Anthropogenic climate change is reducing ice and snow thickness in the Arctic. The loss of summer sea ice has led to increased access to Arctic waters and the development of marine resources, which raises the risk of oil spills. Thinning ice and snow also increases irradiance in the upper ocean which is predicted to increase primary productivity, disfavoring shade-adapted sea-ice algae while benefitting phytoplankton and cryopelagic taxa. Studies have confirmed the lethality of crude oil and its distillates to Arctic phytoplankton; less well-constrained are the sublethal impacts to sea-ice algae in combination with other drivers. This study investigates the combination of two drivers, crude oil exposure and irradiance, on the growth rate and maximum cell concentration of four sea-ice diatoms (Attheya septentrionalis, Fragilariopsis cylindrus, and two strains of Synedropsis hyperborea) isolated from landfast sea ice near Utqiaġvik, Alaska. Crude oil inhibition of growth was complex and dependent on species and irradiance level. A. septentrionalis was generally tolerant to crude oil exposure, but toxicity was enhanced at the highest irradiance. The cryopelagic taxon, F. cylindrus, exhibited strong growth inhibition at TPH concentrations greater than approximately 6 mg L−1. Growth rates of S. hyperborea strains were stimulated at low concentrations of oil at all light levels. A simple numerical model was used to simulate an oil spill under varying snow depths to follow composition of a mock community comprised of these four isolates across a spring season. Results highlight that the reduction of algal biomass accumulation and the community composition change following a crude oil spill are more severe in a simulated low-snow spring, due to the relative sensitivity of F. cylindrus. We show that a brighter Arctic, which is predicted to increase the relative importance of cryopelagic taxa like F. cylindrus, may render the Arctic ecosystem more vulnerable to crude oil spills.

The loss of summer sea ice in the Arctic is driving the development of marine resources and industries such as shipping lanes, fisheries, tourism, and petroleum exploration (Arctic Council, 2009; Conservation of Arctic Flora and Fauna, 2015; Eguíluz et al., 2016; Cao et al., 2022), which increases the risk of oil spills. Spilled oil may accumulate by pooling at the underside of the ice, saturate the highly porous skeletal layer, percolate through the brine channel system, and become encapsulated in growing ice (Arctic Monitoring and Assessment Programme [AMAP], 2007; Petrich et al., 2013; Oggier et al., 2020). Oil encapsulated in ice remains relatively unweathered, given limited accessibility to microbes and contact with liquid water, therefore retaining much of its original volatility (Boccadoro et al., 2018; Bakke et al., 2021). Encapsulated oil can remobilize under or within sea ice through physical processes such as ice deformation, melting, lead expansion or contraction, and spreading with strong currents (AMAP, 2007). Spills, and subsequent redistribution of oil through dynamic ice processes, render the Arctic ecosystem vulnerable to unweathered oil exposure at any time of the year, even after the point-source has been contained. The ability to respond to an oil spill in the Arctic is impeded by the presence of sea ice which acts as a physical barrier to traditional mechanical and chemical cleanup strategies, for example, boom and skim and dispersant application. A worst-case spill scenario, an oil well blowout, in ice-covered waters would likely persist unmitigated through winter until conditions were more favorable for containment and cleanup response during the following summer. Models show that oil released from an uncontained blowout could accumulate and travel with ice movements >1000 km and affect up to 1 million km2 over the course of a year (Blanken et al., 2017). Toxicological research on the impact of crude oil on the marine species and biological communities of the Arctic is necessary to understand the ecological repercussions of an oil spill.

Average annual sea-ice extent has declined by approximately 2 million km2 since 1979 (Onarheim et al., 2018), and summer sea ice is predicted to be absent from all Arctic shelf seas by 2050, regardless of emission scenario (Årthun et al., 2021). Loss of Arctic sea ice will change the relative roles of sympagic and pelagic primary producers dramatically (Tedesco et al., 2019; Lannuzel et al., 2020; and references therein). Despite this habitat loss, sea-ice algae will continue to grow in ice that forms each winter on an annual basis. Bacillariophyta (diatoms) proliferate in polar seas and dominate microalgal biomass and diversity within Arctic sea ice (Poulin et al., 2011; Campbell et al., 2016; Szymanski and Gradinger, 2016). The majority of ice-algal biomass is confined to the lowermost centimeters of sea ice where nutrient exchange with underlying seawater supports growth. The vernal ice-algal bloom generally occurs 1–3 months prior to the subsequent phytoplankton bloom (Hsiao, 1992; Ji et al., 2013; Mundy et al., 2014), fueling the base of the Arctic food web (Kohlbach et al., 2016; Kohlbach et al., 2019) by supporting dependent grazers like the copepod Calanus glacialis (Søreide et al., 2010) and under-ice amphipods (Poltermann, 2001; Gradinger and Bluhm, 2009). Key species like the Arctic cod (Boreogadus saida) link ice-algal production to higher trophic level consumers, for example, ringed seals, which are an important subsistence food with cultural significance in Arctic communities (Steiner et al., 2019). The accumulation of microalgae, their grazers, and the congregation of fish at the ice–water interface, renders the sea-ice ecosystem highly susceptible to crude oil exposure that would also accumulate at this interface. Not surprisingly, most toxicological studies on Arctic biota have focused on fauna with important trophic positions, but less attention has been paid to the photoautotrophs that form the base of the food web.

Crude oil is a complex mixture of organic compounds with large compositional differences between source oils (Faksness and Brandvik, 2008); each oil compound has a different solubility and toxicological impact, making the intercomparison of oil toxicity experiments inherently difficult. Exposure to hydrocarbons can result in a decline in growth rate and biomass for phytoplankton (Hsiao, 1978; Gilde and Pinckney, 2012; Bretherton et al., 2018) and sea-ice algae (Fiala and Delille, 1999; Dilliplaine et al., 2021), a decrease in primary productivity (Hsiao et al., 1978; Lemcke et al., 2019), loss of cell motility (Soto et al., 1975; Garr et al., 2014), and even stimulation of growth (Dunstan et al., 1975; Özhan et al., 2014; Bretherton et al., 2018). Part of the differing responses may be due to differences in source oils and interplay with environmental variables. Oil exposure responses also vary widely across microalgal groups, species, and even strains (Gaur and Kumar, 1981; Huang et al., 2011; Gilde and Pinckney, 2012; Özhan et al., 2014; Bretherton et al., 2020); attempts to generalize responses based on higher taxonomic classification or functional groups, for example, diatoms and dinoflagellates, have yielded conflicting results (Taş et al., 2010; Ozhan and Bargu, 2014; Bretherton et al., 2020), though smaller microalgae are generally more sensitive to crude oil exposure due to the increased surface area to volume ratio (Echeveste et al., 2010).

The mechanisms responsible for crude oil growth inhibition remain poorly understood, but wide-ranging damage has been reported, for example, to nucleic acid synthesis and structure (El-Sheekh et al., 2000; Parab et al., 2008), light-harvesting complex proteins (Kamalanathan et al., 2021), and membranes (Sikkema et al., 1995; Hook and Osborn, 2012; Dilliplaine et al., 2021), including limitations to nutrient uptake and gas exchange (Koshikawa et al., 2007). These cell structures may be damaged directly by contact with crude oil or indirectly via other compounds, such as reactive oxygen species induced by oil exposure (Ozhan et al., 2015; Kamalanathan et al., 2021). Species-specific sensitivity and changes to growth rate can alter microalgal community composition with implications for higher trophic levels (Gilde and Pinckney, 2012). Isolate experiments provide detailed response information and illuminate taxon-specific sensitivity thresholds and mechanisms of response to multiple stressors. The toxicological effects of crude oil have the potential to be enhanced when combined with other drivers such as ultraviolet radiation, salinity, and temperature (Sargian et al., 2007; Bender et al., 2021; DeLorenzo et al., 2021).

The near absence of light during the Arctic polar night limits photosynthesis (Berge et al., 2015). The seasonal transition from winter to spring coincides with the return of the sun above the horizon and supports the growth of sea-ice microalgae that are well adapted to low light, growing at irradiances as low as 0.17 µmols of photons m−2 s−1 (Hancke et al., 2018). This low-light efficiency, combined with their fixed location in the photic zone, provides a growth advantage which allows for accumulation of microalgal biomass in spring prior to the onset of the subsequent phytoplankton bloom (Tremblay et al., 2008; Leu et al., 2011; Song et al., 2016; Ardyna et al., 2020). These adaptations also make them particularly susceptible to photoinhibition and damage when irradiances are excessive (Lund-Hansen et al., 2014; Croteau et al., 2022). Significant decreases in diatom photosynthesis and growth rate due to photoinhibition at light levels above 50 µmol photons m−2 s−1 can occur in ice algae (Arrigo et al., 2010; Juhl and Krembs, 2010). Irradiance values increase at the ice–water interface by several orders of magnitude during the seasonal transition from snow-covered winter ice to summer melt pond formation (Perovich et al., 1998; Nicolaus et al., 2013).

The primary controllers of light availability to sea-ice algae are snow thickness and ice thickness (Veyssière et al., 2022). The extinction coefficient of snow is over an order of magnitude greater than bare ice; therefore, snow cover has a disproportionately larger attenuative effect than ice thickness (Grenfell and Maykut, 1977). The general decline of ice and snow thickness (Kurtz et al., 2011; Webster et al., 2014; Stroeve et al., 2020) is leading to a brighter Arctic that can result in photoinhibition and damage, and may favor replacement of microalgal species that are better adapted for growth at higher irradiances (Croteau et al., 2022). In particular, the current sea-ice algal specialists, like the dominant Nitzschia frigida and ubiquitous Synedropsis hyperborea, are expected to be replaced in the ice by cryopelagic (abundant in both ice and water) taxa, for example, Fragilariopsis cylindrus (Lannuzel et al., 2020). Understanding the acclimatization potential of sea-ice algal specialists, and cryopelagic taxa, to the large range of irradiances encountered on an annual basis, in conjunction with crude oil exposure, is important to better predict how multiple drivers will impact sea-ice algal growth in a brighter Arctic.

The main objectives of our study were to investigate the photoacclimative ability of select sea-ice diatom isolates to realistic irradiances and to determine the effect of crude oil contamination on growth rate and cell concentration. These two easily-measured variables are important to consider with regard to the timing and magnitude of the spring ice-algal bloom. We hypothesized that the combination of the two anthropogenic drivers of increased irradiance and crude oil exposure would have interactive effects that differed by species. To test this hypothesis, we freshly isolated sympagic algae from sea ice collected near Utqiaġvik, Alaska, and exposed them to a range of water-accommodated fraction (WAF) concentrations of Alaska North Slope (ANS) crude oil over a range of irradiances. To contextualize our results, we simulated the toxicological effect of a crude oil spill prior to the onset of the sea-ice algal bloom on biomass and community composition of a simple microalgal community comprised of the four isolates used in this study.

2.1. Isolation and identification of sea-ice algal strains

Cores were collected from landfast sea ice near Utqiaġvik, Alaska (Chukchi Sea; 71°22' N, 156°35' W), on May 25, 2020. The bottom approximately 10 cm of core were melted slowly in the dark, with the addition of 0.2 µm filtered seawater to minimize osmotic shock (Garrison and Buck, 1986). Aliquots of the core melt material were diluted into f/25 +Si medium (Bigelow Labs, ME, USA) according to Guillard and Ryther (1962) using natural seawater (salinity of 36) collected from the Gulf of Alaska at a depth exceeding 500 m. In brief, seawater was filter-sterilized with a 0.2 µm Sterivex filter (MilliporeSigma, St. Louis, MO, USA), macronutrients, and trace metals were added, and the medium was autoclaved. To complete the preparation, filter-sterilized (0.2 µm) vitamins were added after cooling to avoid denaturation of the vitamins. Mixed sea-ice algal cultures were maintained at approximately 15 µmol photons m−2 s−1 in 25 mm culture tubes at 4°C with a 12:12 h light:dark cycle using a low temperature incubator (Percival, Perry, IA, USA) for propagation and isolation.

Sea-ice algae were isolated from these mixed cultures using 1.25% agar plates containing f/25 +Si medium following Kimura and Tomaru (2013). Briefly, 5 mL of diluted cultures were added to the agar plates allowing cells to settle for 12 h before decantation. Plates were left to incubate until colonies were visible, at approximately 1 month. Single colonies were selected using an inoculation loop, resuspended in f/25 +Si medium, and maintained by transfer to fresh medium every 2–4 weeks. Isolates were confirmed to be unialgal by light microscopy (Olympus BX51, Center Valley, PA, USA; Figure 1). Identification to the species level was performed using a combination of light microscopy, scanning electron microscopy (SEM; FEI Quanta 200), and 18S rRNA gene sequencing (Table S1).

Figure 1.

Diatoms isolated from sea ice near Utqiaġvik, Alaska. Differential interference contrast micrographs of (A) Attheya septentrionalis, (B) Fragilariopsis cylindrus, (C) Synedropsis hyperborea strain 1, and (D) Synedropsis hyperborea strain 2. Scales bars are 20 µm.

Figure 1.

Diatoms isolated from sea ice near Utqiaġvik, Alaska. Differential interference contrast micrographs of (A) Attheya septentrionalis, (B) Fragilariopsis cylindrus, (C) Synedropsis hyperborea strain 1, and (D) Synedropsis hyperborea strain 2. Scales bars are 20 µm.

Close modal

2.2. Crude oil WAF preparation and chemical analyses

To test the impact of crude oil contamination on growth physiology of these four sea-ice diatom isolates, oil-contaminated media were generated using ANS crude oil at environmentally relevant temperature. ANS crude oil, collected from Pump Station 1 at the Trans-Alaska Pipeline entry point (provided by Alyeska Pipeline Services), was used as a regionally representative oil for this experiment. The WAF of crude oil was prepared according to the Chemical Response to Oil Spills: Ecological Research Forum (CROSERF; Aurand and Coelho, 2005) and Singer et al. (2000). Briefly, 10.5 L of f/25 +Si medium were added to a 20 L acid-cleaned glass aspirator bottle, and 12 g L−1 of ANS crude oil were gently added to the surface of the water. The mixture was stirred continuously for 48 h on a stir plate to produce a vortex of approximately 25% of the medium height, without the formation of droplets, at −1°C in the dark. WAF was collected from below the oil layer through the spigot and was free of any visible oil.

Samples for chemical analysis were collected in 200 mL acid-cleaned amber glass jars with Teflon-lined lids and stored at −20°C until further processing, approximately 12 months from collection. Concentrations of total petroleum hydrocarbons (TPH) were measured from the undiluted WAF at the onset of the experiment as technical replicates (n = 3; Figure S1). Chemical analyses of the stock WAF (100%) were conducted at Emax Laboratories, Inc. (Torrance, CA), using the Alaska Department of Environmental Conservation (ADEC) methods for n-alkanes in the diesel range organics (DRO; C10–C25), residual range organics (RRO; C25–C36), and TPH (C8–C40) using gas chromatography equipped with a flame ionization detector (GC-FID; Agilent G1530 N; Figure S1), according to methods AK 102/103 (ADEC, 2002a; 2002b). Petroleum concentrations in the dilutions were calculated from the stock concentration as it has been shown to be conserved linearly (Mcfarlin et al., 2011; Gardiner et al., 2013; unpublished data).

2.3. Experiment setup

Experiments were conducted in a walk-in cold room (Environmental Growth Chambers, Chagrin Falls, OH, USA) located at the University of Alaska Fairbanks. Sea-ice algal isolates were transferred to optically clear 96-well flat bottom plates (Eppendorf, Hamburg, DE) and incubated at −1°C to approximate the temperature of the bottom skeletal layer of sea ice. Plates were set on rigid icepacks (40 × 40 × 1 cm; Sonoco ThermoSafe) with a melting point of −1°C to maintain stable temperature. Five light conditions, 3, 10, 20, 50, and 125 µmol photons m−2 s−1, were established using LED lights (54 W cool white; Lithonia Lighting, Conyers, GA, USA) with a 12:12 light:dark cycle. Photosynthetically active radiation (PAR) was measured with a Li-190 R planar light sensor (LI-COR, Lincoln, NE, USA), and intensity was tuned using a combination of voltage regulators and neutral density screens. Cultures were transferred to fresh medium approximately every 1–2 weeks, once exponential growth was achieved, and allowed to acclimate to light conditions over 70 days, ensuring that cultures were photoacclimated prior to the determination of growth physiology.

Two sets of 96-well plates (experimental and unoiled) were prepared for each of the four isolates and five light conditions used in the experiment. Experimental plates contained a linear dilution gradient of WAF, from 0% to 90%, for a total of 12 treatments, each with 8 replicates per plate. Unoiled plates were used exclusively for cell harvest during the course of the experiment and contained oil-free medium (0% WAF). Photoacclimated cultures were transferred to each well, after pipetting to mix, at a 1:9 volume ratio of cell culture:WAF. The range of median initial cell densities within wells was 9.5–11.5 × 103 cells mL−1. Fluorescence was measured daily using a SpectraMax Gemini EM plate reader (Molecular Devices, San Jose, CA, USA) at wavelengths for in vivo chlorophyll a fluorescence (460 ex, 665 em) 8 h after lights were turned off. Experiments were ended after a leveling off, or decline, in fluorescence was observed (Figure S2).

2.4. Dose-response crude oil × irradiance experiments

Cells were harvested from the unoiled plates, immediately after fluorescent measurement, four times during the experiment; four replicate samples were collected from each plate and preserved with a 2% final concentration of Lugol’s solution and stored at 4°C until further processing (<6 months). Each fixed sample was a pool of four adjacent wells, with relative fluorescence averaged among these four wells for calculation of cell concentration. Cells were counted using a Sedgewick rafter counting chamber (Forestry Suppliers Inc., Jackson, MS, USA), with at least 100 cells, or 10 µL total volume, counted for each sample. Fragilariopsis cylindrus and Attheya septentrionalis were enumerated manually using an inverted Nikon Eclipse TE2000-U at 400×. Synedropsis hyperborea cells were enumerated from micrographs taken with a Teledyne Photometrics Prime BSI camera (Tucson, AZ, USA) with automated image analysis using the Cell Counter plugin (v. 3.0.0) for imageJ (Schneider et al., 2012). Image analysis counts closely matched manual counts with a slight underestimation of <5%. The relationships between relative chlorophyll a fluorescence and cell concentration were determined using linear regression with an intercept forced through 0 for each combination of isolate and irradiance (Figure S3). Relative fluorescence measured in the experimental plates was converted to cell concentration by multiplying the blank-corrected value by the empirically determined slope.

Maximum growth rate (µmax, d−1) was calculated as the log-linear change in cell concentration (Ce, cells mL−1) over time (t) using Equation 1:

1

and the all_easylinear() function described by Hall et al. (2014; R package growthrates 0.8.2) for each individual well. Maximum cell concentration was calculated as the maximum observed cell concentration during the experiment. Error was propagated as the sum of squares using calculated errors associated with cell counts and biological replication.

2.5. Data analysis

Photoacclimated growth rate versus irradiance curves were generated for each of the four isolates using the unoiled control wells. The R package “phytotools” (Silsbe and Malkin, 2015) was used to fit the Eilers and Peeters (1988) growth-irradiance model and estimate growth parameters using the Levenberg-Marquardt fitting algorithm (Moré, 1978). This model was selected because it includes a term for photoinhibition, which we expected to occur within our range of irradiances for sea-ice microalgae. The saturation irradiance, ek, was calculated as the intersection of µ and the initial slope of the growth-irradiance curve, α.

Dose-response curves were generated using species-specific maximum growth rates and concentrations using nonlinear models fit with the R package “drc” (Knezevic et al., 2007). Models with and without irradiance as a factor were tested for significance using ANOVA (p < 0.05); irradiance was significant and included as a factor in all models. Three curves (Attheya septentrionalis, irradiance level of 50 µmol photons m−2 s−1; Synedropsis hyperborea strain 1, irradiance level 10 µmol photons m−2 s−1; and S. hyperborea strain 2, irradiance level 20 µmol photons m−2 s−1) were removed before model fitting due to the presence of a positive trend resulting in model direction reversal and fit. A linear regression was fit to these treatments instead. The best fitting model was selected using Akaike’s information criterion (AIC), and visual inspection when values were close (AIC < 10; Table S2). Models for growth rate have a fixed lower limit of 0, with the assumption that there will be no growth at some concentration of WAF (not observed). The lower limit for dose-response curves generated using the maximum concentration was set to the median initial concentration per species. The concentrations required to inhibit growth rate or maximum cell concentration by 10% (IC10) and 50% (IC50) were calculated from the model fits (Knezevic et al., 2007). Due to the variability of the lower limit set point in models used between growth rate and cell concentration, IC values are absolute for growth rate and relative to initial cell concentrations for maximum cell concentration.

2.6. Modeling the impact of oil exposure and snow cover on sea-ice algal growth and diversity

A simple numerical model was used to simulate the growth of a simplified ice algal community over a spring-bloom period in the presence and absence of crude oil exposure with realistic under-ice irradiance. The under-ice irradiance (UIPAR) was calculated using the Beer-Lambert law (Equation 2) and forced with observations of incident irradiance and sea-ice thickness in the Alaskan Arctic (Figure S4). Incident irradiance (PAR; 30 min average) was downloaded from the National Ecological Observatory Network (NEON) at the Utqiaġvik, Alaska, station (BARR) for the years 2020 (Figure S4A) and 2022 (Figure S4B). PAR data were distilled to a single daily irradiance value (I) by averaging the irradiance within a 12 h window centered on solar noon from March 1 to May 15 (Figure S4C). Irradiance values were then averaged across years for the same date and time, while missing data were filled with values from the year with available data. A daily sea-ice thickness (Hi) value was calculated as the average of all daily point measurements (every 15 min) measured at a Sea Ice Mass Balance station in 2014 (Chukchi Sea; 71°22' N, 156°31' W) using an upward-facing acoustic transducer (NSF Arctic Data Center, 2009; Figure S4E). Fixed snow depth (Hs) was chosen to simulate under-ice light intensities under snow depths ranging from 0 to 20 cm (Figure S4D). The integrated PAR attenuation coefficients for snow (Ks, 11) and ice (Ki, 3.6) were derived from McDonald et al. (2015); Ki was adjusted to 2.6, within the reported confidence interval, to include upwelling irradiance as measured in situ (May 2022) using a spherical quantum sensor (LI-COR).

2

The simulated ice-algal community was composed of the four isolates investigated in this study seeded at equal starting concentrations. Nutrient concentrations and fluxes were not considered in this simplified model, as it is intended only to summarize the effect of oil on growth rate based on our experimentally determined values for our isolates, free of nutrient limitation. Therefore, biomass estimations should be considered relative. Cell abundance (Ad) was calculated for each day assuming exponential growth (Equation 3) of cell abundance from the previous time step (Ad−1). Isolate-specific growth rate (µ) was a function of irradiance (UIPAR) and one of two oil concentrations determined empirically in this study: no oil contamination, and at the highest tested concentration (8.7 mg L−1 TPH). A loss term (g) was fixed at a constant fraction of 0.10 of the taxon-specific growth rate (Lavoie et al., 2005). Cell abundance (A) at the initial time step was set at 1000 cells for each of the four algae. Cell concentrations were converted to carbon content (pg C) using the isolate-specific biovolume (BV, µm3) and the relationship derived from all data during exponential growth according to Lomas et al. (2019; Table S1; Text S1).

3

3.1. Characterization of sea-ice diatom isolates

In order to assess the potential for interacting anthropogenic impacts on the Alaskan Arctic ecosystem, sea-ice algae were isolated near Utqiaġvik, Alaska. Four diatoms, Attheya septentrionalis, Fragilariopsis cylindrus, and two strains of Synedropsis hyperborea were selected from among these isolates based on their prevalence in the original samples, pan-Arctic distribution (Hasle et al., 1994; Poulin et al., 2011; Figure 1), and dominance in Arctic ice, sub-ice, and water samples (Hegseth, 1992; von Quillfeldt, 1995; Melnikov et al., 2002). The four diatom isolates were diverse in their morphology (Figure 1). They varied in biovolume, with the largest being A. septentrionalis, followed by F. cylindrus and S. hyperborea (Table S1). The two strains of S. hyperborea varied in morphology and size, with S. hyperborea strain 1 being longer and thinner compared to the shorter and broader S. hyperborea strain 2 (Figure S5).

The sea-ice diatom isolates also varied in light-acclimated growth-irradiance curves (Figure 2). Ranges of ice-algal growth rates were 0.16–0.40 d−1, 0.15–0.50 d−1, 0.15–0.39 d−1, and 0.15–0.36 d−1 for Attheya septentrionalis, Fragilariopsis cylindrus, Synedropsis hyperborea strain 1 and S. hyperborea strain 2, respectively. Ranges fell within those previously reported for A. septentrionalis (Hsiao, 1978; Hegseth, 1992; Basionym = Chaetoceros septentrionalis Østrup) and Arctic strains of F. cylindrus (Pančić et al., 2015); these are the first reports for S. hyperborea. The irradiance at the onset of inhibition, ephot, was indistinguishable with certainty between the two strains of S. hyperborea (Table 1). Other growth-irradiance model parameters differed by strain (Table 1). The parameter α was smallest for F. cylindrus, intermediate for S. hyperborea strains 1 and 2, and greatest for A. septentrionalis, while ek was highest for F. cylindrus, with the lowest ek calculated for S. hyperborea strain 2 (Table 1).

Figure 2.

Growth-irradiance curves for sea-ice diatom isolates. Exponential growth (day−1) versus irradiance (µmols photons m−2 s−1) for four isolates: (A) Attheya septentrionalis, (B) Fragilariopsis cylindrus, (C) Synedropsis hyperborea strain 1, and (D) Synedropsis hyperborea strain 2. Points indicate the light-acclimated growth rates of each strain (mean ± SE; n = 8); lines indicate growth-irradiance model fits.

Figure 2.

Growth-irradiance curves for sea-ice diatom isolates. Exponential growth (day−1) versus irradiance (µmols photons m−2 s−1) for four isolates: (A) Attheya septentrionalis, (B) Fragilariopsis cylindrus, (C) Synedropsis hyperborea strain 1, and (D) Synedropsis hyperborea strain 2. Points indicate the light-acclimated growth rates of each strain (mean ± SE; n = 8); lines indicate growth-irradiance model fits.

Close modal
Table 1.

Growth-irradiance model type and parameter estimates for sea-ice diatom isolates

IsolateModel Parameter Estimatesa
α ± SEµ ± SEek ± SEephot ± SE
Attheya septentrionalis 8.0 ± 0.4 × 10−2 0.40 ± 0.0 4.9 ± 0.2 67.1 ± 5.1 
Fragilariopsis cylindrus 6.5 ± 0.6 × 10−2 0.50 ± 0.1 7.7 ± 0.7 107 ± 32.5 
Synedropsis hyperborea str. 1 7.2 ± 0.3 × 10−2 0.39 ± 0.0 5.5 ± 0.2 55.4 ± 2.4 
Synedropsis hyperborea str. 2 7.8 ± 0.7 × 10−2 0.36 ± 0.1 4.6 ± 0.4 112 ± 52.5 
IsolateModel Parameter Estimatesa
α ± SEµ ± SEek ± SEephot ± SE
Attheya septentrionalis 8.0 ± 0.4 × 10−2 0.40 ± 0.0 4.9 ± 0.2 67.1 ± 5.1 
Fragilariopsis cylindrus 6.5 ± 0.6 × 10−2 0.50 ± 0.1 7.7 ± 0.7 107 ± 32.5 
Synedropsis hyperborea str. 1 7.2 ± 0.3 × 10−2 0.39 ± 0.0 5.5 ± 0.2 55.4 ± 2.4 
Synedropsis hyperborea str. 2 7.8 ± 0.7 × 10−2 0.36 ± 0.1 4.6 ± 0.4 112 ± 52.5 

aInitial slope of the growth-irradiance curve (α; d−1 divided by µmol photons m−2 s−1), maximum growth rate (µ; d−1), saturation irradiance (ek; µmol photons m−2 s−1), and irradiance at the onset of photoinhibition (ephot; µmol photons m−2 s−1). Standard error (SE) calculated using the dependent R package and function FME: modFit for n = 40.

3.2. Sublethal impacts of crude oil modified by irradiance

The mean concentration (± SD; n = 3) of TPHs was 9.7 (± 0.6) mg L−1; of DROs, 3.0 (± 0.0) mg L−1; and of RROs, 0.6 (± 0.3) mg L−1. Approximately 60% of the TPHs were not attributed to the DRO or RRO fractions; nearly all of the remaining TPH concentration was attributed to the lowest molecular weight C8–9 fraction, which is considered to be the range for gasoline organics (C6–C10; Figure S1).

Sea-ice diatom isolates were grown in 96-well plates with a gradient of ANS crude oil-contaminated media under five irradiances. There was interspecific variability in sea-ice algal growth rate responses to crude oil contamination that was modified by irradiance (Figure 3, Table 2). Attheya septentrionalis was relatively insensitive to ANS crude oil, except at the highest irradiance level where growth rates declined by approximately 19% at the highest tested TPH concentration (Figure 3A). The 10% growth inhibition concentration (IC10) of ANS crude oil WAF for A. septentrionalis was calculable for the lowest (3 µmol photons m−2 s−1) and highest (125 µmol photons m−2 s−1) irradiance levels, with a significantly lower IC10 at the highest irradiance level (Table 2, Figure S6). Values of IC10 for Fragilariopsis cylindrus were calculated for all five irradiances, owing to its greater apparent sensitivity (Figure 3B) and to being the only isolate to exhibit inhibition of the growth rate by 50% (IC50) within the tested concentrations (Table 2). There were significant differences in F. cylindrus IC10 by irradiance (p < 0.05), with a lower IC10 for the intermediate (10 and 20 µmol photons m−2 s−1) compared with the high (50 and 125 µmol photons m−2 s−1) irradiance levels. Values of IC50, calculated for F. cylindrus at irradiance levels 10, 50, and 125 µmol photons m−2 s−1, differed significantly from one another (Table 2; Figure S6); the highest irradiance induced the lowest IC50. Synedropsis hyperborea strain 1 was the least sensitive isolate when comparing calculable IC10 values, though both strains of S. hyperborea exhibited relatively low sensitivity to ANS crude oil WAF (Figure 3; Table 2). IC10 fell within the range of tested concentrations only at the lowest irradiance level (3) for both strains of S. hyperborea, with strain 2 exhibiting a significantly lower threshold than strain 1 (Table 2; Figure S6). Both strains of S. hyperborea exhibited a significant (p < 0.05) hormetic response, approaching 20% relative to the control for S. hyperborea strain 2 at some light levels. Three of the dose-response curves were fit with linear models, A. septentrionalis (irradiance of 50 µmol photons m−2 s−1), S. hyperborea strain 1 (irradiance of 10 µmol photons m−2 s−1), and S. hyperborea strain 2 (irradiance of 20 µmol photons m−2 s−1), because they exhibited an increase in growth rate with increasing TPH across the range of tested concentrations (Figure 3A, C, and D). Linear regressions were significant for S. hyperborea strains 1 and 2 (p = 0.0084, R2 = 0.52; p = 0.026, R2 = 0.41, respectively) but not for A. septentrionalis (p = 0.18, R2 = 0.17).

Figure 3.

Dose-response curves of growth rate versus crude oil as modified by irradiance. (A–D) Growth rate (d−1) and (E–H) relative growth rate, proportional to the control (no oil; E–H), of the four isolates, Attheya septentrionalis (A and E), Fragilariopsis cylindrus (B and F), Synedropsis hyperborea strain 1 (C and G), and S. hyperborea strain 2 (D and H), at irradiances of 3 (blue), 10 (purple), 20 (maroon), 50 (orange), and 125 (yellow) µmol photons m−2 s−1. Data points are means; error bars, ± SE (n = 8). The shaded ribbon is the 95% confidence interval of the model fit. Black dashed lines are linear fits; see Table S2 for model types. Solid black line is the 1.00 relative growth rate isoline.

Figure 3.

Dose-response curves of growth rate versus crude oil as modified by irradiance. (A–D) Growth rate (d−1) and (E–H) relative growth rate, proportional to the control (no oil; E–H), of the four isolates, Attheya septentrionalis (A and E), Fragilariopsis cylindrus (B and F), Synedropsis hyperborea strain 1 (C and G), and S. hyperborea strain 2 (D and H), at irradiances of 3 (blue), 10 (purple), 20 (maroon), 50 (orange), and 125 (yellow) µmol photons m−2 s−1. Data points are means; error bars, ± SE (n = 8). The shaded ribbon is the 95% confidence interval of the model fit. Black dashed lines are linear fits; see Table S2 for model types. Solid black line is the 1.00 relative growth rate isoline.

Close modal
Table 2.

Inhibitory concentrations of total petroleum hydrocarbons to growth and maximum cell concentration by species and irradiance

IsolateMeasurea (% inhibited)Inhibitory Concentrationsb (mg L−1) for 5 Irradiances (µmol photons m−2 s−1)
3102050125
Attheya septentrionalis µ (10) >8.7 >8.7 8.2 ± 0.6 >8.7 3.4 ± 0.3 
Ce (10) >8.7 >8.7 >8.7 2.1 ± 2.2 0.5 ± 0.3 
Fragilariopsis cylindrus µ (10) 5.0 ± 1.5 1.8 ± 0.5 1.4 ± 0.4 5.2 ± 0.4 5.3 ± 0.2 
Ce (10) 7.2 ± 0.5 0.1 ± 0.0 0.0 ± 0.0 5.7 ± 1.9 0.3 ± 0.1 
Synedropsis hyperborea strain 1 µ (10) 5.6 ± 0.0 >8.7 >8.7 >8.7 >8.7 
Ce (10) 6.9 ± 0.6 0.3 ± 0.3 2.5 ± 1.4 >8.7 8.5 ± 0.5 
S. hyperborea strain 2 µ (10) 4.4 ± 0.0 >8.7 >8.7 >8.7 >8.7 
Ce (10) 3.5 ± 0.6 4.4 ± 0.9 0.9 ± 0.5 5.3 ± 0.6 7.1 ± 0.4 
F. cylindrus µ (50) >8.7 8.0 ± 0.1 >8.7 8.4 ± 0.0 7.9 ± 0.0 
Ce (50) 8.4 ± 0.3 7.1 ± 0.1 >8.7 8.1 ± 0.1 7.6 ± 0.1 
IsolateMeasurea (% inhibited)Inhibitory Concentrationsb (mg L−1) for 5 Irradiances (µmol photons m−2 s−1)
3102050125
Attheya septentrionalis µ (10) >8.7 >8.7 8.2 ± 0.6 >8.7 3.4 ± 0.3 
Ce (10) >8.7 >8.7 >8.7 2.1 ± 2.2 0.5 ± 0.3 
Fragilariopsis cylindrus µ (10) 5.0 ± 1.5 1.8 ± 0.5 1.4 ± 0.4 5.2 ± 0.4 5.3 ± 0.2 
Ce (10) 7.2 ± 0.5 0.1 ± 0.0 0.0 ± 0.0 5.7 ± 1.9 0.3 ± 0.1 
Synedropsis hyperborea strain 1 µ (10) 5.6 ± 0.0 >8.7 >8.7 >8.7 >8.7 
Ce (10) 6.9 ± 0.6 0.3 ± 0.3 2.5 ± 1.4 >8.7 8.5 ± 0.5 
S. hyperborea strain 2 µ (10) 4.4 ± 0.0 >8.7 >8.7 >8.7 >8.7 
Ce (10) 3.5 ± 0.6 4.4 ± 0.9 0.9 ± 0.5 5.3 ± 0.6 7.1 ± 0.4 
F. cylindrus µ (50) >8.7 8.0 ± 0.1 >8.7 8.4 ± 0.0 7.9 ± 0.0 
Ce (50) 8.4 ± 0.3 7.1 ± 0.1 >8.7 8.1 ± 0.1 7.6 ± 0.1 

aGrowth rate (µ) and maximum cell concentration (Ce).

bEstimated threshold value (>) or mean ± SE (n = 96), where errors were calculated as the propagation of error based on the other model parameter estimates determined during model fit.

While the isolates had varying growth rate responses to ANS crude oil WAF, we observed that maximum cell concentrations also differed between strains and treatments (Figure 4, Table 2). Not all growth curves reached a clear peak, with subsequent decline, by the end of the experiment (Figure S2). The dose-response curves for maximum cell concentration generally exhibited higher sensitivity to ANS crude oil WAF relative to those calculated for growth rate, as reflected by lower IC10 concentrations (Table 2). The maximum cell concentrations of Attheya septentrionalis were the least sensitive to ANS crude oil contamination (Figure 4A); A. septentrionalis had a slight, but significant increase in maximum cell concentration with increasing TPH at irradiances below 50 µmol photons m−2 s−1, and a decrease in maximum cell concentration at higher irradiances. The IC10 values for maximum cell concentration of A. septentrionalis were relatively low and did not differ significantly at irradiance levels of 50 and 125 µmol photons m−2 s−1 (Table 2, Figure S7). The maximum cell concentrations of Fragilariopsis cylindrus were strongly impacted by TPH exposure (Figure 4), particularly at the intermediate and highest irradiances. Again, F. cylindrus was the only isolate with measurable inhibition by 50% (IC50; Figure S7, Table 2) over the range of TPH. IC10 values for the maximum cell concentration of F. cylindrus differed significantly between two groups of irradiance levels; the IC10 at irradiances of 3 and 50 µmol photons m−2 s−1 were significantly higher (p < 0.05) than for 10, 20, and 125 µmol photons m−2 s−1 (Table 2, Figure S7). The IC50 for F. cylindrus mirrored the pattern of IC10 grouping and significance, with the exclusion of irradiance of 20 µmol photons m−2 s−1 which fell outside the range of tested TPH concentrations (Table 2; Figure S7). Maximum cell concentrations varied between strains of Synedropsis hyperborea, with strain 2 growing to approximately half the concentration of strain 1. Although growth rates were relatively constant or even increasing with increasing TPH concentrations for S. hyperborea (Figure 3C and D), maximum cell concentration declined at all irradiance levels for both strains (Figure 4C and D). Nearly all irradiance levels had an IC10 that fell within the range of tested TPH concentrations for both strains of S. hyperborea, reflecting the increased sensitivity of maximum cell concentration relative to growth rate (Table 2).

Figure 4.

Dose-response curves of maximum cell concentration versus crude oil as modified by irradiance. (A–D) Maximum cell concentration (cells × 105 mL−1) and (E–H) relative cell concentration, proportional to the control (no oil), of the four isolates Attheya septentrionalis (A and E), Fragilariopsis cylindrus (B and F), Synedropsis hyperborea strain 1 (C and G), and Synedropsis hyperborea strain 2 (D and H) at irradiances of 3 (blue), 10 (purple), 20 (maroon), 50 (orange), 125 (yellow) µmols photons m−2 s−1. Data points are means; error bars, ± SE (n = 8). The shaded ribbon is the 95% confidence interval of the model fit. See Table S2 for model types. Solid black line is the 1.00 relative cell concentration isoline.

Figure 4.

Dose-response curves of maximum cell concentration versus crude oil as modified by irradiance. (A–D) Maximum cell concentration (cells × 105 mL−1) and (E–H) relative cell concentration, proportional to the control (no oil), of the four isolates Attheya septentrionalis (A and E), Fragilariopsis cylindrus (B and F), Synedropsis hyperborea strain 1 (C and G), and Synedropsis hyperborea strain 2 (D and H) at irradiances of 3 (blue), 10 (purple), 20 (maroon), 50 (orange), 125 (yellow) µmols photons m−2 s−1. Data points are means; error bars, ± SE (n = 8). The shaded ribbon is the 95% confidence interval of the model fit. See Table S2 for model types. Solid black line is the 1.00 relative cell concentration isoline.

Close modal

3.3. Modeling the impact of a crude oil spill on an under-ice bloom

To better understand how a crude oil spill might alter the diversity and biomass of sea-ice algae in the Arctic, we incorporated the dose-response data for these strains into a simple numerical model to simulate the impact of an oil spill prior to the spring bloom. The model was forced with measurements of daily irradiance and sea-ice thickness from the Alaskan Arctic coast, and the under-ice irradiances were calculated assuming three fixed snow depths (Figure S4). The seasonal increase of incident irradiance was similar between the two years of in situ data (Figure S4A and B), so a single daily irradiance value was used to force the model (Figure S4C). The mean daily irradiance ranged from 30% to 89% of the daily maximum incident irradiance (Figure S4A–C). The ice thickness data that was used to force this model reached a maximum thickness of 1.37 m in late April (Figure S4E). Calculated under-ice irradiances reached maximum values of approximately 28, 9, and 3 µmol photons m−2 s−1 under snow depths of 0, 0.1, and 0.2 m, respectively (Figure S4D). The impacts of two oil conditions on sea-ice algal production were evaluated: no oil contamination (0.0 mg L−1 of TPH) and oil contamination equal to the highest tested concentration from this study (8.7 mg L−1 of TPH). Results from the model show that snow depth and crude oil exposure altered simulated community composition and biomass (Figure 5). The relative abundances were distributed more evenly under moderate-to-high snow cover in the absence of crude oil contamination (Figure 5A). Under low-snow conditions, without oil contamination, Fragilariopsis cylindrus dominated the community (Figure 5A), but remained low in the presence of oil (<5%; Figure 5B). Predicted biomass declined with increasing snow depth and in the presence of oil (Figure 5C–E).

Figure 5.

Modeled sea-ice diatom composition and biomass modified by snow depth and crude oil. Relative composition of sea-ice diatoms Attheya septentrionalis (dark gray), Fragilariopsis cylindrus (orange), Synedropsis hyperborea strain 1 (blue), and Synedropsis hyperborea strain 2 (red) as modeled at the end of a sea-ice algal bloom (A) without oil (0 mg L−1) and (B) with oil (8.7 mg L−1) as a static crude oil exposure for a snow depth range of 0–0.2 m. Simulated biomass accumulation under three snow depths (C) 0.0 m, (D) 0.1 m, and (E) 0.2 m at the end of the model run at two different total petroleum hydrocarbon (TPH) concentrations.

Figure 5.

Modeled sea-ice diatom composition and biomass modified by snow depth and crude oil. Relative composition of sea-ice diatoms Attheya septentrionalis (dark gray), Fragilariopsis cylindrus (orange), Synedropsis hyperborea strain 1 (blue), and Synedropsis hyperborea strain 2 (red) as modeled at the end of a sea-ice algal bloom (A) without oil (0 mg L−1) and (B) with oil (8.7 mg L−1) as a static crude oil exposure for a snow depth range of 0–0.2 m. Simulated biomass accumulation under three snow depths (C) 0.0 m, (D) 0.1 m, and (E) 0.2 m at the end of the model run at two different total petroleum hydrocarbon (TPH) concentrations.

Close modal

This study focused on exploring the potential consequences of dual anthropogenic stressors on sea-ice diatoms in the Alaskan Arctic. The combination of two anthropogenic drivers, increasing irradiance from declining snow/ice thickness and crude oil spills, are increasingly likely with a warming Arctic Ocean (Webster et al., 2014; Veyssière et al., 2022). To better understand the impacts of these anthropogenic changes on the sea-ice ecosystem in the Alaskan Arctic, we isolated four diatoms from landfast ice near Utqiaġvik, Alaska, and characterized their responses to a fully factorial gradient of irradiance and crude oil contaminated conditions. The four sea-ice diatoms were freshly isolated for this study to minimize the evolutionary drift that can occur in microalgae over time (Schaum et al., 2013) and were chosen to represent ecologically relevant species with a range of physiological and morphological traits (Figure 1).

4.1. Growth-irradiance curves recapitulate sea-ice diatom niche

Sea-ice algae are exposed to extreme shifts in light intensity, spectrum, and duration, ranging from seasonal variations (e.g., changes in day length and ice thickness) to diel cycles related to solar and lunar zenith angle, to rapid meteorologically driven changes (e.g., cloud cover and precipitation; Johnsen et al., 2021). Photoacclimation is the result of physiological changes that compensate for these shifts in irradiance based on genetic adaptations (Dubinsky and Stambler, 2009). To determine the photophysiology of the four sea-ice diatom isolates, we acclimated cultures to five irradiances and determined their exponential growth rates (Figure 2). We chose ecologically relevant irradiances for our experiment, 3–125 µmol photons m−2 s−1, where the lower end of this range corresponds to returning light near the end of winter and to late spring irradiances under high snow cover (Campbell et al., 2018), while the highest irradiances tested are comparable to bare ice and open water in late spring (Massicotte et al., 2020).

The maximum growth rate of Attheya septentrionalis determined in this study (approximately 0.40 d−1; Figure 2A) was similar to that reported by Hegseth (1992; approximately 0.37 d−1 at −0.5°C) but far greater than reported by Lomas et al. (2019; 0.03 d−1 at 2°C) at an irradiance of 20 µmol photons m−2 s−1; these are the only previously reported growth rates for this species. Lomas et al. (2019) used axenic cultures, which may explain in part the relatively low growth rates recorded for all eleven species tested in that study, as diatoms are known to have beneficial interactions with heterotrophic bacteria (Amin et al., 2012). Our estimates of the maximum growth rate of Fragilariopsis cylindrus (approximately 0.5 d−1; Figure 2B) were similar to those from previous research using Arctic strains (Pančić et al., 2015), but well exceeded those typically reported for Antarctic strains (<approximately 0.31 d−1; Mock and Hoch, 2005; Arrigo et al., 2010; Jabre and Bertrand, 2020; Croteau et al., 2022), with the exception of Sommer (1989) who reported a specific growth rate of 0.69 d−1. We suggest that the vast geographic separation of this bipolar species, or the age of the culture, yield conclusions that differ from more recently collected, regionally specific isolates. To our knowledge, the growth rates reported here are the first to be published for Synedropsis hyperborea; they fall within a typical range of growth rates for Arctic sea-ice diatoms (Lacour et al., 2017).

Arctic diatoms exploit environmental niches, partly defined by light availability, through differences in their photophysiology, for example, the sea-ice environment experiences overall low light levels with low amplitude changes relative to open water (Croteau et al., 2021). Sympagic algae that are well adapted to their environment have a high a and low ek, demonstrating the capacity to utilize photons efficiently to support growth at incredibly low irradiances, that is, 0.17 µmol photons m−2 s−1 (Hancke et al., 2018). In contrast, planktonic taxa have a lower a and higher ek, which shifts their niche preference toward higher light conditions and is generally coupled with mechanisms that allow them to grow uninhibited by supersaturating irradiances, for example, non-photochemical quenching (Croteau et al., 2021). Cryopelagic growth-irradiance parameters can be expected to fall between these two extremes, balancing the contrasting irradiance niches. Fragilariopsis cylindrus is a cryopelagic taxon (Poulin et al., 2011; Hop et al., 2020; Croteau et al., 2022), while Attheya septentrionalis and Synedropsis hyperborea are considered to be sympagic and are often found attached to strands of Melosira arctica or colonies of Nitzschia fridida as epiphytes (Poulin et al., 2011; Poulin et al., 2014; Campbell et al., 2018). We found a range of growth-irradiance parameters, suggesting that all of our strains are adapted to growth in low-light conditions compared to other polar phytoplankton, with ek falling into the lowest quartile out of 18 polar diatoms compiled by Lacour et al. (2017). Of the species tested, F. cylindrus was the only cryopelagic taxon and was distinct from A. septentrionalis and S. hyperborea, with the lowest a and highest ek (Table 1). Different ice types and features (e.g., multi-year vs first year ice, ridges, melt ponds, sediment inclusions) increase heterogeneity of light transmittance through ice and may act as an environmental filter that influences community structure from the bottom up. Knowledge of growth-irradiance parameters for particular taxa may therefore improve our understanding of their fine-scale distribution as it relates to irradiance microhabitats.

For the Arctic strain of Fragilariopsis cylindrus used in this study, the onset of photoinhibition occurred at a much higher irradiance, 100 µmol photons m−2 s−1, than in previous work using an Antarctic strain of F. cylindrus, <50 µmol photons m−2 s−1 (Arrigo et al., 2010; Croteau et al., 2022). This two-fold difference is notable because much of the work being conducted on F. cylindrus utilizes the model strain (CCMP1102), which was collected from the Antarctic but is being used for Arctic-focused studies. For example, Croteau et al. (2022) presents a detailed study using sea-ice, cryopelagic and planktonic microalgae, but their use of the Antarctic strain of F. cylindrus (CCMP1102) demonstrated photoinhibition and a maximum growth rate more similar to the sea-ice algal endmember, Nitzschia frigida, used in their study. Using Arctic-specific isolates of F. cylindrus may better explain its importance as part of under-ice algal blooms in the Arctic (Arrigo et al., 2012) due to a higher maximum growth rate and photoinhibition occurring only at a higher irradiance, as measured in this study. Growth-irradiance data from Hegseth (1992) showed no signs of photoinhibition for Attheya septentrionalis, measured to an irradiance of 400 µmol photons m−2 s−1. Our limited irradiance curves, ending at 125 µmol photons m−2 s−1, present an important caveat, as they introduce uncertainty surrounding the degree of photoinhibition incurred. Attheya spp. have been shown to accumulate rapidly in sea ice with low snow cover, dominating relative abundance and outcompeting pennate diatoms like Nitzschia frigida (Campbell et al., 2018), which suggests that they may occupy a high-light niche within sea-ice associated communities. Together, these data suggest that A. septentrionalis and F. cylindrus are better adapted to an Arctic with thin or bare sea ice than species that are photoinhibited at such high light levels like Synedropsis hyperborea (this study) and N. frigida (Hegseth, 1992; Croteau et al., 2022).

4.2. Chemical composition of WAF from Alaska North Slope crude oil

The relative toxicity of crude oil depends on its chemical composition and concentration, as well as environmental conditions, such as temperature which impacts solubility and viscosity (Atlas, 1981; Das and Chandran, 2011). Variability in source oil composition and WAF preparation methodology, for example, temperature, mixing energy, chemical dispersants, light, and salinity, influence the concentration and composition of the resulting WAF of crude oil. Utilizing the source oil most likely to contaminate a region of interest, as well as preparing the WAF and toxicity tests under environmentally relevant conditions, are therefore important. We prepared our WAF using ANS crude oil at a temperature of −1°C to closely approximate the source oil and conditions under which we might expect an oil spill in the Alaskan Arctic, that is, in the Chukchi or Beaufort seas. The freezing point of seawater, at approximately −1.8°C, is the temperature at the ice–water interface where much of the ice-algal biomass is concentrated. The choice of a slightly warmer temperature was necessary to avoid freeze-thaw cycles due to the small liquid volumes used in the microwell plates and the difficulty in maintaining such a precise temperature in the incubator. Concentrations of TPH generated in our WAF were similar to ranges generated using similar conditions and source oil (Gardiner et al., 2013).

The chemical composition of WAF generated in this study was dominated by short-chain hydrocarbons that can be lost due to volatilization (Das and Chandran, 2011). Crude oil, or contaminated seawater, under sea ice can be expected to experience less volatilization due to low temperatures (Atlas, 1975) and the presence of ice, which inhibits gas exchange (Rutgers van der Loeff et al., 2014). The presence of this lighter fraction of crude oil has been shown to increase toxicity and inhibit biodegradation (Atlas, 1975; Chen et al., 2008; Rodriguez Martinez et al., 2008; Sherry et al., 2014). Our study utilized microwell plates sealed with parafilm to reduce water loss due to evaporation but the relatively large headspace may have allowed some volatilization of crude oil compounds with additional loss by adsorption to the container walls (Berrojalbiz et al., 2011), potentially reducing the effective concentrations during the experiment. Recent studies have also shown considerable biodegradation of petroleum hydrocarbons using Arctic sea-ice, sub-ice, and open-water microbial communities, ranging from 48% to 94% (McFarlin et al., 2014; Garneau et al., 2016; Bakke et al., 2021). Our sea-ice diatom isolates were non-axenic and therefore had associated communities of heterotrophic bacteria which are thought to be the primary source of biodegradation. The anticipated decline in TPH during the study could result in culture recovery, such as was observed by Kamalanathan et al. (2021); however, Garneau et al. (2016) found that sea-ice microbial communities degrade oil much more slowly than the sub-ice community under identical conditions. We did not measure hydrocarbon loss during our experiments due to the small sample volumes, but both biotic and abiotic losses of TPH likely occurred. Therefore, our study presents potentially low estimates of the impacts of ANS crude oil on the sea-ice diatom isolates investigated. Despite this potential underestimation, we observed substantial impacts on microalgal growth.

4.3. Dose response of sea-ice diatoms to ANS crude oil

We used exponential growth rate and maximum cell concentrations of microalgae, which integrate cellular physiology (Sunda and Huntsman, 1995), to provide robust indicators of sublethal crude-oil toxicity. A common approach for investigating toxicological impacts on microalgae is to measure and analyze response variables at single time points (Gilde and Pinckney, 2012; Dilliplaine et al., 2021; Putzeys et al., 2022); a major drawback of this approach is that the conclusions may be time-dependent and conclusions will vary based on the termination date. We used a combination of methods to assess toxicological impacts without a pre-determined termination date. The maximum growth rate was determined by finding the maximum exponential growth over the log-linear portion of the growth curve; the maximum cell concentration was determined from the whole growth curve and occurred after several days to several weeks. Factors that can influence maximum cell concentration include changes in cell nutrient quotas, self-shading, death, and toxicological inhibition of growth. For example, cellular repair from oil-induced damages may increase nutrient quotas leading to a decrease in maximum cell concentration independent of the growth rate. Growth rate and maximum cell concentration therefore represent distinct measures providing information regarding the toxicological impact of crude oil on sea-ice algae. Changes to growth rate and cell concentration can be used to infer how an oil spill may alter the timing and magnitude of the vernal ice algal bloom. We hypothesized that declines in growth rate and maximum cell concentration would be greatest for treatments combining the highest irradiance level and most concentrated crude oil WAF due to the combined stress. The sea-ice diatoms used in this study had a range of crude-oil sensitivities and a complex interaction with irradiance levels that did not always support this hypothesis. In our study, Fragilariopsis cylindrus was the most sensitive isolate assayed, with calculable values for IC10 of the growth rate and cell concentrations for all treatments, and it was the only isolate with measurable values for IC50 (Table 2). IC10 and IC50 are measures used for reporting toxicity data, with the former signifying sensitivity (Beasley et al., 2015). These findings contrasted with the other isolates, Attheya septentrionalis and both strains of Synedropsis hyperborea, which had relatively high tolerances to crude oil in our study with few measurable values for IC10 (Table 2). This variability in sensitivity is similar to previous studies. Özhan et al. (2014) measured near-complete growth inhibition in 5 phytoplankton taxa by a concentration of 10 mg L−1 TPH exposed to Louisiana sweet crude, in contrast to Bretherton et al. (2020) who measured some taxa showing stimulation to growth exceeding 20 mg L−1 TPH of chemically dispersed Macondo crude oil. Our experiments were conducted simultaneously with the same treatment conditions and therefore represent considerable taxon-specific variability.

The mechanisms of crude oil growth inhibition/stimulation are poorly understood despite the widespread observance of these responses. Oxidative stress by the formation of reactive oxygen species is produced during normal metabolic activity but can cause extensive damage and mortality when antioxidant defenses are overwhelmed (Hansel and Diaz, 2021). Ozhan et al. (2015) reported oxidative stress induced by crude oil exposure and large differences in antioxidant efficiency between two phytoplankton taxa, but the role of oxidative stress in growth inhibition remained unknown. Kamalanathan et al. (2021) suggested that oxidative damage is secondary to growth reduction from the damaged light-harvesting complex and electron transport proteins. Differences in light-harvesting complex proteins or susceptibility of antioxidants to disruption by crude oil compounds across species could explain the range of sensitivities displayed here, as well as interactive effects or lack thereof, between irradiance and WAF.

Excessive irradiance alone can impact growth negatively (Section 4.1); we expected to observe the most significant inhibition at the highest irradiance level tested due to an additive or multiplicative effect from combined stressors. Attheya septentrionalis had a significantly lower growth rate IC10 at the highest light level (Table 2, Figure S6), indicating higher sensitivity and supporting our hypothesis. We were unable to determine whether oxidative stress played a role in this enhanced inhibition. We expected the greatest combinatorial effect to be exhibited by the Synedropsis hyperborea isolates because they exhibited photoinhibition in the absence of oil (Figure 2C and D). In contrast, the IC10 of growth rate for S. hyperborea strains indicated that these diatoms were relatively insensitive to crude oil WAF (Figure 3C and D) at all light levels and appeared to be more sensitive to ANS crude oil contamination at the lowest light level tested (Table 2, Figure 3C, D, G and H; Figure S6). This finding directly refutes our hypothesis that combined stressors would enhance growth rate inhibition. Maximum cell concentration of S. hyperborea declined at all irradiance levels, indicating a greater susceptibility to crude oil exposure than indicated by growth rate alone (Figure 4C, D, G and H), possibly indicating a greater metabolic burden in the presence of oil leading to increased nutrient demand. Interestingly, an increase in the maximum cell concentration was observed for A. septentrionalis, with increasing oil for irradiance levels of 3–20 µmol photons m−2 s−1, in spite of declines in growth rate (Figures 3A, E and 4A, E). While not measured directly in the oiled treatments, a decline in cell size or nutrient quota possibly supported a greater cell concentration at the end of the experiment. Mutualistic interactions between microalgae and bacteria might also explain this discrepancy. Bacteria are known to provide microalgae with iron (Amin et al., 2009), vitamin B12 (Croft et al., 2005), CO2, and protection from reactive oxygen species (Hünken et al., 2008), while microalgae provide bacteria with organic carbon and oxygen. Additionally, the bacterial degradation of crude oil results in the production of sugars, lipids, and amino acids that may stimulate growth of microalgae (Goutx et al., 1984). Differences between the bacterial community composition, activity, or abundance could lead to different nutritional availability for microalgal growth leading to the observed differences in maximum cell concentration. Microbial interactions and/or changes in cell nutrient quota may be responsible for the observed discrepancies in A. septentrionalis growth rate and cell concentration data, but were not determined and are beyond the scope of this study.

The mechanisms of crude-oil-induced damage proposed by Kamalanathan et al. (2021) include large scale damage to light-harvesting complex proteins, inhibiting photosynthesis, along with impacts on downstream processes such as the ability to catabolize organic carbon. The greater inhibition of relative growth rate by Synedropsis hyperborea strains at the lowest light level could indicate subtle damage to the light-harvesting complex or catabolic pathway, leading to energy limitation in the presence of oil that could be overcome with greater irradiances. The values for IC10 of Fragilariopsis cylindrus growth indicated the greatest sensitivity at the intermediate irradiances (10 and 20 µmol photons m−2 s−1; Table 2), though they did not differ significantly from the low irradiance (3 µmol photons m−2 s−1; Table 2), and may indicate an energy-dependent growth disruption alleviated at even higher irradiances. The cellular stress response counteracts macromolecular damage, regardless of the stressor, by induction of proteins such as molecular chaperones and removal of macromolecular debris through ubiquitination, and so on, and can convey cross-tolerance to different stressors (Kültz, 2003). Alternatively, deployment of cellular stress response mechanisms by S. hyperborea in the presence of excessive irradiance could stress-harden the taxon preventing combinatorial inhibitory effects. However, threshold-based sensitivity, that is, IC10, provides a limited view of the dose response, especially when applied to biphasic curves such as those of F. cylindrus (Stebbing, 2009).

The shape of dose-response curves also differed by taxa, revealing differences in their underlying physiological responses to crude oil exposure. Monophasic dose responses, like the common sigmoidal log-logistic regression, have a single region of decline in response to a toxic compound. Prior to the decline, a homeodynamic response can compensate for low doses of a toxic compound until a critical threshold is reached (Stebbing, 1981; 2009). Multi-phasic models include the capacity to capture multiple steps in a curve resulting from multiple homeodynamic responses with individual thresholds, or the overemployment of a homeodynamic response that can result in the stimulation of the measured variable (hormesis; Stebbing, 2009). The growth rate and maximum cell concentration of Attheya septentrionalis and cell concentration of Synedropsis hyperborea (Table S2) were best fit with monophasic dose-response curves, suggesting that a single homeodynamic response was overwhelmed over this range of crude oil exposure. In contrast, the dose-response curves of Fragilariopsis cylindrus showed two distinct inhibitory phases, an initial minor inhibition with a plateau phase followed by a steep tipping point at approximately 6 mg L−1 (Figures 3B and 4B). The expectation that cryopelagic taxa, particularly F. cylindrus, will play a greater role in a brighter Arctic may render the microalgal communities here more susceptible to crude oil exposure (Lannuzel et al., 2020). This bi-phasic dose response of F. cylindrus suggests at least two distinct inhibitory mechanisms or homeodynamic responses (Di Veroli et al., 2015). We hypothesize that some cellular function of F. cylindrus is deployed (e.g., repair mechanisms as part of the cellular stress response) by low levels of TPH, partially countering inhibitory effects, until damage becomes excessive and homeostasis is lost.

Some dose-response curves for Synedropsis hyperborea strains 1 and 2 displayed a hormetic response; that is, a stimulation of growth at low concentrations of TPH (Figure 3C, D). According to Stebbing (1981), hormesis is a common characteristic of dose-response curves, and may exceed 200% relative to the control (Agathokleous and Calabrese, 2020). Stimulation of the growth rate of S. hyperborea remained below 20% for both isolates. Previous studies have found microalgal growth to be stimulated to approximately 1 mg L−l of crude oil concentration, with inhibition occurring thereafter (Huang et al., 2011; Ozhan et al., 2014), while some studies have seen stimulation to growth in excess of 20 mg L−1 (Bretherton et al., 2020). Stimulation of S. hyperborea growth rate generally peaked around 0.8–1.6 mg L−1 TPH but remained higher than the control for all concentrations at irradiances of 10 µmol photons m−2 s−1 (strains 1 and 2) and 20 µmol photons m−2 s−1 (strain 2; Figure 3C, D, G, and H). There is no clear indication as to why intermediate irradiances were the most resistant to inhibition for this species nor has a mechanism for light-mediated toxicity tolerance been described, but S. hyperborea may possibly employ an energy-dependent strategy in response to crude oil exposure. No hormetic response was observed for maximum cell concentration in dose-response curves of S. hyperborea (Figure 4C and D); instead, there was a general decline at all light levels. These findings suggest that increased growth rate, stimulated by crude oil exposure, may require a larger nutrient cell quota resulting in a quicker nutrient limitation. This combination of crude oil tolerance for growth rate at light levels common under sea ice suggests that S. hyperborea may be positioned physiologically to experience little interruption after exposure to crude oil when nutrients are replete, but may be more sensitive to sublethal inhibition during winter to early spring when light is restricted and during the spring bloom when nutrients may become limiting. Further research is required to determine the mechanisms behind the various phases of toxicity for these taxa and which cellular processes are disrupted or deployed to counter toxicity-induced cellular stress.

4.4. Model

We used a simple numerical model to simulate the biomass accumulation and relative compositional change of our isolates in sea ice under varying snow depths, with and without oil, to contextualize our results over a spring growth season in the Alaskan Arctic (Figure 5). The light environment was forced with seasonal irradiance and ice thickness data from Utqiaġvik, with variable snow depth; attenuation by snow is much greater than ice and is the controlling factor for transmitted irradiance (Figure S4). In the absence of oil, the diatom community shifts in response to snow depth from dominance by Fragilariopsis cylindrus to Attheya septentrionalis to Synedropsis hyperborea with increasing snow depth (Figure 5A). At the end of the simulation, biomass declined with increasing snow depth and was always lower when exposed to oil at the same snow depth (Figure 5C–E). The impact of oil on community structure and biomass was largest in the zero-snow scenario (Figure 5B and C). For example, F. cylindrus was nearly entirely absent from the community exposed to oil, except at the deepest snow depth where irradiance was also limiting to the other isolates. This model result shows the relative sensitivity of F. cylindrus to crude oil exposure and highlights the threat of a crude oil spill to a key Arctic species.

This model has several important caveats and limitations. For example, the model does not take into account nutrient limitation, changes in salinity, or radiative absorption. Attheya spp. have been shown to outcompete pennate diatoms at higher irradiances when combined with low-salinity and low-nitrogen conditions (Campbell et al., 2018). Nutrient limitation could also impact the final biomass, as a previous study found lower biomass under bare ice in the Canadian Arctic (Campbell et al., 2015). However, in a more locally relevant study in the Alaskan Arctic, removal of thin snow layers (4–5 cm) during peak bloom stimulated algal growth in landfast sea ice from Utqiaġvik, supporting the assumption in our model that nutrients are not limiting in the spring (Juhl and Krembs, 2010). This model is also limited in considering only four strains of sea-ice algae. We expect the true sea-ice algal community to be more diverse, as over 1000 species of algae have been documented in sea ice (Poulin et al., 2011); for example, Synedropsis hyperborea has not been observed to dominate the sea-ice community even under thick snow cover. The oil concentration of 8.7 mg L−1 TPH in our model, and our experiments, was a static exposure. Oil concentrations are likely to decline over time as dilution and or biodegradation occur, but sustained high concentrations of crude oil may be encountered where brine is in contact with oil or when continuously resupplied as during a blowout event. A worst-case Arctic oil spill, as a well blowout, could affect as much as 1.1 million km2 with high concentrations of oil, representing a significant threat to sea-ice microalgal biomass (Blanken et al., 2017), comparable to our model assumptions.

Our study highlights the importance of toxicological studies that use high-resolution doses in order to determine responses and thresholds at low concentrations. We found evidence of stimulation and inhibition that was dependent on sea-ice algal species, oil concentration, and irradiance level. The interaction of crude oil exposure and irradiance was complex and not generalizable across all taxa. Growth rate and maximum cell concentration represent useful measurements because they integrate the physiological responses of an organism and are straightforward to measure, making them ideal candidates for high-throughput assays. The reduction of growth rate and maximum cell concentration may delay the timing and magnitude, respectively, of a spring sea-ice algal bloom. However, growth rate and maximum cell concentration alone do not provide a mechanistic understanding of crude oil stimulation/inhibition, which is required to make broader predictions for unmeasured taxa. We provide evidence that the toxicological impact of crude oil exposure on the sea-ice algal community could shift community structure and inhibit biomass accumulation. The degree of growth inhibition is species-specific; of our isolates tested, the abundant and widespread cryopelagic taxon, Fragilariopsis cylindrus, was the most impacted. The predicted increase in the relative importance of F. cylindrus in a brighter Arctic may make the Arctic ecosystem more prone to disruption from oil spills in the future.

The following datasets were generated:

The supplemental files for this article can be found as follows:

Figures S1–S7. Tables S1–S2. Text S1. Docx

The authors thank the Ukpeaġvik Iñupiat Corporation (UIC) Science for their logistical support in Utqiaġvik during field work and willingness to core and ship ice during a global pandemic. The authors thank Jim Janoso for his microscopy expertise and the two anonymous reviewers whose comments and suggestions improved this manuscript.

This project was funded by a research grant provided through the Coastal Marine Institute (CMI M20AC10007), a joint institute between the Bureau of Ocean Energy Management (BOEM) and the University of Alaska Fairbanks. Further support and match funding was provided by the College of Fisheries and Ocean Sciences at the University of Alaska Fairbanks. Micrographs were captured at the UAF Molecular Imaging Facility, NIGMS P20GM130443.

The authors indicate that they have no known competing interests.

  • Contributed to conception and design: KD, GH.

  • Contributed to acquisition of data: KD.

  • Contributed to analysis and interpretation of data: KD, GH.

  • Drafted and/or revised the article: KD, GH.

  • Approved the submitted version for publication: KD, GH.

Agathokleous
,
E
,
Calabrese
,
EJ.
2020
.
A global environmental health perspective and optimisation of stress
.
Science of The Total Environment
704
:
135263
. DOI: http://dx.doi.org/10.1016/j.scitotenv.2019.135263.
Alaska Department of Environmental Conservation
.
2002
a.
Method ak 102
.
Available at
https://dec.alaska.gov/media/13473/ak-10-2.pdf.
Accessed October 20, 2020
.
Alaska Department of Environmental Conservation
.
2002
b.
Method ak 103
.
Available at
https://dec.alaska.gov/media/13474/ad-10-03.pdf.
Accessed October 20, 2020
.
Amin
,
SA
,
Green
,
DH
,
Hart
,
MC
,
Küpper
,
FC
,
Sunda
,
WG
,
Carrano
,
CJ.
2009
.
Photolysis of iron–siderophore chelates promotes bacterial–algal mutualism
.
Proceedings of the National Academy of Sciences
106
(
40
):
17071
17076
. DOI: http://dx.doi.org/10.1073/pnas.0905512106.
Amin
,
SA
,
Parker
,
MS
,
Armbrust
,
EV.
2012
.
Interactions between diatoms and bacteria
.
Microbiology and Molecular Biology Reviews
76
(
3
):
667
684
. DOI: http://dx.doi.org/10.1128/MMBR.00007-12.
Arctic Council
.
2009
.
Arctic Marine Shipping Assessment 2009 Report
.
Available at
https://www.pmel.noaa.gov/arctic-zone/detect/documents/AMSA_2009_Report_2nd_print.pdf.
Accessed December 19, 2018
.
Arctic Monitoring and Assessment Programme
.
2007
.
Arctic oil and gas
.
Oslo, Norway
:
Arctic Monitoring and Assessment Programme
(
vol. 1–2
).
Ardyna
,
M
,
Mundy
,
CJ
,
Mills
,
MM
,
Oziel
,
L
,
Grondin
,
P-L
,
Lacour
,
L
,
Verin
,
G
,
van Dijken
,
G
,
Ras
,
J
,
Alou-Font
,
E
,
Babin
,
M
,
Gosselin
,
M
,
Tremblay
,
J-É
,
Raimbault
,
P
,
Assmy
,
P
,
Nicolaus
,
M
,
Claustre
,
H
,
Arrigo
,
KR.
2020
.
Environmental drivers of under-ice phytoplankton bloom dynamics in the Arctic Ocean
.
Elementa: Science of the Anthropocene
8
:
30
. DOI: http://dx.doi.org/10.1525/elementa.430.
Arrigo
,
KR
,
Mills
,
MM
,
Kropuenske
,
LR
,
van Dijken
,
GL
,
Alderkamp
,
A-C
,
Robinson
,
DH.
2010
.
Photophysiology in two major southern ocean phytoplankton taxa: Photosynthesis and growth of Phaeocystis antarctica and Fragilariopsis cylindrus under different irradiance levels
.
Integrative and Comparative Biology
50
(
6
):
950
966
. DOI: http://dx.doi.org/10.1093/icb/icq021.
Arrigo
,
KR
,
Perovich
,
DK
,
Pickart
,
RS
,
Brown
,
ZW
,
Van Dijken
,
GL
,
Lowry
,
KE
,
Mills
,
MM
,
Palmer
,
MA
,
Balch
,
WM
,
Bahr
,
F
,
Bates
,
NR
,
Benitez-Nelson
,
C
,
Bowler
,
B
,
Brownlee
,
E
,
Ehn
,
JK
,
Frey
,
KE
,
Garley
,
R
,
Laney
,
SR
,
Lubelczyk
,
L
,
Mathis
,
J
,
Matsuoka
,
A
,
Mitchell
,
BG
,
Moore
,
GWK
,
Ortega-Retuerta
,
E
,
Pal
,
S
,
Polashenski
,
C
,
Reynolds
,
RA
,
Schieber
,
B
,
Sosik
,
HM
,
Stephens
,
M
,
Swift
,
JH.
2012
.
Massive phytoplankton blooms under Arctic sea ice
.
Science
336
(
6087
):
1408
. DOI: http://dx.doi.org/10.1126/science.1215065.
Årthun
,
M
,
Onarheim
,
IH
,
Dörr
,
J
,
Eldevik
,
T.
2021
.
The seasonal and regional transition to an ice-free Arctic
.
Geophysical Research Letters
48
(
1
). DOI: http://dx.doi.org/10.1029/2020GL090825.
Atlas
,
RM.
1975
.
Effects of temperature and crude oil composition on petroleum biodegradation
.
Applied Microbiology
30
(
3
):
396
403
. DOI: http://dx.doi.org/10.1128/am.30.3.396-403.1975.
Atlas
,
RM.
1981
.
Microbial degradation of petroleum hydrocarbons: An environmental perspective
.
Microbiological Reviews
45
(
1
):
180
209
. DOI: http://dx.doi.org/10.1128/MMBR.45.1.180-209.1981.
Aurand
,
D
,
Coelho
,
G.
2005
. Cooperative aquatic toxicity testing of dispersed oil and the “chemical response to oil spills: Ecological effects research forum (CROSERF).”
Lusby, MD
:
Ecosystem Management & Associates, Inc
.
Technical Report no: 07-03, 105 pages + Appendices
.
Bakke
,
I
,
Greer
,
CW
,
Gunnar
,
O.
2021
.
Biodegradation of weathered crude oil by microbial communities in solid and melted sea ice
.
Marine Pollution Bulletin
172
(
August):
112823
. DOI: http://dx.doi.org/10.1016/j.marpolbul.2021.112823.
Beasley
,
A
,
Belanger
,
SE
,
Brill
,
JL
,
Otter
,
RR.
2015
.
Evaluation and comparison of the relationship between NOEC and EC10 or EC20 values in chronic Daphnia toxicity testing
.
Environmental Toxicology and Chemistry
34
(
10
):
2378
2384
. DOI: http://dx.doi.org/10.1002/etc.3086.
Bender
,
ML
,
Giebichenstein
,
J
,
Teisrud
,
RN
,
Laurent
,
J
,
Frantzen
,
M
,
Meador
,
JP
,
Sørensen
,
L
,
Hansen
,
BH
,
Reinardy
,
HC
,
Laurel
,
B
,
Nahrgang
,
J.
2021
.
Combined effects of crude oil exposure and warming on eggs and larvae of an Arctic forage fish
.
Scientific Reports
11
(
1
):
8410
. DOI: http://dx.doi.org/10.1038/s41598-021-87932-2.
Berge
,
J
,
Renaud
,
PE
,
Darnis
,
G
,
Cottier
,
F
,
Last
,
K
,
Gabrielsen
,
TM
,
Johnsen
,
G
,
Seuthe
,
L
,
Weslawski
,
JM
,
Leu
,
E
,
Moline
,
M
,
Nahrgang
,
J
,
Søreide
,
JE
,
Varpe
,
Ø
,
Lønne
,
OJ
,
Daase
,
M
,
Falk-Petersen
,
S.
2015
.
In the dark: A review of ecosystem processes during the Arctic polar night
.
Progress in Oceanography
139
:
258
271
. DOI: http://dx.doi.org/10.1016/j.pocean.2015.08.005.
Berrojalbiz
,
N
,
Dachs
,
J
,
Ojeda
,
MJ
,
Valle
,
MC
,
Castro-Jiménez
,
J
,
Wollgast
,
J
,
Ghiani
,
M
,
Hanke
,
G
,
Zaldivar
,
JM.
2011
.
Biogeochemical and physical controls on concentrations of polycyclic aromatic hydrocarbons in water and plankton of the Mediterranean and Black seas
.
Global Biogeochemical Cycles
25
(
4
):
1
14
. DOI: http://dx.doi.org/10.1029/2010GB003775.
Blanken
,
H
,
Tremblay
,
LB
,
Gaskin
,
S
,
Slavin
,
A.
2017
.
Modelling the long-term evolution of worst-case Arctic oil spills
.
Marine Pollution Bulletin
116
(
1–2
):
315
331
. DOI: http://dx.doi.org/10.1016/j.marpolbul.2016.12.070.
Boccadoro
,
C
,
Krolicka
,
A
,
Receveur
,
J
,
Aeppli
,
C
,
Le Floch
,
S.
2018
.
Microbial community response and migration of petroleum compounds during a sea-ice oil spill experiment in Svalbard
.
Marine Environmental Research
142
(
September
):
214
233
. DOI: http://dx.doi.org/10.1016/j.marenvres.2018.09.007.
Bretherton
,
L
,
Hillhouse
,
J
,
Kamalanathan
,
M
,
Finkel
,
ZV
,
Irwin
,
AJ
,
Quigg
,
A.
2020
.
Trait-dependent variability of the response of marine phytoplankton to oil and dispersant exposure
.
Marine Pollution Bulletin
153
(
January
):
110906
. DOI: http://dx.doi.org/10.1016/j.marpolbul.2020.110906.
Bretherton
,
L
,
Williams
,
A
,
Genzer
,
J
,
Hillhouse
,
J
,
Kamalanathan
,
M
,
Finkel
,
ZV
,
Quigg
,
A.
2018
.
Physiological response of 10 phytoplankton species exposed to macondo oil and the dispersant, corexit
.
Journal of Phycology
54
(
3
):
317
328
. DOI: http://dx.doi.org/10.1111/jpy.12625.
Campbell
,
K
,
Mundy
,
CJ
,
Barber
,
DG
,
Gosselin
,
M.
2015
.
Characterizing the sea ice algae chlorophyll a-snow depth relationship over Arctic spring melt using transmitted irradiance
.
Journal of Marine Systems
147
:
76
84
. DOI: http://dx.doi.org/10.1016/j.jmarsys.2014.01.008.
Campbell
,
K
,
Mundy
,
CJ
,
Belzile
,
C
,
Delaforge
,
A
,
Rysgaard
,
S.
2018
.
Seasonal dynamics of algal and bacterial communities in Arctic sea ice under variable snow cover
.
Polar Biology
41
(
1
):
41
58
. DOI: http://dx.doi.org/10.1007/s00300-017-2168-2.
Campbell
,
K
,
Mundy
,
CJ
,
Landy
,
JC
,
Delaforge
,
A
,
Michel
,
C
,
Rysgaard
,
S.
2016
.
Community dynamics of bottom-ice algae in Dease Strait of the Canadian Arctic
.
Progress in Oceanography
149
:
27
39
. DOI: http://dx.doi.org/10.1016/j.pocean.2016.10.005.
Cao
,
Y
,
Liang
,
S
,
Sun
,
L
,
Liu
,
J
,
Cheng
,
X
,
Wang
,
D
,
Chen
,
Y
,
Yu
,
M
,
Feng
,
K.
2022
.
Trans-Arctic shipping routes expanding faster than the model projections
.
Global Environmental Change
73
(
3
):
102488
. DOI: http://dx.doi.org/10.1016/j.gloenvcha.2022.102488.
Chen
,
Y
,
Cheng
,
JJ
,
Creamer
,
KS.
2008
.
Inhibition of anaerobic digestion process: A review
.
Bioresource Technology
99
(
10
):
4044
4064
. DOI: http://dx.doi.org/10.1016/j.biortech.2007.01.057.
Conservation of Arctic Flora and Fauna
.
2015
. The economics of ecosystems and biodiversity (TEEB) for the Arctic: A scoping study.
Akureyri, Iceland
:
Conservation of Arctic Flora and Fauna
.
Croft
,
MT
,
Lawrence
,
AD
,
Raux-Deery
,
E
,
Warren
,
MJ
,
Smith
,
AG.
2005
.
Algae acquire vitamin B12 through a symbiotic relationship with bacteria
.
Nature
438
(
7064
):
90
93
. DOI: http://dx.doi.org/10.1038/nature04056.
Croteau
,
D
,
Guérin
,
S
,
Bruyant
,
F
,
Ferland
,
J
,
Campbell
,
DA
,
Babin
,
M
,
Lavaud
,
J.
2021
.
Contrasting nonphotochemical quenching patterns under high light and darkness aligns with light niche occupancy in Arctic diatoms
.
Limnology and Oceanography
66
(
S1
):
1
15
. DOI: http://dx.doi.org/10.1002/lno.11587.
Croteau
,
D
,
Lacour
,
T
,
Schiffrine
,
N
,
Morin
,
PI
,
Forget
,
MH
,
Bruyant
,
F
,
Ferland
,
J
,
Lafond
,
A
,
Campbell
,
DA
,
Tremblay
,
,
Babin
,
M
,
Lavaud
,
J.
2022
.
Shifts in growth light optima among diatom species support their succession during the spring bloom in the Arctic
.
Journal of Ecology
110
(
6
):
1356
1375
. DOI: http://dx.doi.org/10.1111/1365-2745.13874.
Das
,
N
,
Chandran
,
P.
2011
.
Microbial degradation of petroleum hydrocarbon contaminants: An overview
.
Biotechnology Research International
2011
:
1
13
. DOI: http://dx.doi.org/10.4061/2011/941810.
DeLorenzo
,
ME
,
Key
,
PB
,
Chung
,
KW
,
Aaby
,
K
,
Hausman
,
D
,
Jean
,
C
,
Pennington
,
PL
,
Pisarski
,
EC
,
Wirth
,
EF.
2021
.
Multi-stressor effects of ultraviolet light, temperature, and salinity on Louisiana sweet crude oil toxicity in larval estuarine organisms
.
Archives of Environmental Contamination and Toxicology
80
(
2
):
461
473
. DOI: http://dx.doi.org/10.1007/s00244-021-00809-3.
Di Veroli
,
GY
,
Fornari
,
C
,
Goldlust
,
I
,
Mills
,
G
,
Koh
,
SB
,
Bramhall
,
JL
,
Richards
,
FM
,
Jodrell
,
DI.
2015
.
An automated fitting procedure and software for dose-response curves with multiphasic features
.
Scientific Reports
5
:
1
11
. DOI: http://dx.doi.org/10.1038/srep14701.
Dilliplaine
,
K
,
Oggier
,
M
,
Gradinger
,
R
,
Eicken
,
H
,
Collins
,
E
,
Bluhm
,
BA.
2021
.
Crude oil exposure reduces ice algal growth in a sea-ice mesocosm experiment
.
Polar Biology
44
:
525
537
. DOI: http://dx.doi.org/10.1007/s00300-021-02818-3.
Dubinsky
,
Z
,
Stambler
,
N.
2009
.
Photoacclimation processes in phytoplankton: Mechanisms, consequences, and applications
.
Aquatic Microbial Ecology
56
(
2–3
):
163
176
. DOI: http://dx.doi.org/10.3354/ame01345.
Dunstan
,
WM
,
Atkinson
,
LP
,
Natoli
,
J.
1975
.
Stimulation and inhibition of phytoplankton growth by low molecular weight hydrocarbons
.
Marine Biology
31
(
4
):
305
310
. DOI: http://dx.doi.org/10.1007/BF00392087.
Echeveste
,
P
,
Agustí
,
S
,
Dachs
,
J.
2010
.
Cell size dependent toxicity thresholds of polycyclic aromatic hydrocarbons to natural and cultured phytoplankton populations
.
Environmental Pollution
158
(
1
):
299
307
. DOI: http://dx.doi.org/10.1016/j.envpol.2009.07.006.
Eguíluz
,
VM
,
Fernández-Gracia
,
J
,
Irigoien
,
X
,
Duarte
,
CM.
2016
.
A quantitative assessment of Arctic shipping in 2010–2014
.
Scientific Reports
6
(
March
):
3
8
. DOI: http://dx.doi.org/10.1038/srep30682.
Eilers
,
PHC
,
Peeters
,
JCH.
1988
.
A model for the relationship between light intensity and the rate of photosynthesis in phytoplankton
.
Ecological Modelling
42
(
3–4
):
199
215
. DOI: http://dx.doi.org/10.1016/0304-3800(88)90057-9.
El-Sheekh
,
MM
,
El-Naggar
,
AH
,
Osman
,
MEH
,
Haieder
,
A.
2000
.
Comparative studies on the green algae chlorella homosphaera and chlorella vulgaris with respect to oil pollution in the river Nile
.
Water, Air, and Soil Pollution
124
(
1
):
187
204
. DOI: http://dx.doi.org/10.1023/A:1005268615405.
Faksness
,
L-G
,
Brandvik
,
PJ.
2008
.
Distribution of water soluble components from arctic marine oil spills—A combined laboratory and field study
.
Cold Regions Science and Technology
54
(
2
):
97
105
. DOI: http://dx.doi.org/10.1016/j.coldregions.2008.03.005.
Fiala
,
M
,
Delille
,
D.
1999
.
Annual changes of microalgae biomass in Antarctic sea ice contaminated by crude oil and diesel fuel
.
Polar Biology
21
(
6
):
391
396
. DOI: http://dx.doi.org/10.1007/s003000050378.
Gardiner
,
WW
,
Word
,
JQ
,
Word
,
JD
,
Perkins
,
RA
,
McFarlin
,
KM
,
Hester
,
BW
,
Word
,
LS
,
Ray
,
CM
.
2013
.
The acute toxicity of chemically and physically dispersed crude oil to key Arctic species under Arctic conditions during the open water season
.
Environmental Toxicology and Chemistry
32
(
10
):
2284
2300
. DOI: http://dx.doi.org/10.1002/etc.2307.
Garneau
,
M-È
,
Michel
,
C
,
Meisterhans
,
G
,
Fortin
,
N
,
King
,
TL
,
Greer
,
CW
,
Lee
,
K.
2016
.
Hydrocarbon biodegradation by Arctic sea-ice and sub-ice microbial communities during microcosm experiments, Northwest Passage (Nunavut, Canada)
.
Federation of European Microbiological Societies Microbial Ecology
92
(
10
):
1
18
. DOI: http://dx.doi.org/10.1093/femsec/fiw130.
Garr
,
AL
,
Laramore
,
S
,
Krebs
,
W.
2014
.
Toxic effects of oil and dispersant on marine microalgae
.
Bulletin of Environmental Contamination and Toxicology
93
(
6
):
654
659
. DOI: http://dx.doi.org/10.1007/s00128-014-1395-2.
Garrison
,
DL
,
Buck
,
KR.
1986
.
Organism losses during ice melting: A serious bias in sea ice community studies
.
Polar Biology
6
(
4
):
237
239
. DOI: http://dx.doi.org/10.1007/BF00443401.
Gaur
,
JP
,
Kumar
,
HD.
1981
.
Growth response of four micro-algae to three crude oils and a furnace oil
.
Environmental Pollution Series A, Ecological and Biological
25
(
1
):
77
85
. DOI: http://dx.doi.org/10.1016/0143-1471(81)90116-1.
Gilde
,
K
,
Pinckney
,
JL.
2012
.
Sublethal effects of crude oil on the community structure of estuarine phytoplankton
.
Estuaries and Coast
35
(
3
):
853
861
. DOI: http://dx.doi.org/10.1007/s12237-011-9473-8.
Goutx
,
MM
,
Berland
,
B
,
Leveau
,
M
,
Bertrand
,
JC.
1984
.
Effects of petroleum biodegradation products on phytoplankton growth
, in
Colloque International de Bacteriologie Marine
.
Brest, France
:
621
627
.
Available at
https://archimer.ifremer.fr/doc/1984/acte-1017.pdf.
Gradinger
,
R
,
Bluhm
,
B.
2009
.
Timing of ice algal grazing by the Arctic nearshore benthic amphipod Onisimus litoralis
.
Arctic
63
(
3
):
355
358
.
Grenfell
,
TC
,
Maykut
,
GA.
1977
.
The optical properties of ice and snow in the Arctic basin
.
Journal of Glaciology
18
(
80
):
445
463
. DOI: http://dx.doi.org/10.3189/S0022143000021122.
Guillard
,
RR
,
Ryther
,
JH.
1962
.
Studies of marine planktonic diatoms: I. Cyclotella nana Hustedt, and Detonula confervacea (cleve) Gran
.
Canadian Journal of Microbiology
8
(
2
):
229
239
. DOI: http://dx.doi.org/10.1139/m62-029.
Hall
,
BG
,
Acar
,
H
,
Nandipati
,
A
,
Barlow
,
M.
2014
.
Growth rates made easy
.
Molecular Biology and Evolution
31
(
1
):
232
238
. DOI: http://dx.doi.org/10.1093/molbev/mst187.
Hancke
,
K
,
Lund-Hansen
,
LC
,
Lamare
,
ML
,
Højlund Pedersen
,
S
,
King
,
MD
,
Andersen
,
P
,
Sorrell
,
BK.
2018
.
Extreme low light requirement for algae growth underneath sea ice: A case study from Station Nord, NE Greenland
.
Journal of Geophysical Research: Oceans
123
(
2
):
985
1000
. DOI: http://dx.doi.org/10.1002/2017JC013263.
Hansel
,
CM
,
Diaz
,
JM.
2021
.
Production of extracellular reactive oxygen species by marine biota
.
Annual Review of Marine Science
13
:
177
200
. DOI: http://dx.doi.org/10.1146/annurev-marine-041320-102550.
Hasle
,
GR
,
Medlin
,
LK
,
Syvertsen
,
EE.
1994
.
Synedropsis gen. nov., a genus of araphid diatoms associated with sea ice
.
Phycologia
33
(
4
):
248
270
. DOI: http://dx.doi.org/10.2216/i0031-8884-33-4-248.1.
Hegseth
,
EN.
1992
.
Sub-ice algal assemblages of the Barents Sea: Species composition, chemical composition, and growth rates
.
Polar Biology
12
(
5
):
485
496
. DOI: http://dx.doi.org/10.1007/BF00238187.
Hook
,
SE
,
Osborn
,
HL
.
2012
.
Comparison of toxicity and transcriptomic profiles in a diatom exposed to oil, dispersants, dispersed oil
.
Aquatic Toxicology
124–125
:
139
151
. DOI: http://dx.doi.org/10.1016/j.aquatox.2012.08.005.
Hop
,
H
,
Vihtakari
,
M
,
Bluhm
,
BA
,
Assmy
,
P
,
Poulin
,
M
,
Gradinger
,
R
,
Peeken
,
I
,
Quillfeldt
,
C
,
Olsen
,
L
,
Zhitina
,
L
,
Melnikov
,
I.
2020
.
Changes in sea-ice protist diversity with declining sea ice in the Arctic Ocean from the 1980s to 2010s
.
Frontiers in Marine Science
7
(
May
):
1
18
. DOI: http://dx.doi.org/10.3389/fmars.2020.00243.
Hsiao
,
SC.
1978
.
Effects of crude oils on the growth of Arctic marine phytoplankton
.
Environmental Pollution
17
(
2
):
93
107
.
Hsiao
,
SC.
1992
.
Dynamics of ice algae and phytoplankton in Frobisher Bay
.
Polar Biology
12
(
6–7
):
645
651
. DOI: http://dx.doi.org/10.1007/BF00236987.
Hsiao
,
SIC
,
Kittle
,
DW
,
Foy
,
MG.
1978
.
Effects of crude oils and the oil dispersant corexit on primary production of Arctic Marine phytoplankton and seaweed
.
Environmental Pollution (1970)
15
(
3
):
209
221
. DOI: http://dx.doi.org/10.1016/0013-9327(78)90066-6.
Huang
,
YJ
,
Jiang
,
ZB
,
Zeng
,
JN
,
Chen
,
QZ
,
Zhao
,
YQ
,
Liao
,
YB
,
Shou
,
L
,
Xu
,
XQ.
2011
.
The chronic effects of oil pollution on marine phytoplankton in a subtropical bay, China
.
Environmental Monitoring and Assessment
176
(
1–4
):
517
530
. DOI: http://dx.doi.org/10.1007/s10661-010-1601-6.
Hünken
,
M
,
Harder
,
J
,
Kirst
,
GO.
2008
.
Epiphytic bacteria on the Antarctic ice diatom Amphiprora kufferathii manguin cleave hydrogen peroxide produced during algal photosynthesis
.
Plant Biology
10
(
4
):
519
526
. DOI: http://dx.doi.org/10.1111/j.1438-8677.2008.00040.x.
Jabre
,
L
,
Bertrand
,
EM.
2020
.
Interactive effects of iron and temperature on the growth of Fragilariopsis cylindrus
.
Limnology and Oceanography Letters
5
(
5
):
363
370
. DOI: http://dx.doi.org/10.1002/lol2.10158.
Ji
,
R
,
Jin
,
M
,
Varpe
,
Ø.
2013
.
Sea ice phenology and timing of primary production pulses in the Arctic ocean
.
Global Change Biology
19
(
3
):
734
741
. DOI: http://dx.doi.org/10.1111/gcb.12074.
Johnsen
,
G
,
Zolich
,
A
,
Grant
,
S
,
Bjørgum
,
R
,
Cohen
,
JH
,
McKee
,
D
,
Kopec
,
TP
,
Vogedes
,
D
,
Berge
,
J.
2021
.
All-sky camera system providing high temporal resolution annual time series of irradiance in the Arctic
.
Applied Optics
60
(
22
):
6456
. DOI: http://dx.doi.org/10.1364/AO.424871.
Juhl
,
AR
,
Krembs
,
C.
2010
.
Effects of snow removal and algal photoacclimation on growth and export of ice algae
.
Polar Biology
33
(
8
):
1057
1065
. DOI: http://dx.doi.org/10.1007/s00300-010-0784-1.
Kamalanathan
,
M
,
Mapes
,
S
,
Hillhouse
,
J
,
Claflin
,
N
,
Leleux
,
J
,
Hala
,
D
,
Quigg
,
A.
2021
.
Molecular mechanism of oil induced growth inhibition in diatoms using Thalassiosira pseudonana as the model species
.
Scientific Reports
11
(
1
):
19831
. DOI: http://dx.doi.org/10.1038/s41598-021-98744-9.
Kimura
,
K
,
Tomaru
,
Y.
2013
.
A unique method for culturing diatoms on agar plates
.
Plankton and Benthos Research
8
(
1
):
46
48
. DOI: http://dx.doi.org/10.3800/pbr.8.46.
Knezevic
,
SZ
,
Streibig
,
JC
,
Ritz
,
C.
2007
.
Utilizing R software package for dose-response studies: The concept and data analysis
.
Weed Technology
21
(
3
):
840
848
. DOI: http://dx.doi.org/10.1614/WT-06-161.1.
Kohlbach
,
D
,
Ferguson
,
SH
,
Brown
,
TA
,
Michel
,
C.
2019
.
Landfast sea ice−benthic coupling during spring and potential impacts of system changes on food web dynamics in Eclipse Sound, Canadian Arctic
.
Marine Ecology Progress Series
627
:
33
48
.
Kohlbach
,
D
,
Graeve
,
M
,
Lange
,
BA
,
David
,
C
,
Peeken
,
I
,
Flores
,
H.
2016
.
The importance of ice algae-produced carbon in the central Arctic Ocean ecosystem: Food web relationships revealed by lipid and stable isotope analyses
.
Limnology and Oceanography
61
(
6
):
2027
2044
. DOI: http://dx.doi.org/10.1002/lno.10351.
Koshikawa
,
H
,
Xu
,
KQ
,
Liu
,
ZL
,
Kohata
,
K
,
Kawachi
,
M
,
Maki
,
H
,
Zhu
,
MY
,
Watanabe
,
M.
2007
.
Effect of the water-soluble fraction of diesel oil on bacterial and primary production and the trophic transfer to mesozooplankton through a microbial food web in Yangtze estuary, China
.
Estuarine, Coastal and Shelf Science
71
(
1–2
):
68
80
. DOI: http://dx.doi.org/10.1016/j.ecss.2006.08.008.
Kültz
,
D.
2003
.
Evolution of the cellular stress proteome: From monophyletic origin to ubiquitous function
.
Journal of Experimental Biology
206
(
18
):
3119
3124
. DOI: http://dx.doi.org/10.1242/jeb.00549.
Kurtz
,
NT
,
Markus
,
T
,
Farrell
,
SL
,
Worthen
,
DL
,
Boisvert
,
LN.
2011
.
Observations of recent Arctic sea ice volume loss and its impact on ocean-atmosphere energy exchange and ice production
.
Journal of Geophysical Research: Oceans
116
(
C4
):
1
19
. DOI: http://dx.doi.org/10.1029/2010JC006235.
Lacour
,
T
,
Larivière
,
J
,
Babin
,
M.
2017
.
Growth, Chl a content, photosynthesis, and elemental composition in polar and temperate microalgae
.
Limnology and Oceanography
62
(
1
):
43
58
. DOI: http://dx.doi.org/10.1002/lno.10369.
Lannuzel
,
D
,
Tedesco
,
L
,
van Leeuwe
,
M
,
Campbell
,
K
,
Flores
,
H
,
Delille
,
B
,
Miller
,
L
,
Stefels
,
J
,
Assmy
,
P
,
Bowman
,
J
,
Brown
,
K
,
Castellani
,
G
,
Chierici
,
M
,
Crabeck
,
O
,
Damm
,
E
,
Else
,
B
,
Fransson
,
A
,
Fripiat
,
F
,
Geilfus
,
NX
,
Jacques
,
C
,
Jones
,
E
,
Kaartokallio
,
H
,
Kotovitch
,
M
,
Meiners
,
K
,
Moreau
,
S
,
Nomura
,
D
,
Peeken
,
I
,
Rintala
,
JM
,
Steiner
,
N
,
Tison
,
JL
,
Vancoppenolle
,
M
,
Van der Linden
,
F
,
Vichi
,
M
,
Wongpan
,
P.
2020
.
The future of Arctic sea-ice biogeochemistry and ice-associated ecosystems
.
Nature Climate Change
10
(
11
):
983
992
. DOI: http://dx.doi.org/10.1038/s41558-020-00940-4.
Lavoie
,
D
,
Denman
,
K
,
Michel
,
C.
2005
.
Modeling ice algal growth and decline in a seasonally ice-covered region of the Arctic (Resolute Passage, Canadian Archipelago)
.
Journal of Geophysical Research
110
(
C11
):
C11009
. DOI: http://dx.doi.org/10.1029/2005JC002922.
Lemcke
,
S
,
Holding
,
J
,
Møller
,
EF
,
Thyrring
,
J
,
Gustavson
,
K
,
Juul-Pedersen
,
T
,
Sejr
,
MK.
2019
.
Acute oil exposure reduces physiological process rates in Arctic phyto- and zooplankton
.
Ecotoxicology
28
(
1
):
26
36
. DOI: http://dx.doi.org/10.1007/s10646-018-1995-4.
Leu
,
E
,
Søreide
,
JE
,
Hessen
,
DO
,
Falk-petersen
,
S
,
Berge
,
J.
2011
.
Consequences of changing sea-ice cover for primary and secondary producers in the European Arctic shelf seas: Timing, quantity, and quality
.
Progress in Oceanography
90
(
1–4
):
18
32
. DOI: http://dx.doi.org/10.1016/j.pocean.2011.02.004.
Lomas
,
MW
,
Baer
,
SE
,
Acton
,
S
,
Krause
,
JW.
2019
.
Pumped up by the cold: Elemental quotas and stoichiometry of cold-water diatoms
.
Frontiers in Marine Science
6
(
286
). DOI: http://dx.doi.org/10.3389/fmars.2019.00286.
Lund-Hansen
,
LC
,
Hawes
,
I
,
Sorrell
,
BK
,
Nielsen
,
MH.
2014
.
Removal of snow cover inhibits spring growth of Arctic ice algae through physiological and behavioral effects
.
Polar Biology
37
(
4
):
471
481
. DOI: http://dx.doi.org/10.1007/s00300-013-1444-z.
Massicotte
,
P
,
Amiraux
,
R
,
Amyot
,
MP
,
Archambault
,
P
,
Ardyna
,
M
,
Arnaud
,
L
,
Artigue
,
L
,
Aubry
,
C
,
Ayotte
,
P
,
Bécu
,
G
,
Bélanger
,
S
,
Benner
,
R
,
Bittig
,
HC
,
Bricaud
,
A
,
Brossier
,
E
,
Bruyant
,
F
,
Chauvaud
,
L
,
Christiansen-Stowe
,
D
,
Claustre
,
H
,
Cornet-Barthaux
,
V
,
Coupel
,
P
,
Cox
,
C
,
Delaforge
,
A
,
Dezutter
,
T
,
Dimier
,
C
,
Domine
,
F
,
Dufour
,
F
,
Dufresne
,
C
,
Dumont
,
D
,
Ehn
,
J
,
Else
,
B
,
Ferland
,
J
,
Forget
,
MH
,
Fortier
,
L
,
Galí
,
M
,
Galindo
,
V
,
Gallinari
,
M
,
Garcia
,
N
,
Ribeiro
,
CG
,
Gourdal
,
M
,
Gourvil
,
P
,
Goyens
,
C
,
Grondin
,
PL
,
Guillot
,
P
,
Guilmette
,
C
,
Houssais
,
MN
,
Joux
,
F
,
Lacour
,
L
,
Lacour
,
T
,
Lafond
,
A
,
Lagunas
,
J
,
Lalande
,
C
,
Laliberté
,
J
,
Lambert-Girard
,
S
,
Larivière
,
J
,
Lavaud
,
J
,
LeBaron
,
A
,
Leblanc
,
K
,
Le Gall
,
F
,
Legras
,
J
,
Lemire
,
M
,
Levasseur
,
M
,
Leymarie
,
E
,
Leynaert
,
A
,
Lopes dos Santos
,
A
,
Lourenço
,
A
,
Mah
,
D
,
Marec
,
C
,
Marie
,
D
,
Martin
,
N
,
Marty
,
C
,
Marty
,
S
,
Massé
,
G
,
Matsuoka
,
A
,
Matthes
,
L
,
Moriceau
,
B
,
Muller
,
PE
,
Mundy
,
CJ
,
Neukermans
,
G
,
Oziel
,
L
,
Panagiotopoulos
,
C
,
Pangrazi
,
JJ
,
Picard
,
G
,
Picheral
,
M
,
Pinczon du Sel
,
F
,
Pogorzelec
,
N
,
Probert
,
I
,
Quéguiner
,
B
,
Raimbault
,
P
,
Ras
,
J
,
Rehm
,
E
,
Reimer
,
E
,
Rontani
,
JF
,
Rysgaard
,
S
,
Saint-Béat
,
B
,
Sampei
,
M
,
Sansoulet
,
J
,
Schmechtig
,
C
,
Schmidt
,
S
,
Sempéré
,
R
,
Sévigny
,
C
,
Shen
,
Y
,
Tragin
,
M
,
Tremblay
,
JE
,
Vaulot
,
D
,
Verin
,
G
,
Vivier
,
F
,
Vladoiu
,
A
,
Whitehead
,
J
,
Babin
,
M
.
2020
.
Green edge ice camp campaigns: Understanding the processes controlling the under-ice Arctic phytoplankton spring bloom
.
Earth System Science Data
12
(
1
):
151
176
. DOI: http://dx.doi.org/10.5194/essd-12-151-2020.
Mcdonald
,
S
,
Koulis
,
T
,
Ehn
,
J
,
Campbell
,
K
,
Gosselin
,
M
,
Mundy
,
CJ.
2015
.
A functional regression model for predicting optical depth and estimating attenuation coefficients in sea-ice covers near resolute passage, Canada
.
Annals of Glaciology
56
(
69
):
147
154
. DOI: http://dx.doi.org/10.3189/2015AoG69A004.
McFarlin
,
KM
,
Perkins
,
RA
,
Gardiner
,
WW
,
Word
,
JD
,
Word
,
JQ.
2011
.
Toxicity of physically and chemically dispersed oil to selected Arctic species
, in
International Oil Spill Conference
. Vol:
2011
(
1
):
149
.
McFarlin
,
KM
,
Prince
,
RC
,
Perkins
,
R
,
Leigh
,
MB.
2014
.
Biodegradation of dispersed oil in Arctic seawater at −1°C
.
PLoS One
9
(
1
):
1
8
. DOI: http://dx.doi.org/10.1371/journal.pone.0084297.
Melnikov
,
IA
,
Kolosova
,
EG
,
Welch
,
HE
,
Zhitina
,
LS.
2002
.
Sea ice biological communities and nutrient dynamics in the Canada basin of the Arctic Ocean
.
Deep Sea Research Part I: Oceanographic Research Papers
49
(
9
):
1623
1649
. DOI: http://dx.doi.org/10.1016/S0967-0637(02)00042-0.
Mock
,
T
,
Hoch
,
N.
2005
.
Long-term temperature acclimation of photosynthesis in steady-state cultures of the polar diatom Fragilariopsis cylindrus
.
Photosynthesis Research
85
(
3
):
307
317
. DOI: http://dx.doi.org/10.1007/s11120-005-5668-9.
Moré
,
JJ.
1978
.
The Levenberg-Marquardt algorithm: Implementation and theory
.
Numerical Analysis
630
:
105
116
. DOI: http://dx.doi.org/10.1007/BFb0067700.
Mundy
,
CJ
,
Gosselin
,
M
,
Gratton
,
Y
,
Brown
,
K
,
Galindo
,
V
,
Campbell
,
K
,
Levasseur
,
M
,
Barber
,
D
,
Papakyriakou
,
T
,
Bélanger
,
S.
2014
.
Role of environmental factors on phytoplankton bloom initiation under landfast sea ice in resolute passage, Canada
.
Marine Ecology Progress Series
497
:
39
49
. DOI: http://dx.doi.org/10.3354/meps10587.
Nicolaus
,
M
,
Petrich
,
C
,
Hudson
,
SR
,
Granskog
,
MA.
2013
.
Variability of light transmission through Arctic land-fast sea ice during spring
.
Cryosphere
7
(
3
):
977
986
. DOI: http://dx.doi.org/10.5194/tc-7-977-2013.
NSF Arctic Data Center
.
2009
.
Automated ice mass balance site (SIZONET)
.
Arctic Data Center
. DOI: http://dx.doi.org/10.18739/A2BM0G.
Oggier
,
M
,
Eicken
,
H
,
Wilkinson
,
J
,
Petrich
,
C
,
O’Sadnick
,
M.
2020
.
Crude oil migration in sea-ice: Laboratory studies of constraints on oil mobilization and seasonal evolution
.
Cold Regions Science and Technology
174
(
September
):
102924
. DOI: http://dx.doi.org/10.1016/j.coldregions.2019.102924.
Onarheim
,
IH
,
Eldevik
,
T
,
Smedsrud
,
LH
,
Stroeve
,
JC.
2018
.
Seasonal and regional manifestation of Arctic sea ice loss
.
Journal of Climate
31
(
12
):
4917
4932
. DOI: http://dx.doi.org/10.1175/JCLI-D-17-0427.1.
Ozhan
,
K
,
Bargu
,
S.
2014
.
Distinct responses of gulf of Mexico phytoplankton communities to crude oil and the dispersant corexit® Ec9500A under different nutrient regimes
.
Ecotoxicology
23
(
3
):
370
384
. DOI: http://dx.doi.org/10.1007/s10646-014-1195-9.
Özhan
,
K
,
Miles
,
SM
,
Gao
,
H
,
Bargu
,
S.
2014
.
Relative phytoplankton growth responses to physically and chemically dispersed South Louisiana sweet crude oil
.
Environmental Monitoring and Assessment
186
(
6
):
3941
3956
. DOI: http://dx.doi.org/10.1007/s10661-014-3670-4.
Ozhan
,
K
,
Parsons
,
ML
,
Bargu
,
S.
2014
.
How were phytoplankton affected by the deepwater horizon oil spill?
BioScience
64
(
9
):
829
836
. DOI: http://dx.doi.org/10.1093/biosci/biu117.
Ozhan
,
K
,
Zahraeifard
,
S
,
Smith
,
AP
,
Bargu
,
S.
2015
.
Induction of reactive oxygen species in marine phytoplankton under crude oil exposure
.
Environmental Science and Pollution Research
22
(
23
):
18874
18884
. DOI: http://dx.doi.org/10.1007/s11356-015-5037-y.
Pančić
,
M
,
Hansen
,
PJ
,
Tammilehto
,
A
,
Lundholm
,
N.
2015
.
Resilience to temperature and pH changes in a future climate change scenario in six strains of the polar diatom Fragilariopsis cylindrus
.
Biogeosciences
12
(
14
):
4235
4244
. DOI: http://dx.doi.org/10.5194/bg-12-4235-2015.
Parab
,
SR
,
Pandit
,
RA
,
Kadam
,
AN
,
Indap
,
MM.
2008
.
Effect of Bombay high crude oil and its water-soluble fraction on growth and metabolism of diatom Thalassiosira sp
.
Indian Journal of Marine Sciences
37
(
3
):
251
255
.
Perovich
,
D
,
Roesler
,
S
,
Pegau
,
WS.
1998
.
Variability in Arctic sea ice optical properties
.
Journal of Geophysical Research: Oceans
103
(
C1
):
1193
1208
.
Petrich
,
C
,
Karlsson
,
J
,
Eicken
,
H.
2013
.
Porosity of growing sea ice and potential for oil entrainment
.
Cold Regions Science and Technology
87
:
27
32
. DOI: http://dx.doi.org/10.1016/j.coldregions.2012.12.002.
Poltermann
,
M.
2001
.
Arctic sea ice as feeding ground for amphipods—Food sources and strategies
.
Polar Biology
24
(
2
):
89
96
. DOI: http://dx.doi.org/10.1007/s003000000177.
Poulin
,
M
,
Daugbjerg
,
N
,
Gradinger
,
R
,
Ilyash
,
L
,
Ratkova
,
T
,
von Quillfeldt
,
C.
2011
.
The pan-Arctic biodiversity of marine pelagic and sea-ice unicellular eukaryotes: A first-attempt assessment
.
Marine Biodiversity
41
(
1
):
13
28
. DOI: http://dx.doi.org/10.1007/s12526-010-0058-8.
Poulin
,
M
,
Underwood
,
GJC
,
Michel
,
C.
2014
.
Sub-ice colonial Melosira arctica in Arctic first-year ice
.
Diatom Research
29
(
2
):
213
221
. DOI: http://dx.doi.org/10.1080/0269249X.2013.877085.
Putzeys
,
S
,
Juárez-Fonseca
,
M
,
Valencia-Agami
,
SS
,
Mendoza-Flores
,
A
,
Cerqueda-García
,
D
,
Aguilar-Trujillo
,
AC
,
Martínez-Cruz
,
ME
,
Okolodkov
,
YB
,
Arcega-Cabrera
,
F
,
Herrera-Silveira
,
JA
,
Aguirre-Macedo
,
ML
,
Pech
,
D.
2022
.
Effects of a light crude oil spill on a tropical coastal phytoplankton community
.
Bulletin of Environmental Contamination and Toxicology
108
(
1
):
55
63
. DOI: http://dx.doi.org/10.1007/s00128-021-03306-4.
Rodriguez Martinez
,
MF
,
Kelessidou
,
N
,
Law
,
Z
,
Gardiner
,
J
,
Stephens
,
G.
2008
.
Effect of solvents on obligately anaerobic bacteria
.
Anaerobe
14
(
1
):
55
60
. DOI: http://dx.doi.org/10.1016/j.anaerobe.2007.09.006.
Rutgers van der Loeff
,
MM
,
Cassar
,
N
,
Nicolaus
,
M
,
Rabe
,
B
,
Stimac
,
I
.
2014
.
The influence of sea ice cover on air-sea gas exchange estimated with radon-222 profiles
.
Journal of Geophysical Research: Oceans
119
(
5
):
2735
2751
. DOI: http://dx.doi.org/10.1002/2013JC009321.
Sargian
,
P
,
Mas
,
S
,
Pelletier
,
É
,
Demers
,
S
.
2007
.
Multiple stressors on an Antarctic microplankton assemblage: Water soluble crude oil and enhanced UVBR level at Ushuaia (Argentina)
.
Polar Biology
30
(
7
):
829
841
. DOI: http://dx.doi.org/10.1007/s00300-006-0243-1.
Schaum
,
E
,
Rost
,
B
,
Millar
,
AJ
,
Collins
,
S.
2013
.
Variation in plastic responses of a globally distributed picoplankton species to ocean acidification
.
Nature Climate Change
3
(
3
):
298
302
. DOI: http://dx.doi.org/10.1038/nclimate1774.
Schneider
,
CA
,
Rasband
,
WS
,
Eliceiri
,
KW.
2012
.
NIH image to imagej: 25 years of image analysis
.
Nature Methods
9
(
7
):
671
675
. DOI: http://dx.doi.org/10.1038/nmeth.2089.
Sherry
,
A
,
Grant
,
RJ
,
Aitken
,
CM
,
Jones
,
DM
,
Head
,
IM
,
Gray
,
ND.
2014
.
Volatile hydrocarbons inhibit methanogenic crude oil degradation
.
Frontiers in Microbiology
5
(
Apr
):
1
9
. DOI: http://dx.doi.org/10.3389/fmicb.2014.00131.
Sikkema
,
J
,
De Bont
,
JAM
,
Poolman
,
B.
1995
.
Mechanisms of membrane toxicity of hydrocarbons
.
Microbiological Reviews
59
(
2
):
201
222
.
Silsbe
,
G
,
Malkin
,
S.
2015
.
Phytotools: Phytoplankton production tools
.
R package version 1.0
.
Available at
https://CRAN.R-project.org/package=phytotools.
Accessed June 20, 2020
.
Singer
,
M
,
Aurand
,
D
,
Bragin
,
G
,
Clark
,
J
,
Coelho
,
G
,
Sowby
,
M,
Tjeerdema
,
R.
2000
.
Standardization of the preparation and quantitation of water-accommodated fractions of petroleum for toxicity testing
.
Marine Pollution Bulletin
40
(
11
):
1007
1016
. DOI: http://dx.doi.org/10.1016/S0025-326X(00)00045-X.
Sommer
,
U.
1989
.
Maximal growth rates of Antarctic phytoplankton: Only weak dependence on cell size
.
Limnology and Oceanography
34
(
6
):
1109
1112
. DOI: http://dx.doi.org/10.4319/lo.1989.34.6.1109.
Song
,
HJ
,
Lee
,
JH
,
Kim
,
GW
,
Ahn
,
SH
,
Joo
,
HM
,
Jeong
,
JY
,
Yang
,
EJ
,
Kang
,
SH
,
Lee
,
SH.
2016
.
In-situ measured primary productivity of ice algae in Arctic sea ice floes using a new incubation method
.
Ocean Science Journal
51
(
3
):
387
396
. DOI: http://dx.doi.org/10.1007/s12601-016-0035-7.
Søreide
,
JE
,
Leu
,
E
,
Berge
,
J
,
Graeve
,
M
,
Falk-Petersen
,
S.
2010
.
Timing of blooms, algal food quality and Calanus glacialis reproduction and growth in a changing Arctic
.
Global Change Biology
16
(
11
):
3154
3163
. DOI: http://dx.doi.org/10.1111/j.1365-2486.2010.02175.x.
Soto
,
C
,
Hellebust
,
JA
,
Hutchinson
,
TC
,
Sawa
,
T.
1975
.
Effect of naphthalene and aqueous crude oil extracts on the green flagellate Chlamydomonas angulosa. I. Growth
.
Canadian Journal of Botany
53
(
2
):
109
117
. DOI: http://dx.doi.org/10.1139/b75-017.
Stebbing
,
ARD.
1981
.
The kinetics of growth control in a colonial hydroid
.
Journal of the Marine Biological Association of the United Kingdom
61
(
1
):
35
63
. DOI: http://dx.doi.org/10.1017/S0025315400045902.
Stebbing
,
ARD.
2009
.
Interpreting ‘dose-response’ curves using homeodynamic data: With an improved explanation for hormesis
.
Dose-Response
7
(
3
). DOI: http://dx.doi.org/10.2203/dose-response.08-020.Stebbing.
Steiner
,
NS
,
Cheung
,
WWL
,
Cisneros-Montemayor
,
AM
,
Drost
,
H
,
Hayashida
,
H
,
Hoover
,
C
,
Lam
,
J
,
Sou
,
T
,
Sumaila
,
UR
,
Suprenand
,
P
,
Tai
,
TC
,
VanderZwaag
,
DL.
2019
.
Impacts of the changing ocean-sea ice system on the key forage fish Arctic cod (Boreogadus Saida) and subsistence fisheries in the western Canadian Arctic—Evaluating linked climate, ecosystem and economic (CEE) models
.
Frontiers in Marine Science
6
(
179
):
1
24
. DOI: http://dx.doi.org/10.3389/fmars.2019.00179.
Stroeve
,
J
,
Liston
,
GE
,
Buzzard
,
S
,
Zhou
,
L
,
Mallett
,
R
,
Barrett
,
A
,
Tschudi
,
M
,
Tsamados
,
M
,
Itkin
,
P
,
Stewart
,
JS.
2020
.
A Lagrangian snow evolution system for sea ice applications (SnowModel-LG): Part II—Analyses
.
Journal of Geophysical Research: Oceans
125
(
10
):
e2019JC015900
. DOI: http://dx.doi.org/10.1029/2019JC015900.
Sunda
,
WG
,
Huntsman
,
SA.
1995
.
Iron uptake and growth limitation in oceanic and coastal phytoplankton
.
Marine Chemistry
50
(
1–4
):
189
206
. DOI: http://dx.doi.org/10.1016/0304-4203(95)00035-P.
Szymanski
,
A
,
Gradinger
,
R.
2016
.
The diversity, abundance and fate of ice algae and phytoplankton in the Bering Sea
.
Polar Biology
39
(
2
):
309
325
. DOI: http://dx.doi.org/10.1007/s00300-015-1783-z.
Taş
,
S
,
Okuş
,
E
,
Ünlü
,
S
,
Altiok
,
H.
2010
.
A study on phytoplankton following ‘Volgoneft-248’ oil spill on the north-eastern coast of the Sea of Marmara
.
Journal of the Marine Biological Association of the United Kingdom
91
(
03
):
715
725
. DOI: http://dx.doi.org/10.1017/S0025315410000330.
Tedesco
,
L
,
Vichi
,
M
,
Scoccimarro
,
E.
2019
.
Sea-ice algal phenology in a warmer Arctic
.
Science Advances
5
(
5
). DOI: http://dx.doi.org/10.1126/sciadv.aav4830.
Tremblay
,
,
Simpson
,
K
,
Martin
,
J
,
Miller
,
L
,
Gratton
,
Y
,
Barber
,
D
,
Price
,
NM.
2008
.
Vertical stability and the annual dynamics of nutrients and chlorophyll fluorescence in the coastal, southeast Beaufort Sea
.
Journal of Geophysical Research: Oceans
113
(
C7
):
1
14
. DOI: http://dx.doi.org/10.1029/2007JC004547.
Veyssière
,
G
,
Castellani
,
G
,
Wilkinson
,
J
,
Karcher
,
M
,
Hayward
,
A
,
Stroeve
,
JC
,
Nicolaus
,
M
,
Kim
,
J-H
,
Yang
,
E-J
,
Valcic
,
L
,
Kauker
,
F
,
Khan
,
AL
,
Rogers
,
I
,
Jung
,
J.
2022
.
Under-ice light field in the western Arctic Ocean during late summer
.
Frontiers in Earth Science
9
(
Feb
):
1
19
. DOI: http://dx.doi.org/10.3389/feart.2021.643737.
von Quillfeldt
,
CH.
1995
.
Distribution of diatoms in the Northeast Water Polynya, Greenland
.
Journal of Marine Systems
10
(
1–4
):
211
240
. DOI: http://dx.doi.org/10.1016/S0924-7963(96)00056-5.
Webster
,
MA
,
Rigor
,
IG
,
Nghiem
,
SV
,
Kurtz
,
NT
,
Farrell
,
SL
,
Perovich
,
DK
,
Sturm
,
M.
2014
.
Interdecadal changes in snow depth on Arctic sea ice
.
Journal of Geophysical Research: Oceans
119
(
8
):
5395
5406
. DOI: http://dx.doi.org/10.1002/2014jc009985.

How to cite this article: Dilliplaine, K, Hennon, G. 2023. Impacts of crude oil on Arctic sea-ice diatoms modified by irradiance. Elementa: Science of the Anthropocene 11(1). DOI: https://doi.org/10.1525/elementa.2023.00074

Domain Editor-in-Chief: Jody W. Deming, University of Washington, Seattle, WA, USA

Associate Editor: Kevin R. Arrigo, Department of Earth System Science, Stanford University, Stanford, CA, USA

Knowledge Domain: Ocean Science

This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International License (CC-BY 4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. See http://creativecommons.org/licenses/by/4.0/.

Supplementary Material