You are on page 1of 9

Marine Pollution Bulletin 139 (2019) 381–389

Contents lists available at ScienceDirect

Marine Pollution Bulletin


journal homepage: www.elsevier.com/locate/marpolbul

Inorganic nutrients have a significant, but minimal, impact on a coastal T


microbial community's response to fresh diluted bitumen

Alice C. Ortmann , Susan E. Cobanli, Gary Wohlgeschaffen, Peter Thamer, Claire McIntyre,
Jennifer Mason, Thomas L. King
Center for Offshore Oil, Gas and Energy Research, Department of Fisheries and Oceans Canada, Bedford Institute of Oceanography, 1 Challenger Drive, Dartmouth, NS B2Y
4A2, Canada

A R T I C LE I N FO A B S T R A C T

Keywords: Microbes capable of degrading hydrocarbons in oil are present in low abundances in coastal waters, but quickly
Biodegradation respond to oil following a spill. When estimating potential biodegradation rates in the laboratory, high con-
Diluted bitumen centrations of inorganic nutrients are often added to prevent nutrient limitation. In this study, we tested the
Dilbit short term response of coastal microbes to fresh diluted bitumen under varying nutrient conditions in a cold
Nutrient limitation
water regime. Total hydrocarbon concentrations changed minimally over five days; however, oil composition
Coastal waters
changed over time and the abundance of microbes increased in all treatments. Addition of phosphate, with or
without nitrogen, resulted in rapid changes in community composition, but after three days treatments no longer
differed. Nutrients were never depleted in any treatment suggesting that, even at low inorganic nutrient con-
centrations, microbial communities can quickly respond to hydrocarbons following a spill.

1. Introduction compounds that are not present in the original oil, but may dissolve
more easily in water. Rates and extent of all degradation processes will
Extraction and transportation of bitumen from oil sands in north vary depending on the oil type and environmental conditions (e.g.
central Canada is expected to increase by 53% between 2016 and 2030 Bagby et al., 2017; Hori et al., 2014; Owsianiak et al., 2009). Following
(CAPP, 2017). Bitumen is a semi-solid material that is diluted with the Deepwater Horizon spill in the Gulf of Mexico, biodegradation ra-
lighter hydrocarbons (diluted bitumen) to enable transport using pi- pidly removed gases and alkanes in the deep water plume (Redmond
pelines to the coast, where it can be transferred to tankers for export and Valentine, 2012; Valentine et al., 2010), with significant biode-
(reviewed in Crosby et al., 2013). Because of the relatively few studies gradation occurring along shorelines (Kostka et al., 2011; Mortazavi
of diluted bitumen, there is concern regarding the fate, behaviour and et al., 2013). However, heavier oils, including diluted bitumen, have
effects of this material if it were to spill, either from a pipeline leak, higher contributions from resins and asphaltenes that are more resistant
during transfer at a port or due to a tanker accident at sea. As a com- to biodegradation (Prince et al., 2003). Twenty years after the Arrow
posite of light and heavy oils, diluted bitumen weathers quickly in- spill of Bunker C fuel oil in Chedabucto Bay, Nova Scotia, petroleum
itially, but then undergoes minimal change over a longer time frame residues ranging from only slightly weathered to highly weathered
(King et al., 2014). could be detected on the shores of the bay (Vandermeulen and Singh,
Physical, chemical and biological processes, collectively known as 1994). Even in 2016, oil could still be detected just below the surface of
weathering, change the physical properties and chemical composition the cobble beach of Black Duck Cove in Chedabucto Bay (MacDonald,
of the oil immediately following a spill. Degradation of the oil includes 2017).
oxidation of compounds either driven by light (photooxidation), che- Few studies have investigated the ability of the microbial commu-
micals or microbial activity (biodegradation). Biodegradation is mainly nity to degrade diluted bitumen. Using microbes enriched from the
due to the actions of specific Bacteria, which can break down the oil Kalamazoo River sediment, which was exposed to diluted bitumen
(Hazen et al., 2016; Head et al., 2006). Some of the compounds in oil following a pipeline spill in 2010, high rates of biodegradation were
can be completely degraded to CO2 through the actions of microbial measured for alkanes and polycyclic aromatic hydrocarbons (PAHs) at
consortia. Other compounds can be modified and transformed into new 25 °C, with slower rates and less overall degradation measured at 5 °C


Corresponding author.
E-mail address: Alice.Ortmann@dfo-mpo.gc.ca (A.C. Ortmann).

https://doi.org/10.1016/j.marpolbul.2019.01.012
Received 4 July 2018; Received in revised form 2 January 2019; Accepted 6 January 2019
Available online 12 January 2019
0025-326X/ Crown Copyright © 2019 Published by Elsevier Ltd. All rights reserved.
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

(Deshpande et al., 2018). In contrast, a coastal microbial community Table 1


that had not previously been exposed to diluted bitumen exhibited slow Mean initial concentrations of nutrients in the original seawater and in the
degradation rates, and minimal degradation was measured at ~18 °C seven treatments in ESAW. Standard deviations are in parentheses for the tri-
over 13 days (King et al., 2014). The difference in these estimated rates plicate samples.
could be due to previous exposure or the ability of freshwater microbial Treatment NO3− (μM) NH4+ (μM) PO4−3 (μM) N:P
communities to degrade diluted bitumen more effectively than marine
Bedford Basin Water 6.70 6.30 1.01 13.24
communities; however, inorganic nutrient concentrations were also
Control 3.43 (0.20) 6.63 (1.75) 3.52 (0.35) 2.91
significantly different between the two experiments. With the microbes Dilbit 3.71 (0.43) 6.16 (0.35) 3.72 (0.07) 2.70
enriched from the freshwater sediments, high concentrations of ni- Nitrate 241.17 (3.63) 5.71 (0.31) 3.52 (0.43) 70.18
trogen and phosphorous were used in the experiments, while the ex- Ammonium 4.27 (0.12) 301.18 (9.27) 3.40 (0.39) 89.97
periment using the coastal microbial community included no additional Phosphate 3.39 (0.41) 6.98 (1.40) 41.24 (2.43) 0.26
Nitrate + phosphate 244.58 (2.63) 5.58 (0.61) 41.68 (2.17) 6.01
nutrients above what was naturally present in the seawater. Because oil
Ammonium + phosphate 3.99 (0.08) 310.95 (7.01) 40.05 (3.32) 7.87
is dominated by carbon, nitrogen and phosphorous are thought to limit
the microbial response (Atlas and Hazen, 2011), thus the low de-
gradation rates measured for the marine community may be due to low flask received 100 mL of ESAW + microbes, 2 mL of nutrient solutions
availability of inorganic nutrients. or distilled water, and 15 μL of fresh Access Western Blend (AWB), a
A short term experiment was carried out to quantify the impact of bitumen product diluted with condensate (dilbit). The controls received
inorganic nutrients on degradation that may occur immediately fol- no AWB or additional nutrients. The concentrations of nitrate, ammo-
lowing a spill of diluted bitumen in coastal waters. Fresh diluted bi- nium and phosphate in the ESAW were chosen to represent the lower
tumen, along with a natural coastal microbial community was added to concentrations detected in the Bedford Basin over a 20-year time series
artificial seawater with low concentrations of nitrate, ammonium and (Li, 2014). The added nitrate and ammonium resulted in the nitrate and
phosphorous and samples were collected over a five day period. Other ammonium concentrations being similar to that in Bushnell Haas
treatments included addition of inorganic nutrients alone, or in com- media, a common media used for isolating hydrocarbon degrading
bination, to determine if total degradation rates could be increased. bacteria (Bushnell and Haas, 1941), although at one half the total ni-
While both abiotic and biological degradation would be occurring in trogen concentration in the Bushnell Haas media, which contains both
the flasks, inorganic nutrients are most likely to impact biological nitrate and ammonium. These concentrations were ~35 times higher
processes, without affecting physical or chemical weathering of the oil. than measured in the Bedford Basin water. Phosphate was added to
The short time frame was chosen to characterize what might happen in maintain the N:P ratio from the Bedford Basin, but were determined to
the first few days following a spill. Given that in an actual spill situa- be ~40 times higher than in Bedford Basin, resulting in concentrations
tion, water movement will introduce nutrients and dilute contaminants, of 20% of that found in Bushnell Haas media. Baffle flasks for each
a longer incubation was not undertaken. Additionally, the short time- treatment were all prepared one at a time, resulting in a staggering of
frame alleviates bottle affects associated with long incubations of small the start times for each treatment. After addition of water, nutrients and
volumes. Previous studies also suggest that diluted bitumen rapidly oil, flasks were capped and sealed with Teflon tape.
weathers, with most of the changes occurring in the first few days (King Following preparation of each treatment, three flasks were sacri-
et al., 2014). The experiment with the Kalamazoo sediments also de- ficed. The remaining flasks were stored in an incubator set to 4 °C until
termined that most of the biodegradation occurred within the first being transported to a shipping container equipped with orbital sha-
10 days, even at 5 °C (Deshpande et al., 2018), therefore, much of the kers, a small heater to prevent freezing, an air temperature probe and
degradation that is likely to occur should be observed in the first few three temperature data loggers (HOBO® pendants) submerged in
days following a spill. The experiment was also carried out at low 100 mL of water. Low light was used during sampling otherwise in-
temperature to reflect winter conditions. cubations occurred in the dark. Shakers were set to 150 rpm to ensure
rapid mixing of the flasks. Three flasks for each treatment were
2. Materials and methods equipped with a Presens O2 optical sensor. A Fibox 4 handheld probe
was used to measure O2 concentrations in the headspace of the flasks
2.1. Sample collection and experimental set-up immediately after the treatment flasks were prepared and sealed,
45 min after the last treatment was placed in the shipping container (3 h
Seawater was collected near the shore of Bedford Basin, Dartmouth, after the first treatment was finished) and then every day for the
Nova Scotia in December 2016. At the time of sampling, temperature, duration of the experiment (t1–t5). For analysis, the measurement
salinity and dissolved oxygen (DO) were measured using a Professional made 0.75–3 h post set up was considered the beginning of the ex-
Plus 2030 handheld probe (YSI). Seawater was taken to the Bedford periment (t0). O2 was measured by taking three consecutive measure-
Institute of Oceanography, where cells were concentrated over 47 mm ments for each flask. Air temperature and barometric pressure were
0.22 μm Durapore® (polyvinylidene fluoride) filters (EMD Millipore). input into the Fibox 4 system prior to collection of data. The system was
One litre of seawater was reduced to 0.2 L using gentle vacuum previously calibrated using air and N2 (0 mg L−1 O2). Triplicate mea-
(5 in. Hg) in 15 separate autoclaved filter holders. Cells were re-sus- surements were averaged for each flask and time point and readings
pended gently using disposable Pasteur pipettes. The 0.2 L from each of were converted to mg L−1. After measuring O2, three flasks from each
the filter holders were combined and diluted to a final volume of 15 L in treatment were collected each day and processed as described below.
artificial seawater (ESAW) at a salinity of 27 (Berges et al., 2001; Flasks with O2 sensors were sacrificed on the final day of the experi-
Harrison et al., 1980). The recipe was modified to reduce the con- ment after measuring O2.
centrations of nitrate, ammonium, phosphate and silicate (Table 1,
control and dilbit). Approximately 2 h passed between collection of
seawater and dilution of cells in ESAW. 2.2. Subsampling incubations
The degradation experiment included 7 treatments: a control with
no additions to confirm that manipulations did not kill the microbial Flasks were sampled with an autoclaved, 15 cm long stainless steel
community, diluted bitumen only (dilbit) and diluted bitumen with needle connected to an all-plastic syringe. A 30 mL sample was ex-
added nitrate, ammonium, phosphate, nitrate + phosphate or ammo- tracted, and ~1 mL was ejected into a microfuge tube. From this vo-
nium + phosphate. For each treatment, 18 baffle flasks were prepared lume, a 980 μL aliquot was added to 20 μL of 25% EM grade glutar-
using the ESAW containing the natural microbial community. Each aldehyde in a cryovial. After 10 min on ice, the fixed sample was flash

382
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

frozen in liquid N2 and stored at −86 °C until prokaryotes were 18 h before the DCM layer was removed, weighed and evaporated using
quantified by flow cytometry. The remaining volume in the syringe was an N-Evaporator (Organomation, Berlin, MA). Samples were diluted to
filtered through a 0.2 μm polycarbonate filter using an acid washed a final volume of 10 mL with DCM and stored at −20 °C. After con-
polypropylene filter holder. The filtrate was collected in an acid cleaned centration to 1 mL, extracts (1 μL) were injected and analyzed by an
plastic vial and frozen at −20 °C for analysis of inorganic nutrients. The Agilent 7890B GC connected to a flame ionization detector (FID) and
filter was removed, folded in half and placed in a cryovial. The filter quantified using an 8 point calibration curve generated from the stock
was flash frozen and stored at −86 °C for community diversity analysis AWB (Cole et al., 2007; King et al., 2015). Raw data were exported as
by DNA sequencing. The remaining 72 mL in each flask was carefully text and spectra were aligned using SpecAlign (Wong et al., 2005) with
poured into a 150 mL amber glass bottle. 10 mL of dichloromethane the average spectrum generated from all sample extracts using peak
(DCM) was added to the flask, rinsing the needle. The DCM was then matching. Spectra were trimmed to retain data from 5 to 20 s and im-
swirled around the flask to dissolve any dilbit stuck to the sides and ported into R for analysis. The remaining sample in DCM was ex-
poured into the amber bottle. The bottles were sealed, shaken and changed into hexanes and purified using a modified solid-phase ex-
stored at 4 °C until they were extracted to analyze hydrocarbons. traction method (EPA, 1996). These extracts were analyzed using an
To quantify prokaryote cell numbers, cryovials were thawed in a Agilent 6890 GC coupled to an Agilent 5973N MS operated in the se-
water bath at 37 °C. Samples were diluted 1:10 in Tris-EDTA (TE, lective ion monitoring mode (SIM). Specific n-alkanes (C10 to C35),
pH = 8) and incubated at room temperature for 10 min with 5 μL of a PAHs and alkylated PAHs were quantified using retention time
1:200 dilution of SYBR Green (Lonza) (Ortmann et al., 2012) and run matching to a seven-point calibration curve for identification (EPA,
on MED for 3 min through a FACSCalibur flow cytometer (BD Bios- 2007). Values were standardized to the concentration of 17α(H),
ciences) equipped with a 488 nm laser. For quality control, 1 μm 21β(H)-hopane, a conservative biomarker (Prince et al., 1994; Venosa
fluorescent beads were added to each sample. The total number of cells et al., 1997). Hydrocarbon analysis was not carried out for control
was determined by gating the signals in a FL1 (green) vs. side scatter samples.
(SSC) plot and converted to cells mL−1 using the run time and the ca-
librated flow rate. Inorganic nutrients were measured using a SEAL 2.3. Data analysis
Analytical AA3 continuous segmented flow autoanalyzer using standard
methods. Detection limits were determined for each run and only va- The control flasks were used to confirm that the manipulation of the
lues above the detection limit were considered non-zero. NO3− was samples had not inhibited cell activity and were not included in the
calculated as the difference between NO2− + NO3− and NO2−. DNA analysis of the various parameters. The effect of nutrient treatment on
was extracted from filters using the QIAGEN MagAttract PowerWater prokaryote abundance, Shannon diversity, PD and TPH was determined
DNA/RNA Kit run on a KingFisher Duo Prime (Thermo) as per the by applying two-way ANOVAs to test the effects of treatment, time and
manufacturer's instructions, with the initial bead-beating in 2 mL tubes the interaction term. Post hoc tests were carried out using a control
on a Disruptor Genie (Scientific Industries). The V4–V5 region of the group, with comparisons to either t0 or the Dilbit treatment following
16S rRNA gene was amplified (Walters et al., 2016) and sequenced on the Holm-Šídák method (SigmaPlot 13, Systat Software, Inc.). For TPH,
an Illumina MiSeq (300 + 300 bp) by the Integrated Microbiome Re- concentrations at each time point were divided by the average con-
source at Dalhousie University, Halifax, NS. centration at t0 for each treatment to compensate for differences in the
Raw demultiplexed FASTQ files were analyzed following the starting TPH concentrations between treatments. ANOVA analysis was
workflow described in the Microbiome Helper SOP (Comeau et al., then carried out on the relative TPH concentrations. After alignment,
2017). Briefly, paired reads were stitched together using PEAR (Zhang trimming and background subtraction, GC-FID spectra were analyzed
et al., 2014) and filtered to remove low quality short reads. Remaining using Principle Components Analysis (PCA) in R (R Core Team, 2017)
sequences were screened to identify and remove chimeras using VSEARCH using the rda function with scaling in the VEGAN package (Oksanen et al.,
(Rognes et al., 2016) and then processed via the open reference picking 2018). O2 concentrations were analyzed using a two-way repeated
pipeline in QIIME (Caporaso et al., 2010) using SORTMERNA (Kopylova measures ANOVA. Post hoc tests were carried out with Dilbit and t0 as
et al., 2012) and SUMACLUST (Mercier et al., 2013) to identify operational the control groups. For inorganic nutrients, 1-way ANOVAs or Kruskal
taxonomic units (OTUs) at 97% similarity. The OTU table was then Wallis analyses were carried out for each treatment and nutrient se-
filtered to remove OTUs representing < 0.1% of the total reads parately to identify changes over time. This was due to the large dif-
(Bokulich et al., 2013). Because samples were incubated in the dark, the ferences in starting concentrations among treatments. The initial mi-
OTU table was filtered to remove photosynthetic taxa including phylum crobial community in the seawater was compared to the microbial
Cyanobacteria, order Chromatiales and family Rhodospirillaceae, all of communities in the experimental flasks at t0 using the SIMPER routine
which decreased over the course of the experiment. The filtering step in Primer (Primer-E, Auckland). Taxa contributing at least 2% to the
removed 119 Cyanobacteria OTUs representing 7.9% of the sequences, differences between communities were identified. The effects of treat-
8 Chromatiales OTUs representing 0.04% of the sequences and 29 ment and time on community structure were identified using 2-way
Rhodospirillaceae OTUs representing 0.4% of the sequences. Ad- ANOSIM analysis in Primer based on the Bray-Curtis dissimilarity ma-
ditionally, taxa identified as mitochondria were removed (12 OTUs, trix. For pairwise comparisons, communities were considered sig-
0.1% of the sequences) to retain only sequences associated with Bac- nificantly different when the R2 value was > 0.3 and the p-value
teria and Archaea. Finally, the OTU table was normalized to a final was < 0.05. Changes in community structure were visualized with
sequencing depth of 12,100 sequences per sample. Due to low sequence Principle Correspondence Analysis (PCO). The rarefied OTU table was
numbers, three samples were omitted from further analysis. Ten sepa- imported into STAMP (Parks et al., 2014) to identify taxa that differed
rate rarefied OTU tables were generated to estimate sample diversity between the two groups of treatment identified by ANOSIM for each
metrics with one being randomly selected for community structure time point. OTUs were collapsed into taxonomic groups at the lowest
analysis. For each OTU table, Shannon diversity and Phylogenetic dis- level possible, which for most OTUs meant identification to the family
tance (PD) were calculated. The mean of the ten estimates was averaged level. For each time point, the two groups were compared using Welch's
to generate a single diversity estimate for each sample. Community two sided t-test with a Benjamini Hochberg FDR correction. Taxa with
structure was measured by calculating a Bray-Curtis dissimilarity ma- corrected p-values < 0.05 and > 1% difference between mean abun-
trix for one randomly selected OTU table. FASTQ files are available dances were considered significantly different. Taxa that changed over
through the NCBI SRA database (accession number SRP145477). time were analyzed in STAMP for each ANOSIM group and the control
Samples for quantifying total petroleum hydrocarbons (TPH) were treatment using ANOVA with Tukey-Kramer post hoc tests and the
placed on a Wheaton R2P roller (Wheaton, Millville, NJ) at 9 rpm for Benjamini Hochberg FDR. Corrected p-values < 0.05 and η2 effects

383
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

sizes of > 0.9 were identified taxa with significant change in abundance
over time. For clarity, mean relative abundances of abundant taxa were
plotted for each group at each time. Abundant taxa were defined as
those with a mean relative abundance of ≥0.5%. To estimate de-
gradation rates, linear regressions of the natural log of the sum of the
alkanes, alkylated PAHs or PAHs normalized to hopane were carried
out against time for each treatment. Significant slopes were determined
using a two-tailed t-test with p < 0.1.

3. Results

The temperature of the seawater at collection was 6.1 °C with a


salinity of 26.27, slightly lower than the ESAW media. DO was mea-
sured at 9.48 mg L−1, compared with 11–12 mg L−1 at t0 (Fig. S1).
ESAW was stored in a 4 °C incubator prior to setting up the experiment, Fig. 1. Abundance of prokaryote cells in the different treatments over time.
minimizing the temperature difference. Samples for prokaryote abun- Maximum cell density in nitrate + phosphate occurred at t3, while maximum
dance, inorganic nutrients and bacterial community structure were densities for ammonium + phosphate occurred at t4.
collected prior to filtering to characterize parameters in the original
seawater. The initial prokaryote abundance in the seawater was 52.2 ± 0.7 to 28.6 ± 2.4. Shannon diversity decreased from t0 to t2
7.61 × 105 cells mL−1, while estimates for all treatments at t0 averaged and then remained relatively constant through to the end of the ex-
2.66 × 105 cells mL−1 ± 6.02 × 104. Thus, the density of prokaryotes periment. Oleispira spp., Glaciecola spp., Sediminicola spp. and
at the beginning of the experiment was only ~35% of the in situ con- Colwelliaceae increased significantly in the control flasks compared to
centrations. t0. Thirty-three taxa decreased significantly over the five days.
At t0, nitrate was lower in the control and Dilbit treatments com-
pared to in situ values, while phosphate was higher and ammonium was
similar to in situ values. Nutrient additions resulted in significantly 3.2. AWB and Nutrient additions
higher nitrate, phosphate or ammonium, depending on the treatment
(Table 1). The N:P ratio was lower than in situ for all treatments except Analysis of O2 concentrations in the headspace identified significant
for the nitrate-only and ammonium-only additions. differences over time (p < 0.001) and among treatments (p < 0.001),
Microbial diversity in the flasks was similar to that in the seawater. but the interaction term was not significant. Generally, O2 decreased
PD averaged 51.7 ± 1.5 in the flasks compared to 49.2 in the seawater, from t0 to t2, but increased at t3 before decreasing by t5 (Fig. S1).
while Shannon diversity in the flasks averaged 7.16 ± 0.04, with 7.16 Holm-Šídák post hoc tests indicated that all time points were sig-
measured in the seawater. The structure of the microbial communities nificantly different from t0. All nutrient treatments except Nitrate were
was slightly different in the flasks compared to the initial seawater determined to have significantly lower O2 concentrations compared to
based on SIMPER analysis using the Bray-Curtis dissimilarity metric the Dilbit treatment.
(average dissimilarity = 21.21%). The 7 OTUs contributing the most to Cells increased in all treatments over the duration of the experi-
the differences between samples included 4 OTUs with higher average ment, with abundances in several treatments exceeding
abundances in the experimental flasks. These OTUs included 2.0 × 106 cells mL−1 (Fig. 1). Abundances in the nitrate + phosphate
Saprospiraceae (flasks = 2.6%, seawater = 1.4%) and treatment peaked at t3 and decreased through the remainder of the
Flavobacteriaceae (Bacteroidetes) (flasks = 4.5%, seawater = 3.3%) experiment. In the ammonium + phosphate flasks, abundances peaked
and two Rhodobacteraceae (Alphaproteobacteria) OTUs, one of which at t4. The other treatments appeared to still be increasing at the end of
was identified as the genus Octadecabacter (flasks = 7.1%, sea- the experiment. Two-way ANOVA of cell abundances identified a sig-
water = 6.1%) and the other that was only identified to the family level nificant interaction between time and treatment (p < 0.001). Holm-
(flasks = 1.7%, seawater = 0.9%). Two Pelagibacteraceae Šídák post hoc tests indicated that for dilbit and ammonium, cells did
(Alphaproteobacteria) OTUs (flasks = 1.8%, seawater = 3.7%) along not grow significantly until after t2, as neither t1 nor t2 were sig-
with the OCS155 family of the Acidimicrobiales (flasks = 1.5%, sea- nificantly different from t0. For all other treatments, t0 and t1 were not
water = 2.3%) had lower abundances in the experimental flasks com- significantly different, but t2–t5 were higher than t0 suggesting a
pared to the whole seawater. Pelagibacteraceae are generally small cells shorter lag period. Significantly higher cell densities were detected at t2
and may have passed easily through the filter (Rappe et al., 2002). (nitrate + phosphate), t3 (nitrate + phosphate) and t4 (ammo-
Some marine Acidimicrobiales may also be small enough to pass nium + phosphate) compared to the dilbit treatment.
through the filter (Ghai et al., 2013). The four most dominant OTUs Inorganic nutrients differed due to treatment by design, so 1-way
were the same in both the flasks and the seawater and included three ANOVAs were carried out for each treatment. Even where concentra-
Rhodobacteraceae, including two Octadecabacter spp., and Flavo- tions were low, nutrients were never depleted during the experiment
bacteriaceae. (Fig. 2). In the dilbit treatment, nitrate decreased between t0 and t1,
but then increased with t1 significantly lower than t4 and t5
3.1. Control treatment (p = 0.008). Phosphate showed no change over time (p = 0.085), but
similar to the controls, ammonium increased between t2 and t3, with
In the control treatment, O2 concentrations in the headspace re- significantly higher concentrations throughout the end of the experi-
mained relatively constant over time with an increase at t3, corre- ment (p < 0.001).
sponding to a decrease in incubation temperature. Cell numbers in- For the nitrate treatment, neither nitrate (p = 0.538) nor phosphate
creased over the experiment, with a final average abundance of (p = 0.288) showed change over time, but ammonium increased be-
2.66 × 106 cell mL−1 ± 7.64 × 105, higher than in the initial sea- tween t1 and t2, with significantly higher concentrations from t3
water. During the experiment, the concentration of inorganic nutrients through t5 relative to t0 and t1 (p < 0.001). The ammonium treatment
changed little, with the exception of ammonium, which increased ~3 had a similar nitrate pattern as dilbit alone, decreasing by t1 and then
fold between t2 and t3. Microbial community diversity decreased in the increasing at t4 and t5 (p = 0.002). A Kruskal-Wallis ANOVA on ranks
control flasks over time, with PD showing a gradual decrease from indicated significant differences in phosphate concentrations due to

384
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

Fig. 2. Concentrations of inorganic nutrients in the flasks over time with values in μM. The top panel includes treatments where additional nutrients were added. The
bottom panel includes treatments where concentrations began at the concentration in the ESAW.

time (p = 0.022); however, Tukey post hoc pairwise tests did not detect treatment. By t3, there were no longer significant differences among
differences among time points. The ammonium treatment, although treatments. Looking at overall community structure, there was a clear
starting with much higher ammonium concentrations, also had in- shift in all treatments over time. Two-way ANOSIM analysis using the
creases between t2 and t3, resulting in significantly higher concentra- Bray-Curtis similarity metric identified significant differences between
tions at t3, t4 and t5 compared to t1 (p = 0.004). The phosphate all pairs of time points (Global R = 0.892, p = 0.001,
treatment had no significant change in concentrations of phosphate ppairwise = 0.464–1.00, p = 0.001). Treatment was also significant
(p = 0.631) or ammonium (p = 0.118), but nitrate concentrations at t5 (Global R = 0.499, p = 0.001), but not all pairwise comparisons were
were significantly higher than at t0 (p = 0.010). No change in the significant resulting in two groups of samples. Dilbit, nitrate and am-
phosphate concentration was detected in the nitrate + phosphate monium samples formed one group (D-N-A), while phosphate, ni-
treatment (p = 0.631). Both nitrate and ammonium increased in this trate + phosphate and ammonium + phosphate (P-NP-AP) made up
treatment with all time points significantly higher than t0 for nitrate the second group. The most obvious difference in these groups was at t1
(p = 0.002) and t2 through t5 significantly higher than t0 and t1 for and t2, where the two groups formed separate clusters (Fig. 4). By t3,
ammonium (p < 0.001). The ammonium + phosphate treatment had the difference in the two groups was reduced, although differences in
significantly higher nitrate concentrations at t5 compared to t1 specific taxa could be detected at t3 and t4. Using STAMP, six taxa were
(p = 0.031), but neither phosphate (p = 0.754) nor ammonium determined to be significantly different between P-NP-AP and D-N-A at
(p = 0.115) showed changes over time. t1. In the P-NP-AP group, Colwelliaceae, Glaciecola spp., Sediminicola
The diversity of the microbial community decreased throughout the spp. and Rhodobacteraceae averaged 2.6, 4.8, 1.2 and 2.7% higher,
experiment in all treatments (Fig. 3). Two-way ANOVA indicated a respectively, when compared to the D-N-A group (Fig. 5). In contrast,
significant interaction between time and treatment for both PD and Pelagibacteriaceae and Flavobacteriaceae were 2.5 and 1.5% higher,
Shannon diversity (pPD = 0.001, pShannon < 0.001). Post hoc tests in- respectively, in the samples without additional phosphate. By t2, only
dicated that the diversity in phosphate, nitrate + phosphate and am- two taxa were found to be significantly different, with Rhodobacter-
monium + phosphate was significantly lower at t1 and t2 compared to aceae 7.4% higher in the phosphate group, while Flavobacteriaceae was
the dilbit treatment indicating a faster decrease compared to the nitrate 1.1% higher without phosphate. Two taxa were significantly different
and ammonium treatments which did not differ from the dilbit between groups at t3, with 4.6% more Colwelliaceae in treatments

Fig. 3. Bacterial and archaeal diversity decreased over time, with significantly lower diversity in phosphate, nitrate + phosphate and ammonium + phosphate
treatments compared to dilbit at t1 and t2. Shannon diversity (A) and Phylogenetic distance (B) show the same patterns.

385
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

Fig. 6. Plot of the first two PCA axes from analysis of the GC-FID spectra. Points
are coloured based on time from dark (t0) to light (t5), with shapes indicating
treatments.
Fig. 4. PCO plot of bacterial and archaeal communities over time based on the
Bray-Curtis dissimilarity metric. Colours indicate time points from dark (t0) to concentrations compared to t0. Compared to dilbit alone, neither ni-
light (t5). Treatments are grouped based on ANOSIM results, which groups trate nor phosphate was significantly different, while ammonium, ni-
treatments with additional phosphate together (P-NP-AP) separate from the trate + phosphate and ammonium + phosphate generally had higher
other treatments (D-N-A).
relative TPH concentrations. Analysis of TPH spectra with PCA in-
dicated a change in the composition of the oil during the experiment.
without additional phosphate and 6.8% more Glaciecola spp. in the There was no pattern due to treatment, but samples generally became
additional phosphate treatments. At t4, Colwelliaceae and Rhodo- more positive along the PCA Axis 1, but more negative on PCA Axis 2,
bacteraceae were 5.3 and 5.0% higher, respectively, in the treatments indicating a change in composition over time (Fig. 6). Measured with
without phosphate. Glaciecola spp. was 8.9% higher with phosphate. GC–MS, normalized alkylated PAHs and PAHs showed no significant
Averaged across groups, Rhodobacteraceae was the most abundant taxa decrease over the five days, regardless of the treatment. p-Values for the
in all treatments at all times, except for the P-NP-AP samples at t5, t-tests for the significance of the slopes ranged from 0.173 to 0.973.
where Glaciecola spp. was higher on average. In all of the oiled treat- Alkylated PAHs averaged 8.52 ± 0.65 at t0 and 9.12 ± 0.91 at t5
ments, Colwelliacea, Octadecabacter spp., and the Gammaproteo- when normalized to hopane. For PAHs, the means were 0.63 ± 0.06 at
bacteria genus HTC (family HTCC2188) rounded out the five most t0 and 0.64 ± 0.04 at t5. For most treatments, the slope of the re-
abundant taxa. Of all the taxa that changed significantly among time gression for alkanes was not significantly different from zero (p-va-
points (nD-N-A = 34 and nP-NP-AP = 41), only three showed significant lues = 0.157–0.329), suggesting minimal degradation occurred. How-
increases over time. Those taxa were Colwelliaceae, Rhodobacteraceae ever, in the nitrate + phosphate and ammonium + phosphate
and Glaciecola spp. (Fig. 5). These three taxa have been associated with treatments, slopes were significant (p = 0.068 and p = 0.070, respec-
hydrocarbon degradation. Other abundant taxa linked with hydro- tively), and the alkane degradation rate for nitrate + phosphate was
carbon degradation included Gammaproteobacteria in the HTCC2188 determined to be 0.0230 d−1 while the rate for ammonium + phos-
family and the genus HTCC2207. phate was 0.0235 d−1.
TPH concentrations at t0 averaged 120.41 μg mL−1 ± 15.94 (Fig.
S2). Because of this variation, the concentrations were divided by the
4. Discussion
average for each treatment at t0 to produce relative TPH concentrations
and two-way ANOVA was carried out using these values. The analysis
The addition of fresh diluted bitumen to a natural coastal microbial
detected significant differences due to both time (p < 0.001) and
community resulted in microbial growth and a shift in the community
treatment (p < 0.001), but the interaction term was not significant
towards known hydrocarbon degrading organisms. The total amount of
(p = 0.461). t2 and t3 had significantly higher relative TPH
petroleum hydrocarbons did not decrease significantly over the short

Fig. 5. Mean relative abundances of taxa of abundant taxa, defined as those with mean abundances ≥0.5% of the total microbial community. Taxa discussed in the
text are highlighted in the legend. Samples are grouped based on ANOSIM analysis as in Fig. 4. A full list of taxa with mean relative abundances and standard
deviations are provided in Table S1.

386
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

time frame of the experiment, but changes in the profile of the GC-FID changed. At t1 there is a clear difference between the treatments
spectra indicated that degradation occurred (Fig. 6). This was sup- amended with phosphate and those that were not (Fig. 5). The differ-
ported by GC–MS analysis in two treatments, nitrate + phosphate and ence in these two clusters is mainly due to larger increases in the re-
ammonium + phosphate, where significant decreases in alkanes were lative abundance of Glaciecola spp., Colwelliaceae and Rhodobacter-
measured over the five days of the experiment. While the patterns ob- iaceae with additional phosphate (Fig. 5). While all three taxa increased
served for TPH and alkanes may appear contradictory, differences in in the treatments without excess phosphate relative to t0, the increases
the methods can explain why this is not true. GC–MS data was stan- were smaller. Glaciecola spp. have previously been observed to increase
dardized relative to hopane concentrations. Thus, the decrease in al- in North Atlantic waters in response to oil additions (Tremblay et al.,
kanes is relative to the stable marker, hopane. This is in contrast to the 2017) and has been documented in hydrocarbon contaminated cold
TPH data, where concentrations are not standardized with a marker, waters, although in a mesocosm experiment Glaciecola spp. was more
but represent total extractable hydrocarbons in the sample. Ad- abundant in the control treatments (Chronopoulou et al., 2015). In the
ditionally, GC–MS is more sensitive than GC-FID. Based on the calcu- present experiment, Glaciecola spp. increased with oil at t1, and was one
lated degradation rates for alkanes in nitrate + phosphate and ammo- of the most abundant taxon at t5, with average relative abundances of
nium + phosphate treatments, over the course of 5 days, there was a 22.9 and 30.3% in dilbit treatment without and with additional phos-
loss of ~41 ng mL−1 of alkanes against a background TPH concentra- phate, respectively. In the controls at t5, Glaciecola spp. was the second
tion of > 100 μg mL−1. most abundant component of the community, but only reached average
The degradation rates calculated in this experiment when both ni- abundances of 18.8%.
trogen and phosphate was added fell between previous estimates for At t5, the control flasks were dominated by organisms that are as-
dilbit degradation rates (Deshpande et al., 2018; King et al., 2014). The sociated with hydrocarbon degradation, including Rhodobacteraceae,
measured rates were ten-fold higher than the rates observed by King Oleispira spp. and Colwelliaceae (Jimenez et al., 2007; Ortmann and Lu,
et al. (2014) with a natural marine microbial community at ~18 °C 2015; Ribicic et al., 2018; Yakimov et al., 2003), although not all
without added nutrients. With the addition of dilbit alone, no sig- Rhodobacteraceae or Colwelliaceae degrade oil and are common
nificant decrease in alkanes was observed, suggesting the additional members of coastal microbial communities (Bowman, 2014; Elifantz
nitrogen and phosphorous stimulated hydrocarbon degradation. Even et al., 2013; Ortmann and Ortell, 2014). Sediminicola spp. were also
with additional inorganic nutrients, the rates observed here were still abundant in the controls at t5, but were also present in the dilbit treated
ten-fold lower than the rates observed for incubations of fresh dilbit and samples, although they averaged only 2.1% vs. 5.3% in the controls.
excess inorganic nutrients with microbial communities enriched from Sediminicola spp. (identified as NS3a marine group of the Flavobacter-
the Kalamazoo River at 5 °C (Deshpande et al., 2018). The main dif- iaceae in the Silva database (Yilmaz et al., 2014)), are commonly as-
ference between studies is the prior exposure of the river sediment sociated with blooms and likely capable of degrading complex organic
community to dilbit, whereas the present study used a community that matter (Kirchman, 2002; Yang et al., 2015). This group may have been
had not previously been exposed to dilbit. While some regions may responsible for the increase in ammonium in many of the treatments,
have previous or continuous exposure to oil over time, priming a re- due to remineralization of organic matter present in the incubations.
sponse to a large spill, other areas may be more pristine. Thus, lower Along with Rhodobacteraceae, other taxa associated with hydro-
degradation rates are likely to occur following a dilbit spill in marine carbon degradation were abundant in the oil exposed flasks at t5.
environments that have not been previously exposed to oil. Octadecabacter spp., within the family Rhodobacteraceae, has pre-
The short time frame of this experiment does not allow us to de- viously been associated with oil contamination under cold conditions
termine if all of the alkanes would have been degraded given sufficient (Brakstad et al., 2008). This organism increased by t1 in the presence of
time. It is also not clear if the advantage of additional nitrogen and oil under all treatments relative to t0, but decreased back down to
phosphorous would have accelerated the removal rates over the long starting abundances by t5. Other taxa included Gammaproteobacteria
term. There are some indications that this would not have occurred. identified as HTCC2207 (SAR92) and HTCC2188 (OM182), which in-
Looking at the microbial community itself, by t5 there was no differ- creased in the presence of oil with larger increases in treatments with
ence in the community composition among treatments (Fig. 5) and additional phosphate (Fig. 5). By t5 HTCC2188 increased 1.3× with
while growth in the nitrate + phosphate and ammonium + phosphate dilbit, and 1.8× with dilbit and additional phosphate. HTCC2207
treatments had decreased and was showing declining cell numbers, was < 0.5% of the community at t0, but increased to 0.9 and 1.2% with
growth was still occurring in the other treatments (Fig. 1). Together, dilbit and with dilbit and phosphate, respectively. A draft genome of
these patterns suggest that a hydrocarbon degrading community de- HTCC2207 (Alteromonadales, Gammaproteobacteria) includes anno-
veloped in all treatments with dilbit regardless of inorganic nutrient tations for a putative alkane monooxygenase gene (Stingl et al., 2007),
concentrations. Additionally, looking at the chemical data, there was no suggesting these organisms can degrade alkanes (Rojo, 2009).
significant difference in TPH concentrations among treatments at t5. HTCC2188 has been detected previously in coastal waters (Liu et al.,
While alkane degradation rates were determined for two of the treat- 2016), but why it responded to dilbit is unclear. This organism is
ments, PAHs and alkylated PAHs did not change. Together, these three identified as either belonging to the OM182 of the Altermonadales
groups of compounds make up approximately 45% of AWB (King et al., (Greengenes, McDonald et al., 2012) or Pseudohongiella spp. in the
2014), with the rest composed of resins and asphaltenes, which are Oceanospirillales (SILVA, Yilmaz et al., 2014). Both of these Gamma-
even more resistant to biodegradation. Thus, it is not surprising that proteobacteria families are associated with hydrocarbon degradation
TPH showed little change over the experiment. While TPH did not (e.g. Kostka et al., 2011; Ortmann and Lu, 2015; Redmond and
change, the composition of the oil as measured by GC-FID did show a Valentine, 2012), so it is possible that HTCC2188 may share this ability.
clear change over time, but no clear difference was observed among The addition of dilbit to a natural microbial community that had not
treatments. This suggests that weathering of the oil was occurring in all previously been exposed to oiling resulted in a rapid shift in the mi-
of the treatments. However, we cannot discount that some of the crobial community. Bacteria commonly associated with both cold water
weathering was abiotic, which should not be impacted by inorganic and hydrocarbons increased, with faster increases in treatments
nutrients. amended with phosphate, with or without inorganic nitrogen. Cell
Although the experiment was carried out at low temperatures, the abundances increased fastest in the ammonium + phosphate and ni-
microbial community structure changed rapidly, with the largest shifts trate + phosphate treatments, in which alkane degradation rates could
between t0 and t1 as well as t1 and t2. After t2 shifts in community be calculated. Although the rates measured here were lower than ob-
structure were smaller. The addition of phosphate, either alone or with served for a freshwater microbial community enriched on dilbit
a nitrogen source appeared to increase the rate at which the community (Deshpande et al., 2018), they were higher than rates measured for a

387
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

coastal community in the absence of additional inorganic nutrients medium for coastal and open ocean phytoplankton. J. Phycol. 16, 28–35.
(King et al., 2014). The results observed here suggest that there may be Hazen, T.C., Prince, R.C., Mahmoudi, N., 2016. Marine oil biodegradation. Environ. Sci.
Technol. 50, 2121–2129.
an initial benefit to increasing inorganic nutrients to stimulate biode- Head, I.M., Jones, D.M., Röling, W.F.M., 2006. Marine microorganisms make a meal of
gradation of dilbit, but in the short time frame of this experiment, the oil. Nat. Rev. Microbiol. 4, 173–182.
advantage was minimal and disappeared by t5. Under natural condi- Hori, T., Kimura, M., Aoyagi, T., Navarro, R., Ogata, A., Sakoda, A., Katayama, Y.,
Takasaki, M., 2014. Biodegradation potential of organically enriched sediments
tions, where water is constantly moving and resupplying nutrients, it is under sulfate- and iron-reducing conditions as revealed by the 16S rRNA deep se-
unlikely that a coastal community at low temperature would become quencing. J. Water Environ. Technol. 12, 357–366.
nutrient limited; hence, there would be no advantage to adding in- Jimenez, N., Vinas, M., Bayona, J.M., Albaiges, J., Solanas, A.M., 2007. The Prestige oil
spill: bacterial community dynamics during a field biostimulation assay. Appl.
organic nutrients in an attempt to increase the rate of dilbit biode- Microbiol. Biotechnol. 77, 935–945.
gradation. Increases in inorganic nutrients during this experiment King, T.L., Robinson, B., Boufadel, M., Lee, K., 2014. Flume tank studies to elucidate the
suggest that organic sources of nutrients may be used by microbial fate and behavior of diluted bitumen spilled at sea. Mar. Pollut. Bull. 83, 32–37.
King, T.L., Robinson, B., McIntyre, C., Toole, P., Ryan, S., Saleh, F., Boufadel, M.C., Lee,
communities during the biodegradation of hydrocarbons.
K., 2015. Fate of surface spills of Cold Lake Blend diluted bitumen treated with
Supplementary data to this article can be found online at https:// dispersant and mineral fines in a wave tank. Environ. Eng. Sci. 32, 250–261.
doi.org/10.1016/j.marpolbul.2019.01.012. Kirchman, D.L., 2002. The ecology of Cytophaga–Flavobacteria in aquatic environments.
FEMS Microbiol. Ecol. 39, 91–100.
Kopylova, E., Noe, L., Touzet, H., 2012. SortMeRNA: fast and accurate filtering of ribo-
Acknowledgements somal RNAs in metatranscriptomic data. Bioinformatics 28, 3211–3217.
Kostka, J.E., Prakash, O., Overholt, W.A., Green, S.J., Freyer, G., Canion, A., Delgardio, J.,
Thanks to S. Ryan and B. Robertson for technical and logistical as- Norton, N., Hazen, T.C., Huettel, M., 2011. Hydrocarbon-degrading bacteria and the
bacterial community response in Gulf of Mexico beach sands impacted by the
sistance. C. Anstey processed inorganic nutrient samples. This research Deepwater Horizon oil spill. Appl. Environ. Microbiol. 77, 7962–7974.
was supported through the Oceans Protection Plan program of the Li, W.K.W., 2014. The state of phytoplankton and bacterioplankton at the Compass Buoy
Government of Canada. Station: Bedford Basin monitoring program 1992–2013. In: Canadian Technical
Report of Hydrography and Ocean Sciences, Ottawa, pp. 122.
Liu, S., Luo, Y., Huang, L., 2016. Dynamics of size-fractionated bacterial communities
References during the coastal dispersal of treated municipal effluents. Appl. Microbiol.
Biotechnol. 100, 5839–5848.
MacDonald, S., 2017. Findings from the “Long-term Persistence of Stranded Oil
Atlas, R.M., Hazen, T.C., 2011. Oil biodegradation and bioremediation: a tale of the two
Workshop, Nova Scotia”. In: 40th AMOP Technical Seminar on Environmental
worst spills in U.S. history. Environ. Sci. Technol. 45, 6709–6715.
Contamination and Response, Calgary.
Bagby, S.C., Reddy, C.M., Aeppli, C., Fisher, G.B., Valentine, D.L., 2017. Persistence and
McDonald, D., Price, M.N., Goodrich, J., Nawrocki, E.P., DeSantis, T.Z., Probst, A.,
biodegradation of oil at the ocean floor following Deepwater Horizon. Proc. Natl.
Andersen, G.L., Knight, R., Hugenholtz, P., 2012. An improved Greengenes taxonomy
Acad. Sci. U. S. A. 114, E9–E18.
with explicit ranks for ecological and evolutionary analyses of Bacteria and Archaea.
Berges, J.A., Franklin, D.J., Harrison, P.J., 2001. Evolution of an artificial seawater
ISME J. 6, 610–618.
medium: improvements in enriched seawater, artificial water over the last two dec-
Mercier, C., Boyer, F., Bonin, A., Coissac, E., 2013. SUMATRA and SUMACLUST: fast and
ades. J. Phycol. 37, 1138–1145.
exact comparison and clustering of sequences. In: Programs and Abstracts of the
Bokulich, N.A., Subramanian, S., Faith, J.J., Gevers, D., Gordon, J.I., Knight, R., Mills,
SeqBio 2013 Workshop. Abstract, pp. 27–29.
D.A., Caporaso, J.G., 2013. Quality-filtering vastly improves diversity estimates from
Mortazavi, B., Horel, A., Beazley, M.J., Sobecky, P.A., 2013. Intrinsic rates of petroleum
Illumina amplicon sequencing. Nat. Methods 10, 57–59.
hydrocarbon biodegradation in Gulf of Mexico intertidal sandy sediments and its
Bowman, J.P., 2014. The family Colwelliaceae. In: Rosenberg, E., DeLong, E.F., Lory, S.,
enhancement by organic substrates. J. Hazard. Mater. 244-245, 537–544.
Stackebrandt, E., Thompson, F. (Eds.), The Prokaryotes: Gammaproteobacteria.
Oksanen, J., Blanchet, F.G., Friendly, M., Kindt, R., Legendre, P., McGlinn, D., Minchin,
Springer, Berlin, pp. 179–195.
P.R., O'Hara, R.B., Simpson, G.L., Solymos, P., Stevens, M.H.H., Szoecs, E., Wagner,
Brakstad, O.G., Nonstad, I., Faksness, L.G., Brandvik, P.J., 2008. Responses of microbial
H., 2018. vegan: Community Ecology Package. R Package Version 2.4-6.
communities in Arctic sea ice after contamination by crude petroleum oil. Microb.
Ortmann, A.C., Lu, Y., 2015. Initial community and environment determine the response
Ecol. 55, 540–552.
of bacterial communities to dispersant and oil contamination. Mar. Pollut. Bull. 90,
Bushnell, L.D., Haas, H.F., 1941. The utilization of certain hydrocarbons by micro-
106–114.
organisms. J. Bacteriol. 41, 653–673.
Ortmann, A.C., Ortell, N., 2014. Changes in free-living bacterial community diversity
Caporaso, J.G., Kuczynski, J., Stombaugh, J., Bittinger, K., Bushman, F.D., Costello, E.K.,
reflect the magnitude of environmental variability. FEMS Microbiol. Ecol. 87,
Fierer, N., Pena, A.G., Goodrich, J.K., Gordon, J.I., Huttley, G.A., 2010. QIIME allows
291–301.
analysis of high-throughput community sequencing data. Nat. Methods 7, 335–336.
Ortmann, A.C., Anders, J., Shelton, N., Gong, L., Moss, A.G., Condon, R.H., 2012.
CAPP (Canadian Association of Petroleum Producers), 2017. 2017 CAPP Crude Oil
Dispersed oil disrupts microbial pathways in pelagic food webs. PLoS One 7, e42548.
Forecast, Markets and Transportation. Calgary. pp. 54.
Owsianiak, M., Chrzanowski, L., Szulc, A., Staniewski, J., Olszanowski, A., Olejnik-
Chronopoulou, P.M., Sanni, G.O., Silas-Olu, D.I., van der Meer, J.R., Timmis, K.N.,
Schmidt, A.K., Heipieper, H.J., 2009. Biodegradation of diesel/biodiesel blends by a
Brussaard, C.P., McGenity, T.J., 2015. Generalist hydrocarbon-degrading bacterial
consortium of hydrocarbon degraders: effect of the type of blend and the addition of
communities in the oil-polluted water column of the North Sea. Microb. Biotechnol.
biosurfactants. Bioresour. Technol. 100, 1497–1500.
8, 434–447.
Parks, D.H., Tyson, G.W., Hugenholtz, P., Beiko, R.G., 2014. STAMP: statistical analysis of
Cole, M.G., King, T.L., Lee, K., 2007. Analytical technique for extraction hydrocarbons
taxonomic and functional profiles. Bioinformatics 30, 3123–3124.
from water using sample container as extraction vessel in combination with a roller
Prince, R.C., Elmendorf, D.L., Lute, J.R., Hsu, C.S., Haith, C.E., Senius, J.D., Dechert, G.J.,
apparatus. In: Canadian Technical Report of Fisheries and Aquatic Sciences. Fisheries
1994. 17alpha(H),21beta(H)-Hopane as a conserved internal marker for estimating
and Oceans Canada, Ottawa, pp. 19.
the biodegradation of crude oil. Environ. Sci. Technol. 28, 142–145.
Comeau, A.M., Douglas, G.M., Langille, M.G.I., 2017. Microbiome helper: a custom and
Prince, R.C., Garrett, R.M., Bare, R.E., Grossman, M.J., Townsend, T., Suflita, J.M., Lee,
streamlined workflow for microbiome research. mSystems 2.1, e00127-16.
K., Owens, E.H., Sergy, G.A., Braddock, J.F., Lindstrom, J.E., Lessard, R.R., 2003. The
Crosby, S., Fay, R., Groark, C., Smith, J.R., Sullivan, T., Pavia, R., Shigenaka, G., 2013.
roles of photooxidation and biodegradation in long-term weathering of crude and
Transporting Alberta oil sands products: defining the issues and assessing the risks.
heavy fuel oils. Spill Sci. Technol. Bull. 8, 145–156.
In: NOAA Technical Memorandum NOS OR&R 44. Department of Commerce, Seattle,
R Core Team, 2017. R: A Language and Environment for Statistical Computing. R
pp. 153.
Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/.
Deshpande, R.S., Sundaravadivelu, D., Techtmann, S., Conmy, R.N., Santo Domingo,
Rappe, M.S., Connon, S.A., Vergin, K.L., Giovannoni, S.J., 2002. Cultivation of the ubi-
J.W., Campo, P., 2018. Microbial degradation of Cold Lake Blend and Western
quitous SAR11 marine bacterioplankton clade. Nature 418, 630–633.
Canadian select dilbits by freshwater enrichments. J. Hazard. Mater. 352, 111–120.
Redmond, M.C., Valentine, D.L., 2012. Natural gas and temperature structured a micro-
Elifantz, H., Horn, G., Ayon, M., Cohen, Y., Minz, D., 2013. Rhodobacteraceae are the key
bial community response to the Deepwater Horizon oil spill. Proc. Natl. Acad. Sci. U. S.
members of the microbial community of the initial biofilm formed in Eastern
A. 109, 20292–20297.
Mediterranean coastal seawater. FEMS Microbiol. Ecol. 85, 348–357.
Ribicic, D., Netzer, R., Hazen, T.C., Techtmann, S.M., Drabløs, F., Brakstad, O.G., 2018.
EPA (US Environmental Protection Agency), 1996. EPA Method 3630C: Silica Gel
Microbial community and metagenome dynamics during biodegradation of dispersed
Cleanup, Part of Test Methods for Evaluating Solid Waster, Physical/Chemical
oil reveals potential key-players in cold Norwegian seawater. Mar. Pollut. Bull. 129,
Methods. Washington, DC. pp. 15.
370–378.
EPA (US Environmental Protection Agency), 2007. EPA Method 8270D: Semivolatile
Rognes, T., Flouri, T., Nichols, B., Quince, C., Mahe, F., 2016. VSEARCH: a versatile open
Organic Compounds by Gas Chromatography/Mass Spectrometry. Washington, DC.
source tool for metagenomics. PeerJ 4, e2584.
pp. 71.
Rojo, F., 2009. Degradation of alkanes by bacteria. Environ. Microbiol. 11, 2477–2490.
Ghai, R., Mizuno, C.M., Picazo, A., Camacho, A., Rodriguez-Valera, F., 2013.
Stingl, U., Desiderio, R.A., Cho, J.C., Vergin, K.L., Giovannoni, S.J., 2007. The SAR92
Metagenomics uncovers a new group of low GC and ultra-small marine
clade: an abundant coastal clade of culturable marine bacteria possessing pro-
Actinobacteria. Sci. Rep. 3, 2471.
teorhodopsin. Appl. Environ. Microbiol. 73, 2290–2296.
Harrison, P.J., Waters, R.E., Taylor, F.J.R., 1980. A broad spectrum artificial seawater
Tremblay, J., Yergeau, E., Fortin, N., Cobanli, S., Elias, M., King, T.L., Lee, K., Greer, C.W.,

388
A.C. Ortmann et al. Marine Pollution Bulletin 139 (2019) 381–389

2017. Chemical dispersants enhance the activity of oil- and gas condensate-degrading Wong, J.W., Cagney, G., Cartwright, H.M., 2005. SpecAlign—processing and alignment of
marine bacteria. ISME J. 11, 2793–2808. mass spectra datasets. Bioinformatics 21, 2088–2090.
Valentine, D.L., Kessler, J.D., Redmond, M.C., Mendes, S.D., Heintz, M.B., Farwell, C., Hu, Yakimov, M.M., Giuliano, L., Gentile, G., Crisafi, E., Chernikova, T.N., Abraham, W.R.,
L., Kinnaman, F.S., Yvon-Lewis, S., Du, M., 2010. Propane respiration jump-starts Lunsdorf, H., Timmis, K.N., Golyshin, P.N., 2003. Oleispira antarctica gen. nov., sp.
microbial response to a deep oil spill. Science 330, 208. nov., a novel hydrocarbonoclastic marine bacterium isolated from Antarctic coastal
Vandermeulen, J.H., Singh, J.G., 1994. Arrow oil spill, 1970–90: persistence of 20-year sea water. Int. J. Syst. Evol. Microbiol. 53, 779–785.
weathered Bunker C fuel oil. Can. J. Fish. Aquat. Sci. 51, 845–855. Yang, C., Li, Y., Zhou, B., Zhou, Y., Zheng, W., Tian, Y., Van Nostrand, J.D., Wu, L., He, Z.,
Venosa, A.D., Suidan, M.T., King, D., Wrenn, B.A., 1997. Use of hopane as a conservative Zhou, J., Zheng, T., 2015. Illumina sequencing-based analysis of free-living bacterial
biomarker for monitoring the bioremediation effectiveness of crude oil con- community dynamics during an Akashiwo sanguine bloom in Xiamen Sea, China. Sci.
taminating a sandy beach. J. Ind. Microbiol. Biotechnol. 18, 131–139. Rep. 5, 8476.
Walters, W., Hyde, E.R., Berg-Lyons, D., Ackermann, G., Humphrey, G., Parada, A., Yilmaz, P., Parfrey, L.W., Yarza, P., Gerken, J., Pruesse, E., Quast, C., Schweer, T., Peplies,
Gilbert, J.A., Jansson, J.K., Caporaso, J.G., Fuhrman, J.A., Apprill, A., Knight, R., J., Ludwig, W., Glockner, F.O., 2014. The SILVA and “All-species Living Tree Project
2016. Improved bacterial 16S rRNA gene (V4 and V4-5) and fungal internal tran- (LTP)” taxonomic frameworks. Nucleic Acids Res. 42, D643–D648.
scribed spacer marker gene primers for microbial community surveys. mSystems 1 Zhang, J., Kobert, K., Flouri, T., Stamatakis, A., 2014. PEAR: a fast and accurate Illumina
(2), e00009–e00015. Paired-End reAd mergeR. Bioinformatics 30, 614–620.

389

You might also like