<<

Anthropocene 13 (2016) 18–33

Contents lists available at ScienceDirect

Anthropocene

journa l homepage: www.elsevier.com/locate/ancene

Review

Assessing the impacts of oil-associated formation and

sedimentation during and after the Deepwater Horizon

a, b c a

Kendra L. Daly *, Uta Passow , Jeffrey Chanton , David Hollander

a

College of Marine Science, University of South , St. Petersburg, FL 33705,

b

Marine Science Institute, University of California, Santa Barbara, CA 93106, United States

c

Earth, , Atmospheric Science, Florida State University, Tallahassee, FL 32306, United States

A R T I C L E I N F O A B S T R A C T

Article history: The Deepwater Horizon oil spill was the largest in US history, unprecedented for the depth and volume of

Received 20 October 2015

oil released, the amount of dispersants applied, and the unexpected, protracted sedimentation of oil-

Received in revised form 27 January 2016

associated marine snow (MOS) to the seafloor. Marine snow formation, incorporation of oil, and

Accepted 28 January 2016

subsequent gravitational settling to the seafloor (i.e., MOSSFA: Marine Oil Snow Sedimentation and

Available online 2 February 2016

Flocculent Accumulation) was a significant pathway for the distribution and fate of oil, accounting for as

much as 14% of the total oil released. Long residence times of oil on the seafloor will result in prolonged

Keywords:

exposure by benthic organisms and economically important fish. of hydrocarbons into

Deepwater Horizon oil spill

the also has been documented. Major surface processes governing the MOSSFA event included

Gulf of Mexico

an elevated and extended Mississippi River discharge, which enhanced production and

Marine oil snow

MOSSFA suspended concentrations, grazing, and enhanced microbial formation.

Bacteria Previous reports indicated that MOS sedimentation also occurred during the Tsesis and Ixtoc-I oil spills;

Plankton thus, MOSSFA events may occur during future oil spills, particularly since 85% of global deep-water oil

exploration sites are adjacent to deltaic systems. We provide a conceptual framework of MOSSFA

processes and identify data gaps to help guide current research and to improve our ability to predict

MOSSFA events under different environmental conditions. Baseline time-series data and model

development are urgently needed for all levels of in regions of hydrocarbon extraction to

prepare for and respond to future oil spills and to understand the impacts of oil spills on the environment.

ã 2016 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND

license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Contents

1. Introduction ...... 19

2. Characterization of oil-associated marine snow: processes and pathways ...... 20

2.1. Marine snow distribution and sedimentation patterns ...... 20

2.2. Factors affecting MOSSFA processes ...... 21

2.3. Field and lab MOS formation ...... 22

2.4. Microbial processes ...... 22

2.5. impacts on MOS ...... 23

2.6. Oil-mineral aggregates (OMAs) ...... 24

2.7. MOS sedimentation ...... 24

3. Temporal and spatial patterns of MOS ...... 24

4. Deep oil plumes ...... 25

5. Oil accumulation rates and fate at the seafloor ...... 26

6. effects ...... 27

7. Proposed framework for MOSSFA investigations and data gaps ...... 28

7.1. Physical and chemical characteristics of oil droplets and associated hydrates (A) ...... 29

7.2. Formation of subsurface oil plumes: droplet and bubble size (B) ...... 29

* Corresponding author. Fax: +1 7275531189.

E-mail address: [email protected] (K.L. Daly).

http://dx.doi.org/10.1016/j.ancene.2016.01.006

2213-3054/ã 2016 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 19

7.3. Important processes in the surface mixed layer (C) ...... 29

7.3.1. Interaction of oil droplets with euphotic zone plankton (C1) ...... 29

7.3.2. Transformation of oil floating at the atmosphere-ocean interface (C2) ...... 29

7.3.3. New sources of material at the surface (C3) ...... 29

7.3.4. Sinking of oil-derived material from the surface (C4) ...... 29

7.4. Processes impacting MOS during sedimentation (D) ...... 29

7.5. Processes impacting MOS on the seafloor (E) ...... 29

7.6. Processes impacting resuspension of oiled (F) ...... 29

8. Conclusions ...... 30

Acknowledgements ...... 30

References ...... 30

2

1. Introduction ultimately covered up to 68,000 square miles (180,000 km ) of

ocean before it was contained (Norse and Amos, 2010), including

The 2010 Deepwater Horizon (DWH) oil spill (Fig. 1), which coastal, deep water, and riverine influenced regions. The outflow

began on April 20th and ended on July 15th, was unprecedented for from the Mississippi River and associated distributary channels

several reasons: (1) the oil outflow originated from an ultra-deep was above climatological mean between mid-May and October of

well at 1500 m, (2) the large volume of oil released (about 2010, which influenced the near surface of oil

4.9 million barrels or 779 million L), (3) the large amount of (Kourafalou and Androulidakis, 2013). Typically, the river

6 À1

dispersants (Corexit EC9500A and EC9527A; about 2.1 million discharges about 130–150 10 t year of (Corbett

gallons or 7.9 million L) released in deep water and at the et al., 2006; Bianchi et al., 2007); hence, riverine suspended

surface (Kujawinski et al., 2011; McNutt et al., 2012; Lubchenco mineral particles likely interacted with oil (Muschenheim and Lee,

et al., 2012), and (4) the unexpected and protracted sedimentation 2002; Khelifa et al., 2005) to form sinking oil-mineral aggregations

event of oil-associated marine snow to the seafloor (Passow, 2014; (OMAs) in addition to MOS. Significant sediment and hydrocarbon

Brooks et al., 2015; Romero et al., 2015). deposition to the seafloor was observed in the DeSoto Canyon

The Oil Budget Calculator Science and Engineering Team region to the east of the Deepwater Horizon wellhead (Brooks et al.,

reported that by the end of the oil spill 17% of the South Louisiana 2015; Romero et al., 2015) and elsewhere (White et al., 2012;

Sweet Crude oil (MS252) had been recovered at the wellhead, 5% Montagna et al., 2013; Valentine et al., 2014; Chanton et al., 2015).

burned, 16% chemically dispersed, 13% naturally dispersed, 23% Based on combined sedimentology, geochemical, and biological

evaporated or dissolved, with 23% remaining in an Other category approaches, this mass deposition was primarily a product of

(Lehr et al., 2010; McNutt et al., 2012). The Oil Budget Calculator marine snow formation and appears to have occurred over a

was designed as a response tool to inform response actions and 4–5 month period during and after the oil spill, far exceeding pre-

cleanup decisions and was not intended to be a research or damage spill sediment accumulation rates (Brooks et al., 2015). A high

assessment tool. Consequently, oil sedimentation and its accumu- degree of patchiness in the deposition of oil on the seafloor,

lation on the seafloor were not specifically considered in the oil combined with the large geographic area of the sedimentation

budget calculation. However, early scientific investigations of the event (including on the upper continental slope and shelf), makes

oil spill noted that high concentrations of oil-associated marine it a difficult task to estimate the total amount of oil transported to

snow (herein marine oil snow or MOS) were observed in the depth. Nevertheless since the oil spill ended, Valentine et al. (2014)

vicinity of surface oil and the sub-surface oil plumes (e.g., Passow estimated that 1.8–14% of the oil was transported to the seafloor

et al., 2012; Daly et al., 2013). Based on images, the oil spill using hopanes as a biomarker-tracer, while Chanton et al. (2015)

3 2

Fig. 1. The surface distribution of Deepwater Horizon oil. Average volume of surface oil (m /km ) in 5 Â 5 km gridded cells between 24 April and 3 August 2010, based on SAR

satellite data (after MacDonald et al., 2015).

20 K.L. Daly et al. / Anthropocene 13 (2016) 18–33

estimated the amount at 0.5–9% using radiocarbon distributions. 2. Characterization of oil-associated marine snow: processes

For comparison, Boehm and Fiest (1980) estimated 1–3% of the oil and pathways

from the Ixtoc-I spill in the southern Gulf of Mexico (1979–1980)

reached the shallower seafloor in that region, while Jernelöv and 2.1. Marine snow distribution and sedimentation patterns

Lindén (1981) estimated 25% of the Ixtoc oil sank to the seafloor by

mass balance. Marine snow is defined as particles >0.5 mm to 10 s of cm in

In May 2010, BP committed $500 M for the Gulf of Mexico size, which may consist of aggregations of smaller organic and

Research Initiative (GOMRI) to investigate the impacts of oil and inorganic particles, including , phytoplankton, micro-

dispersants on the Gulf of Mexico ecosystems. Preliminary findings zooplankton, zooplankton fecal pellets and feeding structures

reported at conferences and in discussion groups indicated that (e.g., larvacean houses), biominerals, terrestrially-derived litho-

MOS led to significant oil sedimentation to the seafloor. The GOMRI genic components, and (Alldredge and , 1988).

funded Marine Oil Snow Sedimentation and Flocculent Accumula- Marine snow occurs throughout the world’s and at all

tion (MOSSFA) Workshop was held during October 2013, with the depths. Marine snow is formed in near surface waters where

goal to obtain community input, particularly on three topic areas: particle abundances vary spatially and seasonally, usually between

À1 À1

(1) factors affecting the formation and sinking of MOS in the water 1 and 14 particles per L (range 0–500 particles L ) (Alldredge

column, (2) the deposition, accumulation, and biogeochemical fate and Silver, 1988). In the Gulf of Mexico, suspended particle

of MOS on the seafloor, and (3) the ecologic impacts of MOS on concentrations off Louisiana during April 1987 were reported to

À1 À1

pelagic and benthic species and communities (MOSSFA Workshop range from 10–60 particles L on the shelf to 0.2 to >3 particles L

Report, 2014; Kinner et al., 2014). The purpose of this review is to on the slope, declining with depth and then increasing near bottom

À1

raise awareness that the settling of MOS may play a significant role (1.6–3.3 particles L ) (Gardner and Walsh, 1990). Aggregates in

in the distribution and fate of spilled oil in both shallow and deep the Mississippi Canyon were assessed for size and settling rates

water environments, and that sinking and sedimentary-oil below the mixed layer at 167 m depth during October 1993

deposition should be considered in future oil spill response (Diercks and Asper, 1997). Aggregates typically ranged between

assessments. Here, we report on the MOSSFA Workshop findings, 0.5 and 3.5 mm in diameter, with one aggregate at 7.3 mm. Settling

À1

provide a summary of the current state of knowledge related to the rates ranged from 10 to 85 m d , with the largest particle having a

Deepwater Horizon marine snow-oil deposition event, present a rate similar to 1–2 mm particles. Particle abundances near coastal

mechanistic framework of MOSSFA processes, and offer examples shelves in this region also may be elevated at midwater depths as a

of data gaps, as well as providing recommendations for future result of benthic resuspension by mesoscale circulation features

research. A comprehensive review of marine snow is outside the and transport (lateral advection) of particles off the shelf and slope

scope of this paper; for general background information see, for (Walsh and Gardner, 1992; Diercks and Asper, 1997). Relatively

example, Alldredge and Silver (1988),Simon et al. (2002), Turner high concentrations of suspended aggregates (20–250 particles

À1

(2002), and Burd and Jackson (2009). L ) were observed in a bottom nepheloid layer in the Mississippi

Canyon following the passage of Hurricane Isaac in August 2012,

with most aggregates <1 mm in diameter (Ziervogel et al., 2015).

Fig. 2. Particle size spectra of marine snow and examples of marine snow images observed by the SIPPER imaging system during August 2010, September 2011, August 2012,

August 2013, and August 2014. The particle size distribution data are from a station 50 km to the east of the DWH site.

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 21

In comparison, within one month after the wellhead was capped 2.2. Factors affecting MOSSFA processes

(August 2010), marine snow concentrations ranged from 10.5 to

À1

64.6 particles L in the upper 20 m of off-shelf waters in the region There are many factors that impact the formation and

of the oil spill, with particle sizes ranging from 0.150 to 17.6 mm in modification of MOS in the (Figs. 3–5). Figs. 3–5

diameter (Daly et al., 2014). Oil concentrations (total were developed as part of the MOSSFA Workshop discussions.

hydrocarbons, TPH, and polycyclic aromatic hydrocarbons, PAH) Fig. 3 focuses on surface processes, Fig. 4 illustrates processes from

were still elevated compared to background levels within 800 km the origin of oil outflow to final oil deposition, and Fig. 5 depicts the

of the wellhead during August, albeit at lower concentrations than impacts of gradients driven by riverine processes on MOS in the

were observed prior to capping the wellhead (Wade et al., 2015). water column, on the seafloor, and within the sediments. These

The August 2010 maximum marine snow concentrations diverse factors emphasize the complex nature of the ocean

À1

(64.6 particles L ) were considerably higher in near surface environment and its interactions with oil blowouts. Factors shown

À1

waters, than during September 2011 (2 particles L ), August in Fig. 3 include (1) riverine gradients of salinity and inputs of

À1 À1

2012 (4.9 particles L ), August 2013 (20.7 particles L ), or August , dissolved , and lithogenic material

À1

2014 (1.5 particles L ) at a deep-water station 50 km to the east of influences particle aggregation (Muschenheim and Lee, 2002),

the DWH site. Integrated abundances (0–140 m) of marine snow because coagulation is a function of particle abundance and size

also showed a similar pattern of maximum concentrations during (Burd and Jackson, 2009). Nutrients enhance phytoplankton

À2

August 2010 (866,862 particles m ) at this site compared to growth and dissolved organic matter forms gel particles, especially

summer concentrations during the following four years in salinity gradients (Wetz et al., 2009; Verdugo et al., 2004).

(September 2011: 190,375, August 2012: 181,644, August 2013: (2) Marine biota may form or destroy marine snow (Alldredge and

À2

238,574, and August 2014: 74,311 particles m ) (Fig. 2). Silver, 1988; Burd and Jackson, 2009). For example, phytoplankton

The dominant factors controlling gravitational settling of and bacteria release ‘sticky’ exopolymeric substances (EPS) due to

surface particles and aggregates vary spatially and temporally oil and dispersant exposure and mucus acts as glue, providing the

(Boyd and Trull, 2007). Sinking speeds, however, are primarily a matrix for aggregates (Passow, 2014). Biological formation also

function of particle size and excess density (i.e., particle includes processes such as incorporation into zooplankton feeding

density > density) (Armstrong et al., 2002; Iverson and structures and feces and microbial generated marine snow.

Ploug, 2010; De La Rocha and Passow, 2007). Sinking speeds of Alternatively, zooplankton feeding on MOS may to fragmen-

marine snow range from 10 s to 100 s of meters per day. Slowly tation of particles. Both microbial degradation of hydrocarbons to

sinking particles are removed by microbial remineralization or non-toxic forms (Kleindienst et al., 2015a) and bioaccumulation

grazing and, therefore, rarely reach great depths. The formation into the food web (Lee et al., 2012; Chanton et al., 2012) occur.

and flux of rapidly sinking marine snow is one of the primary (3) Mediating measures, such as chemical dispersants and burning,

processes by which the marine exports surface affect the properties and behavior of hydrocarbons and influence

and other elements to the deep ocean and seafloor. the formation of MOS (Passow, 2014). Dispersants break down oil

Fig. 3. Image depicts the factors that affect the formation and modification of MOS. The blue triangle represents the area of the oil spill. MOS is affected by the ocean

environment, the state (degree of weathering) and composition of oil, spatial overlap with riverine and shelf processes, the type of marine biota present and their related

processes, and the timing and location of spill mediation measures (e.g., application of dispersants and burning).

22 K.L. Daly et al. / Anthropocene 13 (2016) 18–33

Fig. 4. Conceptual diagram of MOS related processes from the source of oil discharge to the fate of hydrocarbons in sediments. (A) Shows the release of oil at the wellhead and

application of dispersants and (B) represents rising oil droplets and gas bubbles and the formation of a deep oil plume. (C1–C4) Shows surface processes influencing the

formation of MOS: (C1) illustrates wind impacts, a diatom bloom, and application of surface dispersants, (C2) shows oil transformation due to UV and evaporation, (C3)

depicts the role of and oil burning in creating new material sources, and (C4) shows processes impacting sinking MOS particles in surface waters and as particles sink

through (D) a benthic nepheloid layer and deep oil plumes. (E) Shows benthic sedimentation of MOS and flocculation onto corals, and (F) represents resuspension of oiled

sediments due to turbulence. See the text for a more detailed explanation of the figure.

À1

into small droplets, but may inhibit microbial oil degradation (see 68 to 553 m d . Laboratory experiments elucidated some of the

Section 2.5, Kleindienst et al., 2015b). Oil burning leaves about 5% mechanisms leading to the formation of MOS by using roller tanks

of the original volume as burned residue such as and to simulate in situ conditions. Roller tanks allow marine snow to

other pyrogenic by-products (Lehr et al., 2010). Black carbon settle continuously without contact with surfaces (container

particles efficiently absorb organic matter, including PAHs walls) (Ploug et al., 2010). These experiments revealed a number

(Koelmans et al., 2006), and can stimulate transparent exopolymer of important processes, including, (1) diatom aggregates incorpo-

particle (TEP) production and aggregation of marine snow (Mari rated appreciable amounts of oil, either by collision and attach-

et al., 2014). (4) Weathering, or aging, and photochemical ment of oil droplets with phytoplankton cells, or via absorption of

alterations to oil similarly impact MOS properties (Bacosa et al., oil to phytoplankton, (2) oil droplets were incorporated into

2015; Passow, 2014). Photo-oxidation is the process by which colonies of (Trichodesmium) through coagulation,

hydrocarbons, particularly PAHs, react with in the presence and (3) bacteria mediated MOS formation in the absence of other

of sunlight, resulting in chemical and structural changes to oil that particles, possibly through the release of high concentrations of

may lead to increased water solubility or decreased microbial mucus (Passow et al., 2012; Fu et al., 2014; Passow, 2014).

. (5) Physical processes may lead to aggregation of Interactions between oil components, bacteria, and natural

particles due to collision or high turbulence may result in particle suspended matter formed flocs (relatively large, fluffy aggregates)

fragmentation (Hill et al., 2002; Burd and Jackson 2009). Currents, likely due to the formation of macro-gels, such as TEP, from EPS

eddies, , benthic resuspension and cross-shelf flow, released by bacteria in response to the presence of oil (Passow

and other physical processes impact the distribution of particles et al., 2012; Ziervogel et al., 2014). A complicating factor was the

and oil in space and time (e.g., Paris et al., 2012; Smith presence of the dispersant, Corexit, which at near in situ

et al., 2014). These processes all influence the dynamics of sinking concentrations slowed or inhibited the formation of microbial

MOS. MOS (Passow 2014). However, after longer incubations, possibly

once Corexit was degraded, marine snow formed (Fu et al., 2014).

2.3. Field and lab MOS formation

2.4. Microbial processes

Large, mucus-rich MOS particles were observed in surface oil

during May, ranging in size from small compact particles to Numerous microbes are able to utilize oil as a major source of

particles several centimeters in size. Large, stringy marine snow carbon and , including 175 genera of bacteria, several

particles (<10 cm), with mucous threads resembling web-like haloarchaeal genera, and many Eukarya (McGenity et al., 2012).

structures, were also common (Passow et al., 2012). Sinking Microbial response to an oil spill is governed by many factors,

velocities for settling MOS particles were estimated to range from including oil composition, degree of oil weathering, and

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 23

Fig. 5. Diagram illustrates the environmental gradients of material properties and fluxes associated with a of oil released in regions influenced by river outflow

compared to offshore regions not influenced by riverine processes. Gradient shifts include the concentration and composition of suspended particles (clays to carbonate), the

magnitude of particulate organic carbon (POC) and petrochemical fluxes to the seafloor, the depth of the sediment redoxcline, and the tolerance of benthic organisms, such as

foraminifera, to different oxygen levels in sediments. Oil-mineral aggregations (OMAs) may sediment separately or in association with marine oil-snow (MOS). These

environmental gradients overlap and interact with gradients generated by oil spills, e.g., oil and dispersant distributions, causing a complex temporal-spatial distribution of

interactive effects.

environmental conditions, such as temperature and which was abundant during the DWH oil spill (Yan et al. in review,

concentrations. The temporal and spatial variability in microbial Passow unpubl.), provide a biotope for hydrocarbonclastic bacteria,

populations before, during, and after the DWH oil spill, and in which appear to specialize in PAH degradation. While significant

comparison with communities outside of the spill, showed that progress has been achieved in documenting shifts in the bacterial

groups of oil-degrading bacteria responded to the presence of oil community in response to the oil spill, the spatial and temporal

by May 2010, including alkane, PAH, and degraders, and variability of microbial effects on MOS formation and degradation

nitrifying (Dubinsky et al., 2013; Yang et al., 2014; remain poorly known.

Crespo-Medina et al., 2014; King et al., 2015). Overall, the relative

importance of different taxa was governed by changes in 2.5. Plankton impacts on MOS

hydrocarbon composition and supply. Surface oil communities

appeared to have been more diverse than communities in the deep Microplankton, microzooplankton, and mesozooplankton were

oil plumes (Kimes et al., 2014). By September 2010, the pre-spill impacted by the spilled oil, which likely affected marine snow

community was re-established, with low background concen- formation and sedimentation. Crude oil and the dispersant Corexit

trations of oil degrading bacteria. In both surface and deep waters, are reported to have had direct lethal and sublethal effects on

MOS particles were colonized by heterotrophic microbes, which bacteria, phytoplankton, microzooplankton (Garr et al., 2014;

expressed high specific rates of enzymatic activity that were Almeda et al., 2014b; Kleindienst et al., 2015a,b) and mesozoo-

different from those of the surrounding seawater (Ziervogel et al., plankton, including changes in physiology and reproduction

2012; Arnosti et al., 2015). In addition, several bacterial species that (Ortmann et al., 2012; Almeda et al., 2013a,b, 2014c; Cohen

are noted for producing EPS had elevated concentrations in surface et al., 2014; Peiffer and Cohen, 2015). The impact of dispersants,

waters during the oil spill and formed mucus aggregates that however, is being debated. Prince (2015) argues that dispersants

contained oil droplets in lab experiments (Gutierrez et al., 2013). are not significantly toxic under typical oil spill field conditions,

In situ MOS particles also had significant concentrations of they make oil more bioavailable for degradation, and impacts on

glycoprotein (carbohydrate and protein), a primary component are minimized. In contrast, a number of studies have

of EPS (Arnosti et al., 2015). Specific bacteria respond to oil in indicated that toxic effects on pelagic organisms increased with the

different ways, but in general, microbial communities produce addition of Corexit to oil treatments, although results varied

mucus that acts like a biofilm, allowing a complex community to between taxa (e.g., Almeda et al., 2013b, 2014b; Cohen et al., 2014;

establish, which jointly utilizes the different components of oil and Garr et al., 2014). Kleindienst et al. (2015b) also reported that the

its metabolites (McGenity et al., 2012). Bacterial EPS also may presence of Corexit decreased microbial degradation rates of

stimulate non-oil degrading bacterial communities. Mishaman- hydrocarbons, by selecting for bacteria that degrade dispersant

dani et al. (2015) demonstrated that eukaryotic phytoplankton, (i.e., Colwellia) rather than bacteria that degrade hydrocarbons (i.e.,

such as the cosmopolitan marine diatom, Skeletonema costatum, Marinobacter). These authors suggested that dispersants might

24 K.L. Daly et al. / Anthropocene 13 (2016) 18–33

stimulate hydrocarbon degradation in other marine systems that Since the DWH platform was about 75 km offshore directly

do not already have a background population of hydrocarbon- southeast from the Mississippi River, OMAs may have played a

degrading bacteria. The Gulf of Mexico harbors hydrocarbon- role in DWH oil deposition.

degrading bacteria at low densities, due to chronic oil input from

cold seeps. 2.7. MOS sedimentation

Crustacean molts and dead zooplankton typically become part

of the marine snow assemblage, sinking rapidly out of the water The role of MOS in DWH oil sedimentation is supported by data

column. Zooplankton (e.g., dinoflagellates, gelatinous doliolids, from a deployed in late August 2010 about five km

) ingest oil and egest oil in fecal pellets, which sink southwest of the DWH platform, 140 m above the seafloor at

rapidly to the seafloor (Lee et al., 2012; Almeda et al., 2014a,c). 1400 m depth. The trap results showed exceptionally high POC

Størdal et al. (2015) demonstrated that despite reduced feeding sedimentation rates relative to other years, which were due to the

activity by a North Sea in the presence of oil, their fecal sinking of a large diatom bloom that was almost entirely composed

pellets contained oil plus oil-degrading bacteria (Rhodobacter- of Skeletonema sp. (Yan et al. in review, Passow unpubl.), a

aceae), which were indigenous to the copepods. This family of cosmopolitan taxa that thrives under brackish conditions and is

bacteria was a dominant group present after the DWH spill tolerant of the presence of oil (Parsons et al., 2014). Sedimentation

(Dubinsky et al., 2013; Kimes et al., 2014); hence, zooplankton may of oil continued for >5 months after the spill ended, characterized

contribute to oil degradation. In addition, oil may adhere to by numerous smaller events (Yan et al. in review, Passow unpubl.).

zooplankton and be passively absorbed or ingested; thus, Lithogenic minerals (i.e., silts and clays) co-settled with organic

contributing to bioaccumulation of PAHs (Mitra et al., 2012). particles and constituted on average 60% per weight of the settled

Furthermore, carbon isotopic depletion in suspended particulate material. from diatom frustules and organic carbon

matter and zooplankton supports the notion that oil carbon was composed other significant portion, whereas little calcium

incorporated into the lower trophic food web through biodegra- carbonate was observed in trap material.

dation by bacteria (Graham et al., 2010; Chanton et al., 2012;

Cherrier et al., 2014). 3. Temporal and spatial patterns of MOS

2.6. Oil-mineral aggregates (OMAs) Although the overall processes related to the formation of MOS

and OMAs are reasonably well know, less is known about the

Interactions between oil and suspended particulate material, or temporal and spatial variability of their formation and deposition,

OMAs as they have been designated in the literature, have long which are generally regulated by physical processes. Wind forcing

been recognized to result in oil sedimentation in freshwater and was the primary mechanism governing surface oil drift (Le Hénaff

marine systems, especially in tidal and subtidal areas (Lee, 2002; et al., 2012). In addition, freshwater discharge from the nearby

Payne et al., 2003). OMAs are usually smaller (<1 mm, often Mississippi River, which is the world’s third largest river, often

<50 mm) than most marine snow particles (Stoffyn-Egli and Lee, results in strong stratification and enhanced wind-driven coastal

2002) and form primarily in surf zones, near river outflows, jets (Morey et al., 2003; Jochens and DiMarco, 2008). During April

melting glaciers or sea ice, and in semi-enclosed bays, where 2010, freshwater diversions along the lower Mississippi River were

suspended lithogenic particle concentrations are relatively high opened with the intent to minimize the impact of the oil spill on

(Lee and Page, 1997; Payne et al., 2003). Boehm (1987) estimated and (Bianchi et al., 2011). Kourafalou and

that suspended particle concentrations need to be > 10 mg/L for Androulidakis (2013) concluded that in addition to wind forcing,

significant deposition of oil to occur and >100 mg/L for large Mississippi River induced circulation also significantly influenced

deposition events. Since inorganic particles concentrations are near surface oil transport based on satellite and in situ data and

usually <10 mg/L in the open ocean, this pathway was considered numerical simulations. Winds blowing from the south reduced the

to be relatively unimportant offshore of intertidal/subtidal regions. interaction of surface oil with the , consequently only

Offshore suspended drilling muds may be considered an exception. a small amount of oil was trapped in a spin off (Eddy Franklin)

A review of near-shore oil spills indicated that 1–13% of spilled (Le Hénaff et al., 2012). A number of processes, which likely

oil settled to the seafloor due to OMA processes (Lee and Page, influenced weathering of MC252 crude oil, including physical

1997). The interactions between suspended particulate matter, oil agitation, wave conditions, evaporation, photo-oxidation, and

weathering, adsorption, microbial processes, other marine snow dispersability, were summarized in Daling et al. (2014). Three

particles, and grazing zooplankton impact oil packaging and MOS storm events during the oil spill, of which two were named (e.g.,

sedimentation (reviewed in Muschenheim and Lee, 2002). OMAs Hurricane Alex, Tropical Storm Bonnie), led to a change in surface

sink rapidly, in spite of their small size, because of their high oil extent and deep mixing (Goni et al., 2015), but the storm

mineral content. OMA formation increases the surface to volume impacts on the oil spill and oil deposition are for the most part

ratio of oil, thereby extending and enhancing the weathering unknown. Previous studies during hurricanes observed significant

processes of dissolution, evaporation, and biodegradation vertical variations in currents and internal waves (Shay and

(Lee, 2002). Water turbulence (i.e. breaking waves, strong flood Elsberry, 1987; Keen and Allen, 2000) Storm induced resuspension

currents) enhances OMA formation and transport and mineral-oil of flocculent material deposited on the seafloor during the DWH

flocculation rates are typically highest in low to intermediate event was shown to lead to increased microbial activities

salinity waters. Danchuk and Willson (2011) suggested that OMA (Ziervogel et al., 2015). In addition, the Loop Current, eddies,

formation in the vicinity of the Mississippi River would vary cross-shelf flows, and mesoscale circulation are important physical

depending on the time of year and type of oil. Lighter oils had a processes in the NE Gulf of Mexico, which impact biological

higher probability of forming OMAs during winter and spring, production and oil transport (Nababan et al., 2011; Olascoaga et al.,

when there was higher sediment availability and denser, high- 2013; Goni et al., 2015; Jones and Wiggert, 2015; Smith et al., 2014).

viscosity oil may form more OMAs during summer, when salinity Biological factors were another major influence on the spatial—

was higher. OMAs form where oil and suspended minerals temporal patchiness of the formation and fate of MOS. During

(including clay-sized particles) co-occur, whereas lithogenic August 2010, satellite data (MODIS Fluorescence Line Height)

material from, for example, a deep nepheloid layer may be showed unusually high chlorophyll concentrations compared to

scavenged by sinking marine snow passing through that layer. the mean of an eight-year time-series for the NE Gulf of Mexico.

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 25

This chlorophyll maximum was located to the east of the DWH wellhead was capped, the plumes were estimated to have reached

2

wellhead and covered an area >11,000 km (Hu et al., 2011). These up to 300 km in extent, with oil concentrations ranging from 5 to

elevated phytoplankton concentrations appeared to be related to 500 ppb, and had interacted with the slope sediments of the

the oil spill since the high river flow anomalies did not correlate Mississippi and Desoto canyons (Paris et al., 2012; Lindo-Atichati

with the spatial area of the chlorophyll anomaly. Whereas et al., 2014; Romero et al., 2015).

exposure to high concentrations of oil is lethal to most Socolofsky et al. (2011) demonstrated that the deep subsurface

phytoplankton, sensitivity to hydrocarbons varies and some oil plumes derived from a stratification-dominated multiphase

species thrive under low concentrations of oil (Parsons et al., plume, which was characterized by multiple subsurface intrusions

2014; González et al., 2013). In addition to the region of the of dissolved gas and oil, as well as small droplets of liquid oil.

elevated chlorophyll anomaly, nutrients from the Mississippi River Simulations by North et al. (2015) emphasized the importance of

fueled phytoplankton blooms during 2010 (Hu et al., 2011), which oil droplet size and rates of biodegradation in understanding

are typically observed every year in the vicinity of the river plume factors controlling oil transport. Based on a model comparison

(Lohrenz et al., 1997). Large mucus-rich marine snow particles study, DWH oil droplet sizes were predicted to have ranged from

were observed floating in surface waters during May, but had 0.3 to 6 mm without the addition of dispersants and between 0.01

disappeared by June (Passow et al., 2012). Other types of marine and 0.8 mm with the addition of dispersants (Socolofsky et al.,

snow, however, remained in relatively high concentrations at least 2015). Ryerson et al. (2012) estimated that the majority of oil

into August (Daly et al., 2014). Similarly, Patton et al. (1981) noted droplets that arrived at the sea surface were millimeters in

that pancakes and flakes of mousse or weathered oil occurred at diameter based on hydrocarbon mass transport calculations. Aman

the surface during the Ixtoc-I oil spill. In addition, large numbers of et al. (2015) reported that droplets <40 mm formed the deep lateral

dead gelatinous zooplankton (Pyrosoma sp.) were observed floating oil intrusions, whereas larger droplets >100 mm rapidly rose to the

in an area covering a two mile radius on 14 June 2010, about three surface, based on results from high-pressure laboratory experi-

nautical miles southwest of the DWH spill site (R. Amon pers. ments and model simulations (Paris et al., 2012). The simulations

comm.), suggesting that thalacians (e.g., pyrosome, , or doliolid further suggested that the application of dispersants might not

zooplankton) also may have contributed to MOS flux. have played a significant role in the partitioning of surface and

subsea oil, in that the amount of oil reaching the surface may have

4. Deep oil plumes only been reduced by 3%. Details on the impacts of droplet size and

pathways for the formation of MOS await further investigations.

After the DWH explosion, persistent subsurface oil plumes Microbial degradation rates of oil and gas hydrocarbons in the

(Fig. 4) were detected by early May and later verified by chemical deep-water plumes are the subject of debate. Ziervogel and Arnosti

analyses to have occurred primarily between 1000 and 1400 m (2013) observed higher microbial rates of protein and carbohy-

depth (Camilli et al., 2010; Diercks et al., 2010; Spier et al., 2013). drate hydrolysis in the deep oil plume compared to waters outside

Ryerson et al. (2012) estimated that >30% of the hydrocarbon mass the plume, suggesting EPS and oil degradation products enhanced

released may have gone into the deep plumes. The subsurface oil bacterial metabolism in the plume. Hazen et al. (2010) suggested



plumes contained a complex mixture of soluble and insoluble that degradation rate in these deep cold waters (5 C) were

liquid hydrocarbons, including alkanes, monoaromatic hydro- relatively high (turnover rates on order of days) due to several

(e.g., BTEX), PAHs, and gaseous C1–C4 hydrocarbons, factors: (1) the MC252 oil was a light (volatile) crude and thus

including methane, ethane, propane and butane (Spier et al., 2013; more easily degraded, (2) the deep oil plumes included accessible

Reddy et al., 2012). The application of dispersants at the wellhead small oil droplets, and (3) oil from natural seeps in the Gulf of

in deep water was intended to decrease the mean oil droplet size, Mexico maintained an oil-adapted deep-sea microbial community.

which may have decreased the droplet rising velocity and Camilli et al. (2010), however, reported relatively slow rates of

increased the residence time of oil in the water column and hydrocarbon degradation on the order of months. The observed

impacted its transport (Socolofsky et al., 2015). In contrast, high- differences in rate measurements are likely related to variability in

pressure experimental studies of oil droplet size by Aman et al. the temporal and spatial response of the microbial community to

(2015) suggest that the sub-surface plume would have formed oil and Corexit input. Degradation rates also may have been

even without dispersant application. Marine snow particles, which depressed by dispersants (Kleindienst et al., 2015b). Recent topic

formed at the surface and sank to the seafloor, likely interacted reviews suggest that the initial microbial response to the DWH oil

with rising oil droplets and/or the oil plumes at depth (Valentine spill was driven by light hydrocarbons (Valentine et al., 2010) and

et al., 2014). Surface bacteria and phytoplankton-affiliated that overall the microbial response was rapid and robust (Kimes

sequences were found in sediments at and below the subsurface et al., 2014; King et al., 2015). Although little methane emitted at

plume (Mason et al., 2014; Brooks et al., 2015), substantiating the the wellhead made it to the sea surface (Yvon-Lewis et al., 2011),

notion of transport of surface-formed MOS to depths. MOS likely the timing of methane degradation is controversial (Kessler et al.,

formed in the deep-water plumes as well, as many bacteria were 2011; Joye et al., 2011a,b; Crespo-Medina et al., 2014). In addition,

active in these regions (Hazen et al., 2010; Redmond and Valentine, the rates of hydrocarbon degradation under high-pressure

2012; Dubinsky et al., 2013) and because MOS formed in conditions remain poorly understood (Gutierrez and Aitken,

experiments using bacteria from the deep plume (Baelum et al., 2014; Joye et al., 2014; King et al., 2015), even though pressure

2012; Kleindienst et al., 2015b). affects reaction rates of dissolved gasses (Bowles et al., 2011).

The temporal and spatial dynamics of the deep oil plumes were Metabolic function in some bacterial strains which degrade

complex and variable. A coupled 3-D hydrodynamic and oil particle hydrocarbons were recently shown to be inhibited by pressure

tracking model indicated that local topography and hydrodynamic levels typical of the deep oil plumes (Schedler et al., 2014),

processes, such as eddies, were important in governing the DWH emphasizing that pressure must be considered when evaluating

oil distribution and that use of deep dispersants may have led to a degradation rates in deep water. In contrast, methane oxidation,

deeper vertical distribution of suspended oil (Paris et al., 2012). The which is a first order kinetic process, occurs at significantly higher

relatively narrow plumes occurred at multiple depths and rates under deep-sea pressures, preventing methane supersatura-

extended both to the southwest and to the northeast of the tion (Bowles et al., 2011). Nonetheless, early concerns raised about

wellhead, with the maximum oil concentrations in the plume the onset of hypoxic conditions as a result of the large methane

generally flowing along the 1200 m isobath. By the time the releases in the (Joye et al., 2011a,b) were not realized, as

26 K.L. Daly et al. / Anthropocene 13 (2016) 18–33

dissolved oxygen concentrations did not decrease substantially. with river discharge, marine detritus, phytoplankton, microbial,

Instead, the methanotroph community appears to have consumed and petrochemical substances. The relative importance of these

methane relatively rapidly, transforming it into particulate organic materials will vary with distance from rivers and continental

carbon (Redmond and Valentine, 2012; King et al., 2015; Cherrier shelves and water depth. Except for regions influenced by the

et al., 2014). Overall microbial processing of oil components has Mississippi River Plume, Loop Current, or eddies, offshore areas in

been estimated to account for 40–60% of the discharged hydro- the central Gulf of Mexico typically have lower than

carbons (gas and oil) (Joye, 2015). Because marine snow particles shelf regions and phytoplankton communities are dominated by

are considered hot spots of microbial activity (Azam and Long, small cells, primarily cyanobacteria (Wawrik and Paul, 2004;

2001), with greatly enhanced enzyme activities and degradation Muller-Karger et al., 2015). While Fig. 5 shows a simplified pattern

rates compared to surrounding seawater (Smith et al., 1992), MOS of onshore to offshore gradients in properties and productivity, the

and OMAs likely contributed to the high microbial degradation off-shelf region of the oil spill in the NE Gulf of Mexico is much

rates of oil in the deep oil plumes. more environmentally complex and generally has higher produc-

tivity than other regions of the Gulf of Mexico (Nababan et al.,

5. Oil accumulation rates and fate at the seafloor 2011). Sedimentation rates coupled to sediment grain-size and

organic content are the major controls on the rates and depths of

Factors affecting the long-term sediment accumulation rates in oxidation–reduction (redox) reactions in sediments (Hastings

the northeast GOM and during the DWH marine snow-oil event are et al., 2015). Redox reactions, in turn, will influence porewater

shown in Figs. 3–5. The Mississippi River typically dominates chemistry, microbial community structure/function, dissolved and

sediment transport and composition on Mississippi, Louisiana, and solid-phase metal concentrations, degradation rates, and decom-

Texas shelves, as well as the offshore Mississippi Deep-Sea Fan position and transformation of petrochemicals. Furthermore, the

(Balsam and Beeson, 2003). The Apalachicola River to the east on impacts of oil deposition on benthic fauna via smothering, shoaling

the Florida Gulf affects sediment transport and composition of the oxic–anoxic boundary, and toxic effects of oil will affect the

too. Fig. 5 illustrates the large gradients of ocean properties driven depth and occurrence of sediment bioturbation and oil and organic

by riverine processes. Sediment deposition from riverine sources is matter degradation rates.

a function of discharge rate, grain size, clay minerology, sediment MOS deposition was observed in bottom sediments within

resuspension, cross-shelf transport, and slumping (Gardner and 256 km of the DWH platform, as an approximately 1 cm-thick

Walsh, 1990; Corbett et al., 2006; Bianchi et al., 2007). River sedimentary layer (Montagna et al., 2013; Mason et al., 2014;

additions of nutrients also fuels marine phytoplankton blooms of Valentine et al., 2014; Brooks et al., 2015). The Brooks et al. (2015)

diatoms (Lohrenz et al.,1997; Dagg and Breed, 2003). Thus, sources study area extended up to 100 nautical miles (185 km) northeast of

of organic matter accumulated on the seafloor include anthropo- the DWH wellhead where thorium inventories of sediment cores

genic and terrestrially-derived biogenic substances associated indicated that the deposition occurred within four to five months,

Fig. 6. Conceptual model illustrating the complexity of ecosystem response to a disturbance, such as an oil spill. In order to understand the impact and system response,

information is needed at all levels from to species, populations, communities, and ecosystems. The green arrows depict ecosystem impacts flowing down to impact

genetic adaptation and red arrows show impacts of genetic variability on changes at the species, population, and community levels. Impacts may vary for different parts of an

ecosystem and can be positive, negative, or no observable change.

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 27

through the summer and fall of 2010. Mass accumulation rates over abundance of microbial genes for hydrocarbon degradation

2

the past ca. 100 years ranged from 0.05 to 0.16 g/cm /year. In suggested that sediment surface microbes could rapidly deplete

contrast, the 2010 deposition rate was four-fold higher oil in that environment. Ziervogel et al. (2014), however, observed

2

(0.48–2.40 g/cm /year) compared to rates before 2010 or after only a moderate increase in microbial rates in sediments in

the oil spill during 2011 and 2012. Sedimentation at these sites was response to the MOSSFA event several months (November and

principally due to marine snow, with minor input from the December 2010) after the well-head was capped, leading these

extended Mississippi River discharge (Brooks et al., 2015). authors to hypothesize that this oil may have a long residence time

Petrographic investigations, biomarker analyses, and gene on the seafloor. Furthermore, Hastings et al. (2015) recognized,

sequencing indicated that the amorphous aggregates in sediments through the analysis of redox sensitive metals (e.g., Re and Mn),

were derived mainly from surface-derived diatoms, coccolitho- that the intensity of redox reactions and anaerobic conditions

phores, and cyanobacteria sources, which co-occurred with increased for up to three years after the DWH event, likely a result

pyrogenic PAHs from surface oil burning (Brooks et al., 2015; of excess organic matter and hydrocarbon burial and decomposi-

Romero et al., 2015). In addition, Lincoln et al. (2015) observed tion. These redox changes over time and space were associated

increased carbohydrate concentrations associated with relatively with dramatic shifts in the benthic community composition

large particles in sediments. These authors utilized scanning (Schwing et al., 2015). While the fate of DWH oil on the seafloor

confocal laser microscopy and fluorescently-labeled lectins (that remains uncertain, Soto et al. (2014) reported that during the Ixtoc-

bind to specific glycoconjugates common in marine exopolysac- I oil spill total hydrocarbon concentrations in sediments had

charides) to show that the labeled carbohydrates occurred as blebs returned to background levels two years after the release of

with distinct drape and stringer morphologies and that their 3 million barrels of oil.

collapsed appearance suggested a water column origin as opposed

to formation in situ by hydrocarbon-degrading bacteria. These 6. Ecosystem effects

microscopy results provide further evidence that the carbohydrate

source was from surface-derived EPS in marine snow. It is uncertain to what extent the MOSSFA event impacted Gulf

The amount and distribution of DWH oil in sediments was of Mexico ecosystems. The MOSSA Workshop participants devel-

estimated by Chanton et al. (2015) to be 0.5–9% of the total oil oped a conceptual model (Fig. 6) to convey the complexity of an

from the wellhead, using radiocarbon distributions. Their results ecosystem response to a perturbation and the need for information

indicated that oil was deposited within the upper 1 cm of the on all levels: genes, organisms, populations, communities, and

2

sediments over an area of 8400 km , mostly to the southwest and ecosystem, for both water column and seafloor . All levels

as far as 190 km to the southwest of the wellhead (Chanton are connected through food web and environmental processes,

unpubl.). Valentine et al. (2014), using hopane as a tracer, with impacts transmitted from genes up to ecosystem scales and

reported that 1.8–14% of the oil from DWH was in the upper 1 cm vice versa with ecosystems impacting changes in populations and

2

of sediments over an area of 3200 km around the wellhead. DWH ultimately genetic adaptation. Connectivity through currents with

PAH concentrations in sediment cores from the same sites unimpacted regions is an important factor in recovery following

reported in Brooks et al. (2015) were significantly elevated during severe disturbances in marine systems. However, ecosystem

2010 (up to two orders of magnitude greater than background), response to disturbances and long-term changes may involve

À1

ranging from 70 to 524 ng g and diagnostic markers were threshold effects and non-linear dynamics (Folke et al., 2004).

consistent with a MC252 oil origin (Romero et al., 2015). The PAHs Thus, the complexity of these interactions at all temporal and

appeared to be a mix of petrogenic and pyrogenic (burn-derived) spatial scales makes it difficult to assess the impacts of

material. In addition, these authors point out that the deep oil disturbances, such as the DWH oil spill and MOSSFA, on marine

plumes likely impinged on the continental slope and acted as ecosystems.

another mechanism for oil deposition on the seafloor. Mason et al. Baseline data, which are essential for assessing the impacts of

(2014) evaluated 64 sediment cores collected primarily south- disturbances, unfortunately, are not available for most components

southwest of the DWH site and observed the highest PAH of the food web in the NE Gulf of Mexico. Bernhardt and Leslie

À1

concentrations (19,258 ng g ) within five km of the wellhead and (2013) recommend obtaining time-series data for the underlying

À1

the lowest concentrations (18 ng g ) further away. Valentine ecological components of food webs that are important for

et al. (2014) tabulated hopane concentrations for 534 locations, resilience, i.e., population size, species diversity, and functional

noting that the highest concentrations occurred within 40 km of groups. Information also is needed on response patterns to MOS

the wellhead, primarily to the southwest. These authors argued sedimentation for key species, such as those that are ecologically or

that there is a mismatch between the location of the surface oil economically important. Ecologically important species should

slick and where oil was observed on the seafloor, implying that include those that are representative of specific habitats, or

surface oil was not the primary source of the seafloor oil. Instead, foundation species, at all depths of the water column and in the

they suggested that the majority of the oil stemmed from the . Laboratory mesocosm studies would provide information

deep plume, rather than from the surface layer, based on the for the many species that were not adequately studied during the

composition of oil residues at the seafloor. In addition, Kolian DWH spill. Furthermore, laboratory studies would permit data

et al. (2015) reported continued deposition of oil near the collection on lethal and sublethal exposure and impacts to

wellhead and in coastal waters of Mississippi, suggesting that the different flux rates of MOS and concentrations of oil and

wellhead may have continued to leak until at least May 2012. dispersants.

Thus, a number of different processes may have contributed over Several studies have shown that MOSSFA due to the oil spill

different time and space scales resulting in the heterogeneous impacted the (Fisher et al., 2014; Schwing

distribution of oil on the seafloor. et al., 2015). Direct and indirect impacts of MOS may have

Oil deposition in these deep-sea sediments resulted in occurred through ingestion, microbial activity, smothering,

enhanced microbial response and an intensification of redox suboxic and anoxic conditions, transfer through the marine food

reactions and anaerobic conditions relative to unoiled sediments web, and immunotoxicity through gills and transdermal exposure

(Mason et al., 2014; Hastings et al., 2015; King et al., 2015). Both and/or bioaccumulation. Bioaccumulation of hydrocarbons

aerobic and anaerobic processes were detected in surface sedi- through the planktonic food web would increase exposure of

ments, with shifting community compositions. The relatively high higher trophic-level organisms (Meador, 2003). Zooplankton

28 K.L. Daly et al. / Anthropocene 13 (2016) 18–33

ingested oil droplets directly (Lee et al., 2012; Almeda et al., year later there was evidence of recovery at some, but not all sites

2014c) or may have ingested oil-containing marine snow, as many including the site 111 km away. This is consistent with the

zooplankton are known to feed on marine snow (Alldredge and protracted and increasing intensity of sediment redox conditions

Silver, 1988). Typically 70–90% of marine snow particles sinking for up to three years after MOSSFA (Hastings et al., 2015). In

from surface waters are ingested (repackaged) or fragmented by addition, deep-water coral communities were impacted.

mesopelagic zooplankton or remineralized by bacteria (Guidi Oil-containing flocculent material was observed on 90% of the

et al., 2008). It is not known what percent of the sinking MOS was corals within 6 km of the wellhead with less impact on

deposited on the seafloor or the fate of the mesopelagic communities up to 22 km to the southwest (White et al., 2012;

community during and after the oil spill. Carbon and Fisher et al., 2014). Corals that were covered with brown

stable isotope analyses and natural abundance radiocarbon flocculent material (MOS) showed evidence of stress (e.g., excess

analysis have indicated that oil entered particulate organic mucus production, sclerite enlargement, tissue loss) and mortali-

carbon and the planktonic food web in surface waters (Graham ty (DeLeo et al., 2015). Shallow water mesophotic corals at the

et al., 2010; Chanton et al., 2012; Mitra et al., 2012; Cherrier et al., Pinnacles Reef offshore of Alabama were impacted as well, with

2014; Fernandez et al. in review) and mesopelagic fish and shrimp corals having bare skeletons and broken or missing branches

in deep water (>600 m) likely through trophic transfer (Sam- (Silva et al., 2015). Coral branches that were exposed to higher

marco et al., 2013; Quintana-Rizzo et al., 2015). Pelagic fish that concentrations of settling MOS have not recovered to date

swim with their mouths open to maintain a continuous current of (Hsing et al., 2013). Prouty et al. (2014) reported that corals had

water across their gills (e.g., scombrid fishes; Klinger et al., 2015) incorporated oil carbon up to 30 km from the spill site. Experi-

could experience increased oil exposure from suspended or ments using toxicological assays of coral fragments demonstrated

sinking MOS. There is evidence that coastal macrofauna ingested that all three coral species tested exhibited more severe health

petrocarbon as well (Sammarco et al., 2013; Wilson et al., 2015). decline (e.g., polyp retraction, mucous discharge, exposed

The mesopelagic community primarily feeds on zooplankton and skeleton, mortality) when exposed to dispersant and oil-

vertically migrates into the upper 200 m at night, so they may dispersant mixtures than to oil alone (DeLeo et al., 2015). Since

have encountered subsurface oil plumes and food sources these corals are very slow growing and may live >600 years, they

enriched in oil both at depth and at the surface. In turn, surface are highly vulnerable to disturbance and may be very slow to

and mesopelagic fish are eaten by higher trophic animals, such as recover (Prouty et al., 2014). Some fish and use deep corals

tuna, seabirds, and . An unusual mortality event has been as spawning grounds; thus, impacts to corals could lead to larger

reported for bottlenose dolphins (Tursiops truncatus) in coastal ecosystem impacts.

waters of Louisiana, Mississippi, and Alabama starting in 2010 and Sedimentation of MOS also impacted bottom dwelling fish. An

continuing into 2014 (Venn-Watson et al., 2015). This is the longest assessment of offshore fishes between 2011 and 2012 revealed red

die-off in recorded history of the Gulf of Mexico. snapper (Lutjanus campechanus), king snake (Ophichthus rex),

Although the majority of strandings are bottlenose dolphins, and golden tilefish (Lopholatilus chamaeleonticeps) had elevated

spinner dolphins (Stenella longirostris), Atlantic spotted dolphins PAH and metabolite concentrations, with a composition similar to

(Stenella frontalis), and melon-headed (Peponocephala that from the DWH oil (Murawski et al., 2014; Snyder et al., 2015).

electra) also have stranded. While the factors contributing to Elevated PAHs were detected in fish bile, but not in muscle tissue,

the high mortalities are still being investigated, dolphins in at sites near the wellhead and extending to the west Florida shelf.

Barrataria Bay, Louisiana have exhibited moderate to severe lung Skin lesions were present on up to 9% of the fish. Although skin

disease and evidence of hypoadrenocorticism since 2010 consis- lesions may be a short-term consequence of acute PAH contami-

tent with immunotoxic effects of oil. nation, PAH exposure may result in a variety of immunotoxicity

The large sedimentation of MOS clearly had a negative impact population-level effects, including impaired growth, increased

on benthic organisms, including injuries, mortality, and changes disease susceptibility, reduced larval survival, and reduced net

in community composition. The Gulf of Mexico has a high benthic population fecundity. Red snapper and king snake showed

species diversity, with maximum diversity on the mid to upper signs of recovery during 2012 and 2013, while golden tilefish,

continental slope between 1200 and 1600 m depth (Wei et al., which burrows into sediments and likely had a longer exposure to

2010), which coincides with the depth of the DWH well site. PAHs, still had elevated biliary PAH metabolites during 2013

Montagna et al. (2013) and Baguley et al. (2015) reported on the (Snyder et al., 2015). The recovery of the deep-sea ecosystem is

results of benthic fauna sampled at 170 stations after the oil spill, uncertain given the spatial complexity of the sedimentation event

of which 68 stations were located 0.5–125 km from the wellhead. and lower biodegradation rates of oil in sediments (Mason et al.,

These authors observed a severe to moderate impact on benthic 2014; Ziervogel et al., 2014).

meiofauna, with the highest impact within 3 km of the wellhead

and a moderate impact within 60 km to the southwest. 7. Proposed framework for MOSSFA investigations and data

Community changes in severely impacted areas included an gaps

increase in nematode worms (an indicator of organic ), a

decline in harpacticoid copepod abundance, and low meiofauna MOSSFA Workshop participants created a conceptual diagram

and macrofauna diversity compared to unimpacted sites. Overall, (Fig. 4) to (1) better define the processes that impact MOS

copepods had the largest (10-fold), followed formation and sedimentation from the point of the deep-water oil

by polychaete worms, ostracods, and kinorhynchs. The dramatic discharge to MOS sedimentation on the seafloor, (2) identify

increase in nematode abundance could be due to organic significant research questions, (3) determine the space and time

enrichment from sinking MOS and/or a higher tolerance to scales of processes and events, and (4) identify data gaps that may

hydrocarbons. Moreover, Schwing et al. (2015) determined that be important for further defining and modeling of MOSSFA

there was an 80–93% decline in benthic foraminifera following events. Examples of data gaps and questions relating to the

the oil spill up to 111 km from the wellhead, likely due to suboxic MOSSFA processes are provided below to help guide future

conditions (Hastings et al., 2015). Using principal component research. The MOSSFA Workshop Report (2014) provides a more

analysis, Schwing et al. (2015) strongly suggested that oil comprehensive list of questions and discussion of data gaps. The

exposure coupled to the onset of suboxic-anoxic conditions in letter-designated sections below correspond to topics illustrated

the sediments were the factors controlling foraminifera decline. A in Fig. 4.

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 29

7.1. Physical and chemical characteristics of oil droplets and associated 7.3.3. New sources of material at the surface (C3)

hydrates (A) Oil was burned at the surface and pyrogenic particles were

observed in the sediments. The role of aerosols and burning oil

The characteristics of oil released at a point source are by-products in MOS formation and sedimentation, however, are

determined by the type of oil released, water temperature, water unknown. Did Saharan dust input interact with the sea surface

depth (ambient pressure), and the pressure within the petroleum microlayer to influence MOS formation? What was the role of

reservoir. Laboratory experiments and models analyses of oil (i.e., charred combustion products of burning oil) in particle

droplet size distributions are required over a range of conditions formation and sedimentation?

(pressure, turbulence) to predict the fate of bubbles of gas hydrates

and oil droplets as they rise to the surface. During the DWH oil spill,

7.3.4. Sinking of oil-derived material from the surface (C4)

chemical dispersants were applied directly to the deep subsea oil

Increasing the density of marine snow was key to forming

jet to promote smaller oil droplet size and longer residence time in

sinking particles. Although the general physical and biological

the water column. The consequences of this action need further

investigation. factors contributing to MOS formation and sedimentation are

known, there is little information on the composition of MOS

particles and the factors controlling its spatial and temporal

7.2. Formation of subsurface oil plumes: droplet and bubble size (B)

variation and sinking rates. Lithogenic deposition was evident in a

sediment trap deployed five km SW of the DWH site and on the

Deep horizontal intrusions of dissolved gas and oil formed at

seafloor shortly after the oil spill. However, the role of OMAs

mid-water depths, which impacted the mesopelagic environment,

(mineral ballast) in oil sedimentation and whether OMAs

sedimentation of oil, and sediments where plumes impinged on

aggregated with MOS particles over the broad area of the oil spill

the and slope. In addition, MOS particles may

is not known.

have formed in the plumes and marine snow particles that formed

at the surface may have entrained additional oil as they sank

through the plumes. Many questions remain relating to the

7.4. Processes impacting MOS during sedimentation (D)

characteristics of oil in the plumes, the role of the physical

environment in plume formation and spatial extent, the plume’s

Microbial activity, turbulent disruption, and zooplankton

role in the MOSSFA event, and an evaluation of the conditions

fragmentation transform sinking aggregates. Currents laterally

under which subsurface plumes may form during future oil well

advect sinking particles. There is evidence that a subsurface

blowouts.

nepheloid layer occurred off the shelf slope during the oil spill. It is

not known how surface-formed sinking particles interacted with

7.3. Important processes in the surface mixed layer (C)

suspended sediment layers and/or the deep subsurface oil plumes,

or to what extent particles were laterally advected. Clay

All surface ocean processes that impacted MOS formation and

mineralogy, the size range of clay particles, and oil emulsions

sedimentation varied spatially and temporally. Atmosphere-ocean

likely controlled whether oil droplet/particles are buoyant or sink,

interactions and submesoscale physical processes influenced the

as clay particles may stabilize emulsified droplets (Sullivan and

spatial distribution of oil, marine biota, and MOS. Winds and

Kilpatrick, 2002). The extent to which sinking aggregates

current shear spread the surface oil slick. Breaking waves disperse

scavenged organisms, organic detritus, inorganic particles, and

oil into the water column resulting in oil droplets of different sizes.

other oil droplets is uncertain.

The impacts of large storms are unknown. Continued model

development is needed to better understand the role of physical

processes in oil and MOS aggregation and sedimenta- 7.5. Processes impacting MOS on the seafloor (E)

tion, as well as the recovery (replacement via advection) of the

. Although it is clear that a significant percentage of DWH oil

sedimented to the sea floor, an accurate assessment of the amount

of oil and its spatial distribution remain to be determined. The

7.3.1. Interaction of oil droplets with euphotic zone plankton (C1)

dominant factors controlling the persistence and degradation of

Oil that rose to the surface encountered high concentrations of

hydrocarbons in sediments needs further investigation. What are

phytoplankton fueled by nutrients from river discharge. Dis-

the roles of microbial communities, redox interactions, and

persants also were applied at the surface. Laboratory experiments

sediment oxygen demand in the degradation of hydrocarbons?

demonstrated that MOS particles were formed by interactions of

In particular, long-term studies are needed to determine microbial

oil with bacteria, phytoplankton, and zooplankton. Did the oil

rates of hydrocarbon degradation over time and to assess the

spill contribute to the observed enhanced phytoplankton bloom?

recovery rate of benthic communities.

How did variation in oil and dispersant concentrations change

microbial mucus formation and zooplankton behavior and

survival?

7.6. Processes impacting resuspension of oiled sediments (F)

7.3.2. Transformation of oil floating at the atmosphere-ocean interface

The role of water motion in resuspension, disaggregation, and

(C2)

re-aggregation of oiled sediments in benthic boundary layers is

UV light acts to weather oil and evaporation of more volatile

poorly known. What is the role of storm-generated deep currents

constituents acts to change oil composition and density over time.

in resuspending oiled sediments? How do these processes impact

Both evaporation and emulsification increases the density and

benthic communities? There is little information on the role of

viscosity of the surface slick. Why does weathered oil enhance

natural oil seeps in accumulation and re-suspension of oiled

MOS aggregation in experiments? Microbial activity was shown to

sediments. Are marine snow particles produced near natural

create large flocs early during the oil spill. Why did this phenomena seeps?

change over time? How did dispersants impact MOS formation?

30 K.L. Daly et al. / Anthropocene 13 (2016) 18–33

8. Conclusions Acknowledgements

The DWH MOSSFA event was unexpected, but now recognized The MOSSFA workshop was supported by a grant from the Gulf

to be a significant pathway for the distribution and fate of oil. The of Mexico Research Initiative. Research support was provided by

contribution to sedimentation and interactions between MOS and the University of South Florida Division of Sponsored Research and

OMA particles will vary with distance from shore and river water Florida Institute of (FIO)/BP to Daly, and grants from

plumes. Evidence suggests that the prolonged DWH sedimentation the Gulf of Mexico Research Initiative through its consortia: Center

event may have been comprised of numerous smaller sedimenta- for Integrated Modeling and Analysis of Gulf Ecosystems

tion events during and after the spill, with the overall accumula- (C-IMAGE) to Daly, Hollander, and Chanton, Deep-C to Chanton

tion at the seafloor of as much as 14% of the total oil released, and Hollander, and Ecosystem Impacts of Oil & Gas Inputs to the

thereby relocating oil to mesopelagic and deep benthic ecosys- Gulf (ECOGIG) to Passow and Chanton. We thank the workshop

tems. Clearly, understanding the impact of MOS processes on the participants (http://deep-c.org/mossfa/) for their significant con-

transport and fate of oil spilled into the environment is important tributions to the workshop discussions. We are especially grateful

and should be considered for future oil spill assessments. to Dr. Nancy Kinner for facilitating the workshop and Dr. Peter

Consequences of the MOSSFA event include the fact that MOS Kinner for organizing and editing the final report. Meredith Fields

particles in surface waters and aggregates sinking through the and Kathy Mandsager provided excellent logistical support. A.

water column were remineralized by bacteria and/or ingested by Remsen and K. Kramer provided analyses of marine snow. We also

zooplankton. Assimilation of petrocarbon into food web thank the reviewers for thoughtful comments and suggestions.

would be a more favorable pathway for marine food webs than Data are publically available through the Gulf of Mexico Research

bioaccumulation of oil in organisms, which negatively impact Initiative Information and Data Cooperative (GRIIDC) at http://

higher trophic levels. In addition, the accumulation of DWH oil on data.gulfresearchinitiative.org (doi: 10.7266/N78P5XFP, http://dx.

the seafloor may have a relatively long residence time, with doi.org/10.7266/N76T0JKS).

continued metabolism of toxic and carcinogenic hydrocarbons by

microbial communities and, therefore, protracted exposure by References

ecologically sensitive benthic animals and economically and

Alldredge, A.L., Silver, M.W., 1988. Characteristics: dynamics and significance of

recreationally important fish. Responders to future oil incidents

marine snow. Prog. Oceanogr. 20, 41–82.

should consider the possibility of a MOSFFA event when evaluating

Almeda, R., Connelly, T.L., Buskey, E.J., 2014a. Novel insight into the role of

appropriate actions to limit the impacts of oil spills. heterotrophic dinoflagellates in the fate of crude oil in the sea. Sci. Rep. 4, 7560.

doi:http://dx.doi.org/10.1038/srep07560.

The large DWH MOSSFA incident occurred due to a nexus of

Almeda, R., Hyatt, C., Buskey, E.J., 2014b. of dispersant Corexit 9500A and

events in a similar manner to the events which created Superstorm

crude oil to marine microzooplankton. Ecotoxicol. Environ. Saf. 106, 76–7685.

Sandy on the Atlantic coast of the USA in 2012: (1) the DWH site Almeda, R., Baca, S., Hyatt, C., Buskey, E.J., 2014c. Ingestion and sublethal effects of

physically and chemically dispersed crude oil on marine planktonic copepods.

was located in one of the most productive regions of the Gulf of

Ecotoxicology 23, 988–1003.

Mexico governed by Mississippi River outflow, (2) the spill

Almeda, R., Wambaugh, Z., Chai, C., Wang, Z., Liu, Z., Buskey, E.J., 2013a. Effects of

occurred during spring and summer when bacteria and phyto- crude oil exposure on bioaccumulation of polycyclic aromatic hydrocarbons and

plankton concentrations were at a maximum and surface survival of adult and larval stages of gelatinous zooplankton. PLoS One 8 (10),

e74476. doi:http://dx.doi.org/10.1371/journal.pone.0074476.

community activity rates were relatively high, (3) the large

Almeda, R., Wambaugh, Z., Chai, C., Wang, Z., Liu, Z., Buskey, E.J., 2013b. Interactions

Mississippi River discharge enhanced phytoplankton production

between zooplankton and crude oil: toxic effects and bioaccumulation of

and suspended particle concentrations, and (4) there was polycyclic aromatic hydrocarbons. PLoS One 8 (6), e67212. doi:http://dx.doi.

org/10.1371/journal.pone.0067212.

enhanced microbial mucus formation, especially in the presence

Aman, Z.M., Paris, C.B., Maya, E.F., Johns, M.L., Lindo-Atichati, D., 2015. High-

of weathered oil. Nevertheless, the DWH event was not unique as

pressure visual experimental studies of oil-in-water dispersion droplet size.

several reports indicate that oil sedimentation was observed Chem. Eng. Sci. 127, 392–400.

Armstrong, R.A., Lee, C., Hedges, J.I., Honjo, S., Wakeham, S.G., 2002. A new,

during the Tsesis, Ixtoc-I, and other oil spills (Johansson et al.,1980;

mechanistic model for organic carbon fluxes in the ocean: based on the

Boehm and Fiest, 1980; Jernelöv and Lindén, 1981; Patton et al.,

quantitative association of POC with ballast minerals. Deep-Sea Res. II 49, 219–

1981; Teal and Howarth, 1984; Vonk et al., 2015). Indeed, such 236.

Arnosti, C., Ziervogel, K., Yang, T., Teske, A., 2015. Oil-derived marine aggregates—

observations suggest that MOSSFA events have a high probability

hot spots of polysaccharide degradation by specialized bacterial communities.

of occuring during future oil spills in coastal and deep-water sites,

Deep-Sea Res. II doi:http://dx.doi.org/10.1016/j.dsr2.2014.12.008.

particularly since 85% of deep- water oil exploration sites Azam, F., Long, R.A., 2001. Sea snow microcosms. Nature 414, 495–498.

Bacosa, H.P., Erdner, D.L., Liu, Z., 2015. Differentiating the roles of photooxidation

worldwide are adjacent to deltaic systems (Weimer and Pettingill,

and biodegradation in the weathering of Light Louisiana Sweet crude oil in

2007). Despite the large academic and government response to the

surface water from the Deepwater Horizon site. Mar. Pollut. Bull. 95, 265–272.

oil spill, there remain signi cant data gaps in understanding the Baelum, J., Borglin, S., Chakraborty, R., Fortney, J., Lamendella, R., Mason, O., Auer, M.,

dominant factors controlling the temporal and spatial variability of Zemla, B.M., Conrad, M., Malfatti, S., Tringe, S., Holman, H., Hazen, T., Jansson, J.,

2012. Deep-sea bacteria enriched by oil and dispersant from the Deepwater

the DWH MOSSFA event. In this paper, we provide a conceptual

Horizon spill. Environ. Microbiol. 14, 2405–2416.

framework for understanding MOSSFA processes to help guide

Baguley, J.G., Montagna, P.A., Cooksey, C., Hyland, J.L., Bang, H.W., Morrison, C.,

current research on the DWH oil spill and which could be applied Kamikawa, A., Bennetts, P., Saiyo, G., Parsons, E., Herdener, M., Ricci, M., 2015.

Community response of deep-sea soft-sediment metazoan meiofauna to the

to other locations. Mechanistic models of MOS formation and

Deepwater Horizon blowout and oil spill. Mar. Ecol. Prog. Ser. 528, 127–140.

sedimentation coupled to circulation models need to be developed

Balsam, W.L., Beeson, J.P., 2003. Sea-floor sediment distribution in the Gulf of

to use as a tool to manage risk and improve our ability to predict Mexico. Deep-Sea Res. I 50, 1421–1444.

Bernhardt, J.R., Leslie, H.M., 2013. Resilience to in coastal marine

the extent of MOSSFA events under different conditions and

ecosystems. Annu. Rev. Mar. Sci. 5, 371–392.

environments. Furthermore, baseline time-series data are urgently

Bianchi, T.S., Galler, J.J., Allison, M.A., 2007. Hydrodynamic sorting and transport of

needed for all levels of the ecosystem in regions of hydrocarbon terrestrially derived organic carbon in sediments of the Mississippi and

Atchafalaya Rivers. Estuarine Coastal Shelf Sci. 73, 211–222.

extraction to prepare for and respond to oil spills and to

Bianchi, T.S., Cook, R.L., Perdue, E.M., Kolic, P.E., Green, N., Zhang, Y., Smith, R.W.,

understand the impacts of oil spills on the environment,

Kolker, A.S., Ameen, A., King, G., Ojwang, L.M., Schneider, C.L., Normand, A.E.,

particularly in sensitive ecosystems such as the Arctic. Hetland, R., 2011. Impacts of diverted freshwater on dissolved organic matter

and microbial communities in Barataria Bay, Louisiana, U. S. A. Mar. Environ.

Res. 72, 248–257.

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 31

Boehm, P.D., Fiest, D.L., 1980. Aspects of the transport of petroleum hydrocarbons to Folke, C., Carpenter, S., Walker, B., Scheffer, M., Elmqvist, T., Gunderson, L., Holling, C.

the offshore benthos during the Ixtoc-I blowout in the Bay of Campeche. In: S., 2004. Regime shifts, resilience, and biodiversity in ecosystem management.

Proceedings of the Symposium on the Preliminary Results from the September, Annu. Rev. Ecol. Evol. Syst. 35, 557–581.

1979 Pierce/Research IXTOC-1Cruises. Key Biscayne, Florida, June 9-10, 1980. Fu, J., Gong, Y., Zhao, X., Reilly, S.E., Zhao, D., 2014. Effects of oil and dispersant on

Publications Office, NOAA/RD/MP3, Office of Assessment, formation of marine oil snow and transport of oil hydrocarbons. Environ. Sci.

NOAA, US Dept. of Commerce, 325 Broadway, Boulder, CO 80303, USA. Technol. 48, 14392–14399.

Boehm, P.D., 1987. Transport and transformation processes regarding hydrocarbon Gardner, W.D., Walsh, I.D., 1990. Distribution of macroaggregates and fine-grained

and metal pollutants in offshore sedimentary environments. In: Boesch, D.F., particles across a and their potential role in fluxes. Deep-Sea

Rabalais, N.N. (Eds.), Long-Term Environmental Effects of Offshore Oil and Gas Res. 37 (3), 401–411.

Development. Elsevier Applied Science, London & , pp. 233–287. Garr, A.L., Laramore, S., Krebs, W., 2014. Toxic effects of oil and dispersant on marine

Bowles, M.W., Samarkin, V.A., Joye, S.B., 2011. Improved measurement of microbial microalgae. Bull. Environ. Contam. Toxicol. 93, 654–659.

activity in deep-sea sediments at in situ pressure and methane concentration. Goni, G.J., Trinanes, J.A., MacFadyen, A., Streett, D., Olascoaga, M.J., Imhoff, M.L.,

Limnol. Oceanogr.: Methods 9, 499–506. Muller-Karger, F., Roffer, M.A., 2015. Variability of the Deepwater Horizon

Boyd, P.W., Trull, T.W., 2007. Understanding the export of biogenic particles in surface oil spill extent and its relationship to varying ocean currents and

oceanic waters: is there consensus? Prog. Oceanogr. 72, 276–312. extreme weather conditions. In: Ehrhardt, M. (Ed.), Mathematical Modelling

Brooks, G.R., Larson, R.A., Schwing, P.T., Romero, I., Moore, C., Reichart, G.-J., Jilbert, and Numerical Simulation of Oil Pollution Problems, The Reacting Atmosphere

T., Chanton, J.P., Hastings, D.W., Overholt, W.A., Marks, K.P., Kostka, J.E., Holmes, 2. Springer International doi:http://dx.doi.org/10.1007/978-3-319-16459-5_1.

C.W., Hollander, D., 2015. Sedimentation pulse in the NE Gulf of Mexico González, J., Fernández, E., Figueiras, F.G., Varela, M., 2013. Subtle effects of the

following the 2010 DWH Blowou. PLoS One doi:http://dx.doi.org/10.1371/ water soluble fraction of oil spills on natural phytoplankton assemblages

journal.pone.0132341 24 pp. enclosed in mesocosms. Estuarine Coastal Shelf Sci. 124, 13–23.

Burd, A.B., Jackson, G.A., 2009. Particle aggregation. Annu. Rev. Mar. Sci. 1, 65–90. Graham, W.M., Condon, R.H., Carmichael, R.H., D’Ambra, I., Patterson, H.K., Linn, L.J.,

Camilli, R., Reddy, C.M., Yoerger, D.R., Van Mooy, B.A.S., Jakuba, M.V., Kinsey, J.C., Hernandez Jr., F.J., 2010. Oil carbon entered the coastal planktonic food web

McIntyre, C.P., Sylva, S.P., Maloney, J.V., 2010. Tracking hydrocarbon plume during the Deepwater Horizon oil spill. Environmental Research Letters 5 doi:

transport and biodegradation at Deepwater Horizon. Science 330, 201–204. http://dx.doi.org/10.1088/1748-9326/5/4/045301 6 pp.

Chanton, J.P., Cherrier, J., Wilson, R.M., Sarkodee-Adoo, J., Bosman, S., Mickle, A., Guidi, L., Gorsky, G., Claustre, H., Miquel, J.C., Picheral, M., Stemmann, L., 2008.

Graham, W.M., 2012. Radiocarbon evidence that carbon from the Deepwater Distribution and fluxes of aggregates >100 mm in the upper kilometer of the

Horizon spill entered the planktonic food web of the Gulf of Mexico. Environ. South-Eastern Pacific. Biogeosciences 5, 1361–1372.

Res. Lett. 7 doi:http://dx.doi.org/10.1088/1748-9326/7/4/045303 4. Gutierrez, T., Aitken, M.D., 2014. Role of methylotrophs in the degradation of

Chanton, J., Zhao, T., Rosenheim, B.E., Joye, S., Bosman, S., Brunner, C., Yeager, K.M., hydrocarbons during the Deepwater Horizon oil spill. Int. Soc. Microb. Ecol. 8,

Diercks, A.R., Hollander, D., 2015. Using natural abundance radiocarbon to trace 2543–2545.

the flux of petrocarbon to the seafloor following the Deepwater Horizon oil spill. Gutierrez, T., Berry, D., Yang, T., Mishamandani, S., McKay, L., Teske, A., Aitken, M.D.,

Environ. Sci. Technol. 49, 847–854. 2013. Role of bacterial exopolysaccharides (EPS) in the fate of the oil released

Cherrier, J., Sarkodee-Adoo, J., Guilderson, T., Chanton, J.P., 2014. Fossil carbon in during the Deepwater Horizon Oil Spill. PLoS One 8 (6), e67717. doi:http://dx.

particulate organic matter in the Gulf of Mexico following the Deepwater doi.org/10.1371/journal.pone.0067717.

Horizon event. Environ. Sci. Technol. Lett. 1, 108–112. doi:http://dx.doi.org/ Hastings, D.W., Schwing, P.T., Brooks, G.R., Larson, R.A., Morford, J.L., Roeder, T.,

10.1021/ez400149c. Quinn, K.A., Bartlett, T., Romero, I.C., Hollander, D.J., 2015. Changes in

Cohen, J.H., McCormick, L.R., Burkhardt, S.M., 2014. Effects of dispersant and oil on sedimentary redox conditions following the BP DWH blowout event. Deep-Sea

survival and swimming activity in a marine copepod. Bull. Environ. Contam. Res. II doi:http://dx.doi.org/10.1016/j.dsr2.2014.12.009.29.

Toxicol. 92, 381–387. Hazen, T.C., Dubinsky, E.A., DeSantis, T.Z., Andersen, G.L., Piceno, Y.M., Singh, N.,

Corbett, D.R., McKeeb, B., Allison, M., 2006. Nature of decadal-scale sediment Jansson, J.K., Probst, A., Borglin, S.E., Fortney, J.L., Stringfellow, W.T., Bill, M.,

accumulation on the western shelf of the Mississippi . Cont. Shelf Res. Conrad, M.S., Tom, L.M., Chavarria, K.L., Alusi, T.R., Lamendella, R., Joyner, D.C.,

26, 2125–2140. Spier, C., Baelum, J., Auer, M., Zemla, M.L., Chakraborty, R., Sonnenthal, E.L.,

Crespo-Medina, M.C., Meile, D., Hunter, K.S., Diercks, A.-R., Asper, V.L., Chanton, J.P., D’Haeseleer, P., Holman, H.-Y.N., Osman, S., Lu, Z., Van Nostrand, J.D., Deng, Y.,

Orphan, V.J., Shiller, A.M., Battles, J.J., Joung, D.-J., Amon, R.M.W., Bracco, A., Zhou, J., Mason, O.U., 2010. Deep-sea oil plume enriches indigenous oil-

Montoya, J.P., Villareal, T.A., Vossmeyer, A., Wood, A.M., Joye, S.B., 2014. The rise degrading bacteria. Science 330, 204–208.

and fall of methanotrophy following a deepwater oil-well blowout. Nat. Geosci. Hill, P.S., Khelifa, A., Lee, K., 2002. Time scale for oil droplet stabilization by mineral

7, 423–427. doi:http://dx.doi.org/10.1038/ngeo2156. particles in turbulent suspensions. Spill Sci. Technol. Bull. 8 (1), 73–81.

Daling, P.S., Leirvik, F., Almås, I.K., Brandvik, P.J., Hansen, B.H., Lewis, A., Reed, M., Hsing, P.-Y., Fu, B., Larcom, E.A., Berlet, S.B., Shank, T.M., Govindarajan, A.F.,

2014. Surface weathering and dispersibility of MC252 crude oil. Mar. Pollut. Bull. Lukasiewicz, A.J., Dixon, P.M., Fisher, C.R., 2013. Evidence of lasting impact of the

87, 300–310. Deepwater Horizon oil spill on a deep Gulf of Mexico coral community.

Daly, K.L., Schwierzke-Wade, L., Dreger, K., Murasko, S., Outram, D., Remsen, A., Elementa: Sci. Anthropocene 1, 000012. doi:http://dx.doi.org/10.12952/journal.

2013. Plankton dynamics following the BP oil spill. Gulf of Mexico Oil Spill and elementa.000012.

Ecosystem Science Conference, January 21–23, New Orleans, LA. Hu, C., Weisberg, R.H., Liu, Y., Zheng, L., Daly, K.L., English, D.C., Zhao, J., Vargo, G.A.,

Daly, K., Kramer, K., Remsen, A., 2014. Oil sedimentation pathway: marine snow 2011. Did the northeastern Gulf of Mexico become greener after the Deepwater

distributions in the NE Gulf of Mexico, 2010–2013. Gulf of Mexico Oil Spill and Horizon oil spill? Geophys. Res. Lett. 38, L09601. doi:http://dx.doi.org/10.1029/

Ecosystem Science Conference, January 27–30, Mobile, AL. 2011GL047184.

Dagg, M.J., Breed, G.A., 2003. Biological effects of Mississippi River nitrogen on the Iverson, M.H., Ploug, H., 2010. Ballast minerals and the sinking carbon flux in the

northern Gulf of Mexico—a review and synthesis. J. Mar. Syst. 43, 133–152. ocean: carbon-specific respiration rates and sinking velocity of marine snow

Danchuk, S., Willson, C.S., 2011. Influence of seasonal variability of lower Mississippi aggregates. Biogeosciences 7, 2613–2624.

River discharge, temperature, suspended sediments, and salinity on oil–mineral Jernelöv, A., Lindén, O., 1981. Ixtoc I: a case study of the world’s largest oil spill.

aggregate formation. Water Environ. Res. 83 (7), 579–587. AMBIO 10 (6), 299–306.

De La Rocha, C., Passow, U., 2007. Factors influencing the sinking of POC and the Jochens, A.E., DiMarco, S.F., 2008. Physical oceanographic conditions in the

efficiency of the biological carbon pump. Deep-Sea Res. II 54, 639–658. deepwater Gulf of Mexico in summer 2000–2002. Deep-Sea Res. II 55, 2541–

DeLeo, D.M., Ruiz-Ramos, D.V., Baums, I.B., Cordes, E.E., 2015. Response of deep- 2554.

water corals to oil and chemical dispersant exposure. Deep-Sea Res. II doi: Johansson, S., Larsson, U., Boehm, P., 1980. The Tsesis oil spill impact on the pelagic

http://dx.doi.org/10.1016/j.dsr2.2015.02.028i. ecosystem. Mar. Pollut. Bull. 11 (10), 284–293.

Diercks, A.-R., Asper, V.L., 1997. In situ settling speeds of marine snow aggregates Jones, E.B., Wiggert, J.D., 2015. Characterization of a high chlorophyll plume in the

below the mixed layer: and Gulf of Mexico. Deep-Sea Res. I 44 (3), northeastern Gulf of Mexico. Remote Sens. Environ. 159, 152–166.

385–398. Joye, S.B., 2015. Deepwater Horizon, 5 years on. Science 349, 592–593.

Diercks, A.R., Highsmith, R.C., Asper, V.L., Joung, D., Zhou, Z., Guo, L., Shiller, A.M., Joye, S.B., Teske, A.P., Kostka, J.E., 2014. Microbial dynamics following the Macondo

Joye, S.B., Teske, A.P., Guinasso, N., Wade, T.L., Lohrenz, S.E., 2010. oil well blowout across Gulf of Mexico environments. BioScience 64, 766–777.

Characterization of subsurface polycyclic aromatic hydrocarbons at the Joye, S.B., MacDonald, I.R., Leifer, I., Asper, V., 2011a. Magnitude and oxidation

Deepwater Horizon site. Geophys. Res. Lett. 37, L20602. doi:http://dx.doi.org/ potential of hydrocarbon gases released from the BP oil well blowout. Nat.

10.1029/2010GL045046. Geosci. 4, 160–164.

Dubinsky, E.A., Conrad, M.E., Chakraborty, R., Bill, M., Borglin, S.E., Hollibaugh, J.T., Joye, S.B., Leifer, I., MacDonald, I.R., Chanton, J.P., Meile, C.D., Teske, A.P., Kostka, J.E.,

Mason, O.U., Piceno, Y.M., Reid, F.C., Stringfellow, W.T., Tom, L.M., Hazen, T.C., Chistoserdova, L., Coffin, R., Hollander, D., Kastner, M., Montoya, J.P., Rehder, G.,

Andersen, G.L., 2013. Succession of hydrocarbon-degrading bacteria in the Solomon, E., Treude, T., Villareal, T.A., 2011b. Comment on a persistent oxygen

aftermath of the Deepwater Horizon oil spill in the Gulf of Mexico. Environ. Sci. anomaly reveals the fate of spilled methane in the deep Gulf of Mexico. Science

Technol. 47, 10860–10867. 332, 1033. doi:http://dx.doi.org/10.1126/science.1203307.

Fernández, A., Rogers, K.R., Weber, S.C., Chanton, J.P., Montoya, J.P., 2016. Deepwater Keen, T.R., Allen, S.E., 2000. The generation of internal waves on the continental

Horizon oil and methane carbon entered the food web in the Gulf. Limnol. shelf by Hurricane Andrew. J. Geophys. Res. 105 (C11), 26203–26224.

Oceanogr. (in review). Kessler, J.D., Valentine, D.L., Redmond, M.C., Du, M., Chan, E.W., Mendes, S.D., Quiroz,

Fisher, C.R., Hsing, P.-Y., Kaiser, C.L., Yoerger, D.R., Roberts, H.H., Shedd, W.W., Cordes, E.W., Villanueva, C.J., Shusta, S.S., Werra, L.M., Yvon-Lewis, S.A., Weber, T.C.,

E.E., Shank, T.M., Berlet, S.P., Saunders, M.G., Larcom, E.A., Brooks, J.M., 2014. 2011. A persistent oxygen anomaly reveals the fate of spilled methane in the

Footprint of Deepwater Horizon blowout impact to deep-water coral deep Gulf of Mexico. Science 331, 312–315.

communities. Proc. Natl. Acad. Sci. 111 (32), 11744–11749.

32 K.L. Daly et al. / Anthropocene 13 (2016) 18–33

Khelifa, A., Stoffyn-Egli, P., Hill, P.S., Lee, K., 2005. Effects of salinity and clay type on Meador, J.P., 2003. Bioaccumulation of PAHs in . In: Douben, P.

oil–mineral aggregation. Mar. Environ. Res. 59, 235–254. E.T. (Ed.), PAHs: An Ecological Perspective. John Wiley and Sons, IncHoboken, NJ,

Kimes, N.E., Callaghan, A.V., Suflita, J.M., Morris, P.J., 2014. Microbial transformation pp. 147–172.

of the Deepwater Horizon oil spill—past, present, and future perspectives. Front. Mishamandani, S., Gutierrez, T., Berry, D., Aitken, M.D., 2015. Response of the

Microbiol. 5 doi:http://dx.doi.org/10.3389/fmicb.2014.00603. bacterial community associated with a cosmopolitan marine diatom to crude oil

King, G.M., Kostka, J.E., Hazen, T.C., Sobecky, P.A., 2015. Microbial responses to the shows a preference for the biodegradation of aromatic hydrocarbons. Environ.

Deepwater Horizon oil spill: from coastal wetlands to the deep sea. Annu. Rev. Microbiol. doi:http://dx.doi.org/10.1111/1462-2920.12988.

Mar. Sci. 7, 377–401. Mitra, S., Kimmel, D.G., Snyder, J., Scalise, K., McGlaughon, B.D., Roman, M.R., Jahn, G.

Kinner, N.E., Belden, L., Kinner, P., 2014. Unexpected sink for Deepwater Horizon oil L., Pierson, J.J., Brandt, S.B., Montoya, J.P., Rosenbauer, R.J., Lorenson, T.D., Wong,

may influence future spill response. Eos 95 (21), 1. F.L., Campbell, P.L., 2012. Macondo-1 well oil-derived polycyclic aromatic

Kleindienst, S., Paul, J.H., Joye, S.B., 2015a. Using dispersants after oil spills: impacts hydrocarbons in mesozooplankton from the northern Gulf of Mexico. Geophys.

on the composition and activity of microbial communities. Nat. Rev. Microbiol. Res. Lett. 39, L01605. doi:http://dx.doi.org/10.1029/2011GL049505.

13, 388–396. doi:http://dx.doi.org/10.1038/nrmicro3452. Montagna, P.A., Baguley, J.G., Cooksey, C., Hartwell, I., Hyde, L.J., Hyland, J.L., Kalke, R.

Kleindienst, S., Seidel, M., Ziervogel, K., Grim, S., Loftis, K., Perkins, M.J., Field, J., D., Kracker, L.M., Reuscher, M., Rhodes, A.C.E., 2013. Deep-sea benthic footprint

Sogin, M.L., Dittmar, T., Passow, U., Medeiros, P.M., Joye, S.B., 2015b. Chemical of the Deepwater Horizon blowout. PLoS One 8 (8), e70540. doi:http://dx.doi.

dispersants can suppress the activity of natural oil-degrading microorganisms. org/10.1371/journal.pone.0070540.

Proc. Natl. Acad. Sci. 112 (48), 14900–14905. Morey, S.L., Schroeder, W.W., O’Brien, J.J., Zavala-Hidalgo, J., 2003. The annual cycle

Klinger, D.H., Dale, J.J., Machado, B.E., Incardona, J.P., Farwell, C.J., Block, B.A., 2015. of riverine influence in the eastern Gulf of Mexico basin. Geophys. Res. Lett. 30

Exposure to Deepwater Horizon weathered crude oil increases routine (16), 1867. doi:http://dx.doi.org/10.1029/2003GL017348.

metabolic demand in chub , Scomber japonicas. Mar. Pollut. Bull. 98, MOSSFA Workshop Report, 2014. Marine Oil Snow Sedimentation and Flocculent

259–266. Accumulation (MOSSFA) Workshop, October 22–23, 2013, Tallahassee ,FL,

Koelmans, A.A., Jonker, M.T.O., Cornelissen, G., Bucheli, T.D., Van Noort, P.C.M., 37 pp. http://deep-c.org/images/documents/

Gustafsson, Ö., 2006. Black carbon: the reverse of its dark side. Chemosphere 63, MOSSFA_WorkshopReport2014.01.17.pdf.

365–377. Muller-Karger, F.E., Smith, J.P., Werner, S., Chen, R., Roffer, M., Liu, Y., Muhling, B.,

Kolian, S.R., Porter, S.A., Sammarco, P.W., Birkholz, D., Cake Jr., E.W., Subra, W.A., Lindo-Atichati, D., Lamkin, J., Cerdeira-Estrada, S., Enfield, D.B., 2015. Natural

2015. Oil in the Gulf of Mexico after the capping of the BP/Deepwater Horizon variability of surface oceanographic conditions in the offshore Gulf of Mexico.

Mississippi Canyon (MC-252) well. Environ. Sci. Pollut. Res. 22, 12073–12082. Prog. Oceanogr. 134, 54–76.

Kourafalou, V.H., Androulidakis, Y.S., 2013. Influence of Mississippi River induced Murawski, S.A., Hogarth, W.T., Peebles, E.B., Barbeiri, L., 2014. Prevalence of external

circulation on the Deepwater Horizon oil spill transport. J. Geophys. Res. Ocean skin lesions and polycyclic aromatic hydrocarbon concentrations in Gulf of

118, 3823–3842. Mexico fishes, post-Deepwater Horizon. Trans. Am. . Soc. 143 (4), 1084–

Kujawinski, E.B., Kido Soule, M.C., Valentine, D.L., Boysen, A.K., Longnecker, K., 1097.

Redmond, M.C., 2011. Fate of dispersants associated with the Deepwater Muschenheim, D.K., Lee, K., 2002. Removal of oil from the sea surface through

Horizon oil spill. Environ. Sci. Technol. 45, 1298–1306. particulate interactions: review and prospectus. Spill Sci. Technol. Bull. 8 (1), 9–

Lee, K., 2002. Oil–particle interactions in aquatic environments: influence on the 18.

transport, fate, effect and remediation of oil spills. Spill Sci. Technol. Bull. 8 (1), Nababan, B., Muller-Karger, F.E., Hu, C., Biggs, D., 2011. Chlorophyll variability in the

3–8. northeastern Gulf of Mexico. Int. J. Remote Sens. 32 (23), 8373–8391.

Lee, R.F., Page, D.S., 1997. Petroleum hydrocarbons and their effects in subtidal Norse, E.A., Amos, J., 2010. Impacts, perception, and policy implications of the

regions after major oil spills. Mar. Pollut. Bull. 34 (11), 928–940. Deepwater Horizon oil and gas disaster. Environ. Law Rep. 40,11058–11073 ISSN

Lee, R.F., Köster, M., Paffenhöfer, G.A., 2012. Ingestion and defecation of dispersed oil 0046-2284.

droplets by pelagic tunicates. J. Plankton Res. 34, 1058–1063. North, E.W., Adams, E.E., Thessen, A.E., Schlag, Z., He, R., Socolofsky, S.A., Masutani, S.

Le Hénaff, M., Kourafalou, V.H., Paris, C.B., Helgers, J., Aman, Z.M., Hogan, P.J., M., Peckham, S.D., 2015. The influence of droplet size and biodegradation on the

Srinivasan, A., 2012. Surface evolution of the Deepwater Horizon oil spill patch: transport of subsurface oil droplets during the Deepwater Horizon spill: a

combined effects of circulation and wind-induced drift. Environ. Sci. Technol. model sensitivity study. Environ. Res. Lett. 10 doi:http://dx.doi.org/10.1088/

46, 7267–7273. 1748-9326/10/2/024016.

Lehr, B., Bristol, S., Possolo, A., 2010. Oil Budget Calculator Deepwater Horizon. Olascoaga, M.J., Beron-Vera, F.J., Haller, G., Triñanes, J., Iskandarani, M., Coelho, E.F.,

Technical Documentation, A Report to the National Incident Command, Haus, B.K., Huntley, H.S., Jacobs, G., Kirwan Jr., A.D., Lipphardt Jr., B.L., Özgökmen,

November 2010. http://www.restorethegulf.gov/sites/default/files/documents/ T.M., Reniers, A.J.H.M., Valle-Levinson, A., 2013. Drifter motion in the Gulf of

pdf/OilBudgetCalc_Full_HQ-Print_111110.pdf. Mexico constrained by altimetric Lagrangian coherent structures. Geophys. Res.

Lincoln, S.A., Holland, D.J., Freeman, K.H., 2015. Marine snow and methanotrophy: Lett. 40, 6171–6175. doi:http://dx.doi.org/10.1002/2013GL058624.

microbial responses and hydrocarbon sedimentation following the Deepwater Ortmann, A.C., Anders, J., Shelton, N., Gong, L., Moss, A.G., Condon, R.H., 2012.

Horizon oil spill. International Meeting of Organic Geochemistry, September Dispersed oil disrupts microbial pathways in pelagic food webs. PLoS One 7 (7),

2015, Prague, The Czech Republic. 42548e. doi:http://dx.doi.org/10.1371/journal.pone.0042548.

Lindo-Atichati, D., Paris, C.B., LeHénaff, M., Schedler, M., ValladaresJuárez, A.G., Paris, C.B., Le He’naff, M., Aman, Z.M., Subramaniam, A., Helgers, J., Wang, D.-P.,

Müller, R., 2014. Simulating the effects of droplet size, high-pressure Kourafalou, V.H., Srinivasa, A., 2012. Evolution of the macondo well blowout:

biodegradation, and variable flow rate on the subsea evolution of deep plumes simulating the effects of the circulation and synthetic dispersants on the subsea

from the Macondo blowout. Deep-Sea Res. II doi:http://dx.doi.org/10.1016/j. oil transport. Environ. Sci. Technol. 46, 13293–13302.

dsr2.2014.01.011i. Parsons, M.L., Turner, R.E., Overton, E.B., 2014. Sediment-preserved diatom

Lohrenz, S.E., Fahnensteil, G.L., Redalje, D.G., Lang, G.A., Chen, X., Dagg, M.J., 1997. assemblages can distinguish a petroleum activity signal separately from the

Variations in of northern Gulf of Mexico continental shelf nutrient signal of the Mississippi River in coastal Louisiana. Mar. Pollut. Bull. 85,

waters linked to nutrient inputs from the Mississippi River. Mar. Ecol. Prog. Ser. 164–171.

155, 45–54. Passow, U., Ziervogel, K., Asper, V., Dierks, A., 2012. Marine snow formation in the

Lubchenco, J., McNutt, M.K., Dreyfus, G., Murawski, S.A., Kenedy, D.M., Anastas, P.T., aftermath of the Deepwater Horizon oil spill in the Gulf of Mexico. Environ. Res.

Chu, S., Hunter, T., 2012. Science in support of the Deepwater Horizon response. Lett. 7, 11. doi:http://dx.doi.org/10.1088/1748-9326/7/3/035301.

Proc. Natl. Acad. Sci. 109 (50), 20212–20221. Passow, U., 2014. Formation of rapidly-sinking, oil-associated marine snow. Deep-

MacDonald, I.R., Garcia-Pineda, O., Beet, A., Daneshgar Asl, S., Feng, L., Graettinger, Sea Res. II doi:http://dx.doi.org/10.1016/j.dsr2.2014.10.001.

G., French-McCay, D., Holmes, J., Hu, C., Huffer, F., Leifer, I., Mueller-Karger, F., Patton, J.S., Rigler, M.W., Boehm, P.D., Fiest, D.L., 1981. Ixtoc I oil spill: flaking of

Solow, A., Silva, M., Swayze, G., 2015. Natural and unnatural oil slicks in the Gulf surface mousse in the Gulf of Mexico. Nature 290, 235–238.

of Mexico. J. Geophys. Res. Oceans doi:http://dx.doi.org/10.1002/2015JC011062. Payne, J.R., Clayton Jr., J.R., Kirstein, B.E., 2003. Oil/suspended particulate material

Mari, X., Lefèvre, J., Torréton, J.P., Bettarel, Y., Pringault, O., Rochelle-Newall, E., interactions and sedimentation. Spill Sci. Technol. Bull. 8 (2), 201–221.

Marchesiello, P., Menkes, C., Rodier, M., Migon, C., Motegi, C., Weinbauer, M.G., Peiffer, R.F., Cohen, J.H., 2015. Lethal and sublethal effects of oil, chemical dispersant,

Legendre, L., 2014. Effects of soot deposition on particle dynamics and microbial and dispersed oil on the ctenophore leidyi. Aquat. Biol. 23, 237–

processes in marine surface waters. Glob. Biogeochem. Cycles doi:http://dx.doi. 250.

org/10.1002/2014G 2014GB004878. Ploug, H., Terbrüggen, A., Kaufmann, A., Wolf-Gladrow, D., Passow, U., 2010. A novel

Mason, O.U., Scott, N.M., Gonzalez, A., Robbins-Pianka, A., Bælum, J., Kimbrel, J., method to measure particle sinking velocity in vitro, and its comparison to

Bouskill, N.J., Prestat, E., Borglin, S., Joyner, D.C., Fortney, J.L., Jurelevicius, D., three other in vitro methods. Limnol. Oceanogr.: Methods 8, 386–393.

Stringfellow, W.T., Alvarez-Cohen, L., Hazen, T.C., Knight, R., Gilbert, J.A., Jansson, Prince, R.C., 2015. Oil spill dispersants: boon or bane? Environ. Sci. Technol. 49,

J.K., 2014. Metagenomics reveals sediment microbial community response to 6376–6384.

Deepwater Horizon oil spill. Int. Soc. Microb. Ecol. J. 8, 1464–1475. Prouty, N.G., Fisher, C.R., Demopoulos, A.W.J., Druffel, E.R.M., 2014. Growth rates and

McGenity, T.J., Folwell, B., McKew, B.A., Sanni, G.O., 2012. Marine crude-oil ages of deep-sea corals impacted by the Deepwater Horizon oil spill. Deep-Sea

biodegradation: a central role for interspecies interactionsAquat. Biosyst. 8. . Res. II doi:http://dx.doi.org/10.1016/j.dsr2.2014.10.021.

19 pp http://www.aquaticbiosystems.org/content/8/1/10. Quintana-Rizzo, E., Torres, J.J., Ross, S.W., Romero, I., Watson, K., Goddard, E.,

13 15

McNutt, M.K., Chen, S., Lubchenco, J., Hunter, T., Dreyfus, G., Murawski, S.A., Hollander, D., 2015. d C and d N in deep-living fishes and shrimps after the

Kennedy, D.M., 2012. Applications of science and engineering to quantify and Deepwater Horizon oil spill, Gulf of Mexico. Mar. Pollut. Bull. 94, 241–250.

control the Deepwater Horizon oil spill. Proc. Natl. Acad. Sci. 109 (50), 20222– Reddy, C.M., Arey, J.S., Seewald, J.S., Sylva, S.P., Lemkau, K.L., Nelson, R.K., Carmichael,

20228. C.A., McIntyre, C., Fenwick, J., Ventura, G.T., Van Mooy, B.A.S., Camilli, R., 2012.

Composition and fate of gas and oil released to the water column during the

K.L. Daly et al. / Anthropocene 13 (2016) 18–33 33

Deepwater Horizon oil spill. Proc. Natl. Acad. Sci. doi:http://dx.doi.org/10.1073/ Valentine, D.L., Kessler, J.D., Redmond, M.C., Mendes, S.D., Heintz, M.B., Farwell, C.,

pnas.1101242108. Hu, L., Kinnaman, F., Yvon-Lewis, S., Du, M., Chan, E.W., Tigreros, F.G., Illanueva,

Redmond, M.C., Valentine, D.L., 2012. Natural gas and temperature structured a C.J., 2010. Propane respiration jump-starts microbial response to a deep oil spill.

microbial community response to the Deepwater Horizon oil spill. Proc. Natl. Science 330, 208–211.

Acad. Sci. 109 (50) doi:http://dx.doi.org/10.1073/pnas.1108756108. Venn-Watson, S., Garrison, L., Litz, J., Fougeres, E., Mase, B., Rappucci, G., Stratton, E.,

Romero, I.C., Schwing, P.T., Brooks, G.R., Larson, R.A., Hastings, D.W., Ellis, G., Carmichael, R., Odell, D., Shannon, D., Shippee, S., Smith, S., Staggs, L., Tumlin,

Goddard, E.A., Hollander, D.J., 2015. Hydrocarbons in deep-sea sediments M., Whitehead, H., Rowles, T., 2015. Demographic clusters identified within the

following the 2010 Deepwater Horizon Blowout in the Northeast Gulf of Mexico. northern Gulf of Mexico common bottlenose (Tursiops truncates)

PLoS One doi:http://dx.doi.org/10.1371/journal.pone.0128371 23 pp. unusual mortality event: January 2010–June 2013. PLoS One 10 (2), e0117248.

Ryerson, T.B., Camilli, R., Kessler, J.D., Kujawinski, E.B., Reddy, C.M., Valentine, D.L., doi:http://dx.doi.org/10.1371/journal.pone.0117248.

Atlas, E., Blake, D.R., de Gouw, J., Meinardi, S., Parrish, D.D., Peischl, J., Seewald, J. Verdugo, P., Alldredge, A.L., Azam, F., Kirchman, D.L., Passow, U., Santschi, P.H., 2004.

S., Warneke, C., 2012. Chemical data quantify Deepwater Horizon hydrocarbon The oceanic gel phase: a bridge in the DOM–POM continuum. Mar. Chem. 92,

flow rate and environmental distribution. Proc. Natl. Acad. Sci. 109 (50), 20246– 67–85.

20253. Vonk, S.M., Hollander, D.J., Murk, A.-T.J., 2015. Was the extreme and wide-spread

Sammarco, P.W., Kolian, S.R., Warby, R.A.F., Bouldin, J.L., Subra, W.A., Porter, S.A., marine oil-snow sedimentation and flocculent accumulation (MOSSFA) event

2013. Distribution and concentrations of petroleum hydrocarbons associated during the Deepwater Horizon blow-out unique? Mar. Pollut. Bull.100 (1), 5–12.

with the BP/Deepwater Horizon Oil Spill, Gulf of Mexico. Mar. Pollut. Bull. 73, doi:http://dx.doi.org/10.1016/j.marpolbul.2015.08.023.

129–143. Wade, T.L., Sericano, J.L., Sweet, S.T., Knap, A.H., Guinasso Jr., N.L., 2015. Spatial and

Schedler, M., Hiessl, R., Juárez, A.G.V., Gust, G., Muller,̈ R., 2014. Effect of high temporal distribution of water column total polycyclic aromatic hydrocarbons

pressure on hydrocarbon-degrading bacteria. AMB Express 4, 77. doi:http://dx. (PAH) and total petroleum hydrocarbons (TPH) from the Deepwater Horizon

doi.org/10.1186/s13568-014-0077-0. (Macondo) incident. Mar. Pollut. Bull. doi:http://dx.doi.org/10.1016/j.

Schwing, P.T., Flower, B.P., Romero, I.C., Brooks, G.R., Larson, R.A., Hastings, D.W., marpolbul.2015.12.002.

Hollander, D.J., 2015. Effects of the Deepwater Horizon oil blowout on deep sea Walsh, I.D., Gardner, W.D., 1992. A comparison of aggregate profiles with sediment

benthic foraminifera in the Northeastern Gulf of Mexico. PLoS One 10 (3) doi: trap fluxes. Deep-Sea Res. 19 (11–12), 1817–1834.

http://dx.doi.org/10.1371/journal.pone.0120565. Wawrik, B., Paul, J.H., 2004. Phytoplankton community structure and productivity

Shay, L.K., Elsberry, R.L., 1987. Near-inertial response to Hurricane along the axis of the Mississippi River plume in oligotrophic Gulf of Mexico

Frederic. J. Phys. Oceanogr. 17, 1249–1269. waters. Aquat. Microb. Ecol. 35, 185–196.

Silva, M., Etnoyer, P., MacDonald, I.R., 2015. Coral injuries observed at mesophotic Wei, C., Rowe, G.T., Hubbard, G.F., Scheltema, A.H., Wilson, G.D.F., Petrescu, I., Foster,

reefs after the Deepwater Horizon oil discharge. Deep-Sea Res. Part II doi: J.M., Wicksten, M.K., Chen, M., Davenport, R., Soliman, Y., Wang, Y., 2010.

http://dx.doi.org/10.1016/j.dsr2.2015.05.013. Bathymetric zonation of deep-sea macrofauna in relation to export of surface

Simon, M., Grossart, H.P., Schweitzer, B., Ploug, H., 2002. of phytoplankton production. Mar. Ecol. Prog. Ser. 399, 1–14.

organic aggregates in aquatic ecosystems. Aquat. Microb. Ecol. 28, 175–211. Weimer, P., Pettingill, H.S., 2007. Global overview of deep-water exploration and

Smith, D.C., Simon, M., Alldredge, A.L., Azam, F., 1992. Intense hydrolytic enzyme production. Am. Assoc. Pet. Geol. Stud. Geol. 56, 7–11. doi:http://dx.doi.org/

activity on marine aggregates and implications for rapid particle dissolution. 10.1306/1240888St56285.

Nature 359, 139–141. Wetz, M.S., Meredith, C.R., Paerl, H.W., 2009. Transparent exopolymer particles

Smith, R.H., Johns, E.M., Goni, G.J., Trinanes, J., Lumpkin, R., Wood, A.M., Kelble, C.R., (TEP) in a river-dominated : spatial–temporal distributions and an

Cummings, S.R., Lamkin, J.T., Privoznik, S., 2014. Oceanographic conditions in assessment of controls upon TEP formation. Estuaries 32, 447–455.

the Gulf of Mexico in July 2010, during the Deepwater Horizon oil spill. Cont. White, H.K., Hsing, P.-Y., Cho, W., Shank, T.M., Cordes, E.E., Quattrini, A.M., Nelson, R.

Shelf Res. 77, 118–131. K., Camilli, R., Demopoulos, A.W.J., German, C.R., Brooks, J.M., Roberts, H.H.,

Snyder, S.M., Pulster, E.L., Wetzel, D.L., Murawski, S.A., 2015. PAH Exposure in Gulf of Shedd, W., Reddy, C.M., Fisher, C.R., 2012. Impact of the Deepwater Horizon oil

Mexico demersal fishes, Post-Deepwater Horizon. Environ. Sci. Technol. 49, spill on a deep-water coral community in the Gulf of Mexico. Proc. Natl. Acad.

8786–8795. Sci. 109 (50), 20303–20308.

Socolofsky, S.A., Adams, E.E., Sherwood, C.R., 2011. Formation dynamics of Wilson, R.M., Cherrier, J., Sarkodee-Adoo, J., Bosman, S., Mickle, A., Chanton, J.P.,

subsurface hydrocarbon intrusions following the Deepwater Horizon blowout. 2015. Tracing the intrusion of fossil carbon into coastal Louisiana macrofauna

14 13

Geophys. Res. Lett. 38, L09602. doi:http://dx.doi.org/10.1029/2011GL047174. using natural C and C abundances. Deep-Sea Res. doi:http://dx.doi.org/

Socolofsky, S.A., Adams, E.E., Boufadel, M.C., Aman, Z.M., Johansen, Ø., Konkel, W.J., 10.1016/j.dsr2.2015.05.014.

Lindo, D., Madsen, M.N., North, E.W., Paris, C.B., Rasmussen, D., Reed, M., Yan, B., Passow, U., Chanton, J., Nöthig, E.-M., Asper, V., Sweet, J., Pitiranggon, M.,

Rønningen, P., Sim, L.H., Uhrenholdt, T., Anderson, K.G., Cooper, C., Nedwed, T.J., Diercks, A., Pak, D., 2016. Natural marine particles mediated sedimentation of

2015. Intercomparison of oil spill prediction models for accidental blowout. oil and drilling mud components from the DWH spill. Proc. Natl. Acad. Sci. (in

Mar. Pollut. Bull. 96, 110–126. review).

Soto, L.A., Botello, A.V., Licea-Durán, S., Lizárraga-Partida, M.L., Yáñez-Arancibia, A., Yang, T., Nigro, L.M., Gutierrez, T., D'Ambrosio, L., Joye, S.B., Highsmith, R., Teske, A.,

2014. The environmental legacy of the Ixtoc I oil spill in Campeche Sound, 2014. Pulsed blooms and persistent oil-degrading bacterial populations in the

southwestern Gulf of Mexico. Front. Mar. Sci. 1 doi:http://dx.doi.org/10.3389/ water column during and after the Deepwater Horizon blowout. Deep-Sea Res.

fmars.2014.00057. II doi:http://dx.doi.org/10.1016/j.dsr2.2014.01.014i.

Spier, C., Stringfellow, W.T., Hazen, T.C., Conrad, M., 2013. Distribution of Yvon-Lewis, S.A., Hu, L., Kessler, J., 2011. Methane flux to the atmosphere from the

hydrocarbons released during the 2010 MC252 oil spill in deep offshore waters. Deepwater Horizon oil disaster. Geophys. Res. Lett. 38, L01602. doi:http://dx.

Environ. Pollut. 173, 224–230. doi.org/10.1029/2010GL045928.

Stoffyn-Egli, P., Lee, K., 2002. Formation and characterization of oil-mineral Ziervogel, K., Arnosti, C., 2013. Enhanced protein and carbohydrate hydrolyses in

aggregates. Spill Sci. Technol. Bull. 8 (1), 31–44. plume-associated deepwaters initially sampled during the early stages of the

Størdal, I.F., Olsen, A.J., Jenssen, B.M., Netzer, R., Hansen, B.H., Altin, D., Brakstad, O. Deepwater Horizon oil spill. Deep-Sea Res. II doi:http://dx.doi.org/10.1016/j.

G., 2015. Concentrations of viable oil-degrading microorganisms are increased dsr2.2013.09.003i.

in feces from Calanus finmarchicus feeding in petroleum oil dispersions. Mar. Ziervogel, K., Joye, S.B., Arnosti, C., 2014. Microbial enzymatic activity and secondary

Pollut. Bull. 98, 69–77. production in sediments affected by the sedimentation event of oily-particulate

Sullivan, A.P., Kilpatrick, P.K., 2002. The effects of inorganic solid particles on water matter from the Deepwater Horizon oil spill. Deep-Sea Res. II doi:http://dx.doi.

and crude oil emulsion stability. Ind. Eng. Chem. Res. 41, 3389–3404. org/10.1016/j.dsr2.2014.04.003.

Teal, J.M., Howarth, R.W., 1984. Oil spill studies: a review of ecological effects. Ziervogel, K., McKay, L., Rhodes, B., Osburn, C.L., Dickson-Brown, J., Arnosti, C., Teske,

Environ. Manage. 8 (1), 27–44. A., 2012. Microbial activities and dissolved organic matter dynamics in oil-

Turner, J.T., 2002. Zooplankton fecal pellets, marine snow, and sinking contaminated surface seawater from the Deepwater Horizon oil spill site. PLoS

phytoplankton blooms. Aquat. Microb. Ecol. 27, 57–102. One 7 (4), e34816. doi:http://dx.doi.org/10.1371/journal.pone.0034816.

Valentine, D.L., Fisher, G.B., Bagby, S.C., Nelson, R.K., Reddy, C.M., Sylva, S.P., Woo, M. Ziervogel, K., Dike, C., Asper, V., Montoya, J., Battles, J., D’Souza, N., Passow, U.,

A., 2014. Fallout plume of submerged oil from Deepwater Horizon. Proc. Natl. Diercks, A., Esch, M., Dewald, C., Arnosti, C., 2015. Enhanced particle fluxes and

Acad. Sci. U. S. A. 111 (45), 15906–15911. doi:http://dx.doi.org/10.1073/ heterotrophic bacterial activities in Gulf of Mexico bottom waters following

pnas.1414873111. storm-induced sediment resuspension. Deep-Sea Res. II doi:http://dx.doi.org/ 10.1016/j.dsr2.2015.06.017.