ARQUIPÉLAGO
Life and Marine Sciences
SCOPE
ARQUIPÉLAGO - Life and Marine Sciences, publishes annually original scientific
articles, short communications and reviews on the natural and marine environment of the
archipelago of the Azores and surrounding region of the Atlantic Ocean, Macaronesian
and other oceanic Atlantic islands and seamounts.
PUBLISHER
Universidade dos Açores,
Rua da Mãe de Deus, 13A
PT-9501-801 Ponta Delgada, Azores, Portugal.
EDITOR
Helen Rost Martins
Phone: + 351 292 207 800 / 828 - Fax: +351 292 207 811
E-mail: [email protected]
INTERNET RESOURCES
http://www.arquipelago.info
Journal information, instructions to authors and free access to all papers since 1999.
FINANCIAL SUPPORT
Fundo de Apoio à Comunidade Científica (FACC)
EDITORIAL SECRETARIAT
Helen R. Martins, Filipe M. Porteiro, José Nuno Pereira & Emmanuel Arand.
EDITORIAL COMMITTEE
Paulo V. Borges, Angra do Heroísmo; José Azevedo, Ponta Delgada; Joël Bried, Horta.
ADVISORY BOARD
Miguel A. Alcaraz, Barcelona, Spain; Alan B. Bolten, Gainesville, Florida, USA; António B. de
Sousa, Lisboa, Portugal; Manuel Afonso-Dias, Faro, Portugal; Malcolm R. Clarke, Pico, Azores,
Portugal; Richard D.M. Nash, Bergen, Norway; Erik Sjögren, Uppsala, Sweden; Charles H.J.M.
Fransen, Leiden, Netherlands; John Robert Press, London, UK; George R. Sedberry, Georgia, USA.
Indexed in:
Aquatic Sciences and Fisheries Abstracts (ASFA), Biological Abstracts, BIOSIS Previews,
Zoological Record, ISI Web of Knowledge
Cover design: E. Arand. Photo: Hatchling of C. caretta on its way to the sea, Cape Verde Isl. - H.R. Martins
The cave‐adapted arthropod fauna from Madeira archipelago ARTUR R.M. SERRANO & PAULO A.V. BORGES Serrano, A.R.M. & P.A.V. Borges 2010. The cave-adapted arthropod fauna from
Madeira archipelago. Arquipelago. Life and Marine Sciences 27: 1-7.
This work provides an overview of the hypogean fauna from the Madeira archipelago,
presenting a list of obligated cave-dwelling species. A total of 6 troglobiont species in 5
orders have been described to date. The cave fauna in Madeira can be considered poor
when compared with either the local epigean fauna or the cave fauna of other Macaronesian
archipelagos. Curious is the occurrence of one wood-louse cave species (Trichoniscus
bassoti), which apparently is the only troglobite living in more than one Macaronesian
archipelago (Canaries and Madeira). Major problems related to the conservation of cave
fauna are discussed, but it is clear that the protection of this specialized fauna requires the
adequate management of surface habitats.
Key words: Cavalum, Coleoptera, lava tubes, Machico, troglobiont species
Artur R.M. Serrano, Centro de Biologia Ambiental/Departamento de Biologia Animal,
Faculdade de Ciências da Universidade de Lisboa, R. Ernesto de Vasconcelos, C2, PT1749-016 Lisboa, Portugal. Paulo A.V. Borges (e-mail: [email protected]) Azorean
Biodiversity Group, Departamento de Ciências Agrárias, Universidade dos Açores, TerraChã, PT-9700-851 Angra do Heroísmo, Terceira, Portugal.
INTRODUCTION
There are nearly 3891 taxa (species and
subspecies) of terrestrial arthropods belonging to
462 families and 2118 genera recorded for
Madeira and Selvagens, and a large proportion of
these (921 species; 77 subspecies) are endemic to
the archipelago (Borges et al. 2008). Most of the
endemic species and subspecies are known from
the native forest and new species and subspecies
continue to be found, with an average of 90 new
taxa described per decade (Borges et al. 2008).
Madeira Islands have a volcanic origin and
consequently have several habitats in which it is
possible to find cave-adapted fauna (e.g. lavatubes, volcanic pits, mesovoid shallow substratum
– MSS). Interestingly and in contrast to the
archipelagos of the Azores (see Borges & Oromí
1994; Borges et al. 2007) and the Canaries (see
Oromí 1992; 2008), the cave inhabiting fauna of
Madeira archipelago has been less studied and
explored (but see a recent compilation in Serrano
& Borges in press). With the exception of the
recent study of Nunes (2005) no other exhaustive
study of the cave biota of Madeira is available.
The so-called MSS (“Milieu souterrain superficiel” or “Mesovoid Shallow Substratum”, see
Borges & Oromí 1994) could be a potential
interesting habitat to explore cave-adapted fauna
of Madeira archipelago, but almost nothing is
known about the arthropod fauna of this habitat.
Moreover, no organized information on the
Madeira archipelago cave fauna is available in the
project Encyclopaedia Biospeleologica (see Juberthie & Decu 1994; 1998).
Cave ecosystems support large numbers of
arthropod species, which represent most of
biomass in this environment (Culver & Pipan
2009). In many of their ecological roles, cave
arthropods are unique and no other animal group
could substitute them. However, most people are
unaware of the importance of cave arthropods and
frequently these small invertebrates have been
overlooked in conservation management projects
1
both in caves and elsewhere. In this work we
expand the data available in Serrano & Borges (in
press) providing some comments on the
conservation status of the species and their
habitat.
MATERIAL AND METHODS
The Madeira archipelago is located in the Atlantic
Ocean, southwest of the Iberian Peninsula,
between the latitudes 32o30´ and 33o31´ northern
latitude and 16o30´and 17o30´ western longitude .
The distance between this archipelago and Ponta
de Sagres (Portugal - the closest point of
mainland Europe) is about 1,000 km. The largest
island is Madeira (730 km2) and Porto Santo (70
km2) is located 60 km to the northeast of Madeira.
To the southeast the Madeira archipelago
continues along the Desertas sub-archipelago,
composed of three small islands: Ilhéu Chão (ca.
0.5 km2, 100 m maximum altitude), Deserta
Grande (ca. 10 km2, 479 m maximum altitude)
and Bugio (ca. 3 km2, 388 m maximum altitude).
This archipelago is part of the African plate and
includes a single volcanic group (the MadeiraPorto Santo) which is interpreted as being a longlived “hot spot” coming from the mantle
(Carvalho & Brandão 1991). The emerged part of
the Madeira Island dates back to the PostMiocene, <5.6 Ma (Ribeiro et al. 2005).
THE CAVE ENVIRONMENT
The volcanic activity in Madeira archipelago
ended 6000-7000 years BP (Geldmacher et al.
2000). The known lava tube caves and pits are
poorly studied from a geological perspective, and
most probably are the result of the Mio-Pliocenic
volcanic activity. Madeira is the only island with
true lava tube caves known locally by different
names such as “grutas”, “cavernas” and/or
“furnas” (Gouveia 1963). No lava tube caves are
known for Porto Santo, but only some pits locally
known as “lapas” normally used by humans and
cattle as refuges (Gouveia 1963). Cave and pit
formations are unknown in the neighbouring
islets of Desertas. The cavities of Madeira
archipelago are poorly known and only two
works devoted to this subject are available
(Gouveia 1963; Relvas & Monteiro 1987). The
speleometry of a total of six cavities and four
“lapas” were given for Madeira and Porto Santo,
respectively. (Gouveia 1963) (see Table 1). A
topographic characterisation was done also for the
complex of Cavalum caves located in Madeira
Island (Relvas & Monteiro 1987).
Table 1. Volcanic cavities of Madeira and Porto Santo islands, with their location and length (m) (modified from
Serrano & Borges in press).
2
Madeira Island
Volcanic cavities
Location
Gruta do Cavalum or Cavalão 1 or I
Gruta do Cavalum or Cavalão 2 or II
Gruta do Cavalum or Cavalão 3 or III
Gruta do Cavalum or Cavalão 4 or IV
Caverna dos Landeiros
Furna do Sítio da Queimada de Baixo
Furna do Convento – Santa Cruz
Gruta do Cardal
Furna do Sr. Frederico
Southern slope of the streamlet of Machico (Machico)
Southern slope of the streamlet of Machico (Machico)
Southern slope of the streamlet of Machico (Machico)
Southern slope of the streamlet of Machico (Machico)
“Sítio dos Landeiros” (Machico)
Left slope of the streamlet of Queimada (Água de Pena)
Sea-shore of Santa Cruz
Right slope of the streamlet of S. Vicente (S. Vicente)
Near S. Vicente cemetery
Porto Santo
Volcanic cavities
Location
Lapa 1
Lapa 2
Lapa 3
Lapa 4
Pico da Ana Ferreira
Pico da Ana Ferreira
Pico da Ana Ferreira
Pico da Ana Ferreira
Length
83
85
40
25
85
41
35
81
30
Length
8
6
3
3.5
ARTHROPOD SAMPLING AND DATA
Field work in several caves of Madeira Island has
been carried out by one of the authors (ARMS)
between 1991 and 1993. The sampling of cave
arthropods was performed using direct search and
pitfall traps baited with TURQUIN (1000 ml of
dark beer, 5 ml acetic acid, 5 ml formalin and 10
g of hydrate chloral) and a piece of Danish-Blue
cheese suspended from the edge of the trap. A
few drops of liquid detergent were added to
reduce surface tension. Sets of eight traps (radius
70 mm and depth 100 mm) were dug into the soil
(with the rim at the surface level) or placed inside
cracks and fissures (for more details see also
Serrano 1993; and Serrano & Borges 1995).
In addition to the results obtained with field
work we revise the available literature. In this
revision we only consider the troglobite/troglobiont fauna (i.e. hypogean), species which clearly
show absence or at least clear reduction of eyes,
lack of pigmentation, slender and long body and
appendages, wings absent or rudimentary (for
insects) and exclusive presence in subterranean
habitats. We also refer to some cases of troglophilic species, i.e. cave organism that may
complete its life cycle in a cave, but can also
survive in above ground habitats. No mention is
made to trogloxene species that are organisms
that uses caves for shelter but does not complete
its life cycle in them (see Culver & Pipan 2009).
The use of the term “subterranean” includes both
the species that live in the caves and MSS
(hypogean) and those occurring in the soil
(endogean).
RESULTS
COMMENTED LIST OF SPECIES
PALPIGRADI
Eukoeneniidae. Eukoenenia madeirae Strinati &
Condé, 1995 is considered a troglobitic species.
Described originally from specimens collected at
Cavalum II (Strinati & Condé 1995), it was
recently captured only at Cavalum I (Nunes
2005). Interestingly, there is a congeneric epigean
species in Madeira, Eukoenenia mirabilis (Grassi
& Calandruccio, 1885), but Eukoenenia madeirae
differs from the epigean relative by the patterns of
setae distribution in urosternite VI.
PSEUDOSCORPIONES
Chthoniidae. In a recent intensive survey of the
arthropod fauna of Madeiran lava tubes Nunes
(2005) found a new pseudoscorpion species,
Paraliochthonius cavalensis Zaragoza, 2004.
Only one specimen of this troglobitic species was
found in Cavalum III lava tube (Zaragoza et al.
2004; Nunes 2005). Interestingly, there is a
congeneric epigean species on the island,
Paraliochthonius hoestlandti Vachon, 1960, but
the new cave species has no eyes and markedly
elongated legs (Zaragoza et al. 2004).
ARANEAE
Linyphiidae. The two representatives of Araneae
belong to the genus Centromerus (Linyphiidae)
(Wunderlich 1992; 1995) and may have a
common epigean ancestor species (e.g. C.
variegatus). Centromerus sexoculatus Wunderlich, 1992 was found in the lava tubes of the
Cavalum complex (Cavalum I) near Machico and
C. anoculus Wunderlich, 1995 in Cardais cave
near S. Vicente. These two species are cavedwellers with reduced eyes as well as pronounced
depigmentation. Cave species of the genus
Centromerus are not present in other
Macaronesian archipelagos (Wunderlich 1993).
ISOPODA
Trichoniscidae. Curious is the occurrence of one
wood-louse cave species (Trichoniscus bassoti
Vandel, 1960), which apparently is the unique
troglobite living in more than one Macaronesian
archipelago (Canaries and Madeira) (Oromí
1992). The presence in more than one archipelago
could be explained by: i) the secondary dispersal
to an additional island. For instance, in the Azores
many troglobionts occur near the entrances of
caves and it could happen that their eggs or even
juveniles have been transported by a bird or by
wind in debris; ii) poor taxonomic resolution, i.e.,
in each island we have different taxa but with
very similar morphologies.
3
Fig 1. a) Thalassophilus pieperi Erber (Coleoptera, Carabidae); b) Medon vicentensis Serrano
(Coleoptera, Staphylinidae) (Photos by Enésima Mendonça).
4
This species was found only in “furnas” of
Cavalum I, II and III (Machico) and Vandel
(1960) refers that it is close to T. jeanneli Vandel
and T. halophilus Vandel, 1951.
COLEOPTERA
Carabidae. The other three known cave-dwelling
species belong to the order Coleoptera. Thalassophilus pieperi Erber, 1990 (Carabidae) (Fig.
1a) was found in the complex of “furnas” of
Cavalum I, II, III (Machico), in Cardais Cave (S.
Vicente) and recently also in Landeiros Cave
(Nunes 2005). The species shows depigmentation
and reduced eyes. In spite of never having been
found in caves, there is another Madeiran species
of Thalassophilus (T. coecus), which also shows
adaptations to a hypo-endogean habitat. When
comparing five species of the genus, Erber (1990)
concluded that these two species are very close.
The presence of T. pieperi in cavities separated
by several kilometres could indicate a
subterranean connection between the caves (e.g.,
through MSS) or even a previously greater
distribution of the species than the actual one in
Madeira Island. It is strange that an island with a
high number of epigean Trechus species (Lompe
1999) has just one record of a facultative
troglophilic species living in the entrance of
Cardais cave (T. fulvus Dejean, 1831) (Serrano &
Borges 1995) and in Landeiros cave (Nunes
2005). This species, as the former authors pointed
out, probably arrived recently to the island or, if
not, by severe competition with other carabid
species, remained limited to an empty biotope
like the entrance of caves (in the mainland T.
fulvus is sometimes also found in this type of
biotopes).
Staphylinidae. The last cave-dweller found in
Madeira is the rove-beetle Medon vicentensis
Serrano, 1993 (Fig. 1b) (Serrano 1993). This
species is eyeless and depigmented, but with
more or less normal legs and antennae. It was
found only in Cardais Cave (S. Vicente). Rovebeetles as cave-dwellers are only known from the
Canary and Madeira archipelagos in the context
of the oceanic Macaronesia. Medon vicentensis
belongs to the Medon ferrugineus group (species
characterized by dilated protarsi in the males and
the morphology of the aedeagus), as defined by
Assing (2006), which includes a total of ten
species in the Atlantic Islands, the Western
Mediterranean, and Europe. Originally Serrano
(1993) erroneously attributed Medon vicentensis
to the M. fusculus group (sensu Assing 2006).
The epigean fauna of Madeira includes three
additional species that are not included in the
Medon ferrugineus group: Medon apicalis
(Kraatz, 1857) (belonging to the M. apicalis
group sensu Assing 2006); Medon ripicola
(Kraatz, 1854) (belonging to the M. fusculus
group sensu Assing 2006); and the endemic
Medon indigena (Wollaston, 1857) which is
considered as a separate monophyletic group
(Assing 2006). Two additional cave adapted
Medon species are known from the Canary
Islands (M. antricola from Hierro; M. feloi from
La Palma) (see Assing 2006).
DISCUSSION
While the endemic epigean arthropod fauna of the
Madeira archipelago is very rich (many hundreds
of species distributed by several orders) (see
Borges et al. 2008), the known cave adapted
fauna is very poor. On the other hand, the
knowledge of the endogean fauna in the last three
decades, mainly belonging to the genus Geostiba
Thomson (Coleoptera Staphylinidae), has greatly
improved (e.g. Franz 1981; Assing & Wunderle
1996).
The current apparent poverty of the hypogean
fauna is due probably to the small number of
available cavities, to the lack of sampling, or to
both causes. In accordance with these
assumptions, there are indications, based on
recent sampling effort in Cavalum tubes, that the
arthropod fauna of Madeira caves is much richer
than previously expected (Nunes 2005). As was
shown also by Oromí (1992), the fauna of the
MSS is apparently unknown for the Madeira
archipelago. We say “apparently” because some
species collected in litter or even in deeper layers
of soil can also potentially be found in the MSS
(e.g. Talassophilus coecus Jeannel, Trechus
myniops Wollaston, Geostiba spp.). These species
show distinct adaptations to the hypo-endogean
5
life: some depigmentation and reduction of eyes.
However, until now the known hypogean species
of the archipelago were all found only in the
cavities of Madeira Island.
Most caves have unique communities of
invertebrates, but very few are obligate subterranean species as showed by our samples and
Nunes (2005) survey. Due to their island endemic
status and occurrence in only a few caves, most
troglobites may represent an important fraction of
the threatened species in Madeira. The most
effective way to protect those species is to create
special measures for the cave systems, since they
usually cover a small area. Limiting access to
caves with gates could be a solution. However,
surface habitats have also to be carefully
managed due to the danger of inputs of toxic
contaminants or extreme eutrophication. Despite
the widespread destruction of Madeira´s natural
vegetation on the lowlands and particularly in the
south part of the island, approximately 73% of
Madeira’s territory corresponds to protected areas
(Laurisilva Forest is a Natural World Heritage
site under UNESCO). Nevertheless, there is no
indication that cave fauna is safeguarded by the
above-cited legislation. The inconspicuousness of
many cave arthropod species and the lack of
standardized population studies make it difficult
to establish their threatened status. However, it is
a priority to produce lists of threatened Madeira
arthropods and to increase efforts to monitor their
populations.
In contrast to recent efforts in the Azores, we
do not have knowledge of significant efforts in
the protection of cave systems in Madeira Island.
Unfortunately one of the most interesting caves of
Madeira (Gruta dos Cardais, S. Vicente) was
partially destroyed and modified as show-cave in
1992-1993 with the support of the Regional
Government and without any study of environmental impact. At least owing to its biological
interest the cave of Cardais (the remaining
undisturbed tubes) and “Furnas” of Cavalum must
be protected in order to avoid the potential
extinction of troglobitic species.
6
ACKNOWLEDGEMENTS
We would like to thank Dr. Manuel Biscoito
(Museu Municipal do Funchal) for facilities and
logistic support in Madeira. We also thank Mr. J.
Silva of the same Museum for assistance in field
work. We thank Pedro Oromí and Volker Assing
acting as referees for their critical comments and
improvements on the manuscript. Clara S. Gaspar
and Enésima Mendonça produced figure 1.
REFERENCES
Assing, V. 2006. A revision of Western Palaearctic
Medon: the species of the Atlantic Islands, the
Western Mediterranean, and Europe, except for the
southeast (Insecta: Coleoptera: Staphylinidae:
Paederinae). Bonner Zoologische Beitrage 54: 2595.
Assing, V. & P. Wunderle 1996. A revision of the
Madeiran species of the genus Geostiba Thomson,
1858 (Coleoptera: Staphylinidae). Révue Suisse de
Zoologie 103: 119-150.
Borges, P.A.V. & P. Oromí 1994. The Azores. Pp. 605610 in: Juberthie, C. & V. Decu (Eds),
Encyclopaedia Biospeologica. Tome I. Société de
Biospéologie, Moulis.
Borges, P.A.V., P. Oromí, A.R.M. Serrano, I.R
Amorim & F. Pereira 2007. Biodiversity patterns
of cavernicolous ground-beetles and their
conservation status in the Azores, with the
description of a new species: Trechus isabelae n.
sp. (Coleoptera: Carabidae: Trechinae). Zootaxa
1478: 21–31.
Borges, P.A.V., A.M.F. Aguiar, M. Boieiro, M. CarlesTolrá & A.R.M. Serrano 2008. The arthropods
(Arthropoda) of Madeira and Selvagens
archipelagos. Pp. 245-270 in: Borges, P.A.V., C.
Abreu, A.M.F. Aguiar, P. Carvalho, R. Jardim, I.
Melo, P. Oliveira, C. Sérgio, A.R.M. Serrano & P.
Vieira (Eds). A list of the terrestrial fungi, flora
and fauna of Madeira and Selvagens archipelagos.,
Direcção Regional do Ambiente da Madeira and
Universidade dos Açores, Funchal and Angra do
Heroísmo.
Carvalho, A.M.G & J.M. Brandão 1991. Geologia do
Arquipélago da Madeira. Publicações Avulsas,
Museu Nacional de História Natural, Universidade
de Lisboa, 170 pp. [In Portuguese]
Culver, D.C. & T. Pipan 2009. Biology of caves and
other subterranean habitats. Oxford University
Press, Oxford. 256 pp.
Erber, D. 1990. Thalassophilus pieperi n. sp., a new
cavernicolous carabid beetle from Madeira.
Bocagiana 140: 1-12.
Franz, H. 1981. Neue blinde, subterrane Coleopteren
von den Makaronesischen Inseln (Coleoptera:
Staphylinidae et Curculionidae). Entomologica
Scandinavica Supplement 15: 328-332. [In
German]
Geldmacher, J., P. Bogaard, K. Hoernle & H.
Schmincke 2000. The 40Ar/39Ar age dating of the
Madeira Archipelago and hotspot track (eastern
North Atlantic). Geochemistry, Geophysics,
Geosystems, 1:1008.
Gouveia, F.A.R. 1963. Principais grutas existentes na
ilha da Madeira. Boletim da Sociedade Portugeusa
de Espeleologia, 2ª série, 1: 34-40. [In Portuguese]
Juberthie C. & V. Decu (Eds) 1994. Encyclopaedia
Biospeologica. Tome I. Société de Biospéologie,
Moulis.
Juberthie C. & Decu V. (Eds) 1998. Encyclopaedia
Biospeologica. Tome II. Sociétè de Biospéleologie,
Moulis.
Lompe, A. 1999. Revision der Trechus–Arten Madeiras
und Porto Santos (Coleoptera, Carbidae).
Entomologische Blätter 95: 93-143. [In German]
Nunes, E. 2005. Os Tubos de Lava de Machico (Ilha da
Madeira,
Portugal):
Biodiversidade
e
Conservação. BSc Thesis, Dpto. de Biologia,
Universidade da Madeira. 50 pp. [In Portuguese]
Oromí, P. 1992. La Fauna Subterránea en las Islas
Macaronésicas. Pp. 193-205 in: Actas do 3º
Congresso Nacional de Espeleologia e do 1º
Encontro Internacional de Vulcanoespeleologia
das Ilhas Atlânticas (30 de Setembro a 4 de
Outubro de 1992), Angra do Heroísmo. [In
Spanish]
Oromí, P. 2008. Biospeleology in Macaronesia.
Association for Mexican Cave Studies Bulletin 19:
98-104.
Relvas, P. & V. Monteiro 1987. Grutas do Cavalum.
Machico – Madeira. Algarocho 9: 5-9. [In Portuguese]
Ribeiro, L., S. Echchakroun, J. Mata, A. Boven, N.
Mattielli, J. Hus & C. Maerschalk 2005. Elemental
and lead isotopic evidence for coeval
heterogeneities at Madeira/Desertas mantle source.
Pp. 485-488 in: Acts of VIII Congresso de
Geoquímica dos Países de Língua Portuguesa
(Portugal) (ed. Anonymous).
Serrano, A.R.M. 1993. Medon vicentensis n. sp., a new
species of eyeless rovebeetle (Coleoptera:
Staphylinidae: Paederinae) from a cave in the
island of Madeira. Bocagiana, 165: 1-7.
Serrano, A.R.M. & P.A.V. Borges 1995. A new
subspecies of Trechus fulvus Dejean, 1831
(Trechus fulvus madeirensis n. ssp.) from the
Madeira Island with some biogeographical
comments. Boletim do Museu Municipal do
Funchal, Supl. 4º: 663-670.
Serrano, A.R.M. & P.A.V. Borges (in press). The
Madeira archipelago. In: Juberthie C. & V. Decu
(Eds), Encyclopaedia Biospeologica. Tome Ia
Amérique_et_Europe. Soc._de_Biospéologie, Moulis.
Strinati, P. & B. Condé 1995. Grottes et Palpigrades de
Madère. Mémoires de Biospéologie 22:161-168.
Vandel, A. 1960. Les Isopodes terrestres de l’archipel
madérien. Mémoires du Muséum national d'histoire
naturelle, sér. A, Zoologie 22: 1-155.
Wunderlich, J. 1992. Die Spinnen-Fauna der
Makaronesischen Inseln. Taxonomie, Ökologie,
Biogeographie und Evolution. Beiträge zur
Araneologie 1: 1-619. [In German]
Wunderlich, J. 1993. The Macaronesian cave-dwelling
spider fauna (Arachnida: Araneae). Memoires of
the Queensland Museum 33: 681-686.
Wunderlich, J. 1995. Zur Ökologie, Biogeographie,
Evolution und Taxonomie einiger Spinnen der
Makaronesischen Inseln. Beiträge zur Araneologie
4: 385-439. [In German]
Zaragoza, J.A., D. Aguín-Pombo & E. Nunes 2004.
Paraliochthonius cavalensis, nueva especie
cavernícola de Madeira (Arachnida, Pseudoscorpiones, Chthoniidae). Revista Ibérica de
Aracnología 9:343-351. [In Spanish]
Accepted 29 March 2010.
7
Phytobenthic communities of intertidal rock pools in the eastern islands of Azores and their relation to position on shore and pool morphology FRANCISCO M. WALLENSTEIN, S.D. PERES, E.D. XAVIER & A.I. NETO Wallenstein, F.M., S.D. Peres, E.D. Xavier & A.I. Neto 2010. Phytobenthic
communities of intertidal rock pools in the eastern islands of Azores and their
relation to position on shore and pool morphology. Arquipelago. Life and Marine
Sciences 27: 9-20.
This study aimed to characterize algal composition inside rock-pools from two islands of
the Azores archipelago (São Miguel and Santa Maria) and relate it to shore height and pool
morphology. Pools were categorized as upper, medium and lower intertidal according to the
surrounding communities. Maximum depth and surface area were used to reflect
morphology and qualitative sampling to evaluate algal species richness. PRIMER software
assessed the similarity across islands, sites, shore heights and pool morphology. Eighty
eight algal taxa were identified in pools from São Miguel and 52 from Santa Maria.
Rhodophycean species dominated rock-pool flora on both islands. Differences were found
across islands and sites. Higher species richness was observed at medium intertidal pools.
Algae composition was not affected by shore height in pools from Santa Maria. São
Miguel’s medium and lower pools were grouped separately from upper ones. Pool
morphology did not influence significantly the algae composition.
Key words: algae diversity, depth, shore height, spatial variability, surface area
Francisco M. Wallensteina,b (e-mail: [email protected]), Heriot-Watt University, School
of Life Sciences, John Muir Building, Edinburgh EH14 4AS, United Kingdom; CIRN –
Centro de Investigação de Recursos Naturais, Universidade dos Açores, PT-9501-801
Ponta Delgada, Açores, Portugal; Sara D. Peresa, Emanuel D. Xaviera & Ana I. Netoa,b,
a
Universidade dos Açores, Apartado 1422, PT-9501-801 Ponta Delgada, Açores, Portugal;
b
CIIMAR - Centro Interdisciplinar de Investigação Marinha e Ambiental, Rua dos Bragas,
289, PT-4050-123 Porto, Portugal.
INTRODUCTION
Rock-pools are well defined bodies of water,
influenced by sea and atmospheric conditions,
that develop an obvious distinction in flora and
fauna within restricted areas and for which one
single classification system becomes difficult
(Ganning 1971). Pool structure is determined by a
complex set of biological and physical factors
that interact to develop a patchy habitat
(Benedetti-Cecchi & Cinelli 1996). The patchiness observed in intertidal pools is a result of
small scale disturbances that affect communities
and is responsible for the variability observed
among pools even on the same site (Dethier 1984;
Metaxas & Scheibling 1993; Therriault & Kolasa
1999). Additionally, succession in recruitment
and competition for space may contribute to
increased variability of biotic communities in
rock-pools (Astles 1993; Metaxas & Scheibling
1993; Araújo et al. 2006).
Physical harshness during low tide determines
the distribution of intertidal algae (Ganning 1971;
Underwood 1980; Huggett & Griffiths 1986;
Metaxas & Sheibling 1993) and most physicalchemical conditions within individual pools relate
9
directly to their location on the shore relative to
water level, weather conditions, tidal height and
timing, and biological composition (Huggett &
Griffiths 1986). The algal composition in rockpools exhibits a marked gradient in many places,
with green algae dominating pools that occur
higher on the shore, whereas brown and red algae
are dominant at lower shore levels, where
common species from the adjacent subtidal
communities occur (Metaxas & Scheibling 1993).
A direct consequence of this is a decline in
diversity with shore height (Femino & Mathieson
1980; Huggett & Griffiths 1986; Wolfe & Harlin
1988; Kooistra et al. 1989, Therriault & Kolasa
1999), although variations in species variability
across shore levels may not be significant
(Metaxas et al. 1994).
Martins et al. (2007) showed that depth is more
important than area in explaining species
diversity and community composition in both
early successional and mature pools. Shallow
pools and their biota experience more extreme
variations in its physical and chemical conditions
(Ganning 1971; Metaxas & Scheibling 1993); on
the contrary, deeper pools are more stable and
develop a thermal stratification, thus allowing the
existence of more ecological niches (Martins et
al. 2007). The lack of influence of area on the
abundance of organisms in pools is in agreement
with the work of Underwood & Skilleter (1996),
who found little evidence that diameter (a
surrogate for area) of rock-pools leads to
significant differences in the abundance of most
taxa.
The sole study on rock-pool macroalgae
communities in the Azores was done by Neto &
Baldwin (1990) on Flores Island. The present
work aims to provide additional information on
the algal flora of littoral rock-pools from the
Azores, while analysing its spatial variability and
relating it with shore height and pool morphology.
MATERIAL AND METHODS
THE AZORES ARCHIPELAGO
The Azores are centrally located in the North
Atlantic (37º40' N, 25º 31' W, Fig. 1). The islands
lack a continental shelf, thus presenting a res10
Fig. 1. Location of the Azores Archipelago and of the
surveyed sites at the Island of São Miguel (1-3) and
Santa Maria (4-9): 1) Mosteiros; 2) Fenais da Luz; 3)
Maia; 4) Emissores; 5) Anjos; 6) São Lourenço; 7)
Maia; 8) Ribeira Seca; 9) Ilhéu da Vila.
tricted coastal extension that reaches a depth of
1000 m only 200 m offshore (Morton et al. 1998),
and are exposed to medium/high levels of wave
action (Macedo 2002). Shore geomorphology
alternates between high cliffs and rocky
cobble/boulder beaches (Borges 2004). Tidal
range is small (< 2 m, see Instituto Hidrográfico
1981), and therefore the extensive bedrock
platforms that favour the occurrence of rockpools are scarce and heterogeneous. Santa Maria
has steeper and narrower shores than São Miguel
and consequently a lower number of pools.
STRUCTURE OF INTERTIDAL COMMUNITIES
Intertidal communities in the Azores are
organized into three major zones: (i) an
uppermost zone (spray and splash) with littorinids
[Littorina striata King, Melharaphe neritoides
(L.)]; followed by (ii) a barnacle zone
[Chthamalus stellatus (Poli)]; and (iii) an algae
dominated zone (Neto 1992, 2000; Wallenstein &
Neto 2006). The algal dominated zone can be
further subdivided into three main bands: (a) an
upper band of Ulva spp. that overlaps with the
lower limit of the C. stellatus zone (Neto 1992) at
about 1.5m (±0,6m) above low water level,
followed by (b) an algae turf dominated zone at
about 1m (±0,5m) above low water level, with
occasional occurrence of frondose algae, namely
Fucus spiralis Linné and Gelidium microdon
Kützing (Neto 1992; Wallenstein & Neto 2006),
and (c) a frondose algae dominated zone located
in the lower limit of the intertidal zone at about
0,7m (±0,3m) above low water level, establishing
the transition to the subtidal (Neto 1992, 2000;
Wallenstein & Neto 2006), co-dominated by a
variable set of species.
FIELD WORK
Surveys took place at six sites on Santa Maria
during June and July 2005 and at three sites on
São Miguel during August through September of
the same year (Fig. 1), chosen directionally.
Surveyed sites present a variable number of
intertidal rock-pools of different shape and depth
occurring at different shore heights. Rock-pools
are not a common habitat in the Azores thus
imposing uneven sampling designs - on Santa
Maria four pools were surveyed at each site,
while on São Miguel four pools were surveyed at
Maia, 14 at Fenais da Luz and 44 at Mosteiros.
Based on the adjacent exposed bedrock algae
community distribution, pools were categorized
as: upper shore pools (U) when located where
green algae dominate; mid shore pools (M) when
located where turfs dominate; and low shore
pools (L) when located where frondose algae
dominate. Measures of depth and maximum and
minimum diameter of all pools were recorded.
Surface area of each rock-pool was calculated
based on its maximum and minimum diameter
adjusting it to the nearest circular shape. Pools
were numbered and a sample of all species
present inside each was collected. Identification
to species level was made in situ whenever
possible, otherwise taken to the laboratory for
diagnosis. Species were grouped into the classes
Rhodophyceae, Phaeophyceae and Chlorophyceae for data treatment and analysis. Species
nomenclature follows Guiry [cited 2007].
DATA ANALYSIS
Species-accumulation plots were built for (i) the
total number of species, for (ii) the number of
species occurring inside rock-pools at just one
island (hereafter referred to as exclusive species)
and for (iii) the number of species inside rockpools at each shore level, to assess whether a
reliable number of pools were sampled.
Presence/absence data was analysed using the
software PRIMER 6.1.5 (Clarke & Warwick
2001). Species richness was assessed by the total
number of species for each island and shore
height. ANOSIM (non parametric analysis of
similarity) and MDS (non metric multidimensional scaling) analysis were based on the
Bray Curtis similarity matrix. ANOSIM tested
differences between islands, locations, and shore
height, while MDS analysis was used to identify
rock-pool grouping patterns according to pool
depth and surface area, since proper replication
for these two factors was not possible. The
SIMPER routine was based on the presence/
absence data matrix and used to identify species
that contributed most for the differences between
the relevant factors identified with the ANOSIM
procedures.
RESULTS
On Santa Maria 10 pools were categorized as L,
seven as M and seven as U and on São Miguel
eight pools were categorized as L, 38 as M and 16
as U.
A total of 104 algae taxa were identified, of
which 52 recorded in Santa Maria, and 88 in São
Miguel (Table 1 in Appendix): 26 Chlorophyceae, 23 Phaeophyceae and 55 Rhodophyceae. A total of 16 taxa were exclusively
found inside rock-pools from Santa Maria and 52
exclusively inside rock-pools from São Miguel.
Rhodophyceae species were dominant inside
rock-pools from both islands, followed by
Phaeophyceae on Santa Maria and Chlorophyceae
on São Miguel (Fig. 2a). Additionally,
Rhodophyceae species dominate low and mid
shore rock-pools both in São Miguel and Santa
Maria, while a large number of upper shore rockpools were dominated by Phaeophyceae species
in Santa Maria and by Chlorophyceae species in
São Miguel (Fig. 2b, c).
Cumulative species plots for Santa Maria
(Fig. 3a) show that the total number of species
does not stabilize with increasing area, although
the number of exclusive species does tend to
stabilize around 10 when more than 10 pools have
been sampled. At São Miguel (Fig. 3b) both the
total and the exclusive number of species tend to
stabilize at 80 and 50 species respectively,
11
Fig. 3. Cumulative number of species (y-axis) relative
Fig. 2. Relative proportion (%) of major groups of
algae on rock-pools from Santa Maria (SMA) and São
Miguel (SMG): (a) all shore levels considered; (b) at
different shore heights in Santa Maria; (c) at different
shore heights in São Miguel. (L – lower shore; M –
mid-shore; U – upper shore; black bars – Rhodophyceae; dashed bars – Phaeophyceae; white bars Chlorophyceae).
when over 50 pools have been sampled.
Considering rock-pools from separate shore
height categories on both islands, total number of
species does not tend to stabilize (Figs. 3a, b).
Mid shore rock-pools presented species
richness (translated by the total number of
species) than upper shore rock-pools on both
islands (Table 2). The ANOSIM tests (Table 3)
showed that rock-pools from São Miguel and
Santa Maria differ significantly. Rock-pools on
the latter island showed higher significant
differences across survey sites but lower
differences across shore levels than those on São
Miguel. However, on São Miguel low shore (L)
12
to the number of rock-pools sampled (x-axis): (a) for
Santa Maria; (b) for São Miguel (black circles – total nº
of species; inverted white triangle – nº of exclusive
species; black squares - total nº of species at upper
shore levels; white diamonds - total nº of species at mid
shore levels; black triangles - total nº of species at low
shore levels).
and mid shore (M) rock-pools are not significantly different, but do differ significantly from
upper shore rock-pools (U), a pattern clearly
observed at Mosteiros, where most pools from
São Miguel were sampled and replication was
highest. Consistent with the ANOSIM results, the
average similarity results given by the SIMPER
routine (Table 4) evidences a higher similarity
between rock-pools within Santa Maria than
within São Miguel.
Species with a cumulative contribution of at
least 70% were selected as relevant for the
separation of rock-pools from both islands (Table
4). According to this criterion, Pterocladiella
capillacea is the only species in common between
both islands. Rock-pools on São Miguel are
Table 2. Total number of species (S) for different
shore heights (U – upper-shore; M – mid-shore;
L- low-shore) in Santa Maria and São Miguel.
Island
Shore level
S
L
M
U
L
M
U
12
13
10
13
14
7
Santa Maria
São Miguel
pairwise comparison between levels of each factor) for
each of the factors considered (999 random
permutations from a large number possible;
*significance <5%).
R-Value
Significance
0.605
(%)
0.1*
0.581
0.306
0.1*
0.1*
0.123
0.333
5.5
0.1*
0.048
0.453
29.2
0.1*
Factor “island”
Global
Factor “survey site”
Global Santa Maria
Global São Miguel
Factor “shore height”
Global Santa Maria
Global São Miguel
Pairwise Tests S. Miguel
M, L
M, U
L, U
Global Mosteiros
Pairwise Tests Mosteiros
L, M
M, U
L, U
of rock-pools within Santa Maria and São Miguel. their
respective average abundances (Ab) and percentage
contribution (%Con) for similarity from SIMPER
routine applied to the factor “island”.
Av. similarity:
Table 3. One-way ANOSIM results (global and
Tests
Table 4. Species that contribute most for the similarity
0.32
0.1*
0.439
0.1*
-0.008
0.761
0.691
47.6
0.1*
0.4*
characterized by a lower number of contributing
species (4 as opposed to 7 in rock-pools from
Santa Maria; see Table 4). Sargassum cymosum
and Cladophora prolifera are the most representative species in rock-pools from Santa
Maria, whereas Pterocladiella capillacea and
Ulva rigida are the most representative species
from this habitat on São Miguel. On this island
the SIMPER routine evidenced that M and L
pools are more similar when compared to U pools
(M>L>>U, Table 5, next page). Pterocladiella
capillacea is the species contributing to the
similarity between rock-pools across the whole
intertidal, although with varying average abun-
Pterocladiella capillacea
Cladophora prolifera
Sargassum cymosum
Stypocaulon scoparium
Padina pavonica
Cystoseira abies-marina
Chondria dasyphylla
Ulva rigida
Corallina elongata
Ulva intestinalis
Santa Maria
São Miguel
42.79
Ab %Con
0.63
7.1
0.92
16.9
0.83
14.56
0.75
11.3
0.63
7.58
0.58
6.41
0.5
5.08
29.51
Ab
%Con
0.75 21.35
0.75
0.58
0.42
19.46
12.77
10.55
dance values. In Mosteiros there is no species that
contributes to the similarity between rock-pools
on all shore height categories and there is a large
number of species shared by L and M pools
(frondose and turf forming), while U pools are
characterized by a lower number of species,
usually green and filamentous algae (Ulva spp.
and Aglaothamnion sp.). Additionally, mid-shore
rock-pools present a higher average similarity
than both upper-shore pools and lower-shore ones
(Table 6, next section).
Surface area and maximum depth do not seem
to influence algal community composition in
rock-pools from Santa Maria and São Miguel, as
shown by the nmMDS (Fig. 4, next section).
DISCUSSION
The higher number of taxa in rock-pools from
São Miguel is likely to reflect the higher number
of pools sampled there, mainly due to the
contribution of the survey site Mosteiros. The
qualitative inventory of algae inside intertidal
rock-pools on São Miguel was achieved when
about 50 pools were sampled. In the survey
conducted on Santa Maria, where only 24 pools
were considered, the total number of species
never stabilised, suggesting that the minimum
number of pools required for qualitative
13
Fig. 4. nmMDS plots for Santa Maria (stress 0,11): (a) bubble size
according to surface area (m2); (b) bubble size according to maximum depth
(m); and São Miguel (stress 0,22); (c) bubble size according to surface area
(m2); (d) bubble size according to maximum depth (m).
Table 5. Species that contribute most for the similarity of
rock-pools within shore levels on São Miguel, their
respective average abundances (Ab) and percentage
contribution (%Con) for similarity from SIMPER routine
applied to the factor “shore height” (L – lower intertidal
pools; M – medium intertidal pools; U – upper intertidal
pools).
Av. similarity:
L
M
U
31.81
Ab %Con
33.38
Ab %Con
24.02
Ab %Con
Species contributing for three shore height categories
P. capillacea
0.5
5.14 0.61
7.91
0.38
5.95
Species contributing for only two shore height categories
C. prolifera
0.75 12.9 0.68 10.25
S. cymosum
0.75 12.79 0.53
5.59
C. elongata
0.63 9.18 0.84 15.67
U. clathrata
0.5
5.14 0.39
3.21
C. pellucida
0.5
4.87 0.45
3.96
U. rigida
0.5
5.74
0.5
11.64
Species contributing for only one shore height category
S. scoparium
0.75 12.99
C. clavulatum
0.58
6.27
A. fragilissima
0.45
4.2
Ulva sp.
0.45
3.79
Herposiphonia sp.
0.45
3.61
Cystoseira sp.
0.39
3.28
U. compressa
0.75
Aglaothamnion sp.
0.44
39.08
8.44
14
assessments was not achieved at Santa
Maria due to the low occurrence of such
habitats and suggesting also that for the
azorean rock-pools this number is 50.
Nevertheless, the total number of
exclusive species stabilized at about 10
in Santa Maria and 50 in São Miguel.
This suggests lower variability of algae
communities inside rock-pools at Santa
Maria.
On the northwest coast of continental
Portugal Rhodophyceae are dominant at
all shore levels, except in pools located
at 2m and 3m up on the shore where
green algae dominate (Araújo et al.
2006). In other regions, however, the
pattern is one of monospecific green
algae communities in upper shore pools
and dominance of red and brown algae
in low shore rock-pools (Femino &
Mathieson 1980; Wolfe & Harlin 1988;
Kooistra et al. 1989). Although rockpools on São Miguel did not exhibit
monospecific communities, green algae
are dominant in pools located higher on
the shore, and red algae dominate the
mid and low intertidal pools.
Table 6. Species that contribute most for the similarity of
rock-pools within shore levels on Mosteiros, their respective
average abundances (Ab) and percentage contribution (Cont.)
for similarity from SIMPER routine applied to the factor
“shore height” (L – lower intertidal pools; M – medium
intertidal pools; U – upper intertidal pools).
Av. similarity:
L
32.82
Ab
%Cont
M
38.50
Ab
%Cont
U
29.98
Ab %Cont
Species contributing for only two shore height categories
C. elongata
0.8
14.63
0.83
12.11
C. prolifera
0.8
13.57
0.83
12.88
S. scoparium
0.8
13.29
0.55
4.36
C. pellucida
0.8
13.01
0.59
5.92
Herposiphonia sp. 0.6
7.29
0.59
5.4
Species contributing for only one shore height category
U. clathrata
0.6
6.78
C. clavulatum
0.66
6.71
A. fragilissima
0.59
6.29
S. cymosum
0.55
5.14
Ulva sp.
0.55
4.92
P. capillacea
0.52
4.34
U. compressa
1
The fact that surveyed bedrock shores in Santa
Maria exhibit a steeper slope and smaller
extension is the probable cause of having upper
shore pools mostly dominated by frondose
Phaeophyceae species. Surveyed pools were
consequently closer together, thus causing smaller
variation in physical-chemical conditions across
pools at different shore heights. This might
explain the absence of significant differences
across shore height levels at Santa Maria.
Additionally, the survey in this island occurred
entirely during the summer period when
desiccation and light temporarily eliminate the
upper shore green algae communities that are
more common in the winter period (Wallenstein
et al. 2008). The range of physical-chemical
conditions experienced inside pools is related to
their position on the shore and thus community
distribution patterns recognized on the exposed
intertidal zone influences communities inside
rock-pools (Huggett & Griffiths 1986).
Several studies report a diversity decrease
inside rock-pools with shore height (e.g. Metaxas
et al. 1994; Araújo et al. 2006). To clarify the
relationship between algal diversity and shore
68.98
height, cumulative richness curves must be
computed for each tidal level. This is not
always possible or easy (Metaxas &
Scheibling 1993) because of the greater
replication needed at each shore level. This
is the case on Azorean shores, where the
number of natural pools is limited. Mosteiros is the only survey site where rockpool replication for shore height was
possible. At this site, mid-shore rock-pools
exhibited the highest diversity and species
richness, with the upper shore dominated
by opportunistic Chlorophyceae species,
with fast growing life strategies (Larsson
et al. 1997; Björk et al. 2004). This is
analogous to the situation described by
Connell (1975): on a gradient of environmental stress diversity tends to be highest
at intermediate shore levels, as sensitive
species are less likely to survive under the
harsh conditions of upper shore levels and
out-compete the pioneer species on the
lower shore, but both co-exist on the midshore. Differences reported in this study
between rock-pools from the two islands seem to
reflect differences on bedrock communi-ties
evidenced by parallel biotope characterization
studies conducted in Santa Maria and São Miguel
(Wallenstein et al. 2008). This evidence is also
supported by studies of Dethier (1981; 1984) and
Astles (1993) that reveal a relation between
differences in algae composition of rock-pools
and that of adjacent bedrock, namely by
facilitating recruitment (Metaxas & Scheibling
1993; Underwood & Skilleter 1996; Martins et al.
2007). Significant differences across sites on both
islands are likely to be related to the variability of
algal communities across Azorean shores
associated to a highly variable morphology of
bedrock platforms that may play a determinant
role in influencing the tidal input and thus cause
differences in the community composition of
rock-pools (Metaxas et al. 1994). Variability of
local intrinsic factors (e.g. wave action, temperatures, predation and herbivory) might be the
main causes of variability in rock-pools (Dethier
1984; Astles 1993).
In the present study no significant relation was
found between pool morphology (surface area
15
and depth) and the algae communities present.
Martins et al. (2007) report a higher number of
species in deeper pools (disregarding surface
area) and link this observation to the higher
number of niches in those pools. Pool
morphology, however, affects the water volume
inside it and the correspondent exposure to light
and air, thus significantly affecting factors such as
the water temperature (Femino & Mathieson
1980). Considering surface area and depth
separately might neglect their joint effect on
species diversity, since volume is related to both
factors (Wolfe & Harlin 1988). A properly
replicated set of rock-pools for all ‘surface area x
depth categories’ combinations would be
required, but this is virtually impossible to
achieve in natural rock-pools on Azorean shores.
The great variability encountered inside Azorean
rock-pools is also reported elsewhere (e.g.
Underwood & Skilleter 1996; Araújo et al. 2006;
Martins et al. 2007). Pools at the same height and
closer
together
may
present
different
communities, and two pools at different height
categories may be very similar, given the
periodicity of tidal inputs (Metaxas & Scheibling
1993). Unique species assemblages inside each
rock-pool make it difficult to establish an
experimental design and a proper replication
scheme that would suit the need to generalize
about the factors that are known to affect
physical-chemical and biotic conditions inside
rock-pools.
CONCLUSIONS
Natural rock-pools are highly variable in the
Azores, as in most places in the world, due to a
complex interaction of physical-chemical and
biotic factors that are difficult to control and
virtually impossible to replicate. The present
study indicates that 50 is the minimum number of
pools required for qualitative assessments of
intertidal rock-pool algae community composition
in the Azores. In general terms, shore height
proved to be the main factor affecting rock-pool
biodiversity and community composition.
Replication was found to be difficult for
natural pools. Differences found between the two
16
islands are likely to be related to differences on
adjacent bedrock communities. Differences
across sites on both islands are likely to be related
to the variability of algal communities across
Azorean shores associated to a highly variable
morphology of bedrock platforms.
ACKNOWLEDGEMENTS
The authors wish to thank Dr. José Azevedo for
helping with valuable discussion and help with
experimental design and field work. We also
thank André Amaral, Catarina Santos, Eunice
Nogueira, Joana Pombo, João Ferreira, Rodrigo
Reis, Ruben Maciel and Tito Silva for helping
with the field work. A special acknowledgment to
Marlene Terra and Nuno Álvaro for helping in the
field, laboratory and image processing. This work
was developed under the project POCTIMGS/
54319/2002 – Biotope – Classification, Mapping
and Modelling of Azores Littoral Biotopes, and
supported also by CIRN-Centro de Investigação
de Recursos Naturais, Fundação para a Ciência e
Tecnologia (Portugal).
REFERENCES
Araújo R., I. Sousa-Pinto, I. Bárbara & V. Quintino
2006. Macroalgal communities of intertidal rock
pools in the northwest coast of Portugal. Acta
Oecologica 30: 192–202.
Astles, K.L. 1993. Patterns of abundance and
distribution of species in intertidal rock pools.
Journal of the Marine Biological Association of the
United Kingdom 73: 555-569.
Benedetti-Cecchi, L. & F. Cinelli 1996. Patterns of
disturbance and recovery in the littoral rock pools:
non-hierarchical competition and spatial variability
in secondary succession. Marine Ecology Progress
Series 135:145-161.
Björk, M., L. Axelsson & S. Beer 2004. Why is Ulva
intestinalis the only macroalga inhabiting isolated
rock pools along the Swedish Atlantic coast?
Marine Ecology Progress Series 284:109-116.
Borges, P.J.A. 2004. Ambientes litorais nos Grupos
Central e Oriental do Arquipélago dos Açores conteúdos e dinâmica de microescala. PhD Thesis,
Azores University. [In Portuguese].
Clarke, K.R. & R.M. Warwick 2001. Change in marine
communities: an approach to statistical analysis
and interpretation (2nd edition). PRIMER-E.
144 pp.
Connell, J.H. 1975. Some mechanisms producing
structure in natural communities: a model and
evidence from field experiments. Pp 460-490 in:
Cody M.L. & J.M. Diamond (Eds). Ecology and
evolution of communities. Belknap Press,
Cambridge. 545 pp.
Dethier, M.N. 1981. Heteromorphic algal life histories:
the seasonal pattern and response to herbivory of
the brown crust, Ralfsia californica. Oecologia 49:
333-339.
Dethier, M.N. 1984. Disturbance and recovery in
intertidal pools: maintenance of mosaic patterns.
Ecological Monographs 54(1): 99-118.
Femino, R. & A. Mathieson 1980. Investigations of
New England marine algae IV. The ecology and
seasonal succession of tide pool algae at Bald Head
Cliff, York, Maine, USA. Botanica Marina 23:
319-332.
Ganning, B. 1971. Studies on chemical, physical and
biological conditions in Swedish rock-pool
ecosystems. Ophelia 9: 51-105.
Guiry, M.D. & Guiry, G.M. (Internet). AlgaeBase:
world-wide electronic publication. National
University of Ireland, Galway; (cited 15 June
2007). Available from: http://www.algaebase.org/
Huggett, J. & C.L. Griffiths 1986. Some relationships
between elevation, physico-chemical variables and
biota of intertidal rock pools. Marine Ecology
Progress Series 29: 189-197.
Instituto Hidrográfico 1981. Roteiro do Arquipélago
dos Açores. PUB. (N)-IH-128-SN. Instituto
Hidrográfico, Lisboa.
Kooistra, W.H.C.F., A.M.T. Joosten & C. van den
Hoek 1989. Zonation patterns in intertidal pools
and their possible causes: a multivariate approach.
Botanica Marina 32: 9-26.
Larsson, C., L. Axelsson, H. Ryberg & S. Beer 1997.
Photosynthetic carbon utilization by Enteromorpha
intestinalis (Chlorophyta) from a Swedish rockpool. European Journal of Phycology 32: 49-54.
Macedo, F.L.W.F.M. 2002. Biótopos do Intertidal
rochoso da ilha de São Miguel, Açores. Marine
Biology Final Dissertation Thesis. Azores
University. [In Portuguese].
Martins, G.M., S.J. Hawkins, R.C. Thompson & S.R.
Jenkins 2007. Community structure and
functioning in intertidal rock pools: effects of pool
size and shore height at different successional
stages. Marine Ecology Progress Series 329:
43-55.
Metaxas, A. & R.E. Scheibling 1993. Community
structure and organization of tidepools. Marine
Ecology Progress Series 98: 187-198.
Metataxas, A., H.L. Hunt, R.E. Scheibling 1994.
Spatial and temporal variability of macrobenthic
communities in tide pools on a rocky shore in Nova
Scotia, Canada. Marine Ecology Progress Series
105: 89-103.
Morton, B, J.C. Britton & A.M.F. Martins 1998.
Coastal Ecology of the Azoress. Sociedade Afonso
Chaves. 249 pp.
Neto, A.I. & H.P. Baldwin 1990. Algas marinhas das
ilhas do Corvo e Flores. Relatórios e comunicações
do Departamento de Biologia 18:19-25.
Neto, A.I. 1992. Contribution to the taxonomy and
ecology of Azorean benthic marine algae.
Biological Journal of the Linnean Society 46: 163176.
Neto, A.I. 2000. Ecology and dynamics of two
intertidal algal communities on the littoral of the
island of São Miguel (Azores). Hydrobiologia 432:
135-147.
Steneck, R.S. & M.N. Dethier 1994. A functional
group approach to the structure of algal-dominated
communities. Oikos 69: 76-498.
Therriault, T.W. & J. Kolasa 1999. Physical
determinants of richness, diversity, eveness and
abundance in natural aquatic microcosms.
Hydrobiologia 412: 123-130.
Underwood, A.J. 1980. The effects of grazing by
gastropods and physical factors on the upper limits
of distribution of intertidal macroalgae. Oecologia
46: 201-213.
Underwood, A.J. & G.A. Skilleter 1996. Effects of
patch-size on the structure of assemblages in rock
pools. Journal of Experimental Marine Biology and
Ecology 197: 63–90.
Wallenstein, F.F.M.M. & A.I. Neto 2006. Intertidal
rocky shore biotopes of the Azores: a quantitative
approach. Helgoland Marine Research 60: 196206.
Wallenstein, F.F.M.M., A.I. Neto, N.V. Álvaro & C.I.
Santos 2008. Algae based biotopes of the Azores
(Portugal): spatial and seasonal variation. Aquatic
Ecology 42(4): 547-559.
Wolfe, J. & M. Harlin 1988. Tidepools in Southern
Rhode Island, U.S.A. II. Species diversity and
similarity analysis of macroalgal communities.
Botanica Marina 31: 537-546.
Accepted 14 June 2010.
17
APPENDIX
Table 1. List of taxa identified in the islands of Santa Maria and São Miguel, and corresponding authorities (1/3).
Phaeophyceae
Chlorophyceae
Class
18
Species
Blidingia minima (Nägeli ex Kützing) Kylin
Bryopsis cupressina J.V. Lamouroux
Bryopsis hypnoides J.V. Lamouroux
Bryopsis plumosa (Hudson) C. Agardh
Chaetomorpha aerea (Dillwyn) Kützing
Chaetomorpha pachynema (Montagne) Kützing
Cladophora albida (Nees) Kutzing
Cladophora coelothrix Kützing
Cladophora hutchinsiae (Dillwyn) Kützing
Cladophora laetvirens (Dillwyn) Kützing
Cladophora lehmanniana (Lindenberg) Kützing
Cladophora liebetruthii Grunow
Cladophora pellucida (Hudson) Kützing
Cladophora prolifera (Roth) Kützing
Cladophora sp.
Codium adhaerens C. Agardh
Derbesia tenuissima (Moris & De Notaris) P.L. Crouan & H.M.
Crouan
Ulva clathrata (Roth) C. Agardh
Ulva compressa Linnaeus
Ulva intestinalis Linnaeus
Ulva lingulata A.P.de Candolle
Ulva prolifera O.F. Müller
Ulva ralfsii (Harvey) Le Jolis
Ulva rigida C. Agardh
Ulva sp.
Valonia utricularis (Roth) C. Agardh
Bachelotia antillarum (Grunow) Gerloff
Cladostephus spongiosus (Hudson) C. Agardh
Colpomenia sinuosa (Mertens ex Roth) Derbès & Solier
Cystoseira abies-marina (S.G. Gmelin) C. Agardh
Cystoseira humilis Schousboe ex Kützing
Cystoseira sp.
Dictyota dichotoma (Hudson) J.V. Lamouroux
Ectocarpus sp.
Feldmannia irregularis (Kützing) G. Hamel
Fucus spiralis Linnaeus
Halopteris filicina (Grateloup) Kützing
Hincksia sp.
Sargassum cymosum C. Agardh
Sargassum vulgare C. Agardh
Scytosiphon lomentaria (Lyngbye) Link
Santa Maria
São Miguel
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
Rhodophyceae
Phaeophyceae
Table 1. List of taxa identified in the islands of Santa Maria and São Miguel, and corresponding authorities (2/3).
Sphacelaria cirrosa (Roth) C. Agardh
Sphacelaria rigidula Kützing
Sphacelaria sp.
Sphacelaria tribuloides Meneghini
Stypocaulon scoparium (Linnaeus) Kützing
Zonaria tournefortii (J. V. Lamouroux) Montagne
Acrochaetium crassipes (Børgesen) Børgesen
Acrosorium venulosum (Zanardini) Kylin
Aglaothamnion sp.
Amphiroa fragilissima (Linnaeus) J.V. Lamouroux
Amphiroa rigida J.V. Lamouroux
Anotrichium furcellatum (J. Agardh) Baldock
Antithamnionella spirographidis (Schiffner) E.M. Wollaston
Asparagopsis armata Harvey
Boergeseniella fruticulosa (Wulfen) Kylin
Botryocladia sp.
Centroceras clavulatum (C. Agardh) Montagne
Ceramium botryocarpum A.W. Griffiths ex Harvey
Ceramium ciliatum (J. Ellis) Ducluzeau
Ceramium codii (H. Richards) Mazoyer
Ceramium diaphanum (Lightfoot) Roth
Ceramium echionotum J. Agardh
Ceramium rubrum C. Agardh
Ceramium sp.
Chondracanthus acicularis (Roth) Frederiq
Chondria coerulescens (J. Agardh) Falkenberg
Chondria dasyphylla (Woodward) C. Agardh
Corallina elongata J. Ellis & Solander
Dasya corymbifera J. Agardh
Diplothamnion sp.
Erythrocystis montagnei (Derbès & Solier) P.C. Silva
Erythrotrichia carnea (Dillwyn) J. Agardh
Falkenbergia rufolanosa (Harvey) F. Schmitz
Gelidium arbusculum Bory de Saint-Vincent ex Børgesen
Gelidium pusillum (Stackhouse) Le Jolis
Grateloupia filicina (J.V. Lamouroux) C. Agardh
Halarachnion ligulatum (Woodward) Kützing
Haliptilon virgatum (Zanardini) Garbary & H.W. Johansen
Herposiphonia sp.A
Heterosiphonia sp.B
Hypnea arbuscula P. Dangeard
Hypnea musciformis (Wulfen) J. V. Lamouroux
Jania adhaerens J.V. Lamouroux
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
19
Rhodophyceae
Table 1. List of taxa identified in the islands of Santa Maria and São Miguel, and corresponding authorities (3/3).
20
Jania capillacea Harvey
Jania pumila J.V. Lamouroux
Jania rubens (Linnaeus) J.V. Lamouroux
Laurencia minuta H. Vandermeulen, D.J. Garbary & M.D. Guiry
Laurencia tenera C.K. Tseng
Laurencia viridis Gil-Rodríguez & Haroun
Lophocladia sp. (Mertens ex C. Agardh) F. Schmitz
Monosporus pedicellatus (J.E. Smith) Solier
Peyssonnelia rubra (Greville) J. Agardh
Plocamium cartilagineum (Linnaeus) P.S. Dixon
Polysiphonia denudata (Dillwyn) Greville ex Harvey
Polysiphonia furcellata (C. Agardh) Harvey
Polysiphonia sp.
Porphyra sp.
Pterocladiella capillacea (S.G. Gmelin) Santelices & Hommersand
Rhodymenia holmesii Ardissone
Spyridia filamentosa (Wulfen) Harvey
Symphyocladia marchantioides (Harvey) Falkenberg
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
X
Check‐list of interstitial polychaetes from intertidal and shallow subtidal soft bottoms of Tenerife, Canary Islands RODRIGO RIERA, JORGE NÚÑEZ & MARÍA DEL CARMEN BRITO
Riera, R., J. Núñez & M.C. Brito 2010. Check-list of interstitial polychaetes from
intertidal and shallow subtidal soft bottoms of Tenerife, Canary Islands.
Arquipelago. Life and Marine Sciences 27: 21-39.
A check-list of polychaete species from two stations on the south coast of Tenerife (Los
Abrigos and Los Cristianos) at two different tidal levels, intertidal and shallow subtidal (3
m depth) is presented. A total of 47 species were collected, the hesionid Microphthalmus
pseudoaberrans Campoy & Viéitez, 1982 and the spionids Rhynchospio glutaea (Ehlers,
1897) and Spio filicornis (O.F. Müller, 1776) being the most abundant. With 18 species the
family Syllidae is the most diverse, followed by the Spionidae and Paraonidae with 6 and 5
species, respectively. The interstitial polychaetes found are represented by both meiofaunalsized and small-sized macrofaunal species.
Key words: carbonates, ecology, granulometry, organic matter, Polychaeta, sand
Rodrigo Riera (e-mail: [email protected]), Centro de Investigaciones Medioambientales del Atlántico (CIMA SL), Arzobispo Elías Yanes, 44, ES-38206 La Laguna,
Tenerife, Canary Islands, Spain; Jorge Núñez & María del Carmen Brito, Benthos Lab,
Department of Animal Biology, Faculty of Biology, University of La Laguna, ES-38206 La
Laguna, Tenerife, Canary Islands, Spain.
INTRODUCTION
In spite of the abrupt and hilly coastal
morphology in the Canarian archipelago, the
presence of volcanic and organogenic sandy
beaches is frequent. The most extensive beaches
are located in the western islands (Lanzarote and
Fuerteventura). In contrast, beaches are smaller
and scarcer in the eastern islands, particularly in
La Gomera and El Hierro. This study is focussed
on the interstitial polychaetes of two beaches of
Tenerife that forms together with Gran Canaria
the central block of the Canarian archipelago. The
former two islands are the most inhabited and
thus affected by tourism, for example with the
presence of artificial modifications of the littoral
(harbours, docks, coastal avenues, etc.). The
south coast of Tenerife is modified in most of its
extension, with several artificial beaches and
dikes. These beaches are composed of sands from
dredged material.
This study represents the first characterization of
the interstitial polychaete fauna from the Canary
Islands, with special emphasis on the diversity of
coastal habitats like. The present check-list is not
limited only to meiofaunal polychaetes, since
animals belonging to macrofaunal-sized species
were also present (e.g. Capitomastus minimus,
among others).
MATERIAL AND METHODS
Two sandy beaches were sampled, Los Abrigos
(SE Tenerife) and Los Cristianos (SW Tenerife)
(Fig. 1). Los Abrigos beach is characterized by
volcanic sands with a low content of carbonates,
whilst in Los Cristianos beach organogenic sands
are present with a high content of carbonates
(Riera 2004).
Samples from Los Abrigos and Los Cristianos
were collected in the intertidal at mean low tide
21
level and shallow subtidal (3 m depth) by means
of PVC cores to a sediment depth of 30 cm
(475 cm3). Replicates were collected for faunistic
analysis and for abiotic factors (granulometry,
organic matter and carbonates). Samples were
fixed and preserved in 4% neutralized formalinseawater solution. Thereafter, samples were
sieved through a 63 µm mesh and then transferred
to 70% ethanol. Several whole specimens were
mounted in jelly glycerine; examination was
made by means of a compound microscope
provided with differential interference contrast
optics (Nomarski).
The studied material is deposited in the
collections of the Department of Animal Biology
(Zoology) of the University of La Laguna, Canary
Islands. Systematics adopted in the present study
followed Rouse & Fauchald (1997) proposal.
Fig. 1. Map of the investigated areas showing the location of sampling stations: CI and CS,
Los Cristianos intertidal and subtidal; AI and AS, Los Abrigos intertidal and subtidal samples.
22
RESULTS
SPECIES CHECK-LIST
FAMILY CAPITELLIDAE Grube, 1862
Capitomastus Eisig, 1887
Capitomastus minimus (Langerhans, 1880)
Capitella minimus Langerhans 1880: 299, figs. 4,
12.
Capitomastus minimus: Hartmann-Schröder
1971: 396, fig. 140a-c.
Material examined: Los Abrigos beach: subtidal,
13 specimens; Los Cristianos beach: subtidal, 4
specimens.
Ecology: Los Abrigos subtidal: in well sorted fine
sands with 0.77% organic matter and 9.57%
carbonates. Los Cristianos subtidal: in well sorted
fine sands with 0.02% organic matter and 22.56%
carbonates. Canary Islands: in seagrass meadows
(5-18 m depth) (Brito 2002), being a dominant
species. Mediterranean Sea: more frequent in
subtidal soft-bottoms (Capaccioni 1987),
although also present on hard substrates (Sardá
1985; Alós 1988).
Distribution: East Atlantic (Rullier & Amoreux
1970). Mediterranean (Desbrúyeres et al. 1972).
Red Sea (Ben-Eliahu 1976). Canary Islands:
Lanzarote, Tenerife (Brito 2002).
Notomastus Sars, 1851
Notomastus latericeus Sars, 1851
Notomastus latericeus: Fauvel 1927: 143, fig. 49
a-h.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands with 1.68% organic matter and
3.08% carbonates. Canary Islands: in seagrass
meadows (8-14 m depth), as a non-dominant and
solely in subtidal soft-bottoms (Brito 2002).
Atlantic-Mediterranean region: in subtidal soft
bottoms (Bellan 1964) and hard substrates (Sardá
1984). In deep sediments at depths of 4.000 m
(Amoreux 1971) and can be considered to be
euryhaline (Viéitez 1976).
Distribution: Cosmopolitan of cold and warm
waters (Rullier & Amoureux 1970; Desbrúyeres
et al. 1972; Ben-Eliahu 1976). Canary Islands:
Lanzarote, Fuerteventura, Gran Canaria and
Tenerife (Langerhans 1884; Núñez et al. 1984).
FAMILY OPHELIIDAE Malmgren, 1867
Ophelia Savigny in Lamarck, 1818
Ophelia bicornis Savigny in Lamarck, 1818
Ophelia bicornis: Fauvel 1927: 130, fig. 46 a, f.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands with 1.68% organic matter and
3.08% carbonates. Canary Islands: in Cymodocea
seagrass meadows at 14 m depth (Brito 2002).
Atlantic-Mediterranean region: in intertidal and
subtidal soft bottoms, more abundant in coarse
sands (Junoy 1988).
Distribution: East Atlantic (Rioja 1917).
Mediterranean (Bellan 1964). Canary Islands:
Tenerife (Brito 2002).
Pararicia Solís-Weiss & Fauchald, 1989
Pararicia sp.
Material examined: Los Cristianos beach:
subtidal, 2 specimens.
Ecology: Los Cristianos subtidal: in well sorted
fine sands with 0.71-0.86% organic matter and
24.27-26.84%.carbonates
Distribution: First record of this genus for the
Canary Islands (Tenerife).
FAMILY ORBINIIDAE Hartman, 1942
Schroederella Laubier, 1962
Schroederella laubieri Badalamenti & Castelli,
1991
Schroederella laubieri Badalamenti & Castelli,
1991: 218, figs. 1, 2 a-f.
23
Material examined: Los Abrigos beach: intertidal,
1 specimen; subtidal, 27 specimens; Los Cristianos beach: subtidal, 27 specimens.
Ecology: Los Abrigos: in well sorted fine and
medium sands with 0.50-1.33% organic matter
and 1.54-5.30% carbonates. Los Cristianos: in
well sorted fine sands, with 0.006-0.86% organic
matter and 24.27-25.30% carbonates. Canary
Islands: in shallow subtidal soft-bottoms at depths
between 5 and 15 m (Brito 2002). Mediterranean
Sea: in subtidal coarse sand bottoms and
Cymodocea nodosa meadows (Badalamenti &
Castelli 1991).
Distribution: Mediterranean (Badalamenti &
Castelli 1991). Atlantic Ocean, Canary Islands:
Lanzarote, Fuerteventura and Tenerife (Brito
2002).
Scoloplos Blainville, 1828
Scoloplos armiger (O.F. Müller, 1776)
Scoloplos armiger: Fauvel 1927: 20-21, fig.6 k-q.
Material examined: Los Abrigos beach: subtidal,
2 specimens; Los Cristianos beach: subtidal, 4
specimens.
Ecology: Los Abrigos subtidal, in well sorted fine
sands, with 0.50% organic matter and 1.54%
carbonates. Los Cristianos subtidal, in well sorted
fine sands, with 0.73%. organic matter and
24.96%
carbonates.
Atlantic-Mediterranean
region: in soft-bottoms, being more abundant in
medium sand bottoms and can be considered to
be stenobathic (Moreira 1999).
Distribution: Cosmopolitan (Intes & Le Loeuff
1977). First record of this species for the Canary
Islands (Tenerife).
FAMILY PARAONIDAE Cerruti, 1909
Acmira (Hartley, 1981)
Aricidea (Acmira) assimilis Tebble, 1959
Aricidea (Acesta) assimilis: Strelzov 1979: 108,
figs. 7, 16, 39.
Material examined: Los Cristianos beach:
subtidal, 1 specimen.
Ecology: Los Cristianos subtidal: in well sorted
fine sands, with 0.73% organic matter content and
24
24.96% carbonates. Canary Islands: it can reach
high abundances in Lanzarote and Fuerteventura,
being dominant along the year (Brito 2002).
Atlantic-Mediterranean region: in soft-bottoms,
from shallow subtidal to 1.000 metres depth
(Strelzov 1979).
Distribution: East Atlantic (Day 1961; Strelzov
1979). Mediterranean (Campoy 1982). East
Pacific (Hobson & Banse 1981). Canary Islands:
Tenerife (Brito 2002).
Aricidea (Acmira) catherinae Laubier, 1967
Aricidea (Acmira) catherinae: Gaston 1984: 56,
fig. 2-43; Montiel, Hilbig & Rozbaczylo 2002:
136, fig. 2 e-g.
Material examined: Los Abrigos beach: subtidal,
2 specimens.
Ecology: Los Abrigos: in well sorted fine sands,
with 0.54% organic matter and 5.98% carbonates.
Canary Islands: in subtidal soft-bottoms (Brito
2002). Altantic-Mediterranean region: in subtidal
soft-bottoms, being euribathic (Gaston 1984) and
eurihaline (Capaccioni 1987).
Distribution: Amphiatlantic (Pettibone 1963;
Campoy 1982). East Pacific (Hartman 1963).
Mediterranean (Desbrúyeres et al. 1972). Canary
Islands (Tenerife) (Brito 2002).
Cirrophorus Ehlers, 1908
Cirrophorus armatus (Glémarec, 1966)
Cirrophorus armatus: Strelzov 1968: 131, fig. 47
a-e; Hartmann-Schröder 1996: 383, fig. 180.
Material examined: Los Cristianos beach:
subtidal, 1 specimen.
Ecology: Los Cristianos, in well sorted fine
sands, with 0.81% organic matter and 24.10%
carbonates.
Canary Islands: in bare soft-bottoms and
Cymodocea nodosa meadows (Brito 2002).
Mediterranean sea: in muddy and sandy seabeds
(Capaccioni 1987).
Distribution: East Atlantic (Glémarec 1966).
Mediterranean (Harmelin 1969). Canary Islands:
Lanzarote, Fuerteventura and Tenerife (Brito
2002).
Cirrophorus furcatus Hartman, 1957
Cirrophorus furcatus: Strelkov 1979: 140, figs.
18.5, 50a-e.
Material examined: Los Abrigos beach: subtidal,
6 specimens.
Ecology: Los Abrigos subtidal: in well sorted fine
sands, with 0.85% organic matter and 7.18%
carbonates. Mediterranean sea: in soft-bottoms
with diverse granulometry (Desbrúyeres et al.
1972).
Distribution: East Atlantic (Laubier & Ramos
1973). Mediterranean (Desbrúyeres et al. 1972).
East Pacific (Hartman 1969). Canary Islands:
Tenerife. First record of this species in the Canary
Islands.
Cirrophorus perdidoensis McLelland & Gaston,
1994
Cirrophorus perdidoensis: McLelland & Gaston
1994: 525, fig. 1 a-e.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 1.68% organic matter and
3.08% carbonates. Canary Islands: in softbottoms, being constant along the year (Brito
2002). Eastern Atlantic Ocean: in shallow
subtidal soft-bottoms (McLellan & Gaston 1994).
Distribution: Amphiatlantic (McLelland &
Gaston 1994; Brito 2002). Canary Islands:
Lanzarote, Fuerteventura, Tenerife and La Palma.
Levinsenia Mesnil, 1897
Levinsenia canariensis (Brito & Núñez, 2002)
Periquesta canariensis Brito & Núñez, 2002:
284, figs. 2-3.
Levinsenia canariensis: Giere et al. 2008: 312,
fig. 3 A-C.
Material examined: Los Abrigos beach: intertidal,
3 specimens; subtidal, 71 specimens.
Ecology: Los Abrigos intertidal: well sorted
medium sands, with 1.30% organic matter and
4.44% carbonates. Los Abrigos subtidal: in well
sorted fine sands, with 1.54% organic matter and
6.84% carbonates. Canary Islands: in bare softbottoms and Cymodocea nodosa meadows, being
more abundant during summer (Brito 2002).
Distribution: Selvagens Islands (Núñez et al.
2001; Brito & Núñez 2002). Canary Islands:
Lanzarote, Gran Canaria, Tenerife and El Hierro.
FAMILY SABELLIDAE Malmgren, 1867
Desdemona Banse, 1957
Desdemona sp.
Material examined: Los Abrigos beach: intertidal,
2 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.33-1.62% organic matter
and 5.47-5.64%.carbonates.
Distribution: First record of this genus for the
Canary Islands (Tenerife).
Novafabricia Fitzhugh, 1990
Novafabricia sp.
Material examined: Los Abrigos beach: intertidal,
1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.33% organic matter and
5.47% carbonates.
Distribution: First record of this genus for the
Canary Islands (Tenerife).
FAMILY SPIONIDAE Grube, 1850
Dispio Hartman, 1951
Dispio uncinata Hartman, 1951
Dispio uncinata Hartman 1951: 87, fig. 22, figs.
1-5. fig. 23, figs. 1-4;
Material examined: Los Abrigos beach: intertidal,
1 specimen, subtidal, 5 specimens; Cristianos:
intertidal, 1 specimen, subtidal, 1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.33% organic matter content
and 5.30% carbonates. Los Abrigos subtidal: in
well sorted medium sands, with 1.31-1.54%
organic matter and 5.13-6.84% carbonates. Los
Cristianos intertidal, in well sorted fine sands,
with 0.93% organic matter and 23.59%
carbonates. Los Cristianos subtidal: in well sorted
fine sands, with 0.86% organic matter and
25
24.27% carbonates. Canary Islands: in bare softbottoms and Cymodocea nodosa meadows, more
abundant in intertidal and shallow seabeds (Brito
2002). Atlantic-Mediterranean region: in sandy
bottoms, from the intertidal to 100 m depth
(Bellan 1969; Ibáñez & Viéitez 1973).
Distribution: Cosmopolitan (Hartman 1969;
Ibáñez & Viéitez 1973; Uebelacker 1984).
Canary Islands: Tenerife and La Palma (Brito
2002).
Pseudopolydora Czerniavksky, 1881
Pseudopolydora sp.
Material examined: Los Cristianos beach:
intertidal, 1 specimen.
Ecology: Los Cristianos intertidal, it was
collected in well sorted fine sands. The organic
matter content was 0.86% and 24.27% of
carbonates.
Distribution: First record of this genus for the
Canary Islands (Tenerife).
Rhynchospio Hartman, 1936
Rhynchospio glutaea (Ehlers, 1897)
Rhynchospio glutaea: Imajima 1991: 10, fig.4a-q.
Material examined: Los Abrigos beach: intertidal,
8 specimens; subtidal, 74 specimens; Los
Cristianos beach: intertidal, 21 specimens,
subtidal, 114 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.03% organic matter content
and 5.81% carbonates. Los Abrigos subtidal: in
well sorted medium sands, with 0.67-0.91%
organic matter and 3.78-6.32% carbonates. Los
Cristianos intertidal, in well sorted fine sands,
with 0.006% to 0.29%, organic matter and 15.7818.97% carbonates. Los Cristianos subtidal: in
well sorted fine sands, with 0.46-0.73% organic
matter and 23.78-24.96% carbonates. Canary
Islands: in Cymodocea nodosa meadows at 6 m
depth (Brito 2002). East Atlantic Ocean: in
sponges and laminarians (Blake et al. 1996).
Distribution: Cosmopolitan (Day, 1967; Carrasco,
1974; Blake, 1983). Canary Islands: Fuerteventura and Tenerife (Brito 2002).
26
Scolelepis Blainville, 1828
Scolelepis (Scolelepis) squamata (O.F. Müller,
1806)
Scolelepis squamata: Day, 1967: 483, fig. 18.7
c-h.
Material examined: Los Abrigos beach: intertidal,
4 specimens; Los Cristianos beach: intertidal, 16
specimens, subtidal, 1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 0.86-1.11% organic matter
and 3.56-5.81% carbonates. Los Cristianos
intertidal: in well sorted medium sands, with
0.006-0.93% organic matter and 17.89-18.29%
carbonates. Los Cristianos subtidal: in well sorted
fine sands, with 0.73% organic matter content and
24.96% carbonates. Canary Islands: in subtidal
soft-bottoms (Herrando-Pérez et al. 2001).
Atlantic Ocean: in sandy soft-bottoms at a depth
of 25 metres (Maciolek 1987).
Distribution: Cosmopolitan of warm and tropical
waters (Day 1973; Maciolek 1983; Parapar 1991).
Canary Islands: Tenerife and La Gomera
(Herrando et al. 2001).
Spio Fabricius, 1785
Spio decoratus Bobretzky, 1870
Spio decoratus: Dauvin 1989: 167, fig. 1; Parapar
1991: 156, fig. 44 c.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 1.68% organic matter content
and 3.08% carbonates. Canary Islands: in bare
soft-bottoms and Cymodocea nodosa meadows, at
depths between 8 and 13 m (Brito 2002).
Altantic-Mediterranean region: between hard
substrates and sandy bottoms (Bellan 1971), more
abundant in muddy and muddy sand bottoms with
a low organic matter content (Capaccioni 1987),
and harbour areas (Tena 1992).
Distribution: Atlantic (López-Jamar 1978).
Mediterranean (San Martín & Alvarado 1982).
Canary Islands: Gran Canaria, Tenerife and La
Gomera (Brito 2002; Herrando et al. 2001).
Spio filicornis (O.F. Müller, 1766)
Spio filicornis: Fauvel 1927: 43; fig. 15 a-g; Day
1967: 481, fig. 18.6 l-o.
Material examined: Los Abrigos beach: intertidal,
5 specimens, subtidal, 334 specimens; Los
Cristianos beach: intertidal, 17 specimens,
subtidal, 269 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 0.76-1.30% organic matter
and 3.26-4.44% carbonates. Los Abrigos subtidal:
in well sorted medium sands, with 0.76-0.91%
organic matter and 4.56-6.32% carbonates. Los
Cristianos intertidal: in well sorted fine sands,
with 0.76-1.24% organic matter and 19.8723.59% carbonates. Los Cristianos subtidal: in
well sorted fine sands, with 0.006-0.86% organic
matter and 19.98-24.27% carbonates. Canary
Islands: in bare soft bottoms and Cymodocea
nodosa meadows, at 5-10 m depth (Brito 2002).
Eastern Atlantic Ocean: higher densities in the
subtidal (Parapar 1991). Western Atlantic: at
depths of 70 m (Maciolek 1990).
Distribution: Amphiatlantic (Maciolek 1990;
Parapar 1991). Western Pacific (Day 1967).
Canary Islands: Fuerteventura, Gran Canaria and
Tenerife (Brito 2002).
FAMILY CIRRATULIDAE Carus, 1863
Cauleriella Chamberlin, 1919
Cauleriella bioculata (Keferstein, 1862)
Cauleriella bioculata: Hartmann-Schröder 1971:
355.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 1.31% organic matter and
5.13% carbonates. Canary Islands: in Cymodocea
nodosa meadows at 15 m depth (Brito 2002) and
as a component of the enbobiontic fauna of
sponges (Pascual 1996). Atlantic-Mediterranean
region: in hard substrates (Sardá 1982);
considered to be eurihaline (Rullier 1972).
Distribution: Eastern Atlantic (Langerhans 1881),
Mediterranean (Alós et al. 1982). Black Sea
(Rullier 1972). Eastern Pacific (HartmannSchröder 1971). Canary Islands: Tenerife
(Langerhans 1881; Núñez et al. 1984; Pascual
1996; Brito 2002).
Cirriformia Hartman, 1936
Cirriformia tentaculata (Montagu, 1808)
Cirriformia tentaculata: Day 1967: 515, fig. 20.4
a-d.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted fine
sands, 0.85% organic matter and 7.18%
carbonates.
Canary Islands: in intertidal rocky substrates,
with preference to altered habitats (Núñez pers.
comm.). Atlantic-Mediterranean region: in rocky
substrates; abundant in subtidal soft-bottoms with
a high organic matter content (Sardá 1984).
Distribution: Cosmopolitan of warm and tropical
waters (Núñez et al. 1984). Canary Islands:
Tenerife (Kirkegaard 1959; Núñez et al. 1984).
Aphelochaeta Blake, 1991
Aphelochaeta marioni (Saint-Joseph, 1894)
Tharyx marioni: Fauvel 1927: 100, fig. 35 a-b.
Material examined: Los Abrigos beach: intertidal,
1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.30% organic content and
4.44% carbonates. Atlantic-Mediterranean region:
in muddy and sandy bottoms (Laubier 1966;
Parapar 1991), hard substrates (Martín 1986).
Tharyx marioni is an euryhaline (Holthe 1977;
López-Jamar 1981) and eurybathic species
(Hartman & Fauchald 1971).
Distribution: Amphiatlantic (Fauvel 1927;
Amoureux 1976). Mediterranean (Desbrúyeres et
al. 1972). Indian Ocean (Intes & Le Loeuff 1977).
First record of this species for the Canary Islands
(Tenerife).
FAMILY PROTODRILIDAE Czerniavsky, 1881
Protodrilus Hatschek, 1880
Protodrilus cf. rubropharyngeus Jägersten, 1940
27
Protodrilus rubropharyngeus: von Nordheim
1989: 252, tab. I.
Material examined: Los Abrigos beach: intertidal,
29 specimens; subtidal, 1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 0.32-1.03% organic matter
and 4.44-5.81% carbonates. Los Abrigos subtidal:
in well sorted medium sands, with 0.78% organic
matter and 4.56% carbonates. Atlantic Ocean:
more abundant in coarse sandy bottoms (von
Nordheim 1989).
Distribution: East Atlantic (von Nordheim 1989).
West Pacific (Wu et al. 1980). First record of this
species for the Canary Islands (Tenerife).
and 4.44-5.81% carbonates. Los Abrigos subtidal:
in well sorted fine sands, with 0.24-0.51%
organic matter and 3.56-4.61% carbonates. Los
Cristianos intertidal: in well sorted fine sands,
with 0.006-0.81% organic matter and 15.4617.78% carbonates. Los Cristianos subtidal: in
well sorted fine sands, with 1.01% organic matter
and 26.84% carbonates. Canary Islands: in sandy
bottoms at 22 m depth (Brito 2002). AtlanticMediterranean region: in well oxygenated sandy
bottoms, with a low content of organic matter
(Campoy & Viéitez 1982).
Distribution: East Atlantic (Campoy & Viéitez
1982). Madeira (Núñez et al. 1995).
Mediterranean (Capaccioni 1983). Canary
Islands: Tenerife and La Palma (Brito 2002).
FAMILY HESIONIDAE Grube, 1850
Hesionides Friedich, 1937
Hesionides arenaria Friedich, 1937
Hesionides arenaria: Hartmann-Schröder 1971:
134, fig. 44 a-c.
Material examined: Los Abrigos beach: intertidal,
1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 0.85% organic matter and
5.47% carbonates. Canary Islands: in intertidal
soft-bottoms (Schmidt & Westheide 2000), with
preference to exposed sandy beaches. AltanticMediterranean region: in coarse sand bottoms
(Campoy 1982); considered to be estenobathic
(Hartmann-Schröder 1971).
Distribution: Cosmopolitan (Schmidt & Westheide 2000). Canary Islands: Tenerife.
Microphthalmus Mecznikow, 1865
Microphthalmus pseudoaberrans Campoy &
Viéitez, 1982
Microphthalmus pseudoaberrans: Campoy &
Viéitez 1982: 224, fig. 23 a-n.
Material examined: Los Abrigos beach: intertidal,
822 specimens.; subtidal, 9 specimens.; Los
Cristianos beach: intertidal, 14 specimens;
subtidal, 1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 0.76-1.03% organic matter
28
FAMILY NEREIDIDAE Johnston, 1865
Platynereis Kinberg, 1866
Platynereis dumerilii (Audouin & Milne
Edwards, 1834)
Platynereis dumerilii: Fauvel 1923: 359, fig. 141
a-f.
Material examined: Los Cristianos beach:
intertidal, 1 specimen; subtidal, 1 specimen.
Ecology: Los Cristianos intertidal: in well sorted
fine sands, with 1.24% organic matter and
23.59% carbonates. Los Cristianos subtidal: in
well sorted fine sands, with 0.71% organic matter
and 26.84% carbonates. Canary Islands: in
intertidal algae, endobiontic of sponges and
subtidal soft-bottoms (Núñez 1990). AtlanticMediterranean region: inside polluted harbours
(Desbrúyeres et al. 1972; Bellan 1980), with
preference to photophilic algae (Harmelin 1964;
Laubier 1966). It has been recorded in softbottoms (Sardá 1985), being an euryhaline
(Hartmann-Schröder 1971) and an eurybathic
species (Hartman 1965).
Distribution: Cosmopolitan (López 1995).
Canary Islands: Lanzarote, Fuerteventura, Gran
Canaria, Tenerife, La Gomera, La Palma and El
Hierro (Langerhans 1881; May 1912; Fauvel
1914; Núñez et al. 1981; Kirkegaard 1983; Núñez
et al. 1984; Núñez 1990; Pascual 1996).
Perinereis Kinberg, 1866
Perinereis cultrifera (Grube, 1840)
Perinereis cultrifera: Fauvel 1914: 190, fig. 16,
figs. 1-13.
Material examined Los Abrigos beach: subtidal, 1
specimen.
Ecology: Los Abrigos subtidal: in well sorted fine
sands, with 0.50% organic matter and 1.54%
carbonates. Canary Islands: in rocky bottoms and
intertidal coarse sandy bottoms, as well as,
Cymodocea nodosa meadows at 12 m depth
(Núñez 1990). Atlantic-Mediterranean region: in
soft and hard bottoms, being more abundant in
rocky substrates (Campoy 1982; López 1995). In
soft-bottoms, muddy-sand bottoms (Desbrúyeres
et al. 1972; Sardá 1985) and in seagrass meadows
(Cymodocea nodosa and Posidonia oceanica)
(Alós & Pereira 1984; Baratech 1985).
Distribution: Cosmopolitan (López 1995).
Canary Islands: Lanzarote, Fuerteventura, Gran
Canaria, Tenerife, La Gomera, La Palma and El
Hierro (Langerhans 1881; May 1912; Núñez et al.
1981; Kirkegaard 1983; Núñez et al. 1984;
Talavera et al. 1984 Hartman-Schröder 1988;
Núñez 1993).
FAMILY SYLLIDAE Grube, 1850
SUBFAMILY EXOGONINAE Langerhans, 1879
Brania Quatrefages, 1866
Brania arminii (Langerhans, 1881)
Brania arminii: San Martín 2003: 153, figs.75-76.
Material examined: Los Abrigos beach: subtidal,
2 specimens.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 1.54% organic matter and
6.84% carbonates. Canary Islands: in intertidal
and shallow subtidal hard bottoms, among
photophilic algae and Spondylus and Vermetus
shells (Núñez 1990). Atlantic-Mediterranean
region: in shallow subtidal hard and soft bottoms
(Campoy 1982; Sardá 1984; Martín 1986); in
Posidonia oceanica meadows (San Martín &
Viéitez 1984) and “Amphioxus” sands (Besteiro
1986).
Distribution: Circumtropical (Capaccioni 1987).
Canary Islands: Lanzarote, Fuerteventura,
Tenerife, La Gomera and El Hierro (Langerhans
1881; Núñez et al. 1984; Núñez et al. 1992).
Exogone Örsted, 1845
Exgone (Exogone) breviantennata HartmannSchröder, 1959
Exogone breviantennata: Hartmann-Schröder
1959: 125, figs. 75-78.
Material examined: Los Abrigos beach: subtidal,
1 specimen; Los Cristianos beach: intertidal, 2
specimens, subtidal, 9 specimens.
Ecology: Los Abrigos subtidal: in well sorted fine
sands, with 0.51% organic matter and 4.61%
carbonates. Los Cristianos intertidal: in well
sorted fine sands, with 0.61-0.81% organic matter
and 17.78-18.63% carbonates. Los Cristianos
subtidal: in well sorted fine sands, with 0.0060.71% organic matter and 23.45-26.84%
carbonates. Canary Islands: in sandy bare bottoms
and Cymodocea nodosa meadows (Brito et al.
2000). In intertidal and subtidal rocky bottoms,
among algae and vermetid tubes (Núñez 1990),
and endobiontic of sponges (Pascual 1996).
Mediterranean sea: in intertidal and subtidal soft
and hard-bottoms (San Martín & Aguirre 1991).
Distribution: Circumtropical (Núñez 1990).
Canary Islands: Lanzarote, Fuerteventura, Gran
Canaria, Tenerife, La Palma and El Hierro
(Núñez et al. 1992; Pascual et al. 1996; Brito et
al. 2000; Brito 2002).
Exogone (Exogone) naidina Örsted, 1845
Exogone naidina: San Martín 1984: 208, fig. 46.
Material examined: Los Abrigos beach: intertidal,
1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.03% organic matter and
5.81% carbonates. Canary Islands: in intertidal
and subtidal hard and soft-bottoms, among
Spondylus shells and vermetid tubes (Núñez
1990). Altantic-Mediterranean region: in hard
substrates (Aguirrezabálaga 1984; San Martín
1984); and subtidal soft-bottoms (Moreira 1999).
Distribution: Cosmopolitan (Campoy 1982).
Canary Islands: Tenerife (Núñez 1990; Núñez et
al. 1992; Pascual 1996).
29
Parapionosyllis Fauvel, 1923
Parapionosyllis macaronesiensis Brito, Núñez &
San Martín, 2000
Parapionosyllis macaronesiensis: Brito, Núñez &
San Martín, 2000: 1147, fig. 1.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 0.91% organic matter and
6.32% carbonates. Canary Islands: in sandy bare
bottoms and Cymodocea nodosa meadows (Brito
2002). Madeira: in organogenic coarse sands,
with a high percentage of carbonates (59%)
(Núñez et al. 1995).
Distribution: Atlantic Ocean: Madeira (Núñez et
al. 1995). Canary Islands: Tenerife (Brito et al.
2000).
Parapionosyllis minuta (Pierantoni, 1903)
Pionosyllis minuta Pierantoni, 1903: 239, tab. 10,
fig. 5.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 1.54% organic matter and
6.84% carbonates. Atlantic-Mediterranean region:
as endobiontic of sponges (Pérès 1954), and in
sandy bare bottoms, in Caulerpa and seagrass
(Zostera marina and Posidonia oceanica)
meadows (Cognetti 1957; Viéitez 1976; San
Martín 1984).
Distribution: East Atlantic (San Martín 1984).
Mediterranean (Campoy 1982). First record of
this species for the Canary Islands (Tenerife).
Parapionosyllis abriguensis Riera, Núñez &
Brito, 2006
Parapionosyllis abriguensis: Riera, Núñez &
Brito 2006: 20, fig. 1.
Material examined: Los Abrigos beach: subtidal,
4 specimens.
Ecology: Los Abrigos subtidal: in well sorted fine
sands, with 0.11-1.31% organic matter and 5.137.18% carbonates.
Distribution: Canary Islands: Tenerife.
30
Erinaceusyllis San Martín, 2003
Erinaceusyllis cryptica (Ben-Eliahu, 1977)
Erinaceusyllis cryptica: San Martín 2003: 233,
fig. 124.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Cristianos subtidal: in well sorted
fine sands, with 0.006% organic matter and
19.32% carbonates. Canary Islands: as a
component of the Dendrophyllia ramea
community (circalittoral bottoms) (Núñez et al.
1992). Atlantic-Mediterranean region: in seagrass
meadows (Cymodocea, Posidonia and Zostera)
(Baratech & San Martín 1987; Parapar 1991) and
as endobiontic of sponges (Alós et al. 1982).
Distribution: Amphiatlantic (Perkins 1981;
Núñez 1990). Mediterranean (Campoy 1982).
Red Sea (Ben-Eliahu 1977). Canary Islands:
Tenerife, Lanzarote and El Hierro (Núñez 1990;
Núñez et al. 1992; Brito 2002).
Prosphaerosyllis San Martín, 1984
Prosphaerosyllis xarifae (Hartmann-Schröder,
1960)
Sphaerosyllis xarifae: Hartmann-Schröder 1960:
103, figs. 14, 15, figs. 121-124.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted fine
sands, with 0.50% organic matter and 1.54%
carbonates. Canary Islands: in soft and hard
bottoms, as well as a component of the
Dendrophyllia ramea community (Núñez 1990).
Atlantic-Mediterranean region: in hard substrates,
among algae and corals (Hartmann-Schröder
1979; Campoy 1982).
Distribution: East Atlantic (Hartmann-Schröder
1979). Mediterranean (San Martín 1984). Red Sea
(Hartmann-Schröder 1960). Canary Islands
(Tenerife) (Núñez 1990; Núñez et al. 1992).
SUBFAMILY EUSYLLINAE Malaquin, 1893
Perkinsyllis San Martín, López & Aguado, 2009
Perkinsyllis spinisetosa (San Martín, 1990)
Pionosyllis spinisetosa: San Martín 1990: 592,
figs. 2-3.
Material examined: Los Abrigos beach: intertidal,
1 specimen, subtidal, 40 specimens; Los Cristianos beach: subtidal, 4 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.30% organic matter and
4.44% carbonates. Los Abrigos subtidal: in well
sorted fine sands, with 0.54-1.23% organic matter
and 4.44-5.98% carbonates. Los Cristianos
subtidal: in well sorted fine sands, with 0.541.01% organic matter and 23.08-26.84%
carbonates. Canary Islands: in intertidal pools
(Núñez 1990), sandy bare bottoms and
Cymodocea nodosa meadows (Brito 2002). East
Atlantic Ocean: among algae (López & San
Martín 1994). Caribbean Sea: in sandy subtidal
bottoms (San Martín 1990).
Distribution: Amphiatlantic (San Martín 1990;
López & San Martín 1994; Núñez et al. 1995).
Canary Islands: Lanzarote and Tenerife (Núñez
1990; Núñez et al. 1996; Brito 2002).
Neopetitia San Martín, 2003
Neopetitia abadensis Riera, Núñez & Brito, 2007
Neopetitia abadensis: Riera, Núñez & Brito
2007: 221, figs 1-2.
Material examined: Los Abrigos: intertidal, 37
specimens; subtidal, 9 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 0.78-1.62% organic matter
and 4.44-5.64% carbonates. Los Abrigos subtidal:
in well sorted medium sands, with 0.65-168%
organic matter and 3.08-4.56% carbonates.
Distribution: Canary Islands: Tenerife.
Streptosyllis Webster & Benedict, 1884
Streptosyllis bidentata Southern, 1914
Streptosyllis bidentata: Southern 1914: 28, pl. 3,
fig. 4 a-f.
Material examined: Los Abrigos beach: intertidal,
11 specimens, subtidal, 28 specimens; Los
Cristianos beach: intertidal, 3 specimens, subtidal,
56 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 0.76-1.33% organic matter
and 4.44-5.30% carbonates. Los Abrigos subtidal:
in well sorted medium sands, with 0.78-1.68%
organic matter and 3.08-4.56% carbonates. Los
Cristianos intertidal: in well sorted fine sands,
with 0.006-1.24% organic matter and 19.7823.59% carbonates. Los Cristianos subtidal: in
well sorted fine sands, with 0.34-0.86% organic
matter and 23.59-24.27% carbonates.
Canary Islands: in sandy bare bottoms and
Cymodocea nodosa meadows (Brito 2002).
Atlantic-Mediterranean region: in sandy bottoms
from 2 to 24 m depth (Fauvel 1927; San Martín
1984) and very coarse sand with shells at 30 m
depth (Campoy 1982).
Distribution: East Atlantic Ocean. Mediterranean
(Brito et al. 2000). Canary Islands: Lanzarote,
Fuerteventura, Gran Canaria, Tenerife, La Palma
and El Hierro (Brito et al. 2000).
Streptosyllis campoyi Brito, Núñez & San Martín,
2000
Streptosyllis campoyi: Brito, Núñez & San Martín
2000: 611, fig. 5 a-l.
Material examined: Los Abrigos beach: intertidal,
3 specimens, subtidal, 21 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.06-1.33% organic matter
and 4.44-5.30% carbonates. Los Abrigos subtidal:
in well sorted medium sands, with 0.33-0.78%
organic matter and 3.56-5.47% carbonates .
Canary Islands: in sandy bare bottoms and
Cymodocea nodosa meadows (Brito 2002).
Atlantic-Mediterranean region: in “Amphioxus”
sands (Campoy 1982).
Distribution: East Atlantic (Campoy 1982).
Canary Islands: Lanzarote, Tenerife, La Palma
and La Gomera (Brito et al. 2000).
Anoplosyllis Claparède, 1868
Anoplosyllis edentula Claparède, 1868
Anoplosyllis edentula Claparède, 1868: 524; San
Martín 2003: 134, fig. 65.
Material examined: Los Abrigos beach: intertidal,
1 specimen, subtidal, 2 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.33% organic matter and
5.30% carbonates. Los Abrigos subtidal: in well
31
sorted medium sands, with 0.76-1.31% organic
matter and 4.44-5.13% carbonates.
Canary Islands: in sandy bare bottoms and
Cymodocea nodosa meadows (Brito 2002).
Atlantic-Mediterranean region: in soft-bottoms
with diverse granulometry (Cognetti 1957),
including polluted harbours (Cognetti-Varriale
1972), in intertidal pools, among algae and
endobiontic of sponges (Campoy 1982; San
Martín 1984).
Distribution: East Atlantic. Mediterranean.
Pacífico (Núñez 1990). Canary Islands: Lanzarote
and Tenerife (Campoy 1982; Núñez et al. 1984;
Núñez 1990; Brito 2002).
Syllides Örsted, 1845
Syllides japonicus Imajima, 1966
Syllides japonicas: Imajima 1966: 112, fig. 36 ah; San Martín 2003: 142, fig. 69.
Material examined: Los Abrigos beach: intertidal,
1 specimen, subtidal, 16 specimens.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.33% organic matter and
5.30% carbonates. Los Abrigos subtidal: in well
sorted fine sands, with 0.45-0.85% organic matter
and 4.44-7.18% carbonates. Canary Islands: in
sandy bare bottoms and Cymodocea nodosa
meadows (Brito 2002). Atlantic-Mediterranean
region: in sandy subtidal bottoms (Baratech &
San Martín 1987) and Posidonia meadows (San
Martín 1984).
Distribution: West Atlantic (Baratech & San
Martín 1987). Mediterranean (San Martín 1984).
Pacific Ocean (Imajima 1966). Canary Islands:
Lanzarote and Tenerife (Brito et al. 2000).
SUBFAMILY SYLLINAE Grube, 1850
Haplosyllis Langerhans, 1879
Haplosyllis aff. spongicola Grube, 1855
Haplosyllis spongicola: Núñez 1990: 365, fig.
112 a-c.
Material examined Los Cristianos beach:
intertidal, 1 specimen.
Ecology: Los Cristianos intertidal: in well sorted
fine sands, with 0.81% organic matter and
32
17.78% carbonates. Canary Islands: in shallow
hard substrates, among algae, vermetids and
corals, as well as, in Dendrophyllia ramea
community (circalittoral bottoms) and as
endobiontic of sponges (Núñez 1990, Pascual
1996). Atlantic-Mediterranean region: in soft and
hard bottoms, being less frequent as endodiontic
of sponges (Laubier 1966; Campoy 1982) and
soft-bottoms with diverse granulometry, from
muds to coarse sands (Uebelacker 1984) and can
be considered euribathic (Gardiner 1976).
Distribution: Cosmopolitan of warm and tropical
waters (Núñez 1990). Canary Islands: Lanzarote,
Fuerteventura, Gran Canaria, Tenerife, La
Gomera and El Hierro (Pascual 1996).
Syllis Savigny in Lamarck, 1818
Syllis armillaris (O.F. Müller, 1771)
Syllis armillaris: San Martín 1984: 381, figs. 99100.
Material examined: Los Abrigos beach: subtidal,
4 specimens.
Ecology: Los Abrigos subtidal: in well sorted fine
sands, with 0.74-0.85% organic matter content
and 4.44-7.86% carbonates. Canary Islands: in
intertidal and shallow rocky bottoms, among
algae, inside sponge, as well as in the
Dendrophyllia ramea community (circalittoral
bottom) (Núñez 1990). Atlantic-Mediterranean
region: in rocky substrates (San Martín 1984),
“Amphioxus” sands (Campoy 1982), muddy sands
(Desbrúyeres et al. 1972) and seagrass meadows
(Cymodocea nodosa and Zostera marina)
(Schlenz 1965).
Distribution: Cosmopolitan (Núñez 1990).
Canary Islands: Fuerteventura, Gran Canaria,
Tenerife, La Gomera and El Hierro (Núñez et al.
1992; Pascual 1996).
Syllis garciai (Campoy, 1982)
Syllis garciai: San Martín 1984: 364, fig. 92.
Material examined: Los Abrigos beach: subtidal,
7 specimens.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 0.76-1.31% organic matter
and 5.67-9.57% carbonates. Canary Islands: in
sandy bare bottoms at 10 m depth (Brito 2002). It
is frequent in soft and hard bottoms, being
encountered in the Dendrophyllia ramea
community (circalittoral bottoms) (Núñez 1990).
Atlantic-Mediterranean region: among algae,
vermetid tubes (Alós 1988; López 1995), sandy
bare bottoms (Besteiro, Urgorri & Parapar 1987)
and seagrass meadows (Cymodocea nodosa and
Posidonia
oceanica)
(Giangrande
1985;
Giangrande & Gambi 1986).
Distribution: Amphiatlantic (Besteiro 1986; San
Martín 1990). Mediterranean (Campoy 1982).
Canary Islands: Tenerife and Lanzarote (Núñez
1990; Núñez et al. 1992; Núñez et al. 1997).
Syllis prolifera Krohn, 1852
Syllis prolifera: San Martín 1984: 331, figs.78-79.
Material examined: Los Abrigos beach: intertidal,
1 specimen.
Ecology: Los Abrigos intertidal: in well sorted
medium sands, with 1.33% organic and 5.30%
carbonates. Canary Islands: among photophilic
algae and vermetid tubes (Núñez 1990). AtlanticMediterranean region: in intertidal (Parapar 1991)
and subtidal (López 1995) rocky bottoms, among
photophilic algae and calcareous algae, as well as,
endobiontic of sponges (Koukouras et al. 1985;
Voultsiadou-Koukoura et al. 1987).
Distribution: Cosmopolitan (López & San Martín
1994). Canary Islands: Lanzarote, Fuerteventura,
Gran Canaria, Tenerife, Gomera and El Hierro
(Langerhans 1881; Kirkegaard 1983; Núñez et al.
1984; Núñez 1990; Núñez et al. 1992).
FAMLIY DORVILLEIDAE Chamberlin, 1919
Protodorvillea Pettibone, 1961
Protodorvillea kefersteini (McIntosh, 1869)
Protodorvillea kefersteini: Hartmann-Schröder
1996: 276, fig. 123.
Material examined: Los Abrigos beach: subtidal,
1 specimen.
Ecology: Los Abrigos subtidal: in well sorted
medium sands, with 1.68% organic matter and
3.08% carbonates. Canary Islands: in sandy bare
bottoms and Cymodocea nodosa bottoms (Núñez
et al. 1996). Atlantic-Mediterranean region: in
soft-bottoms with diverse granulometry (Bellan
1964; Moreira 1999).
Distribution: Amphiatlantic (Fauvel 1923;
Perkins 1979). Mediterranean (Campoy 1982).
East Pacific (Orensanz 1973). Canary Islands:
Lanzarote and Tenerife (García-Valdecasas 1985;
García-Valdecasas et al. 1986; Núñez et al. 1996).
DISCUSSION
In this work a total of 47 interstitial polychaete
species were collected from the intertidal and
shallow subtidal (3 m depth) of two stations on
the south coast of Tenerife. The family Syllidae
was the most diverse with 18 species, of which 8
belong to the subfamily Exogoninae, 6 to the
Eusyllinae and 4 to the Syllinae. The second most
diverse family was Spionidae with 6 species,
followed by Paraonidae with 5 species.
In terms of specimens, the most abundant
species was the hesionid Microphthalmus pseudoaberrans, which clearly dominated the intertidal
of Los Abrigos, representing more than 80% of
the total number. The other sampling stations
(Los Abrigos subtidal, and Los Cristianos
intertidal and subtidal) were dominated by
spionids, such as Rhynchospio glutaea and Spio
filicornis. This can be partially explained by the
presence of medium sands in the intertidal of Los
Abrigos, whilst the remaining stations were
characterized by fine sands, except for the
subtidal of Los Abrigos, which presented
alternating fine and medium sands all over the
sampling period (Riera 2004).
The interstitial polychaete fauna of the Canary
Islands consists of typical meiofaunal species,
such as small-sized syllids and macrofaunal
species, like the spionids Spio filicornis and
Scolelepis squamata. The latter species reach
only small sizes in the Canarian archipelago
compared to other biogeographical regions, e.g.
Atlantic-Mediterranean area. The main reason for
this is the presence of oligotrophic waters in the
Canarian archipelago (Barton et al. 1998),
although sporadically influenced by eutrophic
waters from the Saharian upwelling.
This study represents the first qualitative
characterization of the interstitial polychaete
fauna from the Canary Islands. Future research
33
must be conducted in order to determine the
temporal and spatial variations of the interstitial
fauna in the intertidal beaches of the Canarian
archipelago.
ACKNOWLEDGEMENTS
The authors are much indebted to all Spanish
polychaetologists for their exchange of ideas and
continuous encouragement. We thank also our
colleagues of the benthos laboratory, Óscar
Monterroso, Lisandra Núñez and Alejandro
Martínez, for taxonomical and ecological
discussions during the last years. To an
anonymous referee who greatly improved a first
version of the manuscript.
REFERENCES
Aguirrezabálaga, F. 1984. Contribución al estudio de
los Anélidos Poliquetos de la costa de Guipúzcoa.
Munibe 36: 119-130. [In Spanish]
Alós, C. 1988. Anélidos Poliquetos del Cabo de Creus
(Alt Empordá). PhD-thesis. University of
Barcelona, 838 pp. [In Spanish]
Alós, C., A. Campoy & F. Pereira 1982. Contribución
al estudio de los Anélidos Poliquetos endobiontes
de esponjas. Actas II Simposio Ibérico del Estudio
del Bentos Marino, 3: 139-157. [In Spanish]
Alós, C. & F. Pereira 1984. Estudio de la poblaciones
de Anélidos Poliquetos de Posidonia oceanica.
Posidonia Newsletter 2(1): 5-16. [In Spanish]
Amoreux, L. 1971. Annélides Polychètes captures au
large de la cote D´Arcachon. Inventarie
taxonomique. Bulletin de la Société Linnéenne de
Bordeaux 1(7): 147-164. [In French]
Armonies, W. & K. Reise 2000. Faunal diversity across
a sandy shore. Marine Ecology Progress Series
196: 49-57.
Badalamenti, F. & A. Castelli 1991. Schoederella
laubieri, a new species of the subfamily
Protoariicinae (Polychaeta, Orbiniidae), with some
notes on the genus Schoerederella Laubier, 1962.
Bolletino di Zoologia 61: 217-227.
Banse, K. 1971. A new species, and additions to the
descriptions of six other species of Syllides Örsted
(Syllidae: Polychaeta). Journal of the Fisheries
Research Board of Canada 28(1): 1469-1481.
Baratech, L. 1985. Contribución al conocimiento de los
Anélidos Poliquetos de las costas andaluzas. MSc-
34
thesis, University Autónoma of Madrid, 306 pp. [In
Spanish]
Baratech, L & G. San Martín 1987. Contribución al
conocimiento de los Anélidos Poliquetos
(Annelida: Polychaeta) de las costas andaluzas.
Boletín del Instituto Español de Oceanografía 4(2):
37-48. [In Spanish]
Barton, E.D., J. Arístegui, P. Tett, M. Cantón, J.
García-Braun, S. Hernández-León, et al. 1998. The
transition zone of the Canary Current upwelling
region. Progress in Oceanography 41: 455-504.
Bellan, G. 1964. Contribution à l'étude systématique,
bionimique et écologique des annélides Polychètes
de la Méditerranée. PhD-thesis, University of
d'Aix-Marseille, 371 pp. [In French]
Bellan, G. 1969. Annélides polychètes recueillies dans
l'archipel de Madère au cours de la campagne
scientifique du navire océanographique "Jean
Charcot" (juillet 1966). Cahiers de Biologie Marie
10(1): 35-57. [In French]
Bellan, G. 1971. Étude sommaire des Polychétes des
horizons supérieurs de l´étage infralittoral sur
substrat dur dans la région provencale. Rapport
Proc. Verb, Commission Internationale pou
l’Exploration Scientifique de la mer Méditerranée
(CIESM) 20: 217-219. [In French]
Bellan, G. 1980. Annélides Polychaétes des substrats
solides de trois milieux pollués sur les côtes de
Provence (France): Cortiou, Golfe de Fos, Vieux
Port de Marseille. Tethys 9(3): 267-278. [In
French]
Ben-Eliahu, M.N. 1976. Polychaete crytofauna from
Rims of similar intertidal vermetid reefs on the
mediterranean coast of Israel and in the Gulf of
Elat. 3 Serpulidae (Polychaeta Sedentaria). Israel
Journal of Zoology 25: 103-119.
Ben-Eliahu, M.N. 1977. Polychaete crytofauna from
Rims of similar intertidal vermetid reefs on the
mediterranean coast of Israel and in the Gulf of
Elat. 6. Syllinae and Eusyllinae (Syllidae:
Polychaeta Errantia). Israel Journal of Zoology 26:
1-58.
Berkeley, E. & C. Berkeley 1948. Canadian Pacific
Fauna. 9 Annelida. 9b (1). Polychaeta Errantia.
Journal of Fisheries Research Board of Canada,
100 pp.
Besteiro, C. 1986. Fauna mesopsámmica de las arenas
de Amphioxus de la ría de Ferrol (Galicia). PhDThesis, University of Santiago de Compostela, 260
pp. [In Spanish]
Besteiro, C., V. Urgorri & J. Parapar 1987.
Aportaciones nuevas para la fauna ibérica:
Anélidos Poliquetos. Cahiers de Biologie Marine
28: 491-504.
Blake, J.A. 1983. Polychaetes of the family Spionidae
from South America, Antarctica and adjacent seas
and islands. Antartic Research Series 39: 205-287.
Blake, A., B. Hilbig & P.H. Scott 1996. Taxonomic
atlas of the benthic fauna of the Santa Maria Basin
and Western Santa Barbara Channel. Vol 6: The
Annelida. Part 3. Polychaeta: Orbiniidae to
Cossuridae. Santa Barbara Museum of Natural
History, Santa Barbara, California, 418 pp.
Brito, M.C. & J. Núñez 2002. A new genus and species
of Questidae (Annelida: Polychaeta) from the
Central Macaronesian Region and a cladistic
analysis of the family. Sarsia 87: 281-289.
Brito, M.C., J. Núñez & G. San Martín 2000.
Parapionosyllis macaronesiensis, a new species of
Exogoninae (Polychaeta: Syllidae) from the
Macaronesian Region. Proceedings of the Biological Society of Washington 113(4): 1147-1150.
Brito, M.C., J. Núñez & G. San Martín 2000. The
genus Streptosyllis Webster & Benedict, 1884
(Polychaeta: Syllidae: Eusyllinae) from the Canary
Islands, with description of a new species. Bulletin
of Marine Science 67(1): 603-615.
Brito, M.C. 2002. Estudio de las comunidades
intersticiales del sebadal (Cymodocea nodosa) en
Canarias, con especial referencia a los anélidos
poliquetos. PhD-thesis, University of La Laguna,
618 pp. [In Spanish]
Campoy, A. 1981. Anélidos Poliquetos de la Península
Ibérica: family Paraonidae. Boletín de la Real
Sociedad Española de Historia Natural (Biología):
79: 15-26. [In Spanish]
Campoy, A. 1982. Fauna de España. Fauna de
Anélidos Poliquetos de la Península Ibérica. PhDThesis, University of Navarra, 781 pp. [In Spanish]
Campoy, A. & Viéitez, J.M. 1982. Microphthalmus
pseudoaberrans
n.
sp.
(Polychaeta,
Microphthalminae) de las rías bajas gallegas. Actas
II Simposio Ibérico del Estudio del Bentos Marino,
III: 159-164. [In Spanish]
Capaccioni, R. 1983. Anélidos Poliquetos del Mar
Menor: Faunística y Ecología. MSc-thesis,
University of Valencia, 205 pp. [In Spanish]
Capaccioni, R. 1987. Anélidos Poliquetos de la
Ensenada de los Alfaques (Delta del Ebro,
Mediterráneo Occidental). PhD-thesis, University
of Valencia, 533 pp. [In Spanish]
Carrasco, F.D. 1974. Spionidae (Polychaeta)
provenientes de la Bahía de La Concepción y
lugares adyacentes. Boletín de la Sociedad de
Biología de Concepción 48: 185-201.
Castelli A. 1985. Paraonidae (Annelida, Polychaeta)
des fonds meubles infralittoraux des cotes toscanes.
Cahiers de Biologie Marine 26: 267-279. [In
French]
Clapárede, E. 1868. Les annélides chétopodes du Golfo
de Naples. Secunde Partie. Mémoires de la Société
de Physique et d'Histoire Naturelle de Genève
20(1): 1-225. [In French]
Cognetti, G. 1957. I Sillidi del Golfo de Napoli.
Publicazione Stazione Zoologica di Napoli 30(1):
1-100. [In Italian]
Cognetti-Varriale, A.M. 1972. Biologia delle secche
della Meloria. IV. Ricerche preliminari sulla
distribuzione dei Policheti. Bollettino di Pesca
Piscicultura e Idrobiologia 27(2): 263-283. [In
Italian]
Dauvin, J.C. 1989. Sur la présence de Spio decoratus
Bobretzky, 1871 en Manche et remarques sur Spio
martinensis Mesnil, 1896 et Spio filicornis (O.F.
Müller, 1776). Cahiers de Biologie Marine 30:
167-180. [In French]
Day, J.H. 1961. The polychaete fauna of South Africa,
Part 6. Sedentary species dregded off Cape coasts
with a few new records from the shore. Journal of
the Linnean Society of London (Zoology) 44: 463560.
Day, J.H. 1967. A monograph on the Polychaeta of
Southern Africa. Part I: Errantia. Trustees British
Museum of Natural History, London, 458 pp.
Day, J.H. 1973. Polychaeta collected by U.D. Gairward
at Ratnagiri, south of Bombay. Proceedings of the
Biological Society of Washington 6: 103-106.
Desbrúyeres, D., J.M. Guille & J.M. Ramos 1972.
Bionomie benthique du plateau continental de la
côte catalane espagnole. Vie et Milieu 23 (3B):
335-366. [In French]
Fauvel, P. 1914. Annélides Polychétes non-pélagiques
provenant des Campagnes de l´Hirondelle et de la
Princesse Alice (1885-1910). Result des
campagnes scientiflque du Prince Albert ler de
Monaco 48: 1-151. [In French]
Fauvel, P. 1923. Faune de France. 4: Polychètes
errantes. Le Chevalier (ed.). París, 488 pp. [In
French]
Fauvel, P. 1927. Faune de France. 16: Polychétes
Sedentaires. Le Chevalier (ed.). París, 494 pp. [In
French]
Foster, N.M. 1971. Spionidae (Polychaeta) of the Gulf
of Mexico and the Caribbean sea. Studies on the
fauna of Curaçao and other Caribbean Islands 36
(129): 1-183.
García-Valdecasas, A. 1985. Estudio faunístico de la
cueva. submarina “Túnel de la Atlántida”, Jameos
del Agua, Lanzarote. Naturalia Hispanic, 27: 3–56.
[In Spanish]
García-Valdecasas, A., J.L. Fernández & J. Bedolla
1986. The fauna of the submarine cave “Túnel de
la Atlántida”, Lanzarote. Actas IX Congreso
Internacional de Espeleología, Barcelona. 2:
145-147.
35
Gardiner, S.L. 1976. Errant Polychaete Annelids from
North Carolina. Journal of the Elisha Mitchell
Scientific Society 91(3): 77-220.
Gaston, J.M. 1984. Family Paraonidae. Pp. 2-53 in:
Uebelacker and Johnson (Eds). Taxonomic guide to
the polychaetes of northern Gulf of Mexico. U.S.
Department of the Interior, Mineral Management
Services, Barry Vittor and Associates, Mobile,
Alabama, vol. 1, Chpt. 30.
Giangrande, A. 1985. Policheti dei rizomi di P.
oceanica L. Delile di una prateria dell´isola di
Ischia (Napoli). Atti della Società Toscana di
Scienze Naturali Memorie, Serie B, 92: 195-206.
[In Italian]
Giangrande, A. & C. Gambi 1986. Polychétes d´une
pelouse a Cymodocea nodosa (Ucria) Ascher du
Golfe de Salermo (Mer Tyrrhénienne). Vie et
Milieu 36(3): 185-190. [In Italian]
Giere, O., B. Ebbe & C. Erséus. 2008. Questa
(Annelida, Polychaeta, Orbiniidae) from Pacific
regions – new species and reassessment of the
genus Periquesta. Organisms Diversity and
Evolution 7: 304-319.
Glémarec, M. 1966. Paraonidae de Bretagne,
description de Paradoneis armata nov. sp. Vie et
Milieu 17(2a): 1077-1085. [In French]
Harmelin, J.G. 1964. Étude de l´endofaune des mattes
d´herbiers de Posidonia oceanica Delile. Recueil
des Travaux de la Station Marine d’Endou, 35(51):
43-105. [In French]
Harmelin, J.G. 1969. Contribution a l´étude de
l´endofauna des prairies d´Halophila stipulacea de
Méditerranée orientales. I. Annélides Polychaétes.
Recueil des Travaux de la Station Marine
d’Endoume 45(61): 305-316.
Hartman, O. 1951. The littoral marine annelids of the
Gulf of Mexico. Publications of the Institute of
Marine Science 2(1): 7-124.
Hartman, O. 1953. Non-pelagic Polychaeta of the
Swedish Antartic Expedition, 1901-1903. Further
Zoological Research of the Swedish Antarctic
Expedition 4(2): 1-83.
Hartman, O. 1963. Submarine canyons of Southern
California. Part III. Systematics Polychaetes.
University of Southern California. Los Angeles, 93
pp.
Hartman, O. 1965. Catalogue of the Polychaetous
Annelids of the World. Supplement 1960-65 and
index. Allan Hancock Foundation, University of
Southern California. Los Angeles, no.23.
Hartman, O. 1969. Atlas of Sedentariate Polychaetous
Annelids from California. Allan Hancock
Foundation, University of Southern California. Los
Angeles, 812 pp.
Hartmann-Schröder, G. 1959. Zur Öcologie der
36
Polychaeten des Mangrove-Estero-Gebietes von El
Salvador. Beiträge Neotropica Fauna 1: 70-183.
[In German]
Hartmann-Schröder, G. 1960. Polychaeten aus dem
Roten Meer. Kieler Meeresforschungen 16(1): 69125. [In German]
Hartmann-Schröder, G. 1971. Annelida, Borstenwürmer, Polychaeta. Die Tierwelt Deutschland,
Gustav Fischer, Verlag. Jena, 594 pp. [In German]
Hartmann-Schröder, G. 1979. Die Polychaeten der
“Atlantichen Kuppenfahrt” von F.S. “Meteor”
(Fahrt 9c, 1967). 1 Proben aus Schlepperäter.
Meteor Forschungsergebnisse Ergebnisse 31: 6390. [In German]
Hartmann-Schröder, G. 1988. Stygofauna of the
Canary Islands, 13 die Polychaeten der sammerl
reisen 1985 und 1987. Bulletin of Zoology of the
Museum of the University of Amsterdam 22: 177184. [In German]
Hartmann-Schröder,
G.
1996.
Annelida,
Borstenwürmer, Polychaeta. In: Die Tierwelt
Deutschlands, vol. 58. Gustav Fischer Verlag. Jena,
648 pp. [In German]
Herrando-Pérez, S., G. San Martín & J. Núñez 2001.
Polychaete patterns from an oceanic island in the
eastern Central Atlantic: La Gomera (Canary
archipelago). Cahiers de Biologie Marine 42: 275287.
Hobson, K.D. & K. Banse 1981. Sedentariate and
Archiannelid Polychaetes of British Columbia and
Washington. Canadian Bulletin of Fisheries and
Aquatic Sciences 209: 1-144.
Holthe, T. 1977. The Polychaetous Annelids of
Trondheinsfjord, Norway. Gunneria 29: 1-64.
Ibáñez, M. 1973. Contribución al estudio ecológico de
los Anélidos Poliquetos de la Península Ibérica.
PhD-Thesis. University Complutense of Madrid,
125 pp. [In Spanish]
Ibáñez, M. & J.M. Viéitez 1973. Presencia en la costa
cantábrica del Anélido Poliqueto Dispio uncinata:
contribución a su estudio sistemático y ecológico.
Boletín de la Real Sociedad Española de Historia
Natural 71: 5-8.
Imajima, M. 1972. Review of the Annelid Worms of
the Family Nereididae of Japan, with descriptions
of five new species or subspecies. Bulletin of the
Natural Science Museum of Tokyo, Series A, 15(1):
37-53.
Imajima, M. 1990. Spionidae (Annelida, Polychaeta)
from Japan V. The genera Streblospio and Dispio.
Bulletin of the Natural Science Museum of Tokyo,
Series A, 16(4): 155-163.
Imajima, M. 1991. Spionidae (Annelida, Polychaeta)
from Japan VI. The genera Malacoceros and
Rhynchospio. Bulletin of the Natural Science
Museum of Tokyo, Series A, 17(1): 5-17.
Intes, A. & P. Le Loeuff 1977. Les Annélides
Polychétes de Côte d´Ivore. II. Polychétes
Sedentaires. Compte rendu systématique. Cahiers
ORSTOM Séries Oceanográphiques 15(3): 215249.
Junoy, J. 1988. Estudio de la macrofauna intermareal
de sustrato blando de la ría de Foz (Lugo). PhDthesis, University of Alcalá de Henares, 619 pp. [In
Spanish]
Junoy, J. & J.M. Viéitez 1990. Manayunkia aestuarina
(Bourne, 1883) (Polychaeta, Sabellidae) en la ría de
Foz (Galicia, España). Boletín de la Real Sociedad
Española de Historia Natural, 86(1/4): 83-94. [In
Spanish]
Katzmann, W. & L. Laubier 1975. Paraonidae
(Polychétes Sédentaires) de l´Adriatique. Annale
Naturhistorische Museum von Wien 79: 567-588.
Kirkegaard, J.B. 1959. The Polychaete of West Africa.
Part 1. Sedentary species. Atlantide Report 5: 7117.
Kirkegaard, J.B. 1983. The Polychaete of West Africa.
Part. II. Errant species 1. Aphroditidae to
Nereididae. Atlantide Report 14: 7-89.
Koukouras, A., E. Voutsiadou-Koukoura, H.
Chintiroglou & C. Dounas 1985. Benthic bionomy
of north Aegean Sea. III. A comparison of the
macrobenthic assemblages associated with seven
sponge species. Cahiers de Biologie Marine 26:
301-319.
Langerhans, P. 1879. Die Wurmfauna von Madeira.
Zeitschrift für Wissenschaftliche Zoologie 32: 513592. [In German]
Langerhans, P. 1880. Die Wurmfauna Madeiras. II.
Zeitschrift für Wissenschaft Zoologie 33: 267-316.
[In German]
Langerhans, P. 1881. Ueber einige canarische
Anneliden. Nova Acta Leopoldina 42: 93-124. [In
German]
Langerhans, P. 1884. Die Wurmfauna von Madeira. IV.
Zeitchrift für Wissenschaftliche Zoologie 40: 247285. [In German]
Laubier, L. 1966. Le coralligène des Albères.
Monographie biocénotique. Annales de l’Institut
Océanografique 43: 137-316. [In French]
Laubier, L. 1967. Sur quelques Aricidea (Polychétes,
Paraonidae) de Banyuls-Sur-Mer. Vie et Milieu 18:
99-132. [In French]
Laubier, L. & J.M. Ramos 1973. Paraonidae
(Polychétes sédentaires de Mediterranée). Bulletin
of the British Museum (Natural History) 113(168):
1109-1148.
Light, W.J. 1978. Spionidae Polychaeta Annelida.
Pacific Grove. Boxwood Press, 211 pp.
López, E. 1995. Anélidos Poliquetos de sustratos duros
de las Islas Chafarinas. PhD-Thesis, University
Autónoma of Madrid, 672 pp. [In Spanish]
López, E. & G. San Martín 1994. Syllidae (Polychaeta)
recolectados en las islas de Cabo Verde por la “I
Expedición Ibérica”. Revista de Biología Tropical
42(1/2): 127-137. [In Spanish]
López-Jamar, E. 1978. Macrobentos infaunal de la ría
de Pontevedra. Boletín del Instituto Español de
Oceanografía 4(264): 113-130. [In Spanish]
López-Jamar, E. 1981. Spatial distribution of the
infaunal benthic communities of the Ría de Muros,
NW Spain. Marine Biology 63: 29-37.
Maciolek, N. 1983. Systematics of Atlantic Spionidae
(Annelida: Polychaeta) with special reference to
deep-water species. PhD-thesis, University of
Boston, 400 pp.
Maciolek, N. 1987. New species and records of
Scolelepis (Polychaeta: Spionidae) from the east
coast of North America, with a review of the
subgenera. Proceedings of the Biological Society of
Washington 7: 16-40.
Maciolek, N.J. 1990. A redescription of some species
belonging to the genera Spio and Microspio
(Polychaeta: Annelida) and descriptions of three
new species from the northwestern Atlantic ocean.
Journal of Natural History 24:1109-1141
Martín, D. 1986. Anélidos Poliquetos y Moluscos
asociados a algas calcáreas. MSc-thesis,
University of Barcelona, 261 pp. [In Spanish]
Martín, D. 1991. Macroinfauna de una bahía
mediterránea: Estudio de los niveles de
organización de las poblaciones de Anélidos
Poliquetos. PhD-thesis, University of Barcelona,
456 pp. [In Spanish]
Martín, D., T.A. Britayev, G. San Martín & J. Gil 2003.
Inter-population
variability
and
character
description in the sponge-associated Haplosyllis
spongicola complex (Polychaeta: Syllidae).
Hydrobiologia 496: 145-162.
May, W. 1912. Gomera die Wandisel der Kanaren. Ver
Karlsruhe 24: 1-215. [In German]
McLelland, J.A. & G.R. Gaston 1994. Two new
species of Cirrophorus (Polychaeta: Paraonidae)
from northern Gulf of Mexico. Proceedings of the
Biological Society of Washington 107(3): 524-531.
Montiel, A., B. Hilbig & N. Rozbaczylo 2002. New
records to Chile of the Family Paraonidae
(Annelida: Polychaeta). Helgoland Marine
Research 56: 134-139.
Moreira, J. 1999. Anélidos Poliquetos de sustratos
blandos infralitorales de la Ensenada de Baiona
(Galicia). MSc-thesis, University of Vigo, 510 pp.
[In Spanish]
von Nordheim, H. 1989. Six new species of Protodrilus
(Annelida, Polychaeta) from Europe and New
37
Zealand, with a concise presentation of the genus.
Zoologica Scripta, 18(2): 245-268.
Núñez, J. 1990. Anélidos Poliquetos de Canarias:
estudio sistemático de los órdenes Phyllodocida,
Amphinomida y Eunicida. PhD-thesis, University
of La Laguna, 610 pp. [In Spanish]
Núñez, J. 1993. Redescription of Perinereis taorica
Langerhans (Polychaeta: Nereidae) from Tenerife
(Canary Islands). Marine Life 3: 31-35.
Núñez, J., J.J. Bacallado & M.C. Brito 1981. Nereidae
(Polychaeta Errantia) de las costas del archipiélago
canario. Boletín del Instituto Español de
Oceanografía 4: 162-177. [In Spanish]
Núñez, J., M.C. Brito & J.J. Bacallado 1984. Catálogo
provisional de los Anélidos Poliquetos del
archipiélago canario. Cuadernos Marisqueros
Publicaciones Técnicas 7: 113-148. [In Spanish]
Núñez, J., O. Ocaña & M.C. Brito 1997. Two new
species (Polychaeta: Fauveliopsidae and Nerillidae)
and other polychaetes from the marine lagoon cave
of Jameos del Agua, Lanzarote (Canary Islands).
Bulletin of Marine Science 60(2): 252-260.
Núñez, J., M. Pascual, J.D. Delgado & G. San Martín
1995. Interstitial polychaetes from Madeira with a
description of Syllides bansei Perkins, 1981.
Bocagiana 179: 1-7.
Núñez, J., M. Pascual & M.C. Brito 1996. Contribución
al estudio de los Dorvilleidos (Annelida:
Polychaeta) de Canarias. Revista de la Academia
Canaria de las Ciencias 8(2/3/4): 139-151. [In
Spanish]
Núñez, J., R. Riera, M.C. Brito & M. Pascual 2001.
Anélidos Poliquetos intersticiales recolectados en
las Islas Salvajes. Vieraea 29: 29-46. [In Spanish]
Núñez, J & G. San Martín 1996. Anélidos Poliquetos
de las Canary Islands. Family Syllidae. I.
Subfamilies Eusyllinae y Autolytinae. I Congreso
de Oceanografía y Recursos Marinos en el
Atlántico Centro-Oriental (ICCM): 197-217. [In
Spanish]
Núñez, J., G. San Martín & M.C. Brito 1992.
Exogoninae (Polychaeta: Syllidae) from the Canary
Islands. Scientia Marina 56(1): 43-52.
Orensanz, J.M. 1973. Los Anélidos Poliquetos de la
provincia biogeográfica de Argentina. IV.
Lumbrineridae. Physis, Sec A. 32(85): 343-393.
Parapar, J. 1991. Anélidos Poliquetos bentónicos de la
ría de Ferrol (Galicia). PhD-thesis, University of
Santiago de Compostela, 1.104 pp. [In Spanish]
Parapar, J., C. Besteiro & V. Urgorri 1996. Inventario
dos Poliquetos de Galicia (Annelida: Polychaeta).
Cuadernos da Área de Ciencias Biolóxicas.
Inventarios. 16: 1-178. [In Spanish]
Pascual, P. 1996. Poliquetos endobiontes de esponjas
de Tenerife y Madeira. MsC-thesis, University of
38
La Laguna, 278 pp. [In Spanish]
Pascual, M., J. Núñez & G. San Martín 1996. Exogone
(Polychaeta: Syllidae: Exogoninae) endobiontics of
sponges from the Canary Islands and Madeira
island with description of two new species. Ophelia
45(1): 67-80.
Pérès, J.M. 1954. Contribution à l´èude des Annélides
Polychétes de la Méditerranée Occidental. Recueil
des Travaux de la Station Marine d’Endoume
8(13): 83-155. [In French]
Perkins, T. 1979. Lumbrineridae, Arabellidae and
Dorvilleidae (Polychaeta) principally from Florida,
with descriptions of siw new species. Proceedings
of the Biological Society of Washington 92(3): 415465.
Perkins, T. 1981. Syllidae (Polychaeta), principally
from Florida, with descriptions of a new genus and
twenty-one new species. Proceedings of the
Biological Society of Washington 4: 1080-1172.
Pettibone, M.H. 1963. Marine Polychaeta Worms of
the New England Region. Part I: Families
Aphroditidae through Trochochaetidae. Bulletin of
the United States National Museum 227: 1-356.
Riera, R. 2004. Comunidad meiofaunal de las playas
de Los Abrigos del Porís y de Los Cristianos en la
isla de Tenerife. Estructura y dinámica de sus
comunidades. PhD-thesis, University of La
Laguna, 486 pp. [In Spanish]
Riera, R., J. Núñez & M.C. Brito 2006 Parapionosyllis
(Polychaeta: Syllidae: Exogoninae) from Tenerife
(Canary Islands, Spain) with description of a new
species and new records. Zootaxa 70S3: 59-64.
Riera, R., J. Núñez & M.C. Brito 2007. A new species
of the interstitial genus Neopetitia (Polychaeta,
Syllidae, Eusyllinae) from Tenerife, with modified
acicular chaetae in males. Helgoland Marine
Research 61(3): 221-223.
Rioja, E. 1917. Datos para el conocimiento de la Fauna
de Anélidos Poliquetos del Cantábrico. Trabajos
del Museo Nacional de Ciencias Naturales
Zoológicas 29: 1-111. [In Spanish]
Rioja, E. 1931. Estudio de los poliquetos de la
Península Ibérica. Memorias de la Academia de
Ciencias Naturales 2: 1-471. [In Spanish]
Rouse, G.W. &. K. Fauchald 1997. Cladistic and the
polychaetes. Zoologica Scripta 26: 139-204.
Rullier, F. 1972. Expedition francaise sur les récif
coralliens de la Nouvelle-Calédonie. Fundation
Singer-Polinac (Ed.). Paris, 168 pp. [In French]
Rullier, F. & L. Amoreux 1970. Nouvelle contribution
a l´étude de la faune des Annélides Polychétes du
Maroc. Bulletin de la Société de Sciences
Naturelles et Physiques du Maroc 49: 109-142. [In
French]
San Martín, G. 1984. Estudio biogeográfico, faunístico
y sistemático de los Poliquetos de la familia Sílidos
(Polychaeta: Syllidae) en Baleares. PhD-thesis,
Universidad Complutense of Madrid, 529 pp. [In
Spanish]
San Martín, G. 1990. Eusyllinae (Syllidae: Polychaeta)
from Cuba and Gulf of Mexico. Bulletin of Marine
Science 46(3): 590-619.
San Martín, G. 1991. Sphaerosyllis and Parapionosyllis
(Polychaeta: Syllidae) from Cuba and Florida.
Ophelia supplement 5: 321-328.
San Martín, G. 2003. Annelida, Polychaeta II: Syllidae.
In: Ramos, M.A et al. (Eds), Fauna Ibérica, vol. 21.
537 pp.
San Martín, G. & O. Aguirre 1991. Ciclo anual de los
Poliquetos asociados al alga Mesophyllum
lichenoides (Ellis) en una playa mediterránea.
Boletín del Instituto Español de Oceanografía 7(1):
157-170. [In Spanish]
San Martín, G. & R. Alvarado 1982. Notas sobre los
Poliquetos de la isla de Cabrera (Baleares). Boletín
de la Real Sociedad Española de Historia Natural
79: 221-234. [In Spanish]
San Martín, G., G. González & E. López-Jamar 1985.
Aspectos sistemáticos y ecológicos sobre algunas
especies de Sílidos (Polychaeta: Syllidae) de las
costas gallegas. Boletín del Instituto Español de
Oceanografía 2(2): 27-36. [In Spanish]
San Martín, G. & J.M. Viéitez 1984. Anélidos
Poliquetos de los rizomas de Posidonia oceanica
en las costas del Cabo de Palos (Murcia, España).
International Workshop of Posidonia oceanica
beds, 149-157 pp. [In Spanish]
San Martín, G., J.M. Viéitez & A. Campoy 1981.
Contribución al estudio de la fauna de Anélidos
Poliquetos de las costas españolas: Poliquetos
Errantes recoletados en la bahía de Palma de
Mallorca. Boletín del Instituto Español de
Oceanografía 6: 63-87. [In Spanish]
Sardá, R. 1982. Anélidos Poliquetos de la Bahía de
Algeciras. Nota sobre una colección próxima a la
desembocadura del Rio Guadarranque. Oecologia
acuática (Supplement III) 165-170. [In Spanish]
Sardá, R. 1984. Contribución al conocimiento de los
Anélidos Poliquetos litorales ibéricos. Estudio
sobre la fauna de Anélidos Poliquetos de las zonas
mediolitoral e infralitoral, en la región del
Estrecho de Gibraltar. PhD-thesis, University of
Barcelona, 901 pp. [In Spanish]
Sardá, R. 1985. Anélidos Poliquetos del estrecho de
Gibraltar. I. Amphinomida, Sphinterida y
Phyllodocida. Miscellania Zoologica 9: 65-78.
Schlenz, R. 1965. Données sur les peuplements des
sédiments a petites phanérogames marines (Zostera
nana Roth et Cymodocea nodosa Ascherson)
comparés a ceux des habitats voisins dépourvus de
végétation. Recueil des Travaux de la Station
Marine d’Endoume 39: 96-125. [In French]
Schmidt, H. & W. Westheide 2000. Are the
polychaetes Hesionides arenaria and Stygocapitella subterranea true cosmopolitan species?results of RAPD-PCR investigations. Zoologica
Scripta 29(1): 17-27.
Strelzov, V.E. 1968. Polychaetous annelids of the
family Paraonidae (Polychaeta, Sedentaria) of the
Barents Sea. Murmanskii Morskoi Biologicheskii
Institut 17(21): 1- 123. [In Russian]
Strelzov, V.E. 1979. Polychaete worms of the family
Paraonidae Cerruti, 1909 (Polychaeta, Sedentaria).
Oxonian Press. New Delhi, 212 pp.
Talavera, J.A., J. Núñez &. M.C. Brito 1984. Estudio
de las poblaciones de anélidos presentes en una
charca supralitoral de la Isla de Tenerife
(Canarias). Actas del IV Simposio Ibérico de
Estudios del Bentos Marino 3: 165-176. [In
Spanish]
Tena, J. 1992. Anélidos Poliquetos del antepuerto de
Valencia: Ecología y aspectos tróficos. MsC-thesis,
University of Valencia, 270 pp. [In Spanish]
Tena, J. 1996. Faunística y Ecología de los Anélidos
Poliquetos de los fondos de sustrato duro del
archipiélago de las Chafarinas (SW Mar de
Alborán). Ph-D thesis, University of Valencia, 549
pp. [In Spanish]
Uebelacker, J.M. 1984. Family Syllidae. Pp. 30-151 in:
Uebelacker, J.M. & P.G. Johnson (Eds).
Taxonomic guide to the polychaetes of the nothern
Gulf of Mexico. U.S. Department of the Interior
Minerals Management Services, Barry A. Vittor &
associates. Alabama, vol 4, chapter 30.
Viéitez, J.M. 1976. Ecología de Poliquetos y Moluscos
de la playa de Meira (Ría de Vigo). I. Estudio de
las comunidades. Investigaciones Pesqueras 40(1):
223-248. [In Spanish]
Viéitez, J.M. 1978. Comparación ecológica de dos
playas de las rías de Pontevedra y Vigo. Ph-D
thesis, Universidad Complutense of Madrid, 273
pp. [In Spanish]
Voultsiadou-Koukoura, M.E., A. Koukouras & A.
Eleftheriou 1987. Macrofauna associated with the
sponge Verongia aerophoba in the North Aegean
Sea. Estarine and Coastal Shelf Science 24: 265278.
Westheide, W. 1974. Interstitielle Fauna von
Galapagos. XI. Pisionidae, Hesionidae, Pilargidae,
Syllidae (Polychaeta). Mikrofauna Meeresbunden
44: 195-338. [In German]
Wu, B.L., R. Sun & M. Chen 1980. On archiannelids of
the Yellow Sea. 1. Protodrilidae. Acta
Oceanologica Sinica 2: 132-148.
Accepted 8 May 2010.
39
Experimental harvesting of juvenile common octopus Octopus vulgaris, for commercial ongrowing in the Azores CHRISTOPHER K. PHAM & EDUARDO ISIDRO Pham, C.K. & E. Isidro 2010. Experimental harvesting of juvenile common
octopus Octopus vulgaris, for commercial ongrowing in the Azores. Arquipelago.
Life and Marine Sciences 27: 41-47.
Octopus aquaculture is currently restricted to ongrowing of sub-adult to commercial size
because culture of paralarvae remains a bottleneck. In most countries, commercial
ongrowing rely upon existing pot fisheries for octopuses for obtaining their specimens. In
the Azores, such fishery does not exist and effective methods of harvest are required if
farming is to be implemented. In this study, we investigated the potential of obtaining subadult octopuses on the coast of Faial Island, Azores. Two sets of traps (n=30) consisting of
3 PVC tubes within cement blocks were set-up on two different substrates; soft sediment
(Pedro Miguel) and rocky-sand (Pasteleiro) at depth varying between 10 and 30 metres.
From June to August 2006, 11 hauls per site were performed. A total of 191 octopuses
(from 1.1 to 989 g; average = 135.3 g) were captured. Catches in the soft sediment site were
significantly higher than in the other location (CPUE: mean ± SD: 0.33 ± 0.17 vs. 0.15 ±
0.17 octopus trap-1 hour-1*100). The catch was initially dominated by octopus of 300-400 g
but as fishing continued, this size classes disappeared and was replaced by smaller
individuals. As a result, half of the catch at both sites (51.8%) was composed of specimens
with a weight equal or inferior to 50 grams. The occurrence of summer recruitment event
combined with a natural displacement of larger individuals into deeper waters is most
probably responsible for this pattern. Our results showed that in shallow water and during
this period of the year, individuals inferior to 50 grams are far more abundant than larger
octopuses and should be the target size class for ongrowing activities.
Key words: aquaculture, catch, rocky-sediment, soft-sediment, traps, tubes
Christopher K. Pham (e-mail: [email protected]) & Eduardo Isidro, Center of
Institute of Marine Research (IMAR), Departamento de Oceanografia e Pescas,
Universidade dos Açores. PT-9901-862 Horta, Açores, Portugal.
INTRODUCTION
The contribution of aquaculture towards total
seafood production is increasing rapidly,
becoming a significant economic activity for
many nations (Soto 2008). Throughout its wide
distribution, the common octopus, Octopus
vulgaris (Cuvier 1797) is a highly valued protein
source (Hastie et al. 2009). In the Azores, O.
vulgaris is harvested on a small scale (Carreira et
al. 2002) and mariculture could represent locally
an additional source of octopuses. Although
cephalopods are increasingly being exploited,
cephalopod mariculture remains poorly developed
when compared to other molluscs (Boyle &
Rodhouse 2005). For octopods, this can be
attributed principally to the difficulty in rearing
early life stages (Iglesias et al. 2007). Within the
cephalopods, two general developmental strategies exists (reviewed by Boletzky 1981): (i) the
production of a few large telolecithal eggs resulting in large, adult-like benthic hatchlings (e.g.
cuttlefish and in some octopuses) and (ii) the
production of a high number of small, less yolky
eggs generating small planktonic juveniles (e.g.
loliginids, sepiolids and octopuses), termed
41
paralarvae (Young & Harman 1988). Octopus
vulgaris falls into the latter category, going
through a planktonic stage of 20 to 35 days,
depending on water temperature (Villanueva et al.
1995). Despite the high number of attributes for
mariculture (e.g. fast growth, high fecundity, high
food conversion ratio, etc.), large-scale farming
of O. vulgaris has been strongly constrained by
high mortality rates during the paralarval phase
(Iglesias et al. 2007).
At present, the few companies producing
octopuses are relying upon pot fisheries for
obtaining sub-adults, subsequently grown in
cages (Vaz-Pires et al. 2004). Although, in the
Azores, such fisheries could be developed
(Carreira & Gonçalves 2008), octopuses are
currently caught mostly with spears by
snorkelling divers with no by-catch of small
individuals (Gonçalves et al. 2002). As a result,
octopus farming could only be possible if a cheap
and effective method of obtaining small
specimens can be developed. Trap fishing for
octopuses is successfully performed in many
places (e.g. Portugal [Cunha & Moreno 1994];
Spain [Guerra 1981; 1997]; Canary Islands
[Hérnandez-García et al. 1998]; California [Rasmussen 1997]). In the Azores, an experimental
fishery using Japanese baited pots suggested that
octopus of commercial size can be efficiently
harvested and that CPUE is higher at shallowest
depth (Carreira & Gonçalves 2008). The present
study was conducted to evaluate whether small
octopuses (300-400 g) could be harvested for
commercial ongrowing using simple un-baited
traps capable of capturing individuals of various
sizes. We focused on maintaining a regular
fishing pressure within two areas (of different
substrate type) close to the harbour in shallow
depths and assess their potential for octopus
supply.
MATERIAL AND METHODS
The gear designed to harvest juvenile octopuses
consisted of 30 cement traps (Fig. 1), placed 20
metres away from each other and buoyed at each
end. Three PVC tubes of various diameters (70;
45 and 35 mm) were present on each trap and
served as shelter for the octopus.
42
Fig. 1. Trap deployed to catch juvenile common
octopus O. vulgaris, on the coast of Faial, Azores. PVC
pipes of three different diameters were offered as
shelter (70, 45 and 35 mm).
FISHING SITES
Harvesting effort (11 trips) was conducted
between June and August 2006 in two sites on the
coast of Faial, Azores. The first site (Pedro
Miguel), is located on the east side of the island.
At this location, the gear was systematically
deployed on a soft sediment bottom, approximately 200 metres from a rocky shore, at a depth
varying between 6 to 18 metres. The second site
(Pasteleiro) is located on the south side of the
island. Here, the gear was deployed on rocks yet
with occasional small sandy patches (rocky-sand)
at depths varying between 12 and 25 meters.
Across this location, practical problems arose
because traps got caught amongst rocks and
subsequently lost when hauled up (n=13). Traps
were also lost in Pedro Miguel but less frequently
(n=7). Gear soaking time was constrained by
weather conditions and varied from 4 to 17 days.
Soaking time for each trip was the same among
sites.
BIOLOGICAL SAMPLING
Each trap was brought onboard and whenever an
octopus was present, the position of the tube and
trap number were recorded. Each octopus was
placed into separate closed containers lowered
into a tank previously filled with seawater. On
arrival to the laboratory, each octopus was then
anaesthetised with a solution of MgCl (7.5%)
diluted in seawater (1:1 ratio) (Messenger et al.
1985). This method was successful since it never
resulted in mortality and allowed good manipulation of the animals. Each animal was weighed
(total weight, TW, 0.1 g) and its mantle measured
(dorsal mantle length, DML, 0.1 mm). The sex
was determined in each specimen through the
examination of the third right arm, which is
shorter in males, with a round suckerless tip
(hectocotylus), and usually presents a number of
enlarged suckers when the octopus is fully
mature. Growth rates were subsequently
monitored in the laboratory using the same
anaesthetic (Pham & Isidro 2009).
(Pasteleiro) and the mean CPUE in soft sediment
was significantly different from the rocky-sand
(t = 2.36; p<0.05). In the rocky-sand location, the
CPUE was far more variable, when compared to
the soft sediments.
The weight frequency of the octopuses caught
at both sites is presented Figure 2. In total, half of
the octopuses caught weighted less than 50 g. Sex
ratio was not significantly different from 1:1 for
both areas (P>0.05).
a) Soft sediment (Pedro Miguel)
DATA ANALYSIS
For all fishing trips, the following indices were
calculated:
1. Catch rate (defined as the number of octopus
caught per number of pots):
(no. of octopuses / no. of traps) x 100
2. The catch per unit effort:
b) Rocky sand (Pasteleiro)
(Total Number of Octopus / n° of traps) x 100
For comparing two samples, a two sample t-test
was performed since all assumptions were
constantly met (normality and homogeneity of
variances). Normality was tested using AndersonDarling normality test whilst homogeneity of
variances was estimated by performing a
Levene’s test (Zar 1996). Sex ratio was analyzed
and differences tested using the Chi-square test.
All statistical analyses were done with Minitab
version 13.0 software (Minitab Inc).
RESULTS
SITE DIFFERENCE AND CATCH COMPOSITION
Over the 3 months period, a total of 191
octopuses (ranging from 1 to 989 g, TW) were
caught. A summary of the catches at both sites is
presented in Table 1. The number of octopuses
caught in soft sediments (Pedro Miguel) was
twice the amount captured at the rocky-sand site
Fig. 2. Weight (g) frequency of common octopus
Octopus vulgaris, caught at two different sites: a.) Soft
sediment (Pedro Miguel) and b.) Rocky-sand
(Pasteleiro).
In Pedro Miguel (soft sediment), there was a
decrease in mean octopus weight with number of
fishing trip (Pearson correlation coefficient =
-0.305; p<0.05). In June, catches were dominated
by octopuses larger than 150 g whereas in July
and August, larger octopuses became scarce and
the bulk of the catch was dominated by specimens
43
Table 1. Summary of the experimental harvesting of the common octopus O. vulgaris, on soft sediment
(Pedro Miguel) and rocky-sand (Pasteleiro), around Faial Island, Azores, during summer 2006: mean
values, standard deviation and min-max values are indicated.
Nº of trips / Total nº octopuses
N° of octopus 30 traps-1
Octopus Total Weight (g)
CPUE (nº octopus trap-1 hour-1*100)
Catch rates (%)
Soft Sediment
11 / 136
12 ± 1
150.1 ± 186.03
0.33 ± 0.17
48.3 ± 15.3
smaller than 150 grams. The situation was rather
different at the other location, probably because
the gear position was regularly changed between
fishing trips. This resulted in little pattern as catch
composition varied enormously.
RELATIONSHIP BETWEEN SOAKING TIME & CATCH
Due to unfavourable weather conditions, gear
soaking time could not remain consistent for each
trip and varied between 4 to 17 days. There was
no significant correlation between CPUE and
soaking time (Pearson correlation coefficient =
-0.205; p>0.05).
TUBE DIAMETER SELECTIVITY
One trap never caught more than one individual.
Table 2 displays the mean, minimum and
maximum weights (TW) of the octopuses caught
by the three different tubes. The largest tube (Ǿ
70 mm) caught the highest number of individuals,
representing approximately 50% of the total
catch. The two other small tubes (35 mm and 45
mm) caught the rest of the octopuses in equal
proportions (25% each tubes). Whilst small tubes
were highly size-selective, most exclusively
catching small octopuses (1.1 to 198 g), the 70
mm tube caught both large and small individuals,
ranging from 7 to 900 g (Fig. 3a).
Min-Max
8 – 19
1.1 – 989
0.1 – 0.67
27.6 – 73
Rocky-Sand
11 / 55
5 ± 4.4
98 ± 124.7
0.15 ± 0.17
20.7 ± 17.3
Min-Max
0 – 14
4.6 – 542.7
0 – 0.43
0 – 51.8
a) 70 mm (n=12)
b) 45 mm (n=45)
c) 35 mm (n=44)
Table 2. Mean total weight (TW, grams), standard
deviations (SD), minimum (min) and maximum (max)
weights and associated number (N) of octopuses caught
in PVC tubes of different diameters (mm).
Tube Ǿ
35
45
70
44
Mean TW
20.54
34.23
232.51
SD
17.5
34.26
188.98
Min-Max
1.1 – 82.9
5 – 198
7.6 – 989
N
44
45
102
Fig. 3. Weight (g) frequency of octopuses
caught in a) 70 mm tube, b) 45 mm tube and
c) 35 mm tube for both sites joined together.
DISCUSSION
The mariculture of O. vulgaris for commercial
purposes has been limited to the ongrowing of
sub-adults (e.g. Chapela et al. 2006; Rodriguez et
al. 2006), implying the need for a reliable method
of obtaining undersize animals directly from the
field. The results of the present study provide
information on what to take into account if such
activity is to be developed in the Azores. The gear
used presented a satisfactory catching efficiency
because the CPUEs obtained were similar to
those previously obtained with Japanese baited
pots for similar depth in the Azores (Carreira &
Gonçalves 2008).
Octopus catches in soft sediment were
significantly higher than in the rocky-sand zone.
In shallow water, O. vulgaris is mostly inhabitant
of coral reefs and rocks (Mangold 1983) but the
large amount of natural dens available in such
areas, makes traps ineffective for catching
octopuses even though their abundance is high. In
contrast, in soft sediments, den availability is a
limiting factor (Katsanevakis & Verriopoulos
2004b) and enrichment experiments using
artificial dens increase octopus abundance
(Katsanevakis & Verriopoulos 2004a). As a
result, trap fishing is more efficient on soft
sediment than on rocky shores. Furthermore, gear
operation over soft sediment is more convenient
since fewer traps get caught up in rocks when
hauled up.
Overall, the size composition in our catch
showed a large dispersion (1.1 to 989 g) but was
predominantly composed of small individuals
(<150g). In fact, the proportion of small
octopuses gradually increased throughout the
summer. Similar to O. vulgaris found elsewhere
(Guerra 1977; Hernández-García et al. 2002;
Katsanevakis and Verriopoulos 2004a), in the
Azores, the species spawns principally in spring
(Gonçalves et al. 2002). After an embryonic
period of 125 to 22 days (at 13 and 25°C, respectively; Mangold 1983) and a paralarval
planktonic stage of 33-40 days (at 25°C), individuals assume a benthic existence and recruit
into the population (Itami et al. 1963; Villanueva
1995). Thus, it is clear that the appearance of
small individuals (10 – 50 g) in our late summer
catches reflects the spring spawning event. This is
in agreement with a previous study showing that
despite recruitment occurring all year round,
small animals are more abundant by the end of
the summer (Gonçalves 1993).
The observed gradual disappearance of larger
octopuses on the other hand, can be attributed to a
natural displacement of larger individuals rather
than a depletion effect of our harvesting activity.
Medium and large octopuses are known to
disappear from shallow waters from July onwards
(Katsanevakis & Verriopoulos 2004b). During
such period, when the thermocline is well pronounced, large octopuses seek cooler areas in
deeper waters, to reduce the energy cost of higher
metabolism whilst smaller octopuses remain in
shallow warmer waters to achieve greater growth
rates and reduce predation risks (Sánchez &
Obarti 1993; Katsanevakis & Verriopoulos
2004b). Although our data suggests that such
phenomenon might happen in the Azores, more
research needs to be conducted as it has important
implications for harvesting and fishing activities.
CONCLUSIONS
The results of this study highlighted important
aspects to consider if ongrowing of wild
octopuses should be implemented in the Azores.
Firstly, specimen harvesting should strictly be
undertaken over a soft sediment type substratum,
preferably in the vicinity of a rocky shore.
Secondly, considering the rapid disappearance of
>150 g individuals and the dominance of smaller
octopuses in the summer, commercial culture
should aim at growing animals smaller than 150
g. Thirdly, soaking time should not exceed four
days as longer time do not increase CPUE.
To be economically viable, a commercial aquaculture would require a much higher number of
octopuses than the amount reported in this study
but also within a shorter time frame. It is worth
performing such experiments during other periods
of the year where natural displacements of
animals can be taken into account. Further work
should not solely concentrate on the ecological
implications of such activity but also on its
economic feasibility.
45
ACKNOWLEDGEMENTS
This study was supported by the project EPA-I
PRODESA 2004.91.001646.0. We would like to
thank Mr. José Santos: captain of “MARFISA”,
Frederic Vandeperre, Marta Monteiro, Rodrigo
Delgado and José Nuno Pereira for their
assistance in the field. We are grateful to Gilberto
Carreira and two anonymous reviewers for
insightful comments that improved the manuscript. Thanks to Emmanuel Arand for preparing
Figure 1.
REFERENCES
Boletzky, S.v. 1981. Reflexions sur les stratégies de
reproduction chez les céphalopodes. Bulletin de la
Société Zoologique de France 106: 293-304.
Boyle, P.R. & P.G. Rodhouse 2005. Cephalopods:
Ecology and Fisheries. Blackwell, Oxford.
Carreira, G.P., J.M. Gonçalves & R.D.M. Nash 2002.
Exploitation of octopus in the Azores (NE
Atlantic): current status and experimental fishery.
Bulletin of Marine Science 71(2): 1116.
Carreira, G.P., & J.M. Gonçalves 2008. Catching
Octopus vulgaris with traps in the Azores: first
trials employing the japanese baited pots in the
Atlantic. Marine Biodiversity Records 2:114.
Chapela, A., A.F. González, E.G. Dawe, J.F. Rocha &
A. Guerra 2006. Growth of common octopus
(Octopus vulgaris) in cages suspended from rafts.
Scientia Marina 70: 121–129.
Cunha, M.M., A. Moreno 1994. Octopus vulgaris: its’
potential on the Portuguese coast. ICES C.M,
1994/k: 33: 19 pp.
Gonçalves, J.M. 1993. Octopus vulgaris Cuvier, 1797
(polvo comun): sinopse da biologia e exploraçâo.
PhD
Thesis.
Universidade
dos
Açores.
Departamento de Oceanografía e Pescas, 470 pp.
[In Portuguese]
Gonçalves, J.M., G.P. Carreira, H.R. Martins & R.D.M.
Nash 2002. Biology of Octopus vulgaris
(Cephalopoda: Octopodida) in the Azores (NE
Atlantic). Bulletin of Marine Science 71(2): 1123.
Guerra, A. 1977. Estudio sobre la biologia y estrutura
de las poblaciones del pulpo común (Octopus
vulgaris) de la costa noroccidental africana. PhD
thesis, Univeristy of Barcelona. [In Spanish]
Guerra, A. 1981. The fishery of Octopus vulgaris off
Finisterre (NW of Spain). ICES C.M., 1981/k, 4.
13 pp.
Guerra, A. 1997. Octopus vulgaris: review of the world
46
fishery. Pp. 91-97, in: Lang, M.A., F.G. Hochberg,
R.A. Ambrose & J.M. Engle (Eds). Proceedings of
the workshop on the fishery and market potential of
octopus in California. Smithsonian Institution's
Office of the Provost - Scientific Diving Program,
Washington.
Hastie, L.C., G.J. Pierce, J. Wang, I. Bruno, A.
Moreno, U. Piatkowski, J-P. Robin 2009.
Cephalopods in the North East Atlantic: Species,
Bio-geography,
Ecology,
Exploitation
and
Conservation. Oceanography and Marine Biology: An
Annual Review 47: 111-190.
Hérnandez-García, V., J.L. Hérnandez-López & J.J.
Castro 1998. The octopus (Octopus vulgaris) in the
small-scale trap fishery off the Canary Islands
(Central-East Atlantic). Fisheries Research 35:
183-189.
Hernández-García, V., J.L. Hernández-López & J.J.
Castro-Hdez 2002. On the reproduction of Octopus
vulgaris off the coast of the Canary Islands.
Fisheries Research. 57:197-203.
Iglesias, J., F.J. Sánchez, J.G.F. Bersano, J.F. Carrasco,
J. Dhont, L. Fuentes, F. Linares et al. 2007.
Rearing of Octopus vulgaris paralarvae: present
status, bottlenecks and trends. Aquaculture 266: 1–
15.
Itami, K., Y. Izawa S. Maeda & K. Nakai 1963. Notes
on the laboratory culture of octopus larvae. Bulletin
of the Japanese Society of Scientific Fisheries. 29:
514-520.
Katsanevakis, S. & G. Verriopoulos 2004a. Den
ecology of Octopus vulgaris Cuvier, 1797, on soft
sediment: availability and types of shelter. Scientia
Marina 68(1): 147-157.
Katsanevakis, S. & G. Verriopoulos 2004b. Abundance
of Octopus vulgaris on soft sediment. Scientia
Marina 68(4):553-560.
Mangold, K. 1983. Octopus vulgaris. Pp. 335-364 in:
P.R. Boyle (Ed.), Cephalopod Life Cycles Volume
I. Species Accounts, Academic Press, New York.
Messenger, J.B., M. Nixon & K.P. Ryan 1985.
Magnesium chloride as an anaesthetic for
cephalopods. Comparative Biochemistry and
Physiology 82C: 203-205.
Pham, C.K. & E. Isidro 2009. Growth and mortality of
common octopus (Octopus vulgaris) fed a
monospecific fish diet. Journal of Shellfish
Research 28(3):617-623.
Rasmussen, A. 1997. Octopus fisherman´s perspective.
Pp. 151-155 in: Lang, M.A., F.G. Hochberg, R.A.
Ambrose, J.M. Engle (Eds). Proceedings of the
workshop on the fishery and market potential of
octopus in California. Smithsonian Institution's
Office of the Provost - Scientific Diving Program,
Washington.
Rodríguez, C., J.F. Carrasco, J.C. Arronte & M.
Rodríguez 2006. Common octopus (Octopus
vulgaris Cuvier, 1797) juvenile ongrowing in
floating cages. Aquaculture 254: 293–300.
Sánchez., P. & R. Obarti 1993. The biology of Octopus
vulgaris caught with clay pots on the Spanish
Mediteranean Coast. Pp. 447-487 in: T Okutamis,
R.K. O’Dor and T. Kudobera (Eds). Recent
Advances in Fisheries Biology. Tokai University
Press, Tokyo.
Soto, D. 2008. Status of world aquaculture and its
future development within an ecosystem’s
perspective. Pp. 3-5 in: Pham, C.K., R.M. Higgins,
M. De Girolamo & E. Isidro (Eds). Abstract
Proceedings of the International Workshop:
Developing a Sustainable Aquaculture Industry in
the Azores. Arquipélago. Life and Marine
Sciences. Supplement 7: xiii + 81 pp.
Vaz-Pires, P., P. Seixas, A. Barbosa 2004. Aquaculture
potential of the common octopus (Octopus vulgaris
Cuvier, 1797): a review. Aquaculture 238: 221238.
Villanueva, R. 1995. Experimental rearing and growth
of planktonic Octopus vulgaris from hatching to
settlement. Canadian Journal of Fisheries and
Aquatic Sciences 52: 2639–2650.
Villanueva, R., C. Nozais & S.v. Boletzky 1995. The
planktonic life of octopuses. Nature 377: 107.
Young, R.E., R.F. Harman 1988. "Larva", "Paralarva",
and "Subadult" in cephalopod terminology.
Malacologia 29(1): 201-207.
Zar, J.H. 1996. Biostatistical analysis. Third edition.
Upper Saddle River, NJ: Prentice-Hall.
Accepted 26 March 2010.
47
Designing a shipboard line transect survey to estimate cetacean abundance off the Azores archipelago CLÁUDIA E.S. FAUSTINO, M.A. SILVA, T.A. MARQUES & L. THOMAS Faustino, C.E.S., M.A. Silva, T.A. Marques & L. Thomas 2010. Designing a
shipboard line transect survey to estimate cetacean abundance off the Azores
archipelago. Arquipelago. Life and Marine Sciences 27: 49-58.
Management schemes dedicated to the conservation of wildlife populations rely on the
effective monitoring of population size, and this may require the accurate and precise
estimation of this parameter. Line transect distance sampling can be an effective approach
for estimating abundance. Little information is available regarding cetacean abundance in
the Azores. This paper had two aims: 1) to design a line transect shipboard survey to
estimate the absolute abundance of the most common cetaceans off the Azores; and 2) to
provide a set of potential survey effort scenarios to policy makers and environmental
managers. Three survey scenarios are assessed, and one detailed survey design is presented.
A total of 8,800 km of survey effort is recommended; at this level the expected coefficient
of variation of estimates is less than 0.3 for most species. However, if logistic constraints
prevent this, at least 5,000 km of survey effort should be used to achieve minimum sample
size requirements; this is estimated to take 36 days of effort. It is also recommended to
conduct a pilot survey. This would provide more detailed information that could be used to
improve the survey design of what would be the first survey of this magnitude ever to be
implemented in the Azores.
Key words: distance sampling, effort scenarios, logistics, methods, statistical robustness
Cláudia E.S. Faustino (e-mail: [email protected]), Scottish Oceans Institute, SMRU Limited,
New Technology Centre, North Haugh, St. Andrews, Fife KY16 9SR, United Kingdom;
Mónica A. Silva, Centro do Instituto do Mar (IMAR) da Universidade dos Açores,
Departamento de Oceanografia e Pescas, PT-9901-862 Horta, Portugal; Biology
Department, MS#33, Woods Hole Oceanographic Institution, Woods Hole MA02543, USA;
Tiago A. Marques1 & Len Thomas, Centre for Research into Ecological and Environmental
Modelling, University of St. Andrews, St. Andrews KY16 9LZ, UK; 1Centro de Estatística e
Aplicações da Universidade de Lisboa, PT-1749-016 Lisboa, Portugal.
INTRODUCTION
Many studies of wildlife populations require an
estimate of population density, size, or rate of
population change. Distance sampling can be an
effective approach for estimating such parameters
(Buckland et al. 2001). The most widely used
type of distance sampling is line transect
sampling (Thomas et al. 2010). Here, the observer
travels along a line, recording detected objects
and the distance from the line to each object
detected (hence the name, distance sampling). In
the standard methods, all objects on or near the
line should be detected, but this method allows a
proportion of objects within a certain distance of
the line to be missed (Buckland et al. 2001).
Achieving reliable results from a distance
sampling survey depends greatly on good survey
design. This relies upon two fundamental
principles: replication (i.e. multiple lines) and
randomization. A large enough number of lines
ensures that the variation in the number of objects
detected per unit survey effort (encounter rate)
can be adequately estimated, as well as that the
underlying distribution of distances available for
detection can be safely assumed as known. The
49
transect lines should be randomly positioned so
that each point within the study area has a known,
non-zero probability of being covered by a transect (the “coverage probability”) (Thomas et al.
2010). Additionally, obtaining reliable results
requires good field methods and data analysis
(Thomas et al. 2007).
Survey design encompasses the selection of a
target sample size to achieve a desired level of
precision for the estimates, and the layout of the
transect lines. The best choice for the layout of
the lines will depend namely, on the survey
region, logistics and efficiency (Buckland et al.
2001). Cetacean surveys generally take place in
large study areas and ship time is expensive so
continuous zigzag designs are often preferred as it
maximizes search effort time and can minimise
transit time between transects (Strindberg &
Buckland 2004). Spatial stratification can be used
to improve precision of estimates; the study area
can be divided into blocks that are likely to have
similar animal density and/or detection functions
(defined as the probability of detecting an object
at a given distance from a transect line or point).
Another benefit of stratification is that the study
area is divided into smaller areas, for which
managers may want separate abundance
estimates, or which may provide survey blocks of
a more manageable size (Thomas et al. 2010).
Other species-specific issues may need to be
taken into account during survey design, such as
responsive movement of animals to the
approaching observer (Buckland et al. 2001), and
diving behaviour of the species. Species
behaviour will also influence the choice of
method to use; sperm whales (Physeter
macrocephalus), for instance, are unavailable to
visual observers as they can dive for an hour or
more. However, this species vocalizes during a
considerable part of their dive; therefore, acoustic
surveys are potentially more valuable for sperm
whales than for many other species (Barlow &
Taylor 2005).
Most distance sampling surveys are analysed,
and many are designed, using the software
Distance (Thomas et al. 2010). The design
outputs can be useful in determining if a design is
feasible, and whether there is sufficient effort to
produce enough sightings for reliable analysis.
Once a design is chosen, a single realisation can
50
be generated and exported to be used as the
survey plan (Thomas et al. 2010).
Despite sometimes denoted as poorly
productive, the waters of the Azores contain a
high diversity of cetaceans with 23 species
confirmed to occur in the area (Steiner et al.
2007). In spring and summer, the most common
cetaceans in the Azores are the Atlantic spotted
dolphin (Stenella frontalis), short-beaked
common dolphin (Delphinus delphis), bottlenose
dolphin (Tursiops truncatus), Risso’s dolphin
(Grampus griseus), and the sperm whale
(Physeter macrocephalus) (Silva et al. 2003). The
IUCN Red List of Threatened Species for the
Azores lists the spotted, common and bottlenose
dolphins as Least Concern, Risso’s dolphins as
Data Deficient and sperm whales as Vulnerable
(Cabral et al. 2005). Additionally, all species are
protected under the EU Habitats Directive and
bottlenose dolphins are listed in the Annex B-II,
requiring the designation of Special Areas of
Conservation (DL 49/2005).
Estimates of abundance are missing for most
cetaceans in the Azores. With only a few
exceptions (e.g. Matthews et al. 2001; Silva et al.
2009), studies carried out in the area do not
provide such information, vital to the implementation of management schemes in the Azores.
Despite the current international push to get
information on cetacean distribution to support
marine spatial planning and habitat protection,
and the recommendations that Portugal should
carry out surveys to estimate cetacean abundance,
it has not yet been possible to secure funding to
support this research.
This paper had two aims: 1) to design a line
transect shipboard survey to estimate the absolute
abundance of the most common cetaceans off the
Azores; and 2) to provide information on
alternative scenarios for policy makers and
environmental managers. We provide clear
information and important considerations to take
in when creating a good survey design. We
briefly present the criteria used for choices made
along the iterative process of defining the
elements of a survey design. Three survey effort
scenarios are assessed to illustrate the range of
possibilities between statistical robustness and
logistic/ management restrictions, and one survey
design is presented.
MATERIAL AND METHODS
STUDY AREA
The archipelago of the Azores is composed of
nine volcanic islands divided into three groups,
extending ca. 600 km along a NW- SE axis
(Fig. 1). The Azorean Exclusive Economic Zone
(EEZ) comprises 938,000 km2, ca. 30% of the
European EEZ (Santos & Pinho 2005). The
islands are separated by deep waters (ca. 2,000 m)
with scattered seamounts (Santos et al. 1995).
The high bathymetric amplitude is known to
influence the local and regional circulation
patterns, which in turn influence the distribution
of pelagic organisms. There is a high seasonal
and inter-annual variability in the oceanographic
processes, which in turn influence the overall
circulation in the Azores (Seabra et al. 2006).
METHODS AND LOGISTICS
The survey will use mark recapture distance
sampling (see Laake & Borchers 2004 for
details), with a double platform configuration.
The target species for data collection will be
spotted, common, bottlenose, and Risso’s
dolphins as well as sperm whales. In the Azores,
spotted and common dolphins are eager bow
riders; bottlenose dolphins may also be attracted
to vessels whereas Risso’s dolphins and sperm
whales do not show attraction to vessels (Silva
pers. comm.). Data will be collected for all
species encountered, provided that this does not
compromise data collection for the target species.
The nautical survey will use visual detections for
the species of dolphins and passive acoustic
detections for sperm whales (e.g. Lewis et al.
2007). To give some starting point, it was
assumed that 20 days of ship time were available
to complete the survey. However, an estimate of
the required survey effort for a given precision
was expected to be an output of the design
process, so the final recommended effort is likely
to be substantially different from this.
The research vessel ARQUIPÉLAGO is used as a
model, as it has the desired specifications; it has a
cruising speed of 9.5 knots and maximum speed
of 11 knots, it is able to operate for 2,500 km
without making landfall, and is able to accommodate six scientists.
Fig.1. Location of the Azores archipelago in the
north Atlantic.
The study area will be defined to be as wide as
possible, from a minimum range of 20-30 km
around the islands. The areas between the three
groups of islands will be included if possible as
well as the seamount complex located south to
Pico Island. The survey will be conducted some
time between June and August, as these are the
months with better and more stable sea-state
conditions (Windguru 2007). At this time of year,
the day length is ca. 14 hours, allowing long days
of work. In this period the percentage of days
with sea-state below Beaufort 4 is about 80%
(Silva pers. comm.).
DEFINITION OF THE SURVEY AREA
The initial area considered to define the extent of
the survey was the geographic area defined by
Seabra et al. (2005), 258,228 km2 that enclosed
the effort and sightings recorded between 1999
and 2004, in two major projects conducted at the
Department of Oceanography and Fisheries of the
University of the Azores (DOP/UAç; “Cetamarh”
(1999-2004) and POPA (2001-2004). Different
shapes and widths for the survey areas were
considered, as well as the number of islands
included in a survey sub-area. Two buffer zones
were tested around the islands, one of 10 nautical
miles (nm) (suggested in previous studies) and
another of 12 nm (Territorial Sea). All maps were
projected in the most appropriate way for the
Azores (WGS1984 UTM Zone 26N).
DEFINITION OF THE SURVEY STRATA
Stratification was created to account for
geographical gradients, given the underlying
management interest. Strata were defined as: 1)
seamount complex SE Pico; 2) corridors between
island groups; 3) Western group; 4) Central
51
group; and 5) Eastern group of islands. Within
these, substrata were created to make the subareas more convex, reduce off-effort time (e.g.
Thomas et al. 2007), and to maximize the number
of transects per strata. Buckland et al. (2001)
recommend 10-20 replicates as a minimum and
Thomas et al. (2007) reinforce the use of > 15.
DEFINITION OF THE SURVEY PARAMETERS
Initially, the potential precision associated with
the choices of survey effort (i.e. line length, L)
was investigated using input parameters from
previous studies in the Azores (Projects Cetamarh
2000-2004, Golfinicho 2005-2006, LIFE (19992000)). These previous findings provided the
range of values of encounter rate (ER); a range of
plausible coefficient of variation (CV) and sample
size (n) was used.
The total line length (L) required in a main
survey was determined using the formula
proposed by Buckland et al. (2001), based on a
pilot study. Given a target CV, cvt , where
^
⎡
⎤ seˆ( Dˆ ) ;
⎢⎣cv( D)⎥⎦ = Dˆ
let n0 be the number of animals (or clusters)
counted in a pilot survey, and
the total line
length from pilot survey, then
L=
b
⎡
⎤
⎢⎣cvt ( D)⎥⎦
^
2
x
L0
n0
where,
⎡
⎤
n. var ⎢ f (0)⎥
⎣
⎦
^
var(n)
b≈
+
n
⎡ ^ ⎤
⎢⎣ f (0)⎥⎦
2
n being the number of animals (or clusters) and
f(0) the probability density function of detected
distances from the line, at zero distance. For
simplicity and lacking better information, b=3 is
52
used, following the suggestion of Buckland et al.
(2001).
Lastly, survey effort scenarios were generated
using R (version 2.5.1) (R Development Core
Team 2007). From these, three survey effort
options were chosen aiming to inform project
managers; one illustrating a scenario where the
resulting abundance estimates are robust, another
illustrating a more feasible scenario incorporating
cost-benefit aspects, and a third illustrating a
trade-off of statistical robustness and logistic/
management restrictions.
DEFINITION OF THE SURVEY DESIGN
An equal spaced zigzag line was chosen to create
the survey design in the present study (Strindberg
& Buckland 2004). A survey design was
generated using Distance 6 (Thomas et al. 2010)
for survey option with smaller effort, using a 2
km strip width and a coverage grid with points 2
km apart (9,817 points in total). The survey
region was approximated by a convex hull.
Effort was determined by line spacing, and
proportional effort was allocated to each
substratum. 5,000 simulations were run to
examine the coverage probability (i.e. assess how
even it is), and a minimum of 15 lines per stratum
was ensured. Additionally, on effort time needed
to perform this survey was compared with the
time allocated initially to perform the survey (20
days), to assess the feasibility of the survey
design.
RESULTS
DEFINITION OF THE SURVEY AREA AND STRATA
Oval regions were preferred, islands were
grouped per group (Eastern, Central, Western),
and 12 nm buffers were created (Fig. 2). Within
the five strata initially defined to account for
geographical gradients, a total of 16 substrata
were created. The survey area, strata and
substrata characteristics are summarized in Table
1. Total survey area is ca. 39,300 km2, and the
proportion of the total area represented by
substratum ranged from 2% (corridor SMi-SMa)
to 12.7% (seamounts_S).
Fig. 2. Map of the survey area showing the 16 substrata with respective label.
Table 1. Characteristics of each stratum and substratum defined for the survey design; % refers to total area.
Strata
Substrata
ID
Seamount
Seamount complex North
Seamount complex South
1
2
Corridors
Corridor Western Group to Central Group
Corridor Central Group to S. Miguel island
Corridor S. Miguel island to S. Maria island
3
4
5
Western
Group
Western Group West
Western Group East
6
7
Cental Group Northwest
Cental Group Northeast
Cental Group centre-top
Cental Group centre-bottom
Cental Group South
8
9
10
11
12
S. Miguel island North
S. Miguel island South
S. Maria island North
S. Maria island South
13
14
15
16
Central
Group
Eastern
Group
Survey area
1
Label
seamounts_N
seamounts_S
sum
corridor WG-CeG
corridor CeG-Smi
corridor SMi-SMa
sum
WG_W
WG_E
sum
CeG_ NW
CeG_ NE
CeG_centre_t
CeG_centre_b
CeG_S
sum
SMi_N
SMi_S
SMa_N
SMa_S
sum
survey area
Area (km2)
% of area
2,480.2
4,997.0
7,477.2
4,244.5
2,656.8
797.1
7,698.4
1,878.6
1,854.9
3,733.5
1,207.4
1,817.4
4,370.7
2,882.5
2,615.0
12,893.0
2,624.9
2,355.2
1,306.9
1,234.9
7,522.0
39,316.9
6.31
12.71
19.01
10.79
6.76
2.03
19.58
4.78
4.72
9.49
3.07
4.62
11.11
7.33
6.65
32.79
6.68
5.99
3.32
3.14
19.13
53
DEFINITION OF THE SURVEY PARAMETERS
The survey effort scenarios showed 8,250 km
were necessary to get CV≤0.3 for all species
except sperm whales, and 17,600 km provided
CV=0.36 for these whales and CV≤0.2 for the
remaining species. In order to obtain CV=0.2, the
amount of effort required for each species ranged
from ca. 6,000 km for spotted dolphins and ca.
57,500 km for sperm whales. Regarding sample
size, L needed to provide n=60 differed greatly,
varying between 4,850 km for spotted dolphins,
and ca. 46,500 km for sperm whales.
Given these results, further analysis for the
survey design aimed for CV≈0.2 (set by the
authors to illustrate good statistical robustness)
and n≈60 (practical minimum suggested by
Buckland et al. 2001); the ER used for each target
species corresponded to mean values recorded in
the Azores from June to August. Further, sperm
whales were left out from the decision-making
process given its abundance estimates will not
depend on visual sightings (and therefore on the
available visual-based ER, but on an acousticbased ER).
Table 2. Summary of the three survey design options
defined, coefficient of variation (CV) and sample size
(n) obtained per species. ER (mean number of animals
recorded per 100 km, for June to August). Codes used
for cetacean species: DDE – short-beaked common
dolphin; GGR – Risso’s dolphin; PMA – sperm whale;
SFR – Atlantic spotted dolphin; TTR – bottlenose
dolphin.
Option 1
Option 2
Option 3
L=5,000km
L=17,600km
L=8,800km
Sps
ER
CV
n
CV
n
CV
n
DDE
GGR
SFR
TTR
PMA
0.8
0.4
1.3
0.8
0.1
0.3
0.4
0.2
0.3
0.7
39.9
20.3
62.3
38.1
6.5
0.2
0.2
0.1
0.2
0.4
140.5
71.4
219.3
134.2
23.0
0.2
0.3
0.2
0.2
0.5
70.2
35.7
109.7
67.1
11.5
Lastly, three survey effort options were chosen, to
be presented to project managers: Option 1 – L =
5,000 km: incorporates cost-benefit aspects
54
(based on hypothetical budget) that result in the
possible loss of robustness of one of the target
species; it generates CV≈0.3 for all target species
except for Risso’s dolphins; Option 2 – L=17,600
km: defined as the minimum L that would
provide CV at least equal to 0.2; Option 3 –
L=8,800 km: defined as half the Option 2,
representing a trade-off of statistical robustness
and logistic/ management restrictions. Table 2
summarizes the values considered in the three
survey effort options. Despite not being used for
decision making, the corresponding values for
sperm whales are also shown.
DEFINITION OF THE SURVEY DESIGN
A survey design was generated for the survey
option with smaller effort (Option 1, L=5,000
km). Designs were not generated for the two
other effort options given the minimum number
(i.e. ≥15 lines) of line transects per stratum was
already achieved in Option 1. The coverage
probability generated was quite even (mean 0.49,
range < 0.001 to 0.76, SE=0.05). The angle of the
zigzag lines per substratum varied between 70o
and 175o to the x-axis (Table 3). Line spacing
(mean spacing for each substratum) ranged
between 7.96 km (seamounts_S) and 8.72 km
(corridor SMi-SMa), and the overall mean used in
the survey design was 8.32 km. The mean total on
effort line length generated for the survey design
was 4,956.4 km. The number of transect lines per
substratum ranged from 5 (corridor SMi-SMa)
and 28 (corridor OcG-CeG) and all strata had at
least 20 lines.
Twenty nine (29.7) days of effort would be
needed to complete the survey when L=5,000 km,
sailing at 9 knots with 10 h work per day; number
of days needed to survey each stratum would
range from 2.8 to 9.7 (Western and Central group,
respectively). A survey plan resulting from a
single realization of the chosen survey design is
shown in Figure 3. This gave a total line length of
4,968.2 km with 156 km off-effort (3.14% of the
total line length).
Fig. 3. Survey plan generated from a single realization of the survey design.
Table 3. Survey design summary. Transect length and number of samples (i.e. transect lines) are means; minimum
and maximum in brackets. Survey days refer to proportion of a total 20 days available and the number of days
required when travelling at 9 knots.
DESIGN
DA angle (°)
160
160
Spacing
(Km)
8.07
7.96
corridor OcG-CeG
corridor CeG-Smi
corridor SMi-SMa
160
160
120
8.55
8.70
8.72
OcG_W
OcG_E
70
70
8.01
8.00
CeG_ NW
CeG_ NE
CeG_centre_t
CeG_centre_b
CeG_S
140
165
150
160
160
8.29
8.28
8.16
8.46
8.42
Substrata label
seamounts_N
seamounts_S
SMi_N
SMi_S
SMa_N
SMa_S
175
175
170
170
overal mean
8.29
8.41
8.42
8.48
8.32
Total
On effort
trackline L (Km)
307.5 (288.2 - 313.9)
611.6 (602.7 - 621.8)
sum
554.26 (543.6 - 565.5)
346.4 (333.2 - 358.0)
101.5 (87.8 - 115.7)
sum
239.6 (223.4 - 250.2)
235.6 (215.7 - 248.9)
sum
154.5 (141.9 - 165.2)
227.0 (207.5 - 249.3)
533.8 (511.9 - 547.3)
365.5 (355.5 - 371.5)
330.5 (316.8 - 341.8)
sum
328.2 (318.8 - 348.1)
296.2 (313.8 - 348.1)
166.2 (157.3 - 180.5)
157.7 (148.2 - 166.9)
sum
4,956.4
(4,729.6 - 5,151.5)
# survey
# samplers
9.8 (8 -10)
13.4 (13 - 14)
23.2 (21 - 24)
27.1 (26 - 28)
18.1 (17 - 19)
6.00 (5 - 7)
51.2 (48 - 54)
10.9 (10 - 12)
10.9 (10 - 12)
21.9 (20 - 24)
7.6 (7 - 8)
9.6 (9 - 10)
18.1 (17 - 19)
16.9 (16 - 17)
15.3 (14 - 16)
67.5 (63 - 70)
13.7 (13 - 14)
13.5 (12 - 14)
7.9 (7 - 8)
7.8 (7 - 8)
42.9 (39 - 44)
206.7
(191 - 216)
from
20 days
1.3
2.5
3.8
2.2
1.4
0.4
4.0
1.0
0.9
1.9
0.6
0.9
2.2
1.5
1.3
6.5
1.3
1.2
0.7
0.6
3.8
20.0
at
9 knots
1.8
3.7
5.5
3.3
2.1
0.6
6.0
1.4
1.4
2.8
0.9
1.4
3.2
2.2
2.0
9.7
2.0
1.8
1.0
0.9
5.7
29.7
55
DISCUSSION
Three options for double platform survey effort
were presented to guide project managers in the
implementation of a shipboard survey design in
the Azores. All these excluded sperm whales (P.
macrocephalus) from the target species, given its
estimation will be based on acoustic detections,
for which there are no previous encounter rates
(ER) available for the Azores. Option 1 illustrated
a scenario based on a hypothetical budget, with
the expected cost of losing precision in the
estimates and possibly not allowing adequate
estimates for one target species (Risso’s dolphins,
G. griseus, the species with lowest ER). Option 2
illustrated a scenario where the expected CV
values are low and sample sizes are large. Despite
the statistical robustness, however, this may be an
excessive financial investment for a first survey.
Option 3 illustrated the trade-off between
statistical robustness and logistic/management
restrictions. Given money is a severe constraint in
the process of planning a design, and adding the
fact this would be the first survey of this
magnitude ever to be implemented in the Azores,
Option 3 (L=8,800 km) is the one recommended.
It may be that funds are not available to survey
even the lowest effort scenario we considered. In
this case, consideration could be given to
undertaking a multi-year survey, with different
areas surveyed in different years (see below).
Alternatively, a design with fewer strata might be
used, so that fewer lines are required to achieve
>15 per stratum. However, since density is
expected to differ between the strata suggested
here, this strategy will likely lead to greatly
increased variance. It will also lead to few
observations for fitting the detection function for
many species.
All scenarios generated (double platform
survey efforts), may nonetheless, be biased. ER
values were derived from previous singleplatform surveys, with a large proportion of
sightings that were not identified to species level,
and low height of the observation platform (Silva
pers. comm.). A double platform translates in
practical terms as having more observers
searching for cetaceans, and a (second) higher
platform of observation, increasing the probability of detecting the animals. The use of Mark
56
Recapture Distance Sampling will provide the
baseline of accurate information for future
double-platform surveys in the Azores. It is
important to stress that a small pilot survey
should precede the main survey designed here, in
order to refine field protocols and other practical
matters, as well as potentially provide better
estimates of encounter rate for use in planning the
main survey. In the absence of a pilot survey, the
main survey will likely become a pilot survey
(Buckland et al. 2001).
The information collected will be a single
snapshot in time. Nevertheless, if repeated every
four or five years, it could be possible to detect
trends in the populations of targeted species (e.g.
Taylor & Gerrodette 1993). This would also mean
one could increase the number of available
detections for each species over the years, which
would improve the modelling of detection
functions, allowing to increase a posteriori the
precision of estimates obtained even for the first
survey.
Reinforcing the underlying management
purposes of this work, the study area was created
using a 12 nm buffer around the strata of interest,
as this comprise the Azorean Territorial Sea.
Although there is insufficient data to define
substrata by a biological gradient (e.g. insufficient
data on costal populations), it is well known that
there are differences between geographical
regions. Silva et al. (2003) reported that cetaceans
were not seen equally in all three groups of
islands (Eastern, Central and Western), possibly
due to differences in the abundance or diversity of
food resources. Seamounts in the Azores may act
as feeding stations for some visitor species as
marine mammals, as they may localize pelagic
prey (Morato et al. 2008). Further, corridors
between islands were considered to illustrate an
off-shore habitat, but might nonetheless be
different when compared to other off-shore areas
not between islands.
Sixteen substrata were created. This improved
the survey design by allowing a better adjustment
of the non-convex survey region, providing short
transect length off-effort and thus maximizing
time on-effort (Thomas et al. 2007). The transect
line width is very small compared to the transect
length, so that overlap and other edge effects are
likely negligible (Strindberg & Buckland 2004;
Thomas et al. 2007). The equally spaced zigzag
used generated even close-to coverage
probabilities along the study area, and one
essential requirement for a good survey design,
randomization, was fulfilled (Buckland et al.
2001). Few points had low coverage probability,
possibly derived by the survey algorithm itself.
This unevenness may not affect the precision with
which animal abundance is estimated (Rexstad
2007). Furthermore, a minimum of 20 transect
lines were allocated to all strata, fulfilling a
second essential requirement for a good survey
design, replication (Buckland et al. 2001).
Even though the survey design generated for
the survey option with smaller effort has little offeffort time, the number of days at sea allocated
initially for the survey (20 days) was not
sufficient. To complete the total 5,000 km
transect in 20 days with 10 hours of work, the
average survey speed would have to be 13.4 knots
and this is an excessive survey speed. More days
should therefore be attributed to implement the
survey. The time on effort needed to perform this
survey at 9 knots was 29.7 days, and 20% (i.e. 6
days) should be added to account for bad weather
(Silva, pers. comm.). Therefore, approximately 36
days should be allocated so that the smaller effort
survey option can be conducted. Accounting for
the example of the large scale European survey
SCANS II (average work days of 6.5 h; Macleod
pers. comm.) and considering days with 8 h of
work, one would need 37.2 days to survey the
area, supporting the recom-mendation above.
The uses of alternative methods to estimate the
abundance of cetaceans should also be assessed,
such as passive acoustic, aerial line transects, or
perhaps mark-recapture for some species (see
Borchers et al. 2002; Evans & Hammond 2004 &
Mellinger et al. 2007 for reviews). These may be
particularly effective for some species, such as
sperm whales, not well catered for in the design
suggested here. The possibility to conduct the
survey in more than one year, with only few strata
surveyed at each time (mosaic survey) could also
be considered. A power analysis could be
performed allowing the evaluation of population
trends over time (Taylor & Gerrodette 1993). The
ultimate objective of these surveys would be to
obtain estimates which can be used for the
management of cetacean populations, and being
able to detect changes in abundance over time is a
fundamental requirement for adequate management.
CONCLUDING REMARKS
If logistic constraints persist and it is not possible
to opt for the intermediate effort option proposed
(8,800 km), at least 36 days should be allocated to
allow a feasible implementation of the survey
option with smaller effort (5,000 km). Not being
possible to allocate more survey effort, Risso’s
dolphins could be removed from the target
species given this was the species levelling the
minimum survey effort required to obtain good
precision levels. Careful consideration should
also be given to the field methods, as poor
methods can destroy an otherwise well-designed
survey. Data analysis should also be carefully
performed, although unlike survey design and
field methods, this can be re-done if improved
methods come to light, so it is less critical to get it
right the first time. It is highly recommended to
conduct a pilot survey. This would enable field
methods to be refined, as well as providing more
detailed information that could be used to
improve the survey design (e.g. number of
substrata in the survey area, existence of
biological gradients in the strata, survey effort
based on adequate ER).
ACKNOWLEDGEMENTS
We are grateful to all who contributed to this
work, namely Dr. Kelly Macleod for the constructive discussions to this paper. This research
was partly funded by “Agência Regional da
Energia e Ambiente da Região Autónoma dos
Açores”, through Interreg IIIB. M.A. Silva was
supported by an FCT postdoctoral grant
(SFRH/BPD/29841/2006). IMAR-DOP/UAç is
the R&D Unit #531 and part of the Associated
Laboratory #9 (ISR) funded through the pluriannual and programmatic funding schemes of
FCT-MCTES and DRCT-Azores. We are also
grateful to Rob Williams and an anonymous
reviewer for their constructive comments.
57
REFERENCES
Barlow, J. & B.L. Taylor 2005. Estimates of sperm
whale abundance in the northeastern temperate
Pacific from a combined visual and acoustic
survey. Marine Mammal Science 21(3): 429-445.
Borchers, D.L., S.T. Buckland & W. Zucchini 2002.
Estimating
Animal
Abundance:
Closed
Populations. Springer Verlag, London. 314 pp.
Buckland, S.T., D.R. Anderson, K.P. Burnham, J.L.
Laake, D.L. Borchers & L. Thomas 2001.
Introduction to Distance Sampling. Oxford
University Press, Oxford. 432 pp.
Cabral, M.J., J. Almeida, P.R. Almeida, T. Dellinger,
N. Ferrand de Almeida, M.E. Oliveira, J.M.
Palmeirim, A.I. Queiroz, L. Rogado & M. SantosReis 2005. Livro Vermelho dos Vertebrados de
Portugal. Instituto da Conservação da Natureza.
Lisboa. 660 pp.
Decreto de Lei nº 49/2005, de 24 de Fevereiro 2005.
Ministério do Ambiente e do Ordenamento do
Território.
Evans, P.G.H. & P.S. Hammond 2004. Monitoring
cetaceans in European waters. Mammal Review
34(1): 131–156.
Laake, J.L & D.L. Borchers 2004. Methods for
incomplete detection a distance zero. Pp. 108-189
in Buckland, S.T., D.R. Anderson, K.P. Burnham,
J.L. Laake, D.L. Borchers & L. Thomas 2004.
Advanced Distance Sampling. Oxford University
Press, Oxford. 416 pp.
Lewis, T., D. Gillespie, C. Lacey, J. Matthews, M.
Danbolt, R. Leaper, R. McLanaghan & A. Moscrop
2007. Sperm whale abundance estimates from
acoustic surveys of the Ionian Sea and Straits of
Sicily in 2003. Journal of the Marine Biological
Association of the United Kingdom 87: 353–357.
Matthews, J.N., L. Steiner & J. Gordon 2001. Markrecapture analysis of sperm whale (Physeter
macrocephalus) photo-id data from the Azores
(1987-1995). Journal of Research Management
3(3): 219-226.
Mellinger, D.K., K.M. Stafford, S.E. Moore, R.P.
Dziak & H. Matsumoto 2007. An Overview of
Fixed Passive Acoustic Observation Methods for
Cetaceans. Oceanography 30(4): 36-45.
Morato, T., D.A. Varkey, C. Damaso, M. Machete, M.
Santos, R. Prieto, R.S. Santos & T.J. Pitcher 2008.
Evidence of a seamount effect on aggregating
visitors. Marine Ecology Progress Ser. 357: 23–32.
R Development Core Team 2007. R: A language and
environment for statistical computing. R
Foundation for Statistical Computing, Vienna,
Austria. ISBN 3-900051-07-0.
Rexstad, E. 2007. Non-uniform coverage estimators for
58
distance sampling. CREEM Tech. report 01: 12 pp.
Santos, J.P.M.J. & J.L.S. Pinho 2005. Estudo das
correntes oceânicas na região envolvente da Ilha
Terceira no Arquipélago dos Açores. Universidade
do Minho. CEC - REC 23 (3): 31-41.
Santos, R.S., S. Hawkings, L.R. Monteiro, M. Alves &
E.J. Isidro 1995. Case studies and reviews: Marine
research, resources and conservation in the Azores.
Aquatic Conservation: Marine and Freshwater
Ecosystems 5: 311-354.
Seabra, M.I., M.A. Silva, S. Magalhães, R. Prieto &
R.S. Santos 2005. Desenvolvimento de um SIG
para analisar a influência das características do
habitat na ecologia dos cetáceos. Relatório Final
da Bolsa de Investigação do Projecto CETAMARH
(POCTI/BSE/38991/2001).
Universidade
dos
Acores, Portugal. 45 pp.
Silva, M.A., R. Prieto, S. Magalhães, R. Cabecinhas, A.
Cruz, J.M. Goncalves & R.S. Santos 2003.
Occurrence and distribution of cetaceans in waters
around the Azores (Portugal), Summer and Autumn
1999- 2000. Aquatic Mammals 29(1): 77-83.
Silva, M.A., S. Magalhães, R. Prieto, R. Cabecinhas,
R.S. Santos & P.H. Hammond 2009. Estimating
survival and abundance in a bottlenose dolphin
population taking into account transience and
temporary emigration. Marine Ecology Progress
Series 392: 263-276.
Steiner, L., M.A. Silva, J. Zereba & M.J. Leal 2007.
Bryde’s Whales, Balaenoptera edeni, observed in
the Azores: a new species record for the region.
Journal of the Marine Biological Association.
Biodiversity Records. 6 pp.
Strindberg, S. & S.T. Buckland 2004. Zigzag survey
designs in line transect sampling. Journal of
Agricultural, Biological and Environmental
Statistics 9: 443–461.
Taylor, B.L. & T. Gerrodette 1993. The Uses of
Statistical Power in Conservation Biology: The
Vaquita and Northern Spotted Owl. Conservation
Biology 7 (3): 489-500.
Thomas, L., R. Williams & D. Sandilands 2007.
Designing line transect surveys for complex survey
regions. Journal of Cetacean Research and
Management 9: 1-13.
Thomas, L., S.T. Buckland, E.A. Rexstad, J.L. Laake,
S. Strindberg, S.L. Hedley, J.R.B. Bishop & T.A.
Marques 2010. Distance software: design and
analysis of distance sampling surveys for
estimating population size. Journal of Applied
Ecology 47:5-14. Available at http://www.ruwpa.
st-and.ac.uk/distance/
Windguru (Internet). (cited 15 August 2007). Available
from: http://www.windguru.cz/pt/.
Accepted 30 April 2010.
The loggerhead sea turtle (Caretta caretta) on Sal Island, Cape Verde: nesting activity and beach surveillance in 2009 SÍLVIA P.P. LINO, EUCLIDES GONÇALVES & JACQUIE COZENS Sílvia P.P. Lino, E. Gonçalves & J. Cozens 2010. The loggerhead sea turtle
(Caretta caretta) on Sal Island, Cape Verde: nesting activity and beach
surveillance in 2009. Arquipelago. Life and Marine Sciences 27: 59-63.
Surveys for Caretta caretta nesting activities were performed during the nesting seasons
from the middle of June to end of October 2009 on Sal Island, Cape Verde. A total of 3628
activities were registered: 1071 nests, 2466 turtle tracks and 91 dead turtles. On nesting
beaches still used by locals to catch female turtles for their meat, nightly patrols from 9 pm
to 5 am resulted in a significant reduction in turtle mortality in comparison to non patrolled
beaches. On beaches regularly patrolled, an increment of nests per km was also observed
which allows us to conclude that the presence of trained Rangers does not disturb the turtles
or interrupt the nesting process.
Key words: mortality, poachers, population, patrols, rangers, tracks
Sílvia Lino1,2 (e-mail: [email protected]), 1Departamento de Oceanografia e Pescas, Universidade dos Açores, PT-9901-862 Horta, Portugal; Jacquie Cozens, 2ADTMA-SOS
Tartarugas, Café Cultural, Santa Maria, Sal, Cabo Verde; Euclides Gonçalves, CMS –
Câmara Municipal do Sal, Largo Hotel Atlântico, Espargos, Sal, Cabo Verde.
INTRODUCTION
Cape Verde harbours one of the world’s largest
nesting aggregations of loggerhead sea turtles
(Monzón-Argüello et al. 2007) which means that
protection of nesting habitats in these islands is of
critical concern for marine turtle populations
worldwide. However, we have not been able to
find reliable published data on number of nests.
Of the five species reported to be observed in
Capeverdian waters, the loggerhead, Caretta
caretta; the green turtle, Chelonia mydas; the
leatherback,
Dermochelys
coriacea;
the
hawksbill, Eretmochelys imbricata; and the Olive
Ridley, Lepidochelys olivacea (Merino et al. 2008
in review), only the loggerhead still nests
regularly on all the islands of Cape Verde.
According to Monzón-Argüello et al. (2007), the
majority of nesting activity occurs in the islands
of Boavista, Sal, Santa Luzia and Maio.
Marine turtles are considered endangered species
and have been protected since 2002 by national
laws. All activities that harm these animals in any
way are considered criminal and can lead to
penalties including prison sentences. Despite this,
poachers still go to nesting beaches to kill females
that come ashore to lay their eggs. To diminish
the marine turtle killing on land, military
personnel working for Sal City Hall have
patrolled some beaches on Sal each year since
2001. However, the number of loggerhead nests
and other activities connected with turtles, have
not been recorded consistently over the years.
Since 2008, biologists and volunteers from the
Non Governmental Organization SOS Tartarugas
increased the survey effort on the southern,
southwestern and southeastern beaches with the
aim of dissuading poachers. In that year, an
intensive program to register all the data related
to the nesting activity of loggerhead turtles began
59
as well. Here we present this data, based on the
information collected in the nesting season in
2009 (and 2008 for comparison) in an area of
about 12 km of sandy beaches.
MATERIAL AND METHODS
Data of all turtle activities (tracks, nests, false
crawls, dead animals and dragging tracks from
killed turtles) were recorded by a group of
permanent volunteers during the nesting season of
2009, from 12 June to 30 of October, on beaches
in the southern part of Sal Island, Cape Verde
(Fig. 1).
with red filtered head lamps and radios for
communication. Each of these groups began their
patrol on foot on several sections of the beaches
in a way that the groups would overlap each
other. When sighting a turtle, in order to diminish
the risk of disturbing nesting, the patrollers would
stop at a safe distance and sit in the dark until the
female started to lay the eggs or returned to the
sea. After recording the data in a notebook, the
track was erased with a specific mark in the sand
to avoid being counted twice. Every morning an
extra patrol was made from 6 to 9 am on a quad
bike, covering all the beaches and recording any
activity not registered the previous night. The
exact geographical position of the activities was
recorded using 6 GPS Garmin ETREX. Monthly
surveys were also conducted by a 4x4 pick-up
van in on all beaches around the island in order to
register the activities in beaches without surveillance.
RESULTS
Fig. 1. Locations of all loggerhead turtle activities
registered during the 2009 nesting season in Sal island,
Cape Verde (1. Santa Maria, 2. Algodoeiro, 3. Ponta
Preta, 4. Costa da Fragata, 5. Serra Negra, 6. Murdeira
and 7. Monte Leão).
Night-time beach patrols were undertaken
regularly from 9 pm to 5 am during that period, in
shifts of 4 hours with 2 to 3 persons, equipped
60
A total of 3628 activities for loggerhead sea
turtles were registered on Sal island during the
2009 nesting season (Table 1): 1071 were nests,
2466 were turtle tracks and 91 were dead turtles
(recorded either as cadavers or as drag marks left
by turtles turned upside down and dragged away
by poachers). The first activity, a False Crawl Uturn (FCU), was registered 19 of June during
morning patrol in Algodoeiro beach (southwest)
and the last, on night patrol 23 October, a nest in
Costa da Fragata (southeast). The mortality
registered (Table 1) was mainly due to human
activities: the highest number of turtles was killed
in areas with difficult access (as on the beaches
on the north coast) or in areas where patrols were
not conducted regularly (Serra Negra and Monte
Leão). Algodoeiro was the patrolled beach with
the highest number of turtles killed. Reasons for
these could be associated with the beach
characteristics: i) a vast open area that allows
poachers to watch the movements of the
patrollers and to avoid them and ii) nearer to
settlements where turtles have been use
traditionally as a food resource (turtle meat is still
highly appreciated by capeverdians).
Table 1. Summary of all activities registered during the 2009 nesting season on Sal Island beaches.
Tracks
Nests
Beaches (coast locations)
total
per km
Total
per km
Santa Maria (South)
Algodoeiro (Southwest)
Ponta Preta (Southwest)
Costa da Fragata (Southeast)
Serra Negra (Southeast)
Murdeira (West)
Monte Leão (West)
Other beaches (North)
total
101
786
161
609
581
21
96
202
2557
27
262
94
154
533
23
109
91
146
56
317
47
387
200
6
32
26
1071
15
106
27
98
183
7
36
12
61
Data shows that the areas most frequently used by
female turtles on Sal Island are Algodoeiro and
Costa da Fragata (Table 1; Figure 1). Of the
unprotected beaches, Serra Negra had the highest
number of activities with a total of 533 tracks and
98 nests per km, followed by Monte Leão with
109 tracks and 36 nests per km. Numbers of turtle
mortality show that both areas are also commonly
used by poachers, who usually try to hide the
carapaces behind the bushes or by burying them
in the sand. Beaches in the north have the highest
mortality rates. Regarding nesting density, results
confirm that Costa Fragata and Algodoeiro were
the preferred areas for turtles to lay their eggs
accounting for more than 60% of all nests on the
island. Surprisingly, the northern beaches
contribute only 4% of all nests on Sal, more than
in Murdeira. This is a rocky region with one large
bay and smaller ones flanked with very small
beaches and locally described as being very
important for the nesting turtles. It is the only
marine protected area in Cape Verde islands but
according to our data, Murdeira only presented
about 1% of all turtle nests on Sal.
Table 2. Numbers of female tracks registered during
the nesting seasons 2008 and 2009; *unpublished data
provided by J. Cozens (SOS Tartarugas).
Month
2008*
2009
June
July
August
Septem
ber
October
Total
90
505
451
48
1099
1520
210
617
24
1280
78
3362
Mortality
total per km
0
17
0
6
13
1
11
43
91
0
6
0
2
12
1
13
19
5
Situation
Status
protected
protected
protected
protected
unprotected
unprotected
unprotected
unprotected
-
In comparison with the previous year (Table 2),
2009 had three times more turtle activities
registered. The first nesting month (June) was the
only exception indicating that the season for
nesting started later but had more registered
activities during a longer period of time. By July
the registered activities were already double
compared to the year before and in August
registered activities were three times higher. The
comparison between the two years also allows us
to observe that in 2008, July was the month with
more activities, while in 2009 it was August. The
number of nests registered for 2008 was 346
(Cozens 2009).
DISCUSSION
The overall results demonstrate that constant
protection of nesting beaches is of critical
concern for Sal’s marine turtle population since
turtle killings are seen to decrease, on average,
from 11 per km in unprotected areas to 2 per km
on patrolled beaches. These results also indicate
that a regular presence of Rangers on the nesting
beaches has an effect of dissuading the hunters to
kill female turtles. On the other hand, several
times in wide open extensions of beaches as in
Algodoeiro, turtles were hunted in areas between
two groups of patrols which led us to believe that
poachers observe the Rangers movements and try
to take a turtle when a chance presents itself. The
methods of poachers have been observed to
change with the presence of volunteers: before
SOS Tartarugas patrolled regularly, turtles were
61
dragged on their backs a short distance from the
water, killed immediately and the meat taken,
leaving the carapace on the spot. Nowadays the
turtles are dragged for longer distances, and left
on their backs in hidden places far from the water
mark (where the patrols are made) until they find
the opportunity to kill them. Several turtles were
rescued when Rangers recognizing the drag mark,
followed the trail and found the female, upside
down. More than once the Rangers detected
poachers hidden near the turtle. This indicates
that surveillance alone may not be enough to stop
the killings. This is particularly true on beaches in
the north with very difficult access. Therefore it is
of critical importance to mobilize the local
community to protect turtles, which might result
in the reduction of unnatural mortality. Costa da
Fragata and Algodoeiro, the areas where the
patrol effort was higher, accounted for more than
60% of all nests. These results can be considered
as good indicators that the presence of highly
trained Rangers on the nesting beaches does not
disturb the turtles and allows them to nest
successfully in contrast to areas where poachers
and other human activity (such as in Santa Maria
with loud music from bars and nightclubs, bright
lights in front line developments and great
construction sites), have a negative influence on
the nesting activities.
Serra Negra and Monte Leão are two areas
where protection is not as complete due to poor
access. The 24 dead turtles found (more than 25%
of total mortality) and the high number of
activities per km at these locations shows that
they are very important for turtles and should be
protected.
Although two years of data is not sufficient to
estimate the nesting population on Sal Island the
comparison of the overall activities during the
same period in two consecutive years allowed us
to understand that, as it is described for Florida
beaches by Witherington et al. (2009), the
number of nests varies annually and that 2009
could have been a “maximum peak year”. If we
compare the overall numbers, results show that in
areas with similar number of activities (e.g. Costa
da Fragata and Serra Negra or Ponta Preta and
Monte Leão) the number of nests increases and
the mortality diminishes in patrolled beaches.
According to a NOAA report (2010), in the
62
Eastern Atlantic, the Cape Verde islands support
an intermediate-sized loggerhead nesting
assemblage. It seems that from our data in Sal
Island the turtle nesting has increased from 2008
(346) to 2009 (1071). Although, the number of
nests in 2008 might be too low as some nests
might have been overlooked, we can conclude
that an increase in nesting has occurred, which is
encouraging when some of the world’s
loggerhead nesting is decreasing (Hawkes et al.
2005; Margaritoulis 2005; Witherington et al.
2009). For instance in Florida, on the coast that
hosts between 80–90% of the world’s loggerhead
nesting activity, the nesting declined by 37%
between 1989 and 2007 and by 49% between
1998 and 2007 (Witherington et al. 2009). Our
results suggest that the constant surveillance of
the most important beaches for nesting turtles can
be of great importance not only for the recovery
of the Capeverdian population but also have a
meaningful impact for the marine turtle populations worldwide in years to come.
CONCLUSIONS
Surveillance actions in the most important
beaches for nesting turtles in Sal Island resulted
in the increase of nests and decrease of mortality
by poachers proving that these actions can have a
positive contribution to stop the decline that these
marine populations face nowadays. Combined
efforts should be made by NGOs, Sal City Hall,
the population and the Government Environment
Department in order to protect the Sal beaches in
the future, especially those with difficult access,
like Serra Negra and Monte Leão.
ACKNOWLEDGEMENTS
The authors would like to especially thank:
Victoria Abbott, Peter Aspden, Linda Aspden,
Gwenaelle Barach, Helena Batalha, Stephen
Brown, Floriano Furtado, Neal Clayton, Sandra
da Graça, Ilaria Mura, Anderson Gammon, João
Gouveia, Robert Hallsworth, Faye Heslop, Anna
Heslop, Heidi Karlberg, Filipe Lopes da Silva,
Patrizia Lozzi, Andrea Mason, Edson Mendes,
Mariel Murazzi, Lauren Nadler, Manuel Pereira,
Katie Quin, Paulo Rocha, Case Santos, Joseph
Scarola and Adriana Volpi. Their collaboration
was essential for the data collection during the
nesting season. Our gratitude also to Sal
Municipality, for their ongoing partnership with
SOS Tartarugas, WWF Cabo Verde and António
Ramos Cruz who made the monthly island survey
viable, the 2ª Military Region of Cape Verde
Republican Army Force (2ª Região Militar das
Forças Armadas da República de Cabo Verde),
whose soldiers helped not only by providing
valuable information but also by making sure
SOS volunteers were safe on the beaches. We are
grateful to Sandra Sequeira and Ricardo Medeiros
at Department of Oceanography and Fisheries,
University of the Azores, for their help in plotting
the data. This work was only possible through the
support of the many donors who are funding SOS
Tartarugas.
REFERENCES
Cozens, J. 2009 ADTMA SOS Tartarugas Cabo Verde
Relatório de Campanha de 2008 [Internet].
Available from: http://www.sostartarugas.org/SOS
Tartarugas/Resultados08_pt.pdf (cited 17 June2010)
Hawkes, L.A., A.C. Broderick, M.H. Godfrey & B.J.
Godley 2005. Status of nesting loggerhead turtles
Caretta caretta at Bald Head Island (North
Carolina, USA) after 24 years of intensive
monitoring and conservation. Oryx, 39: 65-72.
Margaritoulis, D. 2005. Nesting Activity and
Reproductive Output of Loggerhead Sea Turtles,
Caretta caretta, Over 19 Seasons (1984-2002) at
Laganas Bay, Zakynthos, Greece: The Largest
Rookery in the Mediterranean. Chelonian
Conservation and Biology,4(4): 916-929.
Merino, S., S. Correia, I. Cruz & M.A. Correia 2008.
The Cape Verde Archipelago and the Protection of
Sea Turtles. ambientalMENTEsustentable, (II), 4:
117-123 pp.
Monzón-Argüello, C., C. Rico, E. Naro-Maciel, N.V.
Cruz, P. López, A. Marco & L.F. López-Jurado
2007. Population genetic analysis of loggerhead
turtles in the Cape Verde islands. Proceedings of
the 27th Annual Symposium on Sea Turtle Biology
and Conservation, Myrtle Beach, South Carolina,
USA, 245.
NOAA Fisheries, office of protected resources, 2010.
Loggerhead Turtle (Caretta caretta) [Internet].
Technical Report available from: http://www.nmfs.
noaa.gov/pr/species/turtles/loggerhead.htm (cited
25 April 2010)
Witherington, B., P. Kubilis, B. Brost & A. Meylan
2009. Decreasing annual nest counts in a globally
important loggerhead sea turtle population.
Ecological Applications, 19:30–54.
Accepted 5 July 2010.
63
SHORT COMMUNICATION First records of some species of Diptera (Insecta) from the Azores JINDŘICH ROHÁČEK & JAROSLAV STARÝ
Roháček, J. & J. Starý 2010. First records of some species of Diptera (Insecta)
from the Azores. Arquipelago. Life and Marine Sciences 27: 65-68.
Jindřich Roháček (e-mail: [email protected]), Department of Entomology, Silesian
Museum, Tyršova 1, CZ-746 01 Opava, Czech Republic; Jaroslav Starý, Department of
Zoology and Laboratory of Ornithology, Faculty of Science of the Palacký University, tř.
Svobody 26, CZ-77146 Olomouc, Czech Republic.
During a collecting trip undertaken by J. Roháček
and M. Vála in the São Miguel Island (Azores) in
August and September 2006, mainly devoted to
acalyptrate flies, three distinctive species of
Diptera were found, two of which proved to be
hitherto unrecorded from the Azorean
archipelago. These additions to the regional fauna
are presented below with a discussion of their
origin.
The voucher specimens of the species recorded
below are deposited in the following collections:
JSO – collection of J. Starý, Olomouc, Czech
Republic, SMOC – Silesian Museum, Opava,
Czech Republic, ZMAN – Zoological Museum,
Amsterdam, Netherlands.
FAMILY TIPULIDAE
Tipula (Tipula) oleracea Linnaeus, 1758
Material examined: Azores: São Miguel I.:
Remédios nr. Lagoa 0.5 km S, 37º45' N 25º34' W,
180 m, meadow, 02.09.2006, 2 males; Lagoa do
Fogo, NW shore, 37º46' N 25º29' W, 575-600 m,
netted on lake shore, 7.09.2006, 2 males 2
females, J. Roháček leg.; Lombadas, Ribeira
Grande river, 37º47' N 25º27' W, 580 m,
sweeping riverside vegetation, 05-09-2006, 3
males, J. Roháček & M. Vála leg., P. Oosterbroek
det. (SMOC, JSO, ZMAN).
Comments: A species native to the West
Palaearctic area; common and widespread there,
known as an agricultural pest. Also widely distributed in the Nearctic (Canada and USA),
probably introduced by accidental transport. The
recent record from Ecuador (Young et al. 2000)
seems to be a similar case. In Macaronesia
recorded from Canary Islands (Gran Canaria)
(Oosterbroek & Eiroa 2004). The larvae live in
soil in gardens, pastures, or meadows, feeding on
roots of grasses, seedlings and crops, thus causing
commercial losses (Young et al. 2000).
Oosterbroek (2009: 195) recorded this species
from the Azores based on some of the specimens
listed above but without giving precise collecting
data. The species was observed flying in numbers
in habitats with low, mostly grassy vegetation,
including those in montane valleys (see Fig. 1).
This is the second species of the Tipulidae from
the Azores. The endemic Tipula (Savtshenkia)
macaronesica Savchenko, 1961 was not
collected.
FAMILY DROSOPHILIDAE
Dettopsomyia nigrovittata (Malloch, 1924)
Material examined: Azores: São Miguel I.: Sete
Cidades 1 km N, 37º52' N 25º47' W, 280-330 m,
sweeping over meadow, 08-09-2006, 1 female;
65
Fig. 1. Valley of the Ribeira Grande river in Lombadas (São Miguel Island, Azores),
habitat of Tipula oleracea (photo by J. Roháček).
Fig. 2. A sugar-beet field after harvest in bottom of a small crater 4 km N of Ponta Delgada
(São Miguel Island, Azores), habitat of Dettopsomyia nigrovittata (photo by J. Roháček).
66
Ponta Delgada 4 km N, 37º46' N 25º41' W, 200230 m, sweeping over field margin, 4.ix.2006, 1
female, both J. Roháček leg. and det. (SMOC).
Comments: A widespread (mainly tropical to subtropical) species known from North and South
America, Hawaii, Japan (including Bonin
Islands), Oriental Region, Australia, Africa and
Canary Islands (Okada 1982; Singh & Fartyal
2002). There is only one previous record from the
whole West Palaearctic area, viz. that from
Canary Islands (Tenerife, see Hackman 1958)
where the species is considered introduced
(Bächli et al. 2004). The specimens examined
were collected in man-affected habitats, at
margins of a field (Fig. 2) and a meadow,
respectively.
grant (either from Europe or North America)
inasmuch as it establishes rich populations
(documented in 3 localities) even in mountain
regions (Fig. 1) without grassland habitats typical
of this species. The drosophilid Dettopsomyia
nigrovitta certainly has a different origin. As it is
originally a tropical-subtropical species it was
most probably introduced from Africa (or from
Canary Islands, see above), possibly with
transport of tropical fruits, ornamental flowers or
other products. The record from the Azores
apparently is the northernmost occurrence of the
species. The bird parasite Ornithomyia chloropus
may have reached the Azores with host birds
from any part of the West Palaearctic area. The
finding in the São Miguel I. represents a new
westernmost distributional limit of this species
(see Petersen 2009).
FAMILY HIPPOBOSCIDAE
Ornithomyia chloropus Bergroth, 1901
Material examined: Azores: São Miguel I: Sete
Cidades 1 km N, 37º52' N 25º47' W, sweeping
vegetation in wet ravine, 31.08.2006, 1 male, J.
Roháček leg. and det. (SMOC).
Comments: A Palaearctic species mainly recorded from North Europe (including Iceland) but
also known from Central Europe, rarely in South
Europe, Near East and North Africa (Büttiker
1994; Petersen 2009). In Macaronesia hitherto
only recorded from the Canary Islands (Tenerife
and La Palma, see Báez 1978). This bloodsucking bird parasite displays a very low host
specificity. Although preferentially living on
various species of Passeriformes it has also
recorded from representatives of Strigiformes,
Falconiformes and Lariformes (Chalupský 1980).
It was therefore not surprising to find it in São
Miguel Island (albeit purely accidental, as
previously mentioned).
DISCUSSION
All the above recorded species belong to
widespread taxa which were obviously recently
introduced into the São Miguel Island. Tipula
oleracea proved to be a very successful immi-
ACKNOWLEDGEMENTS
We are very grateful to Dr. M. Báez (La Laguna,
Tenerife, Canary Island, Spain), Dr. G. Bächli
(Zürich, Switzerland) and Dr. P. Oosterbroek
(Amsterdam, the Netherlands) for valuable comments on the manuscript.
REFERENCES
Báez, M. 1978. Los Hippobóscidos de las Islas
Canarias (Dipt., Hippoboscidae). Boletín de la
Estación Central de Ecología 7(13): 59-72. [In
Spanish]
Bächli, G., C.R. Vilela, S. Andersson Escher & A.
Saura 2004. The Drosophilidae (Diptera) of
Fennoscandia and Denmark. Fauna entomologica
Scandinavica, Vol. 39. Brill, Leiden – Boston, 362
pp.
Büttiker, W. 1994. Die Lausfliegen der Schweiz
(Diptera, Hippoboscidae). Documenta faunistica
Helvetiae 15, Centre suisse de cartographie de la
faune (CSCF), Neuchâtel. 117 pp. [In German]
Chalupský, J. 1980. 7. čeleď Hippoboscidae –
Klošovití. Pp. 447-478 in Chvála, M. (Ed.).
Krevsající mouchy a střečci – Diptera. Fauna
ČSSR, Vol. 22, Academia, Praha, 538 pp. [In
Czech]
Hackman, W. 1958. Drosophilidae. Pp. 47-48 in: Frey,
R. (Ed.): Kanarische Diptera brachycera, von
67
Håkan Lindberg gesammelt. Societas Scientiarum
Fennica. Commentationes Biologicae 17(4). [In
German]
Okada, T. 1982. A revision of the genera Dettopsomyia
Lamb and Styloptera Duda (Diptera, Drosophilidae). Kontyû, Tokyo 50(2): 270-282.
Oosterbroek, P. 2009. New distributional records for
Palaearctic Limoniidae and Tipulidae (Diptera:
Craneflies), mainly from the collection of the
Zoological Museum, Amsterdam. Zoosymposia 3:
179-197.
Oosterbroek, P. & E. Eiroa 2004. On the Tipulidae
(Insecta, Diptera) of Spain, Portugal, and Andorra.
Studia Dipterologica 11: 199-201.
68
Petersen, F.T. 2009 (Internet). Fauna Europaea.
Hippoboscidae. In: Pape, T. (Ed.). Fauna
Europaea: Diptera, Brachycera. Fauna Europaea
version 2.1. (cited 27 January 2010). Available
from: http://www.faunaeur.org.
Singh, B.K. & R.S. Fartyal 2002. Family Drosophilidae
(Insecta: Diptera) in Kumaon Region, India, with
the description of one new species and three new
records. Proceedings of the Zoological Society,
Calcutta 55: 11-18.
Young, C.W., G. Onore & K. Proaño 2000. First
occurrence of Tipula (Tipula) oleracea Linnaeus
(Diptera: Tipulidae) in the New World, with
biological notes. Journal of the Kansas Entomological Society 72: 226-232.
Accepted 16 March 2010.
SHORT COMMUNICATION Shrimps (Crustacea, Decapoda, Caridea) associated with gorgonians at the coast of Senegal PETER WIRTZ & SAMMY DE GRAVE
Wirtz, P. & S. De Grave 2010. Shrimps (Crustacea, Decapoda, Caridea)
associated with gorgonians at the coast of Senegal. Arquipelago. Life and Marine
Sciences 27: 69-71.
Peter Wirtz (e-mail: [email protected]), Centro de Ciências do Mar do Algarve,
Campus de Gambelas, PT-8000-117 Faro, Portugal; Sammy de Grave, Oxford University,
Museum of Natural History, Parks Road, Oxford, OX1 3PW, United Kingdom.
INTRODUCTION
Symbioses are common in the marine environment. Some taxa appear to be particularly
prone to be involved in associations. Crustaceans
probably form more associations with other
classes than any other marine animals (Ross
1983), crustacean – cnidarian associations being
particularly common (Patton 1967). Gorgonianassociated decapods have been described from
both sides of the Atlantic (e.g. Spotte et al. 1994,
1995, Wirtz & d´Udekem d´Acoz 2001, Wirtz et
al. 2009). Gorgonians were therefore searched for
associated decapods during three dives in the area
of NGor, Senegal, i.e. at the western tip of Africa.
MATERIAL AND METHODS
All observations were made while SCUBA
diving. The first dive was at 14°45.643' N,
17°30.710' W, on 11 October 2009. An unidentified green gorgonian (Fig. 1) in 14 m depth was
visually searched and a hand-held aquarium net
was wiped over its surface. The second dive was
at 14º45.673' N, 17º31.079' W, on 12 October
2009. Several large Leptogorgia sp. (Fig. 2) in 25
m depth were visually searched and a hand-held
aquarium net was wiped over their surfaces. The
third dive was at 14º43.806' N, 17º32.046' W, on
20 October 2009 in 29 m depth. An unidentified,
long-armed, red gorgonian (Fig. 3) was visually
searched and a hand-held aquarium net was
wiped over its surface.
Specimens were deposited at the Oxford
University Natural History Museum (OUMNH)
under the numbers 2009-27-02 and -03 (Rapipontonia platalea), 2009-27-04 and -05 and -06
(Pseudocoutierea wirtzi), and 2009-27-07 and -08
(Hippolyte cf. palliola).
RESULTS
Three species of shrimps were found on the
gorgonians sampled.
Hippolyte cf. palliola Kensley, 1970
An ovigerous female of this species was found on
the unidentified green gorgonian in 14 m depth,
and another one on the unidentified red gorgonian
from 29 m depth (Fig. 1).
Hippolyte palliola is known from Guinea to
western South Africa (Crosnier 1971; d'Udekem
d'Acoz 2007) and apparently has not yet been
recorded in association with other invertebrates.
Pseudocoutierea wirtzi d’Udekem d'Acoz, 2001
Numerous animals of this species were found on
the gorgonians checked in all three dives.
69
Fig. 1a) Unidentified green gorgonian, host of three
species of symbiotic shrimps; b) Leptogorgia sp. (25 m
depth); c) unidentified red gorgonian, host of two
symbiotic shrimp species.
This species has so far been recorded only from
the Cape Verde Islands and from São Tomé
Island, where it also lives in large groups on
gorgonians or solitarily on whip-coral (Wirtz &
d'Udekem d'Acoz 2001, 2008)
Rapipontonia platalea (Holthuis, 1951)
Numerous individuals of this species were found
on the unidentified green gorgonian in 14 m depth
and on the Leptogorgia surveyed in 25 m depth.
Rapipontonia platalea is known from the Cape
Verde Islands, from Guinea and from São Tomé
and Príncipe in the Eastern Atlantic and from
Tobago in the Western Atlantic (Hale & De
Grave 2007). Wirtz & d'Udekem d'Acoz (2001)
noted that it lives in symbiosis with black coral
and gorgonians, while Hale & De Grave (2007)
found it on a hydroid encrusted with a zoantharian.
70
DISCUSSION
At present, it appears unlikely that the gorgonians
in any way benefit from the shrimps living on
them. Most gorgonians are unpalatable to
predators such as fish (Epifanio et al. 1999; and
references therein). Gorgonian symbionts might
profit from the fact that their hosts are avoided.
Shrimps living on gorgonians probably feed on
gorgonian tissue and on particles captured by the
gorgonian polyps. It remains to be tested if
symbiotic shrimps perhaps even take up
unpalatable compounds from their hosts and
thereby become unpalatable themselves.
ACKNOWLEDGEMENTS
Many thanks to Philippe and Hilda of the
Nautilus diving base at NGor and to Karl
Wittmann, diving companion of the first author,
for their kind help. The Centro de Ciências do
Mar (CCMAR), Faro, Portugal, partly financed
the travel costs of the first author.
REFERENCES
Crosnier, A. 1971. Sur quelques Crustacés Décapodes
ouest-africains nouveaux ou rarement signalés.
Bulletin du Muséum National d'Histoire Naturelle,
3éme série, Zoologie 9:569-595.
Epifanio, R. de A., D.L. Martins, R. Villaca & R.
Gabriel 1999. Chemical defenses against fish
predation in three Brazilian octocorals: 11, 12epoxypukalide as a feeding deterrent in
Phyllogorgia dilatata. Journal of Chemical Ecology 26: 2255-2265.
Hale, R. & S. De Grave 2007. The first record of
Periclimenes platalea Holthuis, 1951 (Decapoda
Pontoniinae) in the Western Atlantic. Crustaceana
80(8): 1019-1021.
Patton, W.K. 1967. Commensal crustacea. Proceedings
Symposium
Crustacea,
Marine
Biological
Association of India Part III: 1228-1243.
Ross, D.M. 1983. Symbiotic relationships. Pp: 163212 in: Bliss, D (Ed.), The Biology of Crustacea,
New York, Academic Press.
Spotte, S., R.W. Heard & P.M. Bubucis 1994.
Pontoniine
shrimps
(Decapoda:
Caridea:
Palaemonidae) of the northwest Atlantic. IV.
Periclimenes antipathophilus, new species, a black
coral associate from the Turks and Caicos Islands
and eastern Honduras. Bulletin of Marine Science
55: 212-227.
Spotte, S., P.M. Bubucis & R.M. Overstreet 1995.
Caridean shrimps associated with the slimy sea
plume
(Pseudopterogorgia
americana)
in
midsummer at Guyana Island, British Virgin
Islands, West Indies. Journal of Crustacean
Biology 15: 291-300.
Udekem d'Acoz, C. d' 2007. New records of Atlantic
Hippolyte, with the description of two new species,
and a key to all Atlantic and Mediterranean species
(Crustacea, Decapoda, Caridea). Zoosystema 29(1):
183-207.
Wirtz, P., G. de Melo & S. De Grave 2009. Decapoda
from Actiniaria, Gorgoniaria, Antipatharia and
Echinodermata at the coasts of Espirito Santo,
Brazil. Marine Biodiversity Records 2, and 162.
Wirtz, P. & C. d'Udekem d'Acoz 2001. Decapod
crustaceans
associated
with
Antipatharia,
Gorgonaria and Mollusca at the Cape Verde
Islands. Helgoland Marine Research 55: 112-115.
Wirtz, P. & C. d'Udekem d'Acoz 2008. Crustaceans
associated with Cnidaria, Bivalvia, Echinoidea and
Pisces at São Tomé and Príncipe Islands (eastern
central Atlantic). Arquipélago. Life and Marine
Sciences 25: 63-69.
Accepted 25 May 2010.
71
SHORT COMMUNICATION First records of Tarentola mauritanica (Linnaeus, 1758) (Reptilia; Gekkonidae) in the Azores JOÃO P. BARREIROS, R.B. ELIAS, J. LOURENÇO, E. DIAS & P. BORGES
Barreiros, J.P., R.B. Elias, J. Lourenço, E. Dias & P. Borges 2010. First records of
Tarentola mauritanica (Linnaeus, 1758) (Reptilia; Gekkonidae) in the Azores.
Arquipelago. Life and Marine Sciences 27: 73-75.
João Pedro Barreiros (e-mail: [email protected]),aDepartamento de Ciências Agrárias,
Universidade dos Açores, PT-9701-851 Angra do Heroísmo, Portugal; IMAR, Centro do
Imar da Universidade dos Açores, PT-9901-962 Horta, Portugal; Rui B. Eliasa, Joana
Lourençoa & Eduardo Diasa, Centro do Clima Meteorologia e Mudanças Globais
(C-CMMG; CITA-A), Departamento de Ciências Agrárias, Universidade dos Açores,
Terra-Chã, PT-9701-851 Angra do Heroísmo, Portugal; Paulo Borgesa, Azorean
Biodiversity Group (CITA-A), Departamento de Ciências Agrárias, Universidade dos
Açores, Terra-Chã, PT-9701-851 Angra do Heroísmo, Portugal
INTRODUCTION
The Moorish gecko Tarentola mauritanica
(Linnaeus, 1758) is a widespread species native to
the Mediterranean region from southern France to
Greece and northern Africa (Loveridge 1947;
Martínez-Rica 1997; Hódar 2002; Perera et al.
2008, 2010; Plezeguelos et al. 2008). It has
recently been reported as living and breeding in
California (Marhdt 1998) and also as an
introduced species in Madeira (Báez & Biscoito
1993). Tarentola mauritanica is paraphyletic with
respect to T. angustimentalis Steindachner, 1891,
a Canary Islands endemic (Harris et al. 2004a).
Here we report new occurrences of the Moorish
gecko on Terceira Island, Azores archipelago, and
3 other occurrences on the islands of São Miguel
and Faial. The possibility of an already
established breeding population is discussed.
MATERIAL AND METHODS
Two live specimens of Tarentola mauritanica
were collected in the central area of the city of
Fig. 1. A juvenile Tarentola mauritanica collected in
June 2009 inside a house on the outskirts of Angra do
Heroísmo, Terceira Island. Photo by R.E.
73
Table 1. Dates and locality of confirmed sightings/collection of Tarentola mauritanica in the
Azores.
Island
Terceira
Terceira
São Miguel
São Miguel
São Miguel
São Miguel
Faial
Locality
Angra do Heroísmo
Angra do Heroísmo
Ponta Delgada
Ponta Delgada
Fajã de Cima
Pico Salomão
Horta
Date
Jan 2007
Nov 2007
2002
2002-2009
2002-2009
2010
2009
RESULTS AND DISCUSSION
Fig. 2. An adult, tailless, Tarentola mauritanica,
caught in July 2009 inside a box of lettuce imported
from mainland Portugal and put for sale in a supermarket on the outskirts of Angra do Heroísmo, Terceira
Island. Photo by R.E.
Angra do Heroísmo, Terceira Island, Azores. On
1 January 2010 a juvenile of 42 mm Total length
(Fig. 1) (specimen A) was captured by one of the
authors (JL) outside his house after an adult had
been killed and two other animals had been seen
alive in the same house as well as one in a
neighbouring house (pers. comm. to R.E.).
A second specimen (specimen B), caught alive
on 2 November 2007 by one of the authors (JPB)
from a box of lettuce imported from mainland
Portugal, was kept in a terrarium until 1of March
2008. When it died the length without tail was of
93 mm (Fig. 2). Table 1 summarises the known
occurrences of T. mauritanica in the Azores.
The specimens were fixed in 10% formalin and
preserved in 70% alcohol and deposited in the
Arruda Furtado collection of Department of
Agriculture (University of the Azores) under the
provisional catalogue numbers, EFAF_R0001 and
EFAF_R0002, respectively.
74
As reported by Báez & Biscoito (1993) for
Madeira, the Moorish gecko seems also to be an
accidental anthropogenic introduction in the
Azores. Although our specimen B was clearly a
case of an adult imported from mainland Portugal
inside a box of lettuce, specimen A was probably
born on the island, due to its small size and
reports of adults around the same area.
The absence of native populations of geckos in
the Azores implies that the native invertebrate
species did not co-evolve with this predator, in
fact with any similar one, being eventually
vulnerable to increase predation at least in less
impacted and preserved areas. However, it is not
clear which will be the habitat selection of
Tarentola mauritanica in the Azores due to the
species’ rate of adaptability and genetic differences between known populations (see
Carranza 2000, 2002; Harris et al. 2004a,b). If the
species becomes restricted to low altitude urban
areas the impact on native invertebrate species
will be minimal. Inconclusive evidence via
anectodical reports suggests that Tarentola
mauritanica is limited to sites under anthropogenic influence mainly on low altitude urban
place. The antropophilic nature of this species is
suggested by Hódar (2002). Recent reports of live
specimen from Faial and São Miguel islands, in
the Azores (photos and pers. comm. to P. Borges)
show specimen apparently well adapted to coastal
and urban environments such as airports and
industrial areas. Only one sighting is confirmed
from higher altitudes (Pico Salomão, S. Miguel
Island, pers. comm. to P. Borges).
ACKNOWLEDGEMENTS
The authors are grateful to people who are getting
interested and have given information about the
presence of the Moorish gecko in the Azores
islands, mainly São Miguel, Terceira and Faial.
Map generated by Dinis Pereira, C-CMMG;
CITA-A.
REFERENCES
Carranza, S., Arnold, E.N., Mateo, J.A. & L.F. LopezJurado 2000. Long-distance colonization and
radiation in gekkonid lizards, Tarentola (Reptilia:
Gekkonidae), revealed by mitochondrial DNA
sequences. Proceedings of the Royal Society B
267(1444), 637-649.
Carranza, S., Arnold, E.N., Mateo, J.A. & P. Geniez
2002. Relationships and evolution of the North
African geckos, Geckonia and Tarentola (Reptilia:
Gekkonidae), based on mitochondrial and nuclear
DNA sequences. Molecular Phylogenetics and
Evolution 23(2), 244-256.
Báez, M. & M. Biscoito 1993. First record of Tarentola
mauritanica (Linneus, 1758) from the island of
Madeira. Macaronesian Congress, 1993.
Harris, D.J., V. Batista, P. Lymberakis & M.A.
Carretero
2004a.
Complex
estimates
of
evolutionary relationships in Tarentola mauritanica
(Reptilia: Gekkonidae) derived from mitochondrial
DNA sequences. Molecular Phylogenetics and
Evolution, 30(3): 855-859.
Harris, D.J., Batista, V., Carretero, M.A. & N. Ferrand
2004b. Genetic variation in Tarentola mauritanica
(Reptilia: Gekkonidae) across the Strait of
Gibraltar derived from mitochondrial and nuclear
DNA sequences. Amphibia-Reptilia 25(4), 451459.
Hódar, J.A. 2002. Tarentola mauritanica. Pp. 188-190
in: Plezeguelos, J.M., Márquez, R. & Lizana M.
(Eds.). Atlas y Libro Rojo de los Anfíbios y Reptiles
de España. Vol. II, Dirección General de
Conservación de la Naturaleza – Asociación
Herpetologica Española, Madrid.
Loveridge, A. 1947. Revision of the African Lizards of
the Family Gekkonidae. Bulletin of the Museum of
Comparative Zoology 98: 1-469.
Mahrdt, C.R. 1998. Geographic Distribution of
Tarentola mauritanica. Herpetological Review
29(1): 52.
Martínez-Rica, J.P. 1997. Tarentola mauritanica. Pp.
202–204 in Pleguezuelos, J.M. & Martínez-Rica
J.P. (Eds). Distribución y biogeografía de los
Anfibios y Reptiles de España y Portugal.
Monografías de Herpetología 3. Universidad de
Granada-AHE, Granada.
Perera, A., V. Batista & D.J. Harris 2008. Tarentola
mauritanica. Pp. 136-137 in: Loureiro, A., Ferrand
de Almeida, N., Carretero, M.A. & Paulo O.S.
(Eds). Atlas dos Anfíbios e Répteis de Portugal.
Instituto da Conservação da Natureza e
Biodiversidade, Lisboa.
Perera, A., V. Batista & D.J. Harris 2010. Tarentola
mauritanica. Pp. 132-133 in: Loureiro, A., Ferrand
de Almeida, N., Carretero, M.A. & Paulo O.S.
(Eds). Atlas dos Anfíbios e Répteis de Portugal.
Esfera do Caos editores, Lisboa.
Pleguezuelos, J., El Din, S.B. & I. Martínez-Solano
2008. Tarentola mauritanica: IUCN 2010. IUCN
Red List of Threatened Species. Version 2010.1.
Accepted 7 June 2010.
75
EDITORIAL NOTES
ARQUIPÉLAGO – Life and Marine Sciences
Thirty years of natural sciences in the Atlantic Ocean
The journal ARQUIPÉLAGO of the University of the Azores was launched in 1980, divided into
several Series. The precursor for the present Arquipélago – Life and Marine Sciences was the Series of
Natural Sciences (Série Ciências da Natureza) with a varied scope including geosciences, agricultural
sciences, extending into physics, mathematics, etc. The articles were in Portuguese, French, English
and German, with Portuguese as the dominant language. In 1990 (no. 8) the series was named “Life and
Earth Sciences”, biology and geology of the Azores and other Macaronesian islands and the
surrounding ocean were covered within its scope. The journal received a new image with a cover
representing the 9 Azorean islands. With no. 11 (1993) the series was split into A (Life and Marine
Sciences) and B (Geosciences). However, series B was never published and the letter A was eventually
omitted. From no.10 (1992) all the articles have been in English and Portuguese abstracts were
excluded from no. 18 (2001).
From volume 24 (2007) the cover design adopted a full colour image related to one of the articles in the
volume and the Portuguese subtitle was omitted. The journal was simultaneously published on paper
and online, accessible free of charge, and previous volumes become available online in 2010. Its scope
was widened to include all oceanic islands and seamounts of the Atlantic Ocean.
The first supplement of Arquipélago – Life and Marine Sciences, was edited in 1997, presenting a
Check-list of Marine Fishes of the Azores. To date, 7 supplements have been published, mainly
abstracts and proceedings from science meetings and workshops.
The annual volumes have covered 257 articles and short communications during these 30 years. More
than 200 peer reviewers from 27 different countries have been engaged in the process. The Editorial
Secretariat is based in Horta at the Department of Oceanography and Fisheries while the Editorial
Committee has one representative from each of the three University campuses, on three different
islands. The Advisory Board consists of 10 scientists from Europe and USA, specialists in their fields.
Series covers since its inception: from left to right 1980-1985; 1990-1992; 1993-2006;
1997 (first supplement); and 2007.
77
EDITORIAL NOTES
We acknowledge with gratitude the financial support given by Fundação para a Ciência e Tecnologia
(FCT), Lisboa and Secretaria Regonal da Ciência Tecnologia e Equipamentos (SRCTE), Ponta
Delgada.
The Editor of Arquipélago – Life and Marine Sciences, sincerely thanks the scientist listed below who
served as reviewers for the numbers 25 and 26.
25:
António Bivar Sousa, Portugal
Martins Søndergaard, Denmark
Sammy De Grave, UK
Anthony Dixon, UK
Stewart B. Peck, Canada
Octavio Paulo, Portugal
Monika Bright, Austria
António Machado, Canary islands, Spain
Frank Howarth, USA
Ilan Karplus, Israel
Eijiro Nishi, Japan
Torben Lauridsen, Denmark
Ruth Barnich, Germany
Charles Fransen, Netherlands
Brian Moss, UK
Christian Rinke, Austria
26:
John Preece, USA
Paul E. Read, USA
Samantha Hughes, Portugal
Maurice Clarke, Ireland
Felipe Artigas, France
Cláudia Piccini, Uruguay
Joël Bried, Azores, Portugal
78
Ehud Spanier, Israel
Kari Lavalli, USA
Erik Sjögren, Sweden
Nike Bianchi, Italy
Hanno Schäfer, UK
Alvaro Altuna, Spain
Brian Moss, UK
EXCHANGE
The journal is distributed on exchange basis through the
editor.
PURCHASE
Separate numbers of the Journal can be purchased from
the publisher.
OFFPRINTS
Authors receive PDF files of their papers. Orders of
paper copies can be sent with the proofs if desired.
INSTRUCTIONS TO AUTHORS
General. MSS should be submitted to the editor
preferably by e-mail ([email protected]). The text
should be ready for printing and carefully checked for
all errors, including linguistic ones, before submission.
English is the recommended language. Choose Font
Times New Roman, size 12, double spaced paragraph,
and insert page numbers. MSS that are not in
accordance with these and the following instructions
will be returned for revision. All MSS will be peerreviewed by two referees prior to their acceptance by
the Editor.
Style. 1) Title and 2) Author's name, followed by 3)
Abstract (not exceeding 200 words), up to five key
words, and (4) the author's name and professional postal
address; 5) The main text; 6) Acknowledgements; 7)
References; 8) Appendices; 9) Tables with legends; and
the 10) List of figures with legends.
For research articles the text (5) must be organized as
follows: Introduction; Material and methods; Results;
Discussion. The Abstract (3) should convey the scope of
the paper, emphasizing the results and conclusions, not
merely describing its contents. Limit the authors’
professional address (4) to official filiations only.
b) Chapter from a book:
O’Dor, R., H. O. Pörtner & R. E. Shadwick 1990. Squid
as elite athletes: locomotory, respiratory, and circulatory
integration. Pp. 481-503 in: Gilbert, D.L., W.J.
Adelman & J.M. Arnold (Eds). Squid As Experimental
Animals. Plenum Press, New York-London. 516 pp.
c) Article from a journal:
Bentley, M.G., P.J.W. Olive, P.R. Garwood & N.H.
Wright 1984. The spawning and spawning mechanism
of Nephtys caeca (Fabricius, 1780) and Nephtys
homebergi Savigny, 1818 (Annelida: Polychaeta).
Sarsia 69: 63-68.
d) Electronic article, from online-only Journal:
Woo, K.L. 2006. Testing Visual Sensitivity to the Speed
and Direction of Motion in Lizards. Journal of
Visualized Experiments [Internet]. Available from:
http://www.jove.com/index/details.stp?id=127 (cited 18
February 2007).
Use ampersand (&) for all joint authorships in the
reference list and for double authorships in the text. Use
“et al.” in the text for references with three or more
authors and in the reference list for more than seven
authors. Avoid multiple citations in text, preferably
including the earliest and most recent work. The citation
list should be ordered alphabetically by author’s name
and year of publication, followed by increasing number
of co-authors. Journal names should be written in full.
Figures and Tables. All figures and large tables should
be presented with Font Times New Roman, and
separately. Photographs should be clear, have high
contrast, and in digital format (tiff or jpg) with 300 dpi,
with at least 15x15 cm. No original figures should be
submitted before the MS has been accepted.
References. (7) Should correspond with the examples
given below:
E-mail Attachment. All text file formats are accepted.
Attachments must be identified with the author’s last
name, the file name and extension, as follows:
Bianchi_text.doc, Bianchi_fig1.tif, Bianchi_table1.xls.
a) Book:
Etgen, W.M. & P.M. Reaves 1978. Dairy Cattle
Feeding and Management (6th edition). John Wiley &
Sons Inc. New York. 638 pp.
For a complete list of format specifications and
instructions to authors consult our website at:
http://www.arquipelago.info
ISSN 0873-4704
ARQUIPÉLAGO - Life and Marine Sciences
No. 27 - 2010
CONTENTS:
PAGE
SERRANO, ARTUR R.M. & PAULO A.V. BORGES
The cave-adapted arthropod fauna from Madeira archipelago
1
WALLENSTEIN, FRANCISCO M., S.D. PERES, E.D. XAVIER & A.I. NETO
Phytobenthic communities of intertidal rock pools in the eastern islands of Azores
and their relation to position on shore and pool morphology
9
RIERA, RODRIGO, JORGE NÚÑEZ & MARÍA DEL CARMEN BRITO
Check-list of interstitial polychaetes from intertidal and shallow subtidal soft bottoms
of Tenerife, Canary Islands
21
PHAM, CHRISTOPHER K. & EDUARDO ISIDRO
Experimental harvesting of juvenile common octopus Octopus vulgaris,
for commercial ongrowing in the Azores
41
FAUSTINO, CLÁUDIA E.S., M.A. SILVA, T.A. MARQUES & L. THOMAS
Designing a shipboard line transect survey to estimate cetacean abundance
off the Azores archipelago
49
LINO, SÍLVIA P.P., EUCLIDES GONÇALVES & JACQUIE COZENS
The loggerhead sea turtle (Caretta caretta) on Sal Island, Cape Verde: nesting activity
and beach surveillance in 2009
59
SHORT COMMUNICATIONS:
ROHÁČEK, JINDŘICH & JAROSLAV STARÝ
First records of some species of Diptera (Insecta) from the Azores
65
WIRTZ, PETER & SAMMY DE GRAVE
Shrimps (Crustacea, Decapoda, Caridea) associated with gorgonians
at the coast of Senegal
69
BARREIROS, JOÃO P., R.B. ELIAS, J. LOURENÇO, E. DIAS & P. BORGES
First records of Tarentola mauritanica (Linnaeus, 1758) (Reptilia; Gekkonidae)
in the Azores
73
EDITORIAL NOTES
77
Download

Arquipélago - Life and Marine Sciences 27 - 2010