ArticlePDF Available

Widespread active detachment faulting and core complex formation near 13° N on the Mid-Atlantic Ridge

Authors:

Abstract and Figures

Oceanic core complexes are massifs in which lower-crustal and upper-mantle rocks are exposed at the sea floor. They form at mid-ocean ridges through slip on detachment faults rooted below the spreading axis. To date, most studies of core complexes have been based on isolated inactive massifs that have spread away from ridge axes. Here we present a survey of the Mid-Atlantic Ridge near 13 degrees N containing a segment in which a number of linked detachment faults extend for 75 km along one flank of the spreading axis. The detachment faults are apparently all currently active and at various stages of development. A field of extinct core complexes extends away from the axis for at least 100 km. Our observations reveal the topographic characteristics of actively forming core complexes and their evolution from initiation within the axial valley floor to maturity and eventual inactivity. Within the surrounding region there is a strong correlation between detachment fault morphology at the ridge axis and high rates of hydroacoustically recorded earthquake seismicity. Preliminary examination of seismicity and seafloor morphology farther north along the Mid-Atlantic Ridge suggests that active detachment faulting is occurring in many segments and that detachment faulting is more important in the generation of ocean crust at this slow-spreading ridge than previously suspected.
Content may be subject to copyright.
© 2006 Nature Publishing Group
Widespread active detachment faulting and core
complex formation near 138N on the Mid-Atlantic
Ridge
Deborah K. Smith
1
, Johnson R. Cann
2
& Javier Escartı
´
n
3
Oceanic core complexes are massifs in which lower-crustal and
upper-mantle rocks are exposed at the sea floor
1–3
. They form at
mid-ocean ridges through slip on detachment faults rooted below
the spreading axis
2–6
. To date, most studies of core complexes have
been based on isolated inactive massifs that have spread away from
ridge axes. Here we present a survey of the Mid-Atlantic Ridge
near 138N containing a segment in which a number of linked
detachment faults extend for 75 km along one flank of the
spreading axis. The detachment faults are apparently all currently
active and at various stages of development. A field of extinct core
complexes extends away from the axis for at least 100 km. Our
observations reveal the topographic characteristics of actively
forming core complexes and their evolution from initiation within
the axial valley floor to maturity and eventual inactivity. Within
the surrounding region there is a strong correlation between
detachment fault morphology at the ridge axis and high rates of
hydroacoustically recorded earthquake seismicity. Preliminary
examination of seismicity and seafloor morphology farther
north along the Mid-Atlantic Ridge suggests that active detach-
ment faulting is occurring in many segments and that detachment
faulting is more important in the generation of ocean crust at this
slow-spreading ridge than previously suspected.
The existence of low-angle faults extending deep below the axes of
mid-ocean ridges
7,8
has long been inferred from seafloor exposures of
deep-seated rocks. Direct evidence for such faults has only recently
come from bathymetry
1,2
and seafloor sampling and drilling
4,5,9
. The
faults are corrugated parallel to the spreading direction and cap
smooth topographic highs, termed oceanic core complexes, where
deep-seated rocks such as gabbros and serpentinized peridotites are
exposed
4,5,10,11
. Most core complexes have been identified towards the
ends of spreading segments where magma supply appears to be low,
but in places they extend for tens of kilometres parallel to the
axis
4,5,12–14
. In parts of some segments, extension by low-angle
faulting may have accounted for .50% of the total extension by
spreading
3,11,15
.
Because almost all core complexes identified to date are far enough
from the axis to be inactive, the nature of active detachment faults is
controversial. How do they initiate? How do they evolve as they
emerge from the ocean floor? How is active detachment faulting
accommodated along the length of the spreading axis? Is there a
seismic signature to detachment faulting? We answer these questions
using a new survey of the Mid-Atlantic Ridge (MAR) (D.K.S. & J.E.,
unpublished cruise report, RV Knorr Leg 182, June 2005) near 138N
together with existing multibeam bathymetry and hydroacoustically
recorded seismicity, and extend our conclusions to other mid-ocean-
ridge segments in the region.
A bathymetric map of the MAR between the Fifteen-Twenty and
the Marathon fracture zones (Fig. 1) shows an alternation in the
morphology of the spreading axis. Of the two major segments in this
area, that between 148350Nand138500N (labelled the 148N
segment) has a volcanic signature, with closely spaced volcanic ridges
on both flanks, tens of kilometres long, and cut by steep inward-
facing faults. The segment to the south between 138500N and
128400N (the 138N segment) has more chaotic topography. The
west flank, in particular, is distinctive, with widely spaced, narrow
LETTERS
Figure 1 |Location map and bathymetry near 138N on the MAR. a, Satellite
altimetry data
25
. The red stars indicate seven sections of the axis showing
persistent, high levels of hydroacoustically recorded seismic activity
22
. The
inset at the top shows the MAR axis. b, Multibeam bathymetry from refs 24
and 12 and D.K.S. & J.E., unpublished cruise report (RV Knorr Leg 182, June
2005). The ‘148N segment’ has faulted volcanic morphology, while the
segments to the north and south have irregular and blocky topography and
core complexes. The peridotite-hosted Logatchev hydrothermal vent field
26
is marked. Red dots show locations of 292 hydroacoustically detected events
for the period 1999–2005 with 1jerror bars
22
less than ^10 km.
1
Woods Hole Oceanographic Institution, Woods Hole, Massachusetts 02543, USA.
2
School of Earth Sciences, University of Leeds, Leeds LS2 9JT, UK.
3
CNRS, Marine
Geosciences Group, IPGP Case 89, Paris 75252 cedex 05, France.
Vol 442|27 July 2006|doi:10.1038/nature04950
440
© 2006 Nature Publishing Group
ridges, typically ,20 km long, that have steep slopes dipping away
from the spreading axis, and are separated by areas of smoother sea
floor. As shown below, a large part of this smoother sea floor is
corrugated parallel to the spreading direction.
In well-explored oceanic core complexes, the existence of a
corrugated surface is correlated with detachment faulting and
exposures of deep-seated rocks
1–4,6
. Consequently, corrugations
have been used to identify other core complexes in regions with
limited data
2
. Corrugations have a wavelength of hundreds of metres
and an amplitude of tens of metres
1,3
, and can be seen on shaded relief
maps of multibeam bathymetry data. The origin of the corrugations
is still enigmatic. Direct observations have shown that they are not
produced by faults parallel to the spreading direction
15,16
.One
possibility is that they originate by continuous casting
17
of a ductile
footwall by irregularities in a strong and brittle hanging wall.
An inactive, corrugated core complex near 308N (ref. 1) at the
MAR (Fig. 2) shows other characteristics common to core complexes.
First, a steep (25–308) slope faces away from the axis at the older edge
(scarps at slow- and intermediate-spreading ridges typically face
towards the axis
18
). Second, a narrow, linear ridge parallel to the axis
caps the outward-facing slope
1,2
and extends beyond the corrugated
surface. Dredge samples from the linear ridge in Fig. 2 show that it is
composed of basalt
1
and thus is probably volcanic sea floor created at
the axis. Finally, a steep normal fault dips towards the spreading axis,
truncating the corrugated surface on its younger side.
In the 138N segment, corrugated surfaces and outward facing
steep slopes dominate the western flank of the spreading centre,
extending for 75 km along axis and .100 km across axis (Fig. 3a).
The feature centred at 138100N, 45800 0W (labelled ‘1’ on Fig. 3a) is
typical of many of those observed off-axis, and has a similar
morphology to that shown in Fig. 2. The morphology and spatial
scale of this and other corrugated surfaces and ridges in the 138N
segment indicate that the evolution of the western flank involves
repeated detachment faulting and core complex formation.
Two core complexes north of complex ‘1’ at 13820 0N, 44855 0W
(labelled ‘2’ on Fig. 3a, b) and 13830 0N, 44855 0W (labelled ‘3’) have
a different morphology from the core complexes that lie off-axis.
Each extends ,10 km along-axis and is close to the spreading axis.
The corrugated surfaces are domes that meet ridges to the west, and
curve over to the east, dipping at 158to meet the axial valley floor.
Instead of being truncated by later faults, as with the off-axis core
complexes, the slope at the east end of each of the domes intersects
the median valley floor along a curved line ,5 km from the volcanic
axis (Fig. 3). These features suggest that the domes are the surfaces of
detachment faults actively emerging from the valley floor.
We also identify a larger and apparently compound core complex
extending for 30–40 km parallel to the axis (labelled ‘C’ on Fig. 3a,
stretching from 128550Nto138100N). This compound core com-
plex includes complex ‘1’ and another similar complex to the south
(labelled ‘4’ in Fig. 3a) linked by a third complex located ,5kmto
the east. This third complex terminates in a boundary with the
volcanic morphology of the valley floor that is convex towards the
east. A corrugated surface is present over a large part of the linking
complex.
Because many of the core complexes in this segment have straight
narrow ridges at their outer sides, we consider linear ridges as
possible precursors to the emergence of the detachment faults.
Such ridges occur in several places close to the spreading axis
(labelled ‘R’ on Fig. 3a, b), each backed by a deep basin (Fig. 3).
The outer slopes of the ridges dip at 15–208away from the axis and
show a hummocky volcanic morphology similar to those of the
ridges farther from the spreading axis. Presumably they form by
rotation of sections of volcanic sea floor away from the axis. We infer
that the ridges mark the breakaway zones of detachment faults.
The strong morphologic evidence for active detachment faulting led
us to search for associated seismic activity. Traditionally, seismically
active MAR segments have been identified by locating foci of earth-
quakes of magnitude .4.5 detected by land seismometers, but their
locations are poorly constrained, and these earthquakes are few in
number. Between 1999 and 2005 an autonomous hydrophone array
in the North Atlantic
19
recorded the hydroacoustic energy from
thousands of earthquakesof smaller magnitude (.2.5). Hydroacoustic
events are better located than those detected teleseismically
19–21
,and
clearly show the currently active sections of MAR segments
22
.Two
sections of ridge between the Fifteen-Twenty and Marathon fracture
zones (Fig. 1b) show relatively high and persistent seismicity
23
, and
are separated by the seismically quiet 148N segment. Core complexes
have been identified in the active segment just south of the Fifteen-
Twenty fracture zone
24
. The other seismically active section, the 138N
segment, is active along its entire length, with no obvious spatial or
temporal clustering of events. Location errors (Fig. 1b) are large
enough, though, that events cannot be associated with individual
detachments, nor with possible seismic sources associated with core
complex formation.
The seismic evidence supports the morphologic evidence that a
75-km-long chain of detachment faults on the western side of the
axial valley is active. Some of the detachment faults are mature
(complexes ‘1’ and ‘4’ on Fig. 3a), some are in the early stages of
development (‘2’ and ‘3’), and some are in an incipient stage of
evolution, as shown by the narrow ridges close to the spreading axis
(‘R’). As they evolve and spread away from the axis, the detachments
become inactive and new detachments initiate within a few kilometres
of the axis.
There are a number of core complexes at different stages of
evolution, so we can reconstruct the evolution of a core complex
from initiation to maturity to inactivity (Fig. 4). The first stage is the
subsidence of a basin within the axial valley floor, coupled with the
emergence of a narrow basaltic ridge (Fig. 4a). The basin lies
Figure 2 |Bathymetry map of a well-explored, extinct core complex south
of the Atlantis fracture zone on the MAR
3
.Map contour interval is 100 m.
The complex has spread ,30 km east of the axis. Its morphological
characteristics include an outward-facing slope with a volcanically
constructed ridge capping the core complex on the older (outer) side.
A normal (inward-facing) fault scarp cuts the feature on its younger (inner)
side. Corrugations running parallel to the spreading direction cap the
shallow-dipping top. Serpentinized peridotite was recovered by dredging the
corrugated surface
7
. Mass wasting of the massif is indicated by the scoop on
its northwest corner.
NATURE|Vol 442|27 July 2006 LETTERS
441
© 2006 Nature Publishing Group
immediately behind the ridge, and probably forms by outward
footwall rotation as a new fault initiates. The inner (eastern) side
of the basin (and hence the outer side of the ridge) has the volcanic
morphology of the axial valley floor, tilted at 158–258. As the ridge
evolves, the tilt of the outer slope increases to 208–308. A fault scarp
dipping at 158–258emerges on the inner side of the ridge.
The next stage is the emergence of a domal corrugated fault surface
that intersects the valley floor along a line convex towards the
spreadingaxis(Fig.4b).Thecorrugatedsurfacedipsat,158
where it plunges into the valley floor. The dip of the surface gradually
flattens as it emerges due to flexure of the footwall, and by ,5km
from its emergence the crest is nearly horizontal. Below the median
valley floor, the detachment fault may continue at a low angle, but the
early rotation of the footwall to form the basin suggests that the fault
curves and steepens below the sea floor, so that the tip of the fault may
lie several kilometres down.
The resulting mature core complex (Fig. 4c) becomes extinct when
it is cut off by a normal fault. At that stage the domal surface flattens
until it is close to horizontal. In some places, the core complex may
extend as a single elongate unit from tens of kilometres to over a
hundred kilometres from the spreading axis, as observed in the
Australian–Antarctic Discordance
10,11
and the Parece Vela basin
13
.In
our study area, the small spacing between breakaway faults indicates
a short life for any individual complex and regular nucleation of new
detachment faults in the median valley.
Because of the apparent correlation of persistent earthquake
seismicity and core complex morphology, we examined a larger
region of the MAR (248–158N) to look for additional sites of active
core complex formation. We identified five ridge sections with high
levels of hydroacoustically recorded seismic activity
22
(Fig. 1a). The
seismicity is persistent and not triggered by large earthquakes.
Preliminary examination of available bathymetry from these and
other segments in this larger area indicates a correlation between core
complex morphology and seismicity. We suggest that detachment
faulting may be more common than previously suspected in this part
of the MAR. As much as 35% of the spreading axis may be
experiencing detachment faulting and thus, .15% of the new
seafloor accretion may be dominated by core complexes. We also
suggest that the evolution of core complexes we have identified in our
study area and the associated seismicity may be applicable to under-
standing other regions of active detachment faulting both in the
Figure 3 |Multibeam bathymetry showing detachment faults and core
complexes in the 138N segment. a, Map contour interval is 50 m. Linked
and active core complexes extend ,75km along the axis. The dashed line
indicates the spreading axis. R, topographic ridges, inferred to be
breakaways for new detachment surfaces. Numbers, complexes discussed in
the text. C, compound core complex composed of complexes ‘1’ and ‘4’ and a
third linking complex. b, Three-dimensional perspective view of complex ‘2’
with no vertical exaggeration. The corrugated surface and its intersection
with the sea floor are marked.
Figure 4 |Schematic three-dimensional oblique view of the evolution of a
core complex by detachment faulting. Thick grey lines, sea floor (except for
the detachment surface). Thin black lines, surface of the detachment fault
and steeper normal faults, both exposed on the sea floor and below it. Thick
dashed line, spreading axis. a, A breakaway ridge with a basin behind it. The
basin is floored by volcanic sea floor, tilted up towards the breakaway ridge.
The initially steep subsurface normal fault at the breakaway has already
rotated to a shallower angle. A downward-curved detachment has started to
form linked to the breakaway fault. Both faults are below the sea floor. b, The
detachment has emerged. The thin dashed line marks the line of emergence.
The fault has warped into a dome, the ridge has become arched, and little
further rotation of the initial ridge has occurred. c, The detachment fault has
been cut off by a later normal fault, perhaps of the breakaway of the next
detachment. The dome has flattened, and the detachment fault is inactive.
LETTERS NATURE|Vol 442|27 July 2006
442
© 2006 Nature Publishing Group
oceans and on land, where the faults are more accessible but where
erosion severely hinders their interpretation.
Received 9 April; accepted 2 June 2006.
1. Cann, J. R. et al. Corrugated slip surfaces formed at ridge–-transform
intersections on the Mid-Atlantic Ridge. Nature 385, 329–-332 (1997).
2. Tucholke, B. E., Lin, J. & Kleinrock, M. C. Megamullions and mullion structure
defining oceanic metamorphic core complexes on the Mid-Atlantic Ridge.
J. Geophys. Res. 103, 9857–-9866 (1998).
3. Blackman, D. K., Cann, J. R., Janssen, B. & Smith, D. K. Origin of extensional
core complexes: Evidence from the Mid-Atlantic Ridge at Atlantis Fracture
Zone. J. Geophys. Res. 103, 21315–-21333 (1998).
4. MacLeod, C. J. et al. Direct geological evidence for oceanic detachment
faulting: The Mid-Atlantic Ridge, 15845 0N. Geology 30, 279–-282 (2002).
5. Escartı
´n, J., Me
´vel, C., MacLeod, C. J. & McCaig, A. M. Constraints on
deformation conditions and the origin of oceanic detachments: The Mid-
Atlantic Ridge core complex at 15845 0N. Geochem. Geophys. Geosyst. 4, 1067,
doi:10.1029/2002GC000472 (2003).
6. Searle, R. C. et al. FUJI Dome: A large detachment fault near 64 degrees E on
the very slow-spreading Southwest Indian Ridge. Geochem. Geophys. Geosyst. 4,
doi:10.1029/2003GC000519 (2003).
7. Dick, H. J. B., Thompson, W. B. & Bryan, W. B. Low angle faulting and steady-
state emplacement of plutonic rocks at ridge–-transform intersections. Eos 62,
abstr. 406 (1981).
8. Karson, J. A. & Dick, H. J. B. Tectonics of ridge–-transform intersections at the
Kane Fracture Zone. Mar. Geophys. Res. 6, 51–-98 (1983).
9. Schroeder, T. & John, B. E. Strain localization on an oceanic detachment fault
system, Atlantis Massif, 30 degrees N, Mid-Atlantic Ridge. Geochem. Geophys.
Geosyst. 5, 1–-30 (2004).
10. Christie, D. M., West, B. P., Pyle, D. G. & Hanan, B. B. Chaotic topography,
mantle flow and mantle migration in the Australian–-Antarctic Discordance.
Nature 394, 637–-644 (1998).
11. Okino, K., Matsuda, K., Christie, D., Nogi, Y. & Koizumi, K. Development of
oceanic detachment and asymmetric spreading at the Australian–-Antarctic
Discordance. Geochem. Geophys. Geosyst. 5, doi:10.1029/2004GC000793
(2004).
12. Escartı
´n, J. & Cannat, M. Ultramafic exposures and the gravity signature of the
lithosphere near the Fifteen-Twenty Fracture Zone (Mid-Atlantic Ridge,
148-16.58N). Earth Planet. Sci. Lett. 171, 411–-424 (1999).
13. Ohara, Y., Yoshida, T. & Kasuga, S. Giant megamullion in the Perece Vela
Backarc basin. Mar. Geophys. Res. 22, 47–-61 (2001).
14. Cannat, M. et al. Thin crust, ultramafic exposures, and rugged faulting patterns
at the Mid-Atlantic Ridge (228–- 2 4 8N). Geology 23, 49–-52 (1995).
15. Blackman, D. K. et al. Geology of the Atlantis Massif (Mid-Atlantic Ridge,
308N): Implications for the Evolution of an Ultramafic Oceanic Core Complex.
Mar. Geophys. Res. 23, 443–-460 (2003).
16. Tucholke, B. E., Fujioka, K., Ishihara, T., Hirth, G. & Kinoshita, M. Submersible
study of an oceanic megamullion in the central North Atlantic. J. Geophys. Res.
106, 16145–-16161 (2001).
17. Spencer, J. E. Geologic continuous casting below continental and deep-sea
detachment faults and at the striated extrusion of Sacsayhuaman, Peru.
Geology 27, 327–-330 (1999).
18. Macdonald, K. C. Mid-ocean ridges: Fine scale tectonic, volcanic, and
hydrothermal processes within the plate boundary zone. Annu. Rev. Earth
Planet. Sci. 10, 155–-190 (1982).
19. Smith, D. K. et al. Hydroacoustic monitoring of seismicity at the slow-
spreading Mid-Atlantic Ridge. Geophys. Res. Lett. 29, doi:10.1029/
2001GL013912 (2002).
20. Bohnenstiehl, D. R., Tolstoy, M., Dziak, R. P., Fox, C. G. & Smith, D. K.
Aftershock sequences in the mid-ocean ridge environment: An analysis using
hydroacoustic data. Tectonophysics 354, 49–-70 (2002).
21. Fox, C. G., Matsumoto, H. & Lau, T.-K. Monitoring Pacific ocean seismicity from
an autonomous hydrophone array. J. Geophys. Res. 106, 4183–-4206 (2001).
22. Smith, D. K. et al. Spatial and temporal distribution of seismicity along the
northern Mid-Atlantic Ridge (158–- 3 5 8N). J. Geophys. Res. 108, doi:10.1029/
2002JB001964 (2003).
23. Escartin, J., Smith, D. K. & Cannat, M. Parallel bands of seismicity at the
Mid-Atlantic Ridge, 12–-14N. Geophys. Res. Lett. 30, doi:10.1029/
2003GL017226 (2003).
24. Fujiwara, T. et al. Crustal evolution of the Mid-Atlantic Ridge near the Fifteen-
Twenty Fracture Zone in the last 5 Ma. Geochem. Geophys. Geosyst. 4, 1024,
doi:10.1029/2002GC000364 (2003).
25. Smith, W. H. F. & Sandwell, D. T. Global sea floor topography from satellite
altimetry and ship depth soundings. Science 277, 1956–-1962 (1997).
26. Sudarikov, S. M. & Roumiantsev, A. B. Structure of hydrothermal plumes at the
Logatchev vent field, 14 45 0N, Mid-Atlantic Ridge: evidence from geochemical
and geophysical data. J. Vol. Geotherm. Res. 101, 245–-252 (2000).
Acknowledgements We are grateful to the captain and crew of the RV Knorr
(Leg 182, 2005). We are also grateful to P. Lemmond, H. Schouten and M. Tivey
for their help in collecting the data at 138N. We had fruitful discussions with
R. Searle, H. Schouten, M. Tivey and R. Sohn. We also thank J. Goff for
constructive comments on the manuscript. This work was supported by the
National Science Foundation. This is an IPGP contribution.
Author Information Reprints and permissions information is available at
npg.nature.com/reprintsandpermissions. The authors declare no competing
financial interests. Correspondence and requests for materials should be
addressed to D.K.S. (dsmith@whoi.edu).
NATURE|Vol 442|27 July 2006 LETTERS
443
... High-temperature hydrothermal deposits on mid-ocean ridges (MORs) are typically composed primarily of metal-and sulfide-rich minerals, often forming seafloor massive sulfide (SMS) deposits that are analogous to ancient volcanogenic massive sulfide (VMS) deposits that occur on land (Hannington et al., 2005;Jamieson et al., 2015). At slow-spreading ridges, hydrothermal fluids can interact with deep layers of the oceanic crust through crustal-scale faults that exhume lower crust and/or upper mantle material, such as oceanic core complexes or detachment faults (Escartín et al., 2008;MacLeod et al., 2009;Smith et al., 2006). In these structural settings, the interaction of hydrothermal fluid with lower-crustal rocks typically results in hydrothermal deposits at the seafloor that are characterized by relatively high Au, Cu, Zn, Co, and Ni concentrations . ...
Article
Full-text available
The recently discovered Fåvne vent field, located at 3,040 m depth on the slow‐spreading Mohns mid‐ocean ridge between Greenland and Norway, is a high‐temperature (≥250°C) vent field that is characterized by Fe oxyhydroxide‐rich and S‐poor chimneys and mounds. The vent field is located on both the hanging wall and footwall of a normal fault with a ∼1.5 km throw that forms the western edge of the ∼20 km wide ridge axial valley. Data collected during exploration of the site using a remotely operated vehicle as well as mineralogical and geochemical analyses of rock samples and sediments are used to characterize the geological setting of the vent field and composition of the hydrothermal deposits. The chimney walls are highly porous and lack defined chalcopyrite lined conduits, typical of high‐temperature chimneys. Overall, abundant Fe oxyhydroxide precipitation at high‐temperature vents at Fåvne reflects an excess of Fe over reduced S in the fluid, leading to precipitation of Fe oxide and oxyhydroxide minerals at high to moderate temperature vents (>100°C), and as microbially mediated and abiotic precipitation of Fe oxyhydroxide minerals at low‐temperature diffuse vents (<100°C). The mounds and chimneys exhibit low base metal and reduced S concentrations relative to globally averaged seafloor deposits and suggest subseafloor mixing of hydrothermal fluid with seawater, causing metal sulfide precipitation. Cobalt enrichment at Fåvne may reflect a subsurface influence of an ultramafic substrate on circulating fluids, although ultramafic rocks are absent on the seafloor and no other elements typical of ultramafic deposits are present.
... Hence at slow-spreading MOR segments where magma supply is thought to be relatively low such as in the southern MAR, plate spreading is more likely to be accommodated by slip on short-lived normal faults, which excite earthquakes that are teleseismically-detectable owing to magnitudes exceeding ∼3.5. The spatial pattern of earthquakes along the slow-spreading MAR has been used to infer segment-scale changes in the mode of plate spreading (Escartin et al., 2008;Smith et al., 2006), and also to identify thermal anomalies associated with the Iceland mantle plume (Parnell-Turner et al., 2013). ...
Article
Full-text available
Plain Language Summary Mantle plumes are upwellings of hot material that are thought to originate deep inside Earth's mantle, reaching the base of the tectonic plates and creating oceanic islands like Hawaiian‐Emperor seamount chain. Although these seamount chains help us understand processes within the mantle, the precise location of their associated mantle plumes is surprisingly poorly known. Several prominent seamount chains are located in the south Atlantic ocean, however severe sea conditions mean that they have only been sparsely sampled and mapped, and the location of the Shona, Discovery, and Bouvet mantle plumes have been debated since the 1970s. Prominent gaps in records of earthquake seismicity suggest that hot material from the Shona and Discovery plumes is reaching the Mid‐Atlantic ridge today, inhibiting slip on faults. Associated slow mantle wave speeds, elevated seafloor depths, and the composition of basalts dredged from the seafloor are all consistent with the idea that thermal anomalies associated with the Discovery and Shona plumes are beneath the ridge axis today. The Shona plume is likely to exert significant control on how new crust forms in the southern Atlantic Ocean, and could explain the formation of Bouvet Island, which could be associated with a branch of the main Shona hotspot.
... Among them, the youngest ridge-shaped OCC gradually becomes shallower, reaching a depth of 2600 m, and is backed by a topographic low . This geometric feature is similar to that of core complexes found in the 13 • 20 N region of the Mid-Atlantic Ridge (MAR) (Smith et al., 2006;Escartin et al., 2017). ...
Article
Full-text available
Various crustal processes shape both the lower oceanic crust and mid-ocean ridge basalts (MORBs). To better understand how these crustal processes influence MORB compositions, we conducted comprehensive petrographic and geochemical investigations on gabbroic rocks and erupted lavas dredged from a segment of the Central Indian Ridge (CIR) spanning from 7°50'S to 8°30'S. The petrographic and geochemical analyses of the gabbroic rocks revealed evidence of melt-rock reaction through reactive porous flow (RPF) in olivine gabbro and gabbro. This process resulted in distinctive features in clinopyroxene, including disequilibrium textures with a troctolite/anorthosite matrix, complex variations in Mg#-Cr-Ti [Mg#=molar Mg/(Mg+Fe2+)] relationships, and considerable enrichment and fractionation of incompatible trace elements. A significant finding of our study is the close resemblance of trace element ratios in MORB and olivine-hosted melt inclusions to those of melts in equilibrium with clinopyroxene from olivine gabbro and gabbro with Sr anomaly (Sr/Sr* = SrN/sqrt[PrN*NdN]; N refers to chondrite-normalized values) greater than ~0.7. This observation strongly indicates that the composition of MORB is influenced by the melt-rock reaction taking place in the lower oceanic crust. Furthermore, our findings suggest that evolved melts in equilibrium with clinopyroxene having Sr/Sr* values lower than ~0.7 are less likely to erupt onto the seafloor and are instead trapped within the lower oceanic crust. Oxide gabbronorite is characterized by coarse-granular, pegmatitic textures and exhibits mineralogically and chemically more evolved characteristics compared to olivine gabbro and gabbro. It is inferred that the oxide gabbronorite formed through the in-situ freezing of highly evolved melts within a melt-rich layer. Finally, we present a comprehensive model for melt evolution in the lower oceanic crust at the 7°50'S-8°30'S CIR by integrating all petrological and geochemical data obtained from gabbroic rocks, MORB, and olivine-hosted melt inclusions. This holistic model contributes to a better understanding of the intricate processes governing MORB composition in the context of the lower oceanic crust dynamics at slow-spreading ridges.
... While the complete exhumation history of the Schulz Massif remains unknown, the present-day fault dip above the sedimentary basin is ∼13°-16°, with an estimated dip below the surface of around 20°-25° (Bruvoll et al., 2009). However, it is possible that the initial fault was much steeper (∼60°-70°), as is observed in other oceanic core complexes (de Martin et al., 2007;Smith et al., 2006 Based on the calculations, assuming crystallization occurred in the central part of the axial volcanic ridge in the present-day rift valley (scenario 1) at a depth of 4.5 km, the 206 Pb/ 238 U dates (4.6 Ma) suggest a half spreading rate of 7.4 mm/year, slightly lower than the 7.8 mm/year estimated for the last 10.3 Ma from regional models and seafloor magnetic anomalies (Mosar et al., 2002;Müller et al., 2008), but consistent with spreading rates of 7.4 mm/year for the last 1.3 Ma derived from sedimentary stratigraphy (Bruvoll et al., 2009). ...
Article
Full-text available
The ultraslow‐spreading Mohns Ridge is a key supersegment of the Arctic Mid‐Ocean system, where it represents a boundary between the Jan Mayen hotspot in the south and the highly anomalous Knipovich Ridge to the north. Understanding the timing and mode of Plio‐Pleistocene seafloor spreading along this ridge segment is critical for establishing the recent geodynamic evolution of the Norwegian‐Greenland Sea. To investigate magmatic accretion at this ultraslow spreading ridge, we collected samples from the Schulz Massif, which is located off‐axis at 73.4°N and exposes gabbroic intrusives and mantle peridotite. The petrology and petrography of these samples indicate that the exposed crustal section underwent multiple episodes of magmatism, which are characterized by distinct crystal sizes and geochemistry. To calibrate the age of the seafloor, we combined high‐resolution and high‐precision single zircon U‐Pb geochronology. Our data suggest that seafloor spreading has been nearly symmetrical for the last ∼4.6 Myr with a time‐averaged half‐spreading rate of ∼7.4 mm yr⁻¹. Crystal size analysis of olivine in porphyric intrusions suggests that the crustal section was fed crystal‐laden melts with recurrence rates predicted to stabilize fault‐dominated seafloor spreading. Our combined geochronological and crystal size approach gives a critical perspective on the mode of seafloor spreading in the Mohns Ridge and allows insights into accretionary mechanisms and crustal structures during symmetric seafloor spreading.
... During plate separation, the breakaway-the tip of the footwall, where the fault first cuts the seafloor-migrates off-axis (Buck et al., 2005). With increasing displacement on the fault plane, the uplifted footwall starts rotating and bending under its own weight (MacLeod et al., 2009;Smith et al., 2006). The domed corrugated shear planes of such detachments, known as oceanic core complexes, emerge to the seafloor at shallow angles where they expose lower crust and mantle rocks (e.g., Escartín et al., 2008Escartín et al., , 2017Hayman et al., 2011;Zhao et al., 2013). ...
Article
Full-text available
Flip‐flop” detachment mode represents an endmember type of lithosphere‐scale faulting observed at almost amagmatic sections of ultraslow‐spreading mid‐ocean ridges. Recent numerical experiments using an imposed steady temperature structure show that an axial temperature maximum is essential to trigger flip‐flop faults by focusing flexural strain in the footwall of the active fault. However, ridge segments without significant melt budget are more likely to be in a transient thermal state controlled, at least partly, by the faulting dynamics themselves. Therefore, we investigate which processes control the thermal structure of the lithosphere and how feedbacks with the deformation mechanisms can explain observed faulting patterns. We present results of 2‐D thermo‐mechanical numerical modeling including serpentinization reactions and dynamic grain size evolution. The model features a novel form of parametrized hydrothermal cooling along fault zones as well as the thermal and rheological effects of periodic sill intrusions. We find that the interplay of hydrothermal fault zone cooling and periodic sill intrusions in the footwall facilitates the flip‐flop detachment mode. Hydrothermal cooling of the fault zone pushes the temperature maximum into the footwall, while intrusions near the temperature maximum further weaken the rock and promote the formation of new faults with opposite polarity. Our model allows us to put constraints on the magnitude of two processes, and we obtain most reasonable melt budgets and hydrothermal heat fluxes if both are considered. Furthermore, we frequently observe two other faulting modes in our experiments complementing flip‐flop faulting to yield a potentially more robust alternative interpretation for existing observations.
... This causes outward slopes near fault breakaways to steepen, and the area of these slopes eventually rivals and may even exceed the area of inward-facing fault slopes. The interpretation that the outward slopes are backtilted sides of fault blocks, rather than outward-facing faults, is supported by analysis of multibeam bathymetric data and HMR1 long-range sidescan sonar images on the MAR (Cann et al., 2015;Smith et al., 2006Smith et al., , 2008Zheng et al., 2019); these data show that the vast majority of outward slopes are not fault surfaces but are covered by irregular volcanic morphology and volcanic cones. ...
Article
Full-text available
Magma supply likely exerts primary control on seafloor morphology of oceanic crust, but most studies have related morphology to spreading rate. Here we examine global patterns of morphology on mid‐ocean ridge (MOR) flanks in relation to magma supply derived from residual mantle Bouguer gravity anomaly (proxy for relative crustal thickness) and spreading rate. We use multibeam bathymetry to characterize morphology using both qualitative (descriptive) and quantitative approaches, and we compare results to both magma supply and spreading rate. Morphology becomes more isotropic and abyssal hills are more irregular and discontinuous as magma supply decreases, while roughness, area of steeper slopes, and anomalous fabric orientation increase. We interpret these changes to reflect changing magma distribution along‐axis, from large‐volume and spatially extensive to progressively reduced, increasingly localized, and more irregularly emplaced. Observed relations between crustal thickness and morphology imply that average thickness of purely magmatic crust in the Atlantic and parts of the Indian ridge system is significantly less than average seismically determined crust. Thus seismically defined crustal thickness in those regions likely includes significant non‐magmatic components such as serpentinized mantle. Excepting regions of extensive mantle exposure, most morphologic parameters that we examined are sensitive to estimated magma supply but not necessarily to spreading rate alone. We summarize our results in schematic models that relate morphologic variations to changes in magma supply and mantle serpentinization throughout the global MOR system. Finally, we note that combined qualitative and quantitative results of our study may be useful for developing automated morphologic classification schemes.
... Fault strength in this situation is likely reduced by serpentinization. Bathymetry data acquired by autonomous underwater vehicle of the DF2 clearly shows that the footwall is rotated to a low angle (<20°, Zhu et al., 2010), which is a typical angle of the footwall of detachment faults (Escartin et al., 2017; Smith et al., 2006Smith et al., , 2008. Remotely operated Zhao et al. (2013). ...
Article
Full-text available
Oceanic detachment faulting, a major mode of seafloor accretion at slow and ultraslow spreading ridges, is thought to occur during magma‐poor phases and be abandoned when magmatism increases. In this framework, detachment faulting is the result of temporal variations in magma flux, which is inconsistent with recent geophysical observations at the Longqi segment on the Southwest Indian Ridge (49°42′E). In this paper, we focus on this sequentially active detachment faulting system that includes an old, inactive detachment fault and a younger, active detachment fault. We investigate the mechanisms controlling the temporal evolution of this tectonomagmatic system by using 2D mid‐ocean ridge spreading models that simulate faulting and magma intrusion into a visco‐elasto‐plastic continuum. Our models show that temporal variations in magma flux alone are insufficient to match the inferred temporal evolution of the sequentially active detachment system. Rather we find that sequentially active detachment faulting spontaneously occurs at the Longqi segment as a function of lithospheric thickness. This finding is in agreement with an analytical model, which shows that a thicker axial lithosphere results in a smaller fault heave and that a flatter angle in lithosphere thickening away from the accretion axis stabilizes the active fault. A thicker axial lithosphere and its flatter off‐axis angle combined have the potential to modulate sequentially active detachment faulting at the Longqi segment. Our results thus suggest that temporal changes of magmatism are not necessary for the development and abandonment of detachment faults at ultraslow spreading ridges.
Article
During the 41th scientific cruise of the R/V “Professor Logachev” in 2019, two new ore fields were discovered within the Russian exploration area of the Mid-Atlantic Ridge. The “Korallovoye” ore field is located in the middle part of the rift valley ridge and lies on a gentle slope in the depth range of 2800–2850 m. The “Molodezhnoye” ore field is located at a distance of approximately 5 km northeast of the “Korallovoye” field and lies on a less steeper section of the slope between 3500 and 3550 m in depth. The “Korallovoye” and “Molodezhnoye” fields are confined to the outcrop of a gabbro-peridotite massif on the western side of the Mid-Atlantic Rift valley. The total number of discovered ore fields in the Russian exploration area has been increased to 14 sites.
Article
Full-text available
The morphology of the sea floor at mid-ocean-ridge spreading centres provides a key to understanding how ocean crust is constructed. Images of the axial zone of the slow-spreading Mid-Atlantic Ridge, obtained at a range of spatial scales, show that crustal construction is complex and highly variable, reflecting an underlying three-dimensional variability in magmatic, tectonic and hydrothermal processes.
Article
Full-text available
pdf , The following values have no corresponding Zotero field: ID - 513
Article
Full-text available
[1] In February 1999, long-term hydroacoustic monitoring of the northern Mid-Atlantic Ridge (MAR) was initiated. Six autonomous hydrophones were moored between ∼15°N and ∼35°N on the flanks of the MAR. Results from the first year of data reveal that there is significant variability in along-axis event rate. Groups of neighboring segments behave similarly, producing an along-axis pattern with high and low levels of seismic activity at a wavelength of ∼500 km. This broad scale pattern is likely influenced by the axial thermal regime. Several earthquake sequences with variable temporal characteristics were detected, suggesting fundamental differences in the cause of their seismicity. Off-axis, most seismic faulting occurs within a zone < 15 km from the axis center.
Article
Full-text available
Since May 1996, an array of autonomous hydrophone moorings has been continuously deployed in the eastern equatorial Pacific to provide long-term monitoring of seismic activity, including low-level volcanic signals, along the East Pacific Rise between 20°N and 20°S and the Galapagos Ridge. The instruments and moorings were designed to continuously record low-frequency acoustic energy in the SOFAR channel for extended periods and produce results comparable to those previously derived by using the U.S. Navy Sound Surveillance System (SOSUS) in the northeast Pacific. The technology and methodology developed for this experiment, including instrument design, mooring configuration, analysis software, location algorithms (with an analysis of errors), and a predicted error field, are described in detail. Volcanic activity is observed throughout the Pacific, along with seismicity along transform faults, subduction zones, and intraplate regions. Comparison data sets indicate detection thresholds and accuracy better than the land networks for open ocean areas and results comparable to, or better than, SOSUS. Volcanic seismicity along the fast spreading East Pacific Rise appears similar to documented examples in the northeast Pacific but with much shorter durations. One example from the intermediate spreading Galapagos Ridge is comparable to northeast Pacific examples, and several episodes of activity were observed in the Wilkes Transform Fault Zone. A site of continuing off-axis seismicity is located near 18°S and 116°W. Isolated intraplate earthquakes are observed throughout the study area. Earthquake information from this experiment and future observations will be provided through the World Wide Web and earthquake data centers.
Article
The volcanic morphology of a number of segments of the slow spreading Mid-Atlantic Ridge (MAR) have been reinterpreted based on our understanding of dike emplacement, dike propagation, and eruption at the East Rift Zone of Kilauea Volcano, Hawaii and its submarine extension, the Puna Ridge. The styles of volcanic eruption at the submarine Puna Ridge are remarkably similar to those of the axial volcanic ridges (AVRs) constructed on the median valley floor of the MAR. We use this observation to relate volcanic processes occurring at Kilauea Volcano to the MAR. We now consider that volcanic features (e.g., seamounts and lava terraces) built on the flanks of the AVRs are secondary features that are fed from lava tubes or channels, not primary features fed directly from an underlying dike. We examine simple models of pipe flow and conclude that lava tubes can transport lava down the flanks of submarine rifts to build all of the volcanic features observed there. In addition, deep water lava tubes are strong enough to withstand the pressures of a few megapascals that the building of a volcanic structure 150 m high at the end of the tube would generate. The volumes of individual volcanic terraces and seamounts on the Puna Ridge and at the MAR are large (0.1-1 km3) and similar to the volumes of lava flows that are broadly distributed at the subaerial East Rift Zone of Kilauea. This striking difference in the volcanic morphology on a scale of 1-2 km (producing terraces and seamounts underwater and low-relief flows on land) must be related to the enhanced cooling and to the greater mechanical stability of tubes in the submarine environment. We suggest that at the MAR a crustal magma reservoir, most likely located beneath shallow, flat sections of the segment, provides magma to the rift axis through dikes that propagate laterally tens of kilometers. The zone of dike intrusion, at least in the neighborhood of the magma body, is likely narrower than the width resurfaced by flows, yielding a crustal structure that has a rapid vertical transition from lavas to sheeted dikes. At segment ends the zone of dike intrusion is likely to be wider, giving a resulting structure with a more gradual transition from lavas to dikes.
Article
Multibeam bathymetry data obtained along a 2400 km long section of the Mid-Atlantic Ridge (MAR) from 15°N to the Azores platform (40°N) and satellite-derived gravity data were used to calculate the mantle Bouguer anomaly (MBA) along this portion of the MAR. Both data sets were used to determine the relations between gravity anomalies and topographic variations and discuss these in terms of thermal difference. A long-wavelength influence of the Azores hot spot is characterized by a gentle, continuous slope of the average ridge axial depth and a general gradient in the along-axis MBA profile. This thermal influence of the Azores hot spot controls a systematic southward propagation of the spreading segments at least to 26°30'N. South of 26°30'N, the direction of the segment propagation is controlled by the local difference in thermal state between adjacent segments. Except on the Azores platform, the systematic along-axis 11-90 km long wavelength segmentation is independent of the long-wavelength influence of the Azores. At the segment center, the axial morphology is linked to the thermal state of the segments between: (1) "Hotter segments" characterized by a smooth axial morphology, a well-defined shallow "inner valley", high ΔMBA and a long length; (2) "colder segments" which present a rough axial morphology with a deep, wide and well-defined rift valley, a low ΔMBA and a small length. For "hotter segments" the formation of the abyssal hills is mainly due to a magmato-tectonic cycle over periods of 0.3 to 1 Myr, whereas on "colder segments" the axial morphology is mainly controlled by a tectonic rift valley formation. We propose that these different segment types correspond to a temporal evolution of the rift valley morphology over periods of several million years.
Article
The Omori formula n(t)=K(t+c)-1 and its modified form n(t)=K(t+c)-P have been successfully applied to many aftershock sequences since the former was proposed just 100 years ago. This paper summarizes studies using these formulae. The problems of fitting these formulae and related point process models to observational data are discussed mainly. Studies published during the last 1/3 century confirmed that the modified Omori formula generally provides an appropriate representation of the temporal variation of aftershock activity. Although no systematic dependence of the index p has been found on the magnitude of the main shock and on the lowest limit of magnitude above which aftershocks are counted, this index (usually p = 0.9-1.5) differs from sequence to. sequence. This variability may be related to the tectonic condition of the region such as structural heterogeneity, stress, and temperature, but it is not clear which factor is most significant in controlling the p value. The constant c is a controversial quantity. It is strongly influenced by incomplete detection of small aftershocks in the early stage of sequence. Careful analyses indicate that c is positive at least for some sequences. Point process models for the temporal pattern of shallow seismicity must include the existence of aftershocks, most suitably expressed by the modified Omori law. Among such models, the ETAS model seems to best represent the main features of seismicity with only five parameters. An anomalous decrease in aftershock activity below the level predicted by the modified Omori formula sometimes precedes a large aftershock. An anomalous decrease in seismic activity of a region below the level predicted by the ETAS model is sometimes followed by a large earthquake in the same or in a neighboring region.
Article
From a detailed survey and sampling study of corrugated massifs north of the Fifteen- Twenty Fracture Zone on the Mid-Atlantic Ridge, we demonstrate that their surfaces are low-angle detachment fault planes, as proposed but not previously verified. Spreadingdirection– parallel striations on the massifs occur at wavelengths from kilometers to centimeters. Oriented drill-core samples from the striated surfaces are dominated by fault rocks with low-angle shear planes and highly deformed greenschist facies assemblages that include talc, chlorite, tremolite, and serpentine. Deformation was very localized and occurred in the brittle regime; no evidence is seen for ductile deformation of the footwall. Synkinematic emplacement of diabase dikes into the fault zone from an immediately subjacent gabbro pluton implies that the detachment must have been active as a low-angle fault surface at very shallow levels directly beneath the ridge axis. Strain localization occurred in response to the weakening of a range of hydrous secondary minerals at a very early stage and was highly efficient.
Article
THE Mid-Atlantic Ridge (MAR) extends for over 12,000 km from Iceland to the Bouvet Triple Junction in the south Atlantic. Most of our knowledge of the geology of this slow-spreading centre comes either from widely spaced low-resolution echo-sounding lines or from detailed studies of small areas using deep-towed instruments or submarines. High-resolution multibeam bathy-metry1,2 has been insufficient to define the variability of accretionary processes along this plate boundary. Here we present the results of a recent investigation of a ~800-km-long section of the MAR carried out using the multibeam echo-sounder SeaBeam3. Analysis of these data reveals the scale and nature of the segmentation of the MAR between the Kane Fracture Zone (24° N) and latitude 30°40' N.