Decapod crustacean larval community structure of the submarine canyon off Blanes (NW Mediterranean Sea)

Marta Carreton 1, Joan B. Company 1, Alexandra Boné 1, Guiomar Rotllant 1, Guillermo Guerao 2, Nixon Bahamon 1,3, María Inés Roldán 4, Antonina dos Santos 5, 6

1 Institut de Ciències del Mar, CSIC. Passeig Marítim de la Barceloneta 37-49, 08003 Barcelona, Spain.
(MC) (Corresponding author) E-mail: mcarreton@icm.csic.es. ORCID iD: https://orcid.org/0000-0002-9490-1453
(JBC) E-mail: batista@icm.csic.es. ORCID iD: https://orcid.org/0000-0002-5878-7155
(AB) E-mail: alexandrabonefernandez@gmail.com. ORCID iD: https://orcid.org/0000-0002-7551-8045
(GR) E-mail: guio@icm.csic.es. ORCID iD: https://orcid.org/0000-0002-6692-9678
2 Institut de Recerca i Tecnologia Agroalimentàries (IRTA), Ctra. de Poble Nou 5, 43540 Sant Carles de la Ràpita, Tarragona, Spain.
(GG) E-mail: gguerao@gmail.com. ORCID iD: https://orcid.org/0000-0001-9738-7923
3 Centre d’Estudis Avançats de Blanes (CEAB-CSIC). C/ d’accés a la Cala St. Francesc, 14, 17300 Blanes, Girona, Spain.
(NB) E-mail: n.bahamon@csic.es. ORCID iD: https://orcid.org/0000-0002-5802-7367
4 Laboratori d’Ictiologia Genètica, Universitat de Girona, Maria Aurelia Capmany 40, 17003 Girona, Spain.
(MIR) E-mail: marina.roldan@udg.edu. ORCID iD: https://orcid.org/0000-0003-3936-3522
5 Instituto Português do Mar e da Atmosfera (IPMA), Av. Alfredo Magalhães Ramalho,6, 1495-165 Lisboa, Portugal.
6 CIIMAR – Interdisciplinary Centre of Marine and Environmental Research, Porto, Portugal.
(ADS) E-mail: antonina@ipma.pt. ORCID iD: https://orcid.org/0000-0002-2238-9315

Summary: Decapod crustacean fisheries play a crucial role on the northwestern Mediterranean coast due to their high commercial value. Although knowledge of larval ecology and recruitment dynamics of these species is essential to establish a sustainable fisheries management, they are still poorly known. In this paper, we describe the composition, abundance and distribution of decapod crustacean larvae in the submarine canyon off Blanes (northwestern Mediterranean Sea) during summer thermal stratification conditions. Samples were collected in September 2011 with a multi-net system and a 60-cm bongo net at 22 stations with bottom depths of between 100 and 1800 m. A total of 635 larvae from 60 genera were identified. The most relevant taxa were Aristeus antennatus (7.93 individuals/1000 m3), the family Sergestidae (7.24) and Alpheus glaber (6.78). These three taxa were dominant (>20% of total decapod larvae) at more than half of the stations. Decapod larval communities were found to be richer and more diverse at the canyon head, a finding which could be explained by the higher retention rates when compared with the upstream and downstream walls and the canyon axis.

Keywords: decapod larvae; submarine canyons; Mediterranean Sea; zooplankton; Aristeus antennatus.

Estructura de la comunidad larvaria de crustáceos decápodos del cañón submarino de Blanes (Mediterráneo noroccidental)

Resumen: Las pesquerías de crustáceos decápodos juegan un papel crucial en la costa del Mediterráneo noroccidental debido a su alto valor comercial. Aunque la información sobre la ecología larvaria y las dinámicas de reclutamiento de estas especies es esencial para establecer una gestión sostenible de estas pesquerías, estos aspectos son aún poco conocidos. En este estudio, se describe la composición, abundancia y distribución de las larvas de crustáceos decápodos en el cañón submarino de Blanes (Mediterráneo noroccidental) durante las condiciones de estratificación térmica estival. Las muestras fueron tomadas en septiembre de 2011 con un sistema multi-red y una red bongo de 60 cm de diámetro en 22 estaciones con fondos de entre 100 y 1800 m de profundidad. Un total de 635 larvas de 60 géneros fueron identificadas. Los taxones más relevantes fueron Aristeus antennatus (7.93 individuos/1000 m3), la familia Sergestidae (7.24) y Alpheus glaber (6.78). Estos tres taxones fueron dominantes (>20% del total de larvas de decápodos) en más de la mitad de las estaciones. Se encontró que las comunidades larvarias de decápodos son más ricas y diversas en la cabecera del cañón, lo cual podría explicarse por los altos índices de retención que presenta esta zona en comparación con las paredes del cañón y su eje.

Palabras clave: larvas de decápodos; cañones submarinos; Mediterráneo; zooplancton; Aristeus antennatus.

Citation/Como citar este artículo: Carreton M., Company J.B., Boné A., Rotllant G., Guerao G., Bahamon N., Roldán M.I., dos Santos A. 2020. Decapod crustacean larval community structure of the submarine canyon off Blanes (NW Mediterranean Sea). Sci. Mar. 84(1): 71-82. https://doi.org/10.3989/scimar.04972.17B

Editor: J.A. Cuesta.

Received: July 1, 2019. Accepted: December 3, 2019. Published: January 29, 2020.

Copyright: © 2020 CSIC. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International (CC BY 4.0) License.

Contents

Summary
Resumen
Introduction
Materials and methods
Results
Discussion
Conclusions
Acknowledgements
References

INTRODUCTIONTop

Submarine canyons are known to be hotspots for biodiversity (Gili et al. 2000Gili J.M., Pagès F., Bouillon J., et al. 2000. A multidisciplinary approach to the understanding of hydromedusan populations inhabiting Mediterranean submarine canyons. Deep. Sea Res. Part 1 Oceanogr. Res. Pap. 47: 1513-1533., De Leo et al. 2010De Leo F.C., Smith C.R., Rowden A.A., et al. 2010. Submarine canyons: hotspots of benthic biomass and productivity in the deep sea. Proc. Royal Soc. B.: Biol. Sci. 277: 2783-2792.). Their geomorphology shapes the sediment transport patterns on the continental margin, enhancing productivity as they play the role of particulate and organic matter traps (Granata et al. 1999Granata T.C., Vidondo B., Duarte C.M., et al. 1999. Hydrodynamics and particle transport associated with a submarine canyon off Blanes (Spain), NW Mediterranean Sea. Cont. Shelf Res. 19: 1249-1263.). Previous studies on such structures have shown that particles from shallow coastal adjacent areas can be transported into the greater depths of the canyon axis, where they sink to the bottom due to reduced horizontal advection and higher residence time (Ahumada-Sempoal et al. 2015Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2015. Shelf-slope exchanges and particle dispersion in Blanes submarine canyon (NW Mediterranean Sea): A numerical study. Cont. Shelf. Res. 109: 35-45.). This phenomenon fertilizes the sea bottom and benefits benthic communities.

The western Mediterranean continental margin is cut by several submarine canyons, some of which have been thoroughly studied from a fisheries management perspective. These canyons are important fishing grounds for many fish and crustacean species of both ecological and commercial interest, which also use these areas as spawning, nursery or recruitment zones (Sardà et al. 2009Sardà F., Company J.B., Bahamón N., et al. 2009. Relationship between environment and the occurrence of Aristeus antennatus (Risso, 1816) in the Blanes submarine canyon (NW Mediterranean). Progr. Oceanog. 82: 227-238, Fernandez-Arcaya et al. 2013Fernandez-Arcaya U., Rotllant G., Ramirez-Llodra E., et al. 2013. Reproductive biology and recruitment of the deep-sea fish community from the NW Mediterranean continental margin. Progr. Oceanogr. 118: 222-234.). Among these species, decapod crustaceans are the most valuable group for the local fisheries industry, ranking the highest in selling prices. For instance, the deep-sea blue and red shrimp Aristeus antennatus, targeted by the bottom-trawling fleet in the submarine canyon areas, can account for up to 50% of the fishermen’s associations annual revenue and is subject to a local management plan in the area of Palamós, north of Blanes (BOE 2018Boletín Oficial del Estado (BOE). 2018. Orden APM/532/2018, de 25 de mayo, por la que se regula la pesca de gamba rosada (Aristeus antennatus) con arte de arrastre de fondo en determinadas zonas marítimas próximas a Palamós. 128. Sábado 26 de mayo de 2018(Sección III): 55045-55051.).

Although most decapod crustacean species in the northwestern Mediterranean Sea have been widely studied in their adult form (Maynou et al. 1996Maynou F., Conan G.Y., Cartes J.E., et al. 1996. Spatial structure and seasonality of decapod crustacean populations on the northwestern Mediterranean slope. Limnol. Oceanogr. 41: 113-125., Company and Sardà 2000Company J.B., Sardà F. 2000. Growth parameters of deep-water decapod crustaceans in the Northwestern Mediterranean Sea: a comparative approach. Mar. Biol. 136: 79-90., Company et al. 2003Company J.B., Sardà F., Puig P., et al. 2003. Duration and timing of reproduction in decapod crustaceans of the NW Mediterranean continental margin: Is there a general pattern? Mar. Ecol. Prog. Ser. 261: 201-216.), little is still known about their larval abundance and distribution, and even less in the case of deep-sea species due to the difficulty of sampling in offshore locations. This information is essential to assess the species dispersal and connectivity patterns. In the case of A. antennatus, only five larval stages have been described using morphological methods and confirmed by analysing molecular markers: the first protozoea (PZ I), subsequent to the hatched naupliar phase, protozoea II (PZ II), protozoea III (PZ III), mysis I (M I) and mysis II (M II) (Heldt 1955Heldt J.H. 1955. Contribution a l’étude de la biologie des crevettes pénéïdes Aristaeomorpha foliacea (Risso) et Aristeus antennatus (Risso) (Formes larvaires). Bull. Soc. Sci. Nat. Tunisie VIII 1-2: 1-29., Torres et al. 2014aTorres A.P., dos Santos A., Balbín R., et al. 2014a. Decapod crustacean larval communities in the Balearic Sea (western Mediterranean): Seasonal composition, horizontal and vertical distribution patterns. J. Mar. Syst. 138: 112-126., Carreton et al. 2019Carreton M., Company J.B., Planella L., et al. 2019. Morphological identification and molecular confirmation of the deep-sea blue and red shrimp Aristeus antennatus larvae. PeerJ 7: e6063.). As illustrated in a recent study, the PZ I of A. antennatus is morphologically very similar to that of Gennadas elegans, a small mesopelagic shrimp of no commercial interest, and the correct identification of this stage calls for a combination of morphological and molecular methods (Carreton et al. 2019Carreton M., Company J.B., Planella L., et al. 2019. Morphological identification and molecular confirmation of the deep-sea blue and red shrimp Aristeus antennatus larvae. PeerJ 7: e6063., Carreton et al. unpubl data). First, these PZ I are distinguished morphologically from the rest of Dendrobranchiata PZ I by their smooth carapace, third maxilliped present, and antennule and antenna of similar length (dos Santos and Lindley 2001dos Santos A., Lindley J.A. 2001. Crustacea Decapoda: Larvae II. Dendrobranchiata (Aristeidae, Benthesicymidae, Penaeidae, Solenoceridae, Sicyonidae, Sergestidae, and Luciferidae). ICES Identification Leaflets for Plankton.). They are then analysed through barcoding to differentiate whether they belong to A. antennatus or G. elegans. Although further stages can be clearly identified using only morphological characters, this difficulty with the PZI larvae is a drawback in decapod larval ecology studies in the northwestern Mediterranean Sea, where both species are an important part of the plankton community. In this area, previous studies on zooplankton have mainly focused on brachyuran crab larvae (Abelló 1989Abelló P. 1989. Reproduction and moulting in Liocarcinus depurator (Linnaeus, 1758) (Brachyura: Portunidae) in the Northwestern Mediterranean Sea. Sci. Mar. 53: 127-134., Abelló and Guerao 1999Abelló P., Guerao G. 1999. Temporal variability in the vertical and mesoscale spatial distribution of crab megalopae (Crustacea: Decapoda) in the Northwestern Mediterranean. Estuar. Coast. Shelf Sci. 49: 129-139.) and others have given a broader view of the decapod larval community (Fusté 1987Fusté X. 1987. Distribución de larvas de Crustáceos Decápodos de la costa de Cataluña. Inv. Pesq. 51: 277-284.). It was only recently that a study tackled the larval distribution of decapod crustaceans in commercial fishing grounds, showing an occurrence of 25% of A. antennatus larvae in the Balearic Sea with an average abundance of 0.1 ind./1000 m3 (Torres et al. 2013Torres A.P., dos Santos A., Alemany F., et al. 2013. Larval stages of crustacean species of interest for conservation and fishing exploitation in the western Mediterranean. Sci. Mar. 77: 149-160.). However, to our knowledge no studies have been reported in submarine canyon areas of the Mediterranean, despite their critical role in the reproduction of these species (Sardà et al. 2009Sardà F., Company J.B., Bahamón N., et al. 2009. Relationship between environment and the occurrence of Aristeus antennatus (Risso, 1816) in the Blanes submarine canyon (NW Mediterranean). Progr. Oceanog. 82: 227-238). The recurrent meanders and eddies that affect submarine canyons produce a highly fluctuating hydrodynamic circulation pattern and transport (Ahumada-Sempoal et al. 2013Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2013. Northern Current variability and its impact on the Blanes Canyon circulation: A numerical study. Prog. Oceanogr. 118: 61-70.). Settlement of decapod larvae has also been proved to respond to hydrodynamic conditions (Criales et al. 2013Criales M.M., Zink I.C., Haus B.K., et al. 2013. Effect of turbulence on the behavior of pink shrimp postlarvae and implications for selective tidal stream transport behavior. Mar. Ecol. Prog. Ser. 477: 161-176., 2015Criales M.M., Cherubin L.M., Browder J.A. 2015. Modeling Larval Transport and Settlement of Pink Shrimp in South Florida: Dynamics of Behavior and Tides. Mar. Coast. Fish. 7: 148-176.), so it is particularly interesting to acquire knowledge of larval distribution in these areas.

Blanes submarine canyon incises the continental margin of the northwestern Mediterranean Sea, from 41°N to 41.8°N and from 2.7°E to 3.2°E, off Blanes port. It has a north-south orientation and its width increases with depth until it reaches a maximum of 23 km at 2600 m depth (Canals et al. 2013Canals M., Company J.B., Martín D., et al. 2013. Integrated study of Mediterranean deep canyons: Novel results and future challenges. Prog. Oceanogr. 118: 1-27.). The continental shelf in this area of the eastern Iberian Peninsula is relatively narrow, with the upper canyon located only 4 km from the coastline. It is a deep and narrow canyon with a smooth eastern (upstream) wall and a sharp and abrupt western (downstream) wall. Water circulation in the area is generally cyclonic along the continental slope, as a result of the combination of the Atlantic Water arriving through the Gibraltar Strait and the Northern Current coming from north to south from the Gulf of Lions (Flexas et al. 2008Flexas M.M., Boyer D.L., Espino M., et al. 2008. Circulation over a submarine canyon in the NW Mediterranean. J. Geophys. Res.: Oceans 113: 1-18.). The authors suggest that the speed of the Northern Current starts to increase by the end of summer, entering the canyon, and reaches its maximum in the winter months, when the mixed layer becomes thicker. The interaction of the predominant SW flow with the bottom topography of the canyons may produce transient local mesoscale eddies and anticyclonic gyres over the continental shelf, thus causing “inversions” of the circulation pattern in the area northeastward (Rubio et al. 2005Rubio A., Arnau P.A., Espino M., et al. 2005. A field study of the behaviour of an anticyclonic eddy on the Catalan continental shelf (NW Mediterranean). Progr. Oceanogr. 66: 142-156., Ahumada-Sempoal et al. 2013Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2013. Northern Current variability and its impact on the Blanes Canyon circulation: A numerical study. Prog. Oceanogr. 118: 61-70.). Given the particularities of the geomorphology in the submarine canyon, hydrodynamics is likely to influence the larval distribution and abundance of particles and pelagic life forms such as eggs and larvae by modifying the retention rates and transport dynamics (Ahumada-Sempoal et al. 2013Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2013. Northern Current variability and its impact on the Blanes Canyon circulation: A numerical study. Prog. Oceanogr. 118: 61-70.). The anticyclonic gyres produced over the canyon head may force them to aggregate over the shelf break around 150 m depth (Ahumada-Sempoal et al. 2015Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2015. Shelf-slope exchanges and particle dispersion in Blanes submarine canyon (NW Mediterranean Sea): A numerical study. Cont. Shelf. Res. 109: 35-45.).

The present study aimed to assess the composition, distribution and abundance of the decapod larval community in Blanes submarine canyon under summer thermal stratification conditions. In particular, we expected to find the larvae of the deep-sea blue and red shrimp Aristeus antennatus, as the area is one of its main nursery grounds in the northwestern Mediterranean Sea. The linkage of the community structure with hydrological conditions is discussed.

MATERIALS AND METHODSTop

Zooplankton sampling

Sampling was conducted in the framework of the project “Observation, Analysis and Modelling of the Mediterranean Sea” (OAMMS, CTM2008-03983). The cruise was carried out in Blanes submarine canyon (NW Mediterranean) on board the Spanish R/V Garcia del Cid, from 26 to 29 September 2011. Twenty-two stations were sampled within the study area over bottoms of 123 to 1716 m depth. We used a 60-cm bongo at 19 of the stations and a multiple opening/closing net and environmental sensing system (MOCNESS, Wiebe et al. 1976Wiebe P.H., Burt K.H., Boyd S.H., et al. 1976. A multiple opening/closing net and environmental sensing system for sampling zooplankton. J. Mar. Res. 34: 313-326.) at the remaining three (Fig. 1). Three stations (B1, B8 and B9) were sampled at the canyon head, ten were sampled along the canyon’s upstream wall (B2, B3, B4, B10, B11, B12, B13, MN1, MN2 and MN3), six were sampled along its downstream wall (B14 to B19), and three were sampled along the canyon axis (B5, B6 and B7). Sampling depth and gear were established according to the station bottom depth (Tables 1 and 2).

figure

Full size image

Fig. 1. – Sampling locations in the Blanes submarine canyon area. Canyon zones are indicated by rectangles: canyon head (top), downstream wall (bottom left), canyon axis (bottom centre), and upstream wall (bottom right).

Table 1. – Information on the stations sampled with the Bongo net.

Station Lat. (N) Long. (E) Sampling depth (m) Bottom depth (m)
B1 41.637 2.891 100 123
B2 41.427 2.980 300 840
B3 41.422 2.932 300 1242
B4 41.371 2.932 300 940
B5 41.369 2.872 300 1716
B6 41.425 2.869 300 1360
B7 41.488 2.870 300 1202
B8 41.576 2.826 200 394
B9 41.530 2.835 300 775
B10 41.486 2.932 300 958
B11 41.490 2.992 300 566
B12 41.489 3.055 300 445
B13 41.375 3.000 300 650
B14 41.369 2.823 300 1206
B15 41.371 2.764 200 366
B16 41.418 2.748 200 268
B17 41.417 2.811 300 500
B18 41.479 2.812 300 438
B19 41.482 2.751 120 164

Table 2. – Information on the stations sampled with the MOCNESS sampler.

Station Net Mesh size
(µm)
Initial lat.
(N)
Final lat.
(N)
Initial long.
(E)
Final long.
(E)
Net opening depth (m) Net closing depth (m) Bottom depth (m)
MN01 N8 500 41.436 41.434 3.121 3.121 50 0 381
N7 500 41.438 41.436 3.121 3.121 100 50 393
N6 500 41.440 41.438 3.121 3.121 150 100 400
N5 500 41.442 41.440 3.121 3.121 200 150 419
N4 500 41.444 41.442 3.121 3.121 250 200 445
N3 500 41.447 41.444 3.121 3.121 300 250 459
N2 500 41.451 41.447 3.121 3.121 350 300 436
N1 300 41.487 41.451 3.123 3.121 350 0 436
MN02 N9 300 41.300 41.296 3.122 3.122 100 0 1004
N8 500 41.304 41.300 3.123 3.122 200 100 968
N7 500 41.311 41.304 3.124 3.123 300 200 948
N6 500 41.317 41.311 3.124 3.124 400 300 965
N5 500 41.323 41.317 3.124 3.124 500 400 1000
N4 500 41.328 41.323 3.125 3.124 600 500 1045
N3 500 41.335 41.328 3.124 3.125 800 600 1135
N2 500 41.338 41.335 3.124 3.124 900 800 1224
N1 300 41.387 41.338 3.120 3.124 900 0 1321
MN03 N9 300 41.358 41.355 3.058 3.058 50 0 796
N8 500 41.361 41.358 3.058 3.058 100 50 795
N7 500 41.366 41.361 3.058 3.058 200 100 798
N6 500 41.372 41.366 3.058 3.058 300 200 804
N5 500 41.376 41.372 3.058 3.058 400 300 833
N4 500 41.380 41.376 3.058 3.058 500 400 855
N3 500 41.382 41.380 3.058 3.058 600 500 861
N2 500 41.372 41.382 3.057 3.058 700 600 848
N1 300 41.423 41.372 3.057 3.057 700 0 826

The bongo was used to collect integrated samples from a maximum depth of 300 m to the surface in oblique tow at a speed of 2 kn. The two nets of the bongo carried different mesh sizes, 300 µm and 500 µm, to target different larval stages. These two nets will hereinafter be referred to as B300 and B500, respectively. Both nets were equipped with a flow meter for filtered water volume calculations. The MOCNESS was used to collect stratified samples up to 900 m depth, carrying nine nets in a 1 m2 mouth, mesh size being 300 µm for N1 and N9 and 500 µm for N2 to N8. An integrated sample (N1) was taken from the surface to the maximum programmed sampling depth. Nets N2 to N9 were closed remotely from the vessel so that each stratum covered between 50 and 100 m depth of the water column (Table 2).

Physical data gathering

Information on the water column was gathered with a Teledyne RD Instruments Acoustic Doppler Current Profiler at 76.8 kHz. Reanalysis products from Copernicus Marine Environment Monitoring System were plotted in QGIS to provide a representative sea surface current field. Meteorological and oceanographic data collected from the oceanographic buoy moored at Blanes canyon head and led by the Operational Observatory of the Catalan Sea (OOCS, Bahamon et al. 2011Bahamon N., Aguzzi J., Bernardello R., et al. 2011. The new pelagic Operational Observatory of the Catalan Sea (OOCS) for the multisensor coordinated measurement of atmospheric and oceanographic conditions. Sensors 11: 11251-11272.) of CEAB-CSIC were obtained from the OOCS database.

Larval identification

Morphological identification

All plankton samples were fixed in situ in 5% formaldehyde buffered with sodium tetraborate. The samples were then sorted in the laboratory using a Leica M205 C stereomicroscope, and the decapod crustacean larvae were counted, staged and identified to the lowest possible taxonomical level. We focused on zoeal stages only (protozoea and mysis in the case of the Dendrobranchiata), not accounting for the decapodid stages in the samples. For the identification of the larvae, we followed identification keys by dos Santos and Lindley (2001)dos Santos A., Lindley J.A. 2001. Crustacea Decapoda: Larvae II. Dendrobranchiata (Aristeidae, Benthesicymidae, Penaeidae, Solenoceridae, Sicyonidae, Sergestidae, and Luciferidae). ICES Identification Leaflets for Plankton. and dos Santos and González-Gordillo (2004)dos Santos A., González-Gordillo J.I. 2004. Illustrated keys for the identification of the Pleocyemata (Crustacea: Decapoda) zoeal stages, from the coastal region of south-western Europe. J. Mar. Biol. Assoc. UK 84: 205-227. and the checklist by González-Gordillo et al. (2001)González-Gordillo J., Dos Santos A., Rodriguez A. 2001. Checklist and annotated bibliography of decapod crustacean larvae from the Southwestern European coast (Gibraltar Strait area). Sci. Mar. 65: 275-305..

Molecular identification

For the larval stages that are not clearly distinguished using the available keys and descriptions (dos Santos and Lindley 2001dos Santos A., Lindley J.A. 2001. Crustacea Decapoda: Larvae II. Dendrobranchiata (Aristeidae, Benthesicymidae, Penaeidae, Solenoceridae, Sicyonidae, Sergestidae, and Luciferidae). ICES Identification Leaflets for Plankton., Heldt 1955Heldt J.H. 1955. Contribution a l’étude de la biologie des crevettes pénéïdes Aristaeomorpha foliacea (Risso) et Aristeus antennatus (Risso) (Formes larvaires). Bull. Soc. Sci. Nat. Tunisie VIII 1-2: 1-29., Gurney 1924Gurney R. 1924. Decapod larvae. Part IX. British Antarctic (Terra Nova) Expedition. Nat. Hist. Rep. Terra Nova Exped. 8: 38-200.), that is, the first protozoeal stage of Gennadas spp. and Aristeus antennatus, a molecular analysis was performed in AllGenetics facilities (A Coruña, Spain) to determine the identity of the larvae. The small size of the larvae and their preservation in buffered formaldehyde called for a special approach in which a pool of larvae had to be analysed together to ensure a sufficient DNA concentration. From a total of 52 protozoea I larvae of the type Gennadas/Aristeus identified, we randomly selected 35 individuals from all stations where they were present (B1, B5, B6, B7, B8, B9, B10, B14, B15, B16, B17, B18 and B19, see Fig. 1) and extracted DNA from the pool. DNA isolation was carried out following a protocol for formalin-fixed material (Campos and Gilibert 2012Campos P.F., Gilibert T.M. 2012. DNA extraction from formalin-fixed material. Methods Mol. Biol. 840: 81-85.). Briefly, the formaldehyde-driven protein-DNA cross-links were reversed by treatment in a heated alkaline buffer (120°C for 25 min in a 0.1 M NaOH, 1% SDS solution, pH=12). The hot alkali treatment was followed by a phenol-chloroform extraction. A negative control that contained no sample was processed along with the sample to check for contamination during DNA isolation.

For library preparation, a short 172-base pair fragment of the cytochrome oxidase I (COI) region was amplified using the primers CrustDF1 (5’ GGT CWA CAA AYC ATA AAG AYA TTG G 3’, Steinke et al. 2016Steinke D., Prosser S.W.J., Hebert P.D.N. 2016. DNA barcoding of marine metazoans; In: Bourlat S (ed), Marine Genomics. Methods in Molecular Biology 1452: 155-168. Humana Press, New York, NY.) and CrustDR_int (designed by AllGenetics), to which the Illumina sequencing primer sequences were attached at their 5’ ends. PCRs were carried out in a final volume of 25 µL containing 2.5 µL of template DNA, 0.5 µM of the primers, 12.5 µL of Supreme NZYTaq 2x Green Master Mix (NZYTech) and ultrapure water up to 25 µL. The reaction mixture was incubated as follows: an initial denaturation at 95ºC for 5 min, followed by 25 cycles of 95°C for 30 s, 50°C for 30 s, 72°C for 30 s and a final extension step at 72°C for 10 minutes. Two negative controls that contained no DNA were included to check for contamination during library preparation. The libraries were then purified and sequenced in a fraction of a MiSeq PE300 run (Illumina).

The raw data were demultiplexed, and the forward and reverse reads performed with FLASH (Magoc and Salzberg 2011Magoc T., Salzberg S.L. 2011. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27: 2957-2963.). We used CUTADAPT (Martin 2011Martin M. 2011. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17: 10-12.) and Qiime (Caporaso et al. 2010Caporaso J.G., Kuczynski J., Stombaugh J., et al. 2010. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 7: 335-336.) to preprocess and quality-filter the data, keeping only sequences with a minimum Phred quality score of 20. The sequences were then dereplicated, clustered at a similarity threshold of 100% and sorted. The impact of point mutations and chimeras generated during PCR and sequencing was reduced by applying filters in the bioinformatic pipeline. The sequences were then assigned to operational taxonomic units (OTUs).

For taxonomic assignment, we created a local reference database by combining the COI sequence data available for the order Decapoda in the DDBJ/ENA/GenBank (http://www.insdc.org) and BOLD (http://www.boldsystems.org) databases. The reference database was length-filtered to exclude sequences shorter than 100 base pairs, and identical sequences were collapsed to a single entry. The taxonomic assignment of the OTU table was performed by querying the clustered centroids against the COI reference database with a 90% similarity threshold. Finally, the OTU table was quality-filtered to keep only the OTUs that matched any reference sequence in the reference database. At the end of the filtering process, the 314320 paired-end raw sequences produced a table with 147323 filtered sequences. Illumina paired-end raw data are available at the Sequence Read Archive (Leinonen et al. 2010Leinonen R., Sugawara H., Shumway M. 2010. International Nucleotide Sequence Database Collaboration. The sequence read archive. Nucleic Acids Res. 39 (Database issue): D19-D21.) operated by the National Center for Biotechnology Information under accession number PRJNA539861.

Data analysis

The number of larvae collected at each station was standardised to number of individuals per 1000 m3. These density data were log transformed (log(x+1)) to comply with statistical restrictions. The data were processed using R (R Development Core Team 2018R Development Core Team. 2018. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.) and QGIS (QGIS Development Team 2019QGIS Development Team. 2009. QGIS Geographic Information System. Open Source Geospatial Foundation.). In order to evaluate the specificity of the two different mesh sizes of B300 and B500 for early (zoeas I to IV) and late (zoeas V onward) larval stages, respectively, the average larval stage was calculated for each net and a one-way ANOVA was performed.

A non-metric multidimensional scaling (NMDS) representation of the data was produced with R. We calculated the Shannon-Wiener diversity index, species richness and the Pielou evenness index for each station, and a multivariate permutational analysis of variance using the Bray-Curtis distance matrix of the communities was carried out using the ‘vegan’ package (Oksanen et al. 2017Oksanen J., Blanchet F.J., Friendly M., et al. 2017. vegan: Community Ecology Package. R package version 2.4-4.) in R. Taxa identified up to taxonomic levels higher than genus were used in the spatial representation of the data but not accounted for in the diversity index calculations. Finally, an analysis of similarity was performed to test for community differences between canyon zones and a similarity percentage analysis (SIMPER) was used to determine species or groups that were responsible for the differences between canyon zones.

RESULTSTop

Identification and taxonomic assignment of decapod larvae

The total number of larvae and density by taxon are shown in Tables 3 (bongo samples) and 4 (MOCNESS samples). A total of 635 decapod larvae belonging to 66 taxa were sorted, and 426 of them were identified at least to the genus level by morphological characters. Molecular analyses were used to discriminate between PZI larvae of A. antennatus and G. elegans.

Table 3. – Total number (N), average density of individuals per 1000 m3 with standard deviation and frequency (F) of decapod larvae caught with bongo net.

B300 B500
Taxon N Average density (ind./1000 m3) F (%) N Average density (ind./1000 m3) F (%)
Suborder Dendrobranchiata
Aristeus antennatus 52 8.00 ± 8.71 68.42 2 0.27 ± 1.17 5.26
Sicyonia carinata 2 0.33 ± 1.42 5.26 2 0.36 ± 1.59 5.26
Solenocera membranacea 9 1.17 ± 2.35 21.05 4 0.52 ± 1.37 15.79
Eusergestes arcticus 0 0.00 0.00 8 1.29 ± 4.47 15.79
Sergestes atlanticus 8 1.30 ± 5.69 5.26 0 0.00 0.00
Sergestidae 51 7.38 ± 8.66 73.68 14 1.99 ± 2.54 42.11
Suborder Pleocyemata
Pasiphaea spp. 2 0.24 ± 1.05 5.26 0 0.00 0.00
Acanthephyra spp. 0 0.00 0.00 1 0.27 ± 1.18 5.26
Alpheus glaber 53 6.84 ± 7.49 78.95 41 4.74 ± 5.33 84.21
Hippolyte spp. 0 0.00 0.00 1 0.33 ± 1.44 5.26
Eualus spp. 1 0.12 ± 0.53 5.26 1 0.10 ± 0.45 5.26
Lysmata spp. 9 1.34 ± 4.33 15.79 5 0.90 ± 3.22 10.53
Eualus sp2. 0 0.00 0.00 1 0.18 ± 0.79 5.26
Processa spp. 11 1.50 ± 3.69 26.32 5 0.83 ± 3.18 10.53
Plesionika spp. 4 0.54 ± 1.39 15.79 5 1.09 ± 3.65 10.53
Pontoninae 0 0.00 0.00 1 0.13 ± 0.56 5.26
Pandalidae 0 0.00 0.00 4 0.56 ± 1.42 15.79
Caridea sp1 2 0.44 ± 1.41 10.53 0 0.00 0.00
Caridea sp2 3 0.74 ± 2.65 10.53 0 0.00 0.00
Caridea sp3 2 0.60 ± 2.61 5.26 0 0.00 0.00
Caridea 6 0.87 ± 1.38 31.58 0 0.00 0.00
Scyllaridae 0 0.00 0.00 1 0.12 ± 0.53 5.26
Scyllarus spp. 5 0.62 ± 1.72 15.79 1 0.13 ± 0.58 5.26
Callianassidae 0 0.00 0.00 2 0.39 ± 1.26 10.53
Upogebia spp. 1 0.21 ± 0.91 5.26 1 0.10 ± 0.45 5.26
Galathea spp. 2 0.44 ± 1.94 5.26 0 0.00 0.00
Calcinus tubularis 6 0.96 ± 3.36 10.53 6 1.07 ± 3.43 10.53
Dardanus spp. 21 2.52 ± 3.68 52.63 24 3.13 ± 3.80 57.89
Dardanus arrosor 0 0.00 0.00 1 0.17 ± 0.75 5.26
Anapagurus spp. 1 0.21 ± 0.91 5.26 0 0.00 0.00
Maja sp1 1 0.21 ± 0.91 5.26 0 0.00 0.00
Maja sp2 1 0.21 ± 0.91 5.26 0 0.00 0.00
Achaeus / Inachus spp. 3 0.49 ± 2.13 5.26 0 0.00 0.00
Ebalia spp. 4 0.72 ± 1.89 15.79 6 0.91 ± 1.86 21.05
Calappa granulata 4 0.67 ± 1.50 21.05 2 0.36 ± 1.59 5.26
Pirimela denticulata 2 0.26 ± 0.77 10.53 0 0.00 0.00
Polybiinae 3 0.62 ± 1.59 15.79 0 0.00 0.00
Parthenopidae 8 1.30 ± 5.69 5.26 10 1.62 ± 4.11 26.32
Eriphia verrucosa 1 0.30 ± 1.31 5.26 0 0.00 0.00
Xantho spp. 6 0.92 ± 2.89 15.79 11 1.06 ± 2.51 21.05
Xantho incisus 0 0.00 0.00 1 0.18 ± 0.79 5.26
Pilumnus spp. 1 0.21 ± 0.91 5.26 1 0.18 ± 0.79 5.26
Goneplax rhomboides 1 0.12 ± 0.53 5.26 0 0.00 0.00
Grapsoidea 2 0.34 ± 1.07 0.00 1 0.18 ± 0.79 0.00
Brachyura sp1 1 0.22 ± 0.97 10.53 0 0.00 5.26
Brachyura sp2 1 0.22 ± 0.97 5.26 0 0.00 0.00
Brachyura 1 0.15 ± 0.66 5.26 0 0.00 0.00
Non-identified 5 0.67 ± 1.57 5.26 2 0.27 ± 0.81 0.00
Total 296 44.01 ± 47.20 165 23.44 ± 24.66

Table 4. – Total number (N) of decapod larvae caught with MOCNESS sampler.

Taxon N Average density (ind./1000 m3) F (%)
Suborder Dendrobranchiata
Aristeus antennatus 2 0.79 ± 1.36 33.33
Sergestidae 14 5.08 ± 1.99 100.00
Solenocera membranacea 35 21.66 ± 61.66 100.00
Suborder Pleocyemata
Acanthephyra spp. 1 0.54 ± 0.93 33.33
Alpheus glaber 40 25.67 ± 22.25 100.00
Athanas nitescens 1 0.88 ± 1.52 33.33
Eualus spp. 4 2.39 ± 2.41 66.67
Processa nouveli 1 0.88 ± 1.52 33.33
Processa spp. 12 5.97 ± 10.34 33.33
Plesionika spp. 13 6.77 ± 11.72 33.33
Pandalidae 3 1.07 ± 1.86 33.33
Scyllarus spp. 5 1.68 ± 2.91 33.33
Scyllaridae 2 0.00 0.00
Polycheles typhlops 2 1.22 ± 2.11 33.33
Galathea spp. 3 0.00 0.00
Calcinus tubularis 2 0.79 ± 1.36 33.33
Dardanus spp. 2 0.53 ± 0.93 33.33
Dardanus arrosor 1 0.00 0.00
Nematopagurus longicornis 2 0.00 0.00
Pagurus spp. 1 0.50 ± 0.86 33.33
Inachus spp. 1 0.88 ± 1.52 33.33
Ebalia spp. 3 1.53 ± 2.72 33.33
Ilia nucleus 4 1.13 ± 1.18 66.67
Polybiinae 2 1.53 ± 2.72 33.33
Parthenopidae 6 2.45 ± 2.36 66.67
Geryonidae 1 0.00 0.00
Xantho spp. 4 0.88 ± 1.52 33.33
Goneplax rhomboides 4 3.06 ± 3.99 66.67
Brachyura 1 0.79 ± 1.36 33.33
Non-identified 2 0.00 0.00
Total 174

The pool sample containing the 35 protozoea individuals showed one high-abundance haplotype and 186 additional haplotypes, all belonging to A. antennatus. A search within the unfiltered OTU table, aimed at detecting possible low-abundance sequences of other species that might have been removed during the filtering process, revealed only sequences belonging to A. antennatus. The best match in the reference databases showed a 100% overlap identity with the pool sample sequences with a query cover of 163/172 base pairs.

Decapod larval communities

Both nets B300 and B500 caught mainly shrimp larvae: B300 caught more Dendrobranchiata (43% of the total catch) and B500 more Caridea (42% of the total catch). Due to the low number of individuals caught in B500 and MOCNESS samples, statistical analysis could only be performed with B300 data.

The NMDS representation for B300 data (Fig. 2, stress=0.1734) shows a grouping of the communities separating two spatial regions: one composed of the upstream wall and one composed of the rest of the canyon zones. After applying a Bonferroni correction to a pairwise analysis by canyon zone, the distance between the communities was only significant between the upstream and downstream walls of the canyon (p=0.018). The SIMPER analysis showed that the downstream wall community is characterized by the presence of A. antennatus and larvae of the family Sergestidae, which together explain 43.2% of the dissimilarity from the community of the upstream wall. The relationship between the NMDS ordination and the environmental variables (sea surface temperature and surface chlorophyll a concentration) was not significant.

figure2

Full size image

Fig. 2. – Non-metric multidimensional scaling analysis (stress=0.1734). Polygon colours indicate canyon zones: upstream wall (white), canyon axis (blue), downstream wall (grey), canyon head (red). Station B13, with no decapod larvae, is not shown. Ariant, Aristeus antennatus; sernid, Sergestidae; serspp, Sergia spp.; seratl, Sergestes atlanticus; solmem, Solenocera membranacea; siccar, Sicyonia carinata; alpgla, Alpheus glaber; euaspp, Eualus spp.; lysspp, Lysmata spp.; prospp, Processa spp.; plespp, Plesionika spp.; passpp, Pasiphaea spp.; carnid, unidentified Caridea; darspp, Dardanus spp.; caltub, Calcinus tubularis; anaspp, Anapagurus spp.; galspp, Galathea spp.; scyspp, Scyllarus spp.; upospp, Upogebia spp.; pornid, Portunidae; pisnid, Pisidae; inanid, Inachidae; xthnid, Xanthoidea; xannid, Xanthidae; xanspp, Xantho spp.; parnid, unidentified Parthenopidae; gonrho, Goneplax rhomboides; ebaspp, Ebalia spp.; pilspp, Pilumnus spp.; calgra, Calappa granulata; eriver, Eriphia verrucosa; moncou, Monodaeus couchii; branid, unidentified Brachyura. Data from B300.

Average larval stages of decapod larvae were significantly different in B300 (3.75) and B500 (6.26) (p=0.003). Early larval stages (I to IV) were extracted from B300 and late stages (VI to IX) from B500, and are represented separately for comparison in Figure 3. Both early and late stages show a higher aggregation at the canyon head and downstream wall, although early stages show higher density values.

figure3

Full size image

Fig. 3. – Density of early (B300, stages I to III) and late (B500, stages VI to decapodid) larval stages of all decapod larvae in individuals (ind.) per 1000 m3.

Decapod larvae density and distribution

Total decapod larval density values ranged from 5.79 (station B4, located at the upstream canyon wall) to 188.99 ind./1000 m3 (station B8, located at the canyon head). The decapod larva distribution in the canyon is shown in Figure 4. The highest average density (7.93 ind./1000 m3) corresponded to the first protozoeal stage of the deep-sea blue and red shrimp Aristeus antennatus. The second most abundant larvae were protozoeas of the family Sergestidae (7.24 ind./1000 m3), which includes the very common small mesopelagic shrimp Eusergestes arcticus. The larvae of the mud shrimp Alpheus glaber were the third most abundant species caught (average density 6.78 ind./1000 m3), followed at a considerable distance by the larvae of the hermit crab taxa, Dardanus spp. These three most abundant taxa were also the most frequent, showing frequencies higher than 60%. Their spatial distribution, presented in Figure 5, shows that both A. antennatus and the Sergestidae are more abundant at the canyon head and the downstream wall. On the other hand, Alpheus glaber larvae are found more abundantly at the canyon head and upstream wall.

figure4

Full size image

Fig. 4. – Total decapod larvae density in individuals (ind.) per 1000 m3 for B300.

figure5

Full size image

Fig. 5. – Larval density of most frequent species (individuals/1000 m3) for B300. A, Aristeus antennatus.; B, Sergestidae; C, Alpheus glaber.

Total decapod larval density and Shannon diversity index were significantly higher at the canyon head than the upstream wall (Fig. 6A-B; p=0.003 and p=0.037, respectively). Likewise, species richness was significantly higher at the canyon head than in all the other canyon zones (Fig. 6C; p=0.008). Species evenness showed no significant differences across canyon zones (Fig. 6D). Excluding station B3, which featured only one species of decapod larvae, species evenness was between 0.76 and 0.97. No decapod larvae were found at station B13. During this particular bongo tow, a large colony of medusae was observed at the surface.

figure6

Full size image

Fig. 6. – Box and whisker diagrams showing median, third quartiles and minimum and maximum values by zone for A, decapod larval density (log(ind./1000 m3)+1); B, Shannon diversity index (H); C, species richness (S); D, Pielou evenness index (J). Data from B300.

As for the stratified MOCNESS samples, the larvae were aggregated in the first 100 m of the water column (Fig. 7). The families Alpheidae and Solenoceridae were the most abundant throughout the MOCNESS samples. One A. antennatus PZ I larva was caught at station MN03 between 50 and 100 m depth. Two rare Polycheles typhlops (Polychelidae) zoea II larvae were found between 300 and 500 m depth at station MN03.

figure7

Full size image

Fig. 7. – Abundance of decapod larvae at MOCNESS stations MN01, MN02 and MN03 in relation to sampling depth strata.

Atmospheric and surface water characteristics

An anticyclonic gyre governed local hydrodynamics over the Blanes canyon on the days of the sample collection, as reported by the reanalysis product (Fig. 8). According to these data, the anticyclonic gyre was weak (~0.10 m s–1), approximately 30 km wide, and covered the upper 50 m depth. According to the OOCS database and the oceanographic buoy, the average daily current was stronger than that of the reanalysis, with values between 0.35 and 0.22 m s–1 from the surface to 50 m depth. The current speed fell from 0.12 m s–1 to 0.06 m s–1 from 60 to 100 m depth. The NE direction of the current was consistent with that of the reanalysis (60°-80°) up to 60 m depth, and then deviated SE (130°-170°) between 70 and 100 m depth. The database reported an average air temperature of 22.6°C, an atmospheric pressure of 1023 hPa and average southward winds of 2.8 m s–1.

figure8

Full size image

Fig. 8. – Average surface current velocity from reanalysis
(http://marine.copernicus.eu/) during the month of September 2011. Red dots represent the sampled stations. Black arrows represent the direction of the current.

DISCUSSIONTop

This study offers initial information on the larval distribution of A. antennatus in the northwestern Mediterranean Sea. The decapod crustacean larval communities described differ from those of previous studies in the northwestern Mediterranean Sea (Fusté 1987Fusté X. 1987. Distribución de larvas de Crustáceos Decápodos de la costa de Cataluña. Inv. Pesq. 51: 277-284., Torres et al. 2014aTorres A.P., dos Santos A., Balbín R., et al. 2014a. Decapod crustacean larval communities in the Balearic Sea (western Mediterranean): Seasonal composition, horizontal and vertical distribution patterns. J. Mar. Syst. 138: 112-126.), mainly due to the predominance of larvae of the deep-sea blue and red shrimp Aristeus antennatus in our study. Our results support the hypothesis that the Blanes submarine canyon is an important spawning area for A. antennatus, because its first protozoeal stage, which occurs shortly after spawning, was found to be the dominant taxon in the decapod larval community at the end of the reproductive period of the species, i.e. late September. In addition, the occurrence of one PZ I larva of the species below the upper 70 m where they have been reported in earlier studies (e.g. Heldt 1955Heldt J.H. 1955. Contribution a l’étude de la biologie des crevettes pénéïdes Aristaeomorpha foliacea (Risso) et Aristeus antennatus (Risso) (Formes larvaires). Bull. Soc. Sci. Nat. Tunisie VIII 1-2: 1-29., Carbonell et al. 2010Carbonell A., dos Santos A., Alemany F., et al. 2010. Larvae of the red shrimp Aristeus antennatus (Decapoda: Dendrobranchiata: Aristeidae) in the Balearic Sea: new occurrences fifty years later. Mar. Biodivers. Rec. 3: E103., Carreton et al. 2019Carreton M., Company J.B., Planella L., et al. 2019. Morphological identification and molecular confirmation of the deep-sea blue and red shrimp Aristeus antennatus larvae. PeerJ 7: e6063.) shows the importance of performing more stratified samplings to further evaluate the vertical distribution pattern of these larvae. These findings also support the model proposed by Carbonell et al. (2010)Carbonell A., dos Santos A., Alemany F., et al. 2010. Larvae of the red shrimp Aristeus antennatus (Decapoda: Dendrobranchiata: Aristeidae) in the Balearic Sea: new occurrences fifty years later. Mar. Biodivers. Rec. 3: E103., which considers that the first protozoeal stage performs an ontogenetic migration through the water column to surface layers, where the subsequent larval stages will develop.

The larvae of the family Sergestidae have been cited as a predominant taxon in the plankton of offshore waters of the northwestern Mediterranean throughout the year (Torres et al. 2014aTorres A.P., dos Santos A., Balbín R., et al. 2014a. Decapod crustacean larval communities in the Balearic Sea (western Mediterranean): Seasonal composition, horizontal and vertical distribution patterns. J. Mar. Syst. 138: 112-126., Simão et al. 2014Simão D.S., Torres A.P. Olivar M.P., et al. 2014. Vertical and temporal distribution of pelagic decapod crustaceans over the shelf-break and middle slope in two contrasting zones around Mallorca (western Mediterranean Sea). J. Mar. Syst. 138: 139-149.). Some genera of this family, particularly Eusergestes arcticus, are accidentally caught in the nets of bottom trawlers in the area (Gorelli et al. 2016Gorelli G., Blanco M., Sardà F., et al. 2016. Spatio-temporal variability of discards in the fishery of the deep-sea red shrimp Aristeus antennatus in the northwestern Mediterranean Sea: implications for management. Sci. Mar. 80: 79-88.). The presence of protozoeal stages of these species in the submarine canyon indicates proximity to the spawning grounds. The mud shrimp Alpheus glaber is very common in Mediterranean waters, and ovigerous females of the species can be found in March and August (Zariquiey-Álvarez 1968Zariquiey Álvarez, R. 1968. Crustáceos decápodos ibéricos. Inv. Pesq. 32: 1-510.). Its larvae have been previously found in the northwestern Mediterranean (Fusté 1987Fusté X. 1987. Distribución de larvas de Crustáceos Decápodos de la costa de Cataluña. Inv. Pesq. 51: 277-284., Torres et al. 2014aTorres A.P., dos Santos A., Balbín R., et al. 2014a. Decapod crustacean larval communities in the Balearic Sea (western Mediterranean): Seasonal composition, horizontal and vertical distribution patterns. J. Mar. Syst. 138: 112-126.), although in lower numbers than those obtained in the present study. The two Polycheles typhlops larvae from this study were identified through molecular methods, described, and included in a phylogenetic analysis (Torres et al. 2014bTorres A.P., Palero F., Dos Santos A., et al. 2014b. Larval stages of the deep-sea lobster Polycheles typhlops (Decapoda, Polychelida) identified by DNA analysis: Morphology, systematic, distribution and ecology. Helgol. Mar. Res. 68: 379-397.).

The highest diversity and species richness found in the submarine canyon head supports the idea that submarine canyons are key structures for many species, which aggregate in these areas for spawning (Sardà et al. 2009Sardà F., Company J.B., Bahamón N., et al. 2009. Relationship between environment and the occurrence of Aristeus antennatus (Risso, 1816) in the Blanes submarine canyon (NW Mediterranean). Progr. Oceanog. 82: 227-238, Fernandez-Arcaya et al. 2013Fernandez-Arcaya U., Rotllant G., Ramirez-Llodra E., et al. 2013. Reproductive biology and recruitment of the deep-sea fish community from the NW Mediterranean continental margin. Progr. Oceanogr. 118: 222-234.). The upstream and downstream walls of the canyon host significantly different communities of decapod larvae. Aside from the influence of the local circulation pattern, this distribution could be explained by a patchy distribution of zooplankton but, especially in the case of the early zoeal stages, it also provides information on the habitat preferences of the species during the spawning period. A paired study of decapod adult and larva distribution is desirable to account for dispersal capacity and habitat preferences of the different life stages of the species.

Modelling exercises of particle dispersal carried out in previous studies (Ahumada-Sempoal et al. 2015Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2015. Shelf-slope exchanges and particle dispersion in Blanes submarine canyon (NW Mediterranean Sea): A numerical study. Cont. Shelf. Res. 109: 35-45.) have suggested that particles released upstream (on the eastern wall of the canyon) show greater retention rates over the canyon head than over its surroundings. In these theoretical experiments, retention of around 40% of the particles takes place at depths shallower than 100 m when the particles are released above the shelf break at 150 m depth. In our study, considering the water circulation described for Blanes canyon (Flexas et al. 2008Flexas M.M., Boyer D.L., Espino M., et al. 2008. Circulation over a submarine canyon in the NW Mediterranean. J. Geophys. Res.: Oceans 113: 1-18.), the general abundance values were higher at the stations that follow the main path of the Northern Current: B12, B11, B8, B19, B16 and B14.

The distribution pattern of the larvae was similar for early and late larval stages, but densities were higher overall for the former, probably due to their high mortality. A similar distribution pattern throughout larval stages indicates a good larval retention in the area. In addition, the circulation pattern in the canyon probably favoured the retention of larvae from adult individuals inhabiting the area as opposed to an import of larvae from nearby areas. The vertical distribution of larvae follows the depth distribution of summer chlorophyll a in the northwestern Mediterranean, with a deep chlorophyll maximum located above 100 m depth according to in situ historical data from the oceanographic buoy of the OOCS located at Blanes canyon head (Bahamon et al. 2011Bahamon N., Aguzzi J., Bernardello R., et al. 2011. The new pelagic Operational Observatory of the Catalan Sea (OOCS) for the multisensor coordinated measurement of atmospheric and oceanographic conditions. Sensors 11: 11251-11272.). A study including more stratified (MOCNESS) samples and thorough in situ oceanographic data needs to be conducted to further explore the influence of water mass circulation in decapod larva distribution.

We expected relatively high densities of decapod larvae in our area due to summer stratification conditions and the role of the canyon as a retention zone (Ahumada-Sempoal et al. 2015Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2015. Shelf-slope exchanges and particle dispersion in Blanes submarine canyon (NW Mediterranean Sea): A numerical study. Cont. Shelf. Res. 109: 35-45.), but the overall larval density values obtained for our study were much lower than those of previous studies in nearby areas (Table 5). These values may be related to the sampling period being at the end of September, when most decapod species have ended their reproduction period and larvae have already settled and/or recruited to the adult habitats. The relatively low total density of crustacean larvae may also be related to the presence of a transient anticyclonic gyre, as reported by the oceanographic reanalysis data. According to Ahumada-Sempoal et al. (2015)Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2015. Shelf-slope exchanges and particle dispersion in Blanes submarine canyon (NW Mediterranean Sea): A numerical study. Cont. Shelf. Res. 109: 35-45., this gyre affects the area during the summer months and is formed over the slope as a consequence of a counter-current flowing northeastwards, parallel to the coast (NE direction), opposed to the southward offshore current generated by northern winds. Meanders and anticyclonic eddies are recurrent in Blanes canyon, producing high fluctuations in the three-dimensional circulation patterns within the canyon (Flexas et al. 2008Flexas M.M., Boyer D.L., Espino M., et al. 2008. Circulation over a submarine canyon in the NW Mediterranean. J. Geophys. Res.: Oceans 113: 1-18.). Transient anticyclonic gyres of this sort may form over the canyon head, altering the general distribution pattern of particles and living organisms such as eggs and larvae (Ahumada-Sempoal et al. 2015Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2015. Shelf-slope exchanges and particle dispersion in Blanes submarine canyon (NW Mediterranean Sea): A numerical study. Cont. Shelf. Res. 109: 35-45.).

Table 5. – Comparison of up-to-date studies on decapod larval communities in the western Mediterranean Sea.

Publication Study area Time of year Bottom depth (m) Sampling method Average decapod larval density (ind./1000 m3)
Fusté 1987Fusté X. 1987. Distribución de larvas de Crustáceos Decápodos de la costa de Cataluña. Inv. Pesq. 51: 277-284. Catalan coast (shelf) September 50 200 Bongo net (integrated) 10000
Torres et al. 2014aTorres A.P., dos Santos A., Balbín R., et al. 2014a. Decapod crustacean larval communities in the Balearic Sea (western Mediterranean): Seasonal composition, horizontal and vertical distribution patterns. J. Mar. Syst. 138: 112-126. Balearic Islands July 250 900 MOCNESS (stratified) 144.5 (Northern area) 85.3 (Southern area)
This study Catalan margin (Blanes submarine canyon) September 150-1700 Bongo net (integrated) 23.38

Larval ecology remains one of the main knowledge gaps in marine science, and additional studies are still needed before we can move on to further analysis of the communities. In coastal species, biophysical models have helped understand the influence of oceanic currents and tides in the transport and settlement of decapod larvae (e.g. Pires et al. 2013Pires R.F.T., Pan M., Santos A.M.P., et al. 2013. Modelling the variation in larval dispersal of estuarine and coastal ghost shrimp: Upogebia congeners in the Gulf of Cadiz. Mar. Ecol. Prog. Ser. 492: 153-168., Criales et al. 2015Criales M.M., Cherubin L.M., Browder J.A. 2015. Modeling Larval Transport and Settlement of Pink Shrimp in South Florida: Dynamics of Behavior and Tides. Mar. Coast. Fish. 7: 148-176.). In the case of deep-sea species larvae, once evidence of larval presence has been reported, a reasonable next step is to explore the combined potential of stratified field samplings, oceanographic data and biophysical modelling to explain their distribution.

CONCLUSIONSTop

The decapod larval community of Blanes canyon at the end of September is dominated by three taxa: the deep-sea blue and red shrimp Aristeus antennatus, the family Sergestidae and the mud shrimp Alpheus glaber. Decapod larval communities are richer and more diverse at the canyon head than on the upstream and downstream walls and in the canyon axis.

ACKNOWLEDGEMENTSTop

This research was carried out within project OAMMS (CTM2008-03983), financed by the Spanish Ministerio de Economía y Competitividad. Molecular analyses were done by AllGenetics (A Coruña, Spain).

Data availability

The datasets generated and/or analysed during the current study are available from the corresponding author on reasonable request.

Conflict of interest

During the writing of this manuscript, MC was supported by a grant from the Spanish Ministerio de Educación (FPU2015). All authors declare that we have no conflict of interest with funding sources or with the submission of this manuscript.

REFERENCESTop

Abelló P. 1989. Reproduction and moulting in Liocarcinus depurator (Linnaeus, 1758) (Brachyura: Portunidae) in the Northwestern Mediterranean Sea. Sci. Mar. 53: 127-134.

Abelló P., Guerao G. 1999. Temporal variability in the vertical and mesoscale spatial distribution of crab megalopae (Crustacea: Decapoda) in the Northwestern Mediterranean. Estuar. Coast. Shelf Sci. 49: 129-139.
https://doi.org/10.1006/ecss.1999.0488

Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2013. Northern Current variability and its impact on the Blanes Canyon circulation: A numerical study. Prog. Oceanogr. 118: 61-70.
https://doi.org/10.1016/j.pocean.2013.07.030

Ahumada-Sempoal M.A., Flexas M.M., Bernardello R., et al. 2015. Shelf-slope exchanges and particle dispersion in Blanes submarine canyon (NW Mediterranean Sea): A numerical study. Cont. Shelf. Res. 109: 35-45.
https://doi.org/10.1016/j.csr.2015.09.012

Bahamon N., Aguzzi J., Bernardello R., et al. 2011. The new pelagic Operational Observatory of the Catalan Sea (OOCS) for the multisensor coordinated measurement of atmospheric and oceanographic conditions. Sensors 11: 11251-11272.
https://doi.org/10.3390/s111211251

Boletín Oficial del Estado (BOE). 2018. Orden APM/532/2018, de 25 de mayo, por la que se regula la pesca de gamba rosada (Aristeus antennatus) con arte de arrastre de fondo en determinadas zonas marítimas próximas a Palamós. 128. Sábado 26 de mayo de 2018(Sección III): 55045-55051.

Canals M., Company J.B., Martín D., et al. 2013. Integrated study of Mediterranean deep canyons: Novel results and future challenges. Prog. Oceanogr. 118: 1-27.
https://doi.org/10.1016/j.pocean.2013.09.004

Campos P.F., Gilibert T.M. 2012. DNA extraction from formalin-fixed material. Methods Mol. Biol. 840: 81-85.
https://doi.org/10.1007/978-1-61779-516-9_11

Carbonell A., dos Santos A., Alemany F., et al. 2010. Larvae of the red shrimp Aristeus antennatus (Decapoda: Dendrobranchiata: Aristeidae) in the Balearic Sea: new occurrences fifty years later. Mar. Biodivers. Rec. 3: E103.
https://doi.org/10.1017/S1755267210000758

Carreton M., Company J.B., Planella L., et al. 2019. Morphological identification and molecular confirmation of the deep-sea blue and red shrimp Aristeus antennatus larvae. PeerJ 7: e6063.
https://doi.org/10.7717/peerj.6063

Company J.B., Sardà F. 2000. Growth parameters of deep-water decapod crustaceans in the Northwestern Mediterranean Sea: a comparative approach. Mar. Biol. 136: 79-90.
https://doi.org/10.1007/s002270050011

Company J.B., Sardà F., Puig P., et al. 2003. Duration and timing of reproduction in decapod crustaceans of the NW Mediterranean continental margin: Is there a general pattern? Mar. Ecol. Prog. Ser. 261: 201-216.
https://doi.org/10.3354/meps261201

Caporaso J.G., Kuczynski J., Stombaugh J., et al. 2010. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 7: 335-336.
https://doi.org/10.1038/nmeth.f.303

Criales M.M., Zink I.C., Haus B.K., et al. 2013. Effect of turbulence on the behavior of pink shrimp postlarvae and implications for selective tidal stream transport behavior. Mar. Ecol. Prog. Ser. 477: 161-176.
https://doi.org/10.3354/meps10141

Criales M.M., Cherubin L.M., Browder J.A. 2015. Modeling Larval Transport and Settlement of Pink Shrimp in South Florida: Dynamics of Behavior and Tides. Mar. Coast. Fish. 7: 148-176.
https://doi.org/10.1080/19425120.2014.1001541

De Leo F.C., Smith C.R., Rowden A.A., et al. 2010. Submarine canyons: hotspots of benthic biomass and productivity in the deep sea. Proc. Royal Soc. B.: Biol. Sci. 277: 2783-2792.
https://doi.org/10.1098/rspb.2010.0462

dos Santos A., González-Gordillo J.I. 2004. Illustrated keys for the identification of the Pleocyemata (Crustacea: Decapoda) zoeal stages, from the coastal region of south-western Europe. J. Mar. Biol. Assoc. UK 84: 205-227.
https://doi.org/10.1017/S0025315404009075h

dos Santos A., Lindley J.A. 2001. Crustacea Decapoda: Larvae II. Dendrobranchiata (Aristeidae, Benthesicymidae, Penaeidae, Solenoceridae, Sicyonidae, Sergestidae, and Luciferidae). ICES Identification Leaflets for Plankton.

Fernandez-Arcaya U., Rotllant G., Ramirez-Llodra E., et al. 2013. Reproductive biology and recruitment of the deep-sea fish community from the NW Mediterranean continental margin. Progr. Oceanogr. 118: 222-234.
https://doi.org/10.1016/j.pocean.2013.07.019

Flexas M.M., Boyer D.L., Espino M., et al. 2008. Circulation over a submarine canyon in the NW Mediterranean. J. Geophys. Res.: Oceans 113: 1-18.
https://doi.org/10.1029/2006JC003998

Fusté X. 1987. Distribución de larvas de Crustáceos Decápodos de la costa de Cataluña. Inv. Pesq. 51: 277-284.

Gili J.M., Pagès F., Bouillon J., et al. 2000. A multidisciplinary approach to the understanding of hydromedusan populations inhabiting Mediterranean submarine canyons. Deep. Sea Res. Part 1 Oceanogr. Res. Pap. 47: 1513-1533.
https://doi.org/10.1016/S0967-0637(99)00119-3

González-Gordillo J., Dos Santos A., Rodriguez A. 2001. Checklist and annotated bibliography of decapod crustacean larvae from the Southwestern European coast (Gibraltar Strait area). Sci. Mar. 65: 275-305.
https://doi.org/10.3989/scimar.2001.65n4275

Gorelli G., Blanco M., Sardà F., et al. 2016. Spatio-temporal variability of discards in the fishery of the deep-sea red shrimp Aristeus antennatus in the northwestern Mediterranean Sea: implications for management. Sci. Mar. 80: 79-88.
https://doi.org/10.3989/scimar.04237.24A

Granata T.C., Vidondo B., Duarte C.M., et al. 1999. Hydrodynamics and particle transport associated with a submarine canyon off Blanes (Spain), NW Mediterranean Sea. Cont. Shelf Res. 19: 1249-1263.
https://doi.org/10.1016/S0278-4343(98)00118-6

Gurney R. 1924. Decapod larvae. Part IX. British Antarctic (Terra Nova) Expedition. Nat. Hist. Rep. Terra Nova Exped. 8: 38-200.

Heldt J.H. 1955. Contribution a l’étude de la biologie des crevettes pénéïdes Aristaeomorpha foliacea (Risso) et Aristeus antennatus (Risso) (Formes larvaires). Bull. Soc. Sci. Nat. Tunisie VIII 1-2: 1-29.

Leinonen R., Sugawara H., Shumway M. 2010. International Nucleotide Sequence Database Collaboration. The sequence read archive. Nucleic Acids Res. 39 (Database issue): D19-D21.
https://doi.org/10.1093/nar/gkq1019

Magoc T., Salzberg S.L. 2011. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27: 2957-2963.
https://doi.org/10.1093/bioinformatics/btr507

Martin M. 2011. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17: 10-12.
https://doi.org/10.14806/ej.17.1.200

Maynou F., Conan G.Y., Cartes J.E., et al. 1996. Spatial structure and seasonality of decapod crustacean populations on the northwestern Mediterranean slope. Limnol. Oceanogr. 41: 113-125.
https://doi.org/10.4319/lo.1996.41.1.0113

Oksanen J., Blanchet F.J., Friendly M., et al. 2017. vegan: Community Ecology Package. R package version 2.4-4.

Pires R.F.T., Pan M., Santos A.M.P., et al. 2013. Modelling the variation in larval dispersal of estuarine and coastal ghost shrimp: Upogebia congeners in the Gulf of Cadiz. Mar. Ecol. Prog. Ser. 492: 153-168.
https://doi.org/10.3354/meps10488

QGIS Development Team. 2019. QGIS Geographic Information System. Open Source Geospatial Foundation.
http://qgis.osgeo.org

R Development Core Team. 2018. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.
http://www.R-project.org.

Rubio A., Arnau P.A., Espino M., et al. 2005. A field study of the behaviour of an anticyclonic eddy on the Catalan continental shelf (NW Mediterranean). Progr. Oceanogr. 66: 142-156.
https://doi.org/10.1016/j.pocean.2004.07.012

Sardà F., Company J.B., Bahamón N., et al. 2009. Relationship between environment and the occurrence of Aristeus antennatus (Risso, 1816) in the Blanes submarine canyon (NW Mediterranean). Progr. Oceanog. 82: 227-238
https://doi.org/10.1016/j.pocean.2009.07.001

Simão D.S., Torres A.P. Olivar M.P., et al. 2014. Vertical and temporal distribution of pelagic decapod crustaceans over the shelf-break and middle slope in two contrasting zones around Mallorca (western Mediterranean Sea). J. Mar. Syst. 138: 139-149.
https://doi.org/10.1016/j.jmarsys.2013.10.008

Steinke D., Prosser S.W.J., Hebert P.D.N. 2016. DNA barcoding of marine metazoans; In: Bourlat S (ed), Marine Genomics. Methods in Molecular Biology 1452: 155-168. Humana Press, New York, NY.
https://doi.org/10.1007/978-1-4939-3774-5_10

Torres A.P., dos Santos A., Alemany F., et al. 2013. Larval stages of crustacean species of interest for conservation and fishing exploitation in the western Mediterranean. Sci. Mar. 77: 149-160.
https://doi.org/10.3989/scimar.03749.26D

Torres A.P., dos Santos A., Balbín R., et al. 2014a. Decapod crustacean larval communities in the Balearic Sea (western Mediterranean): Seasonal composition, horizontal and vertical distribution patterns. J. Mar. Syst. 138: 112-126.
https://doi.org/10.1016/j.jmarsys.2013.11.017

Torres A.P., Palero F., Dos Santos A., et al. 2014b. Larval stages of the deep-sea lobster Polycheles typhlops (Decapoda, Polychelida) identified by DNA analysis: Morphology, systematic, distribution and ecology. Helgol. Mar. Res. 68: 379-397.
https://doi.org/10.1007/s10152-014-0397-0

Wiebe P.H., Burt K.H., Boyd S.H., et al. 1976. A multiple opening/closing net and environmental sensing system for sampling zooplankton. J. Mar. Res. 34: 313-326.

Zariquiey Álvarez, R. 1968. Crustáceos decápodos ibéricos. Inv. Pesq. 32: 1-510.