Wild gregarious settlements of Ostrea edulis in a semi‐enclosed Sea Lough; a case study for unassisted restoration

Smyth, D. M., Horne, N. S., Ronayne, E., Millar, R. V., Joyce, P. W. S., Hayden‐Hughes, M., & Kregting, L. (2020). Wild gregarious settlements of Ostrea edulis in a semi‐enclosed Sea Lough; a case study for unassisted restoration. Restoration Ecology. https://doi.org/10.1111/rec.13124

Published in: Restoration Ecology

Document Version: Peer reviewed version

Queen's University - Research Portal: Link to publication record in Queen's University Belfast Research Portal

Publisher rights © 2020 Society for Ecological Restoration. This work is made available online in accordance with the publisher’s policies. Please refer to any applicable terms of use of the publisher.

General rights Copyright for the publications made accessible via the Queen's University Belfast Research Portal is retained by the author(s) and / or other copyright owners and it is a condition of accessing these publications that users recognise and abide by the legal requirements associated with these rights.

Take policy The Research Portal is Queen's institutional repository that provides access to Queen's research output. Every effort has been made to ensure that content in the Research Portal does not infringe any person's rights, or applicable UK laws. If you discover content in the Research Portal that you believe breaches copyright or violates any law, please contact [email protected].

Download date:03. Oct. 2021 Ostrea edulis unassisted restoration

1 Wild gregarious settlements of Ostrea edulis in a semi-enclosed

2 Sea Lough; a case study for unassisted restoration

3

4 Running Head: Ostrea edulis unassisted restoration

5

6 Authors and Address: David M. Smythad*, Nicholas S. Hornea, Elli Ronayneb, Rachel

7 V. Millara, Patrick W.S. Joyceac, Maria Hayden-Hughesc and Louise Kregtinga

8 a School of Natural and Built Environment, Queen's University Belfast, Northern

9 Ireland,

10 b Marine and Freshwater Research Centre (MFRC), Galway-Mayo Institute of

11 Technology, County Galway, Ireland

12 c Danish Shellfish Centre, DTU Aqua, Øroddevej 80, 7900 Nykøbing Mors, Denmark

13 * Correspondence address: d School of Ocean Science, Bangor University Wales,

14 LL59 5AB. E-mail:[email protected].

15

16 Author Contributions: DS, LK conceived and designed the research; DS,LK,PJ

17 performed field surveys; DS,RM,NH,ER,MHH analysed the data; ER,RM,NH,LK

18 contributed materials/analysis tools; DS,LK,PJ wrote and edited the manuscript.

19

20

21

22

23

24

25

1

Ostrea edulis unassisted restoration

26 Abstract

27 Ostrea edulis was once prolific throughout Europe and considered as the

28 continent’s native oyster. However, O. edulis currently exists in small fragmented

29 assemblages where natural unaided recovery is rarely encountered. This research

30 identified the small semi-enclosed sea lough of on the north east coast of

31 as one of the few locations within Europe where the native oyster

32 displayed gregarious natural rejuvenation. On close examination, four influential

33 parameters appeared to assist in concentrated settlement; raised topographical cultch

34 formations, shell coverage, the number of fecund in-situ adult’s and site protection. If

35 these components were to be combined and managed as part of reintroduction and

36 restoration initiatives, high density settlements and self-sustaining populations may be

37 possible. The research also identified that unregulated harvesting of intertidal O. edulis

38 assemblages has the potential to seriously hinder natural recoveries. Indeed, the

39 findings suggest that a review of policy in regards to intertidal hand gathering is

40 necessary. However, naturally occurring high density settlements recorded during this

41 research should be inspirational to all involved in the restoration of the native oyster.

42

43 Keywords; high density settlements, multiple attachments, native oyster, spatting

44 pools, unregulated harvesting

45

46

47

48

49

50

2

Ostrea edulis unassisted restoration

51 Implications for Practice

52 • Examination of naturally occurring native oyster expansion and the

53 parameters which permit augmentation can provide an insight into which

54 components could be implemented to aid successful assisted restoration.

55 • A suite of abiotic and biotic conditions can accommodate mass gregarious

56 Ostrea edulis settlements; hydrodynamics which allow water retention,

57 adequate coverage of suitable substrate, intertidal topography which provides

58 pool structures, in-situ resident adults and policing of settlement sites.

59 • Unregulated harvesting remains a constant threat to restoration and

60 protection of the species needs to be implemented if a self-sustaining status is

61 to be achieved.

62

63

64 Introduction

65 Oysters as a species have been important biogenic and economic components

66 within estuarine, coastal and offshore habitats throughout the world (Coen & Grizzle

67 2007). The immensity of their topographical coverage and economic productivity

68 supported communities for centuries (Trimble et al. 2009). The Roman philosopher

69 Pliny describes the extensive harvesting of oysters in Brittany, France by Sergius

70 Orata at Baiae in 95 BC (Günther 1897; Marzano 2013). In the Persian Gulf circa 250

71 AD Babylonian fishing fleets were commercially exploiting the pearl oyster Pinctada

72 radiata (Carter 2005). The Gulf fishery developed into a considerable financial concern

73 which was valued at £240,000 per annum in 1829, a figure which represents

74 >£25,000,000 today. However, heavy exploitation resulted in the fishery being

3

Ostrea edulis unassisted restoration

75 economically non-viable by the late 1890s (Bowen 1951). Similarly, in the United

76 States of America during the mid-1800s, the Crassostrea virginica fishery was under

77 considerable strain (Kirby 2004). In 1869 the Maryland fishery landed >150,000 tonnes

78 of oysters worth $5,000,000, a current market equivalent of $80,000,000 (MacKenzie

79 2007). Likewise, in Europe in the mid-1800s, wild Ostrea edulis stocks were

80 undergoing a comparable level of exploitation (Eyton 1958). The scale of harvesting

81 was immense with shell piles at oyster shucking houses regularly containing >1 trillion

82 shells (Goulletquer & Heral 1997). The global demand for oysters led to the classic

83 overfishing scenario whereby demand outweighed resource, culminating in a

84 worldwide collapse of stocks by the early 1900s (Korringa 1946; Thurstan et al. 2013).

85 The rapid large-scale removal of these primary ecosystem components

86 resulted in several detrimental ecological consequences such as; a deterioration in

87 water quality, an increase in invasive pest species and a prevalence of new disease

88 (Conner & Simon 1979; Jackson et al. 2001). The anthropogenic removal of global

89 oyster reef ecosystems was almost total with an estimated 85% loss over the last 200

90 years accompanied with a legacy of 85% habitat loss (Beck et al. 2011).

91 Governments, habitat managers and Non-Governmental Organisations have

92 recognised and accepted that this loss has contributed to the accelerated deterioration

93 of the ecological health of numerous marine ecoregions. As a result, a consensus now

94 exists among concerned stakeholders that active restoration of oyster reefs should be

95 supported and encouraged whenever possible (Alleway & Connell 2015). Indeed, the

96 European Union has identified that biogenic oyster reefs under the Natura 2000 are

97 habitats which require immediate protection, augmentation and restoration (Bostock

98 et al. 2010). As the majority of historic European oyster sites are in a poor state of

99 conservational health, any locations which retain a resident self-sustaining population

4

Ostrea edulis unassisted restoration

100 could offer an important insight into environmental parameters which could be

101 replicated to promote recovery (Carlucci et al. 2010). Presently numerous European

102 countries are attempting to address the dwindling wild stocks through the

103 augmentation of cultured or translocated oysters (Smyth et al. 2018; Pogoda et al.

104 2019).

105 Even though many European O. edulis populations are extremely fragmented

106 and in low densities, the fecundity and life cycle of the oyster suggest that when abiotic

107 and biotic parameters are optimal the potential for large scale recovery should be

108 possible (Taylor & Bushek 2008). However, rejuvenation within wild O. edulis

109 assemblages is rare with only a few remote sites displaying tentative recovery.

110 Nevertheless, those which do exist could offer valuable indicators as to what in-situ

111 conditions are required to reinstate or support a restored population (Laing et al. 2006;

112 Christianen et al. 2018). One such location worthy of investigation is the small, semi-

113 enclosed sea Lough of Strangford on the north east coast of Northern Ireland.

114 The native oyster fishery in was once renowned throughout Ireland

115 for its abundant intertidal and subtidal beds (Day & McWilliams 1991). The

116 Montgomery Records of 1680 (Montgomery 1683) stated that, “the Strangford Lough

117 beds are dredged daily with oyster clumps gathered in deep water as well as on the

118 Lough shore”. Lewis (1837) when discussing the fishery referred to several sites that

119 were renowned for their productivity; Ringhaddy Sound, Cunningburn Milltown and

120 . However, even though Strangford possessed a significant oyster stock it

121 experienced the same overexploitation as the rest of Europe and was considered

122 functionally extinct by the early 1900s (Nunn 1994). Consequently only the occasional

123 solitary individual was recorded up until the late 1990s (Kennedy & Roberts 1999).

124 Though, in 1998 a single spawning event of an O. edulis aquaculture assemblage

5

Ostrea edulis unassisted restoration

125 triggered an unintentional rejuvenation of the wild stock (Kennedy & Roberts 1999).

126 The spawning resulted in the Lough’s population growing from an estimated 100

127 individuals to > 1.5 million in a period of five years (Smyth et al. 2009). Unfortunately,

128 the rapid increase of wild oysters was accompanied by an upsurge in unregulated

129 commercial harvesting and over the following three years the population declined to

130 <400,000 (Smyth et al. 2009). Nonetheless, the potential for gregarious oyster

131 settlements within the Lough was established and a hydrodynamic larval tracking

132 model confirmed the north east coast as a potential O. edulis larval sink (Figure 1)

133 (Smyth et al. 2016).

134 The Lough has benefited from being studied since the early 1960s and is a

135 designated Marine Conservation Zone (Roberts et al. 2011). Intense trawling activity

136 over Modiolus modiolus reefs in the late 1990s was suspected of causing damage to

137 highly diverse biotopes in the north of the Lough (NIOA 2015). As a result the local

138 Government introduced a closed fishing zone in 2008 in an attempt to initiate natural

139 recovery (Roberts et al. 2011). The introduction of this closed zone increased the

140 activities of government enforcement officers and unintentionally the protection of the

141 small remaining resident O. edulis assemblages.

142 Owing to an examination of hydrodynamic regimens in the north of the Lough,

143 and recently reported increases in O. edulis densities (Smyth et al. 2009). It was

144 decided to investigate if there had been any evidence of unassisted high density stock

145 recovery. If verified the conditions under which recovery occurred could be examined

146 with the potential to provide test hypotheses for replication in future restoration

147 programmes. The criteria behind choosing the northern region of Strangford is based

148 on the presumption that established assemblages of oysters and increased policing of

6

Ostrea edulis unassisted restoration

149 the marine environment would inadvertently lead to a rejuvenation and augmentation

150 of wild O. edulis stocks.

151

152 Materials and Methods

153 Study Area

154 Strangford Lough is located on the northeast coast of Northern Ireland and lies

155 between 54o 35’ N and 54o 20’ N and between 5o 41’ W and 5o 34’ W (Figure 1)

156 enclosing an area of 150 km2. The depth ranges from 14-60 m, with substrate varying

157 from bedrock to fine sediments (Kregting & Elsäβer 2014). The Lough can be divided

158 into northern and southern basins. The high retention hydrodynamic regimen in the

159 northern basin is particularly accommodating to larval pooling (Smyth et al. 2016). The

160 entrance to the southern basin is a long, narrow channel, known as The Narrows,

161 where current tidal velocity reaches ~3.5 m/s (Kregting & Elsäβer 2014).

162

163 Site Selection

164 Historical provenance, and recent oyster survey results (Day & McWilliams

165 1991; Smyth et al. 2009; Roberts et al. 2011; Guy et al. 2018) led to the northern basin

166 of Strangford Lough being selected as an area of possible undetected O. edulis bed

167 rejuvenation. Potential sites within the region had to meet habitat suitability

168 requirements as previously outlined in Kennedy and Roberts (1999) these included;

169 sufficient natural shell substrate to accommodate broad-scale settlement, an adequate

170 number of adult oysters for the conveyance of settlement cues and hydrodynamics

171 which would support larval retention. In studies over the preceding 12 years six sites

172 within the northern basin were identified as meeting these requirements to varying

173 degrees.

7

Ostrea edulis unassisted restoration

174 The sites of Ballyreagh, Cunningburn and Greyabbey consistently produced

175 high density oyster settlements on a Mytilus edulis shell dominant substrate (Figures

176 2 to 5). Kircubbin, Pig Island and Ringhaddy where considered low density O. edulis

177 sites in comparison with a predominantly mud and shell fragment substrate (Kennedy

178 & Roberts 1999; Smyth et al. 2018). All sites were intertidal as extensive dive surveys

179 in the region between 2008 and 2011 had shown the subtidal substrate to consist of

180 heavy mud absent of a shell mix and therefore unsuitable for O. edulis larval settlement

181 (Roberts et al. 2011). The one exception was Ringhaddy a subtidal site mentioned in

182 the historical accounts of Lewis (1837). However, when examined in 2010 (Roberts et

183 al. 2011) any available hard substrate had succumbed to heavy siltation leaving

184 potential settlement surfaces in a poor state to accommodate high density larval

185 attachments, thus the site was sampled along the intertidal only.

186 All sites were located in the northern basin of the lough in close proximity to a

187 governmental Department of Agriculture sub-surface coastal monitoring buoy. The

188 recorded physical parameters which encompassed all waters within the survey sites

189 hydrodynamic regime included; water temperature of 2–17.6 °C, a mean salinity of 33,

190 mean nutrient concentrations (µmol l-1) of 2.8 ammonium, nitrate 13.5,

191 phosphorus 2, and silicate 4.3 and a mean nutrient load (ton year-1) of 1,202

192 nitrogen and 126 phosphorus (www.afbini.gov.uk/costal-monitoring).

193

194 Survey Techniques

195 Data collection from 2002-2014 were conducted as per those outlined in

196 Kennedy & Roberts (1999) the protocol was replicated again in 2018. Surveys were

197 undertaken in November during low spring tides of <0.5 m below chart datum. A

198 random belt transect and timed search methodology was employed with sampling

8

Ostrea edulis unassisted restoration

199 taking place parallel to the low water mark within three 30  10m plots. Digital still

200 images were taken in-situ of gregarious settlement to record any O. edulis clusters (>2

201 oysters adhered together) known locally as ‘clocks’ (Figure 4). Oyster size was

202 recorded in-situ using a Vernier© caliper with measurements taken from the umbo to

203 the ventral front edge of the shell.

204 The influence of topographical variations on micro-hydrodynamics and O. edulis

205 larval settlement success has seldom been researched. During the 16-year survey

206 period, accumulations of conjoined oysters were recorded in high densities at sites

207 where high elevations of shell cultch dipped to form depressions within the shell

208 substrate. Therefore, a simple snap-shot topographical survey of site plots was

209 undertaken using a Nedo © F-series builders level and two York© ranging poles. This

210 allowed for the quantification of noticeable elevations and depressions of substrate

211 coverage from the horizontal plane (Figure 5 a-d). An area of >3m2 of elevation /

212 depression was required at sites to be considered for analysis.

213

214 Data Analysis

215 Multi-variate PERMANOVA of Bray-Curtis similarity matrices was used throughout

216 with 9999 permutations to determine the similarities of square root transformed data

217 of single and conjoined oyster settlements per; site, year, percent shell cover /m2,

218 harvesting protection and snapshot topographical features. The total area coverage

219 m2 of conjoined oysters per site and year was also examined using the same analytical

220 techniques. All statistical test were undertaken using PAST 3.25© (Hammer et al.

221 2001).

222

223

9

Ostrea edulis unassisted restoration

224 RESULTS

225 Single Settlement

226 The number of single settlements over the survey period was shown to be significantly

227 different between sites (F = 14.58, P = 0.0001). Pairwise post-hoc analysis (Table S1)

228 identified Ballyreagh, Kircubbin, Ringhaddy and Pig Island as significantly different (P

229 < 0.0005) from all sites. Differences in the number of single settlements at

230 Cunningburn and Greyabbey proved to be non-significant (P > 0.1). Differences in the

231 number of single O. edulis settlements per year over the survey period was shown not

232 to be significant (F = 1.72, P > 0.1).

233 The most northerly site of Ballyreagh produced the highest settlement densities

234 between 2002 and 2010 for single attachments over all five size cohorts (0-29, 30-59,

235 60-89, 90-119, 120-149 mm). However, settlements at Ballyreagh regressed rapidly

236 between 2014 and 2018. Cunningburn and Greyabbey proved to be the most prolific

237 single settlement sites post-2010 (Figure 2). Kircubbin, Pig Island and Ringhaddy

238 remained low density settlement sites throughout the survey period (Figure 2).

239

240 Multiple Attachment

241 The number of multiple attachments/clocks in relation to site during the survey period

242 was shown to be significant (F = 3.28, P < 0.005) (Figure 3). Pairwise post-hoc

243 analysis identified Cunningburn, Kircubbin, Ringhaddy and Pig Island as significantly

244 different (P < 0.05) (Table S2).

245 A significant difference (F = 2.08, P < 0.05) was detected in the number of multiple

246 attachments/clocks in relation to year, post-hoc analysis identified 2018 as the

247 survey year which was significantly different (P < 0.05) (Table S3).

10

Ostrea edulis unassisted restoration

248 The total area of coverage in m2 attributed to conjoined oyster attachment was shown

249 to be significantly different (F = 1.58, P < 0.005) between sites. Post-hoc analysis

250 revealed Cunningburn, Kircubbin, Ringhaddy and Pig Island to be significantly

251 different (P < 0.05) in area m2 of conjoined settlement (Table S4). Analysis of area /m2

252 of conjoined oyster settlement and year proved not to be significant (F = 1.08, P =

253 0.3631).

254

255 Percentage of Shell Coverage

256 The percentage of available shell substrate within a site has been recognised by

257 numerous researchers as an influencing factor in the settlement success of O. edulis

258 (Bayne 1969; Kennedy & Roberts 1999; Christianen et al. 2018). Analysis revealed a

259 highly significant difference (F = 5.296, P = 0.0001) between the total percentage of

260 shell cover (no differentiation was made between shell types) per site and the total

261 density of annual oyster settlement. A pairwise post-hoc identified Kircubbin,

262 Ringhaddy and Pig Island as being significantly different (P <0.05) (Table S5).

263

264 Site Protection from Harvesting

265 Active site protection of commercial molluscan assemblages is a proven management

266 measure which can deter unregulated harvesting (Crawford et al. 2010). During the

267 surveys Cunningburn had an assemblage of O. edulis within a protected site. Site

268 protection and settlement densities per site revealed a significant difference (F =

269 4.041, P = 0.012).

270

11

Ostrea edulis unassisted restoration

271 Density of Fecund Oysters (>60-150mm) in Relation to Total Density

272 The average total density per site of oysters considered fecund (>60mm/year 2, based

273 on Walne1964) was compared in relation to total overall density per site and year no

274 significant difference was detected (F= 8.066, P= 0.061).

275

276 Substrate Topography

277 The survey identified four sites with topographical height variations from the horizontal

278 plane of > 3m2 in area: Cunningburn 68-72mm, Ballyreagh 27-32mm, Pig Island

279 10mm, Kircubbin 10-14mm. No variations in shell height from the horizontal plane

280 were detected at Greyabbey and Ringhaddy. A significant difference (F= 3.029, P=

281 0.0003) was detected in oyster clump/clock density in relation to topographical height.

282 Post-hoc analysis showed Cunningburn to be significantly different (P <0.05) from the

283 remaining five sites and Pig Island significantly different (P <0.05) from Greyabbey

284 and Ringhaddy (Table S6).

285

286 Discussion

287 At Strangford Lough the site of Cunningburn was identified as accommodating

288 the most gregarious settlements over the 16 year survey period. The most prolific

289 solitary and clumped O. edulis settlement areas were all located on substrate

290 complexes dominated by shell. The lowest oyster densities were at sites where the

291 predominant substrate coverage was mixed mud. The most northerly site of

292 Ballyreagh showed high density settlements pre-2010 but experienced a dramatic

293 decline post-2006 particularly within the 30-59 mm size cohort. O. edulis densities

294 continued to decline at Ballyreagh throughout the survey period. With the site

12

Ostrea edulis unassisted restoration

295 eventually becoming the least populated of the six by 2018. The three low density

296 sites, Ringhaddy, Kircubbin and Pig Island remained relatively static in regards to O.

297 edulis settlements throughout the 16 year period.

298 One of the most striking changes in the Strangford Lough oyster population was

299 the rapid density decline observed at Ballyreagh post-2006. Prior to 2002 (Nunn 1994;

300 Kennedy & Roberts 1999) no historical accounts of O. edulis settlements existed for

301 Ballyreagh. However, by 2004 an estimated population of >300,000 O. edulis had

302 settled in densities of >30 m2 (Smyth 2009) although by 2018 <20 individuals

303 remained. Pathogen or pollution induced mortalities can be disregarded as an

304 explanation for population decline as no empty oyster valves were recorded and shell

305 debris would have lingered within the low-velocity, low–flush regimen. Numerous

306 witnessed reports from members of the general public of shellfish gatherers in the

307 vicinity of Ballyreagh between 2004 and 2006 suggest that sporadic subsistence and

308 commercial harvesting was the overriding influencer in population declines (Smyth et

309 al. 2009).

310 Oyster densities at the remaining five sites fluctuated over the 16 years though

311 not to the extent experienced at Ballyreagh this may have been because of an increase

312 in policing. However, it would be assumptive to attribute the maintenance of oyster

313 densities to policing alone. It may be simply that stocks had become non-viable for

314 harvesting due to the low Catch Per Unit Effort (CPUE). Similar harvesting scenarios

315 have been described for other intertidally accessible species; Amiantis purpurata,

316 Cymbula granatina, Lucina pectinata, Panopea abrupta, Perna perno (Kyle et al. 1997;

317 Tomalin et al. 1998; Rondinelli & Barros, 2010). Continuous low-level subsistence

318 gathering of at sites may be impeding recovery and keeping standing stocks stagnant.

319 Monthly observations of accessible assemblages of market sized oysters would assist

13

Ostrea edulis unassisted restoration

320 in the monitoring of shellfish gathering and give a better understanding to the

321 population dynamics of native oysters within Strangford.

322 A further detrimental effect of harvesting pressure on mature oysters is the

323 fragmentation of fecund standing stock (Rondinelli & Barros 2010). This can result in

324 a compromised Allee effect and a decrease in successful spawns. Eventually leaving

325 population dynamics inert and as natural mortality ensues the species eventually

326 becomes reproductively non-functional (Marshall et al. 2003; Turra et al. 2016). Guy

327 et al. (2018) showed that O. edulis is particularly susceptible to this chain of events,

328 as a nearest neighbour of < 1.5 m brooded significantly more larvae than those with a

329 nearest neighbour of > 1.5 m. The findings at the low-density sites and the heavily

330 harvested site of Ballyreagh concur with Guy et al. (2018). As adult oysters were

331 removed from the population post-2006 recruitment slowed significantly with the

332 Ballyreagh stock virtually non-existent by 2018.

333 During the research a significant increase in the gregarious nature of larval

334 settlement at Cunningburn was noticed. The site produced settlement complexes of >

335 16 conjoined oysters and a total area coverage of clumps within the survey plots of

336 6.8 m2. This was unique among survey sites and represents the first documented

337 occurrence of conjoined O. edulis settlements in this density in Ireland. Cunningburn

338 had site properties similar to both Ballyreagh and Greyabbey however the volume of

339 post-2014 oyster clump settlements suggested site specific properties had influenced

340 the attachments. A comparison between sites identified Cunningburn as possessing

341 some distinctive characteristics which may have influenced the gregarious behaviour

342 most notably; topographical shell formations, % shell substrate coverage, number of

343 in-situ fecund adults and protection from harvesting.

14

Ostrea edulis unassisted restoration

344 The topographical nature of the shell substrate accumulations at Cunningburn

345 were markedly different from the other sites. Although shell coverage /m2 was similar

346 at Greyabbey and Ballyreagh the M. edulis concentrations at Cunningburn formed

347 undulating banks which created a network of large intertidal pools. These pools held

348 the majority of oysters found at the site and appeared to have acted as natural spatting

349 ponds. It is possible that the formations provided a containment area for larvae to

350 concentrate in high densities thereby increasing the potential for focused gregarious

351 settlement within these Cunningburn pools, which would have several advantages to

352 reduce mortality. The ponds would provide protection against; air exposure over tidal

353 cycles, low winter temperatures and visual detection from harvesters.

354 The use of natural spatting ponds or “polls” in oyster culture has been

355 considerable, with records dating as far back as A.D. 43. when Romans stocked

356 naturally occurring pools in Northern France to act as primitive mariculture systems

357 (Horváth et al. 2012). Historically, the spatting pond was the stock enhancement

358 technique of choice in many regions of Europe (Hedgecock & Davis 2007). However,

359 concerns regarding the genetic fitness of the progeny, high density disease transfer

360 and the development of improved hatchery culture led to their demise (Lallias et al.

361 2010). The intertidal pool structure and high-density settlements discovered during this

362 study present a hypothesis worthy of testing for future native oyster

363 rejuvenation/reintroduction programmes. If a section within a restoration cultch lay was

364 formed into an undulating bank of shell it could provide the micro hydro-dynamics

365 suitable to replicate the larval settlement at Cunningburn.

366 The availability of shell, as opposed to shell type, has been shown to be

367 influential in settlement success (Smyth et al. 2018). A significant higher density of O.

368 edulis attachments were discovered at Cunningburn post-2014 when compared to

15

Ostrea edulis unassisted restoration

369 other sites. In contrast to the findings of Smyth et al. (2018) the survey showed that

370 when available, at Strangford Lough O. edulis pediveligers attachment material of

371 choice was the living edge of a live oyster. Researchers have emphasised this fact

372 previously in relation to larval settlements within Ostrediae spat (Breitburg et al. 2000;

373 Tamburri et al. 2008).

374 The high density of oysters >60mm at Cunningburn and their associated

375 conspecific chemical cues may have also had an influence on settlement behaviour

376 (Rodriguez et al. 2019). An overview of the three hypotheses related to O. edulis and

377 adult density related chemical cues can help explain. The first indicates conspecifics

378 as the source therefore a high-density mature population is required (Keck et al. 1971).

379 The second suggests the importance of biofilms on oyster shells, therefore high

380 density of shell material is required (Bonar et al. 1990). The third identifies high

381 densities of both live oysters and biofilms on shell, as producing settlement inducing

382 cues which act synergistically (Tamburri et al. 2008). The O. edulis aquaculture and

383 mariculture sector concurs with the hypothesis of high densities of conspecifics as an

384 explanation for increased oyster settlements (Hugh-Jones 2003; Galley et al. 2017).

385 Oyster growers have long considered live and dead oyster valves as the most effective

386 settlement surfaces (Yamaguchi 1994). At the Loch Ryan O. edullis fishery in

387 Scotland, >100 spat have been recorded on a single live oyster (Hugh-Jones 2003).

388 In Ireland, at the Bertraghboy fishery, >80 spat have been recorded regularly on wild

389 O. edulis (Barry 1981). There are similar reports from aquaculturists throughout

390 Europe of high density spat attachments on living O. edulis within hatcheries

391 (Drinkwaard 1998; Culloty & Mulcahy, 2007). The findings of this present research are

392 certainly worthy of further investigation as the densities of live adult O. edulis may be

393 an important factor when attempting to establish gregarious settlement events.

16

Ostrea edulis unassisted restoration

394 The active protection from illicit harvesting at Cunningburn was considered as

395 a dominant influence. The location of the O. edulis population is in close proximity to

396 a private yacht club where trespassers onto the adjacent foreshore are actively

397 challenged by club members. This acts as a powerful deterrent. Nevertheless, as wild

398 stocks at other unrestricted sites decline and CPUE decrease, the concentrated

399 assemblages at Cunningburn will become an attractive proposition worthy of trespass

400 for unscrupulous harvesters. Unregulated harvesting events such as those described

401 at Ballyreagh where, 300,000 individuals were removed, can have a devastating

402 impact on a conservation project (Bostock et al. 2010; Guy et al. 2018). Presently

403 throughout Europe a significant amount of time, effort and finance have been

404 committed to restoring O. edulis to a self-sustaining status (Helmer et al. 2019; Pogoda

405 et al. 2019). It is vital that NGO’s, Habitat Managers and Governments recognise the

406 detrimental consequences of mass removal events and produce preventative

407 measures to counter such actions.

408 Historical European native oyster regions, experience a suite of hindrances in

409 regards to recruitment and reintroduction; a lack of effective larval density, water

410 quality, disease, anthropogenic disturbance and substrate availability (Laing et al.

411 2006). The observations described during this survey have highlighted a number of

412 aspects of O. edulis settlement dynamics which are not fully understood and a

413 considerable knowledge gap exists which therefore warrant trial when investigating

414 restorative gregarious, O. edulis settlements.

415 It may be, that historical accounts such as those described by Orton (1937)

416 which refer to the North Sea grounds as; “being built of oysters, knitted and interlaced

417 with countless other invertebrates with the bottom hardened as a living crust”

418 (Houziaux et al. 2008) are beyond the capabilities of present restoration initiatives.

17

Ostrea edulis unassisted restoration

419 However, the tentative high-density bed settlements and the number of clumped

420 oysters recorded during this research should be considered inspirational to all those

421 involved in the re-establishment of Ostrea edulis. As the task which was once

422 considered almost unachievable has now been revealed as possibly attainable.

423

424

425

426 Acknowledgements: The authors would like to thank the following funding bodies for

427 making this work possible: EPSRC Engineering and Physical Sciences Research

428 Council, the Marine Institute Ireland, the Department for the Economy Northern

429 Ireland, the Bryden Centre, the Departments of Employment and Learning, and

430 Agriculture and Rural Development (EU Building Sustainable Development

431 Programme) Northern Ireland.

432

433

434 Reference

435 Alleway HK, Connell SD (2015) Loss of an ecological baseline through the eradication

436 of oyster reefs from coastal ecosystems and human memory. Conservation

437 Biology 29:795–804

438 Barry MD (1981) Distribution and ecology of oysters, Ostrea edulis (L.) in Kilkieran

439 and Bertraghboy Bays, Connemara, Co. Galway. Report to the Department of

440 Fishery Ireland.

441 Bayne B (1969) The gregarious behaviour of the larvae of Ostrea edulis L. at

442 settlement. Journal of the Marine Biological Association of the United Kingdom

443 49: 327-356

18

Ostrea edulis unassisted restoration

444 Beck MW, Brumbaugh RD, Airoldi L, Carranza A, Coen LD, Crawford C, … Guo X (2011)

445 Oyster Reefs at Risk and Recommendations for Conservation, Restoration, and

446 Management. BioScience 61: 107–116

447 Bonar DB, Coon SL, Walch M, Weiner RM, Fitt W (1990) Control of oyster settlement

448 and metamorphosis by endogenous and exogenous chemical cues. Bulletin of

449 Marine Science 46: 484-498

450 Bostock J, McAndrew B, Richards R, Jauncey K, Telfer T, Lorenzen K, Little D,

451 Ross L, Handisyde N, Gatward I, Corner R (2010) Aquaculture: global status

452 and trends. Philosophical Transactions of the Royal Society B: Biological

453 Sciences 365: 2897–2912

454 Bowen RL (1951) Pearl fisheries of the Persian Gulf. Middle East Journal 5:161

455 Breitburg DL, Coen LD, Luckenbach MW, Mann R, Posey M, Wesson JA (2000)

456 Oyster reef restoration: convergence of harvest and conservation strategies.

457 Journal of Shellfish Research 19:371-377

458 Carlucci R, Sassanelli G, Matarrese A, Giove A, D’Onghia G (2010) Experimental

459 data on growth, mortality and reproduction of Ostrea edulis (L., 1758) in a

460 semi-enclosed basin of the Mediterranean Sea. Aquaculture 306(1):167–176

461 Carter R (2005) The history and prehistory of pearling in the Persian Gulf. Journal of

462 the Economic and Social History of the Orient 48:139-209

463 Christianen MJA, Bergsma JH, Coolen JWP, Didderen K, Dorenbosch M, Driessen

464 FMF, Kamermans P, Reuchlin-Hugenholtz E, Sas H, Smaal A et al. (2018)

465 Return of the native facilitated by the invasive? Population composition,

466 substrate preferences and epibenthic species richness of a recently discovered

467 shellfish reef with native European flat oysters (Ostrea edulis) in the North

468 Sea. Marine Biology Research 14: 590–597

19

Ostrea edulis unassisted restoration

469 Coen LD, Grizzle RE (2007) The Importance of Habitat Created by Molluscan along

470 the Atlantic Coast of the United States. Habitat Management Series 8: 108-110

471 Conner W, Simon J (1979) The effects of oyster shell dredging on an estuarine benthic

472 community. Estuarine and Coastal Marine Science 9: 749-758

473 Crawford B, Herrera M D, Hernandez N, Leclair C R, Jiddawi N, Masumbuko S, Haws

474 M (2010) Small Scale Fisheries Management: Lessons from Cockle Harvesters

475 in Nicaragua and Tanzania. Coastal Management 38:195–215

476 Culloty SC, Mulcahy MF (2007) Bonamia ostreae in the Native oyster Ostrea edulis.

477 Marine Environment and Health Series No. 29, Marine Institute Ireland.

478 Day A, McWilliams P (1991) Ordnance Survey Memoirs of Ireland, Parishes of County

479 Down II 1832-4, 1837 North Down & the . Belfast: The Royal

480 Irish Academy 10: 88-89

481 Drinkward AC (1998) Introductions and developments of oysters in the North Sea

482 area: a review. Helgoland Marine Research 52: 301-303

483 Eyton TC (1958) A History of the Oyster and the Oyster Fisheries. London: John Van

484 Voorst.

485 Galley H, Beaumont AR, Vay LL, King JW (2017) Influence of exogenous chemicals

486 on larval development and survival of the king scallop Pecten maximus (L.)

487 Aquaculture 474: 48-56

488 Goulletquer P, Heral M (1997) Marine Molluscaii Production Trends in France: From

489 Fisheries to Aquaculture, 1:37- 164. Gommer, NOAA Tech. Rep. NMFS 129

490 Günther RT (1897) The Oyster Culture of the Ancient Romans. Journal of the Marine

491 Biological Association 4: 360–365

20

Ostrea edulis unassisted restoration

492 Guy C, Smyth D, Robert D (2018) The importance of population density and inter-

493 individual distance in conserving the European oyster Ostrea edulis. Journal of

494 the Marine Biological Association of the United Kingdom 15: 1-7

495 Hammer Ø, Harper DAT, Ryan D (2001) PAST: Paleontological statistics software

496 package for education and data analysis. Palaeontologia Electronica 4: 9-10

497 Hedgecock D, Davis JP (2007) Heterosis for yield and crossbreeding of the Pacific

498 oyster Crassostrea gigas. Aquaculture 272: 17–29

499 Helmer L, Farrell P, Hendy I, Harding S, Robertson M, Preston J (2019) Active

500 management is required to turn the tide for depleted Ostrea edulis stocks from

501 the effects of overfishing, disease and invasive species PeerJ 7, e6431

502 Horváth Á, Bubalo A, Čučević A, Bartulović V, Kotrik L, Urbányi B, Glamuzina B (2012)

503 Cryopreservation of sperm and larvae of the European flat oyster (Ostrea

504 edulis). Journal of Applied Ichthyology 28: 948-951

505 Houziaux JS, Kerckhof F, Degrendele K, Roche M, Norro A (2008) The Hinder Banks:

506 Yet an important region for the Belgian marine biodiversity? Final report

507 HINDERS project, Belgian Science Policy Office.

508 Hugh-Jones T (2003) The Loch Ryan Native Oyster Fishery. Shellfish News 15: 7-18

509 Jackson JBC, Kirby MX, Berger WH, Bjorndal KA, Botsford LW, Bourque J, Bradbury

510 RH, Cooke R, Erlandson J, Estes A, Hughes T, Kidwell S, Lange C, Lenihan

511 HS, Pandolfi JM, Peterson CH, Steneck RS, Tegner MJ, Warner RR (2001)

512 Historical Overfishing and the Recent Collapse of Coastal Ecosystems.

513 Science 293: 629 637

514 Keck R, Mauer D, Kauer JC, Sheppard WA (1971) Chemical stimulants affecting larval

515 settlement in the American oyster. Proceedings of the National Shellfish

516 Association 61: 24–28

21

Ostrea edulis unassisted restoration

517 Kennedy R, Roberts D (1999) A Survey of the current status of the flat oyster Ostrea

518 edulis in Strangford Lough, Northern Ireland, with a view to the restoration of

519 its oyster beds. Biology and Environment: Proceedings of the Royal Irish

520 Academy 99(B):79-88

521 Kirby MX (2004) Fishing down the coast: Historical expansion and collapse of oyster

522 fisheries along continental margins. Proceedings of the National Academy of

523 Sciences of the United States of America 101(35): 13096-13099

524 Korringa P (1946) The decline of natural oyster beds. Governmental Institute of

525 Fisheries Research Bergen-op-Zoom 10: 36–41

526 Kregting L, Elsäßer B (2014) A hydrodynamic modelling framework for Strangford

527 Lough part 1: tidal model. Journal of Marine Science and Engineering 2: 46-65

528 Kyle R, Robertson WD, Birnie SL (1997) Subsistence shellfish harvesting in the

529 Maputuland Marine Reserve, northern KwaZulu-Natal, South Africa; sandy

530 beach Organisms. Biological Conservation 82:173-182

531 Laing I, Walker P, Areal F (2006) Return of the native – is European oyster (Ostrea

532 edulis) stock restoration in the UK feasible? Aquatic Living Resources 19: 283-

533 287

534 Lallias D, Boudry P, Lapègue S, King JW, Beaumont AR (2010) Strategies for the

535 retention of high genetic variability in European flat oyster (Ostrea edulis)

536 restoration programmes. Conservation Genetics 11:1899–1910

537 Lewis H (1837) A Topographical Dictionary of Ireland. London Press

538 Mackenzie CL (2007) Causes underlying the historical decline in eastern oyster

539 (Crassostrea virginica Gmelin, 1791) landings. Journal of Shellfish Research

540 26: 927–938

22

Ostrea edulis unassisted restoration

541 Marshall DJ, Bolton TF, Keough MJ (2003) Offspring size affects the post‐

542 metamorphic performance of a colonial marine invertebrate. Ecology 84: 3131-

543 3137

544 Marzano A (2013) Harvesting the Sea: The Exploitation of Marine Resources in the

545 Roman Mediterranean. Oxford: Oxford University Press.

546 Montgomery H (1683) The Montgomery Manuscripts. Linehall library archives Belfast.

547 NIAO (2015) Protecting Strangford Lough. A report by the auditor and controller of

548 the Northern Ireland Audit Office http://www.niauditoffice.gov.uk/index/pub/recent

549

550 Nunn J (1994) The marine mollusca of Ireland 1. Strangford Lough. Bulletin of the Irish

551 Biogeographical Society 17: 23-282

552 Orton JH (1937) Oyster Biology and Oyster Culture. Being the Buckland Lectures for

553 1935. London: Edward Arnold & Co.

554 Pogoda B, Brown J, Hancock B, Preston J, Pouvreau S, Kamermans P, Sanderson

555 W, von Nordheim H (2019) The Native Oyster Restoration Alliance (NORA) and

556 the Berlin Oyster Recommendation: bringing back a key ecosystem engineer

557 by developing and supporting best practice in Europe. Aquatic Living

558 Resources 32: 13-22

559 Roberts D, Allcock L, Fariñas-Franco JM, Gorman E, Maggs CA, Mahon AM, Smyth

560 D, Strain E, Wilson CD (2011) Modiolus Restoration Research Project: Final

561 Report and Recommendations. Report prepared for Northern Ireland

562 Environmental Agency.

563 Rodriguez-Perez A, James M, Donnan D, Henrya T, Møllere F, Sanderson WG (2019)

23

Ostrea edulis unassisted restoration

564 Conservation and restoration of a keystone species: understanding the

565 settlement preferences of the European oyster (Ostrea edulis) Marine Pollution

566 Bulletin 138: 312-321

567 Rondinelli SFF, Barros F (2010) Evaluating shellfish gathering (Lucina pectinata) in a

568 tropical mangrove system. Journal of Sea Research 64:.401–407

569 Smyth D, Roberts D, Browne L (2009) Impacts of unregulated harvesting on a

570 recovering stock of native oysters (Ostrea edulis). Marine Pollution Bulletin 58:

571 916-922

572 Smyth D, Kregting L, Elsäßer B, Kennedy R, Roberts D (2016) Using particle dispersal

573 models to assist in the conservation and recovery of the overexploited native

574 oyster (Ostrea edulis) in an enclosed sea lough. Journal of Sea Research 108:

575 50-59

576 Smyth D, Mahon AM, Roberts D, Kregting L (2018) Settlement of Ostrea edulis is

577 determined by the availability of hard substrata rather than its nature:

578 implications for stock recovery and restoration of the European oyster. Aquatic

579 Conservation Marine and Freshwater Ecosystems 28: 662-671

580 Tamburri MN, Luckenbach MW, Breitburg DL, Bonniwell SM (2008) Settlement of

581 Crassostrea ariakensis larvae: effects of substrate, biofilms, sediment and adult

582 chemical cues. Journal of Shellfish Research 27: 601-608

583 Taylor J, Bushek D (2008) Intertidal oyster reefs can persist and function in a

584 temperate North American Atlantic estuary. Marine Ecology Progress Series

585 361: 301-306

586 Thurstan RH, Hawkins JP, Raby L, Roberts CM (2013) Oyster (Ostrea edulis)

587 extirpation and ecosystem transformation in the Firth of Forth, Scotland. Journal

588 for Nature Conservation 21: 253–261

24

Ostrea edulis unassisted restoration

589 Tomalin BJ, Kyle R (1998) Subsistence and recreational mussel (Perna perno)

590 collecting in KwaZulu-Natal, South Africa: fishing mortality and precautionary

591 management. South African Journal of Zoology 33: 12–22

592 Trimble AC, Ruesink JL, Dumbauld BR (2009) Factors preventing the recovery of a

593 historically overexploited shellfish species, Ostrea lurida Carpenter 1864.

594 Journal of Shellfish Research 28: 97-106

595 Turra A, Xavier LY, Pombo M, Paschoal CC, Denadai MR (2016) Assessment of

596 recreational harvesting of the trigonal clam Tivela mactroides: socioeconomic

597 aspects and environmental perception, Fisheries Research 174: 58-67

598 Yamaguchi K (1994) Shell structure and behaviour related to cementation in oysters.

599 Marine Biology 118: 89–100

600 Walne P (1964) Observations on the Fertility of the Oyster (Ostrea Edulis). Journal of

601 the Marine Biological Association of the United Kingdom, 44: 293-310

602 www. afbini.gov.uk/coastalmonitoring.htm.

603

604

605

606

607

608

609

610

611

612

613

25

Ostrea edulis unassisted restoration

614 FIGURES

615 Figure 1. Strangford Lough and the six survey sites with substrate constitution at

616 sample plots.

617

618 Figure 2. Individual Ostrea edulis density per size cohort at sites between 2002 and 619 2018. (Site codes are: C- Cunningburn, B- Ballyreagh, G- Greyabbey, K- Kircubbin, R-Ringhaddy, P- Pig Island).

620

621 Figure 3. Total area m2 of multiple attachment Ostrea edulis settlement per site and

622 year.

623

624 Figure 4. Ostrea edulis clumps/clocks at intertidal pools Cunningburn

625

626 Figure 5. Lateral cross-section schematics and in-situ image of undulating mussel 627 bank (Mytilus edulis) formations at Cunningburn site.

628

629

26

Ostrea edulis unassisted restoration

630

631 Figure 1. Strangford Lough and the six survey sites with substrate constitution at

632 sample plots.

633

634

27

Ostrea edulis unassisted restoration

635

636

Ostrea edulis Size Cohorts 2002-2018 180

160 0-29 mm 140 30-59 mm 120 60-89 mm 100 90-119 mm

80 120-149 mm

60

Number single of attachments 40

20

0

B/02 P/06 K/14 K/02 P/02 B/06 K/06 B/10 K/10 P/10 B/14 P/14 B/18 K/18 P/18

C/06 R/10 C/02 R/02 R/06 C/10 C/14 R/14 C/18 R/18

G/18 G/02 G/06 G/10 G/14

637 Site 638

639 Figure 2. Individual Ostrea edulis density per size cohort at sites between 2002 and 640 2018 (Site codes are: C- Cunningburn, B- Ballyreagh, G- Greyabbey, K- Kircubbin, 641 R-Ringhaddy, P- Pig Island).

642

643

644

645

28

Ostrea edulis unassisted restoration

7.5 7 6.5 6 5.5 5

2 4.5 4 3.5

Area Area m 3 2.5 2 1.5 1 0.5

0

C/14 B/02 G/02 C/02 K/02 R/02 P/02 B/06 G/06 C/06 K/06 R/06 P/06 B/10 G/10 C/10 K/10 R/10 P/10 B/14 G/14 K/14 R/14 P/14 B/18 G/18 C/18 K/18 R/18 P/18

646 Site 647

648 Figure 3. Total area m2 of multiple attachment Ostrea edulis settlement per site and

649 year.

650

651

652

653

654

655

656

657

658

659

660

661

29

Ostrea edulis unassisted restoration

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

678

679

680

681

682

683

684 Figure 4. Ostrea edulis clumps/clocks at intertidal pools Cunningburn

685

686

30

Ostrea edulis unassisted restoration

687

688

689

690

691

692 a. Movement of O. edulis veliger and pediveliger larvae during incoming high spring tide 693

694

695

696

697 b. 698 O. edulis veliger and pediveliger containment during the ebbing tide 699

700

701

702

703 c. 704 O. edulis veliger and pediveliger become more densely concentrated within the 705 intertidal mussel pools during the lowest ebb. The undulating mussel banks maintain this level of tidal retention throughout the tidal cycle. 706

707

708

709

710

711 d. Figure 5. Lateral cross-section schematics and in-situ image of an undulating mussel bank (Mytilus edulis) formation at Cunningburn. 31