFC, 2010a), despite
the fact that this measure has not prevented a
large increase in catches of deep-sea species.
The ineffectiveness of these measures may
reflect that the estimates of fishing effort used
(aggregate power, aggregate tonnage, fishing
days at sea, aggregate number of vessels;
NEAFC, 2010a) do not reflect the killing power
of the deep-water fishing fleets in the region.
Changes in technology within a fishing fleet
over time are likely to increase the efficiency
of fishing vessels and may result in increased
fishing mortality even if vessel numbers,
tonnage or other fleet-related estimates of effort
remain stable. Variation in deep-sea catches
over the last five years may also reflect changes
in the status of fish stocks, changing patterns
of fishing (changes in geographic areas or in
targeted species) or problems in reporting of
deep-sea catches. The NEAFC Commission also
agreed to ban discards in high seas fisheries,
although the ban only applies to Annex IA
species (redfish, herring, blue whiting, mackerel
and haddock) and not to Annex IB species,
which include most deep-sea species.
Country
2004
2005
2006
2007
2008
Total 2004–2008
EC
25,157
69,883
51,346
90,554
42,471
279,681
Faroe Islands
642
756
253
202
261
2,114
Greenland
Iceland
Norway
Russia
0
0
648
56
0
0
620
2,188
1,913
0
963
148
2,391
0
933
366
1,415
0
275
362
5,719
0
3,439
3,120
26,503
73,447
54,623
94,446
45,054
294,073
Total
Table 1. High seas catches of deep-sea species in the NEAFC Regulatory Area 2004-2008 (tonnes). Data extracted from NEAFC
catch statistics.
20 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 21
Rockall Bank
Following evidence presented by the ICES
Working Group on Deep-Water Ecology (WGDEC),
NEAFC closed a number of areas on the Rockall
Bank, including northwest Rockall, The Logachev
Mounds and West Rockall Mounds (Fig. 5). In
addition to this, an area known as the Haddock
Box remained closed to trawling to protect
haddock stocks on the Rockall Bank (WGDEC,
2007). The EU closed the portions of these
areas lying within the EEZs of Member States.
Figure 4. Global distribution
Protection of benthic marine ecosystems
Cold-water coral reefs
To determine whether impacts from existing
deep-sea fisheries in the NEAFC area have
impacted, or have the potential to impact, on
benthic ecosystems, requires the overlaying of
geo-referenced fisheries data onto maps of the
known occurrence of VMEs. In the case of the
northeast Atlantic, the following VMEs are known
to occur in the region beyond areas of national
jurisdiction (WGDEC, 2009):
● cold-water coral reefs;
● other coral-associated benthic habitats
(octocoral meadows or forests, dense stands
of Antipatharia or Stylasterida);
● dense stands of sea pens;
● sponge-associated habitats;
● areas of dense occurrence of xenophyophores;
● dense stands of cerianthid sea anemones;
● serpulid reefs (Filograna);
● deep-water oyster reefs (Wisshak et al., 2009).
Cold-water coral reefs have been studied
more thoroughly in the northeast Atlantic than
anywhere else in the world. In this region, reefs
are mainly formed by the framework-building
coral Lophelia pertusa, with contributions from
other species, mainly Madrepora oculata but
also Solenosmilia variabilis, Desmophyllum
dianthus and Dendrophyllia cornigera (Rogers,
1999; Roberts et al., 2009). For Lophelia
pertusa, the northeast Atlantic is the most
important known area in terms of the number
of observations of the species, especially as
a component of deep-water coral reefs. The
coral occurs on the shelf-break and upper
continental slope of Norway, and continental
Europe, including offshore banks such as the
Faroes, Rosemary, Lousy, Hatton and Rockall
Banks. Elsewhere, the coral occurs along the
Mid-Atlantic Ridge and on the continental slope
of West Africa, in the Mediterranean, along the
continental slope in the northwest Atlantic, in
the southern Atlantic, Indian and North Pacific
Oceans (Rogers, 1999; Davies et al., 2008;
Fig. 4).
of Lophelia pertusa (Davies
et al., 2007).
These systems are especially associated with
areas of elevated or steep topography; the
occurrence of strong bottom currents, especially
those generated by small-scale oceanographic
phenomena such as internal waves; a high
concentration of food; and the occurrence of
hard substrata (e.g. Rogers et al., in press).
22 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Subsequently, further observations indicated
the presence of Lophelia pertusa reefs on the
southwestern Rockall Bank and on the Empress
of Britain Bank (Fig. 6). In addition, significant
areas of Lophelia pertusa reefs were also
detected on the northeastern part of the Rockall
Bank and just outside of the northwestern
Rockall Bank protected area (Fig. 7). ICES
subsequently recommended that all these areas
be closed to bottom fishing. The adjustments
to the northwestern Rockall protected area and
the adoption of the Empress of Britain protected
area were accepted by NEAFC and remain in
place at the present time (NEAFC, 2010a),
while the eastern proposed area is under
consideration (NEAFC, 2007a).
Figure 5. Rockall Bank,
showing closures
to fishing in 2007
(WGDEC, 2007).
The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 23
Hatton Bank
In 2007 NEAFC also closed a portion of the
Hatton Bank because of evidence of the
presence of Lophelia pertusa reefs presented
by ICES (WGDEC, 2005), following a proposal
to close a portion of the bank from Norway. In
2008, this was extended following evidence
gathered by UK and Spanish researchers
indicating areas of Lophelia pertusa reef
to the south of the 2007 closure (WGDEC,
2007; NEAFC, 2007a; Fig. 8). Subsequently,
Spanish scientists obtained evidence of further
occurrences of cold-water coral reefs in three
areas to the west of the existing closure, all
associated with rock outcrops forming ridges
or more irregular topographic features lying
between 700 – >1,700m deep (Fig. 9). Coral
mounds are present in this area and live
Scleractinia (Lophelia pertusa, Solenosmilia
variabilis, Madrepora oculata), Antipatharia and
Octocorallia have been sampled (Munõz et al.,
2008). In 2009, NEAFC agreed to extend the
Hatton Bank protected area in line with ICES
recommendations (NEAFC press release, 2009;
NEAFC, 2010a).
Figure 6. ICES recommendation for the protection of the Empress of
Figure 7. ICES recommendations for adjustment of northwestern
Britain Bank as adopted in NEAFC Recommendation IX 2008 (NEAFC,
Rockall Bank protected area (adopted in 2010) and for a new eastern
2007a).
Rockall Bank protected area.
Mid-Atlantic Ridge
In 2004 NEAFC closed five areas in the
Mid-Atlantic Ridge region to fishing as a
precautionary and interim measure to protect
benthic marine ecosystems. These regions
were chosen to provide a variety of ridge and
seamount habitats placed at different latitudes
along and to either side of the Mid-Atlantic
Ridge. Recent synthesis of current knowledge
on the biogeographic classification of the
deep ocean indicates that the sites selected
represent areas lying in different water masses
along the Mid-Atlantic Ridge (UNESCO, 2009)
and may also represent different bathyal
provinces, and so are likely to exhibit some
general differences in fauna beyond those that
would reflect only variation in the local physical
environment. Beyond this aim of biogeographical
representivity, the choice of sites did not take
into account the (albeit) limited information on
the presence of VMEs at the time (Norwegian
government, 2008). These sites were Hecate
Seamount, a complex of seamounts around
Faraday Seamount, an area of the Reykjanes
Ridge and the Altair and Antialtair Seamounts
to the south (Norwegian government, 2008).
With the exception of the Reykjanes Ridge,
these protected areas were all small given the
conservation objectives for which they were
originally set up (representivity of deep-water
Figure 9. ICES’ proposed extension to Hatton Bank closed areas on
Figure 8. Hatton Bank showing the area protected in 2007 and that
the basis of Spanish data indicating the presence of cold-water coral
proposed for protection by WGDEC (2007) and subsequently protected
habitats in three areas to the west of the existing protected area
Figure 10. Proposed Mid-Atlantic Ridge and seamount protected areas;
by NEAFC until late 2009 (NEAFC, 2007a).
(WGDEC, 2008).
adopted by NEAFC in 2009 (Norwegian government, 2008).
24 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
ecosystems along the Mid-Atlantic Ridge).
On the basis of new scientific observations,
the Norwegian government proposed five new
protected areas along the Mid-Atlantic Ridge,
encompassing some of the existing protected
areas but, because of their larger size, including
a wider depth and range of habitats (Fig. 10).
These sites were chosen to represent colder
northern regions of the Mid-Atlantic Ridge
(Reykjanes Ridge); the zone of the Sub-Arctic
Front around the Charlie Gibbs Fracture Zone,
an area of high biological productivity; and an
area in the southern section of the Mid-Atlantic
Ridge representing the warmer areas just to the
north of the Azores EEZ. The Altair and Antialtair
Seamounts were retained but areas were
expanded to include the seamount flanks.
It was already known that there were a number
of observations of the occurrence of Lophelia
pertusa on the Reykjanes Ridge (e.g. WGDEC,
2006), and further evidence of the presence
of colonies of these corals further south along
the ridge were presented by Mortensen et
al. (2008), based on ROV footage from MarEco, a Census of Marine Life project aimed
at understanding the distribution and ecology
of the animal communities of the Mid-Atlantic
Ridge. Along the Mid-Atlantic Ridge Lophelia
pertusa was only observed as relatively small
colonies, up to 50cm in diameter, although in
places large areas of dead coral skeleton were
also observed. Whether corals had died naturally
or these observations were a result of past
fisheries impacts (deep-water trawl fisheries
have existed on the Mid-Atlantic Ridge since
the 1960s and 70s; Shibanov & Vinnichenko,
2008) is unknown, although it is possible that
small colonies represent recolonisation following
damage from fishing (Mortensen et al., 2008).
Various octocorals were also observed during
the Mar-Eco Mid-Atlantic Ridge investigations
and areas with corals supported a higher
abundance of other megafauna (Mortensen et
al., 2008). Despite existing data, knowledge of
the location of VMEs on the Mid-Atlantic Ridge
is sparse. In 2009 NEAFC agreed to establish a
seasonal closure of an area on the northern MidAtlantic Ridge to protect a spawning aggregation
of blue ling (NEAFC, 2010a; see above).
ICES reported that following the closures in
2004, fishing effort actually increased on the
Faraday, Altair and Antialtair Seamounts (ICES,
2007b). It is not known if NEAFC traced the VMS
signatures that were likely to have been fishing
in the protected areas.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 25
The Wyville-Thomson Ridge
This area is within the European EEZ. In 2003
the area to the south of the Wyville-Thomson
Ridge was closed to fishing as a result of the
discovery of the Darwin Mounds, an area of
low-relief submarine hills with coral growing on
their summits (Davies et al., 2007; Fig. 11). This
area had been impacted by fishing, especially
the French roundnose grenadier fishery in the
Rockall Trough (Wheeler et al., 2005). ICES
has recommended that the Wyville-Thomson
Ridge itself be closed to fishing because of the
presence of Lophelia pertusa (WGDEC, 2009).
Figure 11. Wyville-Thomson Ridge: Protected areas and
proposed protected areas (WGDEC, 2009).
Deep-sea sponges
Sponge grounds occur in the North Atlantic at
depths of between 200 and 1,500m. In the
northeast Atlantic, they comprise two distinct
types of aggregations: those formed by glass
sponges (Hexactinellida) and those formed by
silaceous sponges (Demospongiae). Glasssponge beds, formed mainly by the species
Pheronema carpenteri, are found on the upper
slope at 740–2,000m, depending on location
(WGDEC, 2007), from Iceland to West Africa,
with mass occurrences being reported west of
Scotland, in the Porcupine Seabight (Fig. 12a)
and off Morocco (Le Danois, 1948; Rice et al.,
1990; Barthel et al., 1996). These sponges
live on mud and generate spicule mats, which
are associated with increased biomass of
macrofauna (Bett & Rice, 1992). Demosponges
form dense grounds dominated by a few species
of massive sponge. These are referred to as
‘ostur’ (‘cheese’ in Faroese and Icelandic
because the sponges resemble cheese when
brought out of the water), reflecting the large
by-catches of these sponges taken by fishers
in northern waters. The dominant species
on these grounds include Geodia barreti,
Geodia macandrewi, Geodia mesotriaenia,
Geodia phlegraei, Stryphnus ponderosus
(now recognised as two species) and Thenea
muricata. There is evidence to suggest that
there is a distinct boreal ostur comprising
Geodia barreti, Geodia macandrewi, Stryphnus
ponderosus and other species, found around
the Faroes, Norway, Sweden, parts of the
Barents Sea and south of Iceland, and a coldwater ostur, dominated by Geodia mesotriaenia
and other species, found north of Iceland, in
the Denmark Strait, off east Greenland and
north of Spitzbergen (Klitgaard & Tendal, 2004;
Fig. 12b). Recent surveys have also indicated
that important sponge grounds may occur on
the UK continental slope north of the WyvilleThomson Ridge, while trawl surveys by Spanish
researchers have encountered high by-catches of
the sponges east of the Hatton Bank and in the
Hatton-Rockall Basin (WGDEC, 2007).
Figure 12a
(on left).
Pheronema
carpenteri
distribution in
the Porcupine
Seabight.
Figure 12b
(on the
right). Ostur
distribution in
the northeast
Atlantic.
26 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The sponges that form Ostur have 242 species
identified as associated with them, including
some obligate associates (Klitgaard, 1991,
1995; Warén & Klitgaard, 1991). Knowledge
of these communities is restricted to a very
few studies. Sponge grounds represent
important benthic habitats in the deep waters
of the northeastern Atlantic. They have been
demonstrated to be associated with increased
biomass of associated fauna and so may be
viewed as structural species within distinct
communities. Sponges are especially vulnerable
to bottom fishing gear because of their upright
structure (Freese et al., 1999), the fragile
nature of tissues and skeletons (especially
glass sponges) and susceptibility to smothering
with sediment (WGDEC, 2009). Bottom trawls
in particular can take enormous by-catches of
sponges, with up to 50t in a haul reported from
areas south of Iceland, 12t per tow in areas
of the Norwegian Shelf and 1–3t per tow off
the Faroes. Experimental trawling on sponge
grounds has also shown that 30–60 percent of
colonies left on the seabed can be damaged
(Freese, 2001). Survival of damaged colonies
depends on the extent of damage (Henry & Hart,
2005). Wounded sponges may become infected
by necrotising bacteria and subsequently die
(Freese, 2001). Little is known about the growth
rates of sponges in deep water. In shallow
waters, growth rates of 0.76 to 5.6cm per year
have been estimated, with colonies living for
up to 220 years (Leys & Lauzon, 1998). It is
likely that deep-water sponges, living at the
depths associated with glass and silaceous
sponge grounds in the northeast Atlantic, grow
at much slower rates. Some Canadian sponge
reefs, located on the deep shelf, have existed
at the same locality for up to 9,000 years.
Thus, sponge grounds fit definitions of VMEs
in that they are important as diverse benthic
communities, are vulnerable to trawl damage,
and have a very low resilience to fishing
impacts.
To date, there has not been a systematic
evaluation of interactions between fisheries
and sponge grounds in the NEAFC Regulatory
Area. Localities where it is likely that there are
significant impacts on sponge grounds by deepwater bottom fisheries on the high seas are
on the northern Mid-Atlantic Ridge (Reykjanes
Ridge) and the eastern Hatton Bank and the
Hatton-Rockall Basin. No action has been
taken to study sponge/fishery interactions or
to close any areas to fishing on the basis of
the presence of sponge habitats. Some sponge
grounds may have been protected by existing
closures for Lophelia pertusa but this has not
been evaluated.
Coral gardens
Scleractinia, Octocorallia, Antipatharia and
Stylasterida may form dense stands associated
with many other species of invertebrates and
fish. These habitats have been termed coral
gardens, coral forests or coral meadows.
Defining these habitats is difficult but they
appear to exhibit a high density of corals
compared to the surrounding seabed and
are often associated with higher species
diversity or a distinct community of other
mega- and macrofaunal species. The functional
relationships between the coral stands and
associated fauna are often unclear (see below).
A number of studies have examined the density
of octocorals and other corals forming these
habitats and indicate that densities may vary
between 0.1 – >10 colonies per m2 and are
generally higher than background densities by
a factor of 10 (Orejas et al., 2002; Mortensen
& Buhl-Mortensen, 2004, 2005; Stone, 2006).
To some extent the density of corals in such
habitats reflects size, with larger species of
coral occurring at lower densities than smaller
species or mixed coral gardens with a variety of
taxa (WGDEC, 2007).
Rogers et al. (in press) suggest the following
definition for a coral garden:
“Coral communities formed by one or more
of the coral groups Scleractinia, Octocorallia,
Antipatharia or Stylasterida where the density
of colonies reaches >10 times background
densities, is usually >0.1 colonies per m2,
and where there is an enhanced diversity of
associated fauna or a distinct associated faunal
community compared to the background benthic,
epibenthic and epizoozoan fauna.”
Gardens of Scleractinia, Octocorallia,
Antipatharia and Stylasterida are associated
with distinct communities of animals but there is
less known about these communities than those
associated with cold-water coral reefs. In the
case of octocorals, analyses of just 25 colonies
of octocorals from the Atlantic coast of Canada
identified 114 species of associated animals
(Buhl-Mortensen & Mortensen, 2005). In Alaska
97 percent of juvenile rockfish and 96 percent of
juvenile golden king crab were associated with
emergent epifaunal invertebrate communities,
especially those formed by octocorals and
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 27
sponges (Stone, 2006). Identifying why such
associations occur is difficult. In many cases,
fish may use coral in a similar way to other
complex topography, such as rocks and boulders
on the seabed, for shelter and for foraging.
Other large predators may also use coral habitat
as foraging areas. The endangered Hawaiian
monk seal has been observed as foraging
preferentially for fish among beds of octocorals
and black corals off Hawaii (Parrish et al., 2002).
Coral garden communities are extremely
vulnerable to damage from bottom fishing gear
(Stone, 2006; Edinger et al., 2007a; Waller
et al., 2007; Heifetz et al., 2009). Species
are slow growing and some octocorals and
antipatharians are among the most long-lived
species on the planet (ages of >4,000 years
for Leiopathes spp.; Roark et al., 2009). Such
communities show a very low resilience to
fishing impacts, and observations indicate no
recovery of coral gardens decades after fishing
impacts (Waller et al., 2007). These habitats
therefore meet criteria for classification as
VMEs.
Compilation of records of deep-sea octocorals
indicate that they are distributed throughout
the world’s oceans, including areas such as
the Arctic and off continental Antarctica where
Scleractinia are relatively rare (Fig. 13). While
there is no clear pattern of distribution (see
Fig. 14), high levels of diversity have been
observed in the North Pacific, especially around
the Hawaiian Islands; the southwestern Pacific,
especially off New Zealand; the North Atlantic,
especially on the Mid-Atlantic Ridge; as well
as seamounts and on the flanks of oceanic
islands in the northeastern Atlantic; and to a
lesser extent in the Pacific and Atlantic sectors
of the Southern Ocean (Rogers et al., in press).
Records from the Indian Ocean are sparse.
On the basis of existing knowledge, the NEAFC
Regulatory Area may be considered a globally
important region in terms of abundance and
diversity of octocorals. Within the region,
knowledge of the distribution of coral garden
communities is extremely sparse. Most records
of octocorals come from the northern MidAtlantic Ridge; the Mid-Atlantic Ridge around
the Azores; non-ridge seamounts, such as
the Josephine Bank; along the continental
slope from Norway to West Africa; the slopes
of oceanic islands, including the Azores and
Canary Islands; and around Iceland and the
Faroes (Hall-Spencer et al., 2007; WGDEC,
2007; Rogers et al., in press; Fig. 15). The
distribution of stylasterids and antipatharians is
less well known but probably similar, albeit with
differences in depth distribution (see Rogers et
al., 2007).
At present there has been one study of the
interactions of deep-sea bottom fisheries
with octocorals/antipatharians in the NEAFC
Regulatory Area (Durán Muñoz et al., 2007),
despite the known occurrence of octocoral
gardens in the region (e.g. seamounts on
the high seas and around the Azores). Such
interactions are likely to take place in high seas
deep-water fisheries on the Mid-Atlantic Ridge
and on the Hatton Rockall Banks. Spanish
observer studies indicate limited by-catch from
bottom-fishing gear in the Hatton Bank area but
note that such organisms were rare on trawl
grounds (WGDEC, 2007; Fig. 16). NEAFC has
not undertaken assessment of such interactions
and its discussions to date have mainly been
concentrated on cold-water coral reef habitats
and sponge grounds.
Figure 13. Global
distribution of
records of deepsea octocorals
Figure 15. Distribution of non-reefal
corals in the North Atlantic region
based on ICES data (WGDEC, 2007).
(records >50m
deep). Note each
dot may represent
more than one
record (Rogers et
al., in press)..
Figure 14. Relative diversity of octocoral species in the world’s oceans (darker shades = higher diversity; Rogers et al., in press).
28 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 29
General considerations
ICES was requested to identify the current
temporal and spatial extent of deep-water
fisheries in the northeast Atlantic using VMS
data. In 2009, ICES WGDEEP advised that:
“The quality of the data is not yet sufficient to
provide information on the spatial and temporal
extent of current deep-water fisheries in the NE
Atlantic.”
Figure 16. Distribution of catches of (A) Octocorallia and (B) Antipatharia on the
Hatton and Rockall Banks from bottom trawls (black symbols) and longlines
(red symbols) based on Spanish observer study (WGDEC, 2007).
The reasons for this advice were that there was
high interannual variability in data, suggesting
that data were misreported or missing, and that
in many cases only one species was reported
from catches (70 percent of vessels; WGDEEP,
2009). This is highly unlikely in deep demersal
fisheries (Merrett & Haedrich, 1997) and so
it would seem that catches were misreported
or that a portion of catch was unreported.
Notwithstanding this, these analyses did reveal
some new information on where particular
fish species were being fished and therefore
the potential impacts on VMEs. ICES made a
number of recommendations to NEAFC regarding
VMS data, notably to increase the transmission
rate of VMS units and to note what fishing gear
was deployed in fishing operations (because of
the risk of confusing use of different types of
gear). In 2009, NEAFC agreed to increase the
transmission rate of VMS units from once every
two hours to once every hour (NEAFC, 2010a).
A similar measure has been adopted by NAFO
(NAFO Fisheries Commission, 2009).
NEAFC also requested that ICES assist it in
developing a system for categorising fisheries
using VMS and catch data. However, ICES
reported that only 27 percent of vessels for
which VMS data were available had reported
catch data associated with individual vessels.
This severely limited the possibilities of providing
the requested work.
There has been no consideration of interactions
of fisheries with other potential VMEs in the
NEAFC Regulatory Area.
To date, there have been no impact
assessments of fisheries on VMEs in the
NEAFC Regulatory Area in accordance with
recommendations of the FAO Guidelines for
management of deep-sea fisheries on the high
seas. NEAFC has now identified areas that are
categorised as new/exploratory fishing areas
in the NEAFC Regulatory Area, which include
much of the Mid-Atlantic Ridge and seamount
areas in the high seas of the northeast Atlantic
(NEAFC press release, 2009). According to
NEAFC Recommendation XVI 2008, as from
January 1 2009 all bottom fisheries in areas
that are considered as new/exploratory areas
are subject to impact assessments that are to
be reviewed by PECMAS prior to permission to
fish being given. As the areas to be identified as
new/exploratory fishing areas were not identified
until late October 2008 (PECMAS, 2008a),
assessments were not forthcoming for 2009. A
document confirming the geographic areas (5 by
10 minute boxes of latitude/longitude) classified
as exploratory fishing areas and those classified
as being fished has been published (PECMAS,
2008b). However, no impact assessments were
reported to have been assessed by PECMAS in
2009 (PECMAS, 2009).
Move-on rules
Figure 17. Coral garden habitat from the slope of Faial Island, Azores, depth ~350m. Species include: Acanthogorgia hirsuta, Viminella flagellum, and
Narella sp. Note the lost longline along the top of the rock outcrop (© A.D. Rogers and the Rebikoff Foundation).
30 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The current NEAFC move-on rule is triggered
by a catch of >60kg of live coral (Lophelia
pertusa, antipatharians, gorgonians, cerianthid
sea anemones or sea pens) or >800kg of
live sponge (PECMAS, 2009; NEAFC, 2010a).
These threshold levels were modified following
the adoption of lower VME thresholds by NAFO,
which were reduced following scientific studies
and were based on a simple extrapolation of
threshold values estimated from scientific trawls
of 30 minutes duration (PECMAS, 2009; see
below).
These trigger levels apply equally to a trawl tow
or gillnet or longline set (NEAFC, 2008: Para.
2.2). When an encounter takes place, it is
reported to the flag state and/or Secretary (of
NEAFC) and the vessel moves on 2nm from the
best-guess encounter position. Each year these
encounter reports are reviewed by PECMAS and
ICES and a decision is made on whether the
accumulated evidence from encounters indicates
the presence of a VME (NEAFC, 2008: Para.
2.2). For new fishing areas the encounter rules
are the same except that the 2nm zone around
the encounter position is automatically closed to
fishing and then the temporary closure examined
by PECMAS/ICES at the end of the year prior to
making a decision about maintaining or lifting the
closure (NEAFC, 2008: Para. 3.2).
The threshold values for coral and sponge
by-catch in the NEAFC Regulatory Area that
trigger the move-on rule are not supported by
any explicit assumptions of biomass-density
relationships that produce some critical
threshold for a VME nor any related assumptions
about catch efficiency in fishing gear. There
are significant differences in both the area
impacted and the catch efficiency of bottom
trawl gear, gillnets or longlines for corals and
sponges (WGDEC, 2006). Trigger levels for VME
encounters in the NEAFC rules do not reflect
these variations.
An additional issue related to the NEAFC
encounter rules is the actual quantity of by-catch
that triggers move-on. Limited studies indicate
that by-catch may be a very poor indicator
of seabed species composition and density.
Freese et al. (1999) quantified catch efficiency
of trawl-caught invertebrates by comparing
density estimates based on areas swept by
the trawl, with density estimates from seafloor
imagery at deep-water sites (206–274m depth)
off southeast Alaska. They found that nets
retained only a fraction of the organisms on the
seabed swept during tows and for some sessile
organisms, such as octocorals and sponges, no
quantifiable estimates of retention were made,
presumably because of the size and fragility of
species encountered. Light, flexible and fragile
specimens are either not retained by the net or
once in the net are fragmented and lost through
the meshes.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 31
Fisheries research surveys have recorded
by-catch of benthic invertebrates, especially in
the NAFO area off the coast of eastern Canada.
Spanish surveys in NAFO Division 3LMNO,
based on 30-minute tows at 3 knots (kts), did
encounter large by-catches of sponges of up
to 5t per tow. Thus, the NEAFC threshold level
for sponges would have triggered the VME
encounter move-on protocol in some instances
for the fishing gear deployed. However, few
encounters with live coral, including both
small and large octocorals, antipatharians and
scleractinians, would have triggered a move-on
from these surveys (maximum catch was 69kg;
Murillo et al., 2008; see also WGEAFM, 2008b).
Data from standardised research trawls from the
Canadian Department of Fisheries and Oceans
(DFO) surveys on the eastern coast of Canada
(15-minute tows or standardised to 23,391m2)
indicate that catches of up to 1,578.7kg of
coral per haul of sponges were taken (WGDEC,
2009). Only a few instances of coral by-catch
exceeding the threshold to trigger the move-on
rule have been recorded. One was in an area
east of the Hudson Strait, northwest Atlantic,
where no previous fishing had taken place. Here
by-catch of up to 500kg of large octocorals
(Primnoa resedaeformis and Paragorgia arborea)
were taken per tow in the Northern Shrimp
Survey (Edinger et al., 2007b). The other case
was in the Gulf of Alaska at 365m depth, where
1,000kg of Primnoa were removed in a single
trawl during a National Marine Fisheries Service
(NMFS) survey (Krieger, 2001). In both instances
massive octocoral colonies were the principal
by-catch.
These data cannot be extrapolated to
commercial trawls in terms of tow lengths and
times (mean of four hours in de Cárdenas et al.,
1997). It is not possible to simply extrapolate
catches from short duration tows to longer
duration commercial tows, as done by NAFO and
NEAFC (PECMAS, 2009), because sponge and
coral habitats have an aggregated distribution,
meaning that the relationship between tow
length, gear type and by-catch is not linear
(WGDEC, 2007; WGEAFM, 2008b). However,
a trigger level of 800kg of sponge for bottom
trawls is around one order of magnitude higher
than the level estimated for research trawls
of 15 to 30 minute duration in the northwest
Atlantic (WGEAFM, 2009). An 800kg threshold
would miss areas that Murillo et al. (2008)
considered as having high sponge by-catch.
Given the scientific analyses undertaken by
NAFO (WGEAFM, 2009; Kenchington et al.,
2009a), the current threshold for sponge
by-catch may fail to meet conservation
objectives. Equally, a trigger level established for
bottom trawling is not appropriate for passive
fishing gear such as gillnets or longlines as each
fishing method has varying impact.
For habitat-forming corals the situation is similar.
A 60kg trigger level for octocorals would miss
the majority of coral garden habitats formed
by large octocorals and probably 100 percent
of coral garden habitats formed by small
octocorals, antipatharians and stylasterids. For
corals a trigger level of less than 10kg might
be sufficient to detect VMEs if they are to be
treated as a single category of organisms (even
this is not appropriate for small habitat-forming
octocorals, antipatharians and stylasterids).
Such a weight would seem appropriate given
coral by-catch in stock assessment surveys
in areas such as the northwest Atlantic (e.g.
Edinger et al., 2007b). However, given the
analyses undertaken by scientists for NAFO,
even lower trigger levels would be appropriate
for smaller coral species or species with poor
retention in nets (WGEAFM, 2008b). This is
because the low catch efficiency and sheer
mass of fishing gear means many times the
landed by-catch may be left on the seabed,
destroyed or damaged. Thus, repeated trawling
events that do not trigger the move-on rule may
rapidly destroy a VME given the current threshold
levels set by NEAFC.
For cold-water coral reefs, often the reef
comprises a relatively thin layer of live coral
overlying a dead coral framework which may
form the greater part of the VME. The dead
coral framework is an integral part of the coldwater coral reef VME and is the main habitat
for many reef-associated species (e.g. see
Rogers, 1999; Freiwald et al., 2002). Any
differentiation between quantities of live and
dead coral in the context of the VME move-on
rule is therefore unjustified both scientifically
and for management purposes as they are both
important components of a cold water coral reef.
For trawls, the 2nm move-on rule for fished
areas has no conservation value. If mean tow
time is four hours (de Cárdenas et al., 1997)
and the usual speed of trawling is 3.5kts
(WGDEC, 2009), then a trawl will cover a mean
linear distance of 14km (up to 20nm is reported
for the NAFO Regulatory Area; WGEAFM, 2008b).
VMEs are aggregated in their distribution
and there is no way of ascertaining where a
32 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
VME is actually encountered during normal
fishing operations. Even for static gear such
as longlines, doubts have been raised about
identifying where VMEs occur from the position
of a longline set and the specific location of
VME taxa on the longline (segment and hooks;
Government of New Zealand, 2008a). Only
larger cold-water coral reefs may be detected
through irregularities in bottom topography on
fisheries echosounders but this would require a
constant watch to be maintained during trawling
operations and for the position of any seabed
mound recorded. Therefore, a vessel would have
to move 2nm away from the entire trawl track
for a move-on rule to be effective (see below).
Depth zonation of fauna introduces another
complication. A trawl or set across isobaths will
encounter particular species at a very different
rate than a trawl or set along isobaths.
NEAFC distinguishes between fished and
non-fished areas in the response to a VME
encounter. In a previously non-fished area (an
exploratory fishing ground), when a VME is
encountered the area around it is closed to a
diameter of 2nm. Only later, when the closure
is considered by PECMAS/ICES, can a decision
be made to maintain or lift such a closure.
This makes sense, for a VME when present is
protected immediately from fishing. However,
in an area with a history of fishing, although
the encountering vessel must move on 2nm
(as indicated above, this is unlikely to confer
protection to VMEs from trawling), the area
remains open to fishing until PECMAS/ICES
make a decision at the end of the year as to
whether or not the encounter represents a VME.
The VME encounter area may be trawled time
and time again by any vessel fishing in the area
until the decision to close or not to close is
made. A VME encounter carries the same weight
whether or not it occurs in an area with a fishing
history, and the regulatory response should be
the same. Indeed, it could be argued that a VME
in a fished area may be of greater conservation
value as it may represent important habitat
for target species of fisheries and a significant
proportion of the habitat may have already
been destroyed by fishing activities, thereby
increasing conservation and fisheries value of
the remaining habitat. Thus, discriminating in the
application of the move-on rule between fished
and non-fished areas by NEAFC has a negative
impact on protection of VMEs and is therefore
inconsistent with the UNGA resolutions (UNGA,
2004, 2007, 2009) and FAO Guidelines (FAO,
2009a).
Conclusions
(i) Conduct assessments of whether bottom
fishing activities have SAIs on VMEs.
● NEAFC now requires that impact assessments
are undertaken before bottom fishing is
permitted in exploratory fishing areas.
● To date no impact assessments have been
undertaken in the NEAFC Regulatory Area.
(ii) To implement measures in accordance with
the precautionary approach, ecosystems
approaches and international law and to
sustainably manage deep-sea fish stocks.
● Exploitation of deep-sea species within the
NEAFC Regulatory Area has led to depletion
of populations of several deep-sea species.
Deep-water sharks are now classed as
Endangered or Critically Endangered under
IUCN criteria as a result of targeted fishing
and/or by-catch.
● For the majority of deep-sea fish species,
whether targeted, taken as by-catch, or both,
there are no fishery-independent sources of
data for analyses of stock status or trends.
Reliance on fisheries (CPUE) data alone is
extremely difficult for deep-sea species in this
region because:
> for aggregating species, such data are of
limited value as CPUE values can remain
relatively constant and/or high in spite of
stock depletion;
> ICES and NEAFC suspect misreporting of
catches of deep-sea species in the NEAFC
Regulatory Area; and
> there are few data on by-catch/discards for
many of the deep-sea fisheries in the NEAFC
Regulatory Area.
● As a result of the above, the fisheries for many
deep-sea species in the NEAFC Regulatory
Area, including low-productivity species, are not
sustainably managed. Indeed, for many deepsea species management is impossible in view
of the lack of reliable data on catches and
an absence of scientific survey time series.
NEAFC countries agreed to reduce effort on
deep-sea species; despite this, catches of
deep-sea species increased from 2004–2008.
● Catches for deep-sea species have
exceeded or been in contravention of ICES
recommendations for several species,
including the important redfish fisheries in the
region.
● NEAFC has banned the use of gillnets below
200m depth on the high seas in its Regulatory
Area, which may contribute to the protection of
deep-sea shark species.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 33
● NEAFC agreed in 2009 to close one area for
the protection of a deep-sea fish species
(blue ling) during the spawning season;
the extension of protected areas on the
Hatton Bank, also agreed in 2009, has also
coincidentally protected part of a suspected
spawning area for blue ling.
● Other areas previously closed by NEAFC (e.g.
several closed areas along the Mid-Atlantic
Ridge and Rockall Bank) may have the effect
of reducing the catch of some deep-sea
species.
● Multispecies deep-water trawl fisheries along
the Mid-Atlantic Ridge and to the west of
Hatton Bank are in particular need of improved
data and management.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom
fishing where VMEs are known or likely to
occur, or not authorised to proceed.
● VMEs are present in the deep sea in the
NEAFC Regulatory Area beyond areas of
national jurisdiction. These include coldwater coral reefs, sponge grounds and
coral gardens. There are scattered data
on the presence of other potential VMEs
(e.g. occurrence of high densities of
xenophyophores or reefs formed by Filograna)
in the northeast Atlantic region.
● NEAFC has closed significant areas to fishing
because of the presence of VMEs, specifically
cold-water coral reefs.
● NEAFC has not closed all areas for which there
is strong evidence of the presence of VMEs,
specifically areas outside the current closures
on Rockall Bank.
● NEAFC has not closed areas to fishing
because of the presence of non-cold-water
coral reef VMEs specifically, but existing
closures may confer some degree of
protection to VMEs such as sponges and
coral gardens.
● Attempts to identify where deep-sea bottom
fisheries on the high seas are interacting with
and/or impacting on benthic communities that
constitute VMEs are hampered by a lack of
data on where fisheries are taking place in any
detail and especially at fine geographic scales.
(iv) To establish and implement protocols to
cease fishing where an encounter with VMEs
occurs during fishing activities and to report
such encounters so that appropriate measures
can be adopted with respect to that site.
● The threshold levels set by NEAFC for VME
encounters apply to sponges and corals only.
● The threshold levels for sponges and corals
may have little conservation value and
underestimate the occurrence of VMEs and
would only rarely result in any by-catch levels
triggering move-on in most areas of the North
Atlantic.
● Using the same threshold levels for active and
passive fishing gears does not reflect the large
differences in their impact.
● Using the same threshold levels for different
types of corals (and other VME taxa) is likely
to underestimate the occurrence of VMEs
formed by smaller octocorals, antipatharians
and stylasterids.
● Distinguishing by-catch of live and dead
Lophelia or any other coral is not based on
current knowledge of the structure of coldwater coral reefs, many of which rely on dead
structures for their make-up. Not counting
dead coral as a VME encounter will lead to
significant underestimates of the occurrence
of coral VMEs in any area.
● Differentiating the post-VME encounter
protocol between areas with a fishing history
and those without is inconsistent with
conservation objectives. A VME has the same
conservation value whether or not it is in an
area with a history of fishing.
● The 2nm move-on rule is ineffective from a
conservation perspective as it is impossible to
identify where a VME encounter occurs along a
tow for commercial bottom trawling.
NORTHWEST ATLANTIC OCEAN
The northwestern Atlantic Ocean ranks tenth
in importance in terms of capture fisheries,
producing about 2.2 million tonnes of fish in
2006 (FAO, 2009b). Fishing yields in this region
have been in decline since 2000 (FAO, 2009b).
The RFMO for the area is the Northwest Atlantic
Fisheries Organization (NAFO) (Fig. 18), which
replaced the International Commission for
the Northwest Atlantic Fisheries (ICNAF) in
1979. NAFO’s mandate extends to all fishery
resources within its Regulatory Area except
salmon, tuna, marlins and whales (Bensch et
al., 2008).
At present, regulatory measures (TACs or
quotas) are in place for 11 species or groups
of species, including cod (Gadus morhua),
redfish (Sebastes spp.), American plaice
(Hippoglossoides platessoides), yellowtail
flounder (Limanda feruginea), witch flounder
(Glyptocephalus cynoglossus), white hake
(Urophycis tenuis), capelin (Mallotus villosus),
skates (Rajidae), Greenland halibut (Reinhardtius
hippoglossoides), squid (Illex spp.) and shrimp
(Pandalus spp. and Penaeus spp.) (NAFO CEM,
2009). Several of these species are found in
deep water and are caught in the high seas
portions of the NAFO Regulatory Area, including
redfish, white hake, skates, shrimp and
Greenland halibut.
Many other species occur in the NAFO
Regulatory Area and are targeted or taken
as by-catch in high seas deep-sea fisheries,
including blue antimora (Antimora rostrata),
roughhead grenadier (Macrourus berglax),
roundnose grenadier (Coryphaenoides rupestris),
marlin spike grenadier (Nezumia bairdii), threebearded rockling (Gaidropsarus ensis), silver
rockling (Gaidropsarus argentatus), long fin hake
(Urophycis chesteri), striped wolffish (Anarhichas
lupus), spotted wolffish (Anarhichas minor),
northern wolffish (Anarhichas denticulatus), Arctic
eelpout (Lycodes reticulatus), Esmark’s eelpout
(Lycodes esmarki), spiny eel (Notacanthus
chemnitzii), alfonsino (Beryx splendens and Beryx
decadactylus), slickheads (Alepocephalus spp.),
black scabbardfish (Aphanopus carbo), wreckfish
(Polyprion americanus), black cardinalfish
(Epigonus telescopus), barrelfish (Hyperoglyphe
perciformis), Mediterranean roughy (Hoplostethus
mediterraneus), orange roughy (Hoplostethus
atlanticus), Cornish blackfish (Schedophilus
medusophagus), hagfish (Myxine glutinosa),
large-eyed rabbitfish (Hydrolagus mirabilis),
narrownose chimaera (Harriotta raleighana),
spiny dogfish (Squalus acanthias), black dogfish
(Centroscyllium fabricii), deep-sea catshark
(Apristurus profundorum), great lantern shark
(Etmopterus princeps), bluntnose sixgill shark
(Hexanchus griseus) and Portuguese dogfish
(Centroscymnus coelolepis) (Kulka et al., 2003;
Muñoz et al., 2005; Murua et al., 2005; Murua
& de Cárdenas, 2005; González-Troncoso et
al., 2006; Grant, 2006; Kulka, 2006; González
et al., 2007; Kulka et al., 2007a; Thompson
& Campanis, 2007). This list is by no means
inclusive of all targeted or by-catch species and
for a more extensive list of potentially vulnerable
deep-water species see WGEAFM (2008a). Most
of these species, however, are either so rare or
of such little value that they are discarded as
unwanted by-catch.
Management of fisheries for deep-sea
species of low productivity
Redfish (Sebastes spp.)
(Deep-sea redfish [Sebastes mentella]; golden
redfish [Sebastes marinus]; Acadian redfish
[Sebastes fasciatus]; beaked redfish [Sebastes
mentella and Sebastes fasciatus]).
Figure 18. The NAFO Regulatory Area.
34 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Because the species are very difficult to
distinguish and hybrids are known, all are
grouped together in the single statistical
category ‘redfish’ in the northwest Atlantic.
Redfish are long-lived and slow-growing species
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 35
and are generally viviparous with larval exclusion
occurring immediately before or after birth. For
these reasons the species are regarded as low
productivity or vulnerable species (WGEAFM,
2008a; see above). Management of redfish
stocks in the NAFO area is complicated by large
fluctuations in catches from year to year (NAFO
SC, 2008).
Redfish have been targeted in high seas areas
of the NAFO Regulatory Area, particularly in Area
3M, around the Flemish Cap. Initial catches in
this region rose from around 20,000t in 1985 to
81,000t in 1990, followed by a steep decline in
catches to around 1,000t in 1998/99 (Fig. 19).
As well as being subject to a directed fishery,
redfish were also being caught at this time as a
by catch in the shrimp fisheries in the region. As
stocks declined, fishing effort directed towards
these species also decreased but there was an
increase in catches after 2000 when Russian
and Portuguese fleets increased efforts to catch
redfish in the region. Catches have since risen
A. STACFIS ESTIMATES OF BEAKED
REDFISH COMMERCIAL CATCH
Thousand tons
75
60
45
30
15
0
350
1989 1991 1993 1995 1997 1999 2001 2003 2005 2007
Year
B. BEAKED REDFISH COMMERCIAL
CATCH AND BY-CATCH IN NUMBERS
Commercial
300
By-catch
Millions
250
200
150
100
50
0
1989 1991 1993 1995 1997 1999 2001 2003 2005 2007
to 6,500–8,500t in the last few years and it has
been suggested that this is a result of some
recovery in biomass of the targeted fish stocks,
a claim not wholly supported by scientific survey
data (Morin et al., 2004; Devine & Haedrich,
2010). These catches have been consistently
above recommendations by the Scientific
Committee for the recovery of the stock (e.g.
NAFO Fisheries Commission, 2005; NAFO SC,
2008). In 2009 NAFO increased the quota for
redfish in Area 3M to 8,500t (NAFO TACS, 2009)
despite a recommendation by the Scientific
Committee to maintain catches at 5,000t per
annum (NAFO Fisheries Commission, 2008). In
the last few years the Russian fleet has begun
to target pelagic redfish on the Flemish Cap.
As yet no data are available on this fishery,
and its impacts on the redfish stock within the
area as a whole are unclear (NAFO SC, 2008).
ICES provides advice on pelagic redfish for the
whole North Atlantic area, however, NEAFC has
consistently set TACs for this species several
times greater than scientific recommendations
(NAFO SC, 2008).
In Areas 3LN, which also include high seas
areas such as the Flemish Pass and the
southern tip of the Grand Banks, the fishing
history has been similar. Here, catches
increased sharply from 1985, rising from
21,000t to 79,000t in 1987, followed by a steep
decline to about 450t in 1996 (Fig. 19). During
this period catches were consistently above TAC
levels. In 1998 a moratorium was established
preventing directed fishing of Sebastes spp.
in this region. Despite this, by-catch of redfish
resulted in overall catches actually increasing
from 1998 to 2000 (NAFO SC, 2005). The
moratorium in Areas 3LN has since been lifted
and a precautionary TAC of 3,500t set (NAFO
SC, 2008), although this is allocated as a 10
percent limit on by-catch of other fisheries in
the area. The stock remains at a very low level
compared to its size prior to 1985.
Year
80
Figure 19. Beaked redfish
catches in (A) Area 3M
biomass, (B) Area 3M
numbers (Ávilo de Melo et
al., 2009) and (C) Areas
3LN (NAFO SC, 2008).
TAC/Catch (1,000 tons)
70
60
C. TAC / CATCH (1,000 TONS)
TAC
Catch
50
40
30
20
10
0
1955 1960 1965 1970 1975 1980 1985 1990 19950 2000 2005
Year
Redfish are also fished in Area 3O but there
was little specific information on the current
status of this stock and indeed the Scientific
Committee was unable to provide management
advice for this area because of lack of data
(NAFO SC, 2008; NAFO Fisheries Commission,
2009). Despite this, a TAC of 20,000t was set
by the Fisheries Commission in 2005 and has
remained at this level to the present day (NAFO
SC, 2008; NAFO Fisheries Commission, 2009).
Catches in recent years have varied but have
generally not reached the TAC.
36 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Roundnose grenadier (Coryphaenoides
rupestris)
The roundnose grenadier is long-lived, late
maturing and slow growing and therefore fits
the FAO criteria for a low-productivity species
(see above). Roundnose grenadier were first
exploited by Russian fleets in the late 1960s,
with catches peaking in the early 1970s at
around 80,000t and then declining rapidly.
The fishery developed largely in the absence
of any knowledge of the biology of the species
and by the 1990s had declined markedly
and effort switched to roughhead grenadier
(Macrourus berglax). Analyses of catches
of roundnose grenadier in research trawls
concluded that catches had decreased by 99.6
percent from 1978–2003 (Devine et al., 2006;
Supplementary Information), a decline fitting the
IUCN definition for Critically Endangered (Devine
et al., 2006). Subsequent study indicates that
the decline in roundnose grenadier populations
was a result of overfishing (Devine & Haedrich,
2008). Sufficient demographic data exist for
roundnose grenadier to calculate potential
recovery times. These could be as little as 16
years assuming an unlikely high rate of increase
(56 percent) and no fishing; with even a low
fishing rate (5 percent), recovery time rises to
136 years (Baker et al., 2009). Catch history
in recent years is complicated by evidence of
extensive misreporting of catches of roughhead
grenadier as roundnose grenadier (Murua et
al., 2005; González-Troncoso & Paz, 2007).
The species was taken mainly as by-catch in
Greenland halibut fisheries off Greenland (NAFO
Areas 0 and 1) and further south (Areas 2 and
3). Estimates of population biomass remain at
extremely low levels. Roundnose grenadier is not
under specific management measures by NAFO
although the Scientific Committee recommended
no directed fisheries towards the species off
Greenland and efforts to minimise by-catch
(NAFO SC, 2008). The species is considered
Endangered in the northwest Atlantic (COSEWIC,
2008).
Roughhead grenadier (Macrourus berglax)
The roughhead grenadier is becoming an
important commercial species in the northwest
Atlantic where it is taken as by-catch in the
Greenland halibut fishery, mainly in NAFO Areas
3LMN, much of which lies in the high seas
(Fossen et al., 2003; Costas & Murua, 2008).
As with roundnose grenadier, this is considered
to be a slow-growing species, late to mature and
long-lived (20 years +). The species matures
at 15 years of age but is recruited to fisheries
at eight years old (Devine & Haedrich, 2008).
As with the roundnose grenadier, it has been
estimated that populations of this species
have declined catastrophically in the northwest
Atlantic, with a decrease in catches in DFO
research trawls of 93.3 percent between 1978
and 2003. These data have been somewhat
controversial and indeed contradictory to recent
papers suggesting that populations are stable
or recovering, with increases in biomass in
recent years (e.g. Costas & Murua, 2008; but
subsequent communications found that negative
data from the region, i.e. tows with zero catch,
were not provided to these investigators).
Recovery times are calculated at as little as
19 years with no fishing and a 46 percent
annual rate of increase, but rise to 248 years
if even a low rate of fishing is allowed (Baker et
al., 2009). Recent studies have only covered
trends in populations since 1994, by which
time the northwest Atlantic stocks of M. berglax
had already decreased significantly, probably
as a result of by-catch and discarding in the
halibut fishery, and as a result of environmental
changes affecting population distribution (Devine
& Haedrich, 2008). Catches of M. berglax
peaked at 9,000t in 2000 but have declined
since, with catches at 3,000t and 1,500t in
2004 and 2007, respectively (Costas & Murua,
2008). It has been suggested that decreased
catches may have occurred because of efforts to
regulate the fishery for Greenland halibut in this
area (Costas & Murua, 2008). However, several
papers have identified extensive misreporting of
catches of roughhead grenadier as roundnose
grenadier, although the species are quite
distinct. As yet there is no management regime
in place for roughhead grenadier in the NAFO
Regulatory Area. The species is considered to
be of Special Concern (≈ IUCN Vulnerable) in the
northwest Atlantic (COSEWIC, 2007).
Blue antimora, blue hake or flat-nosed
codling (Antimora rostrata)
This is a widely distributed fish, occurring
at depths between 400 and 4,000m in the
northeast and northwest Atlantic, northern
Mid-Atlantic Ridge and elsewhere. In the NAFO
Regulatory Area, blue hake are caught within the
Canadian EEZ, but also on the eastern margins
of the Grand Banks, the Flemish Pass and
Flemish Cap (Kulka et al., 2003). Blue hake do
not concentrate at sufficient densities to warrant
directed fisheries but are taken as by-catch in
commercial trawl and longline fisheries in the
NAFO Regulatory Area (Kulka et al., 2003).
Devine et al. (2006; supplementary material)
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 37
report a decline in catches of blue hake of 92.7
percent in research trawls from 1978–1994.
There are few other data on blue hake in the
region. The resilience of the species is not
understood and it is not possible to evaluate
the current status of stocks in the northwest
Atlantic. There is no market for blue hake and all
by-catch is presumably discarded.
Wolffish or catfish: striped wolffish
(Anarhichas lupus), spotted wolffish
(Anarhichas minor), northern wolffish
(Anarhichas denticulatus),
Wolffish are a small family of very large
blennies, notable for their ferocious appearance
and disposition. They have very unusual and
conservative life histories that include internal
fertilisation and the production of large eggs
and larvae (2cm long in A. lupus), which are
brooded in nests under rocks by the males for
four to nine months. The level of parental care
is extraordinary; in the striped wolffish, the male
actively aerates and turns the eggs and covers
them in skin mucus to prevent infection (Kulka
et al., 2007b). They are long-lived (>20 years)
and tagging studies have indicated that they
are highly sedentary and apparently territorial.
Juvenile wolffish may settle close to their nest
sites but can also range quite far to establish
new home bases where suitable habitat is
found (natal behaviour; Fuller & Watling,
2008). Wolffish are apex predators and exert
considerable effects on ecosystem structure
and the composition of benthic communities by
feeding on urchins and other grazers (Fuller &
Watling, 2008). In all respects, wolffish fit the
FAO description of a low-productivity species,
mainly because of the life history of the species
(O’Dea & Haedrich, 2003; Kulka et al., 2007b).
The three Atlantic species reviewed here are
distributed in the North Atlantic only and are
found within national waters and in high seas
areas of both the NAFO and NEAFC Regulatory
Areas. The outer edges of the Grand Banks and
the Flemish Cap comprise significant areas of
habitat (Kulka et al., 2007b). The depth range is
considerable as they are found from 20–1,500m
depth; northern wolffish have the narrowest
range (Kulka et al., 2007b). These species
are found in a narrow range of environmental
temperatures (1.5–5oC; Kulka et al., 2007b).
Wolffish have been subject to directed fisheries
off Greenland, but off Canada they are caught
as by-catch, with about 1,000t per year taken in
the 1980s. The species have never been taken
in directed fisheries in this area because they
do not reach sufficient densities to maintain
a commercial fishery, despite their market
appeal. The decline in wolffish populations
in the northwest Atlantic is catastrophic, with
northern wolffish declining by 95 percent in
three generations, and spotted wolffish declining
by >90 percent in Canadian waters (Kulka
et al., 2007b). Striped wolffish has declined
significantly in Canadian waters (O’Dea &
Haedrich, 2003), and in US waters further
south catches declined 94.9 percent between
1983 and 2004 (Fuller & Watling, 2008). Major
contractions in the distribution of these species
have also been identified, with a possible but
unlikely shift in distribution to deeper waters
(Kulka et al., 2007b). At present it would appear
that all three of these wolffish species are
threatened with local extinction in the northwest
Atlantic, an area of global significance for
them. All three are considered species-at–risk
in the northwest Atlantic: A. lupus as Special
Concern (≈ IUCN Vulnerable) and A. minor and A.
denticulatus as Threatened (Baker et al., 2009).
Canadian assessments that established the
threat status of the three wolffish species
state that a definitive cause of their decline
is not apparent. The size and habitat of all
three species makes them highly susceptible
to capture by bottom trawls. A significant part
of the catch of wolffish in the NAFO Regulatory
Area comes from outside the Canadian EEZ
and it is suspected that catches are being
under-reported (Kulka et al., 2007b). Catch
statistics for the period 1995–2002 indicate
that significant quantities of wolffish were taken
in NAFO Areas (18.7 percent), which include
high seas areas on the southern part of the
Grand Banks, and the Flemish Pass (Kulka et
al., 2007b). These catches are retained for
commercial purposes and it is likely that the
fish stocks are continuous across national and
high seas waters (Kulka et al., 2007b). There
is often confusion in the identification of the
three wolffish species and they may therefore
be reported as wolffish or catfish for commercial
purposes (Kulka et al., 2007a). At present there
is no specific management regime for wolffish in
the NAFO Regulatory Area. In Canadian waters,
northern and spotted wolffish must now be
released alive if caught as by-catch. In addition
to direct mortality, it is likely that trawling
damages shelter and nesting sites for wolffish
and this may also be a contributory factor to its
decline (Kulka et al., 2007b; Fuller & Watling,
2008).
38 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Skates (Rajidae)
Skates are low-productivity species
characterised by low rates of growth, high ages
at maturity, low fecundities and high longevity
(McPhie & Campana, 2009). In the northwest
Atlantic attention was drawn to the impacts
fishing had on skates following the suggestion
that the largest species in the area, barndoor
skate (Dipturus laevis), was close to extinction
(Casey & Myers, 1998). However, subsequent
work has suggested that this species has
increased over recent years on the Grand Banks
(Gedamke et al., 2005) and has not become
extinct, although numbers were reduced greatly
in areas for which records are available off
Newfoundland (>99.9 percent decrease in 20–
30 years; Casey & Myers, 1998).
The most commonly landed commercial species
in the northwest Atlantic area include winter
skate (Leucoraja ocellata), little skate (Leucoraja
erinacea), thorny skate (Amblyraja radiata) and
the smooth skate (Malacoraja senta). Analyses
of data on these species from Canadian waters
(Scotia Shelf) indicates that from 1970 to 2006
the abundance of mature winter, thorny and
smooth skate declined by >90 percent, whereas
little skate increased in adundance (McPhie &
Campana, 2009). This pattern of declines in
larger species and increases in smaller has
been observed elsewhere in the world with other
skate species (e.g. Dulvy et al., 2000). About 95
percent of the catches in the NAFO Regulatory
Area are of thorny skate and mainly come from
Areas 3LNOP, which include the high seas.
There are major uncertainties in the catches of
these species prior to 1996 (NAFO SC, 2008),
primarily due to failure to distinguish between
species. Catches are thought to have peaked
at around 31,500t in 1991 and for Areas 3LNO
averaged at about 9,050t from 2000–2007.
Currently, the biomass of the stocks, estimated
from scientific surveys, are low compared to the
1980s but have remained stable in recent years
and even increased slightly from 2005–2007.
Skates came into regulation by NAFO in 2004.
NAFO SC (2008) recommended that catches of
skates be maintained at around 6,000t to allow
continued recovery of the stock but the NAFO
Fisheries Commission set a TAC for 2010 of
12,000t on the basis of evidence of increased
stock size. Winter skate, a mainly shelf species,
is considered a species-at-risk in the northwest
Atlantic with population assessments of Special
Concern (Western Scotian Shelf), Threatened
(Eastern Scotian Shelf), Endangered (Gulf of St
Lawrence) and Data Deficient (Newfoundland)
(COSEWIC, 2005).
Sharks
Within the region, porbeagle shark (Lamna
nasus) is subject to a directed longline fishery
and, as a result of significant declines in
populations, has been subject to a low TAC
within Canadian waters. The species is also
taken as by-catch in pelagic fisheries for tuna
and billfish (NAFO SC, 2008). In recent years
catches from pelagic longlines have increased
while mean size has decreased, indicating a
serious threat to the species. The porbeagle
is a pelagic shark of the high seas where it is
caught in epipelagic waters <200m in depth
(Campana & Joyce, 2004). Good catch statistics
and demographic data exist for the region.
The species is considered Endangered in the
northwest Atlantic (COSEWIC, 2004).
There are few data for other sharks in the
region. The only sharks with commercial value
in the area are the spiny dogfish (Squalus
acanthias), a shallow-water species mainly taken
on the Canadian Shelf (Kulka, 2006). Other
sharks, notably the black dogfish (Centroscyllium
fabricii), have little commercial value and are
only taken as by-catch in the NAFO Regulatory
Area. Several other species are also taken
outside the Canadian EEZ in the high seas
areas of the NAFO Regulatory Area (Murua &
de Cárdenas, 2005; Kulka, 2006), including the
Portuguese shark (Centroscymnus coelolepis),
the deep-sea cat shark (Apristurus profundorum)
and the great lantern shark (Etmopterus
princeps). However, fisheries data for these
are either non-existent or the species are
aggregated as deep-sea sharks/dogfish. Status
of these species within the region is unknown,
but all are relatively rare and of little commercial
interest.
Other species
Rabbitfish (Family Chimaeridae) are also caught
in NAFO Regulatory Areas 3LMNO. Two species,
the large-eyed rabbitfish (Hydrolagus mirabilis)
and the narownose rabbitfish (Harriotta
raleighana), are taken in mixed-species, bottom
trawl fisheries in deep waters of the Grand
Banks and Flemish Cap (González et al., 2007).
As with other Chondrichthyes, these species
have a conservative life history and are lowproductivity species. There is no information on
the state of populations of these species in the
northwest Atlantic.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 39
Protection of benthic marine ecosystems
Table 2. NAFO
seamount areas
protected from bottom
fishing in January
2007 (NAFO CEM,
2009).
Area
Fogo Seamounts 1
Fogo Seamounts 2
Orphan Knoll
Comer Seamounts
Newfoundland
Seamounts
New England
Seamounts
The NAFO Regulatory Area is dominated
oceanographically by cold, low-salinity water
flowing from the north in the Labrador Current.
To the south the region is bound by the Gulf
Stream, which means that the region is
characterised by very strong latitudinal gradients
in temperature. Within the region there are
areas of high primary production generated
by input of nutrients from strong tidal mixing
and, in areas, localised upwelling. The benthic
ecosystems of the region are well studied in the
south, especially on the northeastern shelf of
the USA and the southeastern shelf of Canada,
but less well studied in northern areas. However,
the occurrence of cold-water coral habitats and
deep-sea sponge grounds have been known in
the region for nearly 90 years (e.g. Verrill, 1922;
Deichmann, 1936; Breeze et al., 1997; Breeze &
Davis, 1998; MacIsaac et al., 2001; Mortensen
& Buhl-Mortensen, 2004, 2005; Gass & Willason,
2005; Mortensen et al., 2005; WGDEC, 2009)
and they are presently under active investigation
(Gilkinson & Edinger, 2009). The region has
extensive offshore banks (topographic elevations
associated with or on the continental shelf) but
Coordinate 1 Coordinate 2 Coordinate 3
42º31’33”N
53º23’17”W
41º07’22”N
52º27’49”W
50º00’30”N
45º00’30”W
35º00’00”N
48º00’00”W
43º29’00”N
43º20’00”W
35º00’00”N
57º00’00”W
42º31’33”N
52º33’37”W
41º07’22”N
51º38’10”W
51º00’30”N
45º00’30”W
36º00’00”N
48º00’00”W
44º00’00”N
43º20’00”W
39º00’00”N
57º00’00”W
41º55’48”N
53º23’17”W
40º31’37”N
52º27’49”W
51º00’30”N
47º00’30”W
36º00’00”N
52º00’00”W
44º00’00”N
46º40’00”W
35º00’00”N
64º00’00”W
Coordinate 4
41º55’48”N
52º33’37”W
40º31’37”N
51º38’10”W
50º00’30”N
47º00’30”W
35º00’00”N
52º00’00”W
43º29’00”N
46º40’00”W
35º00’00”N
64º00’00”W
a relatively limited number of large seamounts
(isolated topographic elevations), estimated at 43
for the NAFO Regulatory Area, of which a fraction
have summit depths in the range of fishing as
currently practised (Kulka et al., 2007c).
In reponse to UNGA Resolution 61/105, NAFO
closed an area along the southern slope of
the Grand Banks to protect corals from bottom
fishing, roughly along the 800 – 1,000m isobath.
Subsequent analyses of the closed area showed
that many records of octocorals and other corals
actually occurred in waters shallower than the
shallowest boundary of the closed area, just
outside the closed zone (WGDEC, 2008; Fig.
21). It was suggested that the protected area
should be extended to the 200m-depth contour to
protect all the deep-sea corals in this immediate
area (WGDEC, 2008). So far this has not been
acted upon.
In January 2007, six areas of seamounts were
closed to fishing as a measure to protect benthic
biodiversity (Table 2 and Fig. 20). Although
biological information on these seamounts was
extremely scant, there was some evidence of
coral by-catch in trawls from some of the New
England and Corner Rise Seamounts and it was
suggested that deep-sea coral frameworks may
exist in the Orphan Knoll region (Kulka et al.,
2007c).
Post-January 2008: A more systematic
approach to identification of VMEs
Some of these areas have been subject to
significant fisheries for deep-sea species,
particularly alfonsino (Beryx spp.), most notably a
Russian Fishery from 1978–1996 on the Corner
Rise Seamounts and some exploratory fishing by
other states (Kulka et al., 2007c; Thompson &
Campanis, 2007). Subsequent to the cessation
of the Russian alfonsino fishery, some fishing,
including exploratory fishing trips, occurred on the
Corner Rise Seamounts for alfonsino, wreckfish
and black scabbardfish, mainly by Spanish
vessels but also by Canadian vessels (Kulka et
al., 2007; Thompson & Campanis, 2007). Recent
ROV surveys on the Corner Rise Seamounts
identified the existence of some remaining coral
habitat, but large areas of the summits and upper
flanks of two seamounts, Kükenthal and Yakutat,
were denuded of large sessile animals and
showed evidence of trawl damage in the form of
seabed scars, other damage to the seabed and
coral fragments (Waller et al., 2007) (Fig. 22).
In 2008, the NAFO Scientific Committee and
the Working Group on Ecosystem Approach to
Fisheries Management (WGEAFM) initiated a
new approach to assessing coral and sponge
by-catch data from fisheries research surveys
obtained by Canada and Spain. These research
trawls covered a substantial portion of the NAFO
Regulatory Area, including the high seas regions
subject to bottom trawling (WGEAFM, 2009). Two
approaches have been used to identify VMEs
formed by corals and sponges. The first involved
the examination of cumulative catch data for
VME species by ranking the biomass of VME taxa
in each trawl from lowest to highest and then
plotting the increase in accumulative biomass
with each additional trawl. If VME taxa show an
aggregated distribution, as would be expected
if they form high-density patches of individuals
that structure the habitat, then at some stage
the accumulation curve will show a marked
sudden increase in biomass. This is because,
for the greater part of the seabed, the density
of individuals of VME taxa is low, so biomass
accumulates very slowly. Tows that encounter a
VME generate a large step in biomass of VME
species, identified by the area of maximum
curvature of the accumulation curve. This method
Figure 22. Kükenthal Peak
showing scar marks from
Figure 20. Map of seamount areas protected within NAFO
Regulatory Area (Thompson & Campanis, 2007).
Figure 21. NAFO coral protection area, initiated January
2008, showing records of sea pens (Pennatulacea) in green,
gorgonians in blue and soft corals in red (WGDEC, 2008).
40 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
bottom trawling. © DeepAtlantic Stepping Stones
Research Group.
was applied first to corals and then to sponges
for the NAFO Regulatory Area.
The second method uses Geographical
Information System (GIS) software to analyse
the distribution of density of sponge by-catch
in the NAFO Regulatory Area (WGEAFM, 2009).
Effectively, this identifies a circular search radius
around each point (cell) on a map and then
estimates the number of VME features within
that radius, dividing the number by the area
around the cell to get a unit density. For sponge
studies, a 25km search radius was chosen as it
recognises distinct geological features. Smaller
radii (e.g. 10km) resulted in a highly fragmented
picture of sponge density. Contour maps of
sponge density were then constructed.
Using the accumulation curve method, the
presence of coral VMEs was assessed by NAFO
WGEAFM (2008b). The report immediately
identified that it was likely that different
significance should be attached to the levels of
by-catch of different types of corals as colonies
have different weights and different morphology
(shape and structure), making them more or
less likely to be caught and retained in trawls
(WGEAFM, 2008b). Subsequent analyses
revealed that large catches of corals and sea
pens, indicating the presence of a potential VME,
were actually quite rare events in research trawls.
Identifying significant steps on the accumulative
curve of coral catches was difficult but a highly
conservative point was chosen representing the
97.5 percent quantile (upper bound of the 95
percent quantile). This represented a catch of
1.6kg per trawl for sea pens (Fig. 23) and 0.2kg
per trawl for small gorgonian octocorals (Acanella
arbuscula; Fig. 24). For larger gorgonians,
because they are more prone to breakage and
fragmentation in a trawl, a more precautionary
quantile of 90 percent was set, representing a
catch level of 2kg per trawl (WGEAFM, 2008b; Fig.
25), although in very exceptional circumstances,
where the seabed has not been previously
trawled, by-catch of such species has been high
(Krieger, 2001; Edinger et al., 2007b).
For sponges, the area occupied by weight of
sponges in trawls was broken into 25kg bins and
then plotted using GIS (WGEAFM, 2009; Fig. 26).
The weight at which a marked increase in area
occupied by sponges was found to be between
100 and 75kg of sponge by-catch. Further
analyses indicated that the 75kg weight threshold
for a trawl catch was the level that indicated
potential encounter with a VME. This was not
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 41
0
0
0
FIGURE 23
2
4
6
Weight (kg)
8
0
10
FIGURE 24.
1.0
0.4
0.2
WEIGHT QUANTILES
100% maximum 1.195300
99.5%
0.669198
97.5%
0.239704
90%
0.103750
75% quartile 0.035000
50% median 0.010000
25% quartile 0.003000
0.001000
10%
2.5%
0.001000
0.5%
0.000100
0% minimum 0.000100
0.8
0.6
0.4
0.2
0
2
4
6
Weight (kg)
8
10
0
FIGURE
24.
Figure
23. Cumulative
distribution of
0.2
0.4
0.6
0.8
Weight (kg)
1.0
0.6
0.4
0.2
1.2
0
Figure 24. Cumulative distribution of catch
FIGURE 25
(kg)
1.0 of Acanella arbuscula from research
surveys (WGEAFM, 2008b).
WEIGHT QUANTILES
trawl surveys (WGEAFM, 2008b).
1.195300
0.669198
0.239704
0.103750
0.035000
0.010000
0.003000
0.001000
0.001000
0.000100
0.000100
0
0.2
0.4
0.6
0.8
Weight (kg)
1.0
1.2
(b) from 250 – 25kg, using GIS
density rasters (WGEAFM, 2009).
FIGURE 25
Area (km2)
0 sponges from (a) 3980 – 1kg and
1.0
0.8
0.6
f(x)
f(x)
100% maximum
99.5%
97.5%
0.6
90%
75% quartile
50% median
0.4
25% quartile
10%
Figure 26. Area occupied
2.5%by trawls
0.2
0.5%
with decreasing catch0%weight
of
minimum
0.4
FIGURE
90,000
26 (a)
80,0000.2
70,000
60,000 0
50,000
0
0.2
0.4
0.6
0.8
Weight (kg)
1.0
0.2
f(x)
0.4
0.2
0.2
0.4
0.6
0.8
Weight (kg)
1.0
1.2
Area (km2)
0
1.0
1.2
Keratoisis ornata, Acanthogorgia armata, Paramuricea
40,000
30,000
7
3
CODE
WEIGHT
(kg)
AREA
(km2)
N
TOWS
1
2
3
4
5
9
12
14
16
17
18
19
3,980
1,994
999
500
250
125
62
31
15
7
3
1
505
1,743
6,731
8,623
9,186
11,896
19,204
23,432
28,042
36,895
56,690
82,516
7
21
37
61
73
93
118
155
195
264
352
501
CODE
WEIGHT
(kg)
AREA
(km2)
N
TOWS
5
6
7
8
9
10
11
13
15
250
200
175
150
125
100
75
50
25
9,186
10,143
10,143
10,143
11,898
15,032
18,810
20,368
26,371
73
87
88
89
93
102
113
130
168
Figure 29. Areas where catches of octocorals exceed the threshold values
identified by WGEAFM (WGEAFM, 2008b).
20,000
15,000
10,000
Figure 30. Areas where the weight of catches of sponges exceeded the
threshold identified by WGEAFM. Numbers represent NAFO numbering system
for potential VME locations (WGEAFM, 2009).
5,000
0
250
250
175
150
125
100
75
50
25
Weight (kg)
FIGURE 27
Figure 27. Cumulative distribution of
FIGURE 28
COMBINED DATA
(Weight > 0.5 kg)
1.0
(30-minute tows) and Canadian (15-minute
WEIGHT QUANTILES
100% maximum 5000
99.5%
3370.58
99%
2151.38
953.13
97.5%
95%
430.00
124.46
90%
85%
52.52
80%
25.81
75% quartile 16.10
50% median 3.50
25% quartile 1.25
0.97
20%
15%
0.85
0.8
tows) research trawls. Box shows weights
corresponding to the maximum curvature of
0.6
f(x)
catch cumulation curve (WGEAFM, 2009).
Figure 28. (Detail of Fig. 27) zoomed in to
0.4
show the curve between 10 and 300kg.
Black arrow indicates the first long step in
0.2
the cumulation curve, corresponding to 75kg
of sponge catch, Red arrows indicate the
100 and 125kg points (WGEAFM, 2009).
0
0
1,000
2,000
3,000
4,000
1.0
0.9
f(x)
weight of sponge catches from Spanish
0.8
0.7
0.6
50
100
150 200
Weight (kg)
5,000
Weight (kg)
42 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
250
These areas corresponded to the northern, eastern
and southern flanks of the Flemish Cap, southern parts
of the Flemish Pass and the eastern and southern
flanks of the Grand Banks. They were used to guide the
closure of 11 further high seas areas to bottom fishing
in September 2009 to protect VMEs formed by sponges
and corals (Fig. 31). In addition to the previous NAFO
Regulatory Area, 3O, this makes a total of 12 benthic
areas in the high seas as well as a further six seamount
areas closed to bottom fishing in the NAFO region.
The move-on rule
1
25,000
0
0.6
0.8
Weight (kg)
spp.) from research trawl surveys (WGEAFM, 2008b).
1.2
15
0.4
gorgonians (Paragorgia spp., Primnoa resedaeformis,
20,000
0.6
compatible with the 97.5 percent quantile (close to
1,000kg; Fig. 27) but was compatible with the maximum
curvature of a plot of the cumulative weights of sponge
by-catch in trawls (WGEAFM, 2009; Fig. 28). This
threshold weight appeared to be consistent between
approaches in identifying what weight of sponge by-catch
in a trawl represented a likely encounter with a VME.
Figure 25. Cumulative distribution of catch (kg) of large
WEIGHT QUANTILES
100% maximum 68.580000
99.5%
67.484900
97.5%
34.118850
90%
2.066280
75% quartile 0.325000
50% median 0.030000
25% quartile 0.010000
0.002000
10%
2.5%
0.001000
0.001000
0.5%
0% minimum 0.001000
WEIGHT QUANTILES
10,000
100% maximum 68.580000
0
99.5%
67.484900
3,980 1,994 999 500 250 125 62 31
97.5%
34.118850
Weight (kg)
2.066280
90%
75% quartile 0.325000
50% median 0.030000
25% quartile 0.010000
FIGURE 26 (b)
10%
0.002000
2.5%
0.001000 30,000
0.5%
0.001000
0% minimum 0.001000
0.8
1.2
0
1.0
catch
(kg) of sea pens from research trawl
0.8
1.0
WEIGHT QUANTILES
100% maximum 68.580000
99.5%
67.484900
97.5%
34.118850
90%
2.066280
75% quartile 0.325000
50% median 0.030000
25% quartile 0.010000
0.002000
10%
2.5%
0.001000
0.5%
0.001000
0% minimum 0.001000
0.8
0
0
0.6
0.8
Weight (kg)
FIGURE 25
f(x)
f(x)
0.6
0.4
f(x)
WEIGHT QUANTILES
100% maximum 0.116000
99.5%
4.718880
97.5%
1.602200
90%
0.454600
75% quartile 0.170000
50% median 0.049000
25% quartile 0.010500
0.003094
10%
2.5%
0.001000
0.5%
0.001000
0% minimum 0.000100
0.2
1.0
1.0
0.8
2.5%
0.001000
0.5%
0.000100
0% minimum 0.000100
0.2
2.5%
0.001000
0.5%
0.001000
0% minimum 0.000100
0.2
300
Figure 31. Areas closed to bottom fishing by NAFO, September, 2009 (NAFO
Fisheries Commission, 2009).
The NAFO move-on rule for coral by-catch was slightly
adjusted at the Annual Meeting in September 2009
from 100kg of live coral to 60kg of live coral. This
followed the realisation that within the NAFO area the
by-catch of corals in research trawls would never have
triggered a move-on incident with a threshold value of
100kg. The current threshold levels are still more than
one order of magnitude higher than that estimated
using accumulation curves for large octocorals and
more than two orders of magnitude larger than that for
small corals. The scientists advising NAFO undertook
estimation of a by-catch that signified an encounter
with a VME using extensive datasets from research
trawls and based on a consistent and rigorous
methodological approach (WGEAFM, 2008b, 2009).
A similar inconsistency applies to sponge threshold
levels, with a current trigger level of 800kg, compared
to a threshold weight of 75kg for research trawls, which
was estimated as indicative of the presence of a VME
(WGEAFM, 2009). In spite of the much longer duration
of commercial trawl tows, a linear increase in the ‘catch’
of sponge and corals in commercial fishing gear is not
expected as coral or sponge VMEs occur in discrete
patches (rarely more than 500m across) and encounters
with such VMEs are relatively rare with trawls, although
the chance of encountering more than one patch
increases with increased trawl time (WGEAFM, 2008b).
Recognising this, the USA put forward a proposal at the
Annual Meeting of NAFO in 2009 to establish threshold
limits of 2kg of corals and 75kg of sponges. NAFO,
however, only agreed to reduce the threshold levels
of ‘live’ corals from 100 to 60kg and from 1,000 to
800kg for sponges. NAFO has begun the process of
developing identification guides for VME taxa in the
NAFO Regulatory Area (Kenchington et al., 2009b).
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 43
Conclusions
(i) Conduct assessments of whether bottom fishing
activities have SAIs on VMEs.
● As yet no impact assessments have been
undertaken for bottom fishing operations in the
high seas of the NAFO Regulatory Area.
(ii) To implement measures in accordance with the
precautionary approach, ecosystems approaches
and international law and to sustainably manage
deep-sea fish stocks.
● The deep-water fisheries of the NAFO Regulatory
Area have a record of severe overfishing
and many stocks are at a fraction of historic
abundance and biomass and are in recovery from,
or remain in, a depleted state.
● The only deep-water low-productivity species
managed by NAFO are skates and redfish.
● Other low-productivity species are taken as
by-catch in deep-water fisheries for Greenland
halibut, redfish and skate. Some of these species
are threatened with extinction in the NAFO
Regulatory Area as a result of a combination of
environmental change and overfishing. NAFO has
made no attempt to manage the catch of these
species to ensure that populations remain in a
viable state. This situation is a direct threat to the
biodiversity of high seas deep-sea ecosystems in
the NAFO Regulatory Area.
● New fisheries, such as the pelagic redfish fishery
on the Flemish Cap, are not regulated in a
manner that would be consistent with the FAO
Guidelines in respect of exploratory fisheries.
● Extensive misreporting occurs in the deep-sea
high seas fisheries in the NAFO Regulatory Area.
Under-reporting of catches of some species or
groups of species is suspected in the high seas
deep-water fisheries.
● NAFO also does not identify fish catches to
species-level for several groups of regulated and
unmanaged fish, including redfish, wolffish, skate
and sharks.
● The NAFO Fisheries Commission has consistently
set catches above the levels proposed by the
NAFO Scientific Council for low-productivity deepwater species, in some cases to levels more than
double the recommended TACs.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom fishing
where VMEs are known or likely to occur, or not
authorised to proceed.
● NAFO has identified areas of high concentrations
of corals and sponges using information on
by-catch from research trawls. Ongoing studies
have adopted non-destructive photo survey
methods for assessments.
● This information has been used to designate
areas for protection from bottom fishing.
● In some cases, the exact outline of the protected
areas does not reflect data on the positions of
occurrence of VME taxa. The reasons for this are
unclear.
● Aspects of the scientific advice on the potential
occurrence of VMEs and establishment of
thresholds of coral or sponge catch that indicate
the presence of VMEs have been exemplary.
However, the knowledge gained is not currently
reflected in the move-on rule (see below).
● The ecological and biological significance of lesser
concentrations of corals, sponges and other VME
indicator species found in the NAFO Regulatory
Area have not been assessed to establish
whether additional protective areas are required.
(iv) To establish and implement protocols to cease
fishing where an encounter with VMEs occurs
during fishing activities and to report such
encounters so that appropriate measures can be
adopted with respect to that site.
● The threshold levels set by NAFO for VME
encounters apply to sponges and corals only.
● The threshold levels for corals exceed scientific
estimation of threshold levels that indicate
coral VMEs by more than one or two orders of
magnitude, depending on the category of coral
(sea pens, small corals, large corals).
● The threshold levels for sponges are also set at
more than one order of magnitude above levels
estimated by examination of by-catch data from
research trawls. As for corals, current threshold
levels may have little or no conservation value.
● Using the same threshold levels for active and
passive fishing gears does not reflect their greatly
differing impact.
● Scientific advice to NAFO identified that using
the same threshold levels for different types
of corals was likely to underestimate the
occurrence of VMEs formed by smaller octocorals,
antipatharians and stylasterids.
● Differentiating the post-VME encounter protocol
between areas with a fishing history and those
without is inconsistent. A VME has the same
conservation value whether or not it is in an area
with a history of fishing.
● The 2nm move-on rule is ineffective as a
conservation measure, as it is impossible to
identify where a VME encounter occurs along a
tow for commercial bottom trawling and therefore
has no conservation value. In the case of NAFO,
commercial trawls are up to 20nm long.
44 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
MEDITERRANEAN SEA
The Mediterranean is an enclosed sea with
depths as great as 5,121m and is notable
for being one of the few warm, deep-ocean
basins in the world, with temperatures a
uniform 12.5–14.5oC (Cartes et al., 2004). It
comprises many of the topographic features
found elsewhere in the deep ocean, including
canyons, cold seeps, seamounts and coldwater coral reefs. However, because the
countries surrounding the Mediterranean have
not exercised their rights to claim a 200nm
EEZ, many of these features associated with
the continental slope are in, or partially in,
the high seas and not within waters under
national jurisdiction (Bensch et al., 2008). The
RFMO for the region is the General Fisheries
Commission for the Mediterranean (GFCM)
(Fig. 32), which has been in existence since
1952, making it one of the oldest RFMOs in
the world.
The Mediterranean deep-sea fauna is unique,
as a result of isolation by a shallow sill, current
oceanographic conditions and the historical
impacts of the Messinian Salinity Crisis (5.7–5.4
MYA). Elements of the deep-sea fauna are
impoverished compared to that of the deep
Atlantic Ocean but levels of endemism in the
Mediterranean are high (>26 percent), although
this varies markedly by taxonomic group and
is rather low in fishes. The Mediterranean is
highly oligotrophic and this may explain the lower
densities of fish, macrofauna and meiofauna
compared to adjacent areas of the Atlantic
Ocean (Cartes et al., 2004). Fish and decapod
crustaceans are prevalent in the megafauna of
the deep Mediterranean. Fish diversity is also
lower than the Atlantic and the composition is
different, with a higher relative proportion of
deep-sea sharks. Decapod communities are
dominated by large shrimp of tropical origin.
For these reasons, Mediterranean deep-sea
fisheries are targeted at relatively few species
compared to other oceans, with an emphasis
on hake (Merluccius merluccius) and deep-water
shrimps (Aristeus antennatus, Aristeomorpha
foliacea) (Bensch et al., 2008). However, other
deep-water species are also fished, including
blue whiting (Micromesistius poutassou),
Figure 32. Map of
the Mediterranean
showing the GFCM
Regulatory Area
(Bensch et al.,
2008).
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 45
greater forkbeard (Phycis blennoides), angler
fish (Lophius spp.), conger eel (Conger conger)
and blackspot sea bream (Pagellus bogaraveo),
as well as crustaceans including the shrimps
Parapenaeus longirostris, Pasiphaea spp.,
Acanthephyra eximia and Plesionika spp., the
Norway lobster Nephrops norvegicus, and the
crabs Geryon longipes and Paramola cuvieri
(Cartes et al., 2004).
Management of fisheries for deep-sea
species of low productivity
None of the main target species from deep
waters in the Mediterranean are low-productivity
species. However, heavy fishing pressure in
the Mediterranean has resulted in stocks
of high-productivity species such as hake
being overexploited (e.g. off northern Spain
Subdivision 6, Gulf of Lions Subdivision 7, and
elsewhere; GFCM SCMEE, 2008a) and some
drastically so (e.g. northern Levant Sea; GFCM
SCMEE, 2008a). Growth rates of the shrimp
Aristeus antennatus are lower than other
penaeids and some stocks in the Mediterranean
are overexploitated (Cartes et al., 2004). In
2005, at the 29th Session of the GFCM, it
was decided not to allow fishing to extend
beyond 1,000m depth in the Mediterranean, a
decision partially reflecting the lack of species
of commercial interest living there (Cartes et
al., 2004). The juveniles of Aristeus antennatus
recruit almost exclusively below depths of
1,000m and juveniles and males generally
occur below these depths. This may explain the
high resilience of this species to exploitation
in many areas, as a substantial part of the
population lives beyond the depths at which
fishing takes place or is allowed (Cartes et
al., 2004). Aristeus anntenatus in the Gulf of
Lions undergoes periodic crashes in landings
as a result of formation of cold, dense water at
the surface of this area in winter, followed by
cascading of the water into deep water. Three
or four years after such cascading events the
populations undergo good recruitment, probably
because of transport of organic material into the
deep sea during such events (Company et al.,
2008). Other deep-water crustacean fisheries,
such as those for deep-water pink shrimp
Parapenaeus longirostris, are also depleted
or overexploited in the region (GFCM SCMEE,
2008a).
While the serious by-catch problems associated
with shallow-water fisheries in the Mediterranean
on sharks, turtles, cetaceans, birds and
pinnipeds are recognised (GFCM Scientific
Advisory Committee, 2008a), those affecting
deep-water species are not. By-catches of
deep-water species are especially associated
with trawl fisheries, which in the Mediterranean
are almost all multispecies fisheries. Discard
levels are high. One study of discards in six
Spanish ports and one Italian indicate discards
of 13–62 percent of catch in waters of 150–
350m depth and 14–43 percent in waters of
>350m depth (Carbonell, 1997; Carbonell et al.,
1998). Studies of the deep-water trawl fleet of
Alacante (southeast Spain) indicated discards
of 34.6 percent of the total catch (Soriano &
Sánchez-Lizaso, 2000). In this fishery, of the 95
species taken in trawls, 89 are discarded. The
impacts of such multispecies trawl fisheries on
biodiversity are significant. Experimental trawl
surveys in the Gulf of Lions in recent years
recorded only 13 species of elasmobranchs,
whereas 25 species were recorded in 1957–
1960 (Tudela, 2000). The area is subject to an
intensive trawl fishery and the main target stock
of hake is overexploited.
By-catch for deep-water trawl, gillnet and longline
fisheries is poorly documented but there are
reports of significant by-catch of deep-water
sharks in several fisheries. Off the coast of Italy
significant by-catches of the black-mouth cat
shark (Galeus melastomus) and gulper shark
(Centrophorus granulosus) have been reported,
with the former species being taken on its
spawning grounds (Tudela, 2000). Significant
by-catch of the roughshark (Oxynotus spp.) has
been reported from Greek waters (Tudela, 2000).
The angular roughshark Oxynotus centrina
has been classified as Critically Endangered
in the Mediterranean and is now rare or has
been extirpated from many areas where it was
formerly abundant (Cavanagh & Gibson, 2007).
Reports on the status of deep-water sharks
in the Mediterranean indicate that several
are threatened with extirpation as a result,
mainly, of being taken as by-catch. Particularly
notable are three species of angel sharks
in the genus Squatina: Squatina aculeata,
Squatina oculata and Squatina squatina. All
three were historically abundant in the region
but have suffered severe range contractions and
declines (Cavanagh & Gibson, 2007). These
species have relatively small distributions in
the Mediterranean, along the west coast of
Africa and off Europe (Campagno et al., 2005).
IUCN recognises Squatina squatina as Critically
Endangered globally and S. aculeata and S.
46 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
oculata as Endangered globally (Cavanagh &
Gibson, 2007). All three species are Critically
Endangered in the Mediterranean. IUCN has also
classified the rabbitfish Chimaera monstrosa as
Near Threatened in the Mediterranean because
its preferred depth-range lies entirely within
the depth that deep-water fishing takes place
(Cavanagh & Gibson, 2007). Other species of
deep-water sharks taken as by-catch include
Etmopterus spinax and Hexanchus griseus
(Saidi & Bradai, 2008). All of these species, as
with many chondrichthyans, are long-lived with
very low fecundities and therefore are highly
vulnerable to overfishing.
GFCM recognises the issues of by-catch from
Mediterranean deep-sea fisheries and there is
currently a working group within the Scientific
Advisory Committee on by-catch and incidental
catches. Some species of sharks have been
afforded protection by Recommendation GFCM
2005/1 to ban fishing below 1,000m depth,
including the Portuguese dogfish Centroscymnus
coelolepis, which occurs from depths of
~1,300m to >2,800m (Cavanagh & Gibson,
2007). In addition, Recommendation GFCM
2005/1 also requested a minimum of 40mm
mesh size for the cod end of demersal trawls
in an effort to reduce by-catch. However, this
could only confer conservation value to very
small or juvenile deep-water fish. This measure
is currently in the process of being implemented
and will be fully in place for the EU by 2010.
Recommendation GFCM 2009/1 has called for
a reduction in demersal trawling by 10 percent
in the GFCM Regulatory Area. Whether this
reduction will benefit deep-sea species and
habitats depends on where effort reductions
in trawling are implemented. In general, the
issue of by-catch in the deep waters of the
Mediterranean remains under-researched and is
likely to impact a wide range of low-productivity
and VME species (see below).
Protection of benthic marine ecosystems
The VMEs represented within the deep waters
of the Mediterranean are unique, reflecting the
general characteristics of the fauna (see above).
They include communities formed by crinoids
(Leptometra phalangium), octocorals (Funiculina
quadrangularis, Isidella elongata), stony corals
(Lophelia pertusa, Madrepora oculata) and
brachiopods (Gryphus vitreus). These habitats
are associated with a high level of diversity of
associated species and are also associated with
juvenile and adult stages of commercially fished
species (European Commission, 2006). Some
of these habitats have been largely destroyed
in the deep waters of the Mediterranean, for
example, beds of Funiculina quadrangularis and
Isidella elongata have largely disappeared from
many areas as a direct result of bottom trawling
(European Commission, 2006).
Crinoid beds
Communities associated with high densities of
the crinoid Leptometra phalangium occur on the
shelf edge (120–180m depth off Italy) around
the Mediterranean, where there is strong water
movement and a plentiful supply of plankton and
organic matter. They comprise an abundance
of suspension feeding organisms, and the
crinoids introduce three-dimensional complexity
to the environment, enhancing diversity. This
habitat acts as essential fish habitat for a
number of commercial species (Colloca et al.,
2004) including hake, blue whiting and poor cod
(Trisopterus minutus), John Dory (Zeus faber),
red mullet (Mullus barbatus), angler fish (Lophius
spp.), thornback ray (Raja clavata) and the squid
Illex coindetti. Leptometra are extremely fragile
and are easily destroyed by trawling.
Gryphus vitreus (brachiopod)
The brachiopod Gryphus vitreus is associated
with particular soft-bottom communities under
the influence of currents. Its density on the
seabed is an excellent indicator of current speed
and the species can form belt-like zones from
100–250m depth (Emig, 1988), where it may
also be associated with Leptometra phalangium
(Ordines & Massutí, 2009) or Isidella elongata.
The sediments occurring with Gryphus vitreus
may be silty, with a diversity of molluscs, infauna
and epifauna colonising relatively small pieces
of hard substrata (Rosso et al., 2009).
Isidella elongata (octocoral)
This octocoral forms beds in areas of compact
mud on the middle slope from 500m to at
least 1,200m (European Commission, 2006;
Ramírez-Llodra et al., 2008). It forms coral
gardens associated with a high diversity of
benthic invertebrates and high densities
of commercial species including shrimps
(Aristeus antennatus, Aristaeomorpha foliacea,
Parapenaeus longirostris), lobsters (Nephrops
norvegicus) and fish (Merluccius merluccius,
Micromesistius poutassou). Both live and
dead coral areas are important as habitat for
other species. This habitat, along with gardens
formed by the species Funiculina quadrangularis,
has been destroyed across large areas of the
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 47
Mediterranean as a result of bottom trawling in
deep water. All have largely disappeared from
some areas.
Funiculina quadrangularis (sea pen)
Funiculina is a large sea pen reaching 1.5–
2.1m in height that forms dense gardens on
undisturbed sediments and occurs on the
shelf edge and upper slope throughout the
Mediterranean (European Commission, 2006).
The dense beds of sea pens represent essential
habitat for some commercial crustaceans such
as Parapenaeus longirostris and the Norway
lobster Nephrops norvegicus. Communities
formed by this species have been almost
completely destroyed by trawling in many parts
of the Mediterranean.
Lophelia pertusa and other coral
communities
The distribution of the framework-building corals
Lophelia pertusa and Madrepora oculata are
poorly known in the Mediterranean but they have
been identified in scattered localities. These
corals have been observed near the Gibralter
Sill; in the northwest Mediterranean canyons
between Cap de Creus and the Ligurian Sea;
in the Sicilian Channel; on the Apulian Plateau
off southern Italy; and in the southwestern
Adriatic margin off the coast of southeastern
Italy (Fig. 33; Freiwald et al., 2009). The stony
coral communities occur in waters at depths
of ~450m to more than 1,100m in a variety
of environmental settings including canyons,
submarine cliffs, and steep or complex
submarine topography (Taviani et al., 2005;
Freiwald et al., 2009).
The largest complex of coral communities
discovered lies off the coast of Santa Maria de
Luca in Apulia, southern Italy. The reefs were
discovered through coral being caught in the
nets of local fishermen. The corals occur at
depths between 500m and >1,100m in an area
of complex terrain characterised by hummocks
on the seabed and strong current flows
(Taviani et al., 2005). The coral communities
at Santa Maria de Luca are dominated by
Madrepora oculata, with Lophelia pertusa
and Desmophyllum dianthus also occurring.
Associated diversity is lower than northeast
Atlantic cold-water coral reefs but includes
other stony corals (Stenocyathus vermiformis),
octocorals, bivalve molluscs, polychaetes,
including the coral-associated Eunice norvegicus,
and sponges (Taviani et al., 2005).
Fishing is regarded as a major threat to
deep-water stony coral communities in the
Mediterranean, especially given that the
occurrence of these habitats is poorly known,
relatively rare and scattered (Cartes et al.,
2004).
Seeps have been identified in the southeastern
Mediterranean, on the Mediterranean Ridge,
to the south of Crete and Turkey (Anaximander
Mountains) and to the north of Egypt near
the Nile Delta. These seeps are associated
with unique communities of animals with
endosymbiotic bacteria, including bivalve
molluscs (Lucinidae, Vesicomydae, Mytilidae
and Thyasiridae) and siboglinid worms that
utilise hydrogen sulphide or methane (Vanreusel
et al., 2009). Non-symbiont-hosting fauna can
also be abundant at these sites as a result
of heterogenous habitat, elevated topography
and high supplies of food, including polychaete
worms, sponges, echinoids and other species
(Vanreusel et al., 2009). Cold seeps may be
associated with specific geological features
on the seafloor, including mud volcanoes
and pockmarks. Related features are deep
hypersaline basins, which are high-salinity brine
pools lying on the deep seabed that have high
concentrations of associated methane and
hydrogen sulphide (Cartes et al., 2004). Most
of the sites discovered so far lie below 1,000m
depth but those around the Nile Delta are
shallower at 500–800m depths.
Chemosynthetic communities
Cold seeps have been identified in several
areas of the Mediterranean in deep waters.
The contact zone between the Eurasian and
African plates represents one of the world’s
major provinces associated with the seepage
of hydrocarbon associated fluids, particularly
in the deep sea (CIESM, 2006). At cold seeps,
methane-rich fluids provide energy for bacteria,
which may also produce hydrogen sulphide
through the process of sulphate reduction.
Seamounts
Seamounts are not necessarily VMEs in
themselves but often host VMEs such as
cold-water coral reefs or coral gardens. The
seamounts of the Mediterranean are poorly
explored. In the eastern Mediterranean,
Eratosthenes Seamount, a feature with an
elevation of about 1,500m and a summit
depth of 756m, has been subject to limited
studies (Galil & Zibrowius, 1998). It hosts a
diverse community of organisms including the
corals Caryophyllia calveri and Desmophyllum
dianthus as well as black corals and a variety of
polychaetes; the crustaceans Aristaeomorpha
foliacea, Aristaeus antennatus, Plesionika martia
and Polycheles typhlops; and fish, including
Hoplostethus mediterraneus (Galil & Zibrowius,
1998). There are seamounts in other parts of
the Mediterranean, especially in the Tyrhennian
Sea where they form part of the Eolian Arc.
Hydrothermal activity has been identified on
some (e.g. Marsili and Enarete Seamounts;
Uchupi & Ballard, 1989; Eckhardt et al., 1997).
Canyons
Figure 33. Locations where live Lophelia pertusa and Madrepora oculata have been found in the Mediterranean (Freiwald et al., 2009).
48 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Canyons are important ecosystems in the
Mediterranean, especially in the western
Mediterranean basin where they act as a conduit
for organic matter from the shelf into the deep
sea and can have an important influence on
commercial species (e.g. Company et al., 2008).
Canyons can have a higher abundance and
biomass of megafauna than surrounding slope
areas (Sardà et al., 1994) and can be important
in the distribution of suspension-feeding
organisms such as corals (Ramírez-Llodra et al.,
2008).
Conservation measures for deep-sea VMEs
in the GFCM Regulatory Area
Currently, deep-sea ecosystems are underrepresented in the protected areas of the
Mediterranean. Most marine protected areas are
coastal (Abdulla et al., 2008). Recommendation
GFCM/2006/3 established areas protected
from fishing with towed dredges and bottom
trawls around the Lophelia pertusa reefs at
Santa Maria de Luca, the cold seep ecosystems
in the Nile Delta and the benthic communities
of Eratosthenes Seamount (GFCM, 2006),
representing about 15,666km2 of seabed
(Abdulla et al., 2008). Recently, a further
fisheries-restricted area has been proposed in
the Gulf of Lions, specifically to protect spawning
grounds of the hake Merluccius merluccius
(GFCM SCMEE, 2008b). This conservation
measure (Recommendation GFCM/33/2009/1)
is only a freeze on current fishing effort (GFCM,
2009) and so offers limited and temporary
protection to benthic communities. The ban on
fishing below a depth of 1,000m does confer
protection to species of benthic organisms
whose distribution lies partially or wholly below
these depths. However, ecologically important
deep-sea VMEs remain vulnerable, including
coral gardens formed by Isidella elongata,
Funiculina quadrangularis, other corals and other
habitat-forming groups such as crinoids and
brachiopods, which customarily occur shallower
than 1,000m.
To date, there has been almost no response to
UNGA Resolution 61/105 in terms of impact
assessments of deep-sea fisheries in the
Mediterranean on benthic ecosystems. Indeed,
there are few data on many of the important
VMEs of the Mediterranean and their current
and past distributions are not understood.
Many of these habitats are vital to commercial
species, both juveniles and adults (see above).
It is widely acknowledged that several of these
important VMEs have been largely destroyed
across wide areas of the Mediterranean as a
result of bottom fishing, especially trawling.
GFCM has recently established criteria for
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 49
identification of sensitive habitats of relevance
for the management of priority species, including
VMEs formed by corals, crinoids and other
species, as well as seamounts, canyons and
cold seeps (GFCM Scientific Advisory Committee,
2008b). A report has also been made to
the European Commission on Sensitive and
Essential Fish Habitats in the Mediterranean
Sea (European Commission, 2006). This report
also identifies many of the significant deep-sea
VMEs in the Mediterranean area and describes
their relevance to fisheries and impacts on
them from fishing activities. As yet there is no
indication of the development of a systematic
approach to identification of VMEs or the
management of deep-sea fisheries to protect
such habitats.
Conclusions
(i) Conduct assessments of whether bottom
fishing activities have SAIs on VMEs.
● There has been no impact assessment
of fishing in the deep waters of the
Mediterranean on target and by-catch species,
including those that form VMEs.
(ii) To implement measures in accordance with
the precautionary approach, ecosystems
approaches and international law and to
sustainably manage deep-sea fish stocks.
● The Mediterranean is an enclosed sea with
a shallow sill separating it from the Atlantic
as well as a unique palaeoclimatic history.
As a result, it has a characteristic fauna and
distribution of communities in the deep sea
(e.g. a high proportion of chondrichthyans) that
require special management consideration.
● The main target species of Mediterranean
deep-water fisheries are not low productivity.
● Many of the trawl fisheries in the deep
waters of the Mediterranean are multispecies
and have significant impacts on non-target
species, some of which are characterised
by low productivity. Some have been so
severely impacted by fishing that they are
regionally recognised as Critically Endangered,
Endangered or Threatened, raising the risk of
long-term reduction in biodiversity of benthic
communities in the deep Mediterranean.
● At present GFCM has called for a minimum
mesh size of 40mm in the cod end of nets
and a reduction by 10 percent of the effort
of demersal fisheries in the Mediterranean.
The 40mm mesh size requirement will not
prevent the continued decline of the majority
of threatened deep-sea species and its
conservation value in waters down to 1,000m
depth in preventing environmental impacts by
multispecies demersal trawl fisheries is not
clear. As yet it is unclear where reductions
in fishing effort will take place and whether
benefits will accrue for deep-water species
and habitats.
● The GFCM has banned all forms of fishing
beyond 1,000m depth, affording protection
to species whose depth-range partially or
completely lies below this depth.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom
fishing where VMEs are known or likely to
occur, or not authorised to proceed.
● The Mediterranean has a unique marine
fauna that includes deep-sea VMEs formed
by a variety of taxa, some of which are most
common within, or unique to, the region.
● It is widely acknowledged that these
ecosystems have been seriously impacted
by bottom fishing.
● At present, three areas have been protected
from deep-water dredging and trawl fishing
and fishing is banned below 1,000m depth.
SOUTHEAST ATLANTIC OCEAN
The Southeast Atlantic includes one of the
world’s major eastern boundary current
upwelling systems, which makes it a highly
productive marine region. Nonetheless, it does
not feature among the 10 most important
ocean areas in terms of fish landings (FAO,
2009b). The region includes the coastal areas
off South Africa, Namibia and Angola and
extends westwards to beyond the southern
Mid-Atlantic Ridge (Fig. 34). The high seas
region of the Southeast Atlantic includes a
number of large topographic features including
the southern Mid-Atlantic Ridge, the Walvis
Ridge, the Vavilov Ridge, the Agulhas Ridge
and a number of isolated seamounts (e.g.
Vema Seamount) and rise features (e.g.
Meteor Rise; SEAFO Scientific Committee,
2006; Clark et al., 2007; Bensch et al., 2008).
The RFMO for the region is the South East
Atlantic Fisheries Organisation (SEAFO). Current
Contracting Parties include Angola, the EU,
Namibia, South Africa and Norway. Other states
have signed the Convention but have not ratified
it, including Republic of Korea, Japan, UK, USA
and Iceland (SEAFO Commission, 2008). Several
Fisheries in the region have included those
for small pelagic fish, including sardine,
anchovy, Whitehead’s round herring (Etrumeus
whiteheadi) and horse mackerel, and trawl
fisheries for hake (Merluccius capensis,
Merluccius paradoxus, Merluccius polli), kingklip
(Genypterus capensis), snoek (Thyrsites atun),
sole (Austroglossus microlepis) and monkfish
(Lophius spp.; Boyer & Hampton, 2001). There
are also a number of line fisheries and fisheries
for crustaceans, particularly lobsters (Jasus
spp.). Over the past two decades or so, fisheries
have developed on the continental slope for
orange roughy, alfonsino and deep-sea red crab
(Boyer & Hampton, 2001). Catches of many of
these species have declined in comparison to
historical catches as a result of overexploitation,
and in some cases the decline is also probably
related to environmental change (Boyer &
Hampton, 2001).
Status of deep-sea fisheries on the
high seas
Species managed by SEAFO in the high seas
are: Patagonian toothfish (Dissostichus
eleginoides), orange roughy (Hoplosthethus
atlanticus), alfonsino (Beryx spp.), deepsea red crab (Chaceon spp.), mackerel
(Scomber scombrus), armourhead/boarfish
(Pseudopentaceros spp.), oreo dories
(Oreosomatidae), cardinalfish (Epigonus spp.),
octopus, squid (Ommastrephidae), wreckfish
(Polyprion americanus), skates (Rajidae) and
sharks.
(iv) To establish and implement protocols to
cease fishing where an encounter with VMEs
occurs during fishing activities and to report
such encounters so that appropriate measures
can be adopted with respect to that site.
● No specific measures are in place to detect or
map VMEs in the Mediterranean region or to
manage impacts on them by bottom fisheries
outside current protected areas.
Orange roughy (Hoplostethus atlanticus)
Figure 34. Map of the SEAFO Regulatory Area.
50 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
of the non-ratifying states are actively fishing
in the SEAFO Regulatory Area, notably Japan
and Republic of Korea, and both of these have
undertaken to ratify the Convention in 2010
(SEAFO Commission, 2009).
In 1994, exploration for deep-sea fish stocks
by a commercial company began just inside the
Namibian EEZ. In 1995, spawning aggregations
of orange roughy were identified on a ground
known as Hotspot, a seamount at the southern
edge of the Walvis Ridge (19o20’S, 10o05’E;
Boyer et al., 2001). This was followed by
the discovery of further aggregations on the
continental slope in 1995/96 at the sites known
as Rix (22o30’S, 12o40’E), Johnies (26o20’S,
13o30’E) and Frankies (24o30’S, 13o20’E).
These fisheries were opened to full commercial
fishing in 1997, when they produced 15,500t
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 51
of orange roughy (Branch, 2001). Analyses of
commercial CPUE in 1998 indicated that the
stocks had declined significantly, although
there were objections from industry suggesting
that fishing effort had been directed at other
species rather than roughy and CPUE estimates
were therefore biased (Boyer et al., 2001). The
TAC was therefore only reduced to 12,000t. In
1998/99 it was discovered that the stocks had
declined precipitously, probably more so than
could be explained by fishing alone, and the
possibility that fishing was somehow disturbing
the spawning aggregations was raised (Boyer
et al., 2001). A TAC of 9,000t was issued but
only 2,500t of orange roughy were caught and
the Frankies ground was closed to assess the
impact of fishing on aggregating behavior by
roughy. The TAC was further reduced to 1,200t
in 2000/01. The fishery provided an example
of how uncertainty concerning stock size and
appropriate catch levels can lead to rapid
depletion in a new deep-water fishery based
on a low-resilience species.
Armourhead (Pseudopentaceros richardsoni)
catches are low. Incomplete reporting of catches
remains a problem and countries with a history
of fishing in the SEAFO Regulatory Area, such
as Spain, Portugal, Cyprus, Mauritius, Japan,
Russia, Poland, Norway, South Africa and
Namibia have not supplied SEAFO with historical
catch data (SEAFO Scientific Committee, 2008).
Even if some information is present there is
often no spatial component. There are recent
indications of improvement in this situation, with
increased reporting from Japan and Republic of
Korea.
Statistics for SEAFO refer to armourhead and
boarfish. No specific identification is given
for boarfish and the name may also refer to
armourhead (SEAFO Scientific Committee,
2006). The armourhead is a species associated
with seamounts that has also shown a low
resilience to exploitation because of its
aggregating behaviour, especially in areas such
as the North Pacific. Armourhead have been
fished on the high seas in the southeast Atlantic
since the 1970s when Soviet vessels targeted
the species along the Walvis Ridge. The species
has also been taken by a Japanese trawler on
the Valdivia Seamount in an exploratory fishery
in 1979, along with bluemouth (Helicolenus
dactylopterus). Recent, relatively small, catches
of this species have been reported in the SEAFO
Regulatory Area by Russia, Cyprus, Mauritius
and Namibia (SEAFO Scientific Committee,
2008).
SEAFO has ruled that no directed fisheries
should be undertaken for deep-water sharks in
the area because of their extreme vulnerability
to overfishing (SEAFO Scientific Committee,
2008).
Protection of benthic marine ecosystems
Longline fisheries for Patagonian toothfish
are regarded as one of the most valuable in
the SEAFO Regulatory Area. Catch figures are
available from Japan and Republic of Korea for
this species, with fishing in some years also
recorded for the EU (Spain). SEAFO initially set a
precautionary catch limit of 260t of Patagonian
toothfish for 2009, a level higher than the
reported yearly catches of the species within the
SEAFO Regulatory Area in recent years (SEAFO
Scientific Committee, 2008; SEAFO Scientific
Committee, 2009). For 2010, SEAFO reduced
the quota to 200t (SEAFO Commission, 2009).
Alfonsino (Beryx spp.)
The main species of deep-sea red crabs in the
SEAFO Regulatory Area is Chaceon maritae,
which occurs along the west coast of Africa,
but other species also occur. Deep-sea red
crabs are fished using pots, particularly in area
Division B1, which borders the Namibian EEZ
and which includes several seamounts, e.g.
Valdivia Bank, Maloy and Ewing. They are also
fished further south in Division D1. SEAFO has
recommended precautionary catch limits of
200t for Division B1 and 200t for the rest of
the SEAFO Regulatory Area. Recent catches of
crab by Japanese vessels alone have exceeded
500t for the SEAFO region (2007 figures; SEAFO
Scientific Committee, 2008).
BRAZIL
Ascension
2
1
A1
ANGOLA
St Helena
NAMBIA
Deep-Sea red crab (Chaceon spp.)
Alfonsino have been fished in the southeast
Atlantic Ocean since the 1970s when Soviet
vessels targeted this and other species on the
Vavilov Ridge (Clark et al., 2007). In 1978 about
4,200t of alfonsino and cardinalfish (Epigonus
denticulatus) were caught on Udachnaya
Seamount but the fishery declined until the
1990s when trawlers took mixed catches of
500–1,300t per year (Clark et al., 2007).
Fisheries for alfonsino still continue in the
SEAFO Regulatory Area and, recognising that the
species is vulnerable to overfishing because it
forms easily targeted aggregations, SEAFO has
set a precautionary TAC for the area of 200t
(SEAFO Scientific Committee, 2008; SEAFO
Commission, 2009).
In 2006, SEAFO decided to divide the Regulatory
Area into areas and subareas, with the latter
encompassing seamount areas thought to be
ecologically sensitive (Fig. 35; SEAFO Scientific
Committee, 2006).
SEAFO is a relatively new RFMO faced with a
very large area of ocean and with extremely
limited knowledge of benthic ecology and
the presence of VMEs within the region. It
Patagonian toothfish (Dissostichus
eleginoides)
In the high seas, orange roughy were fished on
the Walvis Ridge in the 1990s but little explicit
information on catches or landings is available
for this fishery. SEAFO has noted that fisheries
for orange roughy in the SEAFO Regulatory
Area are unmanaged and that sufficient data
do not exist for a meaningful assessment of
stock size or appropriate levels of exploitation.
Because of the vulnerability of orange roughy to
overexploitation, SEAFO has set precautionary
TACs of 50t for the entire SEAFO area for 2010
(SEAFO Commission, 2009). No catches were
identified for recent years (since 2005) for this
species from the SEAFO Regulatory Area.
has established a number of precautionary
measures to protect seabed ecosystems
and vulnerable species in response to UNGA
Resolution 61/105 including:
● the closure of seamount areas because of
the likelihood that they host VMEs;
● instruction of observers to collect data on
by-catch of VME species;
● support of research initiatives to improve
knowledge of the distribution of VMEs within
the SEAFO Regulatory Area (e.g. southern
Mar-Eco project);
● adoption of move-on rules for encounters with
VME species;
● banning the use of gillnets for fishing in the
SEAFO Regulatory Area – agreed at the 2009
Annual Meeting of SEAFO (SEAFO Commission,
2009).
Information on catches of other species in
the SEAFO area is poor and/or indicates that
52 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
3
B
4
B1
5
6
C1
C
SOUTH
AFRICA
7
Tristan da Cunha
10
Gough
9
8
11
EEZ
12
D
D1
13
Figure 35. Map showing SEAFO Regulatory Area divisions and subdivisions, with protected seamounts indicated: 1. Dampier Seamount; 2.
Malahit Guyot; 3. Ewing Bank; 4. Valdivia Bank; 5. Molloy Seamount; 6. Vema Seamount; 7. Wust Seamount; 8. Africana Seamount; 9. SchmidttOtt and Erica Seamount; 10. Panzarini Seamount; 11. Discovery Seamount, Junoy Seamount and Shannon Seamount; 12. Schwabenland
Seamount and Herdman Seamount (SEAFO Commission, 2006 – note: closed area 13 not identified in Commission report).
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 53
The SEAFO Fisheries Commission requested
that consideration be given to opening a small
proportion of each of these seamounts to fishing
(SEAFO Commission, 2007) but subsequently
it was agreed that this should only occur after
mapping the geomorphology to identify VMEs
(SEAFO Fisheries Commission, 2008). It was
also conditional on an assessment of the
sustainability of such fishing operations on
target species and of the potential for damage
to VMEs present within the area (i.e. an impact
assessment of the fishery; SEAFO Commission,
2008).
Alongside this modification of the rules regarding
closed areas, Spain has undertaken a joint
expedition with Namibia to map the Valdivia
Seamounts and Ewing Bank. This expedition
surveyed the multibeam bathymetry of these
seamounts, producing maps of Ewing Bank,
Valdivia North Seamount, Valdivia Central
Seamount, Valdivia West Seamount and
Valdivia South Seamount. Trawl surveys were
also undertaken on these seamounts and
data collected on both the fish fauna around
the seamounts and the benthic fauna retained
by the trawls. On Ewing Bank the expedition
recovered about 7kg of benthic invertebrates,
including mainly sea urchins (Hygrosoma
pertersii) and hermit crabs, with an associated
zoanthid. Bamboo corals were also recovered
(González-Porto, 2008). Trawls on Valdivia Bank
recovered over 11,000 specimens of benthic
animals weighing about 32kg, including a very
large number of hydroids and large biomass of
sea anemones as well as other taxa including
sea stars (Echinaster reticulatus) and sponges
(González-Porto, 2008).
The Expedition Report identified four main
communities of fish that occurred in different
depth-zones of the seamounts (Navarro et
al., 2008). The most representative fish of
the shallowest depths (200–500m depth)
were bluemouth (Helicolenus dactylopterus)
and pelagic armourhead or boarfish
(Pseudopentaceros richardsoni). From the
depth-zone 800–1,100m various macrourid
species were found, including Bathygadus
favosus, Cetonurus globiceps, Coelorinchus
labiatus, Gadomus capensis and Nezumia
brevibarbata. Other species included the shark
Etmopterus brachyurus and orange roughy. A
deeper assemblage, roughly corresponding to
the 900–1,300m zone, included Alepocephalus
productus, Coryphaenoides striaturus and
the warty oreo (Allocyttus verrucosus). The
deepest stratum included some species
from stratum three but also others such as
Bathysaurus ferox and Bathygadus favosus. A
comprehensive list of species caught at the
seamounts is presented (Navarro et al., 2008),
which includes commercial and potential bycatch species characterised by low productivity
and low resilience to exploitation. Biological
characteristics of some of the commercial or
important species on the seamounts were also
recorded (length frequency, length/weight ratio,
reproductive characteristics, sex ratio).
The Expedition Report includes new information
on the geomorphology, physical oceanography
and invertebrate and fish communities
associated with the Ewing and Valdivia Bank
seamounts. It was also aimed at identifying
“bioconstructions associated with seamounts
as potential vulnerable marine ecosystems
that could be damaged by fishing gears”. It
concluded that the trawl samples presented
no evidence of potential VMEs. This statement
does not address the limitations of fish trawls in
sampling benthic fauna (see discussion of the
move-on rule in the Northeast and Northwest
Atlantic Ocean sections of the report). It also
suggests that the definition of VMEs in this
case is narrow, as samples from Valdivia
Seamount comprised very large numbers of
erect hydrozoan colonies, sponges and large
sea anemones that may be indicative of the
presence of VMEs (and are included in the VME
indicator species guides in other RFMOs such
as CCAMLR). The SEAFO Scientific Committee
also agreed that “few conclusive results were
obtained” (SEAFO Scientific Committee, 2008).
The SEAFO Scientific Committee (2008) received
VMS data over 2007 and 2008 for vessels pot
fishing for crab and longlining for Patagonian
toothfish and other species. There was evidence
of fishing by these vessels in areas closed
54 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
to bottom fishing, including Molloy, Discovery,
Junoy, Shannon, Schwabenland and Herdman
Seamounts. However, it was viewed as possible
that these vessels were fishing for non-SEAFO
species and no further investigation of these
fishing activities has been presented.
The move-on rule
SEAFO adopted a protocol for encounters with
VME species in Conservation Measure 12/08.
This measure adopts the same threshold levels
as the old NEAFC and NAFO threshold: 100kg
of live coral and 1,000kg of sponges. These
levels have already been discussed in relation
to NEAFC and NAFO as having little conservation
value given the current knowledge on distribution
of VMEs in the deep sea. In 2009, these
threshold levels were revised downwards in line
with NEAFC and NAFO to 60kg of corals and
800kg of sponges (SEAFO Commission, 2009;
note: there is a mistake in this report suggesting
60kg of sponges and 800kg of corals). SEAFO
stated that in 2010 the levels will be revised
according to a more rigorous determination of
appropriate threshold levels and VME indicator
species for the SEAFO region. Triggering the
current threshold requires a fishing vessel to
move 2nm away from the end of the trawl tow
or longline set and to report the encounter to
the Scientific Committee, which then makes an
annual assessment of the likely occurrence of
VMEs within the Regulatory Area.
Conclusions
(i) Conduct assessments of whether bottom
fishing activities have SAIs on VMEs.
● SEAFO requires that impact assessments are
undertaken prior to fisheries commencing in
areas currently closed to fishing because of
the risk of SAIs on VMEs.
(ii) To implement measures in accordance with
the precautionary approach, ecosystems
approaches and international law and to
sustainably manage deep-sea fish stocks.
● The level of information for catches of deepsea species on the high seas of the SEAFO
Regulatory Area is sparse because of a failure
of flag states to provide SEAFO with data,
including historical data, on catches, discards
and areas of fishing. This lack of information
is insufficient for effective management goals
and targets and therefore SEAFO has only set
precautionary TACs for many of the species
within the Regulatory Area.
● There is currently no information available
on how effective the regulation of deep-sea
fisheries by SEAFO has been in preventing
overfishing. Several nations fishing in the
area (e.g. South Korea and Japan) have not
ratified the Convention and Japan is currently
exceeding recommended TACs for deep-sea
red crab.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom
fishing where VMEs are known or likely to
occur, or not authorised to proceed.
● SEAFO has acted rapidly to protect some
localities that are likely to host VMEs within
the Regulatory Area.
● SEAFO has also adopted a number of other
measures to protect VMEs and species,
including the banning of gillnetting and
requirements for impact assessments before
any fisheries can commence in closed areas.
● VMS data suggest evidence of non-compliance
of states in respect of SEAFO closed areas.
(iv) To establish and implement protocols to
cease fishing where an encounter with VMEs
occurs during fishing activities and to report
such encounters so that appropriate measures
can be adopted with respect to that site.
● The current move-on rules for SEAFO are
based on high threshold levels unlikely to
trigger a VME-encounter action. No scientific
bases for these threshold levels are given. In
2009, SEAFO reduced the threshold levels in
line with NEAFC and NAFO and have stated
that they will revise these further following an
analyses of appropriate threshold levels for
VME indicator species.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 55
NORTH PACIFIC OCEAN
The North Pacific is the most important
area in the world in terms of marine capture
fisheries with about 24.7 million tonnes of
fish landed in 2006 (FAO, 2009). Important
commercial species in the region include
Japanese anchovy (Engraulis japonicus),
Alaska pollock (Theragra chalcogramma)
and large-head hairtail (Trichiurus lepturus;
FAO, 2009). Until recently no RFMO existed
to regulate fisheries for non-highly-migratory
and non-anadromous species on the high seas
of the North Pacific Ocean. In 2006, Japan,
Russia, South Korea and the USA initiated
negotiations to establish a new RFMO to
regulate fisheries in this area, known as the
North Pacific Fisheries Commission (NPFC).
The new RFMO is still under negotiation,
although interim measures to manage high
seas bottom fisheries in the northwest Pacific
were adopted in 2007 in response to UNGA
Resolution 61/105. No interim measures have
been adopted for bottom fisheries on the high
seas of the northeast Pacific. Vessels from
Japan, South Korea and Russia engage in high
seas bottom fishing in the northwest Pacific.
The high seas area that comprises the NPFC
area under negotiation includes the Emperor
Seamount Chain that extends from the Aleutian
Island Chain in the north 2,000km to the
Hawaiian Ridge (Figs. 36, 37). This area was
one of the first regions of the world to be
subject to deep-sea fisheries that targeted
the seamount-associated slender or pelagic
armourhead, Pseudopentaceros wheeleri.
Initially, large catches of this fish were taken
(133,000t in 1969 by Russia; Sakiura, 1972;
200,000t by Russia and Japan in 1973; Clark
et al., 2007) and catches were maintained at
20,000–30,000t until 1976, when there was a
dramatic decline in the fishery (800,000t taken
in total; Clark et al., 2007). Most catches of
armourhead were taken on Kinmei, Milwaukee,
Colahan and Hancock Seamounts (Clark et
al., 2007). Effort in the fishery then switched
to alfonsino and oreos, although there were
abrupt sporadic increases in armourhead
catches from time to time. Thus the fisheries
have been characterised by a switch between
these species over the last 30 years, although
catches of oreos and especially alfonsino never
approached the landings levels of armourhead.
Initial catches in the alfonsino fishery are not
well recorded, rarely exceeding a few hundred
tonnes annually, but total Pacific catches are
thought to be in the region of 80,000t (Clark
et al., 2007). In the mid-1960s precious corals
were discovered on Milwaukee Bank and along
the Emperor Seamount Chain. A substantial
tangle-net fishery for Corallium secundum was
soon developed by Japan and Taiwan, with an
estimated removal of 150,000kg in the late
1960s; within a few years, however, the fishery
had declined drastically (Humphreys, 2008).
Nonetheless, by 1983 70 percent of the world’s
catch of red coral came from this area; in part
this resulted from the discovery of a slopedwelling Corallium sp. that yielded 200,000kg to
the fishery in 1980 alone (Grigg, 1982). These
tangle-net fisheries were highly destructive, but
this method has been replaced by submersibles
and ROVs in and around Hawaii. Statistics on
the coral fisheries are very poor with respect to
amounts, locations and by-catch, but serious
impacts on VMEs are known to have occurred
(Humphreys, 2008).
seamounts, where it is targeted by bottom
trawl fisheries. The life cycle of armourhead is
unusual. The larvae and post-larvae disperse
away from seamounts in the surface waters of
the temperate and sub-Arctic Pacific and return
after around two years to the seamounts where
they stop growing and reproduce annually,
gradually becoming emaciated before dying
after four to five years (Boehlert & Sasaki,
1988). Thus, seamount populations are
maintained largely through recruitment from
juveniles originating in a different geographic
region. Armourhead have not been assessed
in the North Pacific since the early 1990s
and at present the status of the fish stock is
unknown and insufficient data exist to identify
a sustainable level of exploitation. NOAA
(2008) recommends a new assessment of the
armourhead stock in the North Pacific but the
situation is complicated by the highly episodic
nature of recruitment in this species. There is,
therefore, no current international management
plan for pelagic armourhead.
one year. They recruit back to seamounts and
juveniles are thought to occur above and around
the seamounts, while adults are benthopelagic.
Fish mature at three to four years and live about
15 years. There was a recent assessment of
alfonsino stocks in the North Pacific, Emperor
Seamounts region (Fisheries Agency of
Japan, 2008: Appendix C). This analysis was
complicated by possible changes in the patterns
of fishing, changes in catchability of the species
over time and potential interactions between
alfonsino and armourhead. However, the overall
picture is of a decline of alfonsino stocks over
time (NOAA, 2008). At the end of 2007 the
fishing nations party to the NPFC negotiations
decided to freeze fishing effort to current levels
and to only allow fishing south of 45oN (North
Pacific Fisheries Commission, 2007). Alfonsino
are overexploited in the region and recent
analyses indicate a significant reduction in
fishing effort will be required (Fisheries Agency
of Japan, 2008: Appendix C).
Other target and by-catch species
Alfonsino (Beryx splendens)
This species is considered vulnerable to
overexploitation because it forms aggregations
around seamounts. In the North Pacific region,
alfonsino are thought to spawn in summer
and disperse away from seamounts for about
Reporting on catch of other species by Japan
is very limited and appears to be mainly broad
alfonsino (Beryx decadactylus; Fig. 38), mirror
dory (Zenopsis nebulosa) and cardinalfish
(Epigonus denticulatus and Epigonus
atherinoides). Of these, Beryx decadactylus has
Figure 37.
North Pacific
Fisheries
Commission
region
showing the
position of
fished and
unfished
seamounts.
Management of fisheries for deep-sea
species of low productivity
Fig. 36. NPFC RFMO Map (Zoomed)
Pelagic armourhead (Pseudopentaceros
wheeleri)
This species typically aggregates around
56 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 57
undergone a major decline, while catches of all
the other species have declined, although their
current status is unknown (Fisheries Agency of
Japan, 2008: Appendix F). However, examination
of fish in research trawls from the early 1990s
and those reported from gillnet catches
indicates a high diversity of potential by-catch
species, including deep-water sharks. Species
identified include Allocytus verrucosus, Antigonia
capros, Antigonia spp., Argyropelacus aculeatus,
Benthodesmus pacificus, Brama japonica,
Callionymidae, Chascanopsetta prorigera,
Chaunax sp., Chimaera spp., Chlorophthalmus
sp., Congriscus megastomus, Cookeolus
japonicus, Decapterus tabi, Emmelichthys
struhsakeri, Erilepis zonifer, Etmopterus pusillus
(Fig. 39), Evoxymetopon spp., Helicolenus
spp., Hyperoglyphe japonica, Hoplostethus
crassispinus, Lophiomus miacanthus,
Macrorhamphosus sp., Macrouridae, Malthopsis
spp., Meadia abyssalis, Microstomus shuntovi,
Moridae, Myctophidae, Parabothus coarctatus,
Parapercis spp., Parazen pacificus, Pentaceros
japonicas, Physiculus spp., Plectranthias
kelloggii, Polymixia japonica, Promethichthyus
prometheus, Satyrichthys engyceros, Scorpaena
spp., Sebastidae, Sphoeroides pachygaster,
Squalus mitsukurii, Symphysanodon maunaloae,
Thyrsitoides marleyi, Zenion japonicum
(Fisheries Agency of Japan, 2008: Appendices
A, B, F). Some of these species are mid-water
pelagic fish and not relevant to UNGA Resolution
61/105 and only a few are large enough and
abundant enough to be of any potential interest
to fisheries. Sharks are of particular concern as
they are low-productivity benthic species that,
even if not targeted, are especially vulnerable to
gillnets and longlines, which are commonly used
in the region for fishing. The dominant catch
on the Emperor Seamounts by a South Korean
longliner was 65 percent composed of sharks
(Republic of Korea, 2008).
Figure 38. Broad alfonsino, Beryx decadactylus. © Alex
Rogers
are nonetheless taken in other fisheries as
by-catch (e.g. in the longline fishery for sablefish,
Anoplopoma fimbria; Rodgeveller et al., 2010).
Data are poor, but estimated by-catch of all
grenadiers (mostly giants) from Alaskan waters
has been 10,000–21,000t annually since 1997
(Clausen, 2008).
Figure 39. Deep-water sharks, Etmopterus cf pusillus.
Note, large specimen has squid in mouth. © Alex Rogers
In this region, Russian gillnet vessels target
oreo (Allocyttus verrucosus), mirror dory and
alfonsino, while longliners target rockfish
(primarily Helicolenus spp.), alfonsino, pelagic
armourhead, skilfish (Erilepis zonifera) and
grenadiers (Coryphaenoides spp.). The
Russians also maintain a pot fishery for tanner
(Chioniocetes tanneri), red (Chaceon spp.)
and snow crab (Paralomis spp.). Significant
by-catch from trawl fisheries includes sharks
but there are no data on by-catch from the
gillnet fishery. Longline by-catch species include
escolar (Lepidocybium flavobrunneum), wahoo
(Acanthocybium solandri), dorado (Coryphaena
hippurus), grenadiers (Coryphaenoides spp.)
and codling (primarily Physiculus spp.). Catches
of ‘other’ species are substantial on Russian
vessels, indicating that the fisheries are catching
a mix of species. No real trends can
be identified in catch data.
Biological studies of these fish have only
recently begun, but the fish seem no different
from other grenadiers, i.e. slow-growing, latematuring (age at first maturity 15–36 years),
long-lived (58 years for giants), and hence
can be considered de facto vulnerable and
susceptible to overfishing (Rodgeveller et al.,
2010). There are concerns even now because
the by-catch, which is discarded and presumably
dies, is mostly large females that are segregated
from the males and most abundant at depths
where sablefish and Greenland halibut are
sought (Clausen, 2008). While grenadier stocks
are likely to be found predominantly within EEZs,
it may be possible that in the future a fishery
could develop for these species in the high
seas.
Protection of benthic marine ecosystems
The Pacific is very rich in seamounts but less
than 1 percent have been adequately surveyed
(Humphreys, 2008). Those that have display
great faunal diversity (De Forges et al., 2000).
Japan, South Korea, Russia and the USA (which,
unlike the former three countries, does not
have any vessels bottom fishing on the high
seas) have undertaken impact assessments
for the Emperor Seamounts region. All these
assessments appear to draw heavily from
Japanese work. Seamounts are well-known
localities where VMEs are likely to occur,
including, particularly, corals because of the
availability of hard substrata and the occurrence
of strong currents that carry a supply of food to
suspension-feeding organisms (Rogers, 1994;
Rogers et al., 2007). The Emperor Seamounts
have been well-known for some time as a rich
source of precious corals. However, it is likely
that bottom fishing, including the high seas
bottom drag fishery, which targeted precious
corals in previous decades, will already have
heavily impacted many areas (Humphreys,
2008). The impact assessment undertaken by
Japan for the Emperor Seamounts comprised
ROV surveys, camera drop surveys and an
assessment of snagging points for nets on the
seabed (interpreted as resulting from fishing
gear being caught on corals). Some records of
coral by-catch have also been presented but
these only refer to the presence of coral in
catch and no quantitative data are available.
Figure 41. Koko Seamount showing incidence of net hang ups. Snagging of
nets occurred on Koko Seamount more than any other seamount for which
Grenadiers (Macrouridae): A future fishery?
The seamount fisheries have been of most
interest in the region, but recent interest has
focused on the continental slopes, where two
grenadier species are potentially the target of
significant bottom trawl fisheries in the future.
The popeye (Coryphaenoides cinereus) and giant
(Albatrossia pectoralis) grenadiers are large fish
and both abundant and widespread on the slope
from Alaska through Russian waters to Japan.
The giant grenadier can exceed 2m in length
and 35kg in weight. Its flesh is quite watery, so
it has not met with much favour in the market to
date. Surveys indicate that stock size could be
of the order of 800,000t in the Eastern Bering
Sea; 1,500,000t in the Gulf of Alaska; and
2,000,000t off the Aleutian Islands (Clausen
2008); with perhaps 1,000,000t in the Sea of
Okhotsk (Tuponogov et al., 2008). Because of its
abundance, it is likely that the giant grenadier is
important in the slope ecosystems of the North
Pacific Ocean (Rodgeveller et al., 2010). There is
no directed grenadier fishery in Alaska but they
58 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
there was records.
Figure 40. Koko Seamount showing video survey stations undertaken
by the Fisheries Agency of Japan (2008: Appendix H). Octocoral
gardens were observed at Stations 12 and 15 and Corallium at
Station 11. The proposed protected area is shown in cross-hatch.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 59
During the late 1980s to early 1990s, NOAA
surveys found only sparse patches of octocorals,
presumably the remnants from overfishing
(Humphreys, 2008).
The Fisheries Agency of Japan (2008: Appendix
M) has also reported recent sightings of
Taiwanese vessels fishing for coral, indicating
that the precious coral fishery is still operating
and potentially damaging VMEs present on the
Emperor Seamounts.
Observations
Information from the Japanese ROV and
camera surveys indicates the presence of
octocoral garden communities on the Koko
Seamount, which has historically been the
focus of significant fisheries for precious
corals. Octocoral gardens are classed as
VMEs. The Japanese impact assessment for
trawling states that despite aggregations of
corals existing at Stations 12 and 15 on Koko
Seamount (Fig. 40), it is “not possible to reach
any conclusion they constitute VMEs”. The
assessment notes that the FAO Guidelines on
managing deep-sea fisheries on the high seas
provides no quantitative guidance as to what
constitutes a VME and that the communities
on Koko Seamount do not resemble extremely
high density stylasterid/sponge/bryozoans
communities from the Antarctic. While this
may be true, the Antarctic VMEs comprised
of stylasterids are unusually dense, probably
because of their location on the continental
slope of the Antarctic and the extremely high
seasonal productivity of surface waters there,
and do not constitute any ‘normal’ benchmark
situation. Additionally, the Antarctic areas in the
photographs referred to (Australian Antarctic
Division, 2008) are unlikely to have been fished
with bottom-contact gear. Comparison with coral
garden habitats elsewhere (see discussion
under northeast Atlantic region; Rogers et al.,
in press) suggests that observations on Koko
Seamount do represent VMEs. Data on trawl
hang-ups were also plotted in the Japanese
assessment (Fig. 41). Some of these are
congruent with the coral gardens observed
in the ROV footage of Koko Seamount but
hang-ups can occur for other reasons such as
lodging of the gear on rocks or under ledges or
entanglement with lost fishing gear. The location
of the high coral densities on Koko Seamount
(see photographs in: Fisheries Agency of Japan,
2008: Appendix H; see Fig. 40) follow a pattern
that has been seen on other seamounts: that of
high abundance of organisms close to the edges
of the summit (Rogers, 1994).
The Japanese and other impact assessments
by fishing nations in this region have proposed a
zone protected from fishing on Koko Seamount
to protect the single locality at which Corallium
was observed (Station 11; Fig. 40). This has
little conservation value for the identified
octocorals garden VMEs on the seamount.
Data were also obtained from other seamounts
in the Emporer Seamount Chain, including
Yuryaku, Kammu, Colahan, Jimmu, Suiko,
Showa and Youmei Seamounts. Corals were
present on these but not as abundantly as on
Koko Seamount. However, sampling effort was
extremely low for some sites, comprising just a
few camera drops in some cases. The variability
in coral densities both within a single seamount
and on the different seamounts in this study
is striking. Studies so far are not sufficient
to support the conclusion that there were no
VMEs on other seamounts of the Chain. Some
photographs indicate heavily trawl-impacted
seabed on some of the seamounts investigated.
No other data are presented on the potential
for deep-sea fishing activities to impact benthic
communities on the Emperor Seamounts.
The move-on rule
The fishing nations involved in the NPFC
negotiations initially adopted the NEAFC moveon rule with respect to coral but have lowered
the threshold by-catch limit to 50kg. Points
raised previously in this report with respect to
the move-on rules for NEAFC apply in large part
to the NPFC area. There has been no attempt
to identify VME communities in the region other
than coral communities, and South Korea does
not require its vessels to report encounters with
VMEs.
Conclusions
(i) Conduct assessments of whether bottom
fishing activities have SAIs on VMEs.
● Impact assessments have been undertaken
by Japan, Republic of Korea and Russia for the
Regulatory Area of the North Pacific Fisheries
Commission.
● The impact assessments submitted by
Republic of Korea and Russia appear to draw
heavily on the impact assessment produced by
Japan.
● These assessments have been undertaken
in a region for which there are few data on
60 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
deep-sea benthic ecosystems and where
data on effort and catches for some fisheries
are lacking. Even where there is evidence
of the presence of species associated with
VMEs, interpretation of data has not been
precautionary nor is it in line with studies
elsewhere on what constitutes a VME (see
text and Section (iii) below).
● The impact assessments conclude that in
general SAIs to VMEs do not exist.
(ii) To implement measures in accordance with
the precautionary approach, ecosystems
approaches and international law and to
sustainably manage deep-sea fish stocks.
● The new RFMO is still under negotiation
although interim measures to manage high
seas bottom fisheries in the northwest Pacific
have been adopted.
● The pelagic armourhead fishery has been
severely depleted over the last 40 years yet
there is no stock assessment for the species.
● Alfonsino is overexploited but current
management plans (aimed at maintaining
current levels of fishing effort) do not reflect
an accurate status of the stock.
● For most other species, catch statistics are
unavailable or unreliable and, therefore,
assessment of the effects of fishing mortality
on stocks is not possible. There is no current
plan to change this situation or to plan for
potential grenadier fisheries.
● Overall, impacts on many low-productivity
species, such as sharks, cannot be assessed
on the Emperor Seamount Chain at this time.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom
fishing where VMEs are known or likely to
occur, or not authorised to proceed.
● VMEs are present on the Emperor Seamount
Chain. However, intensive historical bottom
fishing, some targeting precious corals, will
have heavily impacted this and other local
seamounts.
● The fishing nations involved in the NPFC
negotiations have proposed a single protected
area on the Koko Seamount because of the
presence of Corallium at one station. This
protected area does not protect the coral
gardens known to be present elsewhere on
seamount summit edges. Japan and Republic
of Korea have proposed to prohibit their
vessels from engaging in bottom fishing on the
high seas north of 45°N and 40°N latitude,
respectively. Japan further proposes to limit
the allowable depth of bottom fishing to
1,500m.
● Coral fishing is reported as continuing on the
Emperor Seamount Chain.
● Comparison of the Emperor Seamount benthic
communities with those of the Antarctic
continental slope is misleading and does not
reflect the work done to quantify densities of
octocorals in RFMOs elsewhere (e.g. North
Atlantic, northeast Pacific; Stone, 2006;
WGDEC, 2007; Edinger et al., 2009; Rogers et
al., in press).
● The seamounts investigated are likely to
have been heavily impacted by fishing. Where
remnant populations of corals and other
VME species exist, area closures should
be established to allow for some degree of
regeneration.
● Current impact assessments are not adequate
to identify VMEs along the fished seamounts
of the Emperor Seamount Chain and there
have been no analyses of fisheries data to
identify where fishing activities are taking
place on fished seamounts. Given the lack
of data on fishing activities in general, such
assessments are impossible.
● Interim measures consistent with UNGA
Resolutions 61/105 and 64/72 are needed
for the northeast Pacific.
(iv) To establish and implement protocols to
cease fishing where an encounter with VMEs
occurs during fishing activities and to report
such encounters so that appropriate measures
can be adopted with respect to that site.
● The threshold levels set by NPFC for VME
encounters apply to corals only.
● The threshold level for corals do not take
into account the small size and delicate
morphology of coral colonies observed on the
seamounts.
● Using the same threshold levels for active and
passive fishing gears does not reflect large
differences in their impact.
● Differentiating the post-VME-encounter
protocol between areas with a fishing
history and those without does not serve
conservation objectives.
● The 2nm move-on rule is an ineffective means
of conserving deep-sea species because it
is difficult to identify where a VME encounter
occurs along a tow for commercial bottom
trawling.
● Some states (e.g. South Korea) are not
reporting VME encounters even when a VMEencounter protocol is in operation in the
RFMO.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 61
SOUTH PACIFIC OCEAN
The southeast and west-central Pacific is one
of the most important areas of the world in
terms of global fish catches, mainly as a result
of large pelagic fisheries (FAO, 2009b). The
area is geographically vast and only recently
have the high seas fisheries for deep-water
species become subject to management
measures. In May 2007, the countries involved
in negotiating an RFMO in the region adopted
a set of interim measures to implement UNGA
Resolution 61/105. In November 2009 an
agreement to establish the new RFMO, the
South Pacific Regional Fisheries Management
Organisation, was adopted; the RFMO will
be established once countries ratify the
Convention (SPRFMO; Fig. 42). To date, New
Zealand, the Cook Islands, Chile, Columbia and
Peru have signed the Convention out of the 32
states that have participated in consultations
related to the establishment of SPRFMO.
The total reported deep-sea fish catch in the
southern Pacific in 2004 was 426,112t (EEZ and
high seas), about 7 percent of the world’s total
catch (Sissenwine & Mace, 2007). However, this
includes most of the global catch for orange
roughy (Hoplostethus atlanticus), a long-lived
species that is fished generally in aggregations
over seamounts and ridges (Sissenwine & Mace,
2007). Most of the high seas bottom fishing in
the SPRFMO Regulatory Area in recent years has
taken place in the southwestern Pacific, mainly
by Australian and New Zealand vessels, with
the catch averaging several thousand tonnes
per year (Bensch et al., 2008). These high seas
fisheries have mainly occurred in association
with the seamounts of the Norfolk Ridge, the
Northwest Challenger Plateau, the Lord Howe
Rise, the Louisville Ridge and, to a lesser
extent, the Three Kings Ridge and South Tasman
Rise (Clark et al., 2007; SPRFMO, 2007a;
Government of New Zealand, 2009). Some
fishing has taken place in the southeastern
Pacific on the Nazca and Sala Y Gomez Ridges
(Clark et al., 2007). The fisheries have been
conducted mainly with bottom trawls. With the
depletion of deep-sea stocks, and for market
reasons, there has been a shift away from
trawling towards line fishing by some states,
notably New Zealand (Government of New
Zealand, 2009).
Management of fisheries for deep-sea
species of low productivity
Orange roughy (Hoplostethus atlanticus)
The southwest Pacific is the main area where
orange roughy, the iconic species of deep-water
fishing, are caught. The orange roughy has a
very low productivity as a result of its extreme
longevity (isotopic age validation up to 150
years; Andrews & Tracey, 2007), slow growth
rate in relation to size, and late onset of maturity
Figure 42. Map of South Pacific showing SPRFMO Regulatory Area
(still under review; Bensch et al., 2008).
(Rogers, 1994; SPRFMO, 2007a). Orange roughy
breed in aggregations over seamounts but also
demonstrate lengthy periods of low recruitment
to populations, sometimes lasting 10–20 years
(Koslow & Tuck, 2001; Francis & Clark, 2005).
Overall, the extremely conservative life history
of orange roughy reflects low rates of natural
mortality and adaptation to life on seamounts
and other deep-sea habitats. These same
life-history characteristics render this species
very vulnerable to overfishing, especially as it
is easily targeted by modern fishing vessels
when forming aggregations over elevated
topographic features. Other seamount species
also show similar characteristics and modeling
studies have demonstrated that they are more
vulnerable to overfishing than non-seamount
species (Morato et al., 2006; Morato & Clark,
2007).
Orange roughy fisheries have typically followed
a boom-bust pattern globally, with examples
including stocks off Namibia, the southwest
Indian Ocean, and Australia (Branch, 2001;
Lack et al., 2003). In some cases, serial
depletion has occurred, an example being the
Chatham Rise within the New Zealand EEZ where
stocks were successively discovered on small
seamount features, heavily fished and then
depleted as the fishing fleet moved eastwards
searching for new aggregations (Clark, 1999).
Orange roughy fishing peaked in the 1990s and
has since declined (SPRFMO, 2007a). Most
catches came from the Lord Howe Rise and the
Northwest Challenger Plateau, although more
recent fisheries have developed on the Norfolk,
Three Kings and Louisville Ridges. The status
of the high seas stocks of orange roughy in this
region are uncertain and are likely to vary. The
Tasman Sea fisheries for orange roughy are
depleted. Non-standardised CPUE on the Lord
Howe Rise, Northwest Challenger Plateau and
the Louisville Ridge has declined significantly
(Clark, 2004). Elsewhere in the South Pacific,
orange roughy are fished within Chile’s EEZ.
These stocks are also currently overfished
(SPRFMO, 2007a) and were closed to fishing
except for research purposes in 2006 (Clark,
2009).
At present there is no management in place for
high seas stocks of orange roughy, apart from
those in the South Tasman Rise region, where
the fishery is subject to a bilateral arrangement
between Australia and New Zealand to limit
catches. At present the stock status of high
seas populations of orange roughy is unknown,
62 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
so it is unclear whether current levels of
exploitation are sustainable (SPRFMO, 2007a).
Oreos: Black oreo (Allocyttus niger),
smooth oreo (Pseudocyttus maculatus),
spikey oreo (Neocyttus rhomboidalis),
warty oreo (Allocyttus verrucosus), family
Oreosomatidae
In the South Pacific region, oreos occur on the
continental slopes of Australia, New Zealand and
Chile, the Tasman Sea, the Louisville Ridge and
the southern Chatham Rise (SPRFMO, 2007b).
Oreos occur in deep water, close to the seabed,
and are often associated with topographic
features such as pinnacles and canyons. Like
orange roughy, these species aggregate around
submarine features, making them easy targets
for trawlers (SPRFMO, 2007b). They were caught
as by-catch in fisheries for orange roughy but are
now targeted themselves. Like orange roughy,
oreos are extremely long-lived and slow growing,
with ages up to 150 years or more (black oreo;
Smith & Stewart, 1994; Doonan et al., 1995),
as estimated by counts of otolith rings. Genetic
studies indicate that these fish form discrete
populations on large-scale topographic features
such as the New Zealand and Australian slopes
and also at smaller spatial scales.
The major fisheries for oreos in the high
seas include the South Tasman Rise, the
West Norfolk Ridge, the Lord Howe Rise, the
Northwest Challenger Plateau and the Louisville
Ridge. The status of high seas stocks of oreos
are currently uncertain but are likely to vary
(SPRFMO, 2007b). Catches have dropped
markedly in recent years on the South Tasman
Rise (Clark et al., 2007). At present there are no
estimates of stock size in areas beyond national
jurisdication and no management measures in
place for oreos, with the exception of a bilateral
arrangement by Australia and New Zealand with
respect to the South Tasman Rise (SPRFMO,
2007b). Catches are monitored by Australia and
New Zealand for their vessels.
Figure 43. Spikey oreo, Neocyttus rhomboidalis. © Alex
Rogers
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 63
Black cardinalfish (Epigonus telescopus) and
other cardinalfish, family Epigonidae.
Black cardinalfish are found throughout the
Atlantic Ocean and in the Indian Ocean and
southwestern Pacific (SPRFMO, 2007c). The
species is long-lived, with ages being reported
at over 100 years, but commercial catch ages
are generally between 35 and 55 years. The
species is extremely slow growing and does
not mature until it is 40–50cm in length, with
recruitment at 45 years of age. The species is
benthic or bentho-pelagic, forming schools at up
to 150m above the seabed, particularly around
hills or rough seabed topography. Information on
the biology of this species in the South Pacific
region is extremely limited.
Black cardinalfish are taken as by-catch in
fisheries for orange roughy, as with oreos, and
also alfonsino. The largest catches have come
from the northern Challenger Plateau and the
Lord Howe Rise. There is no information on
the status of stocks of black cardinalfish on
the high seas and fisheries are unmanaged.
Some experimental trawl fisheries in the 1970s
on the Louisville and Geracyl Ridges caught
other cardinalfish species, including Epigonus
pectinifer, Epigonus denticulatus, Epigonus parini
and Epigonus geracleus, and estimates at that
time suggested substantial stocks in these
areas (Clark et al., 2007). There was some
fishing on the Geracyl Ridge in the 1970s and
early-1980s and the Louisville Ridge has been
targeted for other species (Clark et al., 2007).
Goldeneye perch or alfonsino (Beryx
splendens)
As described previously, this species is
vulnerable to overfishing as a result of its
aggregating behaviour. In the South Pacific,
alfonsino are found on outer continental shelves,
the slope and on ridges and seamounts. The
majority of catches of this species during the
period 1969–2004 came from the South Pacific,
although most of the catches were from inside
EEZs (SPRFMO, 2007d). In the southwestern
Pacific significant catches of alfonsino have been
taken in high seas areas (SPRFMO, 2007d),
including on the Louisville Ridge (Clark et al.,
2007). There are few data on the alfonsino
stocks that are fished on the high seas and there
are no regulatory measures in place to manage
fisheries in the region (SPRFMO, 2007d).
Bluenose (Hyperoglyphe antarctica)
Bluenose are found across the southern Atlantic,
southern Indian Ocean and southwestern
Pacific. They are found over rough ground from
200–750m depth and are often associated
with seamounts (SPRFMO, 2007e). They live to
about 25 years old and mature at 7–12 years of
age. The species is caught using both bottom
and mid-water trawls as well as a variety of line
gear. Only a small proportion of the current
catch comes from high seas areas (SPRFMO,
2007e). Bluenose are an aggregating species, a
behaviour that can result in an apparently stable
CPUE over several years before a sudden decline
in catches from overexploitation (Government of
New Zealand, 2009).
Foundation lobster (Jasus caveorum)
This lobster is known only from the Foundation
Seamounts and has been fished sporadically
(SPRFMO, 2007f). Other fisheries for Jasus spp.
on seamount localities have resulted in rapid
depletion of stocks. No management is in place
for this species on the high seas. Other lobster
fisheries also probably take place in the South
Pacific Ocean but there is very little information
available on these fisheries (SPRFMO, 2007f).
Other species
A number of other species have been subject
to targeted fishing or are taken in the high
seas area of the South Pacific (Clark et
al., 2007; Government of New Zealand,
2008b, 2009). These include pink mao
mao (Caprodon longimanus), armourhead
(Pseudopentaceros richardsoni and Pentaceros
japonicus), ruby snapper (Etelis carbunculus
and Etelis coruscans), southern blue whiting
(Micromesistius australis), grenadiers (e.g.
Caelorhinchus australis), ribaldo (Mora moro),
giant boarfish (Paristiopterus labiosus), bass
or hapaku (Polyprion oxygeneios, Polyprion
americanus), tarakihi (Nemadactylus spp.),
gemfish (Rexea spp.), kingfish (Seriola lalandi),
toothfish (Dissostichus spp.), rock cod
(Helicolenus spp.), red snapper (Centroberyx
affinis) and sharks (e.g. Dalatias licha, Squalus
acanthias, Galeorhinus galeus). Data on
catches of these species are only available
from some states and include both bottom
trawl and longline fisheries. In some cases,
catches of these species may be very small.
No specific management measures are in place
by SPRFMO for any of these fisheries on the
high seas at the present time (but see New
Zealand impact assessment below). Other
species taken as by-catch from high seas deepsea fisheries include a variety of sharks, rays,
chimaerids and a number of teleost species but
detailed information on catches is not available
64 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
(SPRFMO, 2007a). Few of the species reported
are presently of commercial interest.
Protection of benthic marine ecosystems
Prior to the SPRFMO negotiations there were no
protected areas in the high seas of the South
Pacific. Following UNGA Resolution 61/105,
SPRFMO Contracting Parties agreed to a set of
interim measures to implement the resolution
in May 2007, including an agreement to ‘freeze
the footprint’ of existing high seas bottom
fisheries until 2010 (SPRFMO Interim Measures,
2007). They further developed an Interim
Benthic Assessment Framework, followed by
a Draft Bottom Fishery Impact Assessment
Standard (DBFIAS). These measures established
standards for environmental impact assessment
of deep-sea fisheries on the high seas and
included consideration of the move-on rules
for fishing vessels. The DBFIAS has been
criticised by some member states (e.g. Chile),
while others have adopted their own move-on
rules along with their impact assessments
(e.g. Spain). Efforts to establish and act upon
environmental impact assessments by New
Zealand have met with opposition from the
deep-water fishing industry (Government of New
Zealand, 2009).
Pending the adoption of the final Bottom Fishery
Impact Assessment Standard, the DBFIAS
serves as the standard for impact assessments
for all bottom fisheries in the SPRFMO
Regulatory Area down to 2,000m, based on the
assumption that the deepest depths fished were
1,500m (SPRFMO, 2008). However, it is known
that fishing now takes place down to 2,200m in
areas such as the Antarctic, so that assumption
is not correct on a global scale. The interim
measures adopted in 2007 required all member
states to prepare a benthic impact assessment
of their bottom fisheries regardless of scale or
previous fishing history (SPRFMO, 2008). These
assessments were required prior to bottom
fishing activities taking place. So far only New
Zealand and, more recently, the European Union
have submitted impact assessments to the
SPRFMO Science Working Group, and these
have been placed on the SPRFMO website for
comment. SPRFMO outlined the content of these
assessments as follows.
Details of proposed fishing activity:
● description of vessels used;
● description of the proposed fishing method,
including a gear plan;
● depth-range to be fished;
● target species and potential by-catch;
● period of intended fishing;
● effort (number of vessels, number of tows,
expected tow duration);
● estimated catch and discard of target and
by-catch species.
Mapping of the intended fishing area:
● maps of the intended fishing area;
● mapping of VMEs, potential VMEs or areas
likely to support VMEs in the intended fishing
area;
● any other information useful in assessing likely
impacts of the fishery.
Scoping of issues of concern:
● potential impact of the fishing activity,
including all gear types;
● the risk of loss of fishing gear.
Assessment of:
● intensity or severity of impacts;
● how long the impacts are likely to last;
● spatial extent of impact compared to the
spatial extent of the VME;
● cumulative impact.
Overall assessment of the risk
Interactions with VMEs:
● what interactions will occur between the
fishing gear used and VMEs;
● what is the probability of interaction, its likely
extent and its magnitude;
● what are the characteristics of seabed
habitats likely to be impacted;
● what is the diversity of the fished ecosystem
and likely impacts on diversity of the fishing
activity;
● what is the spatial scale, duration of impact
and cumulative impacts;
● are there any other threats associated with the
proposed fishing plan?
Status of deep-water stocks to be targeted:
● intended target and likely by-catch species;
● historic catches and catch trends in the area
to be fished;
● trends in CPUE in target and likely by-catch
species;
● results of any surveys on stocks to be
targeted;
● results of any stock assessments, if they
exist.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 65
Following this assessment, where a medium
or high risk of impacts of intended fisheries
on VMEs, diversity, or target/by-catch species
is found to exist, a plan of monitoring and
mitigation measures is required. This plan should
include:
● details of methods of collection of VMS data;
● details of the catch and effort data collection
systems;
● details of observer coverage;
● details of any other information provided;
● proposed mitigation measures to prevent or
reduce adverse impacts on VMEs;
● proposed management measures such as
implementation of move-on rules.
New Zealand presented the first comprehensive
impact assessment of its deep-sea bottom
fisheries in the SPRFMO Regulatory Area. The
report follows the DBFIAS guidelines closely
and uses a number of novel approaches, as
well as those adopted by other RFMOs, in
implementing UNGA Resolution 61/105 and
the interim measures agreed by the SPRFMO
process. Under the latter, all countries agreed
to ‘freeze the footprint’ of their bottom fishing
activities until 2010. The footprint was defined
as geographic areas measuring 20 by 20 minute
latitude and longitude ‘blocks’ of ocean space
(a footprint of approximately 1,000km2 in New
Zealand’s case), within which any bottom fishing,
including even a single tow of a trawl net, had
occurred during the period 2002–06.
One novel aspect of the New Zealand impact
assessment was analysis of trawl records
to identify the blocks that have been heavily,
moderately or lightly fished by New Zealand
vessels bottom trawl fishing in the SPRFMO
Regulatory Area (Penney et al., 2009).
Altogether, New Zealand identified 200 such
blocks, with much of the New Zealand fishing
effort in the high seas having been directed
towards seamounts. Heavily trawled blocks
in the SPRFMO area tend to be located over
these features, which have been the source
of most of the deep-water catches in the area.
New Zealand has closed all ‘lightly’ trawled
blocks to bottom fishing, totaling 62 blocks or
approximately 31 percent of the New Zealand
footprint, thus protecting these areas (Penney et
al., 2009). In addition, New Zealand has closed
20 ‘representative’ areas of the remaining 138
blocks of moderately and heavily fished areas,
bringing the total closed area to approximately
40 percent of the footprint or some 40,000km2
(Government of New Zealand, 2009; Penney
et al., 2009). Overall, these closures add up
to protection of a significant area of the deep
seabed that was potentially subject to continued
fishing by New Zealand vessels. New Zealand
considers this its primary tool in protecting deepsea ecosystems and, furthermore, potentially
useful for protecting not only VMEs but also nontargeted by-catch species such as sharks.
New Zealand has also designed a move-on
rule for fishing vessels that encounter VMEassociated species. The determination of the
threshold values is based on an approach
similar to NAFO’s approach of examining
by-catch accumulation curves, in this case
with data from commercial trawlers taken by
observers rather than from fisheries’ survey
trawls (Parker et al., 2009). The threshold value
was taken from an arbitrary cut-off value of the
50th percentile from the biomass accumulation
curve of by-catch (threshold = 30kg for stony
coral, 50kg for sponges and less for other coral
classes; Parker et al., 2009). Determination of
a potential VME encounter considers whether
the threshold is exceeded as well as the number
of VME-associated taxa that are encountered,
regardless of weight, to take into account
impacts on species-diverse habitats (Parker et
al., 2009). Such an approach was only possible
because of the provision of detailed observer
data on by-catch on New Zealand vessels.
Together, the move-on rule and closed areas
represent a serious attempt to implement
the UNGA Resolution 61/105 and the FAO
Guidelines on management of deep-sea
fisheries. However, the high seas areas that
remain open to continued bottom trawl fishing
by New Zealand vessels may contain significant
areas of VMEs. The move-on rule is not applied
to heavily trawled blocks. The New Zealand
government’s view is that such areas are open
to fishing and that VMEs are protected by having
existing areas closed to bottom fishing. It
remains to be seen how effective the measures
adopted by New Zealand will be, particularly if
other states allow their vessels to fish in the
areas closed by New Zealand. New Zealand also
proposes to freeze current catches of deepsea species such as orange roughy. However, it
acknowledges that this freeze in catches, based
on the 2002–2006 figures for catch, is likely
to exceed sustainable levels of exploitation for
species such as orange roughy in the SPRFMO
Regulatory Area. It proposes to undertake stock
assessments of such deep-sea species but
66 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
also points out that the implementation of such
analyses will require international cooperation.
In the meantime, fishing of low-productivity
species such as orange roughy will continue
while a scientifically based harvest plan is being
developed.
Spain (the EU) has also submitted an impact
assessment on its gillnet fisheries in the
SPRFMO Regulatory Area. The document lists
the vessels it is fishing, the intended target
species and the method and depth of fishing
(gillnets; Government of Spain, 2008). It also
outlines a VME-species encounter protocol
that is identical to the old NEAFC and NAFO
thresholds of 1,000kg of live sponges and
100kg of live coral (see previous discussion
on higher threshold limits) and is at variance
with recommendations by SPRFMO and the
government of New Zealand. Spain states that
the by-catch of gillnetting vessels, which are
fishing on similar features to the New Zealand
fleet, including the Challenger Plateau, is
insignificant in terms of VME taxa and that its
fishing operations have low or no impact. The
effects on target or non-target fish species by
gillnets in the SPRFMO area are not considered
in the Spanish impacts assessment and neither
are precautionary management or mitigation
measures. The November 2009 International
Meeting of SPRFMO adopted a resolution to
ban deep-water gillnet fishing in the SPRFMO
Regulatory Area.
Conclusions
(i) Conduct assessments of whether bottom
fishing activities have SAIs on VMEs.
● Only two states (New Zealand and Spain)
have submitted impact assessments of their
bottom fisheries in the South Pacific. None of
the other states whose vessels have engaged
in bottom fishing in the region have submitted
impact assessments to the SPRFMO Science
Working Group.
● The two impact assessments that have been
carried out vary markedly in quality but both
propose that fishing takes place on stocks of
deep-sea fish species that are not subject to
management, i.e. they are unmanaged.
● The New Zealand impact assessment includes
most of the information required by SPRFMO.
(ii) To implement measures in accordance
with the precautionary approach, ecosystems
approaches and international law and to
sustainably manage deep-sea fish stocks.
● High seas deep-sea fisheries in the region
target low-productivity species using bottom
trawls, longlines and traps.
● At present, no management measures are
in place for high seas deep-sea bottom
fisheries on target or by-catch species, with
the exception of bilateral agreements between
Australia and New Zealand for fisheries on the
South Tasman Rise.
● Information on stock size, distribution,
abundance, catch, and the impact of fishing
for most of the deep-sea species taken in the
high seas bottom fisheries in the region is
limited.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom fishing
where VMEs are known or likely to occur, or not
authorised to proceed.
● New Zealand has established novel
measures, including precautionary closures of
approximately 40 percent of the area of the
seabed within its historic fisheries ‘footprint’
to deep-sea fishing by its vessels. However,
the remaining 60 percent of the areas open
to fishing within the New Zealand bottom trawl
fisheries footprint has not been subject to an
impact assessment consistent with the FAO
Guidelines.
(iv) To establish and implement protocols to
cease fishing where an encounter with VMEs
occurs during fishing activities and to report such
encounters so that appropriate measures can be
adopted with respect to that site.
● Both New Zealand and Spain adopted
thresholds for the triggering of a move-on
action for its deep-sea fishing vessels on
encountering VME species.
● The New Zealand rules included threshold
weights for VME species as well as the
diversity of VME species within a catch.
However, the move-on rules were only
applicable to moderately fished or exploratory
fisheries and areas that have historically been
heavily fished will not be subject to move-on
rules.
● The Spanish move-on rules adopted an
encounter protocol identical to the old NEAFC
and NAFO encounter rules. The limited
conservation value of such encounter rules
are discussed in the NEAFC and NAFO
sections of this report.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 67
SOUTHWEST INDIAN OCEAN
The Indian Ocean is globally important for
marine capture fisheries, representing more
than 10 percent of global catches, with the
western Indian Ocean most notable for recent
increases in catches (FAO, 2009b). However,
it is also the region of the world where the
highest proportion of exploited fish stocks
are of unknown or uncertain status (Kimani
et al., 2009), reflecting problems in fisheries
management and ocean governance. Artisanal
fisheries in the Indian Ocean are critical for
the livelihoods and food security of people
in coastal states, particularly island nations
such as the Seychelles. However, there is
evidence that fish catches by the artisanal
sector are grossly under-reported by a factor
of up to five times the FAO statistics. The
offshore fisheries of the western Indian Ocean
are rich but countries within the region have
been unable to develop the infrastructure to
exploit them. Distant-water fishing fleets of
developed countries have gained access to
fish resources through multilateral or bilateral
agreements (Kimani et al., 2009). This
situation is exacerbated by the subsidies to
foreign distant-water fleets, which give them a
competitive advantage over local fishing fleets
(Kimani et al., 2009).
At present, two main agreements exist for the
southern Indian Ocean: the Southwest Indian
Ocean Fisheries Commission (SWIOFC; Fig. 44)
and the South Indian Ocean Fisheries Agreement
(SIOFA; see Fig. 45). SWIOFC was initiated in
2004 to promote sustainable utilisation of
marine living resources and was signed by the
Comoros, France, Kenya, Madagascar, Mauritius,
the early 1980s (Romanov, 2003; Clark et al.,
2007). These fisheries targeted shallow-water
redbait (Emmelichthys nitidus) and rubyfish
(Plagiogeneion rubiginosum), with catches
peaking about 1980 and then decreasing into
the mid-1980s (Clark et al., 2007). Fishing then
switched to the deeper-living alfonsino (Beryx
splendens) in the 1990s as new seamounts
began to be exploited.
Figure 44. SWIOFC’s proposed area of competence
(SWIOFC, 2005).
Mozambique, Seychelles, Somalia, and Tanzania.
At present, SWIOFC is investigating new fisheries
for deep-water species within the EEZ of
Mauritius or Mauritian dependencies (Nazareth
and St Brandon Banks; SWIOFC, 2009). SIOFA
was opened in 2006 and signatories so far
include Australia, the Comoros, France, Kenya,
Madagascar, Mozambique, Mauritius, New
Zealand, Seychelles and the European Union.
SIOFA forms the basis of a regional RFMO for
the management of deep-sea fisheries on the
high seas, but that has not yet entered into
force. Delay in the implementation of the SIOFA
agreement caused sufficient concern among
several deep-water fishing companies operating
in the region for them to form an association
in 2006 to promote technical, research and
conservation activities to provide the future
RFMO with data required for management of
deep-water fisheries (Shotton, 2006). This
association is known as the Southern Indian
Ocean Deepwater Fishers’ Association (SIODFA),
formed by four companies with four deep-water
trawlers flagged to Australia, the Cook Islands
and Mauritius.
Management of fisheries for deep-sea
species of low productivity
Figure 45.
SIOFA area of
competence..
The development of deep-sea fisheries in the
high seas of the Indian Ocean were undertaken
by distant-water fleets of developed countries,
particularly the Soviet Union, which in the early
1970s maintained the largest distant-water
fishing fleet in the world (Romanov, 2003).
Exploratory fishing on the Southwest Indian
Ocean Ridge, the Mozambique Ridge and the
Madagascar Ridge began in the 1970s by the
Soviet fleet and commercial trawling began in
68 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Figure 46. Pelagic armourhead, Pseudopentaceros
richardsoni (top), and alfonsino, Beryx splendens (bottom),
from the Southwest Indian Ocean Ridge. © Alex Rogers.
In the late 1990s a new fishery developed
on the Southwest Indian Ocean Ridge, with
trawlers targeting deep-water species such as
orange roughy (Hoplostethus atlanticus), black
cardinalfish (Epigonus telescopus), southern
boarfish (Pseudopentaceros richardsoni; Fig.
46), oreo (Oreosomatidae) and alfonsino (Clark
et al., 2007). This fishery rapidly expanded, with
estimated catches of orange roughy in the region
of 10,000t, but then rapidly collapsed (Gianni,
2004). Fishing has shifted to the Madagascar
Plateau, Mozambique Ridge and Mid-Indian
Ocean Ridge, targeting alfonsino and rubyfish
(Clark et al., 2007).
Fishing continues along the Southwest Indian
Ocean Ridge, mainly targeting orange roughy
and alfonsino. Recent fishing has also taken
place on the Broken Ridge (eastern Indian
Ocean), Ninety- East Ridge, possibly the Central
Indian Ridge, the Mozambique Ridge and
Plateau and Walter’s Shoal (western Indian
Ocean), where a deep-water fishery for lobster
(Palinurus barbarae) has developed (Bensch et
al., 2008). The banks around Mauritius, within
the EEZ and high seas portions of the Saya da
Malha Bank, have been targeted by fisheries
for shallow-water snappers (Lutjanus spp.) and
emperors (Lethrinidae; SWIOFC, 2009). A new
longline fishery has developed in the northwest
Indian Ocean, mainly by Chinese vessels
targeting deep-water longtail red snapper (Etelis
coruscans; Bensch et al., 2008). There are also
reports of unmanaged gillnet fishing for sharks
in areas of the southern Indian Ocean such as
Walter’s Shoal (Shotton, 2006). New deep-water
fisheries are developing off India, although at
present it is unclear whether the latter is within
or outside the EEZ. SIODFA reports that its
vessels undertake approximately 2,000 deepwater trawl tows per year in the entire Indian
Ocean. By-catch of fish from SIODFA fishing
operations in the region is reported to be small,
especially when fishing below 500m depth
(Shotton, 2006). As with New Zealand vessels
operating in the southern Pacific Ocean, tow
times were typically short, with a duration of
10–15 minutes (Shotton, 2006), reflecting the
highly-targeted nature of roughy and alfonsino
fisheries on seamounts.
Currently, little or no information is available
for the assessment of the impacts of deep-sea
fishing on high seas areas of the Indian Ocean
on populations of either target or
by-catch species. Few scientific surveys have
been undertaken in deep water. What little
information there is suggests that the dominant
slope-dwelling grenadiers in sub-tropical regions
are rather small (Gil et al., 2008). Reporting
of data from commercial fleets is complicated
by issues of confidentiality in those fisheries
where stocks may be located across a wide
area (e.g. the Southwest Indian Ocean Ridge)
and there is no RFMO in force to regulate
fishing. At present, new fisheries are developing
in the region with no apparent assessment of
resource size or appropriate exploitation levels
to ensure sustainability of fisheries. SIODFA has
reported that it is collecting data on both fishing
operations and catches (tow by tow data), as
well as other biological information on target
species, to feed into a future arrangement
(SIOFA) when it is implemented (Shotton, 2006).
Protection of benthic marine ecosystems
At present, the only initiative protecting VMEs in
the high seas region of the Indian Ocean is the
unilateral declaration by SIODFA of 11 Benthic
Protected Areas (BPAs). The companies that
belong to SIODFA have voluntarily closed these
areas to bottom fishing or mid-water trawling
(Shotton, 2006). The BPAs were selected on the
basis of a number of criteria including:
● representivity of seabed type (e.g. seamount,
slope edge, etc.);
● fishing history;
● level of pre-existing knowledge concerning
geology, bathymetry and biology;
● protection of benthic communities;
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 69
● protection of areas of special scientific
interest (e.g. geological features of Atlantis
Bank).
Ten areas in the Indian Ocean were designated
by SIODFA as BPAs (the eleventh is located
in the southeast Atlantic) on the basis of the
knowledge gathered by the members of the
association from various sources as well as the
research and data gathered during commercial
fishing operations. These sites include a
number of seamounts, knolls, ridges and other
topographic features that in some cases are
known or suspected to host VMEs as well as
populations of commercial and non-commercial
fish species (see Fig. 47).
At present little is known concerning the
representivity of the BPAs or whether they offer
protection from bottom fishing. Non-members
of SIODFA are under no legal obligation to avoid
fishing these areas. Currently, a collaborative
international scientific project is underway to
investigate the Southwest Indian Ocean Ridge
and Walter’s Shoal. This project, funded by the
Global Environment Facility and the UK’s Natural
Environment Research Council, will investigate
the ecology and biodiversity of benthic and
pelagic ecosystems, including observations
of birds and cetaceans, associated with five
seamounts, from the Atlantis Bank in the north
to the Coral Seamount in the south. The project
will provide direct observations on the nature of
seabed communities there and how they change
along the ridge. Other information will include
acoustic data on pelagic biomass and records
of birds that are potentially at risk during fishing
operations (particularly longline fishing); both are
data deficient in the region (Shotton, 2006).
BPAs protect a very small area of the seabed
(Figs 47 and 48), estimated to represent
approximately 6 percent of the seamounts at
fishable depths in the region (MCBI, 2009a).
Furthermore, models of habitat suitability in the
Indian Ocean for deep-sea stony corals indicate
that the BPAs are only likely to protect a small
proportion of seamounts that may host VMEs
(see Fig. 48), but useful observations of benthic
communities are sparse. Possible designs for
a representative network of marine-protected
areas on the high seas in the Indian Ocean
need to be evaluated on the basis of increasing
knowledge of deep-sea ecosystems and current
ideas regarding the area and distribution of
protected areas that have conservation value.
impact assessments have been carried out
for deep-sea fisheries in the high seas of the
Indian Ocean.
● The Cook Islands have published information
about vessels authorised to fish in the South
Indian Ocean on the UN FAO website, as called
for in UNGA Resolution 61/105 (paragraph
87) but have not published any information
on impact assessments or conservation
measures adopted with respect to their
flagged vessels; no other country currently
fishing in the region has published any
information whatsoever (FAO, 2010).
(ii) To implement measures in accordance
with the precautionary approach, ecosystems
approaches and international law and to
sustainably manage deep-sea fish stocks.
● Deep-sea fish resources in the high seas
regions of the Indian Ocean have been
severely overexploited in the past.
● In the absence of a RFMO or interim
management measures, as called for in UNGA
Resolution 61/105, deep-sea fisheries on the
Indian Ocean are ongoing and unmanaged,
with the exception of individual state reporting
requirements for some deep-sea fishing
vessels.
● At present, there is little information on
present deep-sea fisheries within the region
in respect of catches of target and by-catch
species or impacts on VMEs or on sustainable
Conclusions
(i) Conduct assessments of whether bottom
fishing activities have SAIs on VMEs.
● No RFMO is in operation in this region, nor
have the flag States whose vessels engage
in bottom fisheries on the high seas region
agreed to or implemented interim measures
for the management of the fisheries, as called
for in UNGA Resolution 61/105. Therefore, no
levels of exploitation for fished stocks. It is,
therefore, not possible to assess the status of
any stocks or species, as has been done for
other localities and RFMOs.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom fishing
where VMEs are known or likely to occur, or not
authorised to proceed.
● The only protected areas are voluntary BPAs
declared by SIODFA. These do not provide
legal protection from fishing activities by
companies outside SIODFA.
● The BPAs have been set up on the basis of
best current knowledge of benthic ecosystems
of the Indian Ocean by the fishing industry.
This information, however, is extremely limited,
so the BPAs only cover a small percentage of
the seamounts at fishable depth in the region
and the conservation value of the BPAs is
unknown.
(iv) To establish and implement protocols to
cease fishing where an encounter with VMEs
occurs during fishing activities and to report such
encounters so that appropriate measures can be
adopted with respect to that site.
● There are currently no encounter protocols in
operation for vessels bottom fishing in deep
water on the high seas of the Indian Ocean.
Figure 48. Habitat
suitability modeling
for stony corals on
the seamounts in the
Figure 47. Map of the
southwest Indian Ocean.
Indian Ocean showing high
As can be seen, the
seas areas; seamounts
BPAs do protect areas of
<2,000m summit depth
suitable habitat, but many
(green dots), seamounts
other areas lie outside
>2,000m depth (red dots)
the protected zones (John
and BPAs (John Guinotte,
Guinotte, Ph.D., MCBI,
Ph.D., MCBI, 2009a).
2009b).
70 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 71
SOUTHERN OCEAN
The Southern Ocean comprises about 6.5
percent of the world’s oceans and is defined
as having a northern boundary at a latitude
of 60oS (Earle & Glover, 2009). However,
it is physically bounded by the Antarctic
Convergence, a major frontal system that
varies in its position but which can be located
as far north as 45oS. This zone, marked by a
steep gradient in temperature, separates the
frigid Antarctic Circumpolar Current from the
warmer Atlantic, Indian and Pacific Oceans
to the north. The Southern Ocean surrounds
the Antarctic continent which, because it
lies underneath a huge weight of ice, has an
unusually narrow and deep shelf ranging from
350–500m deep. Surrounding the continent
of Antarctica are a number of sub-Antarctic
islands, including the South Shetland Islands,
South Orkney Islands, South Georgia, the South
Sandwich Islands, Bouvet Island, the Prince
Edward Islands, Crozet Islands, Kerguelen
Island, Heard and MacDonald Islands, and the
Balleny Islands. These islands are generally
located on large submarine features that
isolate the deep basins of the Southern Ocean,
including the South Georgia Ridge, the East
Scotia Ridge, the America-Antarctic Ridge, the
Atlantic-Indian Ridge, the Southwest Indian
Ridge, the Crozet and Kerguelen Plateaus, the
Southeast Indian Ridge, the South Tasman
Rise and the Pacific-Antarctic Ridge. Large
areas of the coastal seas of Antarctica lie
beneath ice shelves and more than one-half
of the Southern Ocean freezes each winter.
Because of the limited shelf seas, a lack
of the micronutrient iron in surface waters,
harsh environmental conditions and extreme
seasonality, the fisheries in the Southern
Ocean tend to be limited in productivity, with
the exception of the pelagic Antarctic krill
(Euphausia superba).
Management of fisheries for deep-sea
species of low productivity
Large-scale fisheries for finfish in the SubAntarctic/Antarctic commenced in 1969
around South Georgia and other Sub-Antarctic/
Antarctic Islands. Large catches were taken, with
400,000t of marbled notothenia (Notothenia
rossii) taken in 1969/70 and 100,000t in
the following season by Soviet fleets. The
fishery then collapsed after being fished
for a couple of years (Kock et al., 2007).
Other fisheries followed a similar pattern
of collapse after a short period of intense
exploitation, including those for mackerel icefish
(Champsocephalus gunnari), yellow notothenia
(Gobionotothen gibberifrons), Scotia Sea icefish
(Chaenocephalus aceratus), South Georgia
icefish (Pseudochaenichthys georgianus),
Patagonian rockcod (Patagonotothen
brevicauda), spiny icefish (Chaenodraco
wilsoni) and grey notothenia (Lepidonotothen
squamifrons). Overall, by 1992, some 2.08
million tonnes of fish had been extracted from
the Atlantic sector of the Southern Ocean,
with 3 million tonnes taken from the Southern
Ocean overall, not including illegal or unreported
catches (Ainley & Blight, 2009). It has now been
realised that this massive extraction of biomass
has significantly contributed to the decline of
predator (seals and birds) populations in the
Antarctic (Ainley & Blight, 2009). Such mining of
fisheries resources was brought to an end by the
Fig. 49. Map showing the
CCAMLR Regulatory Area.
declaration of 200nm limits around many of the
Sub-Antarctic Islands and the establishment of
the Convention on the Conservation of Antarctic
Marine Living Resources (CCAMLR; Fig 49) in
1982. The only finfish fisheries remaining in
the CCAMLR area at present are for mackerel
icefish, which is taken by bottom and mid-water
trawl fishing around Heard Island (Division
58.5.2) and pelagic trawls around South Georgia
(Division 48.3), and toothfish (Dissostichus
eleginoides and Dissostichus mawsoni), taken
in a number of established and exploratory
fisheries around Antarctica, mainly by longline,
but also by trawl in Heard Island. Toothfish were
initially fished by Russian vessels around South
Georgia in the 1980s but the fishery was only
recently noted because there was a lack of
reporting in the late 1980s (Kock et al., 2007).
Other fisheries of note in the region included
those for Antarctic krill, which reached a peak
of 550,000t in the 1980s but fell dramatically
with the collapse of the Soviet Union. A fishery
for the small mesopelagic lanternfish, Electrona
carlsbergii, also took place in the 1980s–1990s
but was discontinued for commercial reasons.
At present, catches in the Southern Ocean
are a fraction of past fisheries and UN FAO
views the region as one where relatively high
proportions (20 percent or more) of stocks are
moderately or underexploited. There are plans
to increase the exploitation of krill for fishmeal
and pharmaceutical products (CCAMLR Review
Panel, 2008), despite evidence of krill’s key
position in the food chain and of declines in krill
populations over time in some regions of the
Antarctic.
The long period of completely unmanaged
directed fishing in the Southern Ocean meant
that many fish stocks were encountered and
exploited before they came under CCAMLR
management. Rules were initiated in 1991
for new or exploratory fisheries, which require
that any state with vessels that intend to
undertake exploratory fishing activities must
notify the Commission in advance so that such
applications can be assessed and management
measures established prior to exploitation (Kock
et al., 2007). These measures have prevented
the further development of unmanaged directed
fisheries in recent years.
All the fisheries presently targeting finfish in
the CCAMLR Regulatory Area are deep water,
with those at Heard Island including bottom and
mid-water trawl fisheries and at South Georgia,
72 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
pelagic trawls. All other fisheries are based on
bottom longlines, although there have recently
been experimental fisheries using pots. Three
main species are exploited: mackerel icefish
(Champsocephalus gunnari) and Patagonian and
Antarctic toothfish (Dissostichus eleginoides and
Dissostichus mawsoni).
Mackerel icefish (Champsocephalus gunnari)
Mackerel icefish feed on krill and in turn are
an important prey species for other predators
in the Antarctic, such as fur seals and gentoo
penguins (Kock et al., 2007). Recruitment to
stocks of mackerel icefish vary by up to a factor
of 20 and in some years adult mortality can
also be high. Catch limits are set, therefore,
on a two-year projection based on survey
estimates of stock size; the surveys occur
annually (Kock et al., 2007). The species is
generally fished at depths of 180–400m depth
in the Heard Island fishery (MSC, 2006). In
general, although this fishery is classed as fully
exploited and recent TACs have been set at a
very low level, the fishery is regarded as well
managed and received certification from the
Marine Stewardship Council (MSC) in 2006.
Mackerel icefish grow relatively quickly and are
short lived and, overall, the species can be
viewed as one of intermediate productivity. The
stock that is fished using bottom trawl gear does
not, therefore, fall into the scope of the FAO
Guidelines (2009a) because it is fished within
the Australian EEZ around Heard and MacDonald
Islands and not on the high seas.
Toothfish (Dissostichus eleginoides and
Dissostichus mawsoni)
Both of these species are fished with bottom
longlines in the CCAMLR Regulatory Area and
both are long-lived (40–50 years) and slowgrowing species that reach maturity at 6-10
years old. The fish grow to a very large size and
move into deeper waters (up to 3,000m depth)
as they get older (Kock et al., 2007). Aspects of
the life history of toothfish identify it as a lowproductivity species and declines in exploited
populations suggest that it is vulnerable to
overfishing. Most of the fished stocks of
toothfish are managed and have been fished to
planned levels of biomass aimed at sustainable
exploitation over the long term (CCAMLR Review
Panel, 2008). The South Georgia fishery, for
example, has been certified by the MSC and was
recertified in 2009 without condition and is the
first fishery to have received such unconditional
certification. In some cases, though, stocks
have been overexploited in the CCAMLR area,
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 73
a situation that has been aggravated by IUU
fishing, reflecting the difficulties in monitoring,
control and surveillance of fisheries in the
remote Southern Ocean (CCAMLR Review Panel,
2008). This remains a problem in many areas
throughout the Southern Ocean but has reduced
in recent years as a result of catch certification.
Recent evidence suggests that the exploratory
fishery for toothfish in the Ross Sea (Area
88.1) is having significant impacts on toothfish
populations, and further on the wider ecosystem
through impacts on its predators of toothfish
(killer whales, sperm whales and Weddell
seals) and its prey (demersal fish, of which
toothfish can remove up to 70 percent of the
annual production; Ainley et al., 2009; CCAMLR
Scientific Committee, 2008a). Significant
declines in catches of toothfish in long-term
sampling programmes on the coast of the Ross
Sea (DeVries et al., 2007) indicate fisheryinduced declines of toothfish populations and/
or changes in the distribution of toothfish
in response to fishing (density-dependent
behaviour). A draft document supporting the
MSC certification of the Ross Sea fishery as
sustainable has been drawn up (Moody Marine
Ltd, 2008) but has been severely criticised by
Area (Existing
fisheries)
Macrourids/
Macrouridae
48.3 South Georgia
58.5.1 Kerguelen
161t
453t
103t (mainly
M. carinatus)
4t (South
African EEZ)
No data
71t
58.6 Crozet
58.6 + 58.7 Prince
Edward Islands.
58.4.4 Ob & Lena Banks
58.5.2 Heard Island
48.1 Peninsula and
South Shetland Isls.
48.2 South Orkney Isls.
48.4 S Sandwich Isl.
Closed
Closed
16t
Exploratory fisheries
48.2
48.4
Table 3. By-catch of
Macrouridae from
established and exploratory
fisheries in 2007/2008
(CCAMLR Scientific
Committee, 2008)
48.6 Bouvet Is. Sector
58.4.1 South Indian Basin
58.4.2 Prydz Bay
58.4.3a Elan Bank
58.4.3b Banzare Bank
88.1 Balleny Isls.
88.2 Ross Sea
Crab fishery
Crab fishery
0t (no fishing in
2007/2008 season)
36t
12t
0t
7t
112t
17t
three of these areas are overseas territories
and the fisheries take place within the EEZ
of the islands and so are not high seas. For
some areas, assessments on Macrourus
species have been undertaken and catches
are currently within acceptable limits (e.g.
Area 88.1). However, for many of the regions
within the CCAMLR Regulatory Area there are
no fisheries-independent data on macrourid
populations (CCAMLR Review Panel, 2008)
and so it is not possible to assess the overall
impact of fisheries on macrourid populations.
This is aggravated by the fact that by-catch for
grenadiers is not identified to species, but is
usually listed only as Macrouridae. Reasonable
keys to Southern Ocean macrourids are available
(e.g. Gon & Heemstra, 2000) but there is still
confusion over their identification in some
regions of the Southern Ocean.
the Antarctic and Southern Ocean Coalition
(ASOC, 2009) and other NGOs (Moody Marine
Ltd, 2008). These critics point out that many
aspects of the biology of toothfish in the Ross
Sea region are unknown and the species
is vulnerable to overfishing, and, therefore,
management of the fishery is subject to serious
uncertainties, reflected in its continued status
as exploratory.
Other retained and discarded by-catch
species
The information on by-catch from the current
deep-water fisheries in the CCAMLR Regulatory
Area is patchy in respect of species, area and
the interests of the Contracting Parties of the
Commission (CCAMLR Independent Review,
2008). CCAMLR has set catch limits on a
number of deep-water species that are likely to
be of low productivity and thus high vulnerability
and low resilience to fishing pressure. In many
cases, such catch limits are precautionary but
are based on limited scientific information.
There are too few fishery-independent data
to allow an assessment of the impacts of
fishing on non-target fish species. For new and
exploratory fisheries there are by-catch limits
for skates and rays in all management areas
and a by-catch limit of 20t for all other species
combined. If by-catch thresholds are exceeded
for any Regulatory Areas, then fishing must stop
and the vessel responsible must move on 5nm
(CCAMLR Review Panel, 2008).
Macrourids
A number of grenadier species are taken
as by-catch in Southern Ocean fisheries for
toothfish including Macrourus carinatus,
Macrourus holotrachys, Macrourus whitsoni,
Coryphaenoides armatus and Caelorhynchus
marinii (CCAMLR Scientific Committee, 2008a).
However, much confusion exists regarding the
identification of these species and in many
cases they are identified simply as Macrourus
sp. or Coryphaenoides sp. It is recognised that
these are long-lived species with low productivity
(e.g. CCAMLR Scientific Committee, 2008a).
Table 3 shows the reported by-catch of
macrourids in bottom fisheries in the CCAMLR
Regulatory Area. In many cases reported
by-catches are below the maximum allowed
by-catch for individual areas. By-catches of
macrourids are highest in the South Georgia,
Crozet and, particularly, Kerguelen fisheries
for toothfish. Those at Kerguelen represent
a significant proportion of the total catch. All
74 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Table 4. By-catch of skates
and rays from bottom
fisheries in the CCAMLR
Regulatory Area.
In some places, species information has been
recorded. Macrourus whitsoni is the dominant
by-catch species in the Ross Sea longline fishery
for toothfish, averaging abut 10 percent of the
total catch and amounting to 480t in 2005
(Hanchet et al., 2008). Nonetheless, commercial
CPUE catches are not a good estimator of
abundance because these rates are shown
to vary markedly with vessel, area and depth.
Standardised scientific surveys will be required
to properly assess populations.
Skates and rays
(Rajiformes, Bathyraja spp., Raja spp., Bathyraja
eatonii, Bathyraja irrasa, Bathyraja maccaini,
Bathyraja meridionalis, Bathyraja murrayi, Raja
georgiana, Raja taaf)
The other major group of deep-sea fish that are
taken as by-catch in the bottom fisheries of the
Area (Existing fisheries) Rajids
48.3 South Georgia
58.5.1 Kerguelen
58.6 Crozet
58.6 + 58.7 Prince Edward
Islands.
58.4.4 Ob & Lena Banks
58.5.2 Heard Island
48.1 Peninsula and South
Shetland Isls.
48.2 South Orkney Isls.
48.4 S Sandwich Isl.
Exploratory fisheries
48.2
48.4
48.6 Bouvet Is. Sector
58.4.1 South Indian Basin
58.4.2 Prydz Bay
58.4.3a Elan Bank
58.4.3b Banzare Bank
88.1 Balleny Isls.
88.2 Ross Sea
12t (19,558 released)
230t
39t
0t (South African EEZ)
No data
13t (8,586 released)
Closed
Closed
4t (8,276 released)
Crab fishery
Crab fishery
0t (no fishing in 2007/2008 season) Bathyraja eatonii
Bathyraja eatonii, Bathyraja spp.,
0t
Rajiformes
Raja georgiana, Bathyraja eatonii,
Bathyraja maccaini, Bathyraja
0t
irrasa, Bathyraja spp., Raja spp.,
Rajiformes
Raja taaf, Raja georgiana,
2t
Rajiformes
Raja georgiana, Bathyraja maccaini,
1t (155 released)
Bathyraja murrayi, Bathyraja spp.,
Raja spp., Rajiformes
Raja taaf, Raja georgiana, Bathyraja eatonii, Bathyraja maccaini,
4t (7,190 released)
Bathyraja murrayi, Bathyraja irrasa,
Bathyraja meridionalis, Bathyraja
spp., Raja spp., Rajiformes
Raja georgiana, Bathyraja eatonii,
0t
Bathyraja maccaini, Raja spp.,
Rajiformes
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 75
CCAMLR Regulatory Area are skates and rays.
These are low-productivity species because of
their very conservative life histories and are
vulnerable to overfishing.
In most areas the reported by-catch of skates
and rays is low. The exceptions to this are
Kerguelen and Crozet, where the by-catch of
skates and rays is substantial. These fisheries
take place within the EEZ of the islands and
not on the high seas. Presently, measures are
in place to assess skates that are captured
during fishing. If they are injured or dead they
are retained on the vessel but if they are likely
to survive being returned to the sea they are cut
free from the longline and the hook is removed
if it can be done without damaging the animal.
At present, it is unclear what proportions of
animals survive capture and release, however,
a research programme is currently in place
that should produce data to help assess the
survival rates of released skates and rays.
Large numbers are returned to the sea in
some regions (Table 4). As with macrourids,
assessments have been made on skates
and rays for some regions within the CCAMLR
Regulatory Area but in some cases these have
been problematic because of a lack of data
and uncertainties regarding the life history and
growth rates of Southern Ocean species. For
most regions within the CCAMLR area there is
no assessment of skate and ray populations and
so no way of evaluating the state of populations.
In addition, in many areas the skates and rays
are not identified to species; this partially
reflects the systematic problems in this group of
fish in the Southern Ocean, although reasonable
keys do exist to aid identification (e.g. Gon &
Heemstra, 2000).
Blue hake (Antimora rostrata)
3 Note that CCAMLR cannot
designate ASPAs or ASMAs,
although it can designate
CCAMLR protected areas,
which can subsequently be
designated as ASPAs
or ASMAs by the ATCM.
Little is known about this species but it appears
to have undergone major declines in other
regions as a result of by-catch in deep-water
fisheries (e.g. northwest Atlantic; Devine et al.,
2006). By-catches of blue hake are significant
in the bottom fisheries of Kerguelen and Crozet
(45t and 49t, respectively, in 2007/08) and
it is also being taken as by-catch in Regions
88.1 and 88.2. Elsewhere in the CCAMLR
Regulatory Area it is being reported in relatively
low quantities as by-catch. As yet there has
been no assessment of this species in terms
of the impact of fishing on populations and it is
unmanaged in the CCAMLR region. Currently, the
species is of little commercial interest.
Other species
Sleeper sharks, Somniosus antarcticus, are
occasionally taken as by-catch in the Southern
Ocean but CCAMLR has assessed the risks to
this species and concluded that it is ‘low’ from
longline fisheries (CCAMLR Review Panel, 2008).
However, deployment of gillnets by IUU vessels
in the region does pose a risk to these, and to
most, sharks.
A variety of other species are taken as by-catch
in both the bottom trawl and longline fisheries
in the CCAMLR region, although generally not
at the same level as macrourids, skates and
rays and Antimora rostrata. The most significant
species include the moray cods (Muraenolepis
spp.) and various species of Nototheniidae
and Channichthyidae but there are low levels
of catches of other species throughout the
CCAMLR Regulatory Area (CCAMLR Impact
Assessments, 2008). Some of these by-catches
are limited, or the by-catch of ‘other’ species is
limited in exploratory fishing areas. At present
there are no assessments of the impact of
fishing mortality on these species and very little
is known of their relative abundance.
This placed CCAMLR in a unique position
whereby states could use the Commission
to refuse any proposals for MPAs that they
considered might presently, or in the future,
affect their commercial (fishing) activities. The
result of these decisions has been that, even
with the Antarctic Treaty in place, there has been
little development of the legal means required
to initiate a network of MPAs, even in the high
seas, and in recent years only relatively small
areas have been designated, mostly around
overseas territories or in coastal areas (see
Fig. 50). Note that some fisheries protection
measures have been directed at specific
areas of the Southern Ocean, for example the
closure of the Ob and Lena Banks (seamounts;
Statistical Division 58.4.4.) to fishing for
Lepidonotothen squamifrons (Conservation
Measure 32-08 (1997); CCAMLR, 2009a). The
CCAMLR Review Panel (2008) identified that
there were marked differences in views among
Contracting Parties as to how to define ASPAs
and ASMAs and indeed, despite the fact that
CCAMLR had the power to close areas to fishing
for conservation purposes, little action had been
taken. Until 2009, the CCAMLR Regulatory Area
Protection of benthic marine ecosystems
In 1991, it was agreed that the Antarctic
Treaty, through the Protocol on Environmental
Protection, Article 4, Paragraph 1, Annex V,
would acquire the powers to designate “any
area, including any marine area” as an Antarctic
Specially Protected Area (ASPA) or an Antarctic
Specially Managed Area (ASMA) during the
Antarctic Treaty Consultative Meeting (ATCM).
Annex V was adopted in 2002 and placed
the Antarctic Treaty in the unique position of
being able to designate any part of the marine
environment, including the high seas within
the Treaty Area, as a marine protected area
(MPA). However, Annex V, Article 6, Paragraph 2
stipulated that no area was to be closed without
prior approval of CCAMLR, although this was
later modified to include areas where harvesting
or the potential for harvesting existed, or where
CCAMLR-related activities could be prevented
or restricted. Effectively, this gave CCAMLR
powers of veto over any MPA in the Regulatory
Area where Contracting Parties could make a
case that harvesting or some future possibility
of harvesting existed. This meant that any
proposals for MPAs had to enter a process
of dual consideration by the Committee for
Environmental Protection (CEP) and CCAMLR3.
76 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Figure 50. Map of Antarctica showing current protected areas, including sites on
South Orkney Islands, South Shetland Islands, Palmer Archipelago, Marguerite Bay,
Ross Island, Beaufort Island, White Island, Granite Harbour, Edmonson Point, Cape
Hallett, Cape Adare, Sabrina Island, Point Martin, Pointe Geologie, Haswell Island,
Hawker Island and Rookery Island.
was not as active as other RFMOs’, such as
NEAFC and NAFO, in the designation of networks
of MPAs to protect VMEs.
Recently, however, CCAMLR and CEP clarified
their roles in relation to conservation
activities, including the protection of the
marine environment at a workshop (CCAMLR,
2009b). During this meeting it was agreed that
CCAMLR and CEP would work more closely
on the protection of marine areas by adopting
harmonised approaches to data gathering and
designation of protected areas. The Scientific
Committee of CCAMLR will in future lead
work on spatial protection and management
of Antarctic marine biodiversity (CCAMLR,
2009b). To this end, both communities are
adopting a unified approach in the use of
bioregionalisation methods to identify 11
priority representative areas in the Southern
Ocean and coastal Antarctica (CCAMLR, 2008,
2009b; CCAMLR Scientific Committee, 2008b).
This bioregionalisation approach has used a
combination of oceanographic, geomorphological
and environmental data as well as information
on species diversity and biogeography, to
identify how to distribute a system of MPAs
that represent major and rarer ecosystems.
However, other approaches, for example, specific
knowledge of rare or vulnerable ecosystems,
can also be used for designation of MPAs. Two
bioregionalisation workshops have now taken
place, the first in Hobart, Australia (Grant et
al., 2006) and the second in Belgium (Penhale
& Grant, 2007). The CCAMLR Independent
Review (2008) pointed out that there was now
an urgent need to maintain momentum on
the designation of a network of MPAs in the
CCAMLR/Antarctic Treaty Area and that the next
stage, the identification of sites for protection, is
a critical one. This was acknowledged during the
SC-CCAMLR/CEP workshop (CCAMLR, 2009b).
In November 2009, the UK government
designated a large MPA at the South Orkney
Islands. The protected area, which covers
94,000km2 (Fig. 51), protects a range of
marine habitats including shelf and seamounts
as well as habitat for important prey species
such as Antarctic krill and predators such as
Adélie penguins. It also protects an area where
significant concentrations of VMEs have been
located through trawl and video surveys by US
scientists (Lockhart & Jones, 2009). The area
is protected from fishing activities and will come
into force in May 2010.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 77
Figure 51.
New MPA
south of
Coronation
Islands.
Protection of VMEs
In response to UNGA Resolution 61/105,
Conservation Measures (CM) 22-05, 22-06
and 22-07 were adopted by CCAMLR. CM22-05
banned bottom trawling from the Regulatory
Area, apart from areas in which conservation
measures were in place for bottom trawling
(Heard Island) and with the exception of
scientific trawling. CM22-06 required Contracting
Parties to undertake an impact assessment
of all bottom fishing activities in areas where
exploratory fisheries were in operation, to
cease fishing where VMEs were encountered (in
accordance with CM22-07), to carry observers
on all vessels, and to collect data related to the
by-catch of VME taxa. Impact assessments and
other measures in CM22-06 were not required
in fisheries established prior to 2006/07.
CM22-07 outlined the protocols for the moveon rule for CCAMLR, but again this only applies
to exploratory or recent fisheries (those that
started after 2006/07). In addition, a previous
conservation measure, CM22-04 had already
banned gillnetting from the CCAMLR area.
Impact assessments
Despite the requirement for impact
assessments, only five of 11 states (Australia,
Japan, New Zealand, Spain and the UK)
undertaking exploratory bottom fisheries
submitted impact assessments in 2008
(CCAMLR, 2008: 5.8). Furthermore, the impact
assessments varied considerably in substance
because several of them did not follow the proforma that accompanied CM22-06 when it was
agreed. This required the following information.
1. The fishing method.
2. CCAMLR Division (area).
black corals, soft corals, sea fans, sea pens,
hydrocorals, hydroids, bryozoans, crinoids,
basket stars, sea squirts, and species belonging
to chemosynthetic communities. New Zealand
also assessed the distribution of VMEs in
collaboration with scientists, undertaking
scientific surveys of the benthic ecosystems
of the region under the CAML project (Census
of Antarctic Marine Life, part of the Census of
Marine Life programme). It then used fishing
effort rather than by-catch as a conservative
method to assess potential impacts and
concluded, as did the UK, that overall impact
on seabed communities was likely to be small.
Future avenues of research were explored as
well as possible mitigation measures in the
fishery. Some aspects of the analyses might
be disputed, for example the classification
of stony corals as particularly vulnerable to
longline fishing. Most Antarctic Scleractinia
are solitary, not colonial, forms with erect
branching morphology. The extreme fragility of
crinoids to physical impacts was also not taken
into account. The assessment also concluded
that it was probably not possible to accurately
determine VME positions from longline data on
by-catch, as explored in the present report for
other methods of fishing.
3. The year for which the application is made.
4. A detailed description of the fishing gear,
including a diagram of gear configuration.
5. The scale of the proposed activity (number
of hooks/lines to be deployed).
6. Subareas and depths in which fishing was to
take place.
7. Mitigation measures to reduce impacts on
VMEs.
8. Estimated spatial footprint of fishing effort.
9. A summary of potential VMEs present in the
area of fishing.
10. Probability of impacts.
11. Magnitude and severity of impacts on VMEs.
12. Physical and biological/ecological
consequences of impacts.
13. Previous research.
14. Research planned during the season.
15. Future research.
Of the Contracting Parties that completed
impact assessments, New Zealand produced
the most comprehensive assessment and
was the only Contracting Party that followed
the recommended pro forma for assessments
under CM22-06 (Table 5). Its submission
provides a useful model for other Contracting
Parties to follow and, furthermore, offers
much background information to be used for
other impact assessments. Spain and the UK
produced assessments that followed many of
the recommendations of the pro forma but the
assessments for these Contracting Parties
did not contain the same depth of information
as New Zealand, particularly in terms of the
footprint of the fishery and assessment of the
extent and ecological impacts of by-catch of
VME species (Table 5). Australia and Japan did
not follow the pro forma and their assessments
are not comprehensive (Table 5). No impact
assessments were undertaken by Argentina,
Chile, Republic of Korea, Russia, South Africa
or Uruguay despite the fact that all applied to
undertake exploratory fisheries in 2008/09
(CCAMLR Scientific Committee, 2008: Annex 5).
New Zealand provided an assessment of its
past years’ fishing efforts and its intended
fishing operations for 2008/09. This included
a comprehensive literature review and datagathering exercise and an ecological risk
assessment (ERA) for the fishery, including
workshops with participants from the scientific,
management and fishing industries and NGO
communities. These defined VMEs and then
identified VME indicator organisms. VME taxa
included sponges, anemones, stony corals,
78 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
In 2009/10 reporting improved, with Argentina,
Japan, Republic of Korea, New Zealand, Russia,
Uruguay, South Africa, Spain and the UK all
providing assessments (Table 6). Still, not all
were complete and in many cases were quite
vague. Korea, for example, stated that it would
comply with relevant measures but provided
no detail and failed to provide any information
on impacts or future research. Japan, as it had
done in 2007/08, provided a small proportion
of what was required and its commitment to
future research was to send a scientist to a
CCAMLR VME workshop. Argentina, Russia,
South Africa and the UK provided fuller reports,
with reference to gear details, background
literature and information on research plans.
New Zealand’s report was again the most
comprehensive. It included a detailed and
sophisticated research plan with subjects
that included specific studies, inter alia, of D.
Table 5. Summary of the
impact assessments for the
2008/09 season in terms of
providing information on the
15 aspects of the pro forma
for CM22-06 provided by
CCAMLR.
State
Aus
Jpn
NZ
Spa
UK
1.
Yes
Yes
Yes
Yes
Yes
2.
Yes
No
Yes
Yes
Yes
3.
Yes
Yes
Yes
Yes
Yes
4.
No
Yes
Yes
Yes
No*
5.
No
No
Yes
No
Yes
6.
No
No
Yes
No
Yes
7.
No
Yes
Yes
Yes
Yes
mawsoni, macrourids, skates, modeling fishing
distribution vs VMEs, and ecosystem impacts
of bottom fishing. Literature produced and cited
ranged from Working Group reports to papers in
scientific literature, all of it useful in the context
of the Southern Ocean generally.
Australia submitted no assessment for 2009/10
although it had done so in 2008/09. In the
earlier report it stated that while the impacts of
bottom longlining on deep-sea taxa are unknown,
it was likely that only sub-lethal damage would
occur, with some mortality a possibility. The
report cited poor knowledge of which VMEforming groups actually occurred in the area, an
argument that recurred in many of the reports
from Contracting Parties in both years. Australia
did acknowledge that, given the knowledge of
the marine fauna around Heard and MacDonald
Islands on the Kerguelen Ridge or Plateau and
knowledge of the marine fauna located in the
Ross Sea (Australian Antarctic Division, 2008),
the occurrence of VMEs in this region is likely.
Gorgonians are reported as common in the
Heard Island area and bryozoans form extensive
habitats on some of the banks around Heard
and MacDonald Islands (Hibberd & Moore,
2009). VMEs formed by stylasterid corals and
sponges have been recently identified by the
Collaborative East Antarctic Marine Census
(CEAMARC) project in the region (Australian
Antarctic Division, 2008).
Such studies, and the information provided
by New Zealand, suggest it is not possible to
conclude that only negligible damage is expected
in areas where exploratory fishing is planned, as
asserted by many states’ assessments. There
is by now considerable evidence from various
fisheries around the world that benthic longlines,
which is what the vessels of most Contracting
Parties employ (Argentina uses pots), do
damage to VMEs (e.g. Stone, 2006; Edinger et
al., 2007), and indeed by-catch has been welldocumented in the South Georgia fisheries for
Patagonian toothfish. The impact on sessile
epibenthic fauna is less than that of bottom
trawls but damage has nonetheless been shown
to be significant and cumulative.
8.
No
No
Yes
No
Yes
9.
Yes
No
Yes
Yes
Yes
10.
Yes
Yes
Yes
Yes
Yes
11.
No
No
Yes
No
No
12.
No
No
Yes
No
No
13.
No
Yes
Yes
Yes
Yes
14.
Yes
No
Yes
Yes
Yes
15.
No
No
Yes
Yes
Yes
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 79
The move-on rule
The move-on rule has been adopted by CCAMLR
for exploratory fishing areas only. The measures
are detailed in CM22-07 and rely on estimating
the number of ‘VME units’ taken as by-catch
per segment of a longline (1,000 hooks or
1,200m, whichever is shorter). A single VME
unit is 1 litre of organisms in a 10L bucket or
1kg of organisms that do not easily fit into a
bucket. However, some Contracting Parties have
proposed alternative triggers for identification
of VME risk areas (e.g. 15–20 individual VME
taxa per 1,000 hooks; Spain and UK, Impact
Assessments). A ‘risk area’ is designated
where 10 VME units are recovered from a single
segment of longline. Here, a radius of 1nm
from the midpoint of the estimated position of
the line segment is defined as the risk area. A
vessel encountering a risk area should not shoot
any further lines within the risk area and has
to report the encounter area to the flag state
and the Secretariat. Following this action, the
area is closed to fishing. If five or more VME
units are recovered in a segment, that is also
notified to the flag state and Secretariat. If five
or more notifications of catches of five VME
units are recorded in a fine-scale rectangle, the
Secretariat notifies all fishing vessels of the
possibility of occurrence of VMEs within that
area.
VMEs are defined on the basis of organisms
listed in the Benthic Classification Guide. This
is a full-colour set of classification cards with
photographs and defining features of VME
taxa noted on them. Taxa include gorgonians,
hydroids, stylasterids, stony corals, black
corals, bryozoans, sponges, sea anemones,
soft corals, sea pens, sea squirts, stalked
crinoids and basket stars. Most of these are
habitat-forming groups; basket stars associated
with habitat-forming groups tend themselves
Table 6. Summary
of the impact
assessments for the
2009/10 season in
terms of providing
information on the
15 aspects of the
pro forma for CM2206 provided by
CCAMLR.
*Not all details
provided as required
by CCAMLR.
State
Arg
Aus
Jpn
Kor
NZ
Rus
SA
Spa
UK
Ury
1.
2.
3.
Yes Yes Yes
N/A N/A N/A
Yes Yes Yes
Yes Yes Yes
Yes Yes Yes
Yes Yes Yes
Yes Yes Yes
Yes Yes Yes
Yes Yes Yes
Yes Yes Yes
4.
Yes*
N/A
Yes
No
Yes
No
Yes
Yes
No
Yes
5.
Yes
N/A
Yes
No
Yes
No
Yes
Yes
Yes
Yes
to form further fine-scale habitats. Analyses,
based on the first year of operation of these
guides for identification of trigger levels for VME
risk areas, have indicated that they worked
well for most taxa, although there was some
confusion between hydrocorals, stony corals
and precious corals (red corals, octocorals;
Parker et al., 2009). A CCAMLR workshop in
2009 extended the list of potential VME taxa
to include cidaroid sea urchins (pencil urchins),
brachiopods, serpulid worms, barnacles from the
family Bathylasmatidae, the scallop Adamussium
colbecki, pterobranchs and xenophyophores
(CCAMLR, 2009c).
By June 2009, 30 notifications had been
received by the Secretariat from fisheries
operations in the 2008/09 season. These
were from Division 48.6 (one notification;
5.5 VME-indicator units per line segment;
seafloor depth 880–980m), Division 88.1 (18
notifications; 5.0–68.6 VME-indicator units
per line segment; seafloor depth 585–1528m)
and Division 88.2 (11 notifications; 5.1–10.4
VME-indicator units per line segment; seafloor
depth 1,272–1,694m). No notifications were
made for Divisions 58.4.1, 58.4.2 and 58.4.3b.
Seven risk areas were identified in Divisions
88.1 and 88.2 and one small-scale rectangle
was noted as having a risk of encounters of
VMEs in Division 88.2. Trigger levels for actions
related to the closure of risk areas for VMEs are
much lower in the CCAMLR area than for other
RFMOs, including NEAFC, NAFO and NPFC. These
reflect an assessment of the potential retention
of animals taken as by-catch when a longline
encounters a VME. However, research over one
season has indicated that different trigger levels
may be appropriate for different-sized taxa (e.g.
large gorgonians vs small hydrocorals; Mitchell
et al., 2009), as also indicated by research
by NAFO for the northwest Atlantic (WGEAFM,
2008b). Furthermore, it is likely that there are
6.
7.
8.
Yes* No Yes
N/A N/A N/A
Yes Yes Yes
No Yes No
Yes Yes Yes
Yes* No No
Yes* No No
Yes Yes No
Yes No Yes
Yes No Yes
9.
10. 11.
No Yes No
N/A N/A N/A
Yes Yes Yes
Yes Yes No
Yes Yes No
No No No
No No No
Yes Yes No
Yes Yes Yes
No No No
12. 13. 14. 15.
No No Yes Yes
N/A N/A N/A N/A
Yes Yes Yes Yes
No Yes Yes Yes
No Yes Yes Yes
No Yes Yes No
No No Yes No
No Yes Yes Yes
Yes Yes Yes Yes
No No Yes No
80 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
systematic and significant differences in catch
data depending upon vessel, year, gear type and
fisher behaviour, and level of reporting (Hanchet
et al., 2008). The number of line hauls for which
fine-scale VME data was reported (by June,
2009) varied between Contracting Parties with
some reporting on a large proportion of line sets
and others reporting on very few or no sets.
In addition to fishing encounters with VMEs,
the Secretariat received 30 notifications
of encounters with VMEs during research
surveys under CM22-06. These encounters
were reported by the USA in Division 48.1 (17
notifications, seafloor depth 92–642m) and
Division 48.2 (11 notifications, seafloor depth
96–252m) and by Australia in Division 58.4.1
(two notifications, seafloor depth 436–844m).
VMEs were documented with video/photography
or by research trawls.
CCAMLR measures CM22-06 and CM22-07
have been operating over a relatively short time
period. However, given the iterative approach
taken by the CCAMLR Secretariat and the
research undertaken by Contracting Parties in
exploratory fisheries, the current methods of
assessment of encounters with VMEs would
appear likely to improve in coming years. As
more data become available, approaches based
on accumulation curves or GIS-analyses of
density of VME encounters may prove useful in
refining both the trigger levels for designation of
risk areas and for identifying areas with a high
probability of comprising VMEs.
Conclusions
(i) Conduct assessments of whether bottom
fishing activities have SAIs on VMEs.
● All states fishing in the CCAMLR Regulatory
Area now undertake impact assessments for
exploratory or experimental fishing activities.
● Assessments are not undertaken for areas in
which fishing has taken place historically.
● The quality of assessments undertaken to
date are variable with respect to conformity
to CCAMLR requirements.
(ii) To implement measures in accordance
with the precautionary approach, ecosystems
approaches and international law and to
sustainably manage deep-sea fish stocks.
● Historically, fish stocks in the Southern
Ocean were heavily overexploited (mined)
in the 1970s and 1980s.
● CCAMLR has been successful in ending
extreme overexploitation of fish stocks in
the Southern Ocean region. In comparison
to other areas of oceans, stocks of targeted
deep-sea species appear relatively well
managed though in some cases the
management of the fisheries is adversely
impacted by IUU fishing (e.g. Patagonian
toothfish). The ecosystem impacts of targeted
fisheries remain poorly understood.
● By-catch of deep-sea species with a low
productivity, particularly macrourids and
skates/rays, occurs in the bottom fisheries
throughout the CCAMLR Regulatory Area but is
higher in some regions than others (especially
Kerguelen).
● The understanding of the impacts of
fishing on deep-sea by-catch species varies
markedly between sub-regions in the CCAMLR
Regulatory Area and, in general, is poor.
● The management of fisheries to prevent/
reduce by-catch of deep-sea, low-productivity
species also varies markedly between subregions. For by-catch species such as Antimora
rostrata no specific management is in place.
(iii) To ensure that if fishing activities have SAIs
they are managed to prevent such impacts,
including through closing areas to bottom fishing
where VMEs are known or likely to occur, or not
authorised to proceed.
● Closed areas are currently being implemented
around the South Orkney Islands and where
there have been significant by-catches of VMEassociated species or where research has
identified VMEs on the seabed.
● CCAMLR and the CEP are currently working
towards the establishment of 11 representive
MPAs around Antarctica.
(iv) To establish and implement protocols to
cease fishing where an encounter with VMEs
occurs during fishing activities and to report such
encounters so that appropriate measures can be
adopted with respect to that site.
● CCAMLR has initiated measures aimed at
protecting VMEs in areas where exploratory
fisheries are taking place.
● Rules for identification of VME risk areas have
been implemented at a conservative level and
have resulted in the identification of some risk
areas. Research work will continue to refine
and improve the estimation of levels of
by-catch that signify the presence of VMEs.
● At present, not all states are reporting to the
same degree on fine-scale assessment of
by-catch during longline operations in
exploratory areas.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 81
References
Abdulla, A., Gomei, M., Maison, E., Piante, C. (2008) Status
of Marine Protected Areas in the Mediterranean Sea. IUCN,
Malaga and WWF, France, 152 pp.
Ainley, D.G., Ballard, G., Olmastroni, S. (2009) An apparent
decrease in the prevalence of “Ross Sea killer whales”
in the southern Ross Sea. Document SC/61/SM26,
submitted to the Scientific Committee for the Conservation
of Antarctic Marine Living Resources, 16pp.
Ainley, D.G., Blight, L.K. (2009) Ecological repercussions
of historical fish extraction from the Southern Ocean. Fish
and Fisheries 10: 13–38.
Alekseyev, F., Alekseyeva, I., Zakharov, A.N. (1992)
Vitellogenesis, nature of spawning, fecundity and
gonad maturity stages of the roundnose grenadier,
Coryphaenoides rupestris, in the North Atlantic. Journal of
Ichthyology 33: 32–45.
Carbonell, A., Martín, P., de Ranieri, S. and WEDIS team
(1998) Discards of the Western Mediterranean trawl
fleets. Rapport, Commission International pour la Mer
Méditerranée 35: 392–393.
Bensch, A., Gianni, M., Grébroval, D., Sanders, J.S., Hjort,
A. (2008) Worldwide review of bottom fisheries in the
high seas. FAO Fisheries and Aquaculture Technical Paper
No. 522, Food and Agricultural Organisation of the United
Nations, Rome, Italy, 145pp.
Cartes, J.E., Maynou, F., Sardà, F., Company, J.B., Lloris,
D., Tudela, S. (2004) The Mediterranean deep-sea
ecosystems: an overview of their diversity, structure,
functioning and anthropogenic impacts. In: The
Mediterranean deep-sea ecosystems: an overview of their
diversity, structure, functioning and anthropogenic impacts,
with a proposal for conservation. Part 1 IUCN, Málaga and
WWF, Rome, Italy, pp. 9–38.
Bett, B.J., Rice, A.L. (1992) The influence of the
hexactinellid sponge (Pheronema carpenteri) spicules on
the patchy distribution of macrobenthos in the Porcupine
Seabight (bathyal NE Atlantic). Ophelia 36: 217–226.
Bjorndal, T. (2009) Overview, roles and performance of the
North East Atlantic Fisheries Commission. Marine Policy
33: 685–697.
Boehlert, G.W., Sasaki, T. (1988) Pelagic biogeography
of the armourhead Pseudopentaceros wheeleri, and
recruitment to isolated seamounts in the North Pacific
Ocean. Fisheries Bulletin of the United States 86: 453–
465.
Anderson, O.F., Clark, M.R. (2003) Analysis of bycatch in
the fishery for orange roughy, Hoplostethus atlanticus, on
the South Tasman Rise. Marine and Freshwater Research
54: 1–10.
Boyer, D.C., Hampton, I. (2001) An overview of the living
marine resources of Namibia. South African Journal of
Marine Science 23: 5–35.
Arbuckle, M., Atkinson, B., Germani, V. (2008) Performance
Review Panel Report of the North East Atlantic Fisheries
Commission NEAFC. NEAFC, London, UK, 102pp.
Ardron, J.A. (2005) Protecting British Columbia’s Corals
and Sponges from Bottom Trawling. A Report by Living
Oceans Society, Sointula, British Columbia, Canada, 21pp.
ASOC (2009) Letter to Dr Andrew Hough, Moody Marine
Ltd from the Antarctic and Southern Ocean Coalition
Secretariat, Washington DC, USA, 6pp.
Australian Antarctic Division (2008) Notification of
vulnerable marine ecosystems in Statistical Area 58.4.1
CCAMLR SC-CAMLR-XXVII/13, 26 September 2008,
Agenda Item No. 4, Appendix N, 13pp.
Ávilo de Melo, A., Saborido-Rey, F., González-Troncoso, D.,
Skryabin, I., Alpoim, R. (2009) An Assessment of Beaked
Redfish (S. mentella and S. fasciatus) in NAFO Division 3M
Based on Revised 2005–2008 catches (Is a Retrospective
Biased Assessment Necessarily Useless in Terms of
Scientific Advice?). Scientific Council Meeting – June 2009.
Serial No. N5664, NAFO SCR Doc. 09/29, 56pp.
Baker, K.D., Devine, J.A., Haedrich, R.L. (2009) Deep-sea
fishes in Canada’s Atlantic: declines and predicted recovery
times. Environmental Biology of Fishes 85: 79–88.
Carbonell, A. (co-ord.) (1997) Discards of the western
Mediterranean trawl fleets. Final Report to the General
Directorate for Fisheries, EC DGXIV. Project MED/94/027.
Barthel, D., Tendal, O.S., Thiel, H. (1996) A wandering
population of the hexactinellid sponge Pheronema
carpenteri on the continental slope off Morocco, Northwest
Africa. Marine Ecology 17: 603–616.
Althaus, F., Williams, A., Schlacher, T.A., Kloser, R.J., Green,
M.A., Barker, B.A. Bax, N.J., Brodie, P., Schlacher-Hoenlinger,
M.A. (2009) Impacts of bottom trawling on deep-coral
ecosystems of seamounts are long-lasting. Marine Ecology
Progress Series 397: 279-294.
Andrews, A.H., Tracey, D.M. (2007) Age validation of orange
roughy and black cardinalfish using lead-redium dating.
Final Research Report for Ministry of Fisheries Research
Project DEE2005-02 Objective 1, 42pp.
Campana, S.E., Zwanenburg, K.C.T., Smith, J.N. (1990)
210Pb/226Ra determination of longevity in Redfish.
Canadian Journal of Aquatic Science 47: 163–165.
Boyer, D.C., Kirchner, C.H., McAllister, M.K., Staby, A.,
Staalesen, B.I. (2001) The orange roughy fishery of
Namibia: lessons to be learned about managing a
developing fishery. South African Journal of Marine Science
23: 205–221.
Branch, T.A. (2001) A review of orange roughy Hoplostethus
atlanticus fisheries, estimation methods, biology and stock
structure. South African Journal of Marine Science 23:
181–203.
Breeze, H., Davis, D.S. (1998) Deep sea corals. In:
Harrison, W.G., Fenton, D.G. (eds) The gully: a scientific
review of its environment and ecosystem. Canadian stock
assessment secretariat research document 98/83.
Department of Fisheries and Oceans, pp. 113–120.
Breeze, H., Davis, D.S., Butler, M., Kostylev, V. (1997)
Distribution and status of deep sea corals off Nova Scotia.
Marine Issues Committee’s special publication number 1.
Ecology Action Centre, Halifax, Nova Scotia.
Buhl-Mortensen, L., Mortensen, P.B. (2005) Distribution and
diversity of species associated with deep-sea gorgonian
corals off Atlantic Canada. In: Freiwald, A., Roberts, J.M.
(eds) Cold-Water Corals and Ecosystems, Springer-Verlag,
Berlin Heidelberg, pp. 849–879.
Campagno, L., Dando, M., Fowler, S. (2005) Collins Field
Guide to the Sharks of the World. Harper-Collins Publishers
Ltd, London, UK, 368pp.
Campana, S.E., Joyce, W.N. (2004) Temperature and depth
associations of porbeagle shark (Lamna nasus) in the
northwest Atlantic. Fisheries Oceanography 13: 52–64.
82 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Casey, J.M., Myers, R.A. (1998) Near extinction of a large
widely distributed fish. Science 281: 690–692.
Cavanagh, R.D., Gibson, C. (2007) Overview of
the Conservation Status of Cartilaginous Fishes
(Chondrichthyans) in the Mediterranean Sea. IUCN, Gland,
Switzerland and Malaga, Spain, vi + 42pp.
CCAMLR (2008) Report of the twenty-seventh meeting
of the Commission. Hobart, Australia, 27 October – 7
November, 2008, CCAMLR-XXVII, Commission for the
Conservation of Antarctic Marine Living Resources, Hobart,
Tasmania, Australia, 203pp.
CCAMLR (2009a) Schedule of Conservation Measures in
Force 2009/10 Season. Commission for the Conservation
of Antarctic Marine Living Resources, Hobart, Tasmania,
Australia, 232pp.
CCAMLR (2009b) Report of the Joint SC-CAMLR-CEP
Workshop (Baltimore, USA, 3 and 4 April, 2009).
SC-CAMLR-XXVIII/6 1 June 2009 Agenda Item No. 9.
Commission for the Conservation of Antarctic Marine
Living Resources, Hobart, Tasmania, Australia, 20pp +
Appendices.
CCAMLR (2009c) Report of the Workshop on vulnerable
marine ecosystems (La Jolla, CA, USA, 3 to 7 August
2009). SC-CAMLR-XXVIII/10 20 August 2009 Agenda
Item No. 4. Commission for the Conservation of Antarctic
Marine Living Resources, Hobart, Tasmania, Australia,
17pp.
CCAMLR Impact Assessments (2008) Preliminary
assessments of known and anticipated impacts of
proposed bottom fishing activities on vulnerable marine
ecosystems. Collated by the Secretariat. CCAMLR
XXVII/26*, 24 September, 2008, Agenda Item No. 5, SC
Agenda Item No. 4, 107pp.
CCAMLR Review Panel (2008) CCAMLR Performance
Review Panel Report, 168pp.
CCAMLR Scientific Committee (2008a) Report of the 27th
Meeting of the Scientific Committee, Hobart, Australia
27–31 October, 2008. Appendices D to Q Fishery Reports,
Scientific Committee for the Conservation of Antarctic
Marine Living Resources, Hobart, Tasmania, Australia.
CCAMLR Scientific Committee (2008b) Report of the 27th
Meeting of the Scientific Committee, Hobart, Australia 27–
31 October, 2008. SC-CCAMLR-XXVII. Scientific Committee
for the Conservation of Antarctic Marine Living Resources,
Hobart, Tasmania, Australia, 635pp.
CIESM (2006) Executive Summary. Fluid seepages/mud
volcanism in the Mediterranean and adjacent domains,
Bologna 19–22 October, 2005. CIESM Workshop
Monographs No. 29, Monaco, 18pp. http://www.ciesm.
org/online/monographs/bologna06.pdf
Clark, M.R. (1999) Fisheries for orange roughy
(Hoplostethus atlanticus) on seamounts in New Zealand.
Oceanologica Acta 22: 593–602.
Clark, M.R. (2004) Descriptive analysis of orange roughy
fisheries in the New Zealand region outside the EEZ: Lord
Howe Rise, Northwest Challenger Plateau, West Norfolk
Ridge, South Tasman Rise, and Louisville Ridge to the
end of the 2002/03 fishing year. New Zealand Fisheries
Assessment Report 2004/51, 36pp.
Clark, M.R. (2009) Deep-sea seamount fisheries: a review
of global status and future prospects. Latin American
Journal of Aquatic Research 37: in press.
Clark, M.R., Koslow, J.A. (2007) Impacts of fishing on
seamounts. In: Pitcher T.J., Morato, T., Hart, P.J.B., Clark,
M.R., Haggan, N., Santos, R.S. (eds) Seamounts: Ecology,
Fisheries & Conservation, Fish and Aquatic Resources
Series 12, Blackwell Publishing, Oxford, UK, pp. 413–441.
Clark, M.R., O’Driscoll, R. (2003) Deep-water fisheries
and aspects of their impact on seamount habitat in New
Zealand. Journal of Northwest Atlantic Fishery Science 31:
441–458.
Clark, M.R., Rowden, A.A. (2009) Effect of deepwater
trawling on the macro-invertebrate assemblages of
seamounts on the Chatham Rise, New Zealand. Deep-Sea
Research I 56: 1540–1554.
Clark, M.R., Vinnichenko, V.I., Gordon, J.D.M., Beck-Bulat,
G.Z., Kukharev, N.N., Kakora, A.F. (2007) Large-scale
distant-water trawl fisheries on seamounts. In: Pitcher T.J.,
Morato, T., Hart, P.J.B., Clark, M.R., Haggan, N., Santos,
R.S. (eds) Seamounts: Ecology, Fisheries & Conservation,
Fish and Aquatic Resources Series 12, Blackwell
Publishing, Oxford, UK, pp. 361–399.
Clausen, D.M. (2008) The giant grenadier in Alaska. In:
Orlov, A.M., Iwamoto, T. (eds) Grenadiers of the World
Oceans: Biology, Stock assessment, and Fisheries.
American Fisheries Society, Symposium 63, Bethesda,
Maryland, USA, pp. 413–450.
Colloca, F., Carpentieri, P., Balestri, E., Ardizzone,
G.D. (2004) A critical habitat for Mediterranean fish
resources: shelf-break areas with Leptometra phalangium
(Echinodermata: Crinoidea). Marine Biology 145: 1129–
1142.
Company, J.B., Puig, P., Sardà, F., Palanques, A., Latasa,
M., Scharek, R. (2008) Climate influence on deep-sea
populations. PLOSone 3: e1431.
COSEWIC (2003) COSEWIC assessment and status report
on the cusk Brosme brosme in Canada.
Committee on the Status of Endangered Wildlife in
Canada. Ottawa, vi + 30pp.
COSEWIC (2004) COSEWIC assessment and status
report on the porbeagle shark Lamna nasus in Canada.
Committee on the Status of Endangered Wildlife in
Canada. Ottawa viii + 43pp.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 83
COSEWIC (2005) COSEWIC assessment and status
report on the winter skate Leucoraja ocellata in Canada.
Committee on the Status of Endangered Wildlife in
Canada. Ottawa. vii + 41pp.
COSEWIC (2007) COSEWIC assessment and status
report on the roughhead grenadier Macrourus berglax in
Canada. Committee on the Status of Endangered Wildlife
in Canada. Ottawa, vii + 40pp.
COSEWIC (2008) COSEWIC assessment and status report
on the Roundnose Grenadier Coryphaenoides rupestris in
Canada. Committee on the Status of Endangered Wildlife
in Canada. Ottawa, vii + 42pp.
Costas, F., G.-, Murua, H. (2008) An analytical assessment
of the roughhead grenadier stock in NAFO Subareas 2
and 3. In: Orlov, A.M., Iwamoto, T. (eds) Grenadiers of the
World Oceans: Biology, Stock Assessment and Fisheries.
American Fisheries Symposium 63: Springer, Heidelberg.
Davies, A.J., Roberts, J.M., Hall-Spencer, J. (2007)
Preserving deep-sea natural heritage: Emerging issues
in offshore conservation and management. Biological
Conservation 138: 299–312.
Davies, A.J., Wisshak, M., Orr, J.C., Roberts, J.M. (2008)
Predicting suitable habitat for the cold-water reef
framework-forming coral Lophelia pertusa (Scleractinia).
Deep Sea Research I 55: 1048–1062.
de Cárdenas E., Avila de Melo, A., Iglesias, S., Saborido,
F. (1997) Selectivity of 130mm mesh size in deep-sea
bottom trawl fishery in NAFO Regulatory Area. NAFO
Scientific Council Studies 30: 21–25.
de Forges, B.R., Koslow, J.A., Poore, G.C.B. (2000) Diversity
and endemism of the benthic seamount fauna in the
southwest Pacific. Nature 405: 944–947.
Deichman, E. (1936) The Alcyonaria of the western part
of the Atlantic Ocean. Harvard University, Museum of
Comparative Zoology Memoirs 53: 1–317.
Devine, J.A., Haedrich, R.L. (2008) Population trends and
status of two exploited Northwest Atlantic grenadiers,
Coryphaenoides rupestris and Macruorus berglax. In: Orlov,
A.M., Iwamoto, T. (eds) Grenadiers of the World Oceans:
Biology, Stock Assessment and Fisheries. American
Fisheries Symposium 63: Springer, Heidelberg.
Devine, J.A., Haedrich, R.L. (mss submitted) The role of
environmental conditions and exploitation in determining
dynamics of redfish (Sebastes species) in the Northwest
Atlantic.
Devine, J.A., Baker, K.D., Haedrich, R.L. (2006) Deep-sea
fishes qualify as endangered. Nature 439: 29.
DeVries, A.L., Ainley, D.G., Ballard, G. (2007) Decline of the
Antarctic toothfish and its predators in McMurdo Sound
and the Southern Ross Sea, and recommendations for
restoration. CCAMLR WG-EMM 08/07, 20pp.
Doonan, I.J., McMillan, P.J., Kalish, J.M., Hart, A.C. (1995)
Age estimates for black and smooth oreo. New Zealand
Fisheries Assessment Research Document 95/14, 26pp.
Dulvy, N.K., Metcalfe, J.D., Glanville, J., Pawson, M.G.,
Reynolds, J.D. (2000) Fishery stability, local extinctions,
and shifts in community structure in skates. Conservation
Biology 14: 283–293.
Durán Muñoz, P., Sacau, M., Sayago-Gil, M., Patrocinio,
T., Fernández-Salas, L.M., Murillo, F.J., Díaz del Río, V.,
Serrano, A. (2007) First preliminary results from ECOVUL/
ARPA (Estudio de los eCOsistemas VULnerables y los
ARtes de PescA): A Spanish interdisciplinary research
project, focused on the study of the deep-sea vulnerable
ecosystems/habitats in the Hatton Bank area (ICES XIIb
and VIb1). Working document presented to the ICES
Working Group on Deep-water Ecology. Plymouth (UK),
26–28 Feb 2007.
Earle, S., Glover, L.K. (2009) Ocean: An Illustrated Atlas.
National Geographic Society, Washington DC, USA, 352pp.
Eckhardt, J.D., Glasby, G.P., Puchelt, H., Berner, Z. (1997)
Hydrothermal manganese crusts from Enarete and Palinuro
Seamount in the Tyrrhenian Sea. Marine Georesources and
Geotechnology 15: 175–208.
Edinger, E., Baker, K., Devillers, R., et al. (2007a) Coldwater
Corals off Newfoundland and Labrador: Distribution and
Fisheries Impacts. WWF-Canada, Toronto, Canada, 41pp.
Edinger, E.N., Wareham, V.E., Haedrich, R.L. (2007b)
Patterns of groundfish diversity and abundance in relation
to deep-sea coral distributions in Newfoundland and
Labrador waters. In: George, R.Y., Cairns, S.D. (eds)
Conservation and Adaptive Management of Seamount and
Deep-Sea Coral Ecosystems. Rosentiel School of Marine
and Atmospheric Sciences, University of Miami, Miami,
USA, pp. 101–122.
Edinger, E., Wareham, V., Baker, K., Haedrich, R. (2009)
Relationships between deep-sea corals and groundfish.
In: Gilkinson, K., Edinger, E. (eds) The ecology of deepsea corals of Newfoundland and Labrador waters:
biogeography, life history, biogeochemistry, and relation to
fishes. Canadian Technical Report of Fisheries and Aquatic
Science, No. 2830, vi + 136pp.
Emig, C.C. (1988) Distributional patterns along the
Mediterranean continental margin (upper bathyal) using
Gryphus vitreus (Brachiopoda) densities (investigated by
submersible and dredging). Book of Abstracts, Fifth DeepSea Biology Symposium, Brest, France, June 26/July 1,
1988, p. 62.
European Commission (2006) Sensitive and essential
fish habitats in the Mediterranean Sea. Report of the
Mediterranean Subgroup (SGMED 06-01) of the Scientific,
Technical and Economic Committee for Fisheries (STECF),
Commission of the European Communities, Commission
Staff Working Paper, Rome, Italy, 60pp.
FAO (2008) Deep-sea fisheries in the high seas: a trawl
industry perspective on the International Guidelines for the
Management of Deep-sea Fisheries in the High Seas. FAO
Fisheries and Aquaculture Circular, No. 1036, 22pp, Food
and Agricultural Organisation of the United Nations, Rome,
Italy.
Organisation of the UN (FAO), Rome, Italy: ftp://ftp.fao.
org/FI/DOCUMENT/UNGA/deep_sea/UNGA61_105.pdf
(accessed 20 January 2010).
Fisheries Agency of Japan (2008) Report on Identification
of Vulnerable Marine Ecosystems in the Emperor Seamount
and Northern Hawaiian Ridge in the Northwest Pacific
Ocean and Assessment of Impacts Caused by Bottom
Fishing Activities on such Vulnerable Marine Ecosystems or
Marine Species as well as Conservation and Management
Measures to Prevent Significant Adverse Impacts (Bottom
Trawl), December 2008. Appendix A. List of fishes collected
by the Meisyo Maru #128 in 1993, expressed in relative
abundance in weight by seamount depth zone, with related
information, 1p.
Fisheries Agency of Japan (2008) Report on Identification
of Vulnerable Marine Ecosystems in the Emperor Seamount
and Northern Hawaiian Ridge in the Northwest Pacific
Ocean and Assessment of Impacts Caused by Bottom
Fishing Activities on such Vulnerable Marine Ecosystems or
Marine Species as well as Conservation and Management
Measures to Prevent Significant Adverse Impacts (Bottom
Trawl), December 2008. Appendix B. List of recent levels of
Japanese bottom gillnet fishing effort, 2pp.
Fisheries Agency of Japan (2008) Report on Identification
of Vulnerable Marine Ecosystems in the Emperor Seamount
and Northern Hawaiian Ridge in the Northwest Pacific
Ocean and Assessment of Impacts Caused by Bottom
Fishing Activities on such Vulnerable Marine Ecosystems or
Marine Species as well as Conservation and Management
Measures to Prevent Significant Adverse Impacts (Bottom
Trawl), December 2008. Appendix C. Nishimura, A., Yatsu,
A. Application of surplus production models to splendid
alfonsin stock in the Southern Emperor and Northern
Hawaiian Ridge (SE-NHR), 11pp.
Fisheries Agency of Japan (2008) Report on Identification
of Vulnerable Marine Ecosystems in the Emperor Seamount
and Northern Hawaiian Ridge in the Northwest Pacific
Ocean and Assessment of Impacts Caused by Bottom
Fishing Activities on such Vulnerable Marine Ecosystems or
Marine Species as well as Conservation and Management
Measures to Prevent Significant Adverse Impacts (Bottom
Trawl), December 2008. Appendix F. Assessment of
associated species, 5pp.
Fisheries Agency of Japan (2008) Report on Identification
of Vulnerable Marine Ecosystems in the Emperor
Seamount and Northern Hawaiian Ridge in the Northwest
Pacific Ocean and Assessment of Impacts Caused by
Bottom Fishing Activities on such Vulnerable Marine
Ecosystems or Marine Species as well as Conservation
and Management Measures to Prevent Significant Adverse
Impacts (Bottom Trawl), December 2008. Appendix H.
Yamigamoto, T., Takao, Y., Abe, K. Distribution of four
orders of corals observed by ROV survey. Powerpoint
presentation; 23 slides.
FAO (2009b) State of World Fisheries and Aquaculture
2008. Food and Agricultural Organisation of the United
Nations, Rome, Italy, 176pp.
Fisheries Agency of Japan (2008) Report on Identification
of Vulnerable Marine Ecosystems in the Emperor Seamount
and Northern Hawaiian Ridge in the Northwest Pacific
Ocean and Assessment of Impacts Caused by Bottom
Fishing Activities on such Vulnerable Marine Ecosystems or
Marine Species as well as Conservation and Management
Measures to Prevent Significant Adverse Impacts (Bottom
Trawl), December 2008. Appendix M. Information on coral
fisheries in the Emperor Seamount area, 3pp.
FAO (2010) Deep-sea high seas fisheries (UNGA 61/105)
Vessels authorized to conduct bottom fisheries in areas
beyond national jurisdiction. Food and Agricultural
Fosså, J., Mortensen, P., Furevik, D. (2002) The deep-water
coral Lophelia pertusa in Norwegian waters: distribution
and fishery impacts. Hydrobiologia 471: 1–12.
FAO (2009a) International Guidelines for the Management
of Deep-Sea Fisheries in the High Seas. Food and
Agricultural Organisation of the United Nations, Rome, Italy,
73pp.
84 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Fossen, I., Jørgensen, O.A., Gundersen, A.C. (2003)
Roughhead grenadier (Macruorus berglax) in the waters
off East Greenland: distribution and biology. Journal of
Northwest Atlantic Fishery Science 31: 285–298.
Francis, R.I.C.C., Clark, M.R. (2005) Sustainability issues
for orange roughy fisheries. Bulletin of Marine Science 76:
337–351.
Freese, J.L. (2001) Trawl-induced damage to sponges
observed from a research submersible. Marine Fisheries
Review 63: 7–13.
Freese, J.L., Auster, P.J., Heifetz, J., Wing, B.L. (1999)
Effects of trawling on seafloor habitat and associated
invertebrate taxa in the Gulf of Alaska. Marine Ecology
Progress Series 182: 119–126.
Freiwald, A., Beuck, L., Rüggerberg, A., Taviani, M.,
Hebbeln, D. (2009) The white coral community in the
central Mediterranean Sea revealed by ROV surveys.
Oceanography 22: 58–74.
Freiwald, A., Fosså, J.H., Grehan, A., et al. (2004) ColdWater Coral Reefs: Out of Sight – No Longer Out of
Mind, United Nations Environment Programme-World
Conservation Monitoring Centre, Cambridge, UK, 84pp.
Freiwald, A., Hühnerbach, V., Lindberg, B., Wilson, J.B.,
Campbell, J. (2002) The Sula Reef complex, Norwegian
Shelf. Facies 47: 179–200.
Fuller, E. Watling, L. (2008) Before the Secretary of
Commerce: Petition for a rule to list the U.S. population
of Atlantic Wolffish (Anarhichas lupus) as an endangered
species under the Endangered Species Act. Conservation
Law Foundation, Boston, USA, 112pp.
Galil, B., Zibrowius, H. (1998) First benthos samples
from Eratosthenes Seamount, Eastern Mediterranean.
Senckenbergiana Maritima 28: 111–121.
Gass, S.E., Willison, J.H.M. (2005) An assessment of the
distribution of deep-sea corals in Atlantic Canada by using
both scientific and local forms of knowledge. In: Freiwald,
A., Roberts, J.M. (eds) Cold-Water Corals and Ecosystems,
Springer-Verlag, Berlin Heidelberg, Germany, pp. 223–245.
Gedamke, T., DuPaul, W.D., Musick, J.A. (2005)
Observations on the life history of the barndoor skate,
Dipturus laevis, on George’s Bank (Western North Atlantic).
Journal of Northwest Atlantic Fishery Science 35: 67–78.
GFCM (2006) General Fisheries Commission for the
Mediterranean, Report of the Thirtieth Session, Istanbul,
Turkey, 24–27 January, 2006. GFCM Report 30. FAO, Rome,
Italy, 56pp.
GFCM (2009) General Fisheries Commission for the
Mediterranean, Report of the Thirty-third Session, Tunis
23–27 March, 2009. GFCM Report 33. FAO, Rome, Italy,
125pp.
GFCM Scientific Advisory Committee (2008a) Eleventh
Session, Marrakech, Morocco, 1–5 December, 2008,
Report of the Transversal Working Group on Bycatch/
Incidental Catches, Rome, Italy, 15–16 September, 2008.
GFCM:SAC11/2008/Inf.17. FAO, Rome, Italy, 24pp.
GFCM Scientific Advisory Committee (2008b) Eleventh
Session, Marrakech, Morocco, 1–5 December, 2008,
Criteria for the identification of sensitive habitats of
relevance for the management of priority species.
GFCM:SAC11/2008/Inf.20. FAO, Rome, 3pp.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 85
GFCM SCMEE (2008a) Sub-Committee on the Marine
Environment and Ecosystems, Report of the Ninth Meeting
of the SCSA/Working Group on Demersal Species, Izmir,
Turkey, 15–19 September, 2008. GFCM, Rome, Italy, 57pp.
GFCM SCMEE (2008b) General Fisheries Commission
for the Mediterranean, Scientific Advisory Committee,
Eleventh Session, Marrakech, Morocco, 1–5 December,
2008, Report of the Ninth Session of the Sub-Committee
on Marine Environment and Ecosystems (SCMEE) Antalya,
Turkey 13–16 October, 2008. GFCM:SAC11/2008/Inf.5,
FAO, Rome, Italy, 36pp.
Gianni, M. (2004) High seas bottom trawl fisheries and
their impacts on the biodiversity of vulnerable deep-sea
ecosystems: options for international action. International
Union for the Conservation of Nature (IUCN), Gland,
Switzerland.
Gibson, C., Valenti, S.V., Fordham, S.V., Fowler, S.L. (2008)
The Conservation of Northeast Atlantic Chondrichthyans:
Report of the IUCN Shark Specialist Group Northeast
Atlantic Red List Workshop. IUCN Species Survival
Commission Shark Specialist Group. Newbury, UK, viii +
76pp.
Gil, J., Sobrino, I., Baro, J., González, J., Afonso, P.S.,
Balguerias, E. (2008) Distribution and relative abundance
of main grenadiers (Macrouridae, Gadiformes) from
Mozambique. In: Orlov, A.M., Iwamoto, T. (eds) Grenadiers
of the World Oceans: Biology, Stock assessment, and
Fisheries. American Fisheries Society, Symposium 63,
Bethesda, Maryland, USA, pp. 103–118.
Gilkinson, K., Edinger, E. (eds) (2009) The ecology of
deep-sea corals of Newfoundland and Labrador waters:
biogeography, life history, biogeochemistry, and relations to
fishes. Canadian Technical Report of Fisheries and Aquatic
Sciences No. 2830, 136pp.
Gon, O., Heemstra, P.C. (eds) (2000) Fishes of the Southern
Ocean. J.L.B. Smith Institute of Ichthyology, Grahamstown,
South Africa, xviii + 462pp.
González, C., Teruel, J., López, E., Paz, X. (2007) Feeding
habits and biological features of deep-sea species of
the Northwest Atlantic: large-eyed rabbitfish (Hydrolagus
mirabilis), narrownose chimaera (Harriotta raleighana)
and black dogfish (Centroscyllium fabricii). NAFO Scientific
Council Meeting – September, 2007, Serial No. N5423,
NAFO SCR Doc. 07/63, 9pp.
González-Porto, M. (2008) Resultados preliminares
del estudio de los invertebrados bentónicos de la
campaña NAMIBIA-0802. In: Preliminary Report of the
Multidisciplinary Research Cruise on the Walvis Ridge
Seamounts (Atlantic Southeast-SEAFO). Spanish Institute
of Oceanographny of the Canaries and National Marine
Information and Research Centre, Nambia, pp. 85–94.
González-Troncoso, D., Paz, X. (2007) Biomass and length
distribution for roughhead grenadier, thorny skate and white
hake from the surveys conducted by Spain in NAFO 3NO.
Scientific Council Meeting – June 2007, Serial No. N5500,
NAFO SCR Doc. 08/009, 38pp.
González-Troncoso, D., Paz, X., Cardoso, X. (2006)
Persistence and variation in the distribution of
bottom-trawl fish assemblages over the Flemish Cap.
Journal of Northwest Atlantic Fishery Science 37: 103–
117.
Government of New Zealand (2008a) Preliminary
assessment of bottom fishing activities in 2008–2009
for New Zealand: implementing Conservation Measure
22-06 (Bottom Fishing in the Convention Area). Impact
Assessment submitted to CCAMLR, CCAMLR-XXVII/26*,
24 September 2008. Original: English Agenda Item No. 5,
SC Agenda Item No. 4, 70pp.
Government of New Zealand (2008b) New Zealand
National Report on Fishing and Research Activities in the
SPRFMO area during 2008. SP-08-SWG-08. Ministry of
Fisheries, 9pp.
Government of New Zealand (2009) Bottom Fishing
Activities by New Zealand Vessels Fishing in the High Seas
in the SPRFMO Area During 2008 and 2009. Ministry of
Fisheries, 102pp.
Government of Spain (2008) The Fisheries of Spain in the
Regional Organization of Management of Fisheries in the
Pacific South (SPRFMO) During the Season 2009/2010.
Preliminary Assessment of the Risk of Cause Serious
Damage to Vulnerable Marine Ecosystems and Protocol of
Action. SP-08-SWG-DW-02. Ministerio De Medio Ambiente Y
Medio Rural Y Marino, 12pp.
Grant, S., Constable, A., Raymond, B., Doust, S. (2006)
Bioregionalisation of the Southern Ocean: Report of
Experts Workshop, Hobart, September 2006. WWFAustralia and ACE CRC, 46pp.
Grant, S.M. (2006) An exploratory fishing survey and
biological resource assessment of Atlantic hagfish
(Myxine glutinosa) occurring on the southwest slope of the
Newfoundland Grand Bank. Journal of Northwest Atlantic
Fishery Science 36: 91–110.
Henry, L‐A., Hart, M. (2005) Regeneration from injury and
resource allocation in sponges and corals – a review.
International Review of Hydrobiology 90: 125–158.
Klitgaard, A.B. (1995) The fauna associated with outer
shelf and upper slope sponges (Porifera, Demospongiae) at
the Faroe Islands, Northeast Atlantic. Sarsia 80: 1–22.
Hibberd, T., Moore, K. (2009) Field identification guide to
Heard Island and MacDonal Island benthic invertebrates:
a guide for scientific observers aboard fishing vessels.
Australian government, Australia Antarctic Research
Division, Hobart, Tasmania, Australia, 152pp.
Klitgaard, A.B., Tendal, O.S. (2004) Distribution and
species composition of mass occurrences of large
sized sponges in the Northeast Atlantic. Progress in
Oceanography 61: 57–98.
Humphreys, R.L., Jr. (2008) Sources of information for
evaluating the existence of Vulnerable Marine Ecosystems
(VMEs) on seamounts within the Southern EmperorNorthern Hawaiian Ridge. PIFSC Working Paper WP-08-003,
Scientific Working Group Meeting, North Pacific RFMO,
Vladivostok, Russia, 8pp.
ICES (2001) Report of the ICES Advisory Committee on
Fishery Management, Copenhagen 22–31 May, 2001,
Copenhagen 9–17 October 2001. ICES Cooperative
Research Report No. 246, Part 1 of 3, 895pp.
ICES (2007a) 9.3.2.7.b NEAFC request to compile
data on documented historical or present spawning/
aggregation areas of blue ling in the NEAFC Convention
area – additional information supplied by the EC POORFISH
project. ICES Advice Book 9, Agenda Item 7F, Document
AM2007/27, 16pp.
ICES (2007b) Widely Distributed and Migratory Stocks.
9.3.2.4. NEAFC request to continue to provide all available
new information on distribution of vulnerable habitats in
the NEAFC Convention Area and fisheries activities in the
vicinity of such habitats. ICES Advice Book 9, 9pp.
Kock, K.H., Reid, K., Croxall, J., Nicol, S. (2007) Fisheries
in the Southern Ocean: an ecosystem approach.
Philosophical Transactions of the Royal Society B Biological
Sciences 362: 2333–2349.
Koslow, J.A., Boehlert, G.W., Gordon, J.D.M., Haedrich, R.L.,
Lorance, P., Parin, N. (2000) Continental slope and deepsea fisheries: implications for a fragile ecosystem. ICES
Journal of Marine Science 57: 548–557.
Koslow, J.A., Gowlett-Holmes, K. (1998) The seamount
fauna off southern Tasmania: benthic communities, their
conservation and impacts of trawling: Final Report to
Environment Australia and the Fisheries Research and
Development Corporation. Rep. FRDC Project 95/058,
CSIRO, Hobart, Tasmania, Australia, 104pp.
Koslow, J.A., Gowlett-Holmes, K., Lowry, J.K., et al. (2001)
Seamount benthic macrofauna off southern Tasmania:
community structure and impacts of trawling. Marine
Ecology Progress Series, 213: 111–125.
Koslow, J.A., Tuck, G. (2001) The boom and bust of
deep-water fisheries: Why haven’t we done better? NAFO
Scientific Research Council Doc. 141 Ser. 4535, 10pp.
Grigg, R.W. (1982) Precious corals in the Pacific: Economic
and development potential. Infofish 2: 8–11.
ICES WGEF (2007) Report of the Working Group on
Elasmobranch Fishes (WGEF). ICES Advisory Committee on
Fishery Management. ICES CM 2007/ACFM: 27.REF.LRC,
318pp.
Krieger, K. (1998) Primnoa spp. observed inside and
outside a bottom trawl path from a submersible. Abstract.
10th Western Groundfish Conference, Asilomar, California,
USA.
Hall-Spencer, J., Allain, V., Fosså, J.H. (2002) Trawling
damage to Northeast Atlantic ancient coral reefs.
Proceedings of the Royal Society of London, Series B:
Biological Sciences 269: 507–511.
Kelly, C.J., Connolly, P.L., Bracken, J.J. (1996) Maturity,
oocyte dynamics and fecundity of the roundnose grenadier
from the Rockall Trough. Journal of Fish Biology 49 (Suppl.
A): 5–17.
Hall-Spencer, J.M., Rogers, A.D., Davies, J., Foggo, A.
(2007) Historical deep-sea coral distribution on
seamount, oceanic island and continental shelf-slope
habitats in the NE Atlantic. In: George, R.Y., Cairns,
S.D. (eds) Conservation and Adaptive Management of
Seamount and Deep-Sea Coral Ecosystems. Rosenstiel
School of Marine and Atmospheric Science, University of
Miami, Miami, USA, pp. 135–146.
Kelly, C.J., Connolly, P.L., Bracken, J.J. (1997) Age
estimation, growth, maturity and distribution of the
roundnose grenadier from the Rockall Trough. Journal of
Fish Biology 50: 1–17.
Krieger, K.J. (2001) Coral (Primnoa) impacted by fishing
gear in the Gulf of Alaska. In: Willison. J.H.M., Gass, S.E.,
Kenchington, E.L.R., et al. (eds) Proceedings of the First
International Symposium on Deep-Sea Corals, Ecology
Action Centre and Nova Scotia Museum, Halifax, Nova
Scotia, Canada, pp. 106–116.
Hanchet, S., O’Driscoll, R., Ballara, S., Dunn, A. (2008)
Grenadier bycatch in the toothfish longline fishery in
the Ross Sea, Antarctica. In: Orlov, A.M., Iwamoto, T.
(eds) Grenadiers of the World Oceans: Biology, Stock
Assessment, and Fisheries. American Fisheries Society,
Symposium 63, Bethesda, Maryland, USA, pp. 451–462.
Hareide, N-R., Garnes, G., Rihan, D., Mulligan, M., Tyndall,
P., Clark, M., Connolly, P., Misund, R., McMullen, P., Furevik,
D.M., Humborstad, O-B., Høydal, K., Blasdale, T. (2004) A
Preliminary Investigation on Shelf Edge and Deep-water
Fixed Net Fisheries to the West and North of Great Britain,
Ireland, around Rockall and Hatton Bank.
Heifetz, J., Stone, R.P., Shotwell, S.K. (2009) Damage and
disturbance to coral and sponge habitat of the Aleutian
Archipelago. Marine Ecology Progress Series 397: 295303.
86 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Kenchington, E., Cogswell, A., Lirette, C., Murillo-Perez, F.J.
(2009a) The use of density analyses to delineate sponge
grounds and other benthic VMEs from trawl survey data.
NAFO Scientific Council Meeting – June, 2009 Serial No.
N5626, NAFO SCR Doc. 09/6, 15pp.
Kenchington, E., Best, M., Cogswell, A., MacIsaac, K.,
Murillo-Perez, F. J., MacDonald, B., Wareham, V., Fuller, S.
D., Jørgensbye, H. I. Ø., Sklya V., Thompson, A. B. (2009b)
Coral Identification Guide NAFO Area. Sci. Coun. Studies,
42: 1-35. doi:10.2960/S.v42.m1.
Kimani, E.N., Okemwa, G.M., Kazungu, J.M. (2009)
Fisheries in the Southwest Indian Ocean: Trends and
Governance Challenges. In: Laipson, E., Pandya, A. (eds)
The Indian Ocean; Resource and Governance Challenges.
The Henry L. Stimson Centre, Washington, DC, USA, pp.
3–90.
Klitgaard, A.B. (1991) Gnathia abyssorum (G.O. Sars,
1872) (Crustacea, Isopoda) associated with sponges.
Sarsia 76: 33–39.
Kulka, D.W. (2006) Abundance and distribution of demersal
sharks on the Grand Banks with particular reference to the
NAFO Regulatory Area. Scientific Council Meeting – June
2006. Serial No. N5237, NAFO SCR Doc. 06/20, 36pp.
Kulka, D.W., Miri, C., Thompson, A.B. (2007a) Identification
of wolffish, hake and rockling in the Northwest Atlantic.
Northwest Atlantic Fisheries Organization Scientific Council
Studies 40, 5pp.
Kulka, D., Hood, C., Huntington, J. (2007b) Recovery
Strategy for Northern Wolffish (Anarhichas denticulatus)
and Spotted Wolffish (Anarhichas minor), and Management
Plan for Atlantic Wolffish (Anarhichas lupus) in Canada.
Fisheries and Oceans Canada: Newfoundland and Labrador
Region. St. John’s, NL, Canada, x + 103pp.
Kulka, D.W., Simpson, M.R., Inkpen, T.D. (2003)
Distribution and biology of the blue hake (Antimora rostrata
Günther 1878) in the Northwest Atlantic with comparison
to adjacent areas. Journal of Northwest Atlantic Fishery
Science 31: 299–318.
Kulka, D.W., Templeman, N., Janes, J., Power, A., Brodie, W.
(2007c) Information on seamounts in the NAFO Convention
Area. NAFO Scientific Council Meeting – June 2007, Serial
No. N5414, NAFO SCR Doc. 07/61, 17pp.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 87
Lack, M., Short, K., Willock, A. (2003) Managing risk and
uncertainty in deep-sea fisheries: lessons from Orange
Roughy. TRAFFIC Oceania and WWF Endangered Seas
Programme, 72pp.
le Danois, E. (1948) Les profondeurs de la mer, trente ans
de recherches sur la faune sous‐marine au large des côtes
de France, Payot, Paris, France.
Leys, S.P., Lauzon, N.R.J. (1998) Hexactinellid sponge
ecology: growth rates and seasonality in deep-water
sponges. Journal of Experimental Marine Biology and
Ecology 230: 111–129.
Lockhart, S.J., Jones, C.D. (2009) Detection of vulnerable
marine ecosystems in the Southern Scotia Arc (CCAMLR
Sub-Areas 48.1 and 48.2) through research bottom trawl
sampling and underwater imagery. CCAMLR Document WGEMM-09/32, 27pp.
MacIsaac, K., Bourbonnais, C., Kenchington, E., Gordon,
D. Jr, Gass, S. (2001) Observations on the occurrence and
habitat preference of corals in Atlantic Canada. In: Willison,
J.H.M., Hall, J., Gass, S.E., Kenchington, E.L.R., Butler,
M., Doherty, P. (eds) Proceedings of the First International
Symposium on Deep-Sea Corals, Halifax, Canada, pp.
58–75.
MCBI (2009a) Seamount Map of the Indian Ocean (by John
Guinotte), 3pp.
MCBI (2009b) Global habitat suitability for reef-forming
cold-water corals. Final Report to the Pew Charitable Trusts,
12pp.
McPhie, R.P., Campana, S.E. (2009) Reproductive
characteristics and population decline of four species of
skate (Rajidae) off the eastern coast of Canada. Journal of
Fish Biology 75: 223–246.
Merrett, N.R., Haedrich, R.L. (1997) Deep-sea demersal
fish and fisheries. Chapman & Hall, London, UK, 282pp.
Mitchell, R.E., Peatman, T., Pearce, J., Agnew, D.A. (2009)
Analysis of VME data collected by UK vessels fishing in the
Ross Sea during the 2008/09 CCAMLR season. CCAMLR
WS-VME-09/5, 18 July, 2009, Agenda item Nos 5, 6.2.,
10pp.
features in the Northeast Channel (Atlantic Canada).
Marine Biology 144: 1223–1238.
Mortensen, P.B., Buhl-Mortensen, L. (2005) Deep-water
corals and their habitats in The Gully, a submarine canyon
off Atlantic Canada. In: Freiwald, A., Roberts, J.M. (eds)
Cold-Water Corals and Ecosystems, Springer-Verlag, Berlin
Heidelberg, pp. 247–277.
Mortensen, P.B., Buhl- Mortensen, L., Gordon, D.C., et
al. (2005) Evidence of fisheries damage to deep-water
gorgonians in the Northeast Channel, Nova Scotia.
In: Thomas, J., Barnes, P. (eds) Proceeding from the
Symposium on the Effects of Fishing Activities on Benthic
Habitats: Linking Geology, Biology, Socioeconomics and
Management, American Fisheries Society, November 12–
14, 2002, Florida, USA.
Mortensen, P.B., Buhl-Mortensen, L., Gebruk, A.V., Krylova,
E.M. (2008) Occurrence of deep-water corals on the
Mid-Atlantic Ridge based on MAR-ECO data. Deep-Sea
Research II 55: 142–152.
MSC (2006) MSC Assessment Report, The Australian
Antarctic Mackerel Icefish Fishery at Heard and MacDonald
Islands. Final Report. Scientific Certification Systems Inc.,
Marine Fisheries Conservation Programme, Emeryville,
California, USA, 273pp.
Muñoz, P.D., Mandado, M., Gago, A., Gómez, C., Fernández,
G. (2005) Brief results of a trawl experimental survey
at NW Atlantic. Scientific Council Meeting – June 2005
Northwest Atlantic Fisheries Organisation, Serial No.
N5095, NAFO SCR Doc. 05/32, 4pp.
Muñoz, P.D., Murillo, F.J., Serrano, A., Sayago-Gil, M., Parra,
S., Díaz del Río, V., Sacau, M., Patrocinio, T., Cristobo,
J. (2008) A case study of available methodology for the
identification of vulnerable ecosystems/habitats in bottom
deep-sea fisheries: possibilities to apply this method in the
NAFO Regulatory Area in order to select marine protected
areas. NAFO Scientific Committee Working Group on the
Ecosystem Approach to Fisheries Management – May
2008. Serial No. N5491, NAFO SCR Doc. 08/6, 20pp.
Moody Marine Ltd (2008) Public Comment Draft Report for
Ross Sea Toothfish Longline Fishery. Client: Argos Georgia
Ltd, Sanford Ltd, NZLL Ltd, Ref. 82044/Vers. 3, Moody
Marine Ltd, Derby, UK, 79pp + Appendices.
Murillo, F.J., Muñoz, P.D., Sacau, M., González-Troncoso, D.,
Serrano, A. (2008) Preliminary data on cold-water corals
and large sponges bycatch from Spanish/EU bottom trawl
groundfish surveys in NAFO Regulatory Area (Divs. 3LMNO)
and Canadian EEZ (Div. 3L): 2005–2007 period. NAFO
Scientific Committee Working Group on the Ecosystem
Approach to Fisheries Management. Serial No. 5501, NAFO
SCR doc. 08/10, 28pp.
Morato, T., Cheung, W.W.L., Pitcher, T.J. (2006) Vulnerability
of seamount fish to fishing: fuzzy analysis of life history
attributes. Journal of Fish Biology 68: 209–221.
Murua, H., de Cárdenas, E. (2005) Depth distribution of
deepwater species in Flemish Pass. Journal of Northwest
Atlantic Fishery Science 37: 1–12.
Morato, T., Clark, M.R. (2007) Seamount fishes: ecology
and life histories. In: Pitcher T.J., Morato, T., Hart, P.J.B.,
Clark, M.R., Haggan, N., Santos, R.S. (eds) Seamounts:
Ecology, Fisheries & Conservation, Fish and Aquatic
Resources Series 12, Blackwell Publishing, Oxford, UK, pp.
170–188.
Murua, H., González, F., Power, D. (2005) A review of the
fishery and the investigations of roughhead grenadier
(Macrourus berglax) in Flemish Cap and Flemish Pass.
Journal of Northwest Atlantic Fishery Science 37: 13–27.
Morin, B., Méthot, R., Sevigny, J-M., Power, D., Branton, B.,
McIntyre, T. (2004) Review of the structure, the abundance
and distribution of Sebastes mentella and S. fasciatus in
Atlantic Canada in a species-at-risk context. CSAS Res.
Doc. 2004/058, 96 pp.
Mortensen, P.B., Buhl-Mortensen, L. (2004) Distribution of
deep-water gorgonian corals in relation to benthic habitat
NAFO CEM (2009) Northwest Atlantic Fisheries
Organization Conservation and Enforcement Measures.
Serial Number N5614, NAFO/FC Doc. 09/01. Northwest
Atlantic Fisheries Organisation, Dartmouth, Nova Scotia,
Canada, 92pp.
NAFO TACs (2009) Annual Quota Table (Annex I.A.) and
Effort Allocation Scheme (Annex I.B.): http://www.NAFO.
int/about/frames/about.html
88 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
NAFO Fisheries Commission (2005) Report of the Fisheries
Commission Meeting (FC Doc. 05/15), 27th Annual
Meeting, 19–23 September 2005, Tallinn, Estonia, Part 1,
pp. 113–174.
NAFO Fisheries Commission (2008) Report of the Fisheries
Commission 30th Annual Meeting, 22–26 September
2008, Vigo, Spain. Serial No. N5613 NAFO/FC Doc.
08/22. Northwest Atlantic Fisheries Organization, 86pp.
NAFO Fisheries Commission (2009) Report of the Fisheries
Commission and its Subsidiary Body (STACTIC), 31st
Annual Meeting 21–25 September 2009 Bergen Norway.
Serial No. N5735 NAFO/FC Doc. 09/21. Northwest
Atlantic Fisheries Organization, 90pp.
NAFO SC (2005) Northwest Atlantic Fisheries Organization,
Scientific Council Reports 2005, Dartmouth, Nova Scotia,
Canada, 359pp.
NAFO SC (2008) Northwest Atlantic Fisheries Organization,
Scientific Council Reports 2008, Dartmouth, Nova Scotia,
Canada, 328pp.
Navarro, S.J., Alayón, P.J.P., Abellán, L.J.L., Holtzhausen,
H., Jiménez, J.F.G. (2008) Seamount Associated
Species Namibia 08-02. In: Preliminary Report of the
Multidisciplinary Research Cruise on the Walvis Ridge
Seamounts (Atlantic Southeast-SEAFO). Spanish Institute
of Oceanographny of the Canaries and National Marine
Information and Research Centre, Nambia. Annex III, pp.
99–191.
NEAFC (2004) Monthly catches of regulated resources.
North East Atlantic Fisheries Commission, 4pp.
NEAFC (2006a) Recommendation IX – 2007:
Recommendation by the North East Atlantic Fisheries
Commission at its Annual Meeting in November 2006 to
Adopt Conservation and Management Measures by Closing
Certain Areas in the Regulatory Area to Protect Deep-Water
Corals. NEAFC, London, UK, 2pp.
NEAFC (2006b) Agenda Item 4b AM2007/47 rev2. North
East Atlantic Fisheries Commission, 6pp.
NEAFC (2007a) Recommendation IX – 2008:
Recommendation by the North East Atlantic Fisheries
Commission in Accordance with Article 5 of the Convention
on Future Multilateral Cooperation in North East Atlantic
Fisheries at its Annual Meeting in November 2007 to
Adopt Conservation and Management Measures by Closing
Certain Areas in the Regulatory Area in Order to Protect
Deep-Water Corals. NEAFC, London, UK, 2pp.
NEAFC (2007b) Agenda Item 4a AM2008/59 rev 1. North
East Atlantic Fisheries Commission, 6pp.
NEAFC (2008) Recommendation XIII: 2009
Recommendation by the North East Atlantic Fisheries
Commission in Accordance With Article 5 of the Convention
on Future Multilateral Cooperation in North East Atlantic
Fisheries at its Annual Meeting on 10–14 November 2008
to Adopt the Following Recommendation on Operational
Procedures for Fishing in Existing and New Bottom Fishing
Areas. NEAFC, London, UK, 3pp.
NEAFC (2009a) NEAFC Fisheries Status Report 1998–
2007, Kjartan Hoydal (ed). NEAFC, London, UK, 47pp.
to the 28th Annual Meeting, London, 9–13 November
2009, 7pp.
NEAFC (2010a) Report of the 28th Annual Meeting of the
North East Atlantic Fisheries Commission 9–13 November
2009. North East Atlantic Fisheries Commission, London,
25pp + Annexes (113pp total).
NEAFC (2010b) Statistics on quota uptake 2008
Final. AM2009/62 rev.3. North East Atlantic Fisheries
Commission, London, UK, 9pp.
NOAA (2008) North West Pacific Ocean; Reports on
identification of VMEs and assessment of impacts caused
by bottom fishing activities on VMEs and marine species.
NOAA Fisheries, USA, December 23 2008, 47pp.
Norwegian government (2008) Proposal for revision of
areas closed to bottom fisheries in the NEAFC RA on the
mid‐Atlantic Ridge and in the adjacent abyssal plains.
Royal Ministry of Fisheries and Coastal Affairs. Annex 2 to
the Report of the Meeting of the Permanent Committee on
Management and Science (PECMAS), 1718 June, 2008,
NEAFC, London, UK, 13pp.
North Pacific Fisheries Commission (2007) Second Intergovernmental Meeting on Management of High Seas
Bottom Fisheries in the North Western Pacific Ocean
Busan, Republic of Korea 31 January – 2 February
2007 Establishment of new mechanisms for protection
of vulnerable marine ecosystems and sustainable
management of high seas bottom fisheries in the North
Western Pacific Ocean. North Pacific Fisheries Commission,
Japan, 6pp.
O’Dea, N.R., Haedrich R.L. (2003) A review of the status
of the Atlantic Wolffish, Anarhichas lupus, in Canada.
Canadian Field-Naturaliste 116: 423–433.
Ordines, F., Massutí, E. (2009) Relationships between
macro-epibenthic communities and fish on the shelf
grounds of the western Mediterranean. Aquatic
Conservation: Marine and Freshwater Ecosystems 19:
370–383.
Orejas, C., López-González, P.J., Gili, J.M., et al. (2002)
Distribution and reproductive ecology of the Antarctic
octocoral Ainigmaptilon antarcticum in the Weddell Sea.
Marine Ecology Progress Series, 231: 101–114.
Parker, S.J., Mormede, S., Tracey, D.M., Carter, M. (2009)
Evaluation of VME taxa monitoring by observers from five
vessels in the Ross Sea region Antarctic toothfish longline
fisheries during the 2008–09 season. CCAMLR TASO09/8, 19 June 2009, Agenda Item No. 3.2.2., 13pp.
Parker, S.J., Penney, A.J., Clark, M.R. (2009) Detection
criteria for managing trawl impacts in vulnerable marine
ecosystems in high seas fisheries of the South Pacific
Ocean. Marine Ecology Progress Series 397: 309-317.
Parrish, F.A., Abernathy, K., Marshall, G.J., et al. (2002)
Hawaiian monk seals (Monachus schauninslandi) foraging
in deep-water coral beds. Marine Mammal Science 18:
244–258.
PECMAS (2008a) Permanent Committee on Management
and Science of the North-East Atlantic Fisheries
Commission Meeting 28–29 October 2008 at NEAFC
Headquarters, London, UK, 13pp + Annexes.
NEAFC (2009b) Scheme of Control and Enforcement. North
East Atlantic Fisheries Commission, London, 96pp.
NEAFC press release (2009) NEAFC press release relating
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 89
PECMAS (2008b) Mapping Existing Fisheries Areas in the
NEAFC Regulatory Area. Agenda Item 6; AM2008/53 rev
1. North East Atlantic Fisheries Commission, London, UK,
9pp.
PECMAS (2009) Permanent Committee on Management
and Science of the North East Atlantic Fisheries
Commission Meeting 28–29 September, at NEAFC
Headquarters, London, UK, 14pp + Annexes.
Penhale, P., Grant, S. (2007) Workshop on the
bioregionalisation of the Southern Ocean, (Brussels,
Belgium, 13–17 August 2007), Executive Summary, Report
of the Workshop, Annex 9, Report of the Twenty-Sixth
Meeting of the Scientific Committee, Hobart, Australia, 22–
26 October, 2007, SC-CCAMLR-XXVI, Scientific Committee
for the Conservation of Antarctic Marine Living Resources,
Hobart, Tasmania, Australia, pp. 590–676.
Penney, A.J., Parker, S.J., Brown, J.H. (2009) Protection
measures implemented by New Zealand for vulnerable
marine ecosystems in the South Pacific Ocean. Marine
Ecology Progress Series 397: 341-354.
Ramírez-Llodra, E., Ballesteros, M., Company, J.B., Dantart,
L., Sardà, F. (2008) Spatio-temporal variations of biomass
and abundance in bathyal non-crustacean megafauna in
the Catalan Sea (northwestern Mediterranean). Marine
Biology 153: 297–309.
Reed, J.K., Shepard, A.N., Koenig, C.C., et al. (2005)
Mapping, habitat characterization, and fish surveys of the
deep-water Oculina coral reef Marine Protected Area: a
review of historical and current research. In: Freiwald, A.,
Roberts, J.M. (eds) Cold-Water Corals and Ecosystems,
Springer-Verlag, Berlin Heidelberg, pp. 443–465.
Republic of Korea (2008) Reports on identification of VMEs
and assessment of impacts caused by bottom trawl fishing
activities on VMEs and/or marine species. Ministry for
Food, Agriculture, Forestry and Fisheries, Republic of Korea,
22 December 2008, 8pp.
Rice, A. L., Thurston, M. H., New, A. L. (1990) Dense
aggregations of a hexactinellid sponge, Pheronema
carpenteri, in the Porcupine Seabight (northeast Atlantic
Ocean), and possible causes. Progress in Oceanography
24: 179–196.
Roark, E.B., Guilderson, T.P., Dunbar, R.B., Ingram, B.L.
(2006) Radiocarbon-based ages and growth rates of
Hawaiian deep-sea corals. Marine Ecology Progress Series
327: 1–14.
Roark, E.B., Guilderson, T.P., Dunbar, R.B., et al. (2009)
Extreme longevity in proteinaceous deep-sea corals.
Proceedings of the National Academy of Sciences USA
106: 5204–5208.
Roberts, J.M., Harvey, S.M., Lamont, P.A., et al. (2000)
Seabed photography, environmental assessment and
evidence for deep-water trawling on the continental margin
west of the Hebrides. Hydrobiologia 441: 173–183.
Roberts, J.M., Wheeler, A., Freiwald, A., et al. (2009) Cold
Water Corals: The Biology and Geology of Deep-Sea Coral
Habitats, Cambridge University Press, Cambridge, UK,
352pp.
Rodgeveller, C.J., Clausen, D.M., Nagler, J.J., Hutchinson,
C. (2010) Reproductive characteristics and mortality of
female giant grenadiers in the northern Pacific Ocean.
Marine and Coastal Fisheries: Dynamics, Management,
and Ecosystem Science 2: 73-82.
Rogers, A.D. (1994) The biology of seamounts. Advances
in Marine Biology 30: 305–350.
Rogers, A.D. (1999) The biology of Lophelia pertusa
(Linnaeus 1758) and other deep-water reef-forming corals
and impacts from human activities. International Review of
Hydrobiology 84: 315–406.
Rogers, A.D., Baco, A., Griffiths, H.J., et al. (2007) Corals
on seamounts. In: Pitcher T.J., Morato, T., Hart, P.J.B., Clark,
M.R., Haggan, N., Santos, R.S. (eds) Seamounts: Ecology,
Fisheries & Conservation, Fish and Aquatic Resources
Series 12, Blackwell Publishing, Oxford, UK, pp. 141–169.
Rogers, A.D., Taylor, M.L., Kemp, K., Yesson, C., Davies, A.J.
(in press) The Diseases of Deep-Water Corals. In: Downs,
C. (ed.) Diseases of Corals. CRC Press, London / New
York. In press. 97pp (ms) + Figs.
Romanov, E.V. (ed.) (2003) Summary and Review of Soviet
and Ukrainian scientific and commercial fishing operations
on the deepwater ridges of the Southern Indian Ocean. FAO
Fisheries Circular No. 991, 84pp.
Rosso, A., Vertino, A., Di Geronimo, I., Sanfilippo, R., Sciuto,
F., Di Geronimo, R., Violanti, D., Corselli, C., Taviani, M.,
Mastrototaro, F., Tursi, A. (2009) Hard- versus soft-bottom
thanatofacies from the Santa Maria di Leuca deep-water
coral mound province, Recent Mediterranean. Deep-Sea
Research II, doi:10.1016/j.dsr2.2009.08.024.
Rowden, A.A., Clark, M.R., O’Shea, S. (2004) The influence
of deep-water coral habitat and fishing on benthic faunal
assemblages of seamounts on the Chatham Rise, New
Zealand. ICES CM2004/AA:09.
Saidi B., Bradai, M.N. (2008) Elasmobranchs Bycatch
in the Mediterranean Sea. GFCM Scientific Advisory
Committee, SCMEE/SCSA transversal working group on by
catch/incidental catches. General Fisheries Commission
for the Mediterranean, FAO Headquarters, Rome, Italy,
14pp.
Sakiura, H. (1972) The pelagic armourhead Pentaceros
richardsoni, fishing grounds off the Hawaiian Islands, as
viewed by the Soviets. Suisan Shuho 658: 28–31.
Sardà, F., Cartes, J.E., Company, J.B. (1994) Spatiotemporal variations in megabenthos abundance in three
different habitats of the Catalan deep-sea (Western
Mediterranean). Marine Biology 120: 211–219.
SEAFO Commission (2006) Report of the 3rd Annual
Meeting of the Commission, 2006. Walvis Bay, Namibia,
63pp.
SEAFO Commission (2007) Report of the 4th Annual
Meeting of the Commission. Walvis Bay, Namibia, 48pp.
SEAFO Commission (2008) Report of the 5th Annual
Meeting of the Commission. Walvis Bay, Namibia, 59pp.
SEAFO Commission (2009) Report of the 6th Annual
Meeting of the Commission. Walvis Bay, Namibia, 176pp.
SEAFO Scientific Committee (2006) Report of the SEAFO
Scientific Committee 2006. Walvis Bay, Namibia, 59pp.
SEAFO Scientific Committee (2008) Report of the SEAFO
Scientific Committee 2008. Walvis Bay, Namibia, 69pp.
SEAFO Scientific Committee (2009) Report of the SEAFO
Scientific Committee 2009. Walvis Bay, Namibia, 78pp.
90 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
Sherwood, O., Edinger, E. (2009) Carbon-14 composition
of deep-sea corals of Newfoundland and Labrador: proxy
records of seawater 14C and quantification of deep-sea
coral growth rates. In: Gilkinson, K., Edinger, E. (eds) The
ecology of deep-sea corals of Newfoundland and Labrador
waters: biogeography, life history, biogeochemistry, and
relation to fishes. Canadian Technical Report of Fisheries
and Aquatic Sciences No. 2830, pp. 74–84.
Shester, G., Ayers, J. (2005) A cost effective approach to
protecting deep-sea coral and sponge ecosystems with an
application to Alaska’s Aleutian Islands region. In: Freiwald,
A., Roberts, J.M. (eds) Cold-Water Corals and Ecosystems,
Springer-Verlag, Berlin Heidelberg, pp. 1151–1169.
Shibanov, V.N., Vinnichenko, V.I. (2008) Russian
investigations and the fishery of roundnose grenadier in the
North Atlantic. In: Orlov, A.M., Iwamoto, T. (eds) Grenadiers
of the World Oceans: Biology, Stock assessment, and
Fisheries. American Fisheries Society, Symposium 63,
Bethesda, Maryland, USA, pp. 399–412
Shotton, R. (2006) Managment of demersal fisheries
resources of the Southern Indian Ocean. FAO Fisheries
Circular No. 1020, FAO, Rome, Italy, 90pp.
Sissenwine, M.P., Mace, P.M. (2007) Can deep water
fisheries be managed sustainably? In: Report and
documentation of the expert consultation on deep-sea
fisheries in the high seas. FAO Fisheries Reports 838:
61–111.
Smith, D.C., Stewart, B.D. (1994) Development of methods
to age commercially important dories and oreos. Final
Report, Australian Fisheries Research and Development
Corporation, Project 91/36.
Soriano, S., Sánchez-Lizaso, J.L. (2000) Discards of
the upper slope trawl fishery off Alicante province (W.
Mediterranean). 6th Mediterranean Symposium on
Seabirds. Conference on Fisheries, Marine Productivity
and Conservation of Seabirds. Benidorm, Spain. 11–15
October 2000. Book of abstracts, p. 73.
SPRFMO Interim Measures (2007) INTERIM MEASURES
ADOPTED BY PARTICIPANTS IN NEGOTIATIONS TO
ESTABLISH SOUTH PACIFIC REGIONAL FISHERIES
MANAGEMENT ORGANISATION. International Consultations
on the Establishment of the Proposed South Pacific
Regional Fisheries Management Organisation. 3rd Meeting,
Renaca, Chile. 4 May 2007.
SPRFMO (2008) Revised Draft Bottom Fishery Impact
Assessment Standard. Eighth International Meeting:
Science Working Group, SP-08-SWG-DW-01, 39pp.
Stockley, B.M., Menezes, G., Pinho, M.R., Rogers, A.D.
(2005) Genetic population structure of the black-spot sea
bream (Pagellus bogaraveo) from the NE Atlantic. Marine
Biology 146: 793–804.
Stone, R.P. (2006) Coral habitat in the Aleutian Islands of
Alaska: depth distribution, fine-scale species associations,
and fisheries interactions. Coral Reefs 25: 229–238.
Stone, R.P., Masuda, M.M., Malecha, P.W. (2005) Effects of
bottom trawling on soft sediment epibenthic communities
in the Gulf of Alaska. In: Barnes, P.B., Thomas, J.P. (eds)
Benthic habitats and the effects of fishing. American
Fisheries Society, Symposium 41, Bethesda, Maryland,
USA, pp. 461–475.
SWIOFC (2005) South West Indian Ocean Fisheries
Commission, First Session, Mombasa, Kenya, 18–20 April,
2005. Resolution and Statues of the South West Indian
Ocean Fisheries Commission. SAFR/DM/SWIOFC/05/INF
4 E, FAO Rome, Italy, 5pp.
SWIOFC (2009) South West Indian Ocean Fisheries
Commission, Report of the Third Session of the Scientific
Committee, Maputo, Mozambique, 16–19 September
2008, 85pp.
Taviani, M., Freiwald, A., Zibrowius, H. (2005) Deep coral
growth in the Mediterranean Sea: An overview. In: Freiwald,
A., Roberts, J.M. (eds) Cold-water Corals and Ecosystems.
Springer, Heidelberg, Germany, pp. 137–156
SPRFMO (2007a) Information describing orange roughy
Hoplostethus atlanticus fisheries relating to the South
Pacific Regional Fisheries Management Organisation,
Working Draft 4 May 2007. SPRFMO, 23pp.
Taviani, M., Remia, A., Corselli, C., Freiwald, A., Malinverno,
E., Mastrototaro, F., Savini, A., Tursi, A. (2005a) First geomarine survey of living cold-water Lophelia reefs in the
Ionian Sea (Mediterranean basin). Facies 50: 409–417.
SPRFMO (2007b) Information describing Oreosomatidae
(Allocytus niger, Neocyttus rhomboidalis, and Pseudocyttus
maculates) fisheries relating to the South Pacific Regional
Fisheries Management Organisation, Working Draft, 2
September, 2007. SPRFMO-IV-SWG-11, 16pp.
Thompson, A.B., Campanis, G.M. (2007) Information on
fishing on and around the four closed seamount areas in
the NRA. Scientific Council Meeting – June 2007 Northwest
Atlantic Fisheries Organisation, Serial No. N5347, NAFO
SCR Doc. 07/06, 10pp.
SPRFMO (2007c) Information describing black cardinalfish
(Epigonus telescopus) fisheries relating to the South Pacific
Regional Fisheries Management Organisation, Working
Draft, 21 June, 2007, 11pp.
Tudela, S. (2000) Ecosystem effects of fishing in the
Mediterranean: an analysis of the major threats of fishing
gear and practices to biodiversity and marine habitats.
FAO Fisheries Department (EP/INT/759/GEF), Rome, Italy,
45pp.
SPRFMO (2007d) Information describing alfonsino (Beryx
splendens) fisheries relating to the South Pacific Regional
Fisheries Management Organisation, Working Draft, 20
June, 2007, 14pp.
SPRFMO (2007e) Information describing bluenose
(Hyperoglyphe antarctica) fisheries relating to the South
Pacific Regional Fisheries Management Organisation,
Working Draft, 22 June, 2007, 14pp.
SPRFMO (2007f) Information describing Jasus caveorum
fisheries relating to the South Pacific Regional Fisheries
Management Organisation, Revised, 20 February, 2007.
SPRFMO-III-SWG-12, 8pp.
Tupanogov, V.N., Orlov, A.M., Kodolov, L.S. (2008) The
most abundant grenadiers of the Russian Far East EEZ:
distribution and basic biological patterns. In: Orlov, A.M.,
Iwamoto, T. (eds) Grenadiers of the World Oceans: Biology,
Stock assessment, and Fisheries. American Fisheries
Society, Symposium 63, Bethesda, Maryland, USA, pp.
285-316.
Uchupi, E., Ballard, R.D. (1989) Evidence of hydrothermal
activity on Marsili Seamount, Tyrrhenian Basin. Deep-Sea
Research 36: 1443–1448.
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 91
UNESCO (2009) Global Open Oceans and Deep Seabed
(GOODS) Bioregional Classification. Vierros, M., Cesswell,
I., Escobar-Briones, E., Rice, J., Ardron, J. (eds) IOC
Technical Series 84, UNESCO-IOC, Paris, France, 89pp.
UNGA (2004) Resolution 59/25 Sustainable fisheries,
including through the 1995 Agreement for the
Implementation of the Provisions of the United Nations
Convention on the Law of the Sea of 10 December 1982
relating to the Conservation and Management of Straddling
Fish Stocks and Highly Migratory Fish Stocks, and related
instruments. UNGA A/RES/59/25. Available at: http://
www.un.org/Depts/los/general_assembly/general_
assembly_reports.htm, 16pp.
UNGA (2007) Resolution 61/105 Sustainable fisheries,
including through the 1995 Agreement for the
Implementation of the Provisions of the United Nations
Convention on the Law of the Sea of 10 December 1982
relating to the Conservation and Management of Straddling
Fish Stocks and Highly Migratory Fish Stocks, and related
instruments. UNGA A/RES/61/105 Available at: http://
www.un.org/Depts/los/general_assembly/general_
assembly_reports.htm, 21pp.
UNGA (2009) Resolution 64/72 Sustainable fisheries,
including through the 1995 Agreement for the
Implementation of the Provisions of the United Nations
Convention on the Law of the Sea of 10 December 1982
relating to the Conservation and Management of Straddling
Fish Stocks and Highly Migratory Fish Stocks, and related
instruments. UNGA A/RES/64/72 not yet issued (as of 10
January 2010); available as General Assembly document
A/64/L.29 at: http://www.un.org/Docs/journal/asp/
ws.asp?m=A/64/L.29http://www.un.org/Depts/los/
general_assembly/general_assembly_reports.htm, 27pp.
UN SG (2006) Report of the Secretary-General Impacts of
fishing on vulnerable marine ecosystems: actions taken by
States and regional fisheries management organizations
and arrangements to give effect to paragraphs 66 to 69
of General Assembly resolution 59/25 on sustainable
fisheries, regarding the impacts of fishing on vulnerable
marine ecosystems. A/61/154. 14 July 2006, 46pp.
Vanreusel. A., Anderson, A.C., Boetius, A., Connelly, D.,
Cunha, M.R., Decker, C., Hilario, A., Kormas, K.A., Maignien,
L., Olu, K., Pachiadaki, M., Ritt, B., Rodrigues, C., Sarrazin,
J., Tyler, P., Van Gaever, S., Vanneste, H. (2009) Biodiversity
of cold-seep ecosystems along the European margins.
Oceanography 22: 110–127.
Verrill, A.E. (1922) The Alcyonaria of the Canadian Arctic
Expedition, 1913–1918, with a revision of some other
Canadian genera and species. Report of the Canadian
Arctic Expedition 1913–18, vol VIII: Molluscs, Echinoderms,
Coelenterates, etc., Part G: Alcyonaria and Actinaria.
Waller, R., Watling, L., Auster, P., et al. (2007) Anthropogenic
impacts on the Corner Rise Seamounts, north-west Atlantic
Ocean. Journal of the Marine Biological Association of the
UK 87: 1075–1076.
Warén, A., Klitgaard, A.B. (1991) Hanleya nagelfar, a
sponge-feeding ecotype of H hanleyi or a distinct species
of chiton? Ophelia 34: 51-70.
WGDEC (2005) Report of the Working Group on DeepWater Ecology (WGDEC), 8–11 March 2005 ICES
Headquarters, Copenhagen, ICES CM 2005/ACE:02 Ref.
E,G. International Council for the Exploration of the Sea,
Copenhagen, Denmark, 72pp.
Annexes
WGDEC (2006) Report of the Working Group on DeepWater Ecology (WGDEC), 4–7 December 2005, Miami,
U.S.A, ICES CM 2006/ACE:04. International Council for the
Exploration of the Sea, Copenhagen, Denmark, 79pp.
ANNEX 1: UNGA RESOLUTION 61/105,
PARAGRAPHS 83–87
WGDEC (2007) Report of the Working Group on Deep-Water
Ecology (WGDEC), 26–28 February 2007, ICES CM 2007/
ACE:01 Ref LRC. International Council for the Exploration
of the Sea, Copenhagen, Denmark, 57pp.
WGDEC (2008) Report of the ICES-NAFO Joint Working
Group on Deep-Water Ecology (WGDEC) 10–14 March
2008, ICES CM2008/ACOM: 45, Ref. WGECO, WGDEC,
NAFO. International Council for the Exploration of the Sea,
Copenhagen, Denmark, 124pp.
WGDEC (2009) Report of the ICES-NAFO Working Group on
Deep-water Ecology (WGDEC), 9–13 March 2009. ICES CM
2009/ACOM: 23. International Council for the Exploration
of the Sea, Copenhagen, Denmark, 92pp.
WGDEEP (2009) Report of the Working Group on the
Biology and Assessment of Deep-Sea Fisheries Resources
(WGDEEP) 9–16 March 2009, Copenhagen, Denmark. ICES
WGDEEP Report 2009, ICES Advisory Committee, ICES CM
2009/ACOM: 14, 503pp.
WGEAFM (2008a) Report of the NAFO SC Working Group
on Ecosystem Approach to Fisheries Management
(WGEAFM) NAFO Scientific Council Meeting June 2008.
Serial Number 5511, NAFO SCS Doc. 08/10, 70pp.
WGEAFM (2008b) Report of the NAFO SC Working Group
on Ecosystem Approach to Fisheries Management
(WGEAFM) Response to Fisheries Commission Request
9.a. NAFO Scientific Council Meeting October 2008, Serial
Number 5592, NAFO SCS Doc. 08/24, 19pp.
WGEAFM (2009) Report of the NAFO Scientific Council
Working Group on Ecosystem Approach to Fisheries
Management (WGEAFM). Northwest Atlantic Fisheries
Organisation Scientific Council Meeting – June 2008 Ser.
No. N5511 NAFO SCS Doc 08/10, 70pp.
Wheeler, A.J., Bett, B.J., Billett, D.S., et al. (2005) The
impact of demersal trawling on northeast Atlantic coral
habitats: The case of the Darwin Mounds, United Kingdom.
In: Barnes, P.W., Thomas, J.P. (eds) Benthic Habitats and the
Effects of Fishing, American Fisheries Society Symposium
41, Bethesda, Maryland, USA, pp. 807–817.
Whitehead, P.J.P., Bauchot, M-L., Hureau, J-C., Nielsen, J.,
Tortonese, E. (1986) Fishes of the Northeastern Atlantic
and Mediterranean, Vols. 1,2,3. UNESCO, Paris, France.
Wisshak, M., López-Correa, M., Gofas, S., Salas, C., Taviani,
M., Jakobsen, J., Freiwald, A. (2009) Shell architecture,
element composition, and stable isotope signature of the
giant deep-sea oyster Neopycnodonte zibrowii sp. n. from
the NE Atlantic. Deep-Sea Research I 56: 374–407.
92 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The General Assembly
83. Calls upon regional fisheries management
organizations or arrangements with the
competence to regulate bottom fisheries to
adopt and implement measures, in accordance
with the precautionary approach, ecosystem
approaches and international law, for their
respective regulatory areas as a matter of
priority, but not later than 31 December 2008:
(a) To assess, on the basis of the best available
scientific information, whether individual bottom
fishing activities would have significant adverse
impacts on vulnerable marine ecosystems,
and to ensure that if it is assessed that these
activities would have significant adverse
impacts, they are managed to prevent such
impacts, or not authorized to proceed;
(b) To identify vulnerable marine ecosystems
and determine whether bottom fishing activities
would cause significant adverse impacts to such
ecosystems and the long-term sustainability of
deep sea fish stocks, inter alia, by improving
scientific research and data collection and
sharing, and through new and exploratory
fisheries;
(c) In respect of areas where vulnerable marine
ecosystems, including seamounts, hydrothermal
vents and cold water corals, are known to occur
or are likely to occur based on the best available
scientific information, to close such areas to
bottom fishing and ensure that such activities
do not proceed unless conservation and
management measures have been established
to prevent significant adverse impacts on
vulnerable marine ecosystems;
ecosystems are encountered, and to report the
encounter so that appropriate measures can be
adopted in respect of the relevant site;
84. Also calls upon regional fisheries
management organizations or arrangements
with the competence to regulate bottom
fisheries to make the measures adopted
pursuant to paragraph 83 of the present
resolution publicly available;
85. Calls upon those States participating in
negotiations to establish a regional fisheries
management organization or arrangement
competent to regulate bottom fisheries to
expedite such negotiations and, by no later than
31 December 2007, to adopt and implement
interim measures consistent with paragraph
83 of the present resolution and make these
measures publicly available;
86. Calls upon flag States to either adopt
and implement measures in accordance with
paragraph 83 of the present resolution, mutatis
mutandis, or cease to authorize fishing vessels
flying their flag to conduct bottom fisheries in
areas beyond national jurisdiction where there is
no regional fisheries management organization
or arrangement with the competence to
regulate such fisheries or interim measures
in accordance with paragraph 85 of the
present resolution, until measures are taken
in accordance with paragraph 83 or 85 of the
present resolution;
87. Further calls upon States to make publicly
available through the Food and Agriculture
Organization of the United Nations a list of
those vessels flying their flag authorized to
conduct bottom fisheries in areas beyond
national jurisdiction, and the measures they
have adopted pursuant to paragraph 86 of the
present resolution;
(d) To require members of the regional fisheries
management organizations or arrangements
to require vessels flying their flag to cease
bottom fishing activities in areas where, in the
course of fishing operations, vulnerable marine
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 93
ANNEX II: UN FAO GUIDELINES FOR THE
MANAGEMENT OF DEEP-SEA FISHERIES
IN THE HIGH SEAS
Paragraph 47: Impact Assessments
47. Flag States and RFMO/As should conduct
assessments to establish if deep-sea fishing
activities are likely to produce significant
adverse impacts in a given area. Such an impact
assessment should address, inter alia:
i. type(s) of fishing conducted or contemplated,
including vessels and gear-types, fishing
areas, target and potential bycatch species,
fishing effort levels and duration of fishing
(harvesting plan);
ii. best available scientific and technical
information on the current state of fishery
resources and baseline information on the
ecosystems, habitats and communities in the
fishing area, against which future changes are
to be compared;
iii. identification, description and mapping of
VMEs known or likely to occur in the fishing
area;
iv. data and methods used to identify, describe
and assess the impacts of the activity, the
identification of gaps in knowledge, and an
evaluation of uncertainties in the information
presented in the assessment;
v. identification, description and evaluation of
the occurrence, scale and duration of likely
impacts, including cumulative impacts of
activities covered by the assessment on VMEs
and low-productivity fishery resources in the
fishing area;
vi. risk assessment of likely impacts by the
fishing operations to determine which impacts
are likely to be significant adverse impacts,
particularly impacts on VMEs and low
productivity fishery resources; and
vii. the proposed mitigation and management
measures to be used to prevent significant
adverse impacts on VMEs and ensure longterm conservation and sustainable utilization
of low-productivity fishery resources, and the
measures to be used to monitor effects of the
fishing operations.
Paragraph 42: VMEs
42. A marine ecosystem should be classified as
vulnerable based on the characteristics that it
possesses. The following list of characteristics
should be used as criteria in the identification of
VMEs.
i. Uniqueness or rarity – an area or ecosystem
that is unique or that contains rare species
whose loss could not be compensated for by
similar areas. These include:
● habitats that contain endemic species;
● habitats of rare, threatened or endangered
species that occur only in discrete areas; or
● nurseries or discrete feeding, breeding, or
spawning areas.
ii. Functional significance of the habitat –
discrete areas or habitats that are necessary
for the survival, function, spawning/reproduction
or recovery of fish stocks, particular life-history
stages (e.g. nursery grounds or rearing areas),
or of rare, threatened or endangered marine
species.
iii. Fragility – an ecosystem that is highly
susceptible to degradation by anthropogenic
activities.
iv. Life-history traits of component species that
make recovery difficult – ecosystems that are
characterized by populations or assemblages
of species with one or more of the following
characteristics:
● slow growth rates;
● late age of maturity;
● low or unpredictable recruitment; or
● long-lived.
Impacts should be evaluated individually, in
combination and cumulatively.
18. When determining the scale and significance
of an impact, the following six factors should be
considered:
i. the intensity or severity of the impact at the
specific site being affected;
ii. the spatial extent of the impact relative to the
availability of the habitat type affected;
iii. the sensitivity/vulnerability of the ecosystem
to the impact;
iv. the ability of an ecosystem to recover from
harm, and the rate of such recovery;
v. the extent to which ecosystem functions may
be altered by the impact; and
vi. the timing and duration of the impact relative
to the period in which a species needs the
habitat during one or more life-history stages.
19. Temporary impacts are those that are
limited in duration and that allow the particular
ecosystem to recover over an acceptable time
frame. Such time frames should be decided on
a case-by-case basis and should be in the order
of 5–20 years, taking into account the specific
features of the populations and ecosystems.
v. Structural complexity – an ecosystem that is
characterized by complex physical structures
created by significant concentrations of biotic
and abiotic features. In these ecosystems,
ecological processes are usually highly
dependent on these structured systems. Further,
such ecosystems often have high diversity, which
is dependent on the structuring organisms.
20. In determining whether an impact is
temporary, both the duration and the frequency
at which an impact is repeated should be
considered. If the interval between the expected
disturbance of a habitat is shorter than the
recovery time, the impact should be considered
more than temporary. In circumstances of
limited information, States and RFMO/As
should apply the precautionary approach in
their determinations regarding the nature and
duration of impacts.
Examples of potentially vulnerable species
groups, communities, and habitats, as well
as features that potentially support them are
contained in Annex 1.
ANNEX III: UNGA RESOLUTION 64/72,
DECEMBER 2009 – BOTTOM FISHERIES
ON THE HIGH SEAS; KEY PARAGRAPHS
Paragraphs 17–20: Significant Adverse
Impacts
113. Calls upon States to take action
immediately, individually and through regional
fisheries management organizations and
arrangements, and consistent with the
precautionary approach and ecosystem
approaches, to implement the 2008
International Guidelines for the Management
of Deep-sea Fisheries in the High Seas of
the Food and Agriculture Organization of the
United Nations (“the Guidelines”) in order to
sustainably manage fish stocks and protect
17. Significant adverse impacts are those that
compromise ecosystem integrity (i.e. ecosystem
structure or function) in a manner that: (i)
impairs the ability of affected populations to
replace themselves; (ii) degrades the long-term
natural productivity of habitats; or (iii) causes,
on more than a temporary basis, significant loss
of species richness, habitat or community types.
94 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
vulnerable marine ecosystems, including
seamounts, hydrothermal vents and cold water
corals, from destructive fishing practices,
recognizing the immense importance and value
of deep sea ecosystems and the biodiversity
they contain;
114. Reaffirms the importance of paragraphs 80
to 91 of its resolution 61/105 addressing the
impacts of bottom fishing on vulnerable marine
ecosystems and the long-term sustainability of
deep sea fish stocks and the actions called for
in that resolution, and emphasizes the need for
full implementation by all States and relevant
regional fisheries management organizations or
arrangements of their commitments under those
paragraphs on an urgent basis;
119. Considers that, on the basis of the review
carried out in accordance with paragraph 91
of its resolution 61/105, further actions in
accordance with the precautionary approach,
ecosystem approaches and international law,
are needed to strengthen the implementation
of paragraphs 80 and 83 to 87 of its resolution
61/105 and, in this regard, calls on regional
fisheries management organizations or
arrangements with the competence to regulate
bottom fisheries, States participating in
negotiations to establish such organizations
or arrangements, and flag States to take the
following urgent actions in areas beyond national
jurisdiction:
(a) Conduct the assessments called for in
paragraph 83 (a) of its resolution 61/105,
consistent with the Guidelines, and to ensure
that vessels do not engage in bottom fishing
until such assessments have been carried out;
(b) Conduct further marine scientific research
and use the best scientific and technical
information available to identify where
vulnerable marine ecosystems are known
to occur or are likely to occur and adopt
conservation and management measures to
prevent significant adverse impacts on such
ecosystems consistent with the Guidelines,
or close such areas to bottom fishing until
conservation and management measures have
been established, as called for in paragraph
83 (c) of its resolution 61/105;
(c) Establish and implement appropriate
protocols for the implementation of paragraph
83 (d) of its resolution 61/105, including
definitions of what constitutes evidence
of an encounter with a vulnerable marine
ecosystem, in particular threshold levels and
The Implementation of UN ResolutionS 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 95
indicator species, based on the best available
scientific information and consistent with the
Guidelines, and taking into account any other
conservation and management measures
to prevent significant adverse impacts on
vulnerable marine ecosystems, including
those based on the results of assessments
carried out pursuant to paragraph 83 (a) of its
resolution 61/105 and paragraph 119 (a) of
the present resolution;
(d) Adopt conservation and management
measures, including monitoring, control
and surveillance measures, on the basis of
stock assessments and the best available
scientific information, to ensure the long-term
sustainability of deep sea fish stocks and nontarget species, and the rebuilding of depleted
stocks, consistent with the Guidelines; and,
where scientific information is uncertain,
unreliable, or inadequate, ensure that
conservation and management measures be
established consistent with the precautionary
approach, including measures to ensure that
fishing effort, fishing capacity and catch limits,
as appropriate, are at levels commensurate
with the long-term sustainability of such
stocks;
120. Calls upon flag States, members of
regional fisheries management organizations or
arrangements with the competence to regulate
bottom fisheries and States participating in
negotiations to establish such organizations
or arrangements to adopt and implement
measures in accordance with paragraphs 83,
85 and 86 of its resolution 61/105, paragraph
119 of the present resolution, and international
law, and consistent with the Guidelines, and not
to authorize bottom fishing activities until such
measures have been adopted and implemented;
121. Recognizes the special circumstances
and requirements of developing States and
the specific challenges they may face in giving
full effect to certain technical aspects of the
Guidelines, and that implementation by such
States of paragraphs 83 to 87 of its resolution
61/105, paragraph 119 of the present
resolution and the Guidelines should proceed
in a manner that gives full consideration to
the section of the Guidelines on “Special
Requirements of Developing Countries”;
122. Calls upon States and regional fisheries
management organizations or arrangements
to enhance efforts to cooperate to collect and
exchange scientific and technical data and
information related to the implementation of the
measures called for in the relevant paragraphs
of its resolution 61/105 and the present
resolution to manage deep sea fisheries in
areas beyond national jurisdiction and to protect
vulnerable marine ecosystems from significant
adverse impacts of bottom fishing by, inter alia:
(a) Exchanging best practices and developing,
where appropriate, regional standards for
use by States engaged in bottom fisheries
in areas beyond national jurisdiction and
regional fisheries management organizations
or arrangements with a view to examining
current scientific and technical protocols
and promoting consistent implementation of
best practices across fisheries and regions,
including assistance to developing States in
accomplishing these objectives;
(b) Making publicly available, consistent with
domestic law, assessments of whether
individual bottom fishing activities would have
significant adverse impacts on vulnerable
marine ecosystems and the measures
adopted in accordance with paragraphs 83,
85 and 86, as appropriate, of its resolution
61/105, and promoting the inclusion of
this information on the websites of regional
fisheries management organizations or
arrangements;
(c) Submission by flag States to the Food and
Agriculture Organization of the United Nations
of a list of those vessels flying their flag
authorized to conduct bottom fisheries in
areas beyond national jurisdiction, and the
measures they have adopted to give effect
to the relevant paragraphs of its resolution
61/105 and the present resolution;
(d) Sharing information on vessels that are
engaged in bottom fishing operations in areas
beyond national jurisdiction where the flag
State responsible for such vessels cannot be
determined;
129. Decides to conduct a further review
in 2011 of the actions taken by States and
regional fisheries management organizations
and arrangements in response to paragraphs
80 and 83 to 87 of its resolution 61/105 and
paragraphs 117 and 119 to 127 of the present
resolution, with a view to ensure effective
implementation of the measures and to make
further recommendations, where necessary; and
taking into account the discussions occurring
during the workshop described in paragraph 128
of the present resolution.
Acknowledgements
The International Programme on the State of the
Ocean (http://www.stateoftheocean.org)
would like to thank the following for their
support: The Pew Environment Group; the Deep
Sea Conservation Coalition; The JM Kaplan
Fund; The Oakdale Trust; and The John S Cohen
Foundation. Dr Alex Rogers would like to thank
Dr Christopher Yesson, Institute of Zoology,
Zoological Society of London and Roxane Brown,
Communications INC, London, U.K. for their
assistance in the preparation of the report.
123. Encourages States and regional fisheries
management organizations or arrangements
to develop or strengthen data collection
standards, procedures and protocols and
research programmes for identification of
vulnerable marine ecosystems, assessment of
impacts on such ecosystems, and assessment
of fishing activities on target and non-target
species, consistent with the Guidelines and in
accordance with the Convention, including Part
XIII;
96 The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas
The Implementation of UN Resolutions 61/105 and 64/72 in the Management of Deep-Sea Fisheries on the High Seas 97