Next Article in Journal
Treatment of Wastewater Effluent with Heavy Metal Pollution Using a Nano Ecological Recycled Concrete
Next Article in Special Issue
Feeding Ecology of Common Squid Todarodes pacificus in the South Sea of Korea Determined through Stable Isotope and Stomach Content Analyses
Previous Article in Journal
Advances in Ecology Research on Integrated Rice Field Aquaculture in China
Previous Article in Special Issue
Characterization of Trophic Structure of Fish Assemblages in the East and South Seas of Korea Based on C and N Stable Isotope Ratios
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Recovery of Macrobenthic Food Web on Rocky Shores Following the Hebei Spirit Oil Spill as Revealed by C and N Stable Isotopes

1
Department of Marine Bioscience, Gangneung-Wonju National University, Gangneung 25457, Korea
2
School of Earth Sciences & Environmental Engineering, Gwangju Institute of Science and Technology, Gwangju 61005, Korea
*
Author to whom correspondence should be addressed.
Water 2022, 14(15), 2335; https://doi.org/10.3390/w14152335
Submission received: 27 June 2022 / Revised: 18 July 2022 / Accepted: 25 July 2022 / Published: 28 July 2022
(This article belongs to the Special Issue Application of Stable Isotopes in Marine Ecosystems)

Abstract

:
The impact of large-scale oil spills on organisms can lead to modifications of the food web structure. To assess the effects of the Hebei Spirit oil-spill accident on the trophic structure of the macrobenthic community on intertidal rocky shores along Taean Peninsula on the western coast of Republic of Korea 4 years after the Hebei Spirit oil spill, we analyzed carbon and nitrogen stable isotope ratios of macrobenthic consumers and their potential food sources in two heavily oil-impacted and one non-impacted sites. The results show no significant differences in isotopic ratios of feeding groups and their potential food sources between the polluted and reference sites, suggesting similar trophic structures given similar resource use by consumers. Similar isotopic niches and substantial overlap areas of feeding groups between the affected and reference sites suggest that the oil-impacted sites have re-achieved the trophic functions of the natural ecosystem. This study provides valuable information on the ecological processes of trophic recovery in coastal ecosystems impacted by oil spills.

1. Introduction

Coastal environments are threatened by a wide range of anthropogenic issues such as overfishing, eutrophication, chemical pollutants, and habitat destruction leading to the modification of the structure and function of coastal ecosystems [1]. Oil spills in oceans are serious environmental disasters with detrimental biological and socioeconomic effects that can adversely affect biotic communities at regional and national scales [2,3]. Large-scale oil spills can have long-term impacts on the survival and reproduction of marine organisms and on prey–predator interactions and can lead to habitat destruction, resulting in the delayed recovery of marine ecosystems because of the cascading direct/indirect effects [4,5,6]. Such impacts on individual species or populations can lead to the modification of the structure or dynamics of food webs via complex indirect effects through bottom–up or top–down cascades [7,8]. Accordingly, monitoring of long-term consequences of oil spills at ecosystem levels is essential for understanding responses through chronic and indirect long-term impacts, which can help assess the recovery process for coastal habitat protection and management [9,10].
The Hebei Spirit oil spill (HSOS) that occurred in December 2007 severely affected the entire ecosystem, as 10,900 tons of three different forms of crude oil (Iranian Heavy, United Arab Emirates Upper Zakum, and Kuwait Export) was released over 200 km along the western coast of Korea [11]. The spilled oil covered soft and rocky intertidal areas, resulting in toxic effects on living organisms, damage to the marine ecosystem, and collapse of the fisheries and mariculture industries [12]. Although the concentrations of petroleum hydrocarbons have now nearly decreased to pre-oil spill levels in most coastal areas, residual oils in sediments tend to exist for prolonged periods and are, thus, considered to have long-term impacts on various ecosystem components (especially benthic organisms) at individual and community levels [10,13]. After the HSOS, many studies have reported the status of residual oil in the environments, ecological toxicity in organisms, changes in community structure, and recovery processes through integrated ecosystem monitoring [10,13,14,15,16]. However, after the HSOS accident, despite the importance of understanding the functioning of the food web, including diverse trophic levels in the oil-impacted regions, there is little information on the trophic recovery of the benthic food web through quantitative or qualitative assessments of the resilience status of the coastal ecosystems [17].
The analysis of carbon and nitrogen stable isotope ratios can enable the tracking of trophic transfer of organic matter and can help in identifying the trophic structure of aquatic food webs by providing time-integrated information on consumers’ diet [18,19]. This analysis assumes that the stable isotope ratios of a consumer’s tissues reflect those of its dietary sources with predictable trophic enrichments over turnover periods by metabolic processes (≤1‰ for carbon and 2‰ to 4‰ for nitrogen) [20,21,22]. Several studies have demonstrated that organic matter pathways and trophic structures in coastal ecosystems can be influenced by a variety of anthropogenic activities [23,24,25]. Therefore, monitoring isotope ratios of organic matter sources and animals as major ecosystem components may allow us to assess the ecological effects of an oil spill accident, such as the HSOS on the coastal food web and the trophic recovery of the oil-impacted region.
Here, we analyzed carbon and nitrogen stable isotope signatures of macrobenthic consumers and their potential food sources on heavily oil-impacted rocky intertidal shores approximately 4 years after the HSOS, and compared these signatures with those in the non-impacted regions (control sites). We hypothesize that if fauna and flora on the rocky shores are still impacted by residual oils, the isotopic signatures of macrobenthic consumers at the oil-impacted sites will be dissimilar to those at the control sites. Our major objective was to assess the effects of an oil-spill accident on trophic pathways of organic matter and the trophic structure (including isotopic niches of consumers) of macrobenthic communities on rocky shore intertidal habitats. A previous study on the macrobenthic food web impacted by the HSOS only focused on the soft-bottom ecosystem, which showed a similar food web structure between oil-impacted and control sites [17]. To our knowledge, this is the first study to examine the trophic recovery of a rocky shore macrobenthic food web based on the isotopic assessment of the rocky shore ecosystem impacted by the HSOS.

2. Materials and Methods

2.1. Site Description

The Taean Peninsula is located on the western coast of the Korean peninsula and has a semidiurnal and macrotidal regime (spring tidal range of 4–6 m). The sampling sites consist of two heavily oil-impacted sites (Gurepo, 36°53′31″ N 126°11′55″ E and Padori, 36°44′14″ N 126°07′49″ E: referred to as sites G and P, respectively) and one non-impacted reference site (Yeonpo, 36°41′33″ N 126°13′18″ E: referred to as site Y), which are located on the coastal line of the peninsula (Figure 1). The concentrations of total petroleum hydrocarbon in seawater off the Taean coast were 1.5–7300 μg L−1 (mean 732 μg L−1) immediately after the HSOS, which decreased rapidly to 2.0–224 μg L−1 (mean 31 μg L−1) in 1 month in response to clean-up activities and strong tidal currents [26]. Overall, the water quality off the Taean coast—in terms of the concentration of total petroleum hydrocarbons—recovered to below the environmental standard (10 μg L−1) within 1 y after the HSOS [12]. Ecological indices (species richness and diversity) of macrobenthic invertebrates at the oil-impacted sites increased gradually by 1 y after the HSOS, and their mean densities were similar to those at the reference site [11,27].

2.2. Sample Collection and Processing

Two sampling campaigns were carried out at the three sampling sites in July and November of 2013. Macrobenthic invertebrates, organic matter sources, and macroalgae were sampled at each site. The samples of macroinvertebrate consumers were collected using a 0.0625 m2 (25 × 25 cm) quadrat by hand with a stainless-steel knife during low spring tides. Macroinvertebrates were sorted and kept alive overnight in filtered seawater to enable them to evacuate their gut contents at the laboratory. Individuals for macrofauna were identified at the lowest taxonomic level possible under a dissecting microscope; then, their muscle tissues were dissected, lyophilized, and homogenized to a fine powder with a ball mill (Retsch MM200 Mixer Mill, Hyland Scientific, WA, USA) for stable isotope analysis. The prepared samples were kept in a deep freezer (−50 °C) before this.
For collecting suspended particulate organic matter (SPOM), water samples were obtained at each site during flood tides using a van Dorn sampler on board and then filtered through pre-combusted (450 °C for 4 h) Whatman GF/F glass fiber filters (0.70 μm nominal pore size) after removing any zooplankton and large particles using a 200-μm mesh net. Epilithic algae (EA) were obtained from several small-sized rocks by scraping the rock surface using a fine brush. The prepared particles were prefiltered through a 200-μm sieve to remove large materials, and then EA were sampled through the pre-combusted GF/F filters. The filter samples for SPOM and EA were fumed under a 1 N HCl stream overnight to remove inorganic carbonates and then oven-dried for 48 h at 50 °C and wrapped in an aluminum foil. Two dominant macroalgae (Ulva pertusa and Sargassum thunbergii) were sampled by hand during low tides and then were scraped with a razor blade to remove epibionts and rinsed with Milli-Q water. The macroalgal samples were freeze-dried and ground to a fine powder in a ball mill as mentioned above. The treated samples of organic matter sources and macroalgae were kept in a deep freezer (−50 °C) until subsequent analysis.

2.3. Stable Isotope Analysis

For stable isotope analysis, powder and filter samples were packed into tin combustion cups and foils, respectively. The capsulated samples were introduced into an elemental analyzer (Vario MICRO Cube CHNOS elemental analyzer; Elementar Analysensysteme GMbH, Hanau, Germany) to oxidize at a high temperature (1030 °C), and then the resultant CO2 and N2 gases were analyzed for their stable isotope ratios using a continuous flow isotope ratio-mass spectrometry (IsoPrime100; Cheadle Hulme, UK). Stable isotope data are expressed in conventional delta notation (δ, ‰) based on the relative difference between isotopic ratios of the sample and reference gases of Vienna Pee Dee Belemnite for carbon and atmospheric N2 for nitrogen. Data were obtained using the following equation:
δX (‰) = [(Rsample/Rstandard) − 1] × 1000,
where X is 13C or 15N and R values are the 13C/12C or 15N/14N ratios. Sucrose (ANU C12H22O11; NIST, Gaithersburg, MD, USA) and ammonium sulfate ([NH4]2SO4; NIST) with established relationships to international standards were used as reference materials. Standard deviations of repeated measurements of internal urea standard were 0.13‰ and 0.20‰ for δ13C and δ15N, respectively.

2.4. Data Analysis

Permutational multivariate analysis of variance (PERMANOVA) using the software PRIMER version 6 combined with the PERMANOVA + PRIMER add-on [28] was conducted to assess spatial and seasonal differences in the δ13C and δ15N values of organic matter sources (SPOM, EA, and macroalgae) and consumers between the oil-impacted and reference sites between July and November. For consumer species, PERMANOVA was performed on the isotopic data of feeding groups (suspension feeders, deposit feeders, grazers, omnivores, and carnivores) rather than at species levels to assess the community trophic structure impacted by the oil spill. The significance level was set to p < 0.05 for all statistical tests.
To test the similarity in the resource use by consumer feeding groups between the oil-impacted and reference sites, the overall variability in their isotopic niche areas (‰2) was investigated using the package Stable Isotope Bayesian Ellipses in R (software R version 3.5.3) [29,30]. To compare each feeding group in the isotopic niche areas between the sites, we used the indices of total area (TA) and small sample size-corrected standard ellipse area (SEAc) with a confidence interval of 95% as quantitative indicators of the trophic diversity of consumers [29,31]. In addition, the percentage of isotopic niche overlap of each feeding group between the sites was calculated by the ratio between the overlap areas based on the SEAc representing the core isotopic niche with the ‘nicheRover’ package in R [32].

3. Results

3.1. Stable Isotope Ratios of Potential Food Sources

There were no significant differences in the δ13C and δ15N values of all potential food sources (SPOM, EA, U. pertusa, and S. thunbergii) among sites (PERMANOVA, p > 0.746 for all cases) (Table 1), whereas their significant seasonal differences were observed between July and November (PERMANOVA, p < 0.011 for all cases). No significant effects of the interaction term (season × site) were found for all potential food sources (p > 0.688 for all cases). The δ13C values of potential food sources ranged from −23.6 ± 0.4‰ (S. thunbergii at site G) to −12.8 ± 0.5‰ (U. pertusa at site P) in July and from −23.2 ± 0.3‰ (S. thunbergii at site G) to −13.5 ± 0.6‰ (U. pertusa at site Y) in November (Table 1). By contrast, their δ15N values varied from 3.5 ± 0.4‰ (S. thunbergii at site G) to 7.8 ± 1.2‰ (SPOM at site P) in July and from 4.0 ± 0.3‰ (S. thunbergii at site P) to 6.1 ± 0.5‰ (U. pertusa at site P) in November.

3.2. Stable Isotope Ratios of Consumer Species

The mean δ13C and δ15N values of macrobenthic consumers showed similar ranges among sites: −19.2 ± 0.2‰ to −12.7 ± 0.5‰, 4.7 ± 0.6‰ to 12.2 ± 0.7‰, and −19.1 ± 0.4‰ to −12.7 ± 0.3‰, and 5.1 ± 0.2‰ to 12.3 ± 0.4‰, −19.2 ± 0.3‰ to −12.7 ± 0.5‰, and 4.9 ± 0.3‰ to 12.5 ± 0.4‰ at sites G, P, and Y, respectively (Figure 2 and Table 2). Three-way PERMANOVA showed that there were no significant differences in the δ13C and δ15N values among sites (pseudo-F 2, 504 = 1.05, p = 0.390). By contrast, significant differences in the δ13C and δ15N values of macrobenthic consumers were observed between seasons (pseudo-F1, 504 = 47.61, p = 0.001) and among feeding groups (pseudo-F4, 504 = 61.32, p = 0.001). All interaction terms showed significant effects (p < 0.009 for all cases).
PERMANOVA for comparing each feeding group revealed that isotopic values for all feeding groups differed significantly between the two seasons (p = 0.001 for all cases) (Table 3), whereas no significant differences among sites were observed (suspension feeders, pseudo-F2, 117 = 0.55, p = 0.663; deposit feeders, pseudo-F2, 73 = 1.07, p = 0.355; grazers, pseudo-F2, 160 = 1.47, p = 0.212; omnivores, pseudo-F2, 71 = 0.58, p = 0.626; carnivores, pseudo-F2, 79 = 0.79, p = 0.533). No significant effects of the interaction term (season × site) were observed for all feeding groups (p > 0.391 for all cases). All feeding groups in July had higher mean δ13C and δ15N values than those in November. In addition, the mean δ13C and δ15N values of feeding groups showed similar trends among sites ranging from −19.0 ± 0.4‰ (suspension feeder at site Y in November) to −14.3 ± 1.3‰ (grazer in July), and from 5.4 ± 0.6‰ (suspension feeder at site G in November) to 12.2 ± 0.5‰ (carnivore in July), respectively.

3.3. Isotopic Niche Areas

Regardless of the season, the isotopic niche areas (‰2) for each feeding group exhibited similar TA and SEAc values among the three sampling sites (Table 4). By contrast, the omnivore group had different TA (3.27 to 6.38 in July and 2.34 to 5.22 in September) and SEAc values (1.89 to 3.45 in July and 1.24 to 2.53 in September) among the sites. In both July and November, suspension feeders had the lowest TA and SEAc values (2.01 and 0.80 and 1.31 and 0.66 at site Y, respectively), whereas grazers had the highest values (9.45 and 3.45 at site Y for TA and 7.29 and 3.17 at site G for SEAc, respectively). In July, all the feeding groups had substantial overlapping proportions of isotopic niche areas between the polluted and control sites, ranging from 58% for carnivores to 84% for suspension feeders between sites P and Y (Figure 3A). The isotopic overlapping proportions between the impacted and control sites in November showed relatively low values than those in July, ranging from 40% for suspension feeders to 92% for grazers between sites P and Y (Figure 3B).

4. Discussion

Our isotopic data showed that there were no significant differences in the carbon and nitrogen isotope signatures of consumer feeding groups and their isotopic niches between the test and control sites. Oil contamination has been considered to influence multiple trophic levels, and their relationships arise from top–down or bottom–up effects [33]. Therefore, similar isotopic signatures and niches of macrobenthic consumers between polluted and control sites provide evidence for the trophic recovery of rocky shore food webs related to the HSOS [7,17]. A comprehensive assessment of recovery time following the HSOS impacts revealed that damaged intertidal and subtidal benthic communities could recover their ecological status completely to the normal level after approximately 6 years [11]. Although the benthic communities on rocky shores were not consistent between the oiled and control sites at the sampling times used, the high similarity in isotope characteristics of the same macrobenthic feeding groups between the sites suggests almost complete recovery of the functions of the food web in terms of carbon and energy flows.

4.1. Comparison of Isotopic Signatures of Organic Matter Sources and Microbenthic Consumers between Oil-Impacted and Control Sites

Coastal ecosystems impacted by direct or indirect anthropogenic activities can exhibit changes in dietary sources for macrobenthic consumers, thereby causing modification in the trophic structure [1,34]. However, in the present study, regardless of season, the δ13C and δ15N values of consumer species and their potential food sources at the oil-impacted sites were very similar to those at the control site. These isotopic similarities suggest that impacts on the food web through the feeding relationships between basal resources and their consumers after HSOS did not differ between the two sites [17,35]. In addition, the isotopic values of organic matter sources and animals at the two sites were within the ranges previously observed in the natural coastal waters off the Korean Peninsula [36,37,38]. When organic matter sources are still contaminated by crude oil at polluted sites, their δ13C values may be decreased by 1‰ to 4‰ relative to those at the control site [39]. Such altered δ13C values of organic matter sources by petrogenic hydrocarbons might be reflected in their incorporation into the macrobenthic food webs through the feeding of consumer species [17,40]. Here, the spatial isotopic consistency between the two sites suggests that there was no contribution from petrogenic hydrocarbon to the macrobenthic food webs. The concentrations of residual oils in seawater, sediments, and bivalves were estimated quantitatively through long-term assessment, and the results showed that they dropped to background levels [11]. The residual oils in seawater and sediments are decomposed mostly by microbial degradation; therefore, petrogenic hydrocarbon may not be incorporated into the food webs based on the carbon uptake by primary producers [41]. Accordingly, trophic consistencies between the two sites suggest that the rocky shore food web impacted by the HSOS could have been rehabilitated by natural conditions to that of the levels of unaffected environments in response to the modification of energy flow through food web structures.
Seasonal variations in community (both autotrophs and heterotrophs) and food web structures are observed commonly in coastal ecosystems in response to seasonal changes in environmental factors [42,43]. Normal coastal ecosystems constitute very complex environments and a variety of primary producers, resulting in seasonal variations in the trophic structure caused by the changes in the importance of dietary sources of consumer species. Similarly, our study also demonstrated seasonal variations in the isotopic ratios of macrobenthic consumers and their potential food sources regardless of the sites impacted by the HSOS. The seasonal differences in the isotopic ratios of consumer species might be linked to food availability, which can vary with seasons because of the change in the relative contribution of pelagic and benthic sources to the macrobenthic food webs. In addition, isotopic ratios of consumers were well classified by feeding strategies at both sites in both seasons, which was aligned with the general trend in natural ecosystems. The seasonal isotopic shifts of organic matter sources and feeding groups are consistent with the characteristics of the trophic base and feeding strategy of macrobenthic consumers reported for Korean coastal ecosystems [35,36,37,44]. As a result, seasonal variations in the isotopic ratios of all items might be exhibiting the same trend between the two sites, which suggests consistent normal trophic transfers from basal resources to higher trophic levels within the food webs in the oil-impacted sites.

4.2. Isotopic Niche Comparison of Microbenthic Consumers between Oil-Impacted and Control Sites

Trophic functional modifications to macrobenthic communities following an environmental disturbance are likely to occur along with changes in the densities of dominant species [1]. Oil contamination can lead to a trophic niche shift with consequent changes in trophic levels and food chain lengths of consumers [17,45]. The trophic niche of consumers can be estimated by their isotope values, which provide information on the changes in the trophic structure caused by a variety of natural and anthropogenic disturbances [25,46,47]. Our results showed similar isotopic niche parameters in all the feeding groups between the oil-impacted and natural sites, suggesting that consumer species composed of the same feeding groups utilize similar nutritional resources under common trophic pathways regardless of the presence of oil spillage [48]. In addition, here we found high degrees of isotopic niche overlaps (58% to 84% in July and 40% to 92% in November) of each feeding group between the two study sites. The proportion of isotopic niche overlap represents a quantitative measure of trophic redundancy and segregation between consumer species or groups of consumers [49]. These results suggest that the same feeding groups might exploit similar basal resources and show high trophic redundancy between the oil-impacted and control sites [50]. In this respect, the overlap proportions suggest that the macrobenthic food web on the oil-impacted rocky intertidal habitats might be functionally similar to those in the natural habitats approximately 4 years after the HSOS. Accordingly, a comparison of the isotopic niche assessment of feeding groups of the oil-affected and natural sites suggests that the HSOS did not affect the trophic structure of the rocky shore community in the long term because of the trophic recovery of the ecological components.

4.3. Ecological Implication of Trophic Recovery after the HSOS

The density and biomass of macrobenthic invertebrates decrease suddenly after oil spills, which tend to gradually recover with natural attenuation of residual oils with varying recovery periods in different cases [12]. Macrobenthic communities on the oil-impacted rocky shores in our study areas were fully recovered compared with the reference habitats by 67 months after the HSOS [11]. By contrast, Jung et al. [27] reported that, despite the evidence of macrobenthic recovery on the impacted rocky shores, there were noticeable differences in the abundance of dominant species between the oil-impacted and natural sites during the first 5 years after the HSOS because of the proliferation of the Pacific oyster Crassostrea gigas and the high abundance of Lottia spp. and Odostomia aomori at the oil-impacted site. These studies mostly focused on the environmental and structural recovery of the oil-impacted ecosystem by the approach based on residual oils and community analyses. However, assessments of the recovery of community- and ecosystem-level responses following oil spills should consider the structure of the food web [5,7]. Our results show functional recovery of the macrobenthic food web structure on the rocky shores based on similar trophic niches and higher overlap proportions of feeding groups between the oil-affected and natural sites. Following the impact of diverse pollution incidents, the recovery process in marine ecosystems can occur through the recolonization of organisms and increasing robustness of the trophic structure [17]. Overall, despite the differences in the compositions of the macrobenthic community in the oil-affected and natural sites, our results reveal the trophic recovery of rocky shore habitats from the impacts of the HSOS, as evidenced by the high equivalency of the trophic structure between the sites.

5. Conclusions

Our study provides reasonable evidence for the trophic recovery of an ecosystem approximately 4 years after the HSOS as observed by similar trophic characteristics of the oil-impacted rocky habitat to that of the natural habitat, with similar resource use by consumers and their trophic relationships. Furthermore, the similarity in isotopic niche indices of consumer feeding groups between the oil-affected and natural rocky shores suggests that the oil-impacted habitat might achieve the ecological function of energy transfers from basal resources to higher trophic levels in the natural ecosystem. Despite the controversy in the recovery of the macrobenthic community, these results indicate the natural conditions of the rocky shore ecosystem and its services due to the recovery of the food web structure. The recovery assessment based on the comparison of feeding groups between the oil-impacted and natural habitats using stable isotope data can provide valuable information on trophic recovery and ecological function of the ecosystem. Finally, this study enhances our understanding of the ecological processes of trophic recovery in coastal ecosystems that are impacted by oil spills and provides insights into the responses of such ecosystems to large-scale oil contamination using proper and effective assessment procedures.

Author Contributions

Conceptualization, H.J.P. and C.-K.K.; methodology, H.J.P., C.K. and C.-K.K.; formal analysis, H.J.P. and C.K.; investigation, H.J.P.; data curation, H.J.P. and C.K.; writing—original draft preparation, review and editing, H.J.P. and C.-K.K.; visualization, H.J.P. and C.K.; supervision, C.-K.K.; project administration, C.-K.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by “Predictable Effects of Climate Change on Marine Ecosystem in The Tsushima Warm Current System of Korean Waters” program of Korea institute of Marine Science & Technology Promotion (KIMST) funded by the Ministry of Oceans and Fisheries (KIMST-20220558).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Special thanks should be given to Dong Hoon Shin for his effort for field sampling.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kennish, M.J. Environmental threats and environmental future of estuaries. Environ. Conserv. 2002, 29, 78–107. [Google Scholar] [CrossRef]
  2. Mendelssohn, I.A.; Andersen, G.L.; Baltz, D.M.; Caffey, R.H.; Carman, K.R.; Flegger, J.W.; Joye, S.B.; Lin, Q.; Maltby, E.; Overton, E.B.; et al. Oil impacts on coastal wetlands: Implications for the Mississippi River delta ecosystem after the Deepwater Horizon oil spill. BioScience 2012, 62, 562–574. [Google Scholar] [CrossRef]
  3. Depellegrin, D.; Blažauskas, N. Integrating ecosystem service values into oil spill impact assessment. J. Coast. Res. 2013, 29, 836–846. [Google Scholar] [CrossRef]
  4. Suchanek, T.H. Oil impacts on marine invertebrate populations and communities. Am. Zool. 1993, 33, 510–523. [Google Scholar] [CrossRef] [Green Version]
  5. Peterson, C.H.; Rice, S.D.; Short, J.W.; Esler, D.; Bodkin, J.L.; Ballachey, B.E.; Irons, D.B. Long-term ecosystem response to the Exxon Valdez oil spill. Science 2003, 302, 2082–2086. [Google Scholar] [CrossRef] [Green Version]
  6. Moreno, R.; Jover, L.; Diez, C.; Sardà, F.; Sanpera, C. Ten years after the Prestige oil spill: Seabird trophic ecology as indicator of long-term effects on the coastal marine ecosystem. PLoS ONE 2013, 8, e77360. [Google Scholar] [CrossRef]
  7. Carman, K.R.; Fleeger, J.W.; Pomarico, S.M. Response of a benthic food web to hydrocarbon contamination. Limnol. Oceanogr. 1997, 42, 561–571. [Google Scholar] [CrossRef]
  8. McCann, M.J.; Able, K.W.; Christian, R.R.; Fodrie, F.J.; Jensen, O.P.; Johnson, J.J.; López-Duarte, P.C.; Martin, C.W.; Olin, J.A.; Polito, M.J.; et al. Key taxa in food web responses to stressors: The Deepwater Horizon oil spill. Front. Ecol. Environ. 2017, 15, 142–149. [Google Scholar] [CrossRef]
  9. Halpern, B.S.; Selkoe, K.A.; Micheli, F.; Kappel, C.V. Evaluating and ranking the vulnerability of global marine ecosystems to anthropogenic threats. Conserv. Biol. 2007, 21, 1301–1315. [Google Scholar] [CrossRef]
  10. Hawkins, S.J.; Evans, A.J.; Mieszkowska, N.; Adams, L.C.; Bray, S.; Burrow, M.T.; Firth, L.B.; Genner, M.J.; Leung, K.M.Y.; Moore, P.J.; et al. Distinguishing globally-driven changes from regional- and local-scale impacts: The case for long-term and broad-scale studies of recovery from pollution. Mar. Pollut. Bull. 2017, 124, 573–586. [Google Scholar] [CrossRef] [Green Version]
  11. Yim, U.H.; Hong, S.; Lee, C.; Kim, M.; Jung, J.H.; Ha, S.Y.; An, J.G.; Kwon, B.O.; Kim, T.; Lee, C.H.; et al. Rapid recovery of coastal environment and ecosystem to the Hebei Spirit oil spill’s impact. Environ. Int. 2020, 136, 105438. [Google Scholar] [CrossRef] [PubMed]
  12. Hong, S.; Khim, J.S.; Ryu, J.; Kang, S.G.; Shim, W.J.; Yim, U.H. Environmental and ecological effects and recoveries after five years of the Hebei Spirit oil spill, Taean, Korea. Ocean Coast. Manag. 2014, 102, 522–532. [Google Scholar] [CrossRef]
  13. Kim, M.; Jung, J.H.; Ha, S.Y.; An, J.G.; Shim, W.J.; Yim, U.H. Long-term monitoring of PAH contamination in sediment and recovery after the Hebei Spirit oil spill. Arch. Environ. Con. Tox. 2017, 73, 93–102. [Google Scholar] [CrossRef] [PubMed]
  14. Lee, C.-H.; Lee, J.H.; Sung, C.G.; Moon, S.D.; Kang, S.K.; Lee, J.H.; Yim, U.H.; Shim, W.J.; Ha, S.Y. Monitoring toxicity of polycyclic aromatic hydrocarbons in intertidal sediments for five years after the Hebei Spirit oil spill in Taean, Republic of Korea. Mar. Pollut. Bull. 2013, 76, 241–249. [Google Scholar] [CrossRef] [PubMed]
  15. Yu, O.H.; Lee, H.G.; Shim, W.J.; Kim, M.; Park, H.S. Initial impacts of the Hebei Spirit oil spill on the sandy beach macrobenthic community west coast of Korea. Mar. Pollut. Bull. 2013, 70, 189–196. [Google Scholar] [CrossRef]
  16. Jung, Y.H.; Yoon, K.T.; Shim, W.J.; Park, H.S. Short-term variation of the macrobenthic fauna structure on rocky shores after the Hebei Spirit oil spill, west coast of Korea. J. Coast. Res. 2014, 31, 177–183. [Google Scholar] [CrossRef]
  17. Han, E.; Park, H.J.; Bergamino, L.; Choi, K.S.; Choy, E.J.; Yu, O.H.; Lee, T.W.; Park, H.S.; Shim, W.J.; Kang, C.K. Stable isotope analysis of a newly established macrofaunal food web 1.5 years after the Hebei Spirit oil spill. Mar. Pollut. Bull. 2015, 90, 167–180. [Google Scholar] [CrossRef]
  18. Layman, C.A.; Araujo, M.S.; Boucek, R.; Hammerschlag-Peyer, C.M.; Harrison, E.; Jud, Z.R.; Matich, P.; Rosenblatt, A.E.; Vaudo, J.J.; Yeager, L.A.; et al. Applying stable isotopes to examine food-web structure: An overview of analytical tools. Biol. Rev. 2012, 87, 545–562. [Google Scholar] [CrossRef]
  19. Mancinelli, G.; Vizzini, S. Assessing anthropogenic pressures on coastal marine ecosystems using stable CNS isotopes: State of the art, knowledge gaps, and community scale perspectives. Estuar. Coast. Shelf Sci. 2015, 156, 195–204. [Google Scholar] [CrossRef]
  20. Fry, B.; Sherr, E. δ13C measurements as indicators of carbon flow in marine and freshwater ecosystems. Contrib. Mar. Sci. 1984, 27, 49–63. [Google Scholar]
  21. Post, D.M. Using stable isotopes to estimate trophic position: Models, methods, and assumptions. Ecology 2002, 83, 703–718. [Google Scholar] [CrossRef]
  22. McCutchan, J.H., Jr.; Lewis, W.M., Jr.; Kendall, C.; McGrath, C.C. Variation in trophic shift for stable isotope ratios of carbon, nitrogen, and sulfur. Oikos 2003, 102, 378–390. [Google Scholar] [CrossRef]
  23. Vizzini, S.; Mazzola, A. The effects of anthropogenic organic matter inputs on stable carbon and nitrogen isotopes in organisms from different trophic levels in a southern Mediterranean coastal area. Sci. Total Environ. 2006, 368, 723–731. [Google Scholar] [CrossRef] [PubMed]
  24. Quillien, N.; Nordström, M.C.; Schaal, G.; Bonsdorff, E.; Grall, J. Opportunistic basal resource simplifies food web structure and functioning of a highly dynamic marine environment. J. Exp. Mar. Biol. Ecol. 2016, 477, 92–102. [Google Scholar] [CrossRef]
  25. Park, H.J.; Kwak, J.H.; Lee, Y.J.; Kang, H.Y.; Choy, E.J.; Kang, C.K. Trophic structures of two contrasting estuarine ecosystems with and without a dike on the temperate coast of Korea as determined by stable isotopes. Estuar. Coasts 2020, 43, 560–577. [Google Scholar] [CrossRef]
  26. Kim, M.; Hong, S.H.; Won, J.; Yim, U.H.; Jung, J.H.; Ha, S.Y.; An, J.G.; Joo, C.; Kim, E.; Han, G.M.; et al. Petroleum hydrocarbon contaminations in the intertidal seawater after the Hebei Spirit oil spill–Effect of tidal cycle on the TPH concentrations and the chromatographic characterization of seawater extracts. Water Res. 2013, 47, 758–768. [Google Scholar] [CrossRef]
  27. Jung, Y.H.; Park, H.S.; Yoon, K.T.; Kim, H.J.; Shim, W.J. Long-term changes in rocky intertidal macrobenthos during the five years after the Hebei Spirit oil spill, Taean, Korea. Ocean Sci. J. 2017, 52, 103–112. [Google Scholar] [CrossRef]
  28. Anderson, M.J.; Gorley, R.N.; Clarke, K.R. PERMANOVA + for PRIMER: Guide to Software and Statistical Methods; PRIMER-E: Plymouth, UK, 2008. [Google Scholar]
  29. Jackson, A.L.; Inger, R.; Parnell, A.C.; Bearhop, S. Comparing isotopic niche widths among and within communities: SIBER—Stable Isotope Bayesian Ellipses in R. J. Anim. Ecol. 2011, 80, 595–602. [Google Scholar] [CrossRef]
  30. Team, R.C. R: A Language and Environment for Statistical Computing; R Foundation for Statistical Computing: Vienna, Austria, 2019. [Google Scholar]
  31. Layman, C.A.; Arrington, D.A.; Montaña, C.G.; Post, D.M. Can stable isotope ratios provide for community-wide measures of trophic structure? Ecology 2007, 88, 42–48. [Google Scholar] [CrossRef]
  32. Swanson, H.K.; Lysy, M.; Power, M.; Stasko, A.D.; Johnson, J.D.; Reist, J.D. A new probabilistic method for quantifying n-dimensional ecological niches and niche overlap. Ecology 2015, 96, 318–324. [Google Scholar] [CrossRef]
  33. Peterson, C.T.; Grubbs, R.D.; Mickle, A. An investigation of effects of the Deepwater Horizon oil spill on coastal fishes in the Florida big bend using fisher-independent surveys and stable isotope analysis. Southeast. Nat. 2017, 16, G93–G108. [Google Scholar] [CrossRef]
  34. Gedan, K.B.; Silliman, B.R.; Bertness, M.D. Centuries of human-driven change in salt marsh ecosystems. Annu. Rev. Mar. Sci. 2009, 1, 117–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kang, C.K.; Choy, E.J.; Son, Y.; Lee, J.Y.; Kim, J.K.; Kim, Y.; Lee, K.S. Food web structure of a restored macroalgal bed in the eastern Korean peninsula determined by C and N stable isotope analyses. Mar. Biol. 2008, 153, 1181–1198. [Google Scholar] [CrossRef]
  36. Kang, C.K.; Kim, J.B.; Lee, K.S.; Kim, J.B.; Lee, P.Y.; Hong, J.S. Trophic importance of benthic microalgae to macrozoobenthos in coastal bay systems in Korea: Dual stable C and N isotope analyses. Mar. Ecol. Prog. Ser. 2003, 259, 79–92. [Google Scholar] [CrossRef]
  37. Park, H.J.; Han, E.; Lee, Y.J.; Kang, C.K. Trophic linkage of a temperate intertidal macrobenthic food web under opportunistic macroalgal blooms: A stable isotope approach. Mar. Pollut. Bull. 2016, 111, 86–94. [Google Scholar] [CrossRef]
  38. Kim, C.; Kang, H.Y.; Lee, Y.J.; Yun, S.G.; Kang, C.K. Isotopic variation of macroinvertebrates and their sources of organic matter along an estuarine gradient. Estuar. Coasts 2020, 43, 496–511. [Google Scholar] [CrossRef]
  39. Graham, W.M.; Condon, R.H.; Carmichael, R.H.; D’Ambra, I.; Patterson, H.K.; Linn, L.J.; Hernandez, F.J. Oil carbon entered the coastal planktonic food web during the Deepwater Horizon oil spill. Environ. Res. Lett. 2010, 5, 045301. [Google Scholar] [CrossRef]
  40. Spies, R.B.; DesMarais, D.J. Natural isotope study of trophic enrichment of marine benthic communities by petroleum seepage. Mar. Biol. 1983, 73, 67–71. [Google Scholar] [CrossRef]
  41. Leahy, J.G.; Colwell, R.R. Microbial degradation of hydrocarbons in the environment. Microbiol. Rev. 1990, 54, 305–315. [Google Scholar] [CrossRef]
  42. Vizzini, S.; Mazzola, A. Seasonal variations in the stable carbon and nitrogen isotope ratios (13C/12C and 15N/14N) of primary producers and consumers in a western Mediterranean coastal lagoon. Mar. Biol. 2003, 142, 1009–1018. [Google Scholar] [CrossRef]
  43. Kang, H.Y.; Kim, C.; Kim, D.; Lee, Y.J.; Park, H.J.; Kundu, G.K.; Kim, Y.K.; Bibi, R.; Jang, J.; Lee, K.H.; et al. Identifying patterns in the multitrophic community and food-web structure of a low-turbidity temperate estuarine bay. Sci. Rep. 2020, 10, 16637. [Google Scholar] [CrossRef] [PubMed]
  44. Kang, C.K.; Park, H.J.; Choy, E.J.; Choi, K.S.; Hwang, K.; Kim, J.B. Linking inter-tidal and subtidal food webs: Consumer-mediated transport of intertidal benthic microalgal carbon. PLoS ONE 2015, 10, e0139802. [Google Scholar] [CrossRef]
  45. Felder, D.L.; Thoma, B.P.; Schmidt, W.E.; Sauvage, T.; Self-Krayesky, S.L.; Chistoserdov, A.; Bracken-Grissom, H.D.; Fredericq, S. Seaweeds and decapod crustaceans on Gulf Deep Banks after the Macondo oil spill. BioScience 2014, 64, 808–819. [Google Scholar] [CrossRef] [Green Version]
  46. Abrantes, K.G.; Barnett, A.; Bouillon, S. Stable isotope-based community metrics as a tool to identify patterns in food web structure in east African estuaries. Funct. Ecol. 2014, 28, 270–282. [Google Scholar] [CrossRef]
  47. Romano, C.; Fanelli, E.; D’Anna, G.; Pipitone, C.; Vizzini, S.; Mazzola, A.; Badalamenti, F. Spatial variability of soft-bottom macrobenthic communities in northern Sicily (Western Mediterranean): Contrasting trawled vs. untrawled areas. Mar. Environ. Res. 2016, 122, 113–125. [Google Scholar] [CrossRef] [PubMed]
  48. Newsome, S.D.; Martinez del Rio, C.; Bearhop, S.; Phillips, D.L. A niche for isotopic ecology. Front. Ecol. Environ. 2007, 5, 429–436. [Google Scholar] [CrossRef]
  49. Flaherty, E.A.; Ben-David, M. Overlap and partitioning of the ecological and isotopic niches. Oikos 2010, 119, 1409–1416. [Google Scholar] [CrossRef]
  50. Wlodarska-Kowalczuk, M.; Pearson, T.H. Benthic response to chronic natural physical disturbance by glacial sedimentation in an Arctic fjord. Mar. Ecol. Prog. Ser. 2005, 303, 31–41. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Map of the Taean peninsula on the western coast of Korea. Sampling areas (this study, red color) are located at two heavily oil-impacted sites (Gurepo and Padori) and one non-impacted reference site (Yeonpo). Blue color indicates the sampling sites (Malipo and Mongsanpo) investigated by [17].
Figure 1. Map of the Taean peninsula on the western coast of Korea. Sampling areas (this study, red color) are located at two heavily oil-impacted sites (Gurepo and Padori) and one non-impacted reference site (Yeonpo). Blue color indicates the sampling sites (Malipo and Mongsanpo) investigated by [17].
Water 14 02335 g001
Figure 2. Dual isotope plots of δ13C and δ15N values of consumers (empty circles, suspension feeders; filled circles, deposit feeders; empty triangles, grazers; filled triangles, omnivores; stars, carnivores) at the three oil-contaminated sites ((A), Gurepo; (B), Padori; (D), Malipo) and two non-impacted reference sites ((C), Yeonpo; (E), Mongsanpo) in July (blue color) and November (red color) for this study and in spring and summer for [17]. Values of the potential food sources are illustrated by filled squares. SPOM, suspended particulate organic matter; EA, epilithic algae; UP, microphytobenthos; MPB Ulva pertusa; ST, Sargassum thunbergii. Values are means (‰) ± 1 standard deviation.
Figure 2. Dual isotope plots of δ13C and δ15N values of consumers (empty circles, suspension feeders; filled circles, deposit feeders; empty triangles, grazers; filled triangles, omnivores; stars, carnivores) at the three oil-contaminated sites ((A), Gurepo; (B), Padori; (D), Malipo) and two non-impacted reference sites ((C), Yeonpo; (E), Mongsanpo) in July (blue color) and November (red color) for this study and in spring and summer for [17]. Values of the potential food sources are illustrated by filled squares. SPOM, suspended particulate organic matter; EA, epilithic algae; UP, microphytobenthos; MPB Ulva pertusa; ST, Sargassum thunbergii. Values are means (‰) ± 1 standard deviation.
Water 14 02335 g002
Figure 3. Isotopic niche comparisons of each feeding group (suspension feeder, deposit feeder, grazer, omnivore, and carnivore) estimated as the total area (TA, dotted lines) and standard ellipse area (SEAc, solid lines) among the two oil-impacted sites (Gurepo, red lines and Padori, green lines) and one non-impacted reference site (Yeonpo, black lines) in July (A) and November (B) using the Stable Isotope Bayesian Ellipse in R (SIBER) procedure.
Figure 3. Isotopic niche comparisons of each feeding group (suspension feeder, deposit feeder, grazer, omnivore, and carnivore) estimated as the total area (TA, dotted lines) and standard ellipse area (SEAc, solid lines) among the two oil-impacted sites (Gurepo, red lines and Padori, green lines) and one non-impacted reference site (Yeonpo, black lines) in July (A) and November (B) using the Stable Isotope Bayesian Ellipse in R (SIBER) procedure.
Water 14 02335 g003
Table 1. Mean δ13C and δ15N values of potential food sources (SPOM, suspended particulate organic matter; Epilithic algae; U. pertusa, Ulva pertusa; S. thunbergii, Sargassum thunbergii) at the two heavily oil-impacted (Gurepo and Padori) and one non-impacted reference (Yeonpo) sites in July and November. PERMANOVA test of δ13C and δ15N values for each potential food source among the three sites and between the two seasons (significance at p < 0.05). Data represent means ± 1 SD.
Table 1. Mean δ13C and δ15N values of potential food sources (SPOM, suspended particulate organic matter; Epilithic algae; U. pertusa, Ulva pertusa; S. thunbergii, Sargassum thunbergii) at the two heavily oil-impacted (Gurepo and Padori) and one non-impacted reference (Yeonpo) sites in July and November. PERMANOVA test of δ13C and δ15N values for each potential food source among the three sites and between the two seasons (significance at p < 0.05). Data represent means ± 1 SD.
GurepoPadoriYeonpo
Potential Food Source δ13C δ15N δ13C δ15N δ13C δ15N
nMeanSDMeanSDnMeanSDMeanSDnMeanSDMeanSD
July
SPOM8−20.40.67.31.08−20.21.17.81.26−20.30.97.50.8
Epilithic algae6−15.41.07.70.85−15.20.87.51.06−15.70.77.40.9
U. pertusa6−12.90.56.70.46−12.80.56.80.46−13.00.66.90.4
S. thunbergii6−23.60.43.50.46−23.40.53.70.36−23.40.63.80.5
November
SPOM7−20.80.95.40.96−20.60.65.70.98−20.91.05.80.6
Epilithic algae6−14.50.65.90.66−14.90.75.60.66−14.40.65.60.6
U. pertusa6−13.80.55.80.46−13.60.56.10.56−13.60.65.90.4
S. thunbergii6−23.20.34.20.36−23.00.34.00.36−22.90.44.10.3
PERMANOVASeasonSiteInteraction
pseudo-Fp-valuepseudo-Fp-valuepseudo-Fp-value
SPOM30.520.0010.470.7460.010.966
Epilithic algae40.870.0010.110.9110.500.695
U. pertusa32.920.0010.290.7890.210.877
S. thunbergii6.450.0110.160.9200.510.688
Table 2. Mean δ13C and δ15N values of macrobenthic consumers collected at the two heavily oil-impacted (Gurepo and Padori) and one non-impacted reference (Yeonpo) sites in July and November. Data represent means ± 1 SD. Taxa: B, bivalvia; C, crustacea; E, echinodermata; G, gastropoda; P, polychaeta; Pp, polyplacophora.
Table 2. Mean δ13C and δ15N values of macrobenthic consumers collected at the two heavily oil-impacted (Gurepo and Padori) and one non-impacted reference (Yeonpo) sites in July and November. Data represent means ± 1 SD. Taxa: B, bivalvia; C, crustacea; E, echinodermata; G, gastropoda; P, polychaeta; Pp, polyplacophora.
No.Species NameTaxaGurepoPadoriYeonpo
δ13C δ15N δ13C δ15N δ13C δ15N
nMeanSDMeanSDnMeanSDMeanSDnMeanSDMeanSD
July
Suspension feeder
1Balanus albicostatusC4−17.9 0.4 6.0 0.2
2Chthamalus challengeriC4−17.9 0.3 6.2 0.3 4−17.9 0.4 6.2 0.3 4−17.4 0.3 6.2 0.3
3Crassostrea gigasB4−17.5 0.4 7.2 0.5 4−17.4 0.2 6.9 0.6 4−17.7 0.3 7.1 0.5
4Ditrupa arietinaP 3−19.1 0.4 7.1 0.3
5Hydroides ezoensisP3−17.6 0.3 7.0 0.3 3−17.9 0.5 6.8 0.6
6Lasaea undulataB3−18.1 0.5 6.8 0.5 4−18.1 0.4 6.3 0.3 4−18.4 0.4 6.4 0.4
7Mytilus galloprovincialisB4−18.4 0.3 6.4 0.3 4−18.4 0.4 6.5 0.5 4−18.4 0.4 6.8 0.5
Deposit feeder
8Barleeia angustataG4−15.8 0.4 7.9 0.5 4−15.6 0.6 7.9 0.4 4−15.8 0.4 8.1 0.4
9Gammarian amphipodC 4−16.4 0.3 7.6 0.5 4−16.3 0.4 7.8 0.4
10Nainereis laevigataP3−15.3 0.4 7.4 0.4
11Parhyale sp.C4−14.8 0.3 9.5 0.5 4−14.9 0.5 9.7 0.4 4−14.9 0.3 9.6 0.4
Grazer
12Chlorostoma lischkeiG4−16.2 0.5 9.1 0.2 4−15.6 0.5 8.7 0.5
13Chlorostoma turbinataG4−16.6 0.3 9.2 0.7 4−16.6 0.4 8.7 0.6
14Heminerita japonicaG 4−16.0 0.4 7.9 0.5
15Littorina breviculaG4−13.9 0.4 8.6 0.4 4−13.1 0.6 8.8 0.4
16Lottia dorsuosaG4−13.6 0.6 8.6 0.3 4−13.5 0.2 9.0 0.6
17Lottia kogamogaiG 4−14.1 0.5 9.7 0.4
18Lottia tenuisculptaG4−13.1 0.3 9.0 0.5 4−13.0 0.4 8.9 0.3 4−13.0 0.3 9.5 0.6
19Monodonta neritoidesG 4−15.0 0.4 8.9 0.6
20Nipponacmaea schrenckiiG 4−15.7 0.4 9.0 0.6
21Nodilittorina radiataG 4−15.9 0.4 8.6 0.3
22Omphalius rusticusG 4−15.0 0.4 9.0 0.3
23Patelliida conulusG4−13.2 0.4 9.6 0.4
24Patelloida pygmaeaG 4−15.8 0.4 9.0 0.7
25Turbo coronata coreensisG 4−13.2 0.5 10.2 0.5 4−13.4 0.5 10.5 0.5
26Acanthoschitona achatesPp4−13.8 0.4 8.1 0.8 4−13.6 0.4 8.1 0.6
Omnivore
27Asterina pectiniferaE4−12.7 0.5 10.9 0.4 4−12.7 0.3 11.0 0.4 4−12.7 0.5 10.9 0.4
28Hemigrapsus penicillatusC4−17.7 0.3 11.2 0.6 4−16.0 0.5 11.0 0.6 4−15.7 0.4 10.7 0.5
29Pagurus sp.C4−14.7 0.4 10.6 0.4 4−14.4 0.5 11.3 0.3 4−15.3 0.4 10.9 0.3
Carnivore
30Ceratostoma inornatumG4−15.3 0.5 11.9 0.6 4−15.6 0.5 11.8 0.8 4−15.4 0.5 12.0 0.3
31Ceratostoma rorifluumG4−16.5 0.5 12.2 0.7 4−16.5 0.6 11.8 0.5 4−16.2 0.4 12.1 0.6
32Hima fratercula hypoliaG 4−14.6 0.4 12.3 0.4
33Thais clavigeraG4−16.6 0.4 11.9 0.5 4−16.7 0.4 12.0 0.4 4−16.5 0.5 12.5 0.4
November
Suspension feeder
1Balanus albicostatusC4−18.9 0.5 4.7 0.6
2Chthamalus challengeriC4−18.8 0.4 5.1 0.3 4−18.7 0.3 5.1 0.2 4−19.1 0.2 4.9 0.3
3Crassostrea gigasB4−18.2 0.2 5.7 0.6 4−18.0 0.3 6.0 0.5 4−18.5 0.3 6.1 0.3
4Hydroides ezoensisP3−18.1 0.2 5.8 0.5 3−18.2 0.4 5.8 0.5
5Lasaea undulataB4−18.7 0.3 5.9 0.3 4−18.7 0.3 5.5 0.2 4−19.1 0.3 5.4 0.3
6Mytilus galloprovincialisB4−19.2 0.2 5.4 0.4 4−19.0 0.3 5.8 0.4 4−19.2 0.3 6.0 0.3
Deposit feeder
7Barleeia angustataG4−16.2 0.3 6.7 0.3 4−16.5 0.3 6.7 0.4 4−16.8 0.3 7.0 0.2
8Gammarian amphipodC4−17.2 0.4 6.1 0.3 4−17.8 0.3 6.6 0.3 4−17.4 0.6 6.8 0.4
9Nainereis laevigataP3−16.2 0.3 8.4 0.5
10Parhyale sp.C4−15.8 0.2 8.7 0.5 4−16.2 0.5 8.5 0.5 4−15.9 0.4 8.6 0.4
Grazer
11Chlorostoma lischkeiG4−16.9 0.4 8.5 0.6 4−16.2 0.5 8.5 0.4
12Chlorostoma turbinataG4−17.2 0.4 8.8 0.5 4−17.2 0.4 8.4 0.3
13Heminerita japonicaG 4−16.7 0.4 7.8 0.4
14Littorina breviculaG3−15.0 0.3 8.1 0.6 3−14.9 0.3 8.2 0.2
15Lottia dorsuosaG4−14.1 0.4 7.2 0.4 4−14.2 0.5 8.0 0.6
16Lottia tenuisculptaG4−13.8 0.4 7.4 0.6 4−14.0 0.3 7.2 0.2 4−13.8 0.2 8.5 0.6
17Omphalius rusticusG3−15.0 0.6 7.5 0.4 3−15.9 0.4 7.6 0.5
18Turbo coronata coreensisG −14.1 0.3 8.8 0.3 4−14.6 0.4 9.2 0.3
19Acanthoschitona achatesPp4−14.2 0.3 6.2 0.2 4−14.6 0.5 6.9 0.4
Omnivore
20Asterina pectiniferaE4−14.0 0.5 9.7 0.3 4−14.8 0.4 9.9 0.4 4−14.8 0.6 9.9 0.4
21Hemigrapsus penicillatusC4−18.1 0.2 10.0 0.5 4−16.8 0.4 9.8 0.5 4−16.9 0.4 9.8 0.4
22Pagurus sp.C4−15.2 0.3 9.7 0.2 4−15.3 0.4 10.0 0.3 4−15.3 0.5 9.9 0.5
Carnivore
23Ceratostoma inornatumG4−16.0 0.4 11.0 0.4 4−16.2 0.3 10.9 0.2 4−15.9 0.4 11.1 0.4
24Ceratostoma rorifluumG4−17.0 0.2 10.8 0.4 4−17.1 0.5 10.6 0.4 4−17.0 0.3 10.8 0.4
25Hima fratercula hypoliaG 4−15.5 0.4 11.3 0.2
26Thais clavigeraG4−17.2 0.4 11.0 0.3 4−17.1 0.3 11.0 0.2 4−17.0 0.3 11.4 0.5
Table 3. Mean δ13C and δ15N values of consumer feeding groups (suspension feeder, deposit feeder, grazer, omnivore, carnivore) at the two heavily oil-impacted (Gurepo and Padori) and one non-impacted reference (Yeonpo) sites in July and November. PERMANOVA test of δ13C and δ15N values for each consumer feeding group among the three sites and between the two seasons (significance at p < 0.05). Data represent means ± 1 SD.
Table 3. Mean δ13C and δ15N values of consumer feeding groups (suspension feeder, deposit feeder, grazer, omnivore, carnivore) at the two heavily oil-impacted (Gurepo and Padori) and one non-impacted reference (Yeonpo) sites in July and November. PERMANOVA test of δ13C and δ15N values for each consumer feeding group among the three sites and between the two seasons (significance at p < 0.05). Data represent means ± 1 SD.
GurepoPadoriYeonpo
Potential Food Source δ13C δ15N δ13C δ15N δ13C δ15N
nMeanSDMeanSDnMeanSDMeanSDnMeanSDMeanSD
July
Suspension feeder22−17.90.46.60.622−18.10.66.60.516−17.90.56.60.5
Deposit feeder11−15.30.68.32.512−15.60.88.41.012−15.70.78.50.9
Grazer28−14.41.48.90.628−14.91.39.00.736−14.31.39.00.9
Omnivore12−15.02.210.90.512−14.41.511.10.412−14.61.510.80.4
Carnivore12−16.20.712.00.616−15.81.012.00.512−16.00.612.20.5
November
Suspension feeder23−18.70.55.40.619−18.60.55.60.516−19.00.45.60.5
Deposit feeder15−16.40.67.41.212−16.80.87.31.012−16.70.87.50.9
Grazer26−15.21.47.70.920−15.51.38.10.723−14.81.08.10.8
Omnivore12−15.81.89.80.412−15.61.09.90.412−15.71.09.90.4
Carnivore12−16.80.611.00.316−16.50.811.00.312−16.60.611.10.4
PERMANOVASeasonSiteInteraction
pseudo-Fp-valuepseudo-Fp-valuepseudo-Fp-value
Suspension feeder104.900.0010.550.6631.010.391
Deposit feeder35.580.0011.070.3550.360.735
Grazer36.330.0011.470.2120.360.815
Omnivore33.200.0010.580.6260.270.816
Carnivore50.050.0010.790.5330.110.901
Table 4. Isotopic niche areas of each feeding group (suspension feeder, deposit feeder, grazer, omnivore, carnivore) estimated as the total area (TA) and standard ellipse area (SEAc) and isotopic niche overlaps (percentage) between the heavily oil-impacted (Gurepo, G and Padori, P) and non-impacted reference (Yeonpo, Y) sites using the Stable Isotope Bayesian Ellipse in R (SIBER) procedure.
Table 4. Isotopic niche areas of each feeding group (suspension feeder, deposit feeder, grazer, omnivore, carnivore) estimated as the total area (TA) and standard ellipse area (SEAc) and isotopic niche overlaps (percentage) between the heavily oil-impacted (Gurepo, G and Padori, P) and non-impacted reference (Yeonpo, Y) sites using the Stable Isotope Bayesian Ellipse in R (SIBER) procedure.
Isotopic Niche Metrics
Feeding GroupGurepoPadoriYeonpoPercentage Overlap (%)
TASEAcTASEAcTASEAcG and YP and Y
July
Suspension feeder2.320.802.770.992.010.928084
Deposit feeder3.051.512.921.642.671.236082
Grazer8.182.988.312.409.453.237662
Omnivore6.383.454.002.203.271.897160
Carnivore2.731.483.301.702.130.996655
November
Suspension feeder2.040.801.970.701.310.664140
Deposit feeder3.891.833.621.953.211.705483
Grazer7.293.176.022.866.082.714862
Omnivore5.222.532.341.242.551.385792
Carnivore1.370.772.090.911.570.806161
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Park, H.J.; Kim, C.; Kang, C.-K. Recovery of Macrobenthic Food Web on Rocky Shores Following the Hebei Spirit Oil Spill as Revealed by C and N Stable Isotopes. Water 2022, 14, 2335. https://doi.org/10.3390/w14152335

AMA Style

Park HJ, Kim C, Kang C-K. Recovery of Macrobenthic Food Web on Rocky Shores Following the Hebei Spirit Oil Spill as Revealed by C and N Stable Isotopes. Water. 2022; 14(15):2335. https://doi.org/10.3390/w14152335

Chicago/Turabian Style

Park, Hyun Je, Changseong Kim, and Chang-Keun Kang. 2022. "Recovery of Macrobenthic Food Web on Rocky Shores Following the Hebei Spirit Oil Spill as Revealed by C and N Stable Isotopes" Water 14, no. 15: 2335. https://doi.org/10.3390/w14152335

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop