1 s2.0 S0025322724000719 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Marine Geology 471 (2024) 107287

Contents lists available at ScienceDirect

Marine Geology
journal homepage: www.elsevier.com/locate/margo

Methane seeps on the U.S. Atlantic margin: An updated inventory and


interpretative framework
C.D. Ruppel a, *, A.D. Skarke b, N.C. Miller a, M.W. Kidiwela b, c, J. Kluesner d, W. Baldwin a
a
U.S. Geological Survey, Woods Hole, MA, USA
b
Department of Geosciences, Mississippi State University, MS, USA
c
Now at: School of Oceanography, University of Washington, Seattle, WA, USA
d
U.S. Geological Survey, Santa Cruz, CA, USA

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

Editor: Michele Rebesco Since the discovery of >570 methane flares on the northern U.S. Atlantic margin between Cape Hatteras and
Georges Bank in the last decade, the acquisition of thousands of kilometers of additional water column imaging
Keywords: data has provided greater coverage at water depths between the outer continental shelf and the lower continental
Cold seep slope. The additional high-resolution data reveal >1400 gas flares, but the removal of probable duplicates from
Northwest Atlantic Ocean
the combined database of new flares and those recognized in 2014 yields ~1139 unique sites. Most of these sites
Gas hydrate
occur in clusters of 5 or more seeps, leaving about 275 unique locations (including 47 clusters) for seepage along
Fluid migration
Methane the margin. As a function of depth, seep distribution is heavily skewed toward the upper continental slope at
water depths shallower than 400 m on the southern New England margin and ~ 550 m in the Mid-Atlantic Bight,
with additional seeps clustered at ~1100 m and just deeper than ~1400 m in both sectors. Despite little ongoing
tectonic deformation or active faulting on this passive margin, a variety of processes driven from below the
seafloor (e.g., migration of fluids along faults or through permeable strata, seepage above diapirs or other pre-
existing structures) and from above (e.g., erosion, sapping, unroofing) contribute to the development of seeps in
different settings along the margin. In addition, the prevalence of seeps on promontories overlooking shelf-
breaking canyons may be directly related to the three-dimensional nature of the hydrate stability zone in
these locations. As a function of depth, the parts of the slope at the contemporary landward limit of gas hydrate
stability are devoid of seeps, and the upper slope zones with the most concentrated seepage were not within the
gas hydrate stability zone even during the Last Glacial Maximum. Thus, if the large number of upper slope seeps
is at least partially sourced in gas hydrate degradation, the gas emitted at these seeps must have migrated there
from greater depths on the continental slope.

1. Introduction seep clustering relative to seafloor features, and explored possible con­
trols on seep location and evolution. In the intervening decade,
A decade ago, Skarke et al. (2014) published an analysis and asso­ numerous studies have built on the USAM seeps discovery to investigate
ciated database for >570 newly-discovered water column methane changing ocean chemistry and possible sea-air methane exchange
plumes along a ~ 900 km section of the northern U.S. Atlantic margin (Weinstein et al., 2016; Ruppel and Kessler, 2017; Garcia-Tigreros et al.,
(USAM) between Cape Hatteras and Georges Bank. Prior to that study, 2021; Joung et al., 2022); the fate of emitted methane in the water
fewer than five cold seeps had been explicitly mapped on the entire column (Fu et al., 2021); the seafloor processes associated with methane
margin, with most of these located above salt diapirs in the South generation, migration, and seepage (Pohlman et al., 2017; Ruppel and
Atlantic Bight (Van Dover et al., 2003; L. Brothers et al., 2013b). Skarke Waite, 2020; Ruppel et al., 2022); the characteristics of seep benthic
et al. (2014) demonstrated that the newly-mapped water column communities (Quattrini et al., 2015; Meyer et al., 2016; Bourque et al.,
anomalies could be traced to seafloor gas seeps, explored the reasons 2017; McVeigh et al., 2018; Portanier et al., 2023; Turner et al., 2020);
that emissions had to be dominated by methane, described patterns of and the strength and nature of the water column and seafloor oxidative

* Corresponding author at: U.S. Geological Survey, 384 Woods Hole Rd, Woods Hole, MA 02543, USA.
E-mail address: cruppel@usgs.gov (C.D. Ruppel).

https://doi.org/10.1016/j.margeo.2024.107287
Received 26 November 2023; Received in revised form 1 April 2024; Accepted 8 April 2024
Available online 9 April 2024
0025-3227/Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

sinks (e.g., Prouty et al., 2016, 2020; Leonte et al., 2020). interpretation of cold seep distribution on the USAM. We explore the
The USAM seeps were discovered as researchers were beginning to processes contributing to the evolution and persistence of seeps in USAM
routinely record water column imaging data acquired during multibeam settings characterized by different water depths, geologic features, and
echosounder (MBES) bathymetric surveys. As it became possible to re­ sedimentary/erosional regimes. We also explicitly address the role of gas
cord and store the very large water column data (WCD) digital files, hydrate dynamics in contributing to seep evolution and continued
researchers began conducting more frequent post-cruise analyses to emissions, a topic that remains controversial within the seep commu­
locate gas bubble plumes in WCD imagery and link the plumes to sea­ nity. Throughout this paper, we remain consistent with the existing
floor seeps. The seeps database published with Skarke et al. (2014) scientific literature and refer to seep locations (Table 1) using the names
relied exclusively on multibeam WCD acquired by the National Ocean documented in Skarke et al. (2014) or assigned by NOAA OER and
and Atmospheric Administration’s (NOAA’s) Office of Ocean Explora­ related programs during remotely-operated vehicle (ROV) dives.
tion and Research (OER) during surveys with the Okeanos Explorer be­
tween 2011 and 2013. Since that time, many additional research cruises 2. Geologic setting
have been conducted on the USAM. Surveys by NOAA OER, the U.S.
Geological Survey (USGS), and others have acquired thousands of ki­ Seep evolution on the USAM outer shelf and continental slope has
lometers of data to facilitate robust identification of water column gas been influenced by older tectonic features overprinted by patterns of
plumes (e.g., Baldwin et al., 2020b; L. Brothers et al., 2013b; Ruppel more recent deposition, erosion, and mass wasting, especially since the
et al., 2018). As documented in this paper (Fig. 1) and the associated end of the Last Glacial Maximum (LGM; ~20 ka). The USAM is a passive
seeps database release (Skarke et al., 2024), these surveys have identi­ margin formed by the rifting of northwestern Africa from North America
fied hundreds of new water column plumes (Skarke et al., 2024) that starting in the Late Jurassic (Withjack et al., 2020). Onshore, the rifting
significantly expand the original database published by Skarke et al. event produced discrete basins that stretch the entire length of the
(2014). USAM beneath the contemporary coastal plain. The two primary
This paper provides an updated analysis and the first in-depth offshore basins are the Baltimore Canyon Trough, which lies mostly

Fig. 1. The northern U.S. Atlantic margin with the updated seeps database (duplicates removed; Skarke et al., 2024) shown as red circles. The area surveyed for
Skarke et al. (2014) is shown in gray, and the light blue outline encloses the tracklines for the new surveys used to support this study. The white lines labelled MX
denote seismic data collected on the 2018 MATRIX cruise (Majumdar et al., 2022; Baldwin et al., 2020a). Yellow oval on the New Jersey shelf is Baltimore Canyon
Trough (BCT) salt from McKinney et al. (2005). The nearby yellow cross is the approximate location of a BCT-associated salt diapir identified by Grow (1980).
Bathymetry at depths beyond the shelf-break is from Andrews et al. (2016). (For interpretation of the references to colour in this figure legend, the reader is referred
to the web version of this article.)

2
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Table 1 Table 1 (continued )


Key Seep Fields on the USAM. Seep Field Location Characteristics Selected
Seep Field Location Characteristics Selected (Common Name) description1 References
(Common Name) description1 References (Parentheses)
(Parentheses) and Selected
and Selected Dives3
Dives3
BODIE ISLAND (35.735, − 74.817), Bubble streams, (Ruppel et al.,
(Paull et al., ~400 mbsl MDAC, 2019)
1996; Van Upper slope on chemosynthetic EX1903L2 Dive
Dover et al., south side of large mussels, bacterial 14
2003; L. slide north of Keller mats, deep-sea
Brothers et al., Canyon corals
Seeps fed by fluids 2013; (Cantwell,
migrating above Hornbach et al., 2019)
(32.4940, salt diapir; gas 2007; Bourque No dives
(35.8160,
− 76.1908), 2160 hydrate, MDAC2, et al., 2017; Bubble stream Imaged on
− 74.6109), ~1342
mbsl chemosynthetic Cordes, 2019; from EX1206 but not
mbsl
Inner Blake Ridge mussels and Portanier et al., hydroacoustics included in
BLAKE RIDGE Isolated deep seep
over the Blake clams; deep-sea 2023) KELLER DEEP recorded on Skarke et al.
in area unaffected
Ridge diapir; coral; filamentous Alvin Dives: SEEP EX1206 (but only (2014)
by slides between
seafloor within mat; drilled as Site 3710, 3711, (Fig. 2c) discovered during Re-imaged with
Bodie and Kitty
GHSZ 996 of the Ocean 3712, 3909, data analysis in WCD on R/V
Hawk seeps
Drilling Program 3910, 3912, 2018 by Skarke Marcus
Seafloor within
Leg 164 (Diapir) 4798, 4799, et al. (2018)) Langseth
GHSZ
4800 MGL1408, Line
Deep Search 38 on October
2019 ROV 7, 2014.
Jason dive (35.932–74.816),
Vestimentiferan (Cordes, 2019)
J2–1136 ~285 mbsl
worms, bubbles, Dives: Deep
(L. Brothers KITTY HAWK Upper slope on
quill worms, Search 2019
et al., 2013; (Fig. 3a) north side of large
bacterial mats, no ROV Jason dive
Bourque et al., slide north of Keller
mussels J2–1134
2017; Cordes, Canyon
2019) (Skarke et al.,
(32.975, − 75.933), Alvin 3911, 2014, 2018,
~2600 mbsl Bubble streams, 3913, 3914, 2019; Bourque
Above nearly MDAC, bacterial 4801 et al., 2017;
CAPE FEAR
breached salt diapir mats, deep sea NOAA EX 1205 (36.869, − 74.483), Dempoulos
Bubble streams,
Seafloor within corals (Diapir) L1: AUV Sentry ~1450–1620 mbsl et al., 2019;
MDAC,
GHSZ Dives 146, 149, Thalweg of Ruppel and
chemosynthetic
151, 152 NORFOLK unnamed canyon, Waite, 2020;
mussels, bacterial
Deep Search (Figs. 3b, 6, & 9) 18 km south of Portanier et al.,
mats, brine flows
2019 ROV Norfolk Canyon 2023)
(?), mud seeps
Jason dive Seafloor within NOAA EX1302
(Faults, erosion)
J2–1137 GHSZ Dive 6,
No dives (too EX1903L2 Dive
shallow for 19
many ROVs). Alvin Dives:
Discovered on 4802, 4803
(35.5474, Seeping mound in
R/V Armstrong West: (37.5197,
− 74.8384), ~65 layered sediments
SVC2 in 2016 − 74.155) 940 mbsl
SUPERSEEP mbsl on continental
Surveyed with East: (37.5423, Hydrate at
Near head of Keller shelf; vigorous
high-resolution − 74.102) 1035 seafloor (east), (Turner et al.,
Canyon seepage imaged
Kongsberg mbsl extensive mussel 2020; Ruppel
SBP29 on R/V CHINCOTEAGUE Near crest of ~15- beds and MDAC, et al., 2022;
Armstrong in RIDGE km-long ridge bacterial mats, Portanier et al.,
2024 (Fig. 3b) striking NE-SW on effusive gas 2023)
No dives (too mid-slope offshore emissions, quill Alvin Dives:
shallow for Virginia barrier worms 4804, 4806
many ROVs). islands (Faulting?)
(35.563, Discovered on Seafloor within
− 74.8234), 65 Seeping mound in R/V Armstrong GHSZ
mbsl layered sediments SVC2 in April (Skarke et al.,
KELLER MOUND
~2 km north of the on continental 2016 Surveyed (37.5678, 2014, 2018;
(Figs. 3a & 7)
superseep, near shelf; vigorous with high- − 74.2821) Meyer et al.,
head of Keller seepage imaged resolution ~425 mbsl Cap carbonate, 2016; Portanier
Canyon Kongsberg PICK UP STICKS (reported as 450 in bubble streams, et al., 2023)
SBP29 on R/V Skarke et al. (2014) quill worms NOAA EX1302,
Armstrong in Upper slope Dive 10
2024 offshore Virginia Alvin Dive:
(35.7069,- (Cordes, 2019) 4805
74.8122), Vestimentiferan Alvin Dive MDAC, bubble (Prouty et al.,
~285–350 mbsl worms, bubbles, 4961 (38.0495, streams, deep sea 2016; Bourque
PEA ISLAND
Upper slope on deep-sea corals, Deep Search BALTIMORE − 73.822), ~400 m corals, mussels, et al., 2017;
(Fig. 3a)
south side of large MDAC, bacterial 2019 ROV CANYON Overlooking bacterial mats Demopoulos
slide north of Keller mats, no mussels Jason dive (Figs. 3b & 10) Baltimore Canyon (Canyon et al., 2019;
Canyon J2–1133 on the south side Promontory; Portanier et al.,
Unroofing) 2023)
(continued on next page)

3
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Table 1 (continued ) beneath the New Jersey shelf and upper continental slope, and the
Seep Field Location Characteristics Selected Carolina Trough, which is near the base of the North and South Carolina
(Common Name) description1 References upper continental slope. Both basins contain over 10 km of sediment and
(Parentheses) are associated with salt diapirs (Carolina Trough; Dillon et al., 1982) or
and Selected diapirs and sheets (Baltimore Canyon Trough; Grow et al., 1988;
Dives3
McKinney et al., 2005; Grow, 1980), as shown in Figs. 2 and 1,
Alvin Dives: respectively. Like the onshore rift basins, the offshore ones have little
4807, 4808
evidence for oil or thermogenic gases (BOEM, 2021). Trace amounts of
Bubble streams (Skarke et al.,
(39.542, − 72.404), (sonar determined 2014, 2018; thermogenic material were detected a few hundreds of meters below
HUDSON mostly 500–570 – no visual Weinstein et al., seafloor on the Blake Ridge during Ocean Drilling Program (ODP) Leg
CANYON mbsl observations) 2016) 164 (Wehner et al., 2000), but nearly all gas emissions at USAM seeps
(Fig. 3c) Thalweg of Hudson (Hydrate Alvin Dives: are interpreted to be primarily methane of shallow (microbial) origin (e.
Canyon dissociating in 3706, 3707,
thalweg?) 3708, 4809
g., Paull et al., 1995; Skarke et al., 2014; Prouty et al., 2016; Pohlman
(Skarke et al., et al., 2017).
2014, 2018; Because this paper focuses primarily on contemporary cold seep
MDAC, bubble
Quattrini et al., features on the USAM outer shelf and upper continental slope north of
(39.8100, 2015; McVeigh
streams, deep sea Cape Hatteras, we highlight the post-rift geologic history related to
− 69.5907), et al., 2018)
1417–1608 mbsl
corals, bacterial
NOAA EX1302 connections among sedimentation, erosion, the most recent glaciation,
VEATCH CANYON mats, and ocean currents. Sediments have been delivered to the western North
Ridge on east side Dive 14, EX
(Figs. 3c & 11) chemosynthetic
of Veatch Canyon
mussels
1304 L1 Dive Atlantic basin by large, long-lived drainages that are ancestors of the
Seafloor within 13 present-day Hudson River and the Chesapeake (Susquehanna and James
(Faulting, Updip
GHSZ Alvin Dives:
gas migration?)
4810, 4813,
Rivers) and Delaware estuaries, along with smaller rivers from southern
4827, 4828, New England to South Carolina (e.g., D. Brothers et al., 2013a). Some
4835 fine sediment that ends up far offshore has been winnowed by the Gulf
(39.8697, Stream and the Deep Western Boundary Current into large contourite
NEW ENGLAND − 69.2862), 1452 (Ruppel and
deposits like the Chesapeake and Hatteras drifts, which share some
SEEPS #1 mbsl Demopoulos,
Drainages and
MDAC,
2013; Skarke characteristics with the Blake Ridge sediment drift (e.g., McCave and
chemosynthetic Tucholke, 1986; Rebesco et al., 2014).
Referred to as ridges on mid-slope et al., 2014,
mussels, bubble
New England between Veatch
streams, bacterial
2018; Quattrini On the continental shelf and uppermost continental slope, sea level
seep #3 by ( and Shallop et al., 2015) changes partially control where sediment has been deposited. During
mats, hydrate
Quattrini et al., Canyons EX1304L1 Dive
the most recent sea level lowstand at the LGM, Pleistocene shelf-edge
2015) Seafloor within 4
GHSZ deltas formed along much of the margin (Poag and Sevon, 1989; Twi­
(Ruppel and chell et al., 2009). These studies differ in the inferred extent of the
Demopoulos, deltaic deposits, and we here use the analysis of Twichell et al. (2009).
(39.9040, 2013; Skarke
Based on that study, shelf-edge deltas currently lie landward of the
− 69.2552), ~1108 MDAC, et al., 2014,
mbsl chemosynthetic 2018; Quattrini contemporary shelf-break along nearly the entire southern New England
Drainages and mussels, bubble et al., 2015; margin, across Hudson Canyon, between Wilmington and Washington
NEW ENGLAND
SEEPS #2
ridges on mid-slope streams, bacterial McVeigh et al., Canyons, and across the head of the Currituck slide (Twichell et al.,
between Veatch mats, hydrate 2018; Fu et al., 2009), as shown in Fig. 3.
(Fig. 3c)
and Shallop (Pre-existing 2021)
The thick sedimentary sequences deposited near the shelf-break
Canyons geologic EX1304L1 Dive
Seafloor within structure?) 3 north of Cape Hatteras have been prone to slumping, mass wasting,
GHSZ Alvin Dives: and submarine slope failure since at least Miocene time (e.g., Schlee
4815, 4833, et al., 1979; Poag, 1985; Chaytor et al., 2007). The Gulf Stream has not
4834.
usually directly affected the continental slope in this area, and most
West:
(39.9968,
upper slope deformation outside large canyons appears driven by pas­
(Skarke et al., sive collapse of sediments. Slides of presumed Pleistocene or Holocene
− 69.1925),
MDAC, 2014; McVeigh
327–390 mbsl
chemosynthetic et al., 2018;
age have occurred along the upper slope of the entire study area (Twi­
East: chell et al., 2009), although there are key differences along strike. For
mussels, bubble Portanier et al.,
SHALLOP (39.9975,
streams, bacterial 2023) example, the southern New England margin has experienced slope
CANYON − 69.1248), ~320
mbsl
mats Alvin Dives: failures or been highly eroded along most of its length, possibly due to its
(Canyon 4811, 4812,
Overlooking proximity to the Laurentide Ice Sheet termination (Knott and Hoskins,
Promontory) 4814, 4829,
Shallop Canyon,
4830, 4832 1968; Uchupi et al., 2001). The upper slope section from Hudson Canyon
southern New
to Block Canyon is relatively intact (e.g., D. Brothers et al., 2013a)
England margin
compared to the sector from Wilmington Canyon to Cape Hatteras,
1
Depths are nominal. Dives at the same location often report different depth which is the USAM segment that hosts the major shelf-break canyons
ranges, especially when seep fields are on the upper slope and may stretch over (Norfolk, Washington, Baltimore, and Wilmington; Figs. 1–3). Between
tens of meters. Baltimore Canyon and Cape Hatteras, much of the upper slope has been
2
MDAC denotes methane-derived authigenic carbonates.
3 heavily influenced by mass transport processes that have unroofed older
Dives with publicly available data or descriptions are listed. Alvin dive in­
sediments near the shelf-break. Taken together, the older tectonic his­
formation and data are available at https://ndsf.whoi.edu/alvin/dive-log/. “EX”
refers to dives on NOAA Ocean Exploration and Research (OER) cruises. The
tory combined with the overprint of younger deposition, erosion, and
cruises are named with the convention of EXYYNNL#, where YY denotes year, sediment transport contribute to the complexity of this passive margin
NN is expedition number in that area in that year, and # is the leg number. OER and its fluid migration processes.
dives were conducted with the ROV Deep Discoverer. Video for EX dives can be
accessed at: https://oceanexplorer.noaa.gov/data/access/access.html. Deep 3. Data and methods
Search 2019 dive information is available in Cordes (2019).
This paper explores the processes responsible for USAM methane

4
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Fig. 2. Different versions of the USAM seeps database. Green, blue, and purple shading correspond to submarine slide deposits of Pleistocene, older, and unknown
age, respectively, from Twichell et al. (2009). (A) Red circles show the Skarke et al. (2014) seeps database. Yellow crosses are salt diapirs from Dillon et al. (1982).
(B) The total database, including both the Skarke et al. (2014) seeps from (A) plus all new identifications with duplicates removed (Skarke et al., 2024), yields the
seeps shown as red circles. The boxes outline areas corresponding to Fig. 3. Yellow shading on the MAB shelf is salt in the Baltimore Canyon Trough as mapped by
McKinney et al. (2004). (C) Unique seep locations from the total database (Skarke et al., 2024). Red circles denote cluster centers, and yellow circles show other
individual locations. At a larger scale, the unique locations would appear to be evenly spaced within seep fields as an outcome of the clustering algorithm. Named
seeps (Table 1) for Blake Ridge, Cape Fear diapir, and Keller Deep are shown on this map, as is the Currituck slide seep discussed in the text. Updated seeps database
available in Skarke et al. (2024). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

5
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Fig. 3. Close-ups of areas bounded by boxes in Fig. 2b, showing seeps as white circles. The major shelf-breaking canyons are labelled. Pleistocene shelf edge deposits
are taken from Twichell et al. (2009), which is also the source of the slide deposits of Pleistocene (green shading), pre-Pleistocene (blue shading), and undetermined
age (purple shading). Boxes surround named seep fields described in more detail in Table 1. The location of the mega-pockmarks discussed in the text are shown in
(B), although seepage has not been detected here. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

seepage by analyzing water column imaging data that constrain the 3.1. Water column imaging data
locations of seafloor seeps and by integrating several types of seismic
surveys that provide information about geologic structures and fluid MBES water column imaging data were used to infer new seafloor
migration beneath the seafloor. seep locations for this study. WCD detect streams of gas bubbles
emanating from seafloor seeps in real-time, making it possible to

6
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

provisionally identify seeps within minutes while the ship is underway. 3.3. Methods
The physical principle underpinning the detections is the dramatically
lower compressional wavespeed for gas relative to seawater. Acoustic The procedure for identifying seeps in MBES data was described for
signals within an appropriate frequency range (usually 10–80 kHz at the original database by Skarke et al. (2014). For this study, most plumes
continental slope depths) travel through the water column and were identified by Mississippi State University researchers using EM302
encounter the impedance contrast across the interface between ocean and EM122 data analyzed in QPS FMMidwater software. The USGS
water and the interior of a gas bubble, creating a strong reflection that identified flares from R/V Armstrong and R/V Sharp cruises using Caris
reveals water column bubble streams, but not individual bubbles. These and QPS FMMidwater software on Reson 7160, EM122, or EM710 data.
flares are then traced downward and ascribed to emissions at discrete In all cases, individual plumes were identified, traced to their seafloor
seafloor seeps. location, and assigned geographic coordinates (Skarke et al., 2024).
MBES data image the water column and seafloor by transmitting a Plume identifications were assigned a subjective quality factor
pulse that ensonifies a fan-shaped volume below the vessel. For the ranging from 1 (best) to 5. Plume images rated with the highest quality
MBES systems used here, the fan is wide (up to 140◦ ) in the athwartships factor are connected to the seafloor, continue with a plausible
direction and narrow (few degrees) along-track. The reflected energy is morphology (e.g., as a vertical flare unless deflected by the current) into
recorded by narrow (0.5◦ - 4◦ ) beams, resolving seafloor features with the water column, and became weaker with height above the seafloor.
high accuracy (10–40 m; Kongsberg Maritime, 2012). The waveform Some low quality plumes, which include features with low acoustic
data recorded from each beam can be used to identify water column reflection intensity, lack of connectivity to the seafloor, or returns that
flares. are stronger in the water column than close to the seafloor, could be
This study relies mostly on 30 kHz EM302 MBES data collected by schools of fish or other biological phenomena. An additional annotation
the Okeanos Explorer from 2013 to 2014 (NOAA Office of Ocean was added when seeps appeared to be diffuse. Diffuse plumes often have
Exploration and Research, 2013a, b, 2014a, b, c, d), supplemented by low quality factors, and the imagery typically shows short plumes being
12 kHz EM122 data acquired on the R/V Atlantis in 2015 and 2016 and emitted along several tens of meters of continuous seafloor.
EM302 data from the R/V Sikuliaq in 2014. The USGS identified addi­
tional flares using WCD data from the EM122 and EM710 (70 kHz) 4. Results
MBES aboard the R/V Armstrong in 2016 and a portable 44 kHz Reson
7160 MBES pole-mounted aboard the R/V Hugh Sharp during two cruises Skarke et al. (2014) underscore that, for a small subset of the bubble
in 2017 (Baldwin et al., 2020b). Data from the R/V Sikuliaq, R/V Atlantis, plumes investigated by a remotely-operated vehicle (ROV) for the initial
and R/V Armstrong are available from the R2R repository.1 Other USAM USAM seeps database, each bubble stream could be traced to a seafloor
surveys also collected WCD data with MBES systems, particularly seep. In recent years, more ROV and submersible dives at sites identified
several Okeanos Explorer cruises after 2014. However, the geographic by Skarke et al. (2014) have visually confirmed bubble streams emitted
footprints of these later cruises do not substantially expand coverage. from the seafloor at additional sites (e.g., Table 1), and gas recovered
Thus, this study relies on surveys that enhanced coverage compared to from sediments near some seeps has been inferred to be microbial
the original study (Skarke et al., 2014) or that provide more detail in methane (e.g., Pohlman et al., 2017). Exploration of the seafloor at the
specific locations (Fig. 1). The combined surveys used for this study base of water column bubble plumes sometimes fails to discover an
cover an estimated ~77% of the area between the 100 m and 2000 m active seafloor seep. This can be attributed to ephemerality of emissions,
isobaths from Cape Hatteras to the New England seamounts. Of the uncertainty in the seep’s location, or difficulty in finding a discrete hole.
unsurveyed area at those depth ranges, more than half is on the outer Nonetheless, throughout this paper, we assume that identified water
Hudson and New England shelves between 100 m below sea level (mbsl) column plumes are indicative of seafloor seepage of methane during the
and the shelf-break. discrete time interval that the ship acquired the imagery data.
Fig. 2b and c compare the Skarke et al. (2014) seep map (Fig. 2a) to
3.2. Seismic data updated maps that include those seeps plus post-2014 identifications
from the MBES datasets. The original (Skarke et al., 2014) database
The seismic imagery used here includes legacy airgun data released documented 577 identifications of possible seeps. The new database lists
by the Bureau of Ocean Energy Management through the National nearly 1475 additional MBES identifications. However, it is important to
Archive of Marine Seismic Surveys (NAMSS) portal (Triezenberg et al., understand what these seep locations represent.
2016), data from an intermediate-energy (420 in3) airgun survey con­
ducted by the USGS in 2018 (MATRIX; Baldwin et al., 2020a; Majumdar 1. Some identifications are wholly new locations in areas not surveyed
et al., 2022), and data acquired by the USGS using a 2.4 kJ sparker for the Skarke et al. (2014) study. When we use “new seeps” to refer
source in 2015 (Ruppel et al., 2015; Ruppel et al., 2024) and 2016 to those identified since Skarke et al. (2014), the terminology does
(Baldwin and Miller, 2023). These datasets were respectively acquired not imply that the seeps have only recently started emitting gas or
with 4 ms, 1 ms, and 0.5 ms sampling intervals. The legacy airgun data that they have formed in the intervening years.
provide penetration sufficient to image pre-rift basement, while the 2. Some identifications are probably the same seeps as reported in the
sparker data yield outstanding resolution to several hundred meters original database or the same locations reported multiple times from
below seafloor. The full processing sequence for the BOEM data is un­ different cruises in the new database. We explore this quantitatively
known, but the data were downloaded from the NAMSS portal in below as we cull duplicates from the database.
migrated form. MATRIX data processing included normal moveout 3. Some identifications add new sites to existing seep fields where the
correction, stacking, and poststack phase-shift time migration, as well as previous database had already identified multiple seeps. Seep clus­
removal of reverberation for shallow-water lines (Baldwin et al., 2020a). ters are also analyzed further below.
The sparker data were bandpass filtered, normal moveout corrected,
stacked, migrated, and muted above the seafloor reflection (Baldwin and To determine if some plume identifications reported in the combined
Miller, 2023; Ruppel et al., 2024). MBES database (Skarke et al., 2024) could be the same physical seeps,
we ran the QGIS (QGIS.org, 2023) density-based spatial clustering with
noise algorithm on the combined database projected as Universal
Transverse Mercator (UTM) and sought clusters with a minimum of two
identifications within 40 m of each other. The 40 m criterion represents
1
https://www.rvdata.us a reasonable degree of uncertainty in the locations of bubble plumes

7
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

identified in different surveys (e.g., different MBES, different ships, and/ 4.1. Seep depth distribution
or different times). In the field, ROV dives often observe individual
seafloor gas emission points separated by far <40 m, but there is no In Skarke et al. (2014), the analysis of seep depth distributions relied
agreement about whether each hole should be labelled a seep or whether on all locations (including those we now label as duplicates) that had
the seep should constitute all emission loci in a small area. Furthermore, been identified within three depth ranges: Shallower than 180 mbsl,
bubble plumes from individual emission points coalesce in the water 180–600 mbsl, and >600 mbsl. The bins were respectively designed to
column and would be imaged as a single plume at the resolution of WCD. correspond to the continental shelf and very uppermost continental
With these caveats, we found 356 locations in which at least two iden­ slope (up to 180 mbsl); from this depth to approximately the shallowest
tifications were within 40 m of each other for the combined 2014 and limit of gas hydrate stability based on known bottom water temperature
new MBES database (Skarke et al., 2024). These identifications were conditions (550–575 mbsl; e.g., Brothers et al., 2014); and deeper than
considered duplicates, and only one identification was retained at each the minimum depth for hydrate stability. Revising this analysis with the
site. Because several duplicates consisted of more than two identifica­ same depth ranges and new ways of categorizing the gas flare identifi­
tions, the total of 2052 identifications from the combined 2014 and new cations yields the results shown in Table 2.
database was reduced to 1139 unique sites. Note that, when plotted, When duplicates are removed from the new total MBES database
these unique sites can appear to be spaced on a grid within some seep (Skarke et al., 2024) and each remaining identification considered a
fields owing to the adoption of the 40 m criterion for identifying du­ unique seep (Fig. 1b), the upper slope depth range from 180 to 600 mbsl
plicates. For shipboard programs relying on locating seeps, the full (raw) still dominates, but less than in Skarke et al. (2014). Many more iden­
database may therefore be more useful. tifications have been made in deepwater (> 600 mbsl) in the intervening
The next step in analyzing the combined MBES dataset was identi­ years, while the percentage on the outer shelf and uppermost slope (<
fying seep clusters, defined as groups of seeps that are close enough 180 mbsl) has remained nearly constant. When identifications close to
together to meet certain quantitative criteria. We applied the QGIS one another (400 m distance) are collapsed into clusters (Fig. 1c), a
spatial density clustering algorithm to the 1139 non-duplicate locations larger percentage of the seeps is at outer shelf and uppermost slope (<
to identify clusters of seeps within 400 m of each other. This produced 180 mbsl) water depths. As shown in Fig. 1, most of the USAM conti­
47 seep clusters containing at least 5 and as many as 138 individual nental shelf has not been surveyed for seeps, and the percentages re­
identifications, plus 227 identifications that were not part of clusters. ported in Table 1 would be significantly different if more data were
The 400 m criterion was chosen to ensure that all identifications within available, especially for shelf depths <180 mbsl.
one areally-extensive seep field (e.g., Norfolk Seep field, described Figure 4 shows a histogram of the depth distribution of seep locations
below) were identified as the same cluster. With the addition of two for the entire USAM database with clusters and the remaining isolated
identifications that were within 400 m of each other but that did not identifications. When interpreted in this way, most isolated seeps or seep
meet the criteria for a cluster, the database is reduced to 275 unique fields lie at water depths shallower than 350 mbsl. The upper slope in­
locations, of which 47 are considered seep fields (clusters). The largest terval with the fewest seeps is between the contemporary landward limit
cluster is Hudson Canyon, containing 138 (non-duplicative) identifica­ of gas hydrate stability (LLGHS at ~550 mbsl; Brothers et al., 2014;
tions. The mean number of non-duplicative identifications in clusters is Ruppel and Waite, 2020) and 1000 mbsl. Most seep clusters are between
19, and the median is 10. Cluster results are shown in Fig. 2c and re­ 250 and 450 mbsl, although some clusters also occur deeper than 1000
ported in Skarke et al. (2024). mbsl.
The analysis of duplicates and clusters could be done using different The distribution of seeps as a function of depth is dominated by the
approaches, criteria, or algorithms, which would lead to different ab­ ~724 unique seep loci within the Mid-Atlantic Bight (MAB; Fig. 1),
solute numbers of non-duplicate seeps and of clusters plus remaining which we define as the USAM segment between Hudson Canyon and
unclustered seeps. The absolute numbers in each category are not a key Cape Hatteras. Of the three main sectors of the USAM (South Atlantic
focus of this paper, and the spatial and temporal variability in seep Bight or SAB, MAB, and southern New England) examined here, the
processes makes it impossible to count the specific number of seeps in MAB outer shelf and upper continental slope have the greatest spatial
any case. coverage for surveys (Fig. 1) and the most repeated surveys. Survey
Because flare identifications represent emissions at discrete times spatial coverage on the southern New England margin is nearly com­
corresponding to surveys during different cruises over a period of ~7 plete at upper continental slope depths (Fig. 1). Thus, key observations
years, this analysis only captures the time-transitive spatial component about seep distribution north of Cape Hatteras should be considered
of seepage. In other words, the analysis cannot account for temporal robust beyond the shelf-break.
variations, including seeps turning off or on at timescales ranging from As shown in Figs. 2c and 5a, the MAB sector not only hosts the most
seconds to days to years. While we recognize >1000 emission points seeps so far discovered, as well as the seeps that are most evenly
with our criteria, whether they are all active at the same time, whether distributed both along the margin and as a function of water depths.
seepage is shifting periodically among several emission points in the Fig. 5b shows the depths of all known MAB and southern New England
same cluster, and whether a single bubble plume represents multiple margin seeps plotted on composite (averaged) bathymetric profiles. At
emission loci remain impossible to constrain with our data. continental shelf depths, an arbitrary slope up to 50 mbsl was used over
With the refined database, we can examine the overall pattern of 5 km distance to mimic the bathymetric profile, and the southern New
seepage as a function of depth, location, and proximity to certain sea­ England margin bathymetric profile is shifted landward to avoid overlap
floor and subseafloor features. The next section first describes the dis­ with the MAB data. The MAB composite diagram illustrates the
tribution of seeps as a function of depth on the margin and then
progresses south to north along the USAM to highlight how the new
database changes our understanding of the large-scale distribution of Table 2
seeps. We then focus on new identifications in a few significant seep Seep Statistics.
fields and describe several unique locations, using the names commonly Original (Skarke Total database Total database
used by the scientific community. The names, locations, and charac­ et al., 2014) without duplicates with clustering
teristics of key seep fields are summarized in Table 1 and illustrated in <180 mbsl 15.8% 13.5% 30.7%
Fig. 3. Between 180 76.2% 63.9% 55.4%
and 600 mbsl
>600 mbsl 8.0% 22.6% 13.9%
number 577 1139 274

8
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Fig. 4. Histogram showing distribution of unique seep locations as a function of depth for the entire USAM, including the South Atlantic Bight. Blue bars are in­
dividual seeps, and brown bars denote the depths of the centers of seep clusters. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

concentration of seeps at water depths shallower than ~550 mbsl, the 4.2. Regional distribution
nominal LLGHS. On the southern New England margin, seepage is also
primarily an upper slope phenomenon, although the cutoff is shallower The new database (Skarke et al., 2024) significantly expands the
at ~400 mbsl. For the MAB and the southern New England margin, most number of seeps recognized on the continental shelf and uppermost
seeps that lie deeper than the upper slope exist in one of two clusters at continental slope immediately north of Cape Hatteras. This area was
seep locations at ~1000–1100 mbsl (includes Chincoteague and New poorly represented in Skarke et al. (2014) because the EM302 MBES on
England seeps #2, respectively) and at ~1400–1600 mbsl (includes the Okeanos Explorer is primarily operated in deeper waters. Some seeps
Norfolk Seeps and Veatch and New England seeps #1, respectively). mapped near Cape Hatteras since 2014 are at locations and at water
The MAB and southern New England margin correspond to Groups 4 depths landward of the shelf break (nominally 125 m), with some of the
and 1, respectively, of the designations used by D. Brothers et al. (2013a) densest clusters and most spectacular seeps at water depths of 60–75 m.
in describing the geomorphic controls on the USAM continental slope. Based on physical considerations, a fraction of the methane emitted
Fig. 5c shows the relationship of the seep distributions in these two areas from seeps shallower than ~100 mbsl is likely to reach the sea-air
superposed on an interpretation of the gross-scale subsurface structure interface if not consumed aerobically in the water column (Ruppel
in each sector. On the New England margin, the seep cluster just deeper and Kessler, 2017). Thus, continental shelf seeps probably have outsized
than 1000 mbsl roughly lines up with the presence of a buried reef significance for seep methane emissions that reach the atmosphere.
structure, which leads to sedimentation truncation or bypass. The Skarke et al. (2014) noted parts of the MAB where seeps were notably
deeper seeps (> 1400 mbsl) are within the upper wedge of the onlapped lacking and explored possible explanations for this observation. For
sediment, where gas migrating updip might accumulate against the example, over 270 of the ~577 identifications in that study lay on the
buried reef structure. In the MAB, the reef structure occurs much farther upper continental slope between Norfolk and Wilmington Canyons, but
offshore, and the upper slope is instead dominated by the relic Miocene the stretch of slope between Wilmington and Hudson Canyons and from
sedimentary wedge beneath the zone of most active seepage. Hudson to Veatch Canyons appeared devoid of seeps. The new database
(Skarke et al., 2024) covers this area in greater detail and now reveals a
sparse seep distribution (fewer than 10) along the ~155 km segment

9
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Fig. 5. (A) Cumulative number of individual seep locations from deepwater to the shelf as a function of distance from the approximate 50 m isobath for seeps on the
Mid-Atlantic Bight (green) and Southern New England (orange). (B) Depth distribution of seeps in this sector accumulated along a composite bathymetric profile. (C)
Seeps from (B) distributed on scaled schematic diagrams of Brothers et al. (2013a) for the New England and Mid-Atlantic margins. (For interpretation of the ref­
erences to colour in this figure legend, the reader is referred to the web version of this article.)

between Wilmington and Hudson Canyons (Fig. 1c). Most of these seeps 4.3. New emission sites within previously identified seep fields
lie near the heads of small canyons and drainages, but two seeps in the
30-km-long section immediately south of Hudson Canyon are in an area With the availability of new WCD data, many more emission sites
where the upper slope is less disturbed (Fig. 2c), within the prograding have been identified near previously discovered seeps or within already-
section of the Hudson apron described by D. Brothers et al. (2013a). known seep fields. For example, on the outermost shelf and uppermost
Seeps have now also been found in a relatively uneroded section of the slope, the new database identifies many new seeps within previously-
outer shelf and upper continental slope between ~90 and 360 mbsl west mapped dense clusters between Washington and Baltimore Canyons
of Block Canyon, the first seeps discovered between Hudson and Veatch (Figs. 2c and 3b). In the thalweg of Hudson Canyon, where Skarke et al.
Canyons. (2014) describe several dozen seeps at ~525 mbsl, the total database
Since Skarke et al. (2014), numerous surveys have expanded the now contains 138 distinct locations, which form clusters near the LLGHS
dataset for the southern New England margin, with the area between and another grouping at even shallower (~265 mbsl) depth (Fig. 2c).
Veatch and Hydrographer Canyons (e.g., McVeigh et al., 2018; Baldwin The 2014 database also listed only a single seep within the 0.6 km2 area
and Miller, 2023) being especially targeted. Most newly-identified seeps that makes up what is now referred to as the Bodie Island seep field
are on the walls of smaller drainages on the continental slope, not within (Fig. 3a), where three dozen distinct emission loci have now been
the drainages for large canyons. Some of the newly-identified upper mapped.
continental slope seeps in this sector also lie at or near the crests of The addition of new data has also altered perspectives on some of the
ridges that divide drainages, similar to the setting for the Veatch Canyon rarer deepwater (>1000 m) seep fields along the Mid-Atlantic part of the
seeps. margin. For example, Skarke et al. (2014) discuss the Norfolk Seep Field,
which is located in an unnamed canyon ~18 km south of the thalweg of
Norfolk Canyon. Back then, the existing data indicated that ~18 active
seeps were arrayed in two lines intersecting at ~120◦ . The Okeanos

10
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Explorer MBES data analyzed for the new database document >80 seeps Keller Canyon and has informally been called the “superseep.” This
within an area of ~1.3 km2 (Figs. 3b and 6). Seafloor explorations by the feature is ~7 m high, 280 m in the across-shelf direction, and 300 m
NOAA ROV Deep Discoverer had found carbonate outcrops and large along the strike of the shelf-break, for a total area of ~48,200 m2. In
chemosynthetic communities prior to Skarke et al. (2014). A 2019 D2 profile, the superseep looks like two distinct mini-volcanoes, with a flare
dive in a newly discovered part of the seep field discovered similar emitted from one peak. The second feature has been referred to as the
features, as well as a suspected seafloor hydrate mound, small holes Keller Mound and is also ~7 m high. The mound has a lumpy surface and
where fluidized mud was being emitted from the seafloor, and apparent is ~390 m long subparallel to the head of Keller Canyon and has a cross-
brine flows (Skarke et al., 2019; Ruppel and Waite, 2020). This seep field shelf width of ~310 m for a surface area of ~79,200 m2. The top of the
has also been the target of additional studies focused on macrofauna, mound has multiple vents, and several more exist adjacent to the
chemosynthetic communities, biochemical and molecular analyses, and mound. Despite the many cruises that have surveyed comparable loca­
authigenic carbonates (Quattrini et al., 2015; Bourque et al., 2017; tions close to the shelf-break along the USAM, no other similar mounds
Coykendall et al., 2019; Prouty et al., 2020). have yet been discovered at these shallow water depths. There have also
Similarly, a small ridge at ~1100 mbsl offshore Chincoteague Island been no dives on these two features, but their morphology had previ­
had only a few seeps documented in Skarke et al. (2014). With new data ously led us to hypothesize that they might be composed of authigenic
from several USGS cruises and other surveys (e.g., Baldwin et al., 2020b; carbonate. High-resolution imaging obtained by the USGS in 2017 had
Turner et al., 2020), seeps have now been mapped at several places near revealed gas and gas chimneys underneath the entire area, but new data
the spine of the ~15-km-long ridge (Fig. 3b), including within a spec­ have recently shown that the superseep and Keller Mound consist of
tacular seep field on the southeast side (Ruppel et al., 2015; Ruppel layered sediments (Fig. 7b).
et al., 2022). The previous section noted additional seeps discovered within
Hudson Canyon. Of particular significance is the subset of thalweg seeps
4.4. Special locations at water depths of ~400–600 mbsl, which bracket the LLGHS (Skarke
et al., 2014; Weinstein et al., 2016). The updated database now contains
Within the cluster of seeps on the shelf offshore Cape Hatteras near at least 80 distinct seeps at this critical depth range. Most of the major
the head of Keller Canyon (Fig. 3a) are two anomalous features first shelf-breaking canyons within the MAB host seeps arrayed around
discovered in 2016. These seeps are associated with mounds that lie at canyon heads or on promontories overlooking the thalweg (see below)
~65–70 mbsl at locations ~2 km apart (Figs. 3a and 7a). The first at the same range of water depths as this subset of Hudson Canyon seeps.
seeping mound is perched at the edge of a steep slope falling away into Hudson Canyon differs starkly from these other locations in having such

Fig. 6. Norfolk Seeps. (A) Mapped seeps (white circles) draped on shaded seafloor relief showing the arrangement of the seeps down the erosional face and into the
thalweg of the canyon. The seeps are roughly arrayed along the seismic lines depicted on the fig. (B) Three-dimensional composite of MBES WCD data revealing
active seepage at the Norfolk Seeps field in July 2019, as described by Skarke et al. (2019) and imaged by R/V Okeanos Explorer on EX1903L2.

11
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Fig. 7. Superseep and Keller mound features on the continental shelf overlooking Keller Canyon (see Fig. 3a). (A) Large map shows high resolution bathymetry and
locations of non-duplicate seeps (red circles). The green circles mark the central vents for these two mounds. Left inset shows shaded relief map with same green
circles and a yellow trackline corresponding to the location of the profile in (B). Right inset shows a close-up of the bathymetry at Keller mound with the seep
locations plotted from the duplicates removed database, leading to the appearance of equally spaced seeps. (B) Subbottom image between the superseep and Keller
Mound from Kongsberg SBP29 operating on an EM124 multibeam footprint on the R/V Armstrong. Although flares are usually not visible in the water column on such
imagery, the flares are vigorous enough to be detected in this case. The inset shows the 38 kHz EK80 image along the same transect. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

a high concentration of seeps directly within the thalweg. Seep emis­ mbsl; Fig. 2c) is one of the few so far discovered on this margin within a
sions there also vary with time, with a USGS survey in late summer 2015 known large-scale slide deposit (Twichell et al., 2009; Chaytor et al.,
(Ruppel et al., 2015) failing to find active gas flaring from this seep field, 2007). This seep was assigned quality factor 3 when first identified in
which was active both before that survey (2014 surveys on the R/V 2012 MBES data from Okeanos Explorer, and re-examination of the
Endeavor and R/V Sikuliaq) and during a 2016 R/V Atlantis survey. original data has verified that a plume was present at this site. In 2017,
The outermost shelf south of Norfolk Canyon hosts depressions the USGS did several passes over the site and failed to detect a water
elongated parallel to the shelf break and has long been an area of special column plume (e.g., Baldwin et al., 2020b). Review of legacy seismic
interest to geoscientists. These depressions were originally described as data (Triezenberg et al., 2016) passing close to the site reveals a seafloor
en echelon normal faults by Driscoll et al. (2000), but subsequent sur­ feature with signs of fluid or gas migration below. The seep is likely
veys revealed no evidence for normal faulting. Instead, the depressions ephemeral, and its position within Currituck slide deposits may point to
were re-interpreted as mega-pockmarks (Fig. 3b) linked to gas escape a different origin than deepwater seeps associated with faulting, erosion,
(Hill et al., 2004). Newman et al. (2008) detected elevated water column or diapirism on other parts of the margin.
methane near the features, and Garcia-Tigreros et al. (2021) measured
methane concentrations of 25.7 nmol close to the seafloor at 144 mbsl, 5. Discussion
just deeper than the depressions. USGS-led surveys in 2016 and 2017
sought evidence for active seepage (e.g., Baldwin et al., 2020b) from 5.1. Seep evolution
these depressions. Those surveys identified no bubble flares, meaning
that the depressions are not likely to be loci for contemporary ebullitive The USAM differs in several ways from the Cascadia and northern
gas expulsion. The lack of detected flares does not preclude diffuse Gulf of Mexico margins, where hundreds to thousands of seeps have also
methane emission or expulsion of dissolved methane from the seafloor. been mapped (e.g., Brooks et al., 1984; Merle et al., 2021; Johnson et al.,
An isolated deepwater seep identified by Skarke et al. (2014) within 2015, 2019; Kennicutt, 2017; Riedel et al., 2018; Sassen et al., 2001). As
the upper part of the Currituck slide (36.3919◦ N, 74.6854◦ W; 1101 a conventional passive margin, the USAM is tectonically less complex

12
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

than the U.S. Cascadia margin (Rudebusch et al., 2023) and geologically Diapiric structures on the USAM were recognized in the South
less complex than the salt tectonic province of the northern Gulf of Atlantic Bight along a line adjacent to the buried Carolina Trough in the
Mexico (e.g., Roberts and Carney, 1997; Ventress et al., 1989). However, 1970s and 1980s (e.g., Dillon et al., 1982; Fig. 2b), and ODP Leg 164
the USAM seeps still occur in a diverse range of morphologic settings and confirmed the salt composition of these features (e.g., Egeberg, 2000).
are likely connected to various geologic and sedimentary/erosional As noted by Skarke et al. (2014), the Blake Ridge and Cape Fear seeps are
processes, as summarized in Fig. 8. the only ones on the USAM known to have developed over salt diapirs (L.
Skarke et al. (2014) briefly reviewed geologic explanations for USAM Brothers et al., 2013b; Hornbach et al., 2007; Van Dover et al., 2003).
seeps, including the roles of salt diapirs, canyon formation, and disso­ These seeps are the deepest presently known on the margin and have
ciation at the LLGHS. Here we provide detail on a wide range of pro­ long been linked to dissociation of gas hydrate in the sediments above
cesses and provide examples of bottom-up (from features beneath the the rising diapirs.
seafloor) and top-down (from processes associated with sedimentation, During this study, we reviewed additional water column imaging
erosion, or changes in currents) seep initiation. data collected over the other Carolina Trough salt diapirs mapped by
Dillon et al. (1982) and archived in the R2R repository (rvdata.us). We
5.1.1. Salt tectonics also reviewed legacy seismic data through these diapirs (archived in
Salt diapirs and sheets are a common feature of continental passive NAMSS; Triezenberg et al. (2016)), but less thoroughly than Postaagasi
margins (e.g., Pichel et al., 2022; Rowan, 2014), including those ringing (2018) did. Where tracklines crossed close to the Carolina Trough di­
the Atlantic Ocean. Vigorous debate (e.g., Brun and Fort, 2011; Rowan apirs, the additional WCD data either revealed no unmapped seeps or
et al., 2012) continues about the timing of diapirism, the contributing were not collected appropriately for bubble plumes to be found. Based
physical processes, and potential differences between provinces with on the seismic data, four nearly-breached diapirs between 33.5◦ N and
discrete salt diapirs and those underlain by salt sheets (e.g., northern 33.83◦ N (all at 3000 mbsl or more) could be good candidates for seafloor
Gulf of Mexico). Salt emplacement can lead to faulting that provides seep exploration.
migration pathways for seep fluids, and the high thermal conductivity of Currently, there is no evidence for salt underlying seeps north of
salt warms overlying sediments (which may contain gas hydrates), in­ Cape Hatteras on the USAM. An isolated salt diapir and a possible salt
creases local heat flow, and likely enhances advection of fluids (e.g., sheet (Fig. 1) have been recognized in association with the Baltimore
Hornbach et al., 2005; Wilson and Ruppel, 2007). Active salt dissolution Canyon Trough (Grow et al., 1988; Grow, 1980; McKinney et al., 2005),
produces brines that further inhibit gas hydrate stability (Ruppel et al., as shown in Figs. 1 and 2b. The possible salt sheet discussed by
2005). McKinney et al. (2005) is thin (< 40 m), deep (estimated over 2.5 km

Fig. 8. Schematics of geologic and fluid processes responsible for the evolution of seeps on the USAM. Pink features in the sediments in each panel are vapor phase
(free gas) methane. Green circles denote seeps. (A) Salt tectonics could include salt diapirs in the South Atlantic Bight or salt sheets or diapirs in the Mid-Atlantic
Bight. At present, seeps have been found associated with salt diapirs only on the southern part of the margin (Blake Ridge and Cape Fear seeps). (B) Faults or other
pre-existing structures (such as the relic reef in Fig. 5c) may act as migration pathways or channel fluids through overlying sediments, respectively. (C) Updip fluid
migration, particularly related to dissociation of gas hydrates, could occur at the base of the GHSZ, here shown as the red dashed line cutting across stratigraphic
layering. In some locations on the margin, gas appears to be continuous across the updip limit of hydrate stability (Fig. 13d). (D) Unroofing of gas-laden sediments
through the evolution of slumps, slides, or other mass transport deposits can expose gas-charged layers. (E) Erosion, downcutting, and sapping can excavate gas-
charged layers and may be an important process at upper slope canyon heads just beneath the shelf-break. (F) While upper slope benches are not a geologic pro­
cess, they are an important setting for northern USAM seeps. In these settings, relatively intact sediments from the former upper slope host seeps that may be tapping
into older gas, some of which may have accumulated in shelf-edge deltas. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)

13
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

below the seafloor), and so far not associated with seafloor seepage on the southern New England margin and the Chincoteague Ridge field on
the shelf. The isolated diapir near Wilmington Canyon is also beneath the Virginia margin also have seeps arrayed in sublinear configurations
the shelf (Grow, 1980), and no seepage has yet been recognized there. (Fig. 2).
Mosher and Yanez-Carrizo (2021) misidentified as a diapir a feature on a For most USAM seeps, subseafloor imagery is lacking. Determining
seismic line collected by the USGS in 2014 near the deepwater part of the relative importance of faulting or other fluid migration processes in
Hudson Canyon. Because this feature has been re-interpreted as a driving seep formation and seafloor gas emission is therefore not
seamount (Gibson, 2022), it is not considered further in the context of possible. In some cases, seafloor imaging reveals gas-related seismic
seepage. wipeouts below seeps, but no obvious geologic structures that could
promote fluid migration nor a source for the gas. In other cases, the
5.1.2. Seeps associated with faults water column imaging is consistent with seepage, but strata beneath the
Fluid flow along faults feeds some seeps on the USAM. As shown in seep lack fluid flow features or even gas charging. In these cases, it is
Fig. 9, which superposes seep locations on crossing seismic data of possible that gas is supplied through a discrete conduit at resolution
different resolutions, the Norfolk Seeps tap fluids being transported below the capabilities of the seismic imaging to capture.
through faults that disrupt the uppermost hundreds of meters of sedi­
ment. The sediments sit above a mostly continuous stratum that forms 5.1.3. Seep evolution and erosional processes
the base of slide deposits on the slope. Beneath this layer, sediments are
cut by seaward-dipping, en echelon normal faults spaced at 500–750 m. 5.1.3.1. Upper slope unroofing and erosion. Erosional processes that cut
Some of the fluid-related features seen in the seismic data continue from into, disturb, or unroof (through sliding or slumping) methane-charged
below these deeper faults into the shallow faulted section. The linear sediments contribute to the development of seeps within drainages and
orientation of two arms of the Norfolk Seeps field down the eroded slope intervening ridges along the entire northern USAM and at all water
(erosion also contributes to seepage here; see below) and into the can­ depths. Unroofing is especially prevalent on the uppermost slope, where
yon’s floor may also reflect migration of fluids through shallow faults the removal of overlying strata by slides, slumps, or mass transport can
with near-linear seafloor traces. Both the deepwater Veatch seep field on expose underlying gas-charged sediments.

Fig. 9. Seismic lines through the Norfolk Seeps field (see Fig. 6), with green circles showing locations of seeps. Locations of lines shown on a seafloor relief image as
the inset. (A) Sparker high-resolution seismic line crossing nearly perpendicular to the erosional face. (B) Airgun data collected during the MATRIX cruise reveals
fine-scale normal faulting upslope and gas chimneys likely occupying faults beneath and downslope from the seep field. (For interpretation of the references to colour
in this figure legend, the reader is referred to the web version of this article.)

14
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Figure 10 shows data from three seismic lines (Ruppel et al., 2024) (Daigle and Dugan, 2010).
acquired downdip in the MAB in locations where unroofing of upper Among the examples of seep fields for which erosion is likely critical
slope sediments by slope failure processes may be associated with the for its evolution is the Norfolk Seeps field. Previously we discussed the
evolution of seeps. These high-resolution seismic lines, which were role of fluids migrating along underlying faults, but the seafloor
collected between Baltimore Canyon and just south of Washington morphology implies that erosion and slumping have probably unroofed
Canyon, reveal a similar distribution of seeps relative to the underlying gas-charged sediments to provoke seepage as well. Many Norfolk Seeps
geology. In each case, the seeps emerge from gas-charged seafloor updip lie within a heavily eroded sediment face on the north side of the canyon
from finely laminated sediments and mass transport deposits. Strata (Fig. 6). Erosion cools sediments and results in deepening of the base of
visible just below the seafloor on the shallowest parts of Lines 6 (Balti­ the gas hydrate stability zone (GHSZ), but this process is not instanta­
more Canyon) and 24 are erosionally truncated at the seafloor. Line 21 neous, especially where free gas is already within the hydrate stability
(Washington Canyon) reveals more complex structure, and some seeps zone and close to the seafloor (Ruppel and Waite, 2020). Erosion at the
emerge from disrupted material deposited during a slide event. Norfolk Seeps site may have been exacerbated by the flow of gas and
Like wholesale unroofing, more subtle erosional processes also me­ fluids in the sediments or by the accumulation of pore pressure in sed­
chanically remove near-seafloor layers or cut into gas-charged sedi­ iments, as well as by the development of lithified authigenic carbonates
ments. In some cases, layers removed by erosion may have been gas- (mechanically strong) juxtaposed against weaker sediments in the slope.
charged prior to the erosional event, making them more susceptible to The Bodie seep field (~315–460 mbsl; Fig. 3a) and others on the
erosion. In others, erosion may have stripped off less permeable sedi­ margin occupy minor erosional drainages and the sides and tops of
ments, allowing gas to escape from the newly-exposed sediments, which adjacent ridges. The Bodie seep field was likely unroofed during a
may in turn contain fractures associated with gas invasion processes Pleistocene upper continental slope slide event (Twichell et al., 2009).

Fig. 10. High-resolution sparker imagery from the upper slope of the MAB, with seep positions superposed in green. The inset map shows an upper continental slope
relief map with positions of the seismic lines. From top to bottom, the seismic lines are located south of Baltimore Canyon (Line 6), in an unnamed canyon near
Washington Canyon (Line 24), and south of Washington Canyon (Line 21). All seismic lines are plotted at the same scale, and the sections are aligned so that the 0.5 s
crossing of the seafloor is at the same lateral position to demonstrate the pervasiveness of seepage at the same depth range on this section of the margin. The nominal
updip limit of gas hydrate stability is indicated by the red line on the right. Note that sediments deeper than the seep fields on the slope are well-stratified, while a
combination of gas-charging and erosional or unroofing processes affects the areas close to the seep fields. Note that the downslope Washington Canyon seeps exist
within perturbed seafloor sediments that were likely affected by mass transport processes. (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

15
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Seepage originating in the drainages probably results from continued landward. Such erosion can excavate methane rich sediments or at least
erosion and downcutting into underlying gas-charged strata, while remove overburden above migrating methane. A process like ground­
emissions from near the crests of ridges reflects buoyancy driven water sapping may accompany such erosion. As more methane is
migration of gas beneath local bathymetric highs. emitted at seafloor seeps in these places, the sediments can become
On the southern New England margin, the New England seep fields winnowed and less cohesive, in turn enhancing erosion and fluid flow
#1 and #2 and the Veatch Canyon field (Figs. 3c and 11) have settings unless authigenic carbonate formation effectively creates a physical
like that of the Bodie Island seeps but lie ~1000 m deeper. These barrier.
southern New England margin seeps notably occur adjacent to, but not In Skarke et al. (2014), an unnamed canyon just north of Washington
within, parts of the upper slope where slide scarps and major slides have Canyon was used to illustrate the seep distribution phenomenon.
been mapped (Chaytor et al., 2007; Twichell et al., 2009; Fig. 3c) and lie Fig. 12a shows a strike line revealing gas chimneys feeding some of the
well within the GHSZ. These seep fields are therefore degassing hydrate- seeps within the scoop-shaped canyon head. We also recognize Keller
bearing sediments that are eroded, but not necessarily mechanically Canyon, just north of Cape Hatteras, as a location that hosts seeps ar­
unroofed by larger scale mass wasting events. ranged around its head (Fig. 7a). Keller Canyon breaks the shelf edge by
Arguably, nearly the entire upper slope within the northern USAM is only a few kilometers, contrasting with the 10 km or more for the large
heavily deformed by drainages, so the spatial association of upper slope shelf-breaking canyons of the MAB (Norfolk, Washington, Baltimore,
seeps with erosional features and intervening bathymetric highs may not and Wilmington), and may still be cutting into the shelf edge now.
seem remarkable. However, the segment of the northern USAM with the Shallow subseafloor near the head of Keller Canyon is gas-charged and
fewest seep fields—from just south of Hudson Canyon to Alvin Canyon, strata are hard to discern.
around the orographic bend transitioning from the Mid-Atlantic Bight to
southern New England (Fig. 3c) —coincides with the fewest shelf-break 5.1.3.3. Benches between and overlooking canyons in the MAB. Some
canyons and the most intact upper slope sediments (includes the Hudson MAB seep clusters occur on bathymetric benches that slope gently
apron section discussed by D. Brothers et al., 2013a). This observation seaward and that lie just deeper than the shelf-break between canyon
suggests a strong association among seep evolution, unroofing, and heads or adjacent canyons (Fig. 11, inset). These bench-like features are
erosion in drainages. concentrated where Twichell et al. (2009) do not detect downslope slide
deposits, meaning that the benches are likely intact remnants of the
5.1.3.2. Canyon erosion and sapping. Skarke et al. (2014) describe seeps former shelf edge. Seep clusters just south of Keller Canyon (100–125
occurring within bowl-shaped canyon heads on the upper continental mbsl), north of Washington Canyon at 190 to 340 mbsl (Fig. 12a), just
slope. In these canyons, erosional processes may be downcutting into south of Accomac Canyon at 275 to 350 mbsl (Fig. 12b), and on either
underlying gas-charged strata as the fresh erosional face migrates side of Currituck Slide upper headwall (100–130 mbsl) are a few

Fig. 11. Sparker high-resolution seismic line down the southern New England continental slope through the Veatch Canyon seeps, with green circles showing seeps.
The inset map shows a seafloor relief map with the whole seismic line in yellow and the highlighted portion for the main panel in red. The overview seismic line inset
at the left shows the full slope section, with the box corresponding to the main panel. Seeps are arrayed on either side of a ridge on the mid-continental slope, and a
chimney tapping into underlying gas is imaged beneath some of the seeps. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)

16
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Fig. 12. Seismic lines acquired along the strike of the margin on the upper continental slope between Baltimore and Washington Canyons, with seep locations
superposed as green circles. Inset shows the locations of the seismic lines on a seafloor relief map. (A) Sparker high-resolution seismic line showing seeps are arrayed
around the seaward-dipping bowl-shaded canyon head. This section is entirely shallower than the gas hydrate stability zone. Gas chimneys feed some seeps, but some
gas chimneys are not associated with seepage. (B) Airgun seismic line acquired during the 2018 MATRIX cruise. This strike line images below one of the bench
features in the Mid-Atlantic Bight, showing flat lying, but internally disrupted, sediments overlying older hummocky sediments that appear to be the remnants of a
previous shelf edge. Several generations of erosional faces are also visible. Seeps are not underlain by chimneys or other clear signs of gas migration. Gas hydrate is
not stable if the seafloor is shallower than ~0.73 s. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

examples (Fig. 2c). and seepage (Canning et al., 2021; He et al., 2023; Lu et al., 2018). On
Seismic data acquired through some bench-like upper slope features the northern USAM, Twichell et al. (2009), Poag and Sevon (1989), and
show that they generally lie within uniform, gently-dipping sediments Hill et al. (2004) describe the development of Pleistocene shelf-edge
(Fig. 12b), often with no clear gas chimney structures beneath seeps. The deltas near the outfall of major ancestral rivers in the MAB and south­
near-seafloor sediments blanket hummocky sediments that appear to be ern New England. Like modern deltas, the USAM shelf-edge deltas are
the remnants of uppermost slope drainages intersected at high angles by expected to contain ample organic matter whose breakdown leads to
the strike line. The relatively featureless benches may represent pro­ microbial production of methane. Thus, a higher concentration of seeps
graded sediment deposited at the shelf edge during a sea level lowstand. might be expected in and near mapped USAM Pleistocene shelf-edge
Older, eroded, organic-rich sediments in shelf-edge drainages buried deltas.
beneath the blanketing sediments may have accumulated microbial Figure 3 shows the locations of some of the Pleistocene shelf-edge
methane that is only now being released and migrating toward the deposits relative to the distribution of USAM seeps. The 125 mbsl iso­
surface. bath corresponds to the nominal coastline during the sea level lowstand
Bench-like upper slope morphologic features also occur north of at the LGM. In the new database, ~20% of the unique locations for
Wilmington Canyon through the Hudson Canyon segment, on the mapped seeps occur shallower than 125 mbsl, and Table 2 shows ~30%
southern New England margin (e.g., Shallop Canyon seeps), and south of are shallower than ~180 mbsl. While acknowledging that the outer shelf
Cape Hatteras. However, seeps are not as prevalent in these locations. has not been completely surveyed for seeps (especially offshore southern
New England), the area seaward of the shelf-edge delta deposits across
5.1.3.4. Pleistocene shelf features and seepage. River deltas on conti­ Hudson Canyon, along the southern New England margin, and south of
nental shelves are loci of prodigious contemporary methane generation Norfolk Canyon have some of the lowest concentrations of uppermost

17
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

slope seeps on the northern USAM. Conversely, the seeps along the 5.2.2. Special role of shelf-break canyons
upper slope between Baltimore and Accomac Canyons traverse from an The analysis of water column temperature records underscores that
area where shelf-edge deltas have been mapped to one lacking in such gas hydrate is stable within the thalwegs of canyons, even within canyon
deposits. The locations of shallower water seep clusters therefore do not heads penetrating far into the continental shelf, where gas hydrate is not
seem to correlate with shelf-edge delta deposits. stable. Davies et al. (2012, 2021) have explored gas hydrate dynamics
Pockmarks are widespread Pleistocene features on the USAM associated with canyons offshore Mauritania. In USAM shelf-break set­
(Brothers et al., 2014). Pockmarks are sometimes interpreted as the tings, the three-dimensionality of the hydrate system probably explains
consequence of sudden outgassing (Cathles et al., 2010; Hovland et al., the prevalence of seeps on the promontories of shelf-breaking canyons
2002). In contrast, many contemporary upper slope USAM seep fields from at least Keller Canyon to Shallop Canyon. As shown schematically
appear to be emitting methane continuously, as documented during in Fig. 13c, gas hydrate likely exists over free-gas within the canyon’s
repeated surveys over several years. Skarke et al. (2014) analyzed the thalweg, and Davies et al. (2012) argue that substantial accumulations
joint distribution of seeps and pockmarks on the northern USAM and of hydrate and local deepening of the base of the GHSZ occur in these
found no spatial association. In this study, many of the newly identified thalwegs. Beneath adjacent parts of the continental shelf, gas hydrate is
seeps are at water depths >600 mbsl, which is deeper than the Pleis­ not stable. On a cross-section, there is a feather edge of stability (not a
tocene USAM pockmarks. landward edge) on either side of the canyon where it disrupts the con­
tinental shelf, and gas can accumulate below the hydrate-bearing layers
5.2. Role of gas hydrates in the thalweg and sides of the canyon and probably also be driven by
buoyancy to seep points on the adjacent promontories. Gas charging of
When the USAM seeps were first described, it was postulated that sediments that make up the promontories may in turn promote
some seeps could be related to leakage of methane near the LLGHS unroofing discussed earlier for some seep fields in these geographic
(Skarke et al., 2014). This landward edge of stability nominally occurs at positions. The gas-charged strata in the canyon walls would be more
~550 to 575 mbsl on the northern USAM (Brothers et al., 2014). prone to slumping, and the deposition of this canyon wall material in the
thalweg leads to shoaling of the base of hydrate stability (e.g., Ruppel,
5.2.1. Landward limit of gas hydrate stability (LLGHS) 2003; Ruppel and Waite, 2020) and more hydrate dissociation deep in
As a first step in assessing the connection among USAM seep distri­ the GHSZ. Conversely, downcutting and erosion in the thalweg produces
butions and the dynamics of the GHSZ, we used ocean temperature thickening of the GHSZ and may promote more hydrate formation.
profiles from the World Ocean Database (WOD; Boyer et al., 2018) to These concepts require further analysis with support from seismic data,
more accurately determine the LLGHS along the margin. Conductivity- but the schematic model provides a reasonable explanation for the
temperature-depth (CTD) and expendable bathythermograph (XBT) preferential distribution of seeps next to shelf-break canyons along the
data were filtered to include only temperature profiles that continued margin’s strike.
deeper than 350 mbsl in the water column and that were acquired after
1970. The 350 mbsl cutoff is justified based on gas hydrate stability 5.2.3. Seepage and the landward hydrate stability limit
calculations. For gas hydrate of pure methane to be stable at 350 mbsl As described in detail in Ruppel and Kessler (2017) and Ruppel and
and 35‰ salinity, the expression that Ruppel and Waite (2020) provide Waite (2020), the GHSZ (GHSZ) nominally thins to zero at the landward
based on fitting the Sloan and Koh (2007) equation of state yields a edge of stability. However, gas hydrate is likely to be rare there. Not only
temperature of 1.83 ◦ C, which is unrealistically low for water column are methane and hydrate all but excluded within the zone of anaerobic
temperatures at temperate latitudes. Thus, gas hydrate could not be oxidation of methane (AOM) that occupies meters to tens of meters near
stable in sediments at water depths shallower than 350 mbsl. the seafloor away from gas seeps, but any methane hydrate that forms
At depths greater than a few hundred meters in the USAM water near the landward edge of stability is likely to experience pressure-
column, ocean temperatures usually decline with increasing water temperature conditions that oscillate back and forth across the stabil­
depth. If the deepest temperature measurement recorded in a WOD re­ ity boundary, keeping hydrate saturations low and hydrate poised close
cord was less than the calculated gas hydrate stability temperature at to dissociation. As gas hydrates dissolve or dissociate in the sediments,
that depth or the stability temperature at the seafloor depth extracted for the liberated methane gas may remain as bubbles (vapor phase) in the
the location from the Andrews et al. (2016) bathymetric map, then sediments or dissolve in pore waters (aqueous phase), depending on
seafloor at the location is within the GHSZ. When the deepest recorded local solubility conditions (Ruppel and Waite, 2020). In most cases,
temperature in the water column is too warm for seafloor at that same these processes are not expected to provoke seafloor seepage unless gas
temperature to be within the GHSZ, there is more ambiguity: The sea­ finds migration pathways or opens fractures in the sediments (Daigle
floor at such locations could be within the gas hydrate stability field if and Dugan, 2010). In that case, gas can effectively evade the strong AOM
temperatures continue to decrease with depth, but we lack appropriate sink (Stranne et al., 2022). Although gas hydrate is almost assuredly
data. The resulting map (Fig. 13) distinguishes between these ambig­ breaking down near the edge of stability on upper continental slopes
uous locations and locations where gas hydrate is definitely stable at the during the contemporary era of warming intermediate waters (e.g.,
seafloor. Lacking other information, the divide can be loosely inter­ Ruppel, 2011; Ruppel and Kessler, 2017), this degradation does not
preted as the empirically-determined LLGHS. necessarily imply that there will be in situ leakage of methane from the
Even with the relatively sparse distribution of WOD CTD and XBT seafloor.
records at the critical water depths, the most landward locations where Skarke et al. (2014) also briefly considered whether some gas
gas hydrate is stable at the seafloor lie almost exclusively near the 550 liberated by hydrate degradation processes at or deeper than the land­
mbsl isobath from Cape Hatteras to Hydrographer Canyon. From Hud­ ward limit of the GHSZ migrates upslope through permeable strata and
son Canyon northward, a few locations in the shallowest part of canyon is emitted at shallower water depths. Offshore Mauritania, Davies et al.
thalwegs have gas hydrate stable at the seafloor as shallow as ~521 (2024) recently identified a location where gas liberated during hydrate
mbsl, but 550 mbsl appears to be a reliable regional value even on the dissociation may be migrating updip to be vented at pockmarks lying
southern New England margin. South of Cape Hatteras, there are few hundreds of meters shallower than the LLGHS. Imagery acquired by the
ocean water temperature profiles on the upper slope. The landward limit USGS across the deepwater Veatch Canyon seeps (Fig. 11; Baldwin and
of stability is likely deeper on this part of the USAM owing to the Miller, 2023), which are wholly within the GHSZ, hints at potential
impingement of warm Gulf Stream waters. updip migration of gas toward the seep sites. On the MAB upper slope,
the seismic data for sparker Line 24 (Fig. 10; Ruppel et al., 2024) indi­
cate continuous gas beneath the slope from the depth corresponding to

18
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

(caption on next page)

19
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Fig. 13. Hydrate dynamics and USAM gas seepage. (A) Map showing the results of analyzing WOD CTD and XBT data (Boyer et al., 2018) to determine where
temperatures are cold enough for gas hydrate to be stable at the seafloor in the MAB. Seeps are shown as white circles. Red circles denote locations where hydrate is
stable at the seafloor, and blue circles indicate seafloor too warm for gas hydrate stability. The white isobath is 550 mbsl, which closely brackets the landward limit of
the red circles. (B) Seep distribution in the MAB (from Fig. 5b) plotted on nominal ocean thermal structure (black) and shown with hydrate stability zone (pink
shading) for 3.5 wt% NaCl and pure CH4 hydrate. (C) Schematic of hydrate, gas, and seeps on a vertically-exaggerated cross-section through a shelf-break canyon.
The promontories on the left and right are at shelf depth and too shallow for hydrate to be stable in the sediments. The canyon cuts deep enough that the canyon floor
and the deeper parts of the canyon’s sides are within the hydrate stability zone, the base of which is shown by the red dashed line. Slumping into the canyon causes
the shoaling of the BSR, while erosion deepens the BSR. Buoyancy driven gas flows can feed seeps on canyon promontories. (D) On MATRIX line MX09, the seafloor is
shown as blue, and the top of gas chosen with 95% certainty by the machine learning analysis of Majumdar et al. (2022) is shown as red. The red curve continues
across the landward limit of gas hydrate stability (yellow line), consistent with possible updip gas migration. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

the LLGHS to upper slope seep sites. For other places within the MAB, situ by degradation of Pleistocene gas hydrates.
modern high-resolution MCS imagery acquired during the MATRIX Even in the absence of large scale hydrate degradation, Late Pleis­
cruise were analyzed for seismic indicators of gas identified through tocene to contemporary climate change has altered conditions in ways
supervised machine learning (Majumdar et al., 2022). Those seismic that would have led to increased degassing from methane-charged
indicators show different characteristics beneath the upper slope in sediments on the USAM. Slope failures have been common since the
different settings. Fig. 13d shows an example for which these indicators end of the LGM (Twichell et al., 2009), and the role of these processes in
are continuous across the nominal LLGHS offshore Virginia at the unroofing deeper methane-charged sediments was discussed above.
location of MATRIX line MX09 (Fig. 1), consistent with updip migration Warming of bottom waters since the end of the LGM has changed the
of gas toward the upper slope seeps. In contrast, at the Baltimore Canyon thermal boundary condition at the seafloor and led to re-equilibration of
seeps, the seismic indicators for top of gas do not continue across the the conductive geotherm in the sediments. At water depths shallower
landward hydrate stability limit (Majumdar et al., 2022). This obser­ than the LLGHS, this re-equilibration to higher temperatures would
vation may confirm that seeps present on the promontory overlooking decrease methane solubility in pore waters, in turn producing more gas
that major shelf-break canyon may be related more to canyon-based (Ruppel and Waite, 2020). Gas buildup can increase pore pressure and
hydrate dynamics (see above) than to updip migration of fluids cause fracturing (Daigle and Dugan, 2010), even in soft sediments. These
beneath the continental slope. processes in turn enhance migration of gas, which could accumulate
One challenge for testing whether gas emissions on the northern beneath local bathymetric highs, geologic structures, or less permeable
USAM upper continental slope are connected to dissociating gas hydrate strata before eventually forming seeps. The seismic data shown in
is the composition of gases. On the Svalbard margin, seep gas was finally Figs. 8, 9, and 11 show locations where gas chimneys imaged in the
attributed to degrading gas hydrates due to the detection of higher order sediments have not produced a detected seafloor seep, implying that a
thermogenic gases in both the hydrates and seep gas (Berndt et al., variety of factors, including the lithology of near seafloor sediments and
2014). On the USAM, no hydrate samples have been recovered except at the time since the migration and accumulation of the gas, may govern
ODP Leg 164 sites (Paull et al., 1996). These samples all contained where and when seeps form. Some gas may already be leaking diffu­
99.924 to 9.993% microbially-derived methane and only trace amounts sively from locations with underlying gas build-up, and ebullition may
of higher order hydrocarbons (Matsumoto et al., 2000). Ethane (C2) is eventually develop at some of these sites.
the most abundant of the non-methane gases in the hydrate samples and At water depths greater than the landward gas hydrate stability limit,
never exceeds 0.02%. With the nearly pure methane composition of post-LGM warming would have shoaled the base of the GHSZ. As
gases so far analyzed near the USAM seeps (e.g., Pohlman et al., 2017) described in Ruppel and Waite (2020), methane released from dissoci­
and in recovered gas hydrates, the lack of a thermogenic gas tracer to ating hydrate might remain in place if overlying sediments were rela­
link emitted gases to dissociating gas hydrates, and few advances on tively impermeable or clogged with gas hydrate; could migrate into the
using fingerprints such as noble gases to trace seep gases to hydrates GHSZ and combine with free water to form more hydrate; or could
(Hunt et al., 2013), linking seeps to dissociating gas hydrates may migrate completely through the stability zone and be emitted at the
continue to prove challenging on this margin. seafloor if water were limited or throughgoing fractures were present, as
is likely occurring now at the Norfolk, Chincoteague, and Veatch Canyon
5.2.4. Late Pleistocene to contemporary climate change and gas hydrate seeps.
degradation
It is important to consider whether some seepage on the USAM up­ 6. Conclusions
permost slope could be related to degradation of gas hydrates that were
present during the last glaciation, when sea level was ~120 m lower Over the past few years, the acquisition of large amounts of multi­
(corresponding to pressure reduction of ~1.2 MPa compared to beam water column imaging data has made it possible to extend and
contemporary conditions) and when temperatures in the overlying wa­ develop an interpretative framework for the methane seeps discoveries
ters were likely cooler. For pure methane in the presence of seawater, first described by Skarke et al. (2014). Methane is being emitted as
the minimum pressure for gas hydrate to be stable corresponds to 301 m bubbles at >1000 distinct locations along the margin between Cape
water depth for 0 ◦ C and 364 m for 2 ◦ C using equations in Ruppel and Hatteras and Georges Bank, but many of the seeps occur in clusters.
Waite (2020). At present, the summertime temperature at these water Furthermore, there is no agreement on whether two holes a meter apart
depths is ~10 ◦ C offshore Virginia (e.g., Garcia-Tigreros et al., 2020). constitute a single seep or two seeps. Not every part of the margin has
There is little robust information about bottom water temperatures on been completely surveyed, and seeps are only detected if they are active
the USAM at LGM. If gas hydrate were stable at 400 mbsl at the LGM sea when a ship is passing above with sensors running. Only a small fraction
level lowstand, that location would currently be at ~520 mbsl, which is of the margin’s seeps has been visited by systems capable of photo­
shallower than the contemporary LLGHS (550–575 mbsl). Thus, any graphing, sampling, or surveying near seafloor seeps. Thus, the distri­
hydrate between 520 mbsl and the current landward limit should have bution of seep-associated features, such as methane-derived authigenic
broken down over the past ~15 kyr. As noted by Prouty et al. (2016), carbonates, hydrate mounds, seafloor chemosynthetic communities, and
any seeps currently at water depths of 450 mbsl or shallower were un­ seep-dependent water column fauna, remains poorly constrained.
likely to have been within the GHSZ at LGM. Thus, seeps at these depths Most northern USAM seeps lie on the uppermost continental slope at
are almost certainly not emitting vertically-migrating gas generated in depths shallower than a few hundred meters, which is also shallower

20
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

than the contemporary LLGHS. Several processes, including unroofing, for the southern New England margin. Two additional USGS data re­
erosion, and sapping, probably contribute to seep evolution there. leases include the MATRIX seismic data and other geophysical data for
Methane that supplies these seeps may be sourced in situ with accu­ the Mid-Atlantic Bight (Baldwin et al., 2020a, 2020b).
mulations in shelf-edge deltas or other strata or possibly produced by
contemporary methanogenesis. In some places, methane sustaining Acknowledgments
upper slope seeps may be sourced in updip migration of gas released by
dissociation of methane hydrates farther down the slope. The seeps that Open access water column data collected by NOAA’s Office of Ocean
develop on promontories overlooking shelf-break canyons probably Exploration and Research (OER) underpin this study. We particularly
trace at least part of their genesis to the complex configuration of hy­ appreciate efforts by K. Cantwell, E. Lobecker, S. Hoy, and D. Sowers
drate and gas across the thalwegs of the shelf-breaking canyons and the (NOAA). We are grateful to the crews and technical staff for the Okeanos
adjacent continental shelf areas. For seeps at depths of 1000 mbsl or Explorer and the research vessels Atlantis, Armstrong, Endeavor, Sikuliaq,
greater, fluid migration through underlying faults or the existence of and Sharp for many years of work to acquire high-quality water column
buried geologic structures (e.g., the relic reef on the southern New En­ imaging data and/or seismic imagery that supported this research. USGS
gland margin from Fig. 5c; salt diapirs in the South Atlantic Bight) technical staff based in Woods Hole and Santa Cruz ran seismic data
probably plays the most important role in seep evolution, although top- operations on the R/V Endeavor and R/V Sharp in 2015 and 2018,
down processes like erosion enhance methane emissions in some respectively. We especially thank W. Danforth, E. Moore, C. Worley, E.
locations. Bergeron, T. O’Brien, and A. Nichols (USGS) and acknowledge helpful
discussions with D. Hutchinson, J. Chaytor, and A. Demopoulos (USGS).
Funding P. Hart, W. Danforth, T. Himmler, and C. Berndt provided reviews.
Woods Hole Oceanographic Institution gave C.R. and N.C.M. access to
The USGS component of this research was supported by interagency the R/V Armstrong during shakedown cruises in 2016, the first of which
agreements with the U.S. Department of Energy (DE-FE0026195, DE- was conducted in conjunction with R. Corbett and in collaboration with
FE0023495, 89243320SFE000013) between 2015 and present, with the D. Lizarralde. J.H. Edwards assisted with initial processing of the USGS
Bureau of Ocean Energy Management (M17PG00041) for the 2018 sparker data in 2015, and Quality Positioning Services provided access
MATRIX cruise and NOAA’s Office of Ocean Exploration and Research to software used in data interpretation at Mississippi State University
(16–01118) for a 2017 cruise. Data analysis at MSU was in part sup­ (MSU). C.R. thanks D. Orange for inviting USGS personnel aboard the R/
ported by funding from the National Academy of Sciences, Engineering, V Armstrong in 2024 to work with the SBP29 and acquire the data shown
and Medicine (2000007272). in Fig. 7(B). Any use of trade, firm, or product names is for descriptive
purposes only and does not imply endorsement by the U.S. Government.
CRediT authorship contribution statement
References
C.D. Ruppel: Visualization, Methodology, Investigation, Funding
acquisition, Formal analysis, Conceptualization, Data curation, Writing - Andrews, B.D., Chaytor, J.D., Ten Brink, U.S., Brothers, D.S., Gardner, J.V., Lobecker, E.
A., Calder, B.R., 2016. Bathymetric terrain model of the Atlantic margin for marine
original draft, Writing - review & editing, Project Administration. A.D. geological investigations. In: U.S. Geological Survey Open-File Report 2012–1266.
Skarke: Visualization, Methodology, Investigation, Funding acquisition, https://doi.org/10.3133/ofr20121266. Originally posted December 3, 2013;
Formal analysis, Conceptualization, Data curation, Project Administra­ Version 2.0: May 25, 2016.
Baldwin, W.E., Miller, N.C., 2023. High-resolution Multichannel Seismic Reflection Data
tion. N.C. Miller: Visualization, Investigation, Formal analysis, Collected Along the New England Outer Continental Shelf, Slope, and Rise South of
Conceptualization, Data curation. M.W. Kidiwela: Investigation, Martha’s Vineyard and Nantucket, Massachusetts, U.S. Geological Survey Field
Formal analysis. J. Kluesner: Visualization, Investigation, Formal Activity 2016–018-FA. U.S. Geological Survey Data Release. sciencebase.gov. htt
ps://doi.org/10.5066/P9R3VW5H.
analysis, Data curation. W. Baldwin: Visualization, Investigation, Baldwin, W.E., Foster, D.S., Bergeron, E.M., Dal Ferro, P., Mckee, J.A., Moore, E.M.,
Formal analysis, Data curation. Nichols, A.R., O’brien, T.F., Powers, D., Miller, N.C., Ruppel, C.D., 2020a.
Multichannel Seismic-Reflection and Navigation Data Collected Using Sercel GI Guns
and Geometrics GeoEel Digital Streamers during the Mid-Atlantic Resource Imaging
Declaration of competing interest Experiment (MATRIX), USGS Field Activity 2018–002-FA. U.S. Geological Survey
Data Release. sciencebase.gov. https://doi.org/10.5066/P91WP1RZ.
Carolyn Ruppel reports financial support was provided by Bureau of Baldwin, W.E., Moore, E.M., Worley, C.R., Nichols, A.R., Ruppel, C.D., 2020b. Marine
Geophysical Data Collected to Support Methane Seep Research along the U.S.
Ocean Energy Management for scientific activities conducted by herself,
Atlantic Continental Shelf Break and Upper Continental Slope Between the Baltimore
Nathan Miller, and Wayne Baldwin. Carolyn Ruppel reports financial and Keller Canyons. U.S. Geological Survey Data Release. sciencebase.gov. https://do
support was provided by US Department of Energy for scientific activ­ i.org/10.5066/P9Y1MSTN.
ities conducted by herself, Nathan Miller, Jared Kluesner, and Wayne Berndt, C., Feseker, T., Treude, T., Krastel, S., Liebetrau, V., Niemann, H., Bertics, V.J.,
Dumke, I., Dünnbier, K., Ferré, B., Graves, C., Gross, F., Hissmann, K.,
Baldwin. Carolyn Ruppel reports financial support was provided by Hühnerbach, V., Krause, S., Lieser, K., Schauer, J., Steinle, L., 2014. Temporal
NOAA Ocean Exploration and Research. Adam Skarke reports financial constraints on hydrate-controlled methane seepage off Svalbard. Science 343,
support was provided by National Academies of Sciences Engineering 284–287. https://doi.org/10.1126/science.1246298.
Bourque, J.R., Robertson, C.M., Brooke, S., Demopoulos, A.W.J., 2017. Macrofaunal
and Medicine for scientific activities conducted by himself and Maleen communities associated with chemosynthetic habitats from the U.S. Atlantic margin:
Kidiwela. a comparison among depth and habitat types. Deep-Sea Res. II Top. Stud. Oceanogr.
137, 42–55. https://doi.org/10.1016/j.dsr2.2016.04.012.
Boyer, T.P., Baranova, O.K., Coleman, C., Garcia, H.E., Grodsky, A., Locarnini, R.A.,
Data availability statement Mishonov, A.V., Paver, C.R., Reagan, J.R., Seidov, D., Smolyar, I.V., Weathers, K.,
Zweng, M.M., 2018. World Ocean Database 2018. NOAA Atlas NESDIS 87.
The datasets presented in this study can be found in online re­ https://www.ncei.noaa.gov/sites/default/files/2020-04/wod_intro_0.pdf.
Brooks, J.M., Kennicutt, M.C., Fay, R.R., Mcdonald, T.J., Sassen, R., 1984. Thermogenic
positories. Skarke et al. (2024), which is released through the National Gas Hydrates in the Gulf of Mexico. Science 225, 409–411. https://doi.org/10.1126/
Centers for Environmental Information (NCEI), includes the seeps science.225.4660.409.
database (raw), the seeps database with duplicates removed, and the Brothers, D.S., Ten Brink, U.S., Andrews, B.D., Chaytor, J.D., 2013a. Geomorphic
characterization of the U.S. Atlantic continental margin. Mar. Geol. 338, 46–63.
seeps database with clusters and unique sites. The US Rolling Deck to
https://doi.org/10.1016/j.margeo.2012.12.008.
Repository Data or the NCEI archives data for native shipboard systems Brothers, D.S., Ruppel, C., Kluesner, J.W., Ten Brink, U.S., Chaytor, J.D., Hill, J.C.,
for non-NOAA and NOAA cruises, respectively. Ruppel et al. (2024) Andrews, B.D., Flores, C., 2014. Seabed fluid expulsion along the upper slope and
provides processed sparker data on the upper continental slope in the U. outer shelf of the U.S. Atlantic continental margin. Geophys. Res. Lett. https://doi.
org/10.1002/2013gl058048.
S. Mid-Atlantic Bight, and Baldwin and Miller (2023) have sparker data

21
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

Brothers, L.L., Van Dover, C.L., German, C.R., Kaiser, C.L., Yoerger, D.R., Ruppel, C.D., He, S., Maiti, K., Ghaisas, N., Upreti, K., Rivera-Monroy, V.H., 2023. Potential methane
Lobecker, E., Skarke, A.D., Wagner, J.K.S., 2013b. Evidence for extensive methane production in oligohaline wetlands undergoing erosion and accretion in the
venting on the southeastern U.S. Atlantic margin. Geology G34217 (1). https://doi. Mississippi River Delta Plain, Louisiana, USA. Sci. Total Environ. 875, 162685
org/10.1130/g34217.1. https://doi.org/10.1016/j.scitotenv.2023.162685.
Brun, J.-P., Fort, X., 2011. Salt tectonics at passive margins: Geology versus models. Mar. Hill, J.C., Driscoll, N.W., Weissel, J.K., Goff, J.A., 2004. Large-scale elongated gas
Pet. Geol. 28, 1123–1145. https://doi.org/10.1016/j.marpetgeo.2011.03.004. blowouts along the U.S. Atlantic margin. J. Geophys. Res. Solid Earth 109. https://
Bureau of Ocean Energy Management (BOEM), 2021. 2021 Assessment of Technically doi.org/10.1029/2004JB002969.
and Economically Recoverable Oil and Natural Gas Resources of the U.S. Atlantic Hornbach, M.J., Ruppel, C., Saffer, D.M., Van Dover, C.L., Holbrook, W.S., 2005. Coupled
Outer Continental Shelf, OCS Report BOEM 2021–085, p. 83. Available at: https geophysical constraints on heat flow and fluid flux at a salt diapir. Geophys. Res.
://www.boem.gov/sites/default/files/documents/regions/atlantic-ocs-region/ Lett. 32, L24617 https://doi.org/10.1029/2005GL024862.
2021%20Atlantic%20Assessment%20Final_508_0.pdf. Hornbach, M.J., Ruppel, C., Van Dover, C.L., 2007. Three-dimensional structure of fluid
Canning, A., Wehrli, B., Körtzinger, A., 2021. Methane in the Danube Delta: the conduits sustaining an active deep marine cold seep. Geophys. Res. Lett. 34 https://
importance of spatial patterns and diel cycles for atmospheric emission estimates. doi.org/10.1029/2006GL028859.
Biogeosciences 18, 3961–3979. https://doi.org/10.5194/bg-18-3961-2021. Hovland, M., Gardner, J.V., Judd, A.G., 2002. The significance of pockmarks to
Cantwell, K., 2019. Final Project Instructions. EX-19-03 Leg 2: Mid and Southeast (ROV understanding fluid flow processes and geohazards. Geofluids 2, 127–136. https://
& Mapping), June 20 - July 12, 2019. https://doi.org/10.25923/ywdg-9g76. doi.org/10.1046/j.1468-8123.2002.00028.x.
Cathles, L.M., Su, Z., Chen, D., 2010. The physics of gas chimney and pockmark Hunt, A.G., Stern, L., Pohlman, J.W., Ruppel, C., Moscati, R.J., Landis, G.P., 2013. Mass
formation, with implications for assessment of seafloor hazards and gas fractionation of noble gases in synthetic methane hydrate: Implications for naturally
sequestration. Mar. Pet. Geol. 27, 82–91. https://doi.org/10.1016/j. occurring gas hydrate dissociation. Chem. Geol. 339, 242–250. https://doi.org/
marpetgeo.2009.09.010. 10.1016/j.chemgeo.2012.09.033.
Chaytor, J.D., Twichell, D.C., Brink, U.S.T., Buczkowski, B.J., Andrews, B.D., 2007. Johnson, H.P., Miller, U.K., Salmi, M.S., Solomon, E.A., 2015. Analysis of bubble plume
Revisiting submarine mass movements along the U.S. Atlantic Continental margin: distributions to evaluate methane hydrate decomposition on the continental slope.
Implications for Tsunami Hazards. In: Lykousis, V., Sakellariou, D., Locat, J. (Eds.), Geochem. Geophys. Geosyst. 16, 3825–3839. https://doi.org/10.1002/
Submarine Mass Movements and their Consequences: 3 International Symposium. 2015GC005955.
Springer Netherlands, Dordrecht. https://doi.org/10.1007/978-1-4020-6512-5_41. Johnson, H.P., Merle, S., Salmi, M., Embley, R., Sampaga, E., Lee, M., 2019. Anomalous
Cordes, E., 2019. Cruise Report: OER DeepSEARCH 4th Cruise, Deepwater Atlantic concentration of methane emissions at the continental shelf edge of the Northern
Habitats II: Continued Atlantic Research and Exploration in Deepwater Ecosystems Cascadia margin. J. Geophys. Res. Solid Earth 124, 2829–2843. https://doi.org/
with Focus on Coral, Canyon and Seep Communities 2019, p. 127. Available at: htt 10.1029/2018JB016453.
ps://repository.library.noaa.gov/view/noaa/27790. Joung, D.J., Ruppel, C.D., Southon, J., Weber, T.S., Kessler, J., 2022. Negligible
Coykendall, D.K., Cornman, R.S., Prouty, N.G., Brooke, S., Demopoulos, A.W.J., atmospheric release of methane from decomposing hydrates in mid-latitude oceans.
Morrison, C.L., 2019. Molecular characterization of Bathymodiolus mussels and gill Nat. Geosci. https://doi.org/10.1038/s41561-022-01044-8.
symbionts associated with chemosynthetic habitats from the U.S. Atlantic margin. Kennicutt, M.C., 2017. Oil and Gas Seeps in the Gulf of Mexico. In: Ward, C.H. (Ed.),
PLoS One 14, e0211616. https://doi.org/10.1371/journal.pone.0211616. Habitats and Biota of the Gulf of Mexico: Before the Deepwater Horizon Oil Spill:
Daigle, H., Dugan, B., 2010. Origin and evolution of fracture-hosted methane hydrate Volume 1: Water Quality, Sediments, Sediment Contaminants, Oil and Gas Seeps,
deposits. J. Geophys. Res. Solid Earth 115. https://doi.org/10.1029/2010JB007492. Coastal Habitats, Offshore Plankton and Benthos, and Shellfish. New York, NY,
Davies, R., Yang, J., Ireland, M., Berndt, C., Huuse, M., Morales-Maqueda, M., 2024. Springer New York, pp. 275–358. https://doi.org/10.1007/978-1-4939-3447-8_5.
Long-distance migration and venting of methane after marine hydrate dissociation. Knott, S.T., Hoskins, H., 1968. Evidence of Pleistocene events in the structure of the
Nat. Geosci. 17, 32–37. https://doi.org/10.1038/s41561-023-01333-w. continental shelf off the northeastern United States. Mar. Geol. 6, 5–43. https://doi.
Davies, R.J., Thatcher, K.E., Mathias, S.A., Yang, J., 2012. Deepwater canyons: an escape org/10.1016/0025-3227(68)90007-8.
route for methane sealed by methane hydrate. Earth Planet. Sci. Lett. 323-324, Leonte, M., Ruppel, C.D., Ruiz-Angulo, A., Kessler, J.D., 2020. Surface methane
72–78. https://doi.org/10.1016/j.epsl.2011.11.007. concentrations along the Mid-Atlantic Bight driven by aerobic subsurface production
Davies, R.J., Maqueda, M.Á.M., Li, A., Ireland, M., 2021. Climatically driven instability rather than seafloor gas seeps. J. Geophys. Res. Oceans 125. https://doi.org/
of marine methane hydrate along a canyon-incised continental margin. Geology 49, 10.1029/2019JC015989 e2019JC015989.
973–977. https://doi.org/10.1130/G48638.1. Lu, X., Zhou, Y., Zhuang, Q., Prigent, C., Liu, Y., Teuling, A., 2018. Increasing methane
Demopoulos, A.W.J., Mcclain-Counts, J., Bourque, J., Prouty, N.G., Smith, B., Brooke, S., emissions from natural land ecosystems due to sea-level rise. J. Geophys. Res.
Ross, S.W., Ruppel, C., 2019. Examination of Bathymodiolus childressi nutritional Biogeosci. 123, 1756–1768. https://doi.org/10.1029/2017JG004273.
sources, isotopic niches, and food-web linkages at two seeps in the US Atlantic Majumdar, U., Miller, N.C., Ruppel, C.D., 2022. Neural net detection of seismic features
margin using stable isotope analysis and mixing models. Deep Sea Res. Part I 148, related to gas hydrates and free gas accumulations on the northern U.S. Atlantic
53–66. https://doi.org/10.1016/j.dsr.2019.04.002. margin. Interpretation 10, T785–T806. https://doi.org/10.1190/INT-2021-0248.1.
Dillon, W.P., Popenoe, P., Grow, J.A., Klitgord, K.D., Swift, B.A., Paull, C.K., Cashman, K. Maritime, Kongsberg, 2012. Kongsberg EM302 Multibeam Echo Sounder: Product
V., 1982. Growth faulting and salt diapirism: Their relationship and control on the Description. Available at: https://www.kongsberg.com/globalassets/maritime/km-
Carolina Trough, eastern North America. In: Watkins, J.S., Drake, C.L. (Eds.), Studies products/product-documents/306106_em_302_product_specification.pdf.
of Continental Margin Geology, vol. 34, pp. 21–46. Matsumoto, R., Uchida, T., Waseda, A., Uchida, T., Takeya, S., Hirano, T., Yamada, K.,
Driscoll, N.W., Weissel, J.K., Goff, J.A., 2000. Potential for large-scale submarine slope Maeda, Y., Okui, T., 2000. Occurrence, structure, and composition of natural gas
failure and tsunami generation along the U.S. mid-Atlantic coast. Geology 28, hydrate recovered from the Blake Ridge, Northwest Atlantic. In: Paull, C.K.,
407–410. https://doi.org/10.1130/0091-7613(2000)28<407:PFLSSF>2.0.CO;2. Matsumoto, R., Wallace, P.J., Dillon, W.P. (Eds.), Proceedings of the Ocean Drilling
Egeberg, P.K., 2000. Hydrates associated with fluid flow above salt diapirs (Site 996). In: Program, Scientific Results, vol. 164. Ocean Drilling Program, College Station, TX,
Paull, C.K., Matsumoto, R., Wallace, P.J., Dillon, W.P. (Eds.), Proceedings of the pp. 13–28. https://doi.org/10.2973/odp.proc.sr.164.247.2000.
Ocean Drilling Program, Scientific Results. Ocean Drilling Program. https://doi.org/ McCave, I.N., Tucholke, B.E., 1986. Deep current-controlled sedimentation in the
10.2973/odp.proc.sr.164.218.2000. western North Atlantic. In: Vogt, P.R., Tucholke, B.E. (Eds.), The Western North
Fu, X., Waite, W.F., Ruppel, C.D., 2021. Hydrate formation on marine seep bubbles and Atlantic Region. Geological Society of America, Boulder, Colorado, pp. 451–468.
the implications for water column methane dissolution. J. Geophys. Res. Oceans 126, https://doi.org/10.1130/DNAG-GNA-M.451.
e2021JC017363. https://doi.org/10.1029/2021JC017363. McKinney, B.A., Lee, M.W., Agena, W.F., Poag, C.W., 2005. Early to middle jurassic salt
Garcia-Tigreros, F., Leonte, M., Joung, D., Kessler, J.D., Ruppel, C., 2020. Water in baltimore canyon trough. In: U.S. Geological Survey Open File Report,
temperature, salinity, oxygen, dissolved inorganic carbon, pH, and others collected 2004–1435. https://doi.org/10.3133/ofr20041435.
by CTD and Niskin bottles from research vessel Hugh R. In: Sharp in Mid-Atlantic McVeigh, D., Skarke, A., Dekas, A.E., Borrelli, C., Hong, W.L., Marlow, J.J., Pasulka, A.,
Bight from 2017-08-25 to 2017-09-03 (NCEI Accession 0209187). NOAA National Jungbluth, S.P., Barco, R.A., Djurhuus, A., 2018. Characterization of benthic
Center for Environmental Information. https://www.ncei.noaa.gov/archiv biogeochemistry and ecology at three methane seep sites on the Northern U.S.
e/accession/0209187. Atlantic margin. Deep-Sea Res. II Top. Stud. Oceanogr. 150, 41–56. https://doi.org/
Garcia-Tigreros, F., Leonte, M., Ruppel, C.D., Ruiz-Angulo, A., Joung, D.J., Young, B., 10.1016/j.dsr2.2018.03.001.
Kessler, J.D., 2021. Estimating the Impact of Seep methane Oxidation on Ocean pH Merle, S.G., Embley, R.W., Johnson, H.P., Lau, T.-K., Phrampus, B.J., Raineault, N.A.,
and Dissolved Inorganic Radiocarbon along the U.S. Mid-Atlantic Bight. J. Geophys. Gee, L.J., 2021. Distribution of methane Plumes on Cascadia margin and
Res. Biogeosci. 126, e2019JG005621 https://doi.org/10.1029/2019JG005621. Implications for the LLGHS. Front. Earth Sci. 9 https://doi.org/10.3389/
Gibson, J.C., 2022. Controls on Surface and Sedimentary Processes on Continental feart.2021.531714.
Margins from Geophysical Data: New Insights at Cascadia, Galicia, and the Eastern Meyer, K.S., Wagner, J.K.S., Ball, B., Turner, P.J., Young, C.M., Van Dover, C.L., 2016.
North American Margin. Ph.D.. Columbia University, p. 202. https://doi.org/ Hyalinoecia artifex: Field notes on a charismatic and abundant epifaunal polychaete
10.7916/9jd8-m806. on the US Atlantic continental margin. Invertebr. Biol. 135, 211–224. https://doi.
Grow, J.A., 1980. Deep structure and evolution of the Baltimore Canyon trough in the org/10.1111/ivb.12132.
vicinity of the COST no. B-3 well. In: Scholle, P.A. (Ed.), Geological Studies of the Mosher, D.C., Yanez-Carrizo, G., 2021. The elusive continental rise: Insights from
COST No. B-3 Well, United States Mid-Atlantic Continental Slope Area, 833. U.S. residual bathymetry analysis of the Northwest Atlantic margin. Earth Sci. Rev. 217,
Geological Survey Circular, pp. 117–124. https://doi.org/10.3133/cir833. 103608 https://doi.org/10.1016/j.earscirev.2021.103608.
Grow, J.A., Klitgord, K.D., Schlee, J.S., 1988. Structure and evolution of Baltimore Newman, K.R., Cormier, M.-H., Weissel, J.K., Driscoll, N.W., Kastner, M., Solomon, E.A.,
Canyon Trough. In: Sheridan, R.E., Grow, J.A. (Eds.), The Atlantic Continental Robertson, G., Hill, J.C., Singh, H., Camilli, R., Eustice, R., 2008. Active methane
Margin. Geological Society of America, Boulder, Colorado, pp. 269–290. https://doi. venting observed at giant pockmarks along the U.S. mid-Atlantic shelf break. Earth
org/10.1130/DNAG-GNA-I2.269. Planet. Sci. Lett. 267, 341–352. https://doi.org/10.1016/j.epsl.2007.11.053.

22
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

NOAA Office of Ocean Exploration and Research, 2013a. Water Column Sonar Data Ruppel, C., Kluesner, J., Pohlman, J., Brothers, D., Colwell, F., Krause, S., Treude, T.,
Collection (EX1304L1, EM302). NOAA. National Centers for Environmental 2015. Methane hydrate dynamics on the Northern US Atlantic margin. DOE Fire Ice
Information. https://doi.org/10.7289/V51N7Z23. Newslett., 10-13. [Online] Available: https://netl.doe.
NOAA Office of Ocean Exploration and Research, 2013b. Water Column Sonar Data gov/sites/default/files/publication/MHNews_2015_December.pdf#page=10.
Collection (EX1304L2, EM302). NOAA. National Centers for Environmental Ruppel, C., Demopoulos, A., Prouty, N., 2018. Exploring US Mid-Atlantic margin
Information. https://doi.org/10.7289/V5WW7FKS. methane Seeps: IMMeRSS, May 2017. Suppl. Oceanogr. 31 (1) [Online] Available:
NOAA Office of Ocean Exploration and Research, 2014a. Water Column Sonar Data https://tos.org/oceanography/assets/docs/31-1_supplement.pdf.
Collection (EX1403, EM302). NOAA. https://doi.org/10.7289/V53F4MH6. Ruppel, C., Kluesner, J.D., Edwards, J.H., Danforth, B., Miller, N.C., Worley, C.,
NOAA Office of Ocean Exploration and Research, 2014b. Water Column Sonar Data Bergeron, E., O’Brien, T., Nichols, A., 2024. Processed multi-channel seismic
Collection (EX1402L1, EM302). NOAA. https://doi.org/10.7289/V5NC5Z42. reflection data from the U.S. Mid-Atlantic margin (EN555, 2015). Marine Geophys.
NOAA Office of Ocean Exploration and Research, 2014c. Water Column Sonar Data Data Syst. https://doi.org/10.60521/331588.
Collection (EX1404L1, EM302). NOAA. https://doi.org/10.7289/V5PC30BQ. Ruppel, C.D., Waite, W.F., 2020. Timescales and processes of methane hydrate formation
NOAA Office of Ocean Exploration and Research, 2014d. Water Column Sonar Data and breakdown, with application to geologic systems. J. Geophys. Res. Solid Earth
Collection (EX1404L2, EM302). NOAA. https://doi.org/10.7289/V5DZ0690. 125, e2018JB016459. https://doi.org/10.1029/2018JB016459.
Paull, C.K., Spiess, F.N., Ussler III, W., Borowski, W.S., 1995. Methane-rich plumes on the Ruppel, C.D., Shedd, W., Miller, N.C., Kluesner, J., Frye, M., Hutchinson, D., 2022. US
Carolina continental rise: associations with gas hydrates. Geology 23, 89–92. Atlantic margin Gas Hydrates. In: Mienert, J., Berndt, C., Trehu, A.M.,
https://doi.org/10.1130/0091-7613(1995)023<0089:MRPOTC>2.3.CO;2. Camerlenghi, A., Liu, C.-S. (Eds.), World Atlas of Submarine Gas Hydrates in
Paull, C.K., Matsumoto, R., Wallace, P.J., et al., 1996. Proceedings of the Ocean Drilling Continental Margins. Springer Cham, Switzerland, pp. 287–302. https://doi.org/
Program, Initial Reports. Ocean Drilling Program, College Station, TX. 10.1007/978-3-030-81186-0_24.
Pichel, L.M., Huismans, R.S., Gawthorpe, R., Faleide, J.I., Theunissen, T., 2022. Late-syn- Sassen, R., Sweet, S.T., Milkov, A.V., Defreitas, D.A., Kennicutt, M.C. Ii, 2001.
to post-rift salt tectonics on wide rifted margins—insights from geodynamic Thermogenic vent gas and gas hydrate in the Gulf of Mexico slope: is gas hydrate
modeling. Tectonics 41, e2021TC007158. https://doi.org/10.1029/2021TC007158. decomposition significant? Geology 29, 107–110. https://doi.org/10.1130/0091-
Poag, C.W., 1985. Geologic Evolution of the United States Atlantic Margin. Van Nostrand 7613(2001)029<0107:TVGAGH>2.0.CO;2.
Reinhold, New York, N.Y, p. 384. Schlee, J.S., Dillon, W.P., Grow, J.A., 1979. Structure of the Continental Slope off the
Poag, C.W., Sevon, W.D., 1989. A record of Appalachian denudation in postrift Mesozoic Eastern United States. In: Doyle, L.J., Pilkey, O.H. (Eds.), Geology of Continental
and Cenozoic sedimentary deposits of the U.S. Middle Atlantic continental margin. Slopes, vol. 27. SEPM Society for Sedimentary Geology Special Publication,
Geomorphology 2, 119–157. https://doi.org/10.1016/0169-555X(89)90009-3. pp. 95–117. https://doi.org/10.2110/pec.79.27.0095.
Pohlman, J., Ruppel, C.D., Wang, D.T., Ono, S., Kluesner, J., Xu, X., Sylva, S., Casso, M., Skarke, A., Ruppel, C., Kodis, M., Brothers, D., Lobecker, E., 2014. Widespread methane
2017. Natural gas sources from methane seeps on the Northern U.S. Atlantic margin. leakage from the sea floor on the northern US Atlantic margin. Nat. Geosci. 7,
In: EOS Transactions American Geophysical Union, Fall Meeting, OS11B-1133. 657–661. https://doi.org/10.1038/ngeo2232.
Portanier, E., Nicolle, A., Rath, W., Monnet, L., Le Goff, G., Le Port, A.-S., Daguin- Skarke, A., Ruppel, C., Hoy, S., 2019. The 100th NOAA Ship Okeanos Explorer Mission
Thiébaut, C., Morrison, C.L., Cunha, M.R., Betters, M., Young, C.M., Van Dover, C.L., Visits New Methane Plumes Where the U.S. Atlantic Seeps Story Began [Online].
Biastoch, A., Thiébaut, E., Jollivet, D., 2023. Coupling large-spatial scale larval Available: https://oceanexplorer.noaa.gov/okeanos/explorations/ex1903/logs/jul
dispersal modelling with barcoding to refine the amphi-Atlantic connectivity y12/july12.html [Accessed August 8 2020].
hypothesis in deep-sea seep mussels. Front. Mar. Sci. 10 https://doi.org/10.3389/ Skarke, A., Ruppel, C.D., Kidiwela, M., Kodis, M., Danforth, W., Baldwin, W., 2024. U.S.
fmars.2023.1122124. Atlantic Margin Cold Seeps Database from Multibeam Water Column Imagery,
Postaagasi, C., 2018. Salt Tectonism in the Carolina Trough. M.S.. University of South 2011–2016: South Atlantic Bight to Georges Bank. National Center for
Carolina, p. 53 https://scholarcommons.sc.edu/etd/4685/. Environmental Information. https://doi.org/10.25921/xm5q-p940.
Prouty, N.G., Sahy, D., Ruppel, C.D., Roark, E.B., Condon, D., Brooke, S., Ross, S.W., Skarke, A.D., Ruppel, C.D., Kidiwela, M.W., Baldwin, W.E., Danforth, W.W., 2018.
Demopoulos, A.W.J., 2016. Insights into methane dynamics from analysis of Expanded U.S. Atlantic Margin Seep Inventory Yields Insight into Methane
authigenic carbonates and chemosynthetic mussels at newly-discovered Atlantic Dynamics. EOS Transactions of the American Geophysical Union, Washington D.C..
margin seeps. Earth Planet. Sci. Lett. 449, 332–344. https://doi.org/10.1016/j. OS33C-1913.
epsl.2016.05.023. Sloan, E.D., Koh, C.A., 2007. Clathrate Hydrates of Natural Gases, Third edition. CRC
Prouty, N.G., Campbell, P.L., Close, H., Biddle, J.F., Beckmann, S., 2020. Molecular Press, Boca Raton, p. 752.
indicators of methane metabolisms at cold seeps along the United States Atlantic Stranne, C., O’Regan, M., Hong, W.-L., Brüchert, V., Ketzer, M., Thornton, B.F.,
margin. Chem. Geol., 119603 https://doi.org/10.1016/j.chemgeo.2020.119603. Jakobsson, M., 2022. Anaerobic oxidation has a minor effect on mitigating seafloor
QGIS.org, 2023. QGIS Geographic Information System. methane emissions from gas hydrate dissociation. Commun. Earth Environ. 3, 163.
Quattrini, A.M., Nizinski, M.S., Chaytor, J.D., Demopoulos, A.W.J., Roark, E.B., https://doi.org/10.1038/s43247-022-00490-x.
France, S.C., Moore, J.A., Heyl, T., Auster, P.J., Kinlan, B., Ruppel, C., Elliott, K.P., Triezenberg, P.J., Hart, P.E., Childs, J.R., 2016. National Archive of Marine Seismic
Kennedy, B.R.C., Lobecker, E., Skarke, A., Shank, T.M., 2015. Exploration of the Surveys (NAMSS): A USGS Data Website of Marine Seismic Reflection Data within
canyon-incised continental margin of the Northeastern United States reveals the U.S. Exclusive Economic Zone (EEZ). U.S. Geological Survey Data Release.
dynamic habitats and diverse communities. PLoS One 10, e0139904. https://doi. https://doi.org/10.5066/F7930R7P.
org/10.1371/journal.pone.0139904. Turner, P.J., Ball, B., Diana, Z., Fariñas-Bermejo, A., Grace, I., Mcveigh, D., Powers, M.
Rebesco, M., Hernández-Molina, F.J., Van Rooij, D., Wåhlin, A., 2014. Contourites and M., Van Audenhaege, L., Maslakova, S., Young, C.M., Van Dover, C.L., 2020.
associated sediments controlled by deep-water circulation processes: State-of-the-art Methane seeps on the US Atlantic margin and their potential importance to
and future considerations. Mar. Geol. 352, 111–154. https://doi.org/10.1016/j. populations of the commercially valuable deep-sea red crab, Chaceon quinquedens.
margeo.2014.03.011. Front. Mar. Sci. 7, 75. https://doi.org/10.3389/fmars.2020.00075.
Riedel, M., Scherwath, M., Römer, M., Veloso, M., Heesemann, M. & Spence, G. D. 2018. Twichell, D.C., Chaytor, J.D., Brink, U.S., Buczkowski, B.J., 2009. Morphology of late
Distributed natural gas venting offshore along the Cascadia margin. Nat. Commun., Quaternary submarine landslides along the U.S. Atlantic continental margin. Mar.
9, 3264. doi:https://doi.org/10.1038/s41467-018-05736-x. Geol. 264, 4–15. https://doi.org/10.1016/j.margeo.2009.01.009.
Roberts, H.H., Carney, R.S., 1997. Evidence of episodic fluid, gas, and sediment venting Uchupi, E., Driscoll, N., Ballard, R.D., Bolmer, S.T., 2001. Drainage of late Wisconsin
on the northern Gulf of Mexico continental slope. Econ. Geol. 92, 863–879. https:// glacial lakes and the morphology and late quaternary stratigraphy of the New
doi.org/10.2113/gsecongeo.92.7-8.863. Jersey–southern New England continental shelf and slope. Mar. Geol. 172, 117–145.
Rowan, M.G., 2014. Passive-margin salt basins: hyperextension, evaporite deposition, https://doi.org/10.1016/S0025-3227(00)00106-7.
and salt tectonics. Basin Res. 26, 154–182. https://doi.org/10.1111/bre.12043. Van Dover, C.L., Aharon, P., Bernhard, J.M., Caylor, E., Doerries, M., Flickinger, W.,
Rowan, M.G., Peel, F.J., Vendeville, B.C., Gaullier, V., 2012. Salt tectonics at passive Gilhooly, W., Goffredi, S.K., Knick, K.E., Macko, S.A., Rapoport, S., Raulfs, E.C.,
margins: Geology versus models – Discussion. Mar. Pet. Geol. 37, 184–194. https:// Ruppel, C., Salerno, J.L., Seitz, R.D., Sen Gupta, B.K., Shank, T., Turnipseed, M.,
doi.org/10.1016/j.marpetgeo.2012.04.007. Vrijenhoek, R., 2003. Blake Ridge methane seeps: characterization of a soft-
Rudebusch, J.A., Prouty, N.G., Conrad, J.E., Watt, J.T., Kluesner, J.W., Miller, N.C., sediment, chemosynthetically based ecosystem. Deep-Sea Res. I Oceanogr. Res. Pap.
Watson, S.J., Hillman, J.I.T., 2023. Diving deeper into seep distribution along the 50, 281–300. https://doi.org/10.1016/S0967-0637(02)00162-0.
cascadia convergent margin, USA. Front. Mar. Sci. 11, 16. https://doi.org/10.3389/ Ventress, W.P.S., Bebout, D.G., Perkins, B.F., Moore, C.H. (Eds.), 1989. Gulf of Mexico
feart.2023.1205211. Salt Tectonics, Associated Processes and Exploration Potential, 10. SEPM Society for
Ruppel, C., 2003. Thermal State of the Gas Hydrate Reservoir. In: Max, M.D. (Ed.), Sedimentary Geology, p. 165. https://doi.org/10.5724/gcs.89.10.
Natural Gas Hydrate: in Oceanic and Permafrost Environments. Springer Wehner, H., Faber, E., Hufnagel, H., 2000. Characterization of low and high molecular-
Netherlands, Dordrecht. https://doi.org/10.1007/978-94-011-4387-5_4. weight hydrocarbons in sediments from the Blake Ridge, sites 994, 995, and 997. In:
Ruppel, C., 2011. Methane hydrates and contemporary climate change. Nat. Educ. Paull, C.K., Matsumoto, R., Wallace, P.J., Dillon, W.P. (Eds.), Proceedings of the
Knowl. 3. Available: http://www.nature.com/scitable/knowledge/library/metha Ocean Drilling Program, Scientific Results, 164, vol. 164. Ocean Drilling Program,
ne-hydrates-and-contemporary-climate-change-24314790. College Station, TX, pp. 47–58. https://doi.org/10.2973/odp.proc.sr.164.225.2000.
Ruppel, C., Demopoulos, A., 2013. Chemosynthetic Communities and Gas Hydrates at Weinstein, A., Navarrete, L., Ruppel, C., Weber, T.C., Leonte, M., Kellermann, M.Y.,
Cold Seeps South of Nantucket [Online] Available: https://oceanexplorer.noaa. Arrington, E.C., Valentine, D.L., Scranton, M.I., Kessler, J.D., 2016. Determining the
gov/okeanos/explorations/ex1304/logs/july12/july12.html [Accessed 6/29/23].
Ruppel, C., Kessler, J.D., 2017. The interaction of climate change and methane hydrates.
Rev. Geophys. 55 https://doi.org/10.1002/2016RG000534.
Ruppel, C., Dickens, G.R., Castellini, D.G., Gilhooly, W., Lizarralde, D., 2005. Heat and
salt inhibition of gas hydrate formation in the northern Gulf of Mexico. Geophys.
Res. Lett. 32 https://doi.org/10.1029/2004GL021909.

23
C.D. Ruppel et al. Marine Geology 471 (2024) 107287

flux of methane into Hudson Canyon at the edge of methane clathrate hydrate Withjack, M., Malinconico, M., Durcanin, M., 2020. The “Passive” margin of Eastern
stability. Geochem. Geophys. Geosyst. 17 https://doi.org/10.1002/2016GC006421. North America: rifting and the influence of prerift orogenic activity on postrift
Wilson, A., Ruppel, C., 2007. Salt tectonics and shallow subseafloor fluid convection: development. Lithosphere 2020, 8876280. https://doi.org/10.2113/2020/8876280.
models of coupled fluid-heat-salt transport. Geofluids 7, 377–386. https://doi.org/
10.1111/j.1468-8123.2007.00191.x.

24

You might also like