Accepted Manuscript Hydrodynamic patterns favouring sea urchin recruitment in coastal areas: A Mediterranean study case S. Farina, G. Quattrocchi, I. Guala, A. Cucco PII:
S0141-1136(17)30779-1
DOI:
10.1016/j.marenvres.2018.05.013
Reference:
MERE 4527
To appear in:
Marine Environmental Research
Received Date: 18 December 2017 Revised Date:
4 May 2018
Accepted Date: 7 May 2018
Please cite this article as: Farina, S., Quattrocchi, G., Guala, I., Cucco, A., Hydrodynamic patterns favouring sea urchin recruitment in coastal areas: A Mediterranean study case, Marine Environmental Research (2018), doi: 10.1016/j.marenvres.2018.05.013. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
1
Hydrodynamic patterns favouring sea urchin recruitment in coastal
2
areas: a Mediterranean study case.
RI PT
3
S. Farina1*, G. Quattrocchi2, I. Guala1 and A. Cucco2
4 5 6
1
7
2
8
Loc. Sa Mardini, Torre Grande, 09170 Oristano, Italy.
SC
IMC- International Marine Centre, Loc. Sa Mardini, Torre Grande, 09170 Oristano, Italy
9 10
*
[email protected]
14 15 16 17 18 19
EP
13
AC C
12
TE D
11
M AN U
IAMC - Institute for Coastal Marine Environment, CNR - National Research Council of Italy,
ACCEPTED MANUSCRIPT
Abstract In invertebrate fisheries, sea urchin harvesting continues to grow with dramatic
21
consequences for benthic ecosystems. The identification of areas with a marked natural
22
recruitment and the mechanisms regulating it is crucial for the conservation of benthic
23
communities and for planning the sustainable harvesting. This study evaluates the spatial
24
distribution and density of recruits of the edible sea urchin Paracentrotus lividus along the Sinis
25
+Peninsula (Sardinia) and explores its significant relationships with the local oceanographic
26
features. Our results reveal that recruitment is favoured in areas with slow currents and high
27
levels of confinement and trapping of the water masses. Analysis of the residual circulation
28
indicates that the presence of local standing circulation structures promotes the sea urchin
29
recruitment process. Our findings emphasize the importance of managing local sea urchin
30
harvesting as a system of populations with their demographic influence mainly dependent on the
31
most important ecological driver that is the recruitment.
32
Key-words
33
Coastal hydrodynamics, larval dispersal, Paracentrotus lividus, population dynamic, recruitment,
34
sea urchin harvesting, ocean modelling
SC
M AN U
TE D
EP AC C
35
RI PT
20
36
1. Introduction
37
In coastal areas, invertebrate catches rapidly increased over the past decades and artisanal
38
fisheries and harvesting are the main factors determining imbalances in trophic interactions
39
between fishes, invertebrates and macroalgae (Jackson et al., 2001). These effects are
40
particularly intense when considering macrobenthic communities where some species of sea
ACCEPTED MANUSCRIPT
41
urchins with a functional role in the ecosystem are both the prey of commercially important fish
42
species and the target of direct human exploitation (Andrew et al., 2002). Often, the uncontrolled proliferation of these herbivores due to the overfishing of the main
44
predators leads to the creation of permanent barrens (Hereu et al., 2008; McClanahan and Sala,
45
1998; Steneck et al., 2004, 2002). However, intensive sea urchin harvesting for the sale of the
46
gonads in some temperate zones results in the rapid development of large brown algae and
47
subsequent changes in the structure of the associated community (Andrew et al., 2002).
SC
RI PT
43
In the Mediterranean Sea, the species Paracentrotus lividus (Lamarck 1816) is one of the
49
most important functional herbivores of the shallow intertidal and sub-tidal rocky habitats
50
(Boada et al., 2017; Hereu, 2004; McClanahan and Sala, 1997; Prado et al., 2007).At the same
51
time, it is a widely appreciated fishing resource with many reported cases of overexploitation
52
(Pais et al., 2007).
TE D
M AN U
48
The systematic removal of adult sea urchins due to harvesting can compromise the
54
reproduction capacity of populations (Levitan and Sewell, 1998; Levitan et al., 1992; Loi et al.,
55
2017; Pennington, 1985; Tegner and Dayton, 1977). As a consequence, larvae originating near a
56
fishing area could be a critical requirement for avoiding the collapse of the local overexploited
57
populations (Levitan et al., 1992; Pennington, 1985). Fertile size classes, i.e. individuals larger
58
than 3 cm test diameter without spines (TD, hereafter) can produce more than one cohort of
59
mature gametes in a single breeding season (Mita et al., 2007). Spawning occurs between
60
January and June, with a continuous reproductive cycle characterized by one or two peaks, in
61
February-March and in May-June (see reviews in Boudouresque and Verlaque, 2001; Ouréns et
62
al., 2011).
AC C
EP
53
ACCEPTED MANUSCRIPT
In general, sea urchin abundance varies widely from region to region, is mainly linked to
64
larval supply and highly associated with oceanographic features (Fenaux et al., 1988; Prado et
65
al., 2012). The planktonic life-stage ranges between 20 to 40 days. During this time the larvae,
66
driven by the currents, can travel great distances until they find a favourable substrate to adhere
67
to (Fenaux et al., 1988; López et al., 1998; Morgan et al., 2000; Prado et al., 2012; Treml et al.,
68
2012). When larvae find suitable environmental conditions, they swim towards the bottom to
69
undergo metamorphosis (Fenaux and Pedrotti, 1988). In this phase, the presence of adults, as
70
well as the type of substrate (e.g. rugosity, presence of crustose algae, macrophyte canopy), are
71
crucial to the settlement’s success (Boudouresque and Verlaque, 2001; Oliva et al., 2016).
72
However, as an extreme survival action, larvae are also able to metamorphose in pelagic
73
environments (Fenaux and Pedrotti, 1988).
M AN U
SC
RI PT
63
Natural mortality (e.g. predation) and migration are then the habitat-specific regulation
75
processes governing the benthic life stage across the seascape (Boada et al., 2018; Farina et al.,
76
2017, 2016, 2014). Predation represents a bottleneck for urchin populations after the settlement
77
and until individuals reach the refuge size of 5 cm TD (Farina et al., 2009; Guidetti et al., 2004;
78
Hereu et al., 2005). Due to the limited mobility of species, migration regulates population density
79
among patches of contiguous habitats, but it cannot be considered a connecting process between
80
neighbouring populations like the recruitment process (Boada et al., 2018; Ceccherelli et al.,
81
2009).
AC C
EP
TE D
74
82
The planktonic life stage of the sea urchin populations makes them demographically open.
83
Coastal circulation patterns, that depend on coastal morphology, affect the spatial distribution of
84
larvae supply (Morgan et al., 2000; Treml et al., 2012). Some studies have demonstrated how the
85
stock-recruitment relationship showed strong benefits from Marine Protected Areas (MPA)
ACCEPTED MANUSCRIPT
species migration (e.g. Moffit et al., 2009). However, the persistence of sea urchin populations
87
under harvesting pressure may also depend on the local hydrodynamics that affect the spatial
88
distribution of larvae.
RI PT
86
On the island of Sardinia (Western Mediterranean Sea), sea urchin populations have suffered
90
high human pressure for decades (Addis et al., 2014; Pais et al., 2012). In the central western
91
coast of the region, despite the specific management measures adopted inside the local MPA -
92
Peninsula of Sinis - Mal di Ventre Island (see methods), the abundance of sea urchin populations
93
continues to show a dramatic decline (Pieraccini et al., 2016). Along the Sinis Peninsula,
94
populations lacking fertile individuals greater than 5 cm TD are common both inside and outside
95
the MPA (Loi et al., 2015; Pieraccini et al., 2016). Thus, the aim of this research work is to
96
provide spatial information about specific hot spots of recruit density, and potential settlement,
97
as well as about the physical processes regulating recruitment process along this stretch of coast
98
(Puckett et al., 2014; Stenevik et al., 2012). This information and further monitoring will provide
99
a solid data set for enhancing ecosystem-based decision-making strategies for conservational
M AN U
TE D
plans along the coast of the Sinis Peninsula (see Fig. 1).
EP
100
SC
89
In the last years, the prominent role of hydrodynamics in driving dispersal and fluxes in
102
planktonic larvae has been widely investigated through numerical modelling techniques.
103
Individual-based biophysical models have been used to analyse the connectivity among marine
104
populations of both fish and invertebrate species (Koeck et al., 2015; Ospina-Alvarezet al., 2015;
105
Medel et al., 2018; Calò, et al., 2018). In the Mediterranean Sea, modelling applications based on
106
the lagrangian approach were devoted to the investigation of the connectivity patterns between
107
MPAs at basin and/or sub-basin levels (Andrello et al., 2013; Rossi et al., 2014; Dubois et al.,
108
2016; Bray et al., 2017). In the specific case of the sea urchin, a recent work by Paterno et al.,
AC C
101
ACCEPTED MANUSCRIPT
2017 explored the connectivity patterns in populations between the Adriatic and the Ionian Sea.
110
In all these studies, the Euclidean distances between the different marine provinces, between
111
spawning and recruitment areas or between each target subpopulation was enough to consider
112
them potentially separated. In fact, in many works, along with the individual-based model
113
application, population genomics were used to validate the modelling results and confirm a
114
possible hypothesis on connectivity (Calò et al., 2016; Paterno et al., 2017).
SC
RI PT
109
In our study case, the investigated area was limited to a few tens of kilometres, where any
116
possible genetic differentiation or physical separation between the local sea urchin population
117
was a priori excluded. Therefore, the use of individual-based models to explore the local
118
connectivity matrix was considered unnecessary. An alternative approach, based on the Eulerian
119
frame of reference, was selected for investigating the local hydrodynamic patterns favouring the
120
sea urchin recruitment process.
AC C
EP
TE D
M AN U
115
121 122 123
Fig. 1. Geographical setting: coastline of the Sinis Peninsula (Central Western coast of Sardinia). This area includes the Marine Protected Area of the Peninsula del Sinis - Isola di Mal
ACCEPTED MANUSCRIPT
124 125 126
di Ventre and a high-pressure zone of sea urchin harvesting to the north of it. Dots indicate sampling sites: Funtana Meiga (1, 2), Seu (3-9), Sa Benda (10), Porto Suedda (11-15), Su Tingiosu (16-18), Putzu Idu (19-21), Su Pallosu (22, 23), Scal’e Sali (24, 25).
127 Specifically, this work includes the assessment of the hydrodynamic patterns favouring
129
larvae settlement and the indication of potential spatial variability of recruit density in relation to
130
different levels of harvesting (inside and outside the MPA). Biological data were obtained from
131
field work carried out along the whole coastal area over a ten-year period. The adopted dataset,
132
that included the sea urchin density distribution of both adult individuals and recruits, was
133
homogenized considering the type of habitat and the depth of the sampling sites.
M AN U
SC
RI PT
128
Moreover, potential relationships among recruit density, spatial variability and oceanographic
135
features were investigated. A specific set of oceanographic variables, representative of the local
136
hydrodynamics, was obtained by means of a numerical modelling application. A Generalized
137
Linear Mixed Model (GLMM) was adopted to compare the spatial variability of the sea urchin
138
density and oceanographic variables in order to detect significant correlations.
TE D
134
2. Study site
140
The study was conducted along a stretch of coast of about 40 km on the western side of
141
Sardinia, Italy (Sinis coastal area, see Fig. 1) between Cape Seu (39.9080° N, 8.3910° E) and Su
142
Pallosu Bay (40.0379° N, 8.3938° E). The offshore ocean circulation is dominated by low-
143
energetic anticyclonic gyres induced by the baroclinic instabilities of the Algerian current system
144
(Millot, 1999; Olita et al., 2013; Schroeder et al., 2012).
AC C
EP
139
145
In the coastal area, the water current is mainly generated by the action of the frequent and
146
intense wind events, mostly from the north-west (Mistral wind) and, to a lesser degree, from the
ACCEPTED MANUSCRIPT
south-west (Libeccio wind), characterized by a yearly mean speed of 7 m/s and a peak speed
148
higher than 20 m/s (Zecchetto et al., 2016). Most of the coastline is exposed to the wind-waves
149
generated on a wide fetch by the prevailing winds that, in the form of severe winter storms, can
150
produce intense wave fields, with a Significant Wave Height of up to 5 m (Simeone et al., 2012a;
151
Simeone et al., 2016). The tidal force is negligible, with very weak amplitudes of around 0.15 m,
152
and doesn’t produce any remarkable current flows (Cucco et al., 2012a).
SC
RI PT
147
This stretch of coast is characterized by highly variable geomorphology and hydrodynamics,
154
with numerous types of substrates and habitat landscapes (Simeone et al., 2012b; De Falco et al.,
155
2015). The middle and southern parts of the area are included in the Peninsula of Sinis - Mal di
156
Ventre Island MPA, which covers a surface of twenty-five thousand hectares (see Fig. 1).
157
However, the surface area that is fully protected is relatively small (529 ha, Guala et al., 2008a),
158
while the remaining zones are intensively frequented by commercial and recreational fishermen
159
(Pieraccini et al., 2016).
TE D
M AN U
153
In Sardinia, sea urchin harvesting is managed by a regional decree (RAS, Regione Autonoma
161
della Sardegna decree no. 276 of March 3, 1994 and subsequent amendments). Currently, along
162
the whole coast of Sardinia, 189 professional fishermen are authorized to collect sea urchins by
163
scuba diving from November to April with each diver allowed to collect up to 2000 sea urchins
164
per day. In the Peninsula of Sinis - Mal di Ventre Island MPA, the management board allows 55
165
resident professional fishermen to harvest 500 to 1000 sea urchins per day for a shorter period of
166
time, varying from year to year. Despite having more restrictive measures than the rest of
167
Sardinia, in the MPA, the abundance of sea urchins of a commercial size showed a significant
168
decrease after 2007 (Pieraccini et al., 2016). Thus, along the Sinis peninsula, inside and outside
AC C
EP
160
ACCEPTED MANUSCRIPT
169
the MPA, a lack of fertile individuals with a TD > 5 cm is common among populations (Loi et
170
al., 2017; Pieraccini et al., 2016). 3. Material and methods
172
3.1 Sea urchin density distribution
173
Abundance and size-frequency distribution of the sea urchin populations were estimated
174
yearly from 2003 to 2012, both within the MPA borders and outside them. The portion outside
175
was situated along the northernmost coastal area of the domain, which traditionally corresponds
176
to a zone of high harvesting (Loi et al., 2017). Sampling was carried out in 25 sites at a depth of
177
between 2 and 5 m on a calcareous substrate. This is the most common rocky habitat along this
178
stretch of coast. The sampling sites were located inside the MPA (sites numbered from 1 to 15)
179
and outside the MPA (sites numbered from 16 to 25) in order to observe the effects of the
180
different regulations and different levels of harvesting pressure. For each site, sea urchin
181
abundance was assessed in 5 m2 quadrats placed randomly for a total of three times. The sizes of
182
the individuals (TD without spines) were measured with calipers to the closest mm.
TE D
M AN U
SC
RI PT
171
We define the recruits of P. lividus as those individuals that survive approximately one year
184
after their settlement. The size of the one-year-old recruits is estimated using their growth rate
185
(about 1 cm/y during their life stage, according to Ouréns et al., 2013). Therefore, we considered
186
the individuals of P. lividus with TD < 1 cm as the recruits and their density (RD), expressed in
187
terms of number of individuals per square meter (ind/m2), was estimated for each sampling site.
AC C
EP
183
188
Furthermore, since the aggregation of adult sea urchins, intended as individuals with TD > 3
189
cm, provides a protected environment to recruits from both predation (Bonaviri et al., 2012;
190
Oliva et al., 2016) and mechanical removal due to water currents and waves (e.g. Nishizaki and
ACCEPTED MANUSCRIPT
Ackerman, 2007), their densities, expressed as ind/m2, were also calculated for each sampling
192
site. Adult density (AD), along with the oceanographic variables, was considered as one of the
193
predictor factors that can potentially affect recruit distribution density.
RI PT
191
3.2 Oceanographic variables
195
A three-dimensional, finite element hydrodynamic and wind wave model, SHYFEM-WWM
196
(Umgiesser et al., 2004), was adopted to reproduce the hydrodynamic features of the investigated
197
area. The model had already been used successfully for reproducing the wind-wave and the 3D
198
water circulation along the Sinis Peninsula and in the adjacent Oristano Gulf (Cucco et al., 2006;
199
De Falco et al., 2008; Cucco et al., 2016). The model uses an unstructured finite element grid for
200
representing the computational domain. In Fig. 2, the grid geometry and the adopted bathymetric
201
details are shown for the area of interest.
M AN U
SC
194
A simulation run was performed accounting for both oceanic and meteorological seasonal
203
variability. Large-scale atmospheric and oceanographic data provided by already operational
204
ocean and atmospheric models for the biennium 2009 and 2010 were adopted as forcing for the
205
coastal application of SHYFEM-WWM. The 2009-2010 biennium was selected for its
206
meteorological conditions which are highly representative of the local climate (see Appendix
207
A1).
AC C
EP
TE D
202
208
The ocean data used as model boundary conditions, including water levels, current speeds,
209
salinity and temperature fields, were provided by the WMED ocean forecasting system (Sorgente
210
et al., 2016). The atmospheric data, needed for model upper boundaries and including wind
211
speeds, atmospheric pressure and thermal fluxes were provided by SKIRON meteorological
ACCEPTED MANUSCRIPT
forecasting system (Papadopoulos et al., 2001). We refer to Cucco et al. (2012a, 2012b) for a
213
detailed description of the model parameters and the treatment of boundary conditions
M AN U
SC
RI PT
212
218
Fig. 2. Finite element grid and bathymetric data adopted for the numerical simulations: a) the external domain and b) the close-up of the Sinis coastal area with details of the c) bathymetry and sampling site arrangement
EP
215 216 217
TE D
214
The hydrodynamic variables describing the main physical processes influencing the local
220
recruitment of the sea urchin juveniles were defined a priori. The Current Speed (CS, expressed
221
in cm/s), the Significant Wave Height (SWH, expressed in m), the Sea Surface Temperature
222
(SST, expressed in °C) and the Eulerian Ritention Index (ERI, an adimensional variable
223
describing the spatial distribution of the average concentration of a passive tracer exposed to the
224
water circulation patterns), were selected as predictor variables as they are generally considered
225
the main oceanographic factors affecting the larvae fluxes, settlement and the recruitment
226
processes of benthic organisms (e.g. Tracey et al., 2012).
AC C
219
ACCEPTED MANUSCRIPT
From simulation results, the CS, SST and SWH were obtained once an hour for all the
228
elements of the model domain and for the whole simulation period. CS and SST were computed
229
for the surface waters represented by the first 10 m of the water column. This was necessary to
230
include the depths of the sampling sites and to avoid any aliasing errors generated by the model
231
spatial discretization.
RI PT
227
For each element of the domain, the SST and SWH were averaged over the whole simulated
233
period whereas the CS values were averaged only considering the results obtained between
234
January and June, when spawning occurs (see Introduction), for each simulated year.
M AN U
SC
232
A further variable, the ERI, characterized by a higher order of complexity with respect to the
236
previous ones, was selected as predictor factor. We simplified and approximated the sea urchin
237
spawning process by modelling the current-induced transport of a dissolved passive tracer
238
released along the whole coastal area. The retention of this tracer, quantified by the time
239
evolution of its local average concentration, was selected as a proxy for planktonic organism-
240
aggregation capacity, a feature favouring larvae settlement and the recruitment.
TE D
235
Numerically, the ERI was estimated by adding, daily, a unitary concentration of a dissolved
242
passive tracer to all the elements of the model domain characterized by water depths ranging
243
between 0 and 20 m. This corresponds to the habitats and the bathymetric range where sea
244
urchins are most abundant (Boudouresque and Verlaque, 2001). The transport of the tracer
245
induced by the water circulation was simulated for the whole spawning period, from January to
246
June. Its concentration was computed hourly for all the elements of the model domain. The
247
concentration values were then averaged in time and normalized between their minimum and
248
maximum values, obtaining the spatial distribution of an adimensional variable defined as the
AC C
EP
241
ACCEPTED MANUSCRIPT
Eulerian Retention Index. Previous studies on hydrodynamic connectivity between
250
Mediterranean sub-regions (Rossi et al., 2014; Dubois et al., 2016) indicated that the inner shelf
251
of the Sardinian island is relatively isolated from the outer-ocean circulation. This justified the
252
choice of seeding the tracer concentration along the coastal waters only. From a methodological
253
standpoint, ERI is similar to the concept of water residence time (Cucco and Umgiesser; 2015).
254
However, unlike this concept, ERI cannot provide any quantitative information about the time
255
scales of the transport processes. In fact, ERI supplies only relative indications about the
256
different capacities of the local current patterns to promote the trapping of the water masses
257
along the investigated stretch of coast. It does not give information about the lifetimes of the
258
urchins or duration of the simulated larvae or eggs.
M AN U
SC
RI PT
249
The different timing of the two averaging procedures reflects the impacts that the selected
260
physical factors were supposed to generate on the sea urchin recruitment. SST and SWH were
261
expected to mostly influence the annual reproductive cycle, the growth and the fixing ability of
262
recruits (Bulleri et al., 2015; Ouréns et al., 2011). Therefore, while these variables influence the
263
sea urchin population during the entire year, CS and ERI were more related to the transport and
264
settlement of the sea urchin larvae, which generally occur between late January and early June.
EP
TE D
259
3.3 Data analysis
266
A Shapiro-Wilk and Pearson test was applied to the recruit and adult density, ascertaining
267
that the two datasets were assessed as being non-normally distributed. Therefore, the variance of
268
the sea urchin density of both recruits and adults, were analysed considering the “Protection”
269
effect as a fixed factor and the “Site” effect as a random factor nested in the previous one. For
270
this analysis, a preliminary Generalized Linear Mixed Model (GLMM) was chosen as the best
AC C
265
ACCEPTED MANUSCRIPT
271
tool for analysing datasets characterized by non-normal distribution and which involve random
272
effects (Bolker et al., 2009). GLMM was run to model the relationship between density variability for the recruits and the
274
predictors consisting of the oceanographic variables (CS, SWH, SST and ERI) and the adult
275
density (AD). In order to exclude the statistical effects induced by different sampling depths (2
276
or 5 m), the “depth” was set as a further predictor characterized by a random distribution and
277
independent from the response variable RD. Moreover, the whole dataset was scaled to follow
278
Poisson distribution, which is a necessary condition before applying the GLMM based analysis
279
(Bolker et al., 2009). According to the protocol provided by Zuur et al. (2010), before testing the
280
model and providing the final solution, a data exploration of all the predictor variables was
281
carried out by analysing the covariates among all the different datasets to detect possible
282
collinearity. Finally, we used a Multi-model Averaging method for the Best Model (BM)
283
selection, which is a necessary method for providing a robust mean to infer the relative
284
importance of the different predictive variables (Grueber et al., 2011).
287
SC
M AN U
TE D
EP
286
All the described procedures and analyses were performed using the Nortest, lme4 and MultiModel Inference packages (Barton, 2012) in R ( R Development Core Team 2010).
AC C
285
RI PT
273
288
4. Results
289
4.1 Sea urchin density distribution
290
The preliminary GLMM analysis revealed that the “Protection” factor strongly influenced the
291
variability of distribution for recruits (P-value =7.58x10-13) but not for adults (see Table 1). The
ACCEPTED MANUSCRIPT
spatial distribution of the recruits (RD) was highly variable and ranged between 0 and 11 ind/m2
293
between sampling sites. RD values were significantly higher in sampling sites located outside the
294
MPA (sites from 16 to 25) than the ones inside (sites from 1 to 15), with 4±0.8 and 0.8±0.2
295
ind/m2 respectively (see Fig. 3). The highest values were observed at Su Tingiosu and Su Pallosu
296
(sampling sites number 18 and 22) with 6.3±0.9 and 7.2±1.8 ind/m2 respectively (Fig.3).
RI PT
292
SC
297
Fixed effect - Protection SE
Z-value
AD
RD
AD
RD
AD
RD
AD
18.066
-0.049
0.252
0.102
7.169
-0.480
7.58x10-13
0.630
σ2
304
AD
0.008
0.0041
SD
RD
AD
RD
AD
RD
AD
0.089
0.064
0.101
0.014
0.318
0.119
EP
RD
σ2
TE D
SD
Random effect - Site
Table 1. Generalized Linear Mixed Models (GLMM) for both recruit (RD) and adult densities (AD) in relation to the level of protection (fixed factor) and sites (random factor nested to protection). Coefficient estimates (Estimate), standard errors (SE), z-values, and significance level (P-value) are provided for fixed effects, while estimates of the variance (σ2) and standard deviations (SD) are reported for random effect.
AC C
299 300 301 302 303
P-value
RD
Random effect – Protection: Site
298
M AN U
Estimate
305
Conversely, inside the MPA where harvesting was more controlled, lower RD densities were
306
found at Cape Seu and Porto Suedda (sampling sites number 4 and 11) with values 0.2±0.1 and
307
0.2±0.2 ind/m2, and two sites (5 and 15) with an absence of recruits (Fig. 3).
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
308 309 310 311
Fig. 3. Spatial variability of Paracentrotus lividus recruits across the coast (density for m2). Sampling sites from 1 to 15 are placed inside the Marine Protected Area and are indicated in Fig.1. Details of significant differences are reported in Table 1.
312
Adult density (AD: >3cm TD) varied significantly between sites (1 and 27 ind/m2), but no
314
significant differences were found inside and outside the MPA (8.5 ±1.6 ind/m2 and 7.3 ±1.4
315
ind/m2 respectively, see Table 1, Fig 4). In fact, both the highest and the lowest density values
316
were found inside the MPA, near Cape Seu, at sampling sites 4 (20.9 ±2.7 ind/m2) and 8 (5.2
317
±1.3 ind/m2), respectively.
AC C
EP
TE D
313
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
318 319 320
Fig. 4. Density of adult individuals for m2 (TD>3 cm) along the coast inside and outside the Marine Protected Area (grey area). Significant differences between sites are reported in Table 1.
TE D
321 322
4.2 Oceanographic variables
323
In figure 5 the spatial distribution of the four diagnostics, CS, ERI, SST and WHS, used as
EP
predictors and obtained by the oceanographic model results, are reported for the area of interest.
AC C
324
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
325
ACCEPTED MANUSCRIPT
Fig. 5. Spatial patterns of the oceanographic variables used as predictors. Spatial distribution of the semi-annual mean of the surface a) Current Speed (CS) expressed in cm/s, b) normalized tracer concentration (ERI) and of the annual c) Sea Surface Temperature (SST) expressed in degree Celsius and d) Significant Wave Height (SWH) expressed in m.
330
The CS, (see Fig. 5a) varied from a few cm/s up to 28 cm/s with higher values found in the
331
offshore areas and lower values in proximity of the coastline. Exceptions were found near the
332
main capes, Mannu, Seu and Sa Sturaggia (see Fig. 1), where the compression of the flow field
333
promotes the intensification of the CS. Along the coasts, the CS pattern can be clearly divided
334
into three areas: the Su Pallosu and Putzu Idu Bay, located to the north and south of Cape
335
Mannu, characterized by lower CS values (less than 10 cm/s), and the remaining part of the
336
domain, from Sa Sturaggia to Seu Cape, characterized by higher CS values (less than 20 cm/s)
337
with peaks in proximity of the two headlands.
M AN U
SC
RI PT
326 327 328 329
As reported in appendix A2, the averaging procedure adopted to compute the CS, was proven
339
to be statistically confident. This allowed us to deepen the analysis of the surface hydrodynamics
340
throughout the computation of the Residual Circulation (RC). The RC was obtained by averaging
341
the hourly vector fields of the surface currents over time (see Cucco et al., 2006; 2016) obtained
342
by the model results. The RC differs from the average current speed (CS) by providing essential
343
information about the direction of the average circulation.
EP
TE D
338
In Figure 6, the residual surface current field, computed for the first six months of both
345
simulated years, is reported for the interested area. As can be noted, the residual flow is not
346
uniform. Several cyclonic and anticyclonic circulation cells can be detected in the coastal areas
347
of the northern part of the domain. In contrast, a constant north-south residual flow is found in
348
the southern part in both the offshore and coastal areas. The described flow pattern represents the
349
average driver of the passive tracer used to compute the ERI.
350
AC C
344
351 352 353 354
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Fig. 6. Residual sea-surface circulation computed from simulation results obtained for the first half of the year between January and June. Dots refer to the sea urchin sampling sites
355
The spatial distribution of ERI (Fig. 5b) is generally characterized by a cross-shore gradient
356
with lower values found in the offshore areas and higher values found in proximity of the coast.
ACCEPTED MANUSCRIPT
As for the CS, the ERI pattern can be summarily divided into three areas: the two small bays
358
(Putzu Idu and Su Pallosu bays) located at the north and south of Cape Mannu, characterized by
359
higher ERI values (around 0.9), and the remnant part of the coast, south of Sa Sturaggia Cape,
360
where the values of ERI were generally lower (around 0.4).
RI PT
357
The SST spatial distribution (see Fig. 5c) is quite homogeneous, as we expected, due to the
362
small extent of the area of interest. The annual average values ranged between 17.5° C and to 19°
363
C with colder open waters and warmer coastal waters. Similarly, to previous cases, the spatial
364
pattern of the SST in the coastal waters (including the two bays in proximity to Cape Mannu),
365
was characterized by a northern part, which had a higher SST average, and by a southern part
366
with colder waters.
M AN U
SC
361
Finally, the spatial distribution of SWH (see Fig. 5d) ranged between 0.3 and 2 m with the
368
peak values found offshore, in correspondence to Cape Mannu. The lowest values were found in
369
the inner part of the Putzu Idu and Su Pallosu bays, in correspondence with the shallow areas
370
which are sheltered from Mistral and Libeccio main wave regimes, respectively. Along the
371
remaining part of the coast, the SWH is quite homogeneous with values close to 1.5 m as this
372
area is exposed to both the two main wave regimes.
374
EP
AC C
373
TE D
367
4.3 Recruitment model results
375
A GLMM was carried out to relate the sea urchin recruit density to the local hydrodynamic
376
features and to the local abundance of adult sea urchins. A preliminary covariate analysis
377
revealed that the SST was characterized by a high value of the Variance Inflation Factor
ACCEPTED MANUSCRIPT
378
(VIF>3), indicating a strong collinearity with the CS and with the ERI and, consequently, it was
379
excluded from the set of model predictors. Accordingly, the GLMM was conducted using CS, ERI, SWH and AD as predictor variables
381
and RD as the response variable. For the whole dataset, the GLMM results, based on the
382
Akaike's Information Criterion (AIC) and the likelihood ratio tests (Zuur et al., 2009), revealed
383
that only few covariates influenced the RD variability along the selected coastal area. In fact,
384
from the collection of covariates, the CS and the ERI were the only predictors to significantly
385
influence the abundance of recruits (AIC 821, see Appendix A.3).
M AN U
SC
RI PT
380
Finally, a multi-model media selection, taking into account the uncertainty of the model, was
387
developed in order to guarantee the robustness of the GLMM results (Grueber et al., 2011). This
388
technique resulted in a set of the five best models with AIC ranging from 821 for the first model
389
(weight 0.25), up to 823 for the fifth model (weight 0.07) and where, the Current Speed (CS) and
390
the Eulerian Retention Index (ERI) were the most significant explanatory variables with a
391
relative importance of 1 and 0.93, respectively (Table 2).
TE D
386
EP
392
Estimate coefficient
Unconditional SE
P-value
Relative importance
Current Speed
-1.22
0.12337
< 2e-16
1
Eulerian Retention Index
0.13452
0.05142
0.00564
0.93
Significant Wave Height
0.09853
0.05772
0.08781
0.56
Adult density
0.06531
0.10433
0.20265
0.42
AC C
Covariate
393
ACCEPTED MANUSCRIPT
394 395 396 397
Table 2. Relative importance of local hydrodynamic factors in determining sea urchin recruitment along the Sinis coastline, using a multi-model averaging technique. The estimates reflect the weighted average of covariates derived from the six best models identified with a setting of Delta AIC<4; Significant P-values are given in bold (see also Table A3b).
RI PT
398 CS intensity negatively correlated with recruit densities (negative estimate coefficient), ERI
400
is positively associated with them (positive estimate coefficient). Fig. 7 shows the clear opposite
401
trends of recruit density with CS and ERI along the sampling sites.
SC
399
M AN U
402
403 404
AC C
EP
TE D
405 406 407 408 409 410 411 412 413
ACCEPTED MANUSCRIPT
414 415
Fig. 7. Significant correlations between response variable of recruit density (ind/m2) and a) CS (P-value=6.58e-10) and b) ERI (P-value=<2e-16).
416 5. Discussion
418
The recruit density displays a marked spatial heterogeneity in the shallow-sea rocky habitat
419
of the Peninsula of Sinis. In particular, two areas with significant differences can be identified: a
420
low-density area located inside the MPA and a high-density area outside the MPA which
421
corresponds to the non-protected, highly-fished areas in the North of Sinis Peninsula.
422
Conversely, the adult density distribution was found to be similar inside and outside the
423
protected area.
M AN U
SC
RI PT
417
The GLMM analysis highlighted that the hydrodynamic variables Current Speed (CS) and
425
Eulerian Retention Index (ERI), which represent different aspects of the sea surface transport
426
processes, were the most important predictors of the recruit density distribution in the
427
investigated area. In particular, the results revealed that in areas where the average current speeds
428
were low and the waters were more confined or trapped, corresponding to high ERI values,
429
recruitment (of sea urchins <1 cm) could be favoured, since a significant correlation with recruit
430
density was found. These features are typically associated with the presence of eddies or
431
standing circulation structures that force the water masses to recirculate locally. This increases
432
their residence times and favours the sinking of suspended elements, in the case of sediments or
433
biotic aggregates, or settling, in the case of larvae (Monsen et al., 2002; Canu et al., 2012; Cucco
434
et al., 2015).
AC C
EP
TE D
424
435
In our study case, the presence of such dynamic features was confirmed by the analysis of the
436
residual circulation (see figure 6). In particular, residual circulation cells were found in both Su
ACCEPTED MANUSCRIPT
Pallosu and Putzu Idu bays, outside the MPA, where CS values are lower and ERI values higher,
438
respectively. On the contrary, along the rest of the coastal area, where CS values are higher and
439
ERI values lower, the residual flow is uniform, reducing the probability of aggregation or
440
trapping of the suspended particles, including planktonic larvae.
RI PT
437
The spatial distributions of ERI and CS are in line with the results obtained from studies that
442
adopt particle-tracking models to investigate the retention features of the Mediterranean waters.
443
Dubois et al., 2016 found that the retention of numerical particles used to simulate planktonic
444
larvae and eggs was generally favoured by weak current conditions and complex bathymetry. In
445
coastal areas, these give rise to convoluted patterns of circulation which promote the trapping of
446
water masses (Paris and Cowen, 2004; Treml et al., 2012). In particular, along the Western
447
Sardinian shelf, the absence of both energetic large-scale coastal currents and mesoscale
448
structures travelling offshore, reduces the connectivity between the shelf and the open waters
449
(Dubois et al., 2016). This increases the retention of planktonic aggregates in the coastal waters.
450
Within this context, in those stretches of coast, where ERI and CS were high and low
451
respectively, the local circulation further promotes the retention of planktonic organisms such as
452
sea urchin eggs and larvae.
EP
TE D
M AN U
SC
441
Although, the SST was excluded from the GLMM analysis due to collinearity with CS and
454
ERI, its values were generally higher (up to 1 °C more) in those zones characterized by the
455
presence of recirculation cells than in the rest of the coast (see Fig. 3c). This supports the
456
hypothesis that the trapping processes occur in correspondence to the recirculation cells, and
457
indicates that these specific areas can also be more suitable for the reproduction success of
458
populations than the rest of the coast, due to the higher local SST (Ouréns et al., 2013).
AC C
453
ACCEPTED MANUSCRIPT
The results of the GLMM also exclude wind waves as a potential factor influencing the
460
variability in recruit densities, suggesting that the local sea urchin population is capable of
461
surviving in energetic environments. In fact, the distribution of the average SWH, that define
462
areas with a high or low average energy, in terms of wave impact, highlights that the whole
463
stretch of coast is affected by intense storms throughout the year. Implicitly, this confirms that
464
almost everywhere along the western Sardinian coast, the local sea urchin population is subjected
465
to intense wind wave events shaping the population dynamics (see Fig. 3d) thanks to the
466
presence of the Mistral and Libeccio wind regimes.
SC
RI PT
459
Regarding the importance of the biotic variables, the adult sea urchins are known to protect
468
newly settled individuals from predators under the spine canopy, thus favouring the aggregation
469
of juveniles (Oliva et al., 2016; Ourèns et al., 2014). However, according to the similar densities
470
of adults observed inside and outside the MPA (Fig. 2), results of the GLMM exclude AD as an
471
explanatory variable of difference in recruit density variation.
TE D
M AN U
467
Generally, inside MPAs, sea urchin populations suffer a high natural mortality of recruits and
473
middle-sized classes as a consequence of the local fish abundance and sizes of fish present which
474
are generally larger as a direct effect of the protection rules (Guidetti, 2006; Hereu et al., 2012,
475
2005; Sala et al., 1997). However, in the investigated MPA Peninsula of Sinis - Mal di Ventre
476
Island, the effects of protection are not appreciable due to the high-fishing pressure (Casola et al.,
477
2014; Guala et al., 2008b) that compromises the effectiveness of the reserve (Marra et al., 2016;
478
Pieraccini et al., 2016). The abundance of the main specialist predators of small sea urchins (0-
479
1cm TD), the wrasses Coris julis and Thalassoma pavo, was observed to be similar in the rocky
480
habitat substrate inside and outside the MPA (Marra et al., 2016). Accordingly, differences in
481
recruit densities (0-1 cm TD) inside and outside the MPA cannot depend on predation. The
AC C
EP
472
ACCEPTED MANUSCRIPT
482
higher density outside the MPA seems to be primarily caused by the peculiar circulation pattern
483
detected in Su Pallosu and Putzu Idu bays. In general, populations are inclined to a high reproductive capacity in zones where
485
oceanographic forcing is weak, like for example low wave exposure that positively influences
486
the number of gametes released in the environment (Bulleri et al., 2015; Gianguzza et al., 2013;
487
Guettaf et al., 2000; Lozano et al., 1995; Meidel and Scheibling, 1998; Sellem and Guillou,
488
2007). However, local water circulation also makes larval supply available to a large amount of
489
coastal populations (Fenaux et al., 1988; López et al., 1998; Prado et al., 2012). In this sense,
490
larval dispersion in the offshore and coastal areas and their transport induced by the local current
491
field assume a crucial role for population survival (Largier, 2003).
M AN U
SC
RI PT
484
As concerns the coastlines of the Peninsula of Sinis, it is feasible that both larvae produced
493
locally and coming from the nearby areas are mainly trapped in the Su Pallosu and Putzu Idu
494
bays, while they only partially travel down the coast. Then, the higher SST in the recirculation
495
cells and the favourable environmental constraints of the area (e.g. substrate rugosity), could
496
increase the chances of larvae survival and their in situ settling (Boudouresque and Verlaque,
497
2001; Oliva et al., 2016).
EP
TE D
492
It is worth noting that, while larval availability is influenced by large or medium scale factors
499
(e.g. currents, temperature), settlement and early post-settlement events are affected by local
500
scale features such as substrate rugosity and micropredation by small invertebrates (Bonaviri et
501
al., 2012; Jennings and Hunt, 2011; Oliva et al., 2016). This work did not explore either the
502
abundance of settlers, nor the micropredation which entails post-settlement mortality and later
503
demographic stages. However, considering predation and habitat features as constant and similar
504
inside and outside MPA, it is reasonable to suppose that the correlation between local
AC C
498
ACCEPTED MANUSCRIPT
hydrodynamic patterns and post-settlement recruit density could be extended to larval supply and
506
settlers as intermediate phases (Prado et al., 2012). Moreover, our results highlight the possible
507
existence of hot-spots of P. lividus recruitment and their placement outside the local MPA
508
borders, in correspondence to the unprotected northernmost part of the domain.
RI PT
505
According to this hypothesis, this study could explain how the sea urchin populations in this
510
area can sustain the strong human pressure which is present. However, the persistence of the
511
harvesting has led to a deficit of the large adults and the potential reproductive contribution of
512
the population is dependent on the youngest of the species (Loi et al., 2017).
SC
509
Our findings allow us to infer that the success of the recruitment favoured by the water
514
circulation pattern would lead to a faster recovery of the local population in terms of total
515
abundance, thus ensuring the availability of new juveniles for the future stock of local fishing.
M AN U
513
Understanding the effects of demographic drivers of population processes and their spatial
517
and temporal variability is necessary in order to accurately evaluate the long-term population
518
viability and to devise effective management practices. In this sense, an interesting perspective
519
for realizing a systematic conservational plan of sustainable sea urchin harvesting could be based
520
on metapopulation dynamics and their connection through a source-sink model (Hanski and
521
Simberloff, 1997; Hastings and Harrison, 1994; Ilkka, 1999). These concepts emphasize the
522
importance of connectivity in generating a system of discrete local metapopulations: each one
523
determines its own internal dynamics to a large extent, but has a degree of identifiable
524
demographic influence from other populations which are connected via dispersion processes
525
(Knight and Landres, 2002). For instance, the populations of Su Pallosu and Putzu Idu bays are
526
characterized by high densities of recruits and middle-sized classes of sea urchins (Loi et al.,
527
2017). Due to the high harvesting pressure, death rates should exceed birth rates and, as a
AC C
EP
TE D
516
ACCEPTED MANUSCRIPT
consequence, these metapopulations should be classified as “sink”. Conversely, metapopulations
529
whose birth rates will exceed death rates should be referred to as “source”. The phenomenon
530
where “source” populations supplement “sink” populations via dispersing individuals is a rescue
531
effect that should be considered crucial for planning protection plans and sustainable fisheries
532
(Paterno et al. 2017).
RI PT
528
Paradoxically sea urchin harvesting inside MPAs, as an extra-predation pressure, can
534
dramatically compromise the persistence of metapopulations more than in overfished areas.
535
However, in the case of the investigated MPA where there is no significant reserve effect (Marra
536
et al., 2016), the balance between death rate and birth rate along the coast is almost exclusively
537
dependent on the larval supply. Thus, the knowledge of the main routes of the seasonal dispersal
538
of sea urchin larvae through the MPA Peninsula of Sinis - Mal di Ventre Island and the
539
surrounding areas becomes a key factor for the successful management of the resource.
TE D
M AN U
SC
533
Moreover, the presence of residual circulation favouring the trapping of planktonic organisms
541
has a major impact on ecosystem conservation issues. In fact, when the circulation pattern tends
542
to reduce the dispersion of marine organisms, a local increase of the biomasses is found (e.g.
543
Melià et al., 2016). This supports trophic interactions among predators, preys, competitors and
544
macrophytes.
AC C
EP
540
545
Our findings point out the necessity of evaluating ecological processes which drive sea urchin
546
population dynamics, such as recruitment, to differentiate the harvesting in different sectors.
547
However, we also believe that our study could contribute positively to an efficient management
548
plan for biodiversity conservation. In particular, further investigation which extended the study
549
site to the whole western Mediterranean region and, in this case, used an individually based
ACCEPTED MANUSCRIPT
550
modelling approach would improve the understanding of the potential role of larvae dispersal
551
processes in the management of this important resource.
RI PT
552 Acknowledgements
554
This research was carried out thanks to the cooperation between IMC and IAMC-CNR (Oristano,
555
Sardinia). We thank Maura Baroli, Fabio Ledda, Clara Diago, Stefania Coppa, Alessia Iannuzzi,
556
Serena Como, Serena Donadi, Francesco Wrachien, Giorgio Massaro who have collaborated in
557
the field data collection. Finally, we would like to thank Katie Duff for her patient revision of the
558
manuscript.
M AN U
SC
553
559 Supplementary data
561
Supplementary data related to this article can be found at
562
EP
TE D
560
References
564 565 566 567 568 569 570 571 572 573 574 575
Addis, P., Moccia, D., Secci, M., 2014. Effect of two different habitats on spine and gonad colour in the purple sea urchin Paracentrotus lividus. Mar Ecol 1–7. doi:10.1111/maec.12133 Andrello, M., Mouillot, D., Beuvier, J., Albouy, C., Thuiller, W., Manel, S., 2013. Low Connectivity between Mediterranean Marine Protected Areas: A Biophysical Modeling Approach for the Dusky Grouper Epinephelus marginatus. PLoS ONE 8(7): e68564. Andrew, N.L., Agatsuma, Y., Ballesteros, E., Bazhin, A.G., Creaser, E.P., Barnes, D.K.A., Botsford, L.W., Bradbury, A., Campbell, A., Dixon, J.D., Einarsson, S., Gerring, P.K., Hebert, K., Hunter, M., Hur, S.B., Johnson, C.R., Juinio-Menez, M.A., Kalvass, P., Miller, R.J., Moreno, C.A., Palleiro, J.S., Rivas, D., Robinson, S.M.L., Schroeter, S.C., Steneck, R.S., Vadas, R.L., Woodby, D.A., Xiaoqi, Z., 2002. Status and management of world sea urchin fisheries, in: Oceanography and Marine Biology an Annual Review. Taylor &
AC C
563
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
Francis LTD, 11 New Fetter Lane, London ec4p 4ee, England, pp. 343–425. Barton, K., 2012. Package “MuMIn: Multi-model inference” for R, R Package Version 1.6.6 (http://CRAN.R-project.org/package=MuMIn). Boada, J., Arthur, R., Alonso, D., Pagès, J.F., Pessarrodona, A., Oliva, S., Ceccherelli, G., Piazzi, L., Romero, J., Alcoverro, T., 2017. Immanent conditions determine imminent collapses: nutrient regimes define the resilience of macroalgal communities. Proc R Soc B Biol Sci 284, 20162814. doi:10.1098/rspb.2016.2814 Boada, J., Farina, S., Arthur, R., Romero, J., Prado, P., Alcoverro, T., 2018. Herbivore control in connected seascapes: habitat determines when population regulation occurs in the life history of a key herbivore. Oikos. doi:10.1111/oik.02629 Bolker, B.M., Brooks, M.E., Clark, C.J., Geange, S.W., Poulsen, J.R., Stevens, M.H.H., White, J.S.S., 2009. Generalized linear mixed models: a practical guide for ecology and evolution. Trends Ecol Evol 24, 127–135. doi:10.1016/j.tree.2008.10.008 Bonaviri, C., Gianguzza, P., Pipitone, C., Hereu, B., 2012. Micropredation on sea urchins as a potential stabilizing process for rocky reefs. J Sea Res 73, 18–23. doi:10.1016/j.seares.2012.06.003 Boudouresque, C.F., Verlaque, M., 2001. Ecology of Paracentrotus lividus, in: Miller, J. (Ed.), Edible Sea Urchins: Biology and Ecology. Elsevier Science, pp. 177–216. Bray. L., Kassis, D., Hall-Spencer., J.M., 2017. Assessing larval connectivity for marine spatial planning in the Adriatic. Marine Environmental Research, 125, 73–81. http://doi.org/10.1016/J.MARENVRES.2017.01.006 Bulleri, F., Cucco, A., Dal Bello, M., Maggi, E., Ravaglioli, C., Benedetti-Cecchi, L., 2015. The role of wave-exposure and human impacts in regulating the distribution of alternative habitats on NW Mediterranean rocky reefs. Estuar Coast Shelf Sci 1–9. doi:10.1016/j.ecss.2016.02.013 Calò, A., Muñoz, I., Pérez-Ruzafa, Á., Vergara-Chen, C., & García-Charton, J. A., 2016. Spatial genetic structure in the saddled sea bream (Oblada melanura [Linnaeus, 1758]) suggests multi-scaled patterns of connectivity between protected and unprotected areas in the Western Mediterranean Sea. Fisheries Research, 176, 30–38. http://doi.org/https://doi.org/10.1016/j.fishres.2015.12.001 Calò, A., Lett, C., Mourre, B., Pérez-Ruzafa, Á., & García-Charton, J. A. 2018. Use of Lagrangian simulations to hindcast the geographical position of propagule release zones in a Mediterranean coastal fish. Marine Environ Res, 134, 16–27. http://doi.org/https://doi.org/10.1016/j.marenvres.2017.12.011 Casola, E., Lariccia, M., Scardi, M., 2014. Areee Marine Protette e pesca professionale, Unimar: Roma. Canu, D. M., Solidoro, C., Umgiesser, G., Cucco, A., & Ferrarin, C., 2012. Assessing confinement in coastal lagoons. Marine Poll Bull, 64(11), 2391–8. http://doi.org/10.1016/j.marpolbul.2012.08.007 Ceccherelli, G., Pais, A., Pinna, S., Serra, S., Sechi, N., 2009. On the movement of the sea urchin Paracentrotus lividus towards Posidonia oceanica seagrass patches. J Shellfish Res 28, 397–403. doi:http://dx.doi.org/10.2983/035.028.0224 Cucco, A., Perilli, A., De Falco, G., Ghezzo, M., Umgiesser, G., 2006. Water circulation and transport timescales in the Gulf of Oristano. Chem Ecol 22, 307–331. doi:http://doi.org/10.1080/02757540600670364 Cucco, A., Sinerchia, M., Lefrancois, C., Magni, P., Ghezzo, M., Umgiesser, G., Perilli, A.,
AC C
576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
Domenici, P., 2012a. A metabolic scope based model of fish response to environmental changes. Ecol Modell 237–238, 132–141. doi:http://doi.org/10.1016/j.ecolmodel.2012.04.019 Cucco, A., Sinerchia, M., Ribotti, A., Olita, A., Fazioli, L., Perilli, A., Sorgente, B., Borghini, M., Schroeder, K., Sorgente, R., 2012b. A high-resolution real-time forecasting system for predicting the fate of oil spills in the Strait of Bonifacio (western Mediterranean Sea). Mar Pollut Bull 64, 1186–1200. doi:https://doi.org/10.1016/j.marpolbul.2012.03.019 Cucco, A., & Umgiesser, G., 2015. The Trapping Index: How to integrate the Eulerian and the Lagrangian approach for the computation of the transport time scales of semi-enclosed basins. Marine Poll Bull, 98(1), 210–220. Cucco, A., Quattrocchi, G., Satta, A., Antognarelli, F., De Biasio, F., Cadau, E., Umgiesser, G., Zecchetto, S., 2016. Predictability of wind-induced sea surface transport in coastal areas. J Geophys Res Ocean 121, 5847–5871. doi:http://doi.org/10.1002/2016JC01164 De Falco, G., Baroli, M., Cucco, A., Simeone, S., 2008. Intrabasinal conditions promoting the development of a biogenic carbonate sedimentary facies associated with the seagrass Posidonia oceanica. Cont Shelf Res 28, 797–812. De Falco, G., Antonioli, F., Fontolan, G., Lo Presti, V., Simeone, S., Tonielli, R., 2015. Early cementation space dictate the evolution of an overstepping barrier system during the Holocene. Marine Geology, 369, 52-66. Dubois, M., Rossi, V., Ser‐Giacomi, E., Arnaud‐Haond, S., López, C., Hernández‐García, E., & Tittensor, D., 2016. Linking basin-scale connectivity, oceanography and population dynamics for the conservation and management of marine ecosystems. Global Ecology and Biogeography, 25(5), 503–515. http://doi.org/10.1111/geb.1243 Farina, S., Tomas, F., Prado, P., Romero, J., Alcoverro, T., 2009. Seagrass meadow structure alters interactions between the sea urchin Paracentrotus lividus and its predators. Mar Ecol Prog Ser 377, 131–137. doi:10.3354/meps07692 Farina, S., Arthur, R., Pagès, J.F., Prado, P., Romero, J., Vergés, A., Hyndes, G., Heck, K.L., Glenos, S., Alcoverro, T., 2014. Differences in predator composition alter the direction of structure-mediated predation risk in macrophyte communities. Oikos 123, 1311–1322. doi:10.1111/oik.01382 Farina, S., Guala, I., Oliva, S., Piazzi, L., Pires da Silva, R., Ceccherelli, G., 2016. The Seagrass Effect Turned Upside Down Changes the Prospective of Sea Urchin Survival and Landscape Implications. PLoS One 11, e0164294. doi:10.1371/journal.pone.0164294 Farina, S., Oltra, A., Boada, J., Bartumeus, F., Romero, J., Alcoverro, T., 2017. Generation and maintenance of predation hotspots of a functionally important herbivore in a patchy habitat mosaic. Funct Ecol 0–3. doi:0000-0003-0169-8044 Fenaux, L., Cellario, C., Rassoulzadegan, F., 1988. Sensitivity of different morphological stages of the larva of Paracentrotus lividus (Lamarck) to quantity and quality of food., in: Burke, R.D., Mladenov, P.V., Lambert, P., Parsley, R.L. (Eds.), Echinoderm Biology. AA Balkema, Rotterdam, pp. 259–266. Fenaux, L., Pedrotti, M.L., 1988. Metamorphose des Larves d’Echinides SNI en Pleine Eau Metamorphosis of Echinoid Larvae in Midwater. Mar Ecol 9, 93–107. Gianguzza, P., Bonaviri, C., Prato, E., Fanelli, G., Chiantore, M., Privitera, D., Luzzu, F., Agnetta, D., 2013. Hydrodynamism and its in fl uence on the reproductive condition of the edible sea urchin Paracentrotus lividus. Mar Environ Res 85, 29–33. doi:10.1016/j.marenvres.2012.12.007.
AC C
622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
Grueber, C.E., Nakagawa, S., Laws, R.J., Jamieson, I.G., 2011. Multimodel inference in ecology and evolution: Challenges and solutions. J Evol Biol 24, 699–711. doi:10.1111/j.14209101.2010.02210.x Guala I., Massaro G., de Lucia G.A., De Falco G., Domenici P., 2008a. Sinis-Mal di Ventre. In: Vandeperre F., Higgins R., Santos R.S., Marcos C., Pérez-Ruzafa A. (Coord.), 2008. Fishery Regimes in Atlanto-Mediterranean European Marine Protected Areas. EMPAFISH Project, Booklet nº 2. Editum. 47-49. ISBN: 978-84-8371-724-0. Guala I., Massaro G., de Lucia G.A., De Falco G., Domenici P., 2008b. Sinis-Mal di Ventre. In: Planes S., García-Charton J.A., Marcos C., Pérez-Ruzafa A. (Coord.), 2008. Ecological effects of Atlanto-Mediterranean Marine Protected Areas in the European Union. EMPAFISH Project, Booklet nº 1. Editum. 57-60. ISBN: 978-84-8371-723-3. Guettaf, M., San Martin, G. a., Francour, P., 2000. Interpopulation variability of the reproductive cycle of Paracentrotus lividus (Echinodermata: Echinoidea) in the south-western Mediterranean. J Mar Biol Assoc UK 80, 899–907. Guidetti, P., Bianchi, C.N., Chiantore, M., Schiaparelli, S., Morri, C., Cattaneo-Vietti, R., 2004. Living on the rocks: Substrate mineralogy and the structure of subtidal rocky substrate communities in the Mediterranean Sea. Mar Ecol Prog Ser 274, 57–68. doi:10.3354/meps274057 Guidetti, P., 2006. Marine reserves reestablish lost predatory interactions and cause community changes in rocky reefs. Ecol Appl 16, 963–976. doi:10.1890/10510761(2006)016[0963:MRRLPI]2.0.CO;2 Hanski, I., Simberloff, D., 1997. The metapopulation approach, its history, conceptual domain and application to conservation, in: Hanski, I., Gilpin, M.E. (Eds.), Metapopulation Biology. Academic Press, San Diego, pp. 5–26. Hastings, A., Harrison, S., 1994. Metapopulation Dynamics and Genetics. Annu Rev Ecol Syst 25, 167–188. doi:10.1146/annurev.es.25.110194.001123 Hereu, B., 2004. The role of trophic interactions between fishes, sea urchins and algae in the north-western Mediterranean rocky infralittoral. University of Barcelona. Hereu, B., Zabala, M., Linares, C., Sala, E., 2005. The effects of predator abundance and habitat structural complexity on survival of juvenile sea urchins. Mar Biol 146, 293–299. doi:10.1007/s00227-004-1439-y Hereu, B., Zabala, M., Sala, E., 2008. Multiple controls of community structure and dynamics in a sublittoral marine environment. Ecology 89, 3423–3435. Hereu, B., Linares, C., Sala, E., Garrabou, J., Garcia-Rubies, A., Diaz, D., Zabala, M., 2012. Multiple processes regulate long-term population dynamics of sea urchins on Mediterranean rocky reefs. PLoS One 7, 37–41. doi:10.1371/journal.pone.0036901 Ilkka, H., 1999. Metapopulation Ecology. Oxford University Press, Oxford. Jackson, J.B.C., Kirby, M.X., Berger, W.H., Bjorndal, K.A., Botsford, L.W., Bourque, B.J., Bradbury, R.H., Cooke, R., Erlandson, J., Estes, J.A., Hughes, T.P., Kidwell, S., Lange, C.B., Lenihan, H.S., Pandolfi, J.M., Peterson, C.H., Steneck, R.S., Tegner, M.J., Warner, R.R., 2001. Historical Overfishing and the Recent Collapse of Coastal Ecosystems. Science (80- ) 293, 629–637. doi:10.1126/science.1059199 Jennings, L.B., Hunt, H.L., 2011. Small macrobenthic invertebrates affect the mortality and growth of early post-settlement sea urchins and sea stars in subtidal cobble habitat. Mar Ecol Prog Ser 431, 173–182. doi:10.3354/meps09131 Knight, R.L., Landres, P.B., 2002. Central concepts and issues of biological conservation, in:
AC C
668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
Applying Landscape Ecology in Biological Conservation. pp. 22–33. Koeck, B., Gérigny, O., Durieux, E. D. H., Coudray, S., Garsi, L.-H., Bisgambiglia, P.-A. Galgani, F., Agostini, S. 2015. Connectivity patterns of coastal fishes following different dispersal scenarios across a transboundary marine protected area (Bonifacio strait, NW Mediterranean). Estuarine, Coastal and Shelf Science, 154, 234–247. http://doi.org/https://doi.org/10.1016/j.ecss.2015.01.010 Largier, J.L., 2003. Considerations in estimating larval dispersal distances from oceanographic data. Ecol Appl 13, 71–89. doi:10.1890/1051-0761(2003)013[0071:CIELDD]2.0.CO;2 Levitan, D.R., Sewell, M.A., Fu-Shiang Chia, 1992. How distribution and abundance influence fertilization success in the sea urchin Strongylocentrotus franciscanus. Ecology. Levitan, D.R., Sewell, M.A., 1998. Fertilization success in free-spawning marine invertebrates: review of the evidence and fisheries implications. Can Spec Publ Fish Aquat Sci 159–164. Loi, B., Farina, S., Brundu, G., Guala, I., Baroli, M., 2015. Reproductive cycle of Paracentrotus lividus at two sardinian coastal areas, in: 46° Congresso Della Società Italiana Di Biologia Marina. Loi, B., Guala, I., Pires, R., Brundu, G., Baroli, M., Farina, S., 2017. Hard time to be parents ? Sea urchin fishery shifts potential reproductive contribution of population onto the shoulders of the young adults 1–22. doi:10.7717/peerj.3067 López, S., Turon, X., Montero, E., Palacín, C., Duarte, C.M., Tarjuelo, I., 1998. Larval abundance, recruitment and early mortality in Paracentrotus lividus (Echinoidea). Interannual variability and plankton-benthos coupling. Mar Ecol Prog Ser 172, 239–251. doi:10.3354/meps172239 Lozano, J., Galera, J., Lopez, S., Turon, X., Palacin, C., Morera, G., 1995. Biological cycles and recruitment of Paracentrotus lividus (Echinodermata: Echinoidea) in two contrasting habitats. Mar Ecol Prog Ser 122, 179–192. Marra, S., Coppa, S., Camedda, A., Mazzoldi, C., Wrachien, F., Massaro, G., De Lucia, G.A., 2016. Recovery trends of commercial fish: The case of an underperforming mediterranean marine protected area. PLoS One 11, 1–22. doi:10.1371/journal.pone.0146391 McClanahan, T.R., Sala, E., 1997. A Mediterranean rocky-bottom ecosystem fisheries model. Ecol Modell 104, 145–164. Medel, C., Parada, C., Morales, C. E., Pizarro, O., Ernst, B., & Conejero, C. 2018. How biophysical interactions associated with sub- and mesoscale structures and migration behavior affect planktonic larvae of the spiny lobster in the Juan Fernández Ridge: A modeling approach. Progress in Oceanography, 162, 98–119. http://doi.org/https://doi.org/10.1016/j.pocean.2018.02.017 Meidel, S.K., Scheibling, R.E., 1998. Annual reproductive cycle of the green sea urchin, Strongylocentrotus droebachiensis, in differing habitats in Nova Scotia, Canada. Mar Biol 131, 461–478. doi:10.1007/s002270050338 Melià, P., Schiavina, M., Rossetto, M., Gatto, M., Fraschetti, S., Casagrandi, R., 2016. Looking for hotspots of marine metacommunity connectivity: a methodological framework. Sci Rep 6, 23705. doi:10.1038/srep23705 Millot, C., 1999. Circulation in the Western Mediterranean Sea. J Mar Syst 20, 423–442. doi:http://doi.org/10.1016/S0924-7963(98)00078-5 Mita, M., Sato, J., Hirosawa, Y., Nakamura, M., 2007. Gonadal maturation is dependent on body size in the sea urchin, Echinometra tsumajiroi 50, 187–190. doi:10.1080/07924259.2007.9652245
AC C
714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
Moffit, E.A., Bostford, L.W., Kaplan, D.M., M.R., O., 2009. Marine reserve networks for species that move within a home range. Ecol Appl 19, 1835–1847. Monsen, N.E., Cloern, J.E., Lucas, L.V., 2002. A comment on the use of flushing time, residence time, and age as transport time scales. Limnol. Oceanogr. 47 (5), 1545–1553 Morgan, L.E., Wing, S.R., Botsford, L.W., Lundquist, C.J., Diehl, J.M., 2000. Spatial variability in red sea urchin (Strongylocentrotus franciscanus) recruitment in northern California. Fish Oceanogr 9, 83–98. doi:10.1046/j.1365-2419.2000.00124.x Nishizaki, M.T., Ackerman, J.D., 2007. Juvenile-adult associations in sea urchins (Strongylocentrotus franciscanus and S. droebachiensis): Protection from predation and hydrodynamics in S. franciscanus. Mar Biol 151, 135–145. doi:10.1007/s00227-006-0462-6 Olita, A., A. Ribotti, L. Fazioli, A. Perilli, and R. Sorgente, 2013: Surface circulation and upwelling in the Sardinia Sea: A numerical study. Cont. Shelf Res., 71, 95–108. doi:10.1016/j.csr.2013.10.011 Oliva, S., Farina, S., Pinna, S., Guala, I., Agnetta, D., Ariotti, P., Mura, F., Checcherelli, G., 2016. Determinants of Paracentrotus lividus sea urchin recruitment under oligotrophic conditions: implications for conservation management. Mar Environ Res 117, 13–20. Ouréns, R., Fernández, L., Fernández-Boán, M., Naya, I., Freire, J., 2013. Reproductive dynamics of the sea urchin Paracentrotus lividus on the Galicia coast (NW Spain): Effects of habitat and population density. Mar Biol 160, 2413–2423. doi:10.1007/s00227-0132236-2 Ouréns, R., Fernández, L., Freire, J., 2011. Geographic, population, and seasonal patterns in the reproductive parameters of the sea urchin Paracentrotus lividus. Mar Biol 158, 793–804. doi:10.1007/s00227-010-1607-1 Ouréns, R., Freire, J., Vilar, J.A., Fernandez, L., 2014. Influence of habitat and population density on recruitment and spatial dynamics of the sea urchin Paracentrotus lividus: implications for harvest refugia. Mar Sci 71, 1064–1072. Ospina-Alvarez, A., Catalán, I. A., Bernal, M., Roos, D., & Palomera, I., 2015. From egg production to recruits: Connectivity and inter-annual variability in the recruitment patterns of European anchovy in the northwestern Mediterranean. Progress in Oceanography, 138, 431–447. http://doi.org/https://doi.org/10.1016/j.pocean.2015.01.011 Pais, A., Chessa, L. a., Serra, S., Ruiu, A., Meloni, G., Donno, Y., 2007. The impact of commercial and recreational harvesting for Paracentrotus lividus on shallow rocky reef sea urchin communities in North-western Sardinia, Italy. Estuar Coast Shelf Sci 73, 589–597. doi:10.1016/j.ecss.2007.02.011 Pais, A., Serra, S., Meloni, G., Saba, S., Ceccherelli, G., 2012. Harvesting effects on Paracentrotus lividus population structure: a case study from northwestern Sardinia, Italy, before and after the fishing season. J Coast Res 28, 570–575. Papadopoulos, A., Katsafados, P., Kallos, G., 2001. Regional weather forecasting for marine application. Glob Atmos Ocean Syst 8, 219–237. Paris, C. B., & Cowen, R. K., 2004. Direct evidence of a biophysical retention mechanism for coral reef fish larvae. Limnology and Oceanography, 49(6), 1964–1979. http://doi.org/10.4319/lo.2004.49.6.1964 Paterno, M., Schiavina, M., Aglieri, G., Ben Souissi, J., Boscari, E., Casagrandi, R., Chassanite, A., Chiantore, M., Congiu, L., Guarnieri, G., Krsuchel, C., Macic, V., Marino, I., Papetti, C., Patarnello, T., Zane, L., Melià, P., 2017. Population genomics meet Lagrangian simulations: Oceanographic patterns and long larval duration ensure connectivity among
AC C
760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805
ACCEPTED MANUSCRIPT
Paracentrotus lividus populations in the Adriatic and Ionian seas. Ecol Evol 7, 2463–2479. Pennington, J.T., 1985. The ecology of fertilization of echinoid eggs: the consequences of sperm dilution, adult aggregation, and synchronous spawning. Biol Bull 169, 417–430. doi:https://doi.org/10.2307/1541492 Pieraccini, M., Coppa, S., De Lucia, G.A., 2016. Beyond marine paper parks? Regulation theory to assess and address environmental non-compliance. Aquat Conserv Mar Freshw Ecosyst n/a-n/a. doi:10.1002/aqc.2632 Prado, P., Tomas, F., Alcoverro, T., Romero, J., 2007. Extensive direct measurements of Posidonia oceanica defoliation confirm the importance of herbivory in temperate seagrass meadows. Mar Ecol Ser 340, 63–71. Prado, P., Tomas, F., Pinna, S., Farina, S., Roca, G., Ceccherelli, G., Romero, J., Alcoverro, T., 2012. Habitat and scale shape the demographic fate of the keystone sea urchin Paracentrotus lividus in mediterranean macrophyte communities. PLoS One 7, e35170. doi:10.1371/journal.pone.0035170 Puckett, B.J., Eggleston, D.B., Kerr, P.C., Luettich, R.A., 2014. Larval dispersal and population connectivity among a network of marine reserves. Fish Oceanogr 23, 342–361. doi:10.1111/fog.12067
823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851
Rossi, V., Ser‐Giacomi, E., López, C., & Hernández‐García, E. 2014. Hydrodynamic provinces and oceanic connectivity from a transport network help designing marine reserves. Geophysical Research Letters, 41(8), 2883–2891. http://doi.org/10.1002/2014GL059540 Sala, E., 1997. Fish predators and scavengers of the sea urchin Paracentrotus lividus in protected areas of the north-west Mediterranean Sea. Mar Biol 129, 531–539. Sala, E., Boudouresque, C.-F., Harmelin-Vivien, M., 1998. Fishing, trophic cascades, and the structure of algal assemblages: evaluation of an old but untested paradigm. Oikos 82, 425– 439. Schroeder, K., Garcia-Lafuente, J., Josey, S.A., Artale, V., Nardelli, B.B., Carrillo, A., Gacic, M., Gasparini, G.P., Herrmann, M., Lionello, P., Ludwig, W., Millot, C., Ozsoy, E., Pisacane, G., Sanchez-Garrido, J.C., Sannino, G., Santoleri, R., Somot, S., Struglia, M., Stanev, E., Taupier-Letage, I., Tsimplis, M.N., Vargas-Yanes, M., Zervakis, V., Zodiatis, G., 2012. 3 – Circulation of the Mediterranean Sea and its Variability. Clim Mediterr Reg 187–256. doi:http://doi.org/10.1016/B978-0-12-416042-2.00003-3 Sellem, F., Guillou, M., 2007. Reproductive biology of Paracentrotus lividus (Echinodermata: Echinoidea) in two contrasting habitats of northern Tunisia (south-east Mediterranean). J Mar Biol Assoc UK. doi:10.1017/S002531540705521X Simeone S., Palombo L., De Falco G., 2012a. Morphodynamics of a nontidal embayed beach: the case study of Is Arutas (Western Mediterranean). Journal of Coastal Research, 29, 6, 6371 Simeone S., Palombo AGL, Guala I., 2012b. Impact of frequentation on a Mediterranean embayed beach: Implication on carrying capacity. Ocean and Coastal Management, 62, 914 Simeone, S., De Falco, G., Quattrocchi, G., Palombo, L., Cucco, A., 2016. Beaches Morphological Variability Along a Complex Coastline (Sinis Peninsula, western Mediterranean Sea). J Coast Res 1, 1302–1306. doi:http://doi.org/10.2112/SI75-261.1 Sorgente, R., Tedesco, C., Pessini, F., De Dominicis, M., Gerin, R., Olita, A., Fazioli, L., Di Maio, A., Ribotti, A., 2016. Forecast of drifter trajectories using a Rapid Environmental Assessment based on CTD observations. Deep Res Part II Top Stud Oceanogr 133, 39–53.
AC C
EP
TE D
M AN U
SC
RI PT
806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
doi:http://doi.org/10.1016/j.dsr2.2016.06.020 Steneck, R., Graham, M., Bourque, B., Corbett, D., Erlandson, J., Estes, J., Tegner, M., 2002. Kelp forest ecosystems: biodiversity, stability, resilience and future. Environ Conserv 29, 436–459. Steneck, R., Vavrinec, J., Leland, A., 2004. Accelerating Trophic-level Dysfunction in Kelp Forest Ecosystems of the Western North Atlantic. Ecosystems 7, 323–332. Stenevik, E.K., Nash, R.D.M., Vikebø, F., Fossum, P., Bakkeplass, K., 2012. The effects of survey design and circulation pattern on the perceived abundance of herring larvae: A case study for Norwegian spring spawning herring (Clupea harengus). Fish Oceanogr 21, 363– 373. doi:10.1111/j.1365-2419.2012.00631.x Tegner, M., Dayton, P., 1977. Sea urchin recruitment patterns and implications of commercial fishing. Science (80- ) 196, 324–326. doi:10.1126/science.847476 Tracey, S.R., Hartmann, K., Hobday, A.J., 2012. The effect of dispersal and temperature on the early life history of a temperate marine fish. Fish Oceanogr 21, 336–347. doi:10.1111/j.1365-2419.2012.00628.x Treml, E.A., Roberts, J.J., Chao, Y., Halpin, P.N., Possingham, H.P., Riginos, C., 2012. Reproductive output and duration of the pelagic larval stage determine seascape-wide connectivity of marine populations. Integr Comp Biol 52, 525–537. doi:10.1093/icb/ics101 Umgiesser, G., Canu, D. M., Cucco, A., & Solidoro, C., 2004. A finite element model for the Venice Lagoon. Development, set up, calibration and validation. Journal of Marine Systems, 51(1–4), 123–145. http://doi.org/10.1016/j.jmarsys.2004.05.009 Venerables, W., Smith, D., 2010. R Development Core Team. 2010. An Introduction to R. Notes on R: A Programming Environment for Data Analysis and Graphics Version 2.11. 1. Zecchetto, S., della Valle, A., De Biasio, F., Quattrocchi, G., Satta, A., Antognarelli, F., Cadau, E., Cucco, A., 2016. The wind-measuring system in the Gulf of Oristano: a support for coastal-scale oceanographic applications. J Oper Oceanogr 9, 144–154. doi:http://dx.doi.org/10.1080/1755876X.2015.1118806 Zuur, A.F., Ieno, E.N., Walker, N.J., Saveliev, A.A., Smith, G.H., 2009. Mixed effects models and extensions in ecology with R. Springer, New York. Zuur, A.F., Ieno, E.N., Elphick, C.S., 2010. A protocol for data exploration to avoid common statistical problems. Methods Ecol Evol 1, 3–14. doi:10.1111/j.2041-210X.2009.00001.x
AC C
852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888
ACCEPTED MANUSCRIPT Highlights 1. In Sardinia sustainable fishing of the sea urchin Paracentrotus lividus have become a needs 2. Sea urchin abundances are driven by the recruitment processes which dependent on coastal transport
RI PT
3. specific oceanographic variables are used to investigate the spatial variability of recruits 4. Recruitment is favored by low current speeds and by high confinement and trapping of the water masses
SC
5. Results contribute to the development of sustainable management strategies of artisanal
AC C
EP
TE D
M AN U
fishing