ArticlePDF Available

Serpentinization pulse in the actively deforming Central Indian Basin

Authors:

Abstract and Figures

Heat flow in the actively deforming Central Indian Basin is on average 30 mW/m2 higher than the theoretical 55 mW/m2 heat flow expected from plate cooling of a Cretaceous oceanic lithosphere. Strong spatial correlation between the anomaly and the active thrust fault network at local (faults) and regional scales suggests two potential tectonically driven mechanisms activated at the time of initiation of deformation: friction-to-heat conversion or exothermic serpentinization. We quantitatively examine both processes using an updated geometry of the thrust fault network and simple thermal models. Friction generated heat is limited in all cases: at shallow levels, shear stresses remain small, while heat generated at deeper levels does not contribute significantly to the surface heat flow since permanent regime is not reached. In the exothermic serpentinization model, a maximum anomaly of 20 to 30 mW/m2 is reached 2 to 6 Myr after the onset of widespread serpentinization, depending on the efficiency of the water circulation. The amount and timing of heat release can fully explain the present-day surface heat flow of the Central Indian Basin, provided vigorous hydrothermal circulation closely followed the onset of deformation. Based on a reprocessed multichannel seismic line, we suggest that faults cutting through the entire crust and across the Moho discontinuity drive water at mantle levels and trigger the exothermic serpentinization reaction. We interpret sub-Moho reflectors imaged at depths of 8 to 15 km below the top of the crust – and coinciding with the location of the maximum reaction rate coefficient of serpentinization – as serpentinization fronts. We discuss the significance of this pulse of serpentinization in terms of timing of deformation, weakening and transient rheology of the oceanic lithosphere.
Content may be subject to copyright.
Serpentinization pulse in the actively deforming Central Indian Basin
Matthias Delescluse
a,c,
, Nicolas Chamot-Rooke
a,b
a
Ecole normale supérieure, Laboratoire de Géologie, 24 rue Lhomond, 75005, Paris, France
b
CNRS, UMR 8538, France
c
Université Paris XI, Orsay, France
abstractarticle info
Article history:
Received 27 November 2007
Received in revised form 8 September 2008
Accepted 14 September 2008
Editor: C.P. Jaupart
PACS:
91.45.G
91.50.Ln
91.55.Jk
91.55.Ln
91.60.Hg
93.30.Nk
Keywords:
heat ow
oceanic lithosphere
deformation
active faults
friction
serpentinization
multichannel seismic
Heat ow in the actively deforming Central Indian Basin is on average 30 mW/m
2
higher than the theoretical
55 mW/m
2
heat ow expected from plate cooling of a Cretaceous oceanic lithosphere. Strong spatial
correlation between the anomaly and the active thrust fault network at local (faults) and regional scales
suggests two potential tectonically driven mechanisms activated at the time of initiation of deformation:
friction-to-heat conversion or exothermic serpentinization. We quantitatively examine both processes using
an updated geometry of the thrust fault network and simple thermal models. Friction generated heat is
limited in all cases: at shallow levels, shear stresses remain small, while heat generated at deeper levels does
not contribute signicantly to the surface heat ow since permanent regime is not reached. In the exothermic
serpentinization model, a maximum anomaly of 20 to 30 mW/m
2
is reached 2 to 6 Myr after the onset of
widespread serpentinization, depending on the efciency of the water circulation. The amount and timing of
heat release can fully explain the present-day surface heat ow of the Central Indian Basin, provided vigorous
hydrothermal circulation closely followed the onset of deformation. Based on a reprocessed multichannel
seismic line, we suggest that faults cutting through the entire crust and across the Moho discontinuity drive
water at mantle levels and trigger the exothermic serpentinization reaction. We interpret sub-Moho
reectors imaged at depths of 8 to 15 km below the top of the crust and coinciding with the location of the
maximum reaction rate coefcient of serpentinization as serpentinization fronts. We discuss the
signicance of this pulse of serpentinization in terms of timing of deformation, weakening and transient
rheology of the oceanic lithosphere.
© 2008 Elsevier B.V. All rights reserved.
1. Introduction
Anomalously high heat-ow measurements have been acquired in
the northeastern Indian Ocean fordecades (Pollack et al., 1993)andwere
promptly related to the abnormal level of intraplate seismicity recorded
between the India and Australia tectonic plates (Weis sel et al., 1980;
Wiens et al.,1986; Stein et al., 1988). The heat-ow anomaly was shown
to coincide with an area of widespread intraplate deformation that
started in the late Miocene (Cochran et al., 1989). Seismic proles (Bull
and Scrutton, 1992; Chamot-Rooke et al.,1993; VanOrman et al.,1995)
further imaged the highly faulted sediments, crust and mantle of the
Central Indian Ocean, conrming the spatial correlation between high
heat-ow and active deformation. The thermal anomaly was thus seen
as a consequence of friction-to-heat conversion along deeply rooted
reverse faults (Weissel et al., 1980; Geller et al., 1983; Stein et al., 1988;
Gordon et al., 1990). An alternative view was immediately raised:
deformation may have concentrated in areas of high thermal state and
weaker rheology, the high heat ow being a cause rather than a
consequence of the localization of deformation.Stein and Weissel (1990)
ruled out large-scale reheating at the base of the lithosphere showing
that there was no associated bathymetric swell. They further inferred
that the source of additional heat had necessarily to be shallow (Geller
et al., 1983; Stein et al., 1988) in order to satisfy simultaneously the
present-day depth to basement, the deep seismicity (Okal, 1983)
(suggesting low temperatures at depth of 3040 km) and the high
surface heat ow. The topic remained closed and rather unsolved until
Verzhbitsky and Lobkovsky (1993) proposed that the required shallow
heat source may relate to the exothermic serpentinization of mantle
peridotites, as rst mentioned in Fyfe (1974). They provided a crude
estimate of the produced heat and concluded that serpentinization may
be as efcient as other mechanisms to increase signicantly the surface
heat ow in this part of the Indian Ocean.
Substantially more data are available today: (1) Using deep seismic
refraction, Louden (1995) detected a low velocity zone thathe attributed
to partially serpentinized clasts of peridotites in the gabbros layer at the
base of the oceanic crust; (2) At the scale of the thrusts network, one
Earth and Planetary Science Letters 276 (2008) 140151
Corresponding author. Ecole normale supérieure, Laboratoire de Géologie, 24 rue
Lhomond, 75005, Paris, France.
E-mail address: delesclu@geologie.ens.fr (M. Delescluse).
0012-821X/$ see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.epsl.2008.09.017
Contents lists available at ScienceDirect
Earth and Planetary Science Letters
journal homepage: www.elsevier.com/locate/epsl
detailed heat-ow prole across two major active faults was obtained
during a pre-site survey of the ODP Leg 116 drilling in an area close to the
AfanasyNikitin Seamount Chain (Cochran et al., 1989). Distribution of
the observed heat ow along this prole was modeled by Ormond et al.
(1995) as vigorous uid circulation through the Bengal fan sediments,
redistributing an already anomalous basal heat ow. No heat source
origin is proposed to explain the 30 mW/m
2
anomaly, but these data
do conrm the linkwith the faults; (3) Numerous seismic reectors have
been imaged in the mantle, including deep penetrating faults. Acquired
in 1991, Phèdre multichannel seismic proles (Chamot-Rooke et al.,
1993) imaged the deep geometry of the thrust faults along a 2100 km-
long prole across the area. Below the ODP Leg 116 site, two fault
reectors cutacross discontinuous Moho phasesand reach deep into the
mantle, suggesting that uid paths do exist at sub-Moho depth. Deep
Fig. 1. Location of surface heat-ow measurements in the Indian Ocean and some of the seismic tracks where EW thrust faults have been imaged. V: Conrad monotrace prole
(VanOrman et al.,1995); P1: Phèdre Leg1 wide angle prole; P2: Phèdre Leg2 wide angle prole (Chamot-Rooke et al.,1993); K: Eastern CIB seismic prole (Krishna et al., 20 01). Red
symbols cover the Central Indian Basin, green symbols the Wharton Basin, blue symbols the remaining areas. The size of the symbol is proportional to the heat-ow value.
Background is a ltered Sandwell 15.1 satellite (Sandwell and Smith, 1997) gravity eld showing contrasting structural trends on both sides of NinetyEast Ridge. Contour of the
Afanasy Nikitin Seamount is also shown. The ODP Leg 116 site is represented by a star. (For interpretation of the references to color in this gure legend, the reader is referred to the
web version of this article.)
141M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
mantle reectors that do not seem to connect to active faults were also
found. Brightness of the fault planes was further interpreted as
evidences for water circulation (Bull and Scrutton, 1990b).
Discarding the possibility of a pre-deformation large-scale reheating
due to external sources, we quantitatively examine in this paper two
heating mechanisms that may explain the strong link between fault
localization and high heat ow: friction on the thrust fault network and
exothermic serpentinization. Our analysis is based on reprocessed
multichannel seismic data showing newevidences for uid paths down
to the mantle and associated enigmatic deep, bright reectors. We
successively discuss evidences for serpentinization in the Indian Ocean,
the role of uid convection, expansion during serpentinization and
nally other sites of serpentinization in the oceans.
2. Data
2.1. Heat ow versus age analysis
The global heat-ow compilation by Pollack et al. (1993) includes
some 300 measurements acquired in the Indian Ocean during various
surveys since the 1960's, to which we have added measurements from
the pre-site survey of ODP Leg 116 drilling (Cochran et al., 1989)(Fig. 1).
Heat-ow values range from 2 mW/m
2
to more than 200 mW/m
2
.We
show in Fig. 2 the distribution of heat ow as a function of age of the
underlying oceanic lithosphere. The main gure is not different from
what waspublished previously (e.g. Geller et al.,1983;Stein and Weissel,
1990) but we choose to split the data into three subsets: the Central
Indian Basin (including the region of the Afanasy Nikitin Seamount
Chain), theWharton Basin, and theremaining data. The rst two subsets
cover areas that are presently suffering active deformation. Theoretical
plate cooling models are given for comparison. Quality of data was
Fig. 2. Surface heat ow versus age of the oceanic lithosphere inthe Indian Ocean. Colors
and symbols are similar to Fig. 1. A 10 Myr binning is superimposed, as well as two
conductive plate models (dark envelope) for two different plate thickness (95 km and
125 km). CIB: Central IndianBasin; ANS: Afanasy Nikitin Seamount; WB: Wharton Basin.
Fig. 3. Portion of the reprocessed multichannel Phèdre line across the ODP Leg 116 site. Top inset shows the detailed heat-ow measurements obtained above two active faults during
the ODP Leg 116 pre-site survey (Shipboard Scientic Party, 1989). Open dots underline deep crustal and mantle reectors.
142 M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
discussed in a number of papers and they all concluded that the heat
ow was abnormally high (Geller et al.,1983; Steinet al., 1988; Stein and
Weissel, 1990; Williams,1990; Ormond et al.,1995).
The observed heat ow is widely scattered above and below the
plate cooling model predictions. This type of scattering may be the
result of transient hydrothermal circulations, and indeed some of the
temperature gradients were demonstrated to be non-linear (Geller
et al., 1983). An over simplistic view is that the lowest heat-ow values
are sites of recharge of cold uids whereas highest heat-ow values
are hot uids seeping sites. However, the Central Indian Basin
denitely shows a different pattern, the mean heat ow there being
clearly above the theoretical predictions. An alternative to the high
heat-ow anomaly interpretation is that the measurements system-
atically missedthe recharge areas, as suggested by apparently
normal values estimated within two ODP holes (Williams, 1990). This
alternative interpretation was dismissed in Stein and Weissel (1990)
and Ormond et al. (1995), as will be discussed in a latter section.
The high heat-ow anomaly appears to be restricted to the Central
Indian Basin, since the Wharton Basin seems to show the right thermal
age (Stein et al., 1988). Actually, the Cretaceous Central Indian Basin
(mean age of 74 Ma at measurements sites) has a higher surface heat
ow (80 mW/m
2
) than the younger Wharton Basin (mean age of 61 Ma
and average heat ow of 65 mW/m
2
). The simplest interpretation is that
in the Central Indian Basin, convective hydrothermal circulation is
superimposed onto an abnormally hot conductive lithosphere, and that
both processes are intimately linked to the active fault network.
Hydrothermal circulation may also be present in the Wharton Basin,
but the background heat ow there is presently normal.
2.2. Spatial correlation between high heat ow, high straining regions
and active faults
Seismic proling allows us today to draw a quite accurate contour
of the thrust fault network in the Central Indian Basin. It spreads along
a roughly equatorial zone around the Afanasy Nikitin seamount, from
70°E to the NinetyEast Ridge and from 7°S to approximately 5°N. At
large scale, the high heat-ow values cover more or less the thrust
fault network. In our recent kinematic study of IndiaAustralia
intraplate deformation (Delescluse and Chamot-Rooke, 2007), we
showed that the best instantaneous strain rate eld, in terms of style
and amplitude of deformation, is obtained when heat ow is used as a
proxy for rheological strength of the Indian Ocean oceanic lithosphere.
In the northeastern Indian Ocean, this is equivalent to considering the
entire fault network area as weak.
At shorter scale, best evidences for correlation between high heat-
ow and active deformation come from one detailed heat-ow prole
across the ODP Leg 116 site (Fig. 3). This 15 km -lon g prole was sampled
every half a kilometer or less. The 44 heat-ow measurements average
84 mW/m
2
,withapeakvalueat166mW/m
2
between two active reverse
faults both with large offsets. A reprocessed multichannel line of the
Phèdre Cruise over the same area (Fig. 3) shows that the fault reectors
plunge deep into the mantle, below Moho seismic phases. At the scale of
these two faults, the heat-ow prole thus reproduces the Central Indian
Basin large-scale anomaly of 2530 mW/m
2
.Ormond et al. (1995)
successfully modeled the bell shape of the heat-ow anomaly assuming
vigorous uid circulation within the fault planes above an abnormally hot
basement. Williams (1990) also proposed a set of models in which faults
are included as conduits.
These elements suggest that the spatial correlation both global
and local between faulted seaoor and thermal anomaly is not a
coincidence. The question that arises is whether the present-day
deformation is a cause or a consequence of the thermal anomaly. That
phase of deformation began 7.5 (Cochran et al., 1989)to9(Delescluse
et al., 2008) Myr ago as documented by regional late Miocene
unconformities in the Indian Ocean sediments. Stein et al. (1988)
noticed that there was no clear correlation between the exact location
of the earthquakes and the high heat-ow sites, probably because of
the large uncertainties on the position of the epicenters. Nevertheless,
they suggested that the absence of heat-ow anomaly and the low
seismicity in the Wharton Basin was indicating that it is presently
deforming much less than the Central Indian Basin. We know today
that the seismicity is not that low in the Wharton Basin (i.e. Mw= 7.9,
2000; Mw = 6.9, 1995) and that the strike-slip motion accommodated
there is signicant (Deplus et al., 1998). Summing moment tensors of
earthquakes, we actually found that the Wharton Basin is presently
the highest straining region of the Indian Ocean (Delescluse and
Chamot-Rooke, 2007). Strain is thus highest in a region where heat
ow is close to normal, which seems paradoxical both if heating is the
cause or the consequence of the deformation. WB and CIB areas are
actively deforming, but with contrasting style of faulting: mainly EW
reverse faults in the CIB and sinistral shear on NS vertical faults in the
Wharton Basin (Delescluse and Chamot-Rooke, 2007)(Fig. 1). We will
show in a latter section that the contrasting heat-ow anomaly
between the two regions is indirectly related to this contrasting style
of deformation.
3. Origin of the heat source
We review here two different mechanisms that have been
proposed as potential heat sources for the Central Indian Basin. Both
are tectonically driven: friction-to-heat conversion on the thrust fault
network and exothermic serpentinization along the fault planes at
mantle depth. We do not takeinto account the possibility of a hot-spot
reheating of lithosphere here, as it has already clearly been discarded
in Stein and Weissel (1990).
3.1. Friction on faults
Shear friction on faultshas long been recognized as a potential source
of heat, in particular at large continental strike-slip faults (Lachenbruch
and Sass, 1980; Scholz, 2000; D'Alessio et al., 2006) and at subduction
zones (Wang et al.,1995). Crude estimations have been proposed for the
reversefaults of the Indian Ocean (Geller et al.,1983), butwe now have a
much more detailed picture of the thrust network, including fault
spacing, geometryof the faultsat depth, rate of deformation.134 reverse
faults were recognized along one Phèdre deep seismic prole that span
the entire region of deformation in the Central IndianBasin, 30% of them
associated with faultreectors withinthe crust itself with an averagedip
angle of 36° to45° (Chamot-Rooke et al.,1993; Jestin,1994). Twoof these
reectors are clearly imaged in the region of the ODP Leg 116 site (Bull
and Scrutton,1990a), and our reprocessing shows that they cut through
the entire crust, intersect the Moho discontinuity and plunge deep into
the upper mantle.
Our calculation of the friction-related anomaly is based on the
diffusion of heat in a half-space (Carslaw and Jaeger,1959) adapted by
Lachenbruch and Sass (1980, 1992) for the San-Andreas strike-slip
fault. The rationale is to sum the individual contribution of each fault
along the entire prole length. Shear heat ow along a prole normal
to a vertical fault plane is:
qy;tðÞ¼
Qc
2πEix2
2þy2
4Kt

Eix2
1þy2
4Kt

ð1Þ
where yis the horizontal distance to the surface fault trace, tis time, x
l
and x
2
are the minimum and maximum depth of the brittle portion of
the fault, Q
c
is the product of the depth-averaged shear stress by the
fault slip-rate, E
i
is the exponential integral function, Kis the thermal
diffusivity.
This formalism is readily adaptable to a thrust fault with a dip
angle θ. This is done by the discrete summation of innitesimal
vertical planes disposed every dxdepth unit along the fault plane with
length of dx
sin θ
ðÞ to keep the size of the frictional surface identical. This
143M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
also allows for variable shear stress with depth. The heat-ow
anomaly across one thrust fault is then:
q1y;tðÞ¼
tx2x1
dxb
j¼0
Qx
2jdxðÞ
2πEiRj2

EiRj1
 ð2Þ
where R
jl
and R
j2
are functions of distance from the calculation point
(x=0,y) to respectively the top and bottom of each innitesimal
vertical plane. QxðÞ¼ V
cos θ
ðÞτsτs¼μρgx
1μtan θ
ðÞ

is now the product of the
shear stress at depth xon the thrust fault plane of dip θby the fault
slip-rate. Although these equations are strictly valid for innitely long
fault, which is obviously not the case in the Indian Ocean, they still
provide a reasonable estimation of the heat input since the fault
network extends for nearly 2000 km in the EW direction, much larger
than the typical depth of the faults.
The nal thermal anomaly is obtained by summing the effects of
faults spaced at regular intervals to simplify the natural situation. All
calculations were performed using a shortening rate of 1 cm/yr
integrated along the entire length of the prole crossing the fault
network, which can be regarded as an upper limit (Chamot-Rooke
et al., 1993; Delescluse and Chamot-Rooke, 2007). For simplicity, all
faults dip in the same direction, based on the observation that very
large portion of the deformed oceanic crust (100200 km) actually
show uniform fault dips (Jestin, 1994). The depth of the brittle
lithosphere is chosen to be 40 km, based on the maximum depth of
earthquakes in the area (Stein and Weissel, 1990). The shear stress
follows a standard Byerlee's law with null pore uid pressure, which
again will tend to overestimate the frictional heat due to unrealistic
high shear stress at depth. A standard faults spacing of 7 kmwas used,
but larger spacing was also tested since large offset faults are rather
2030 km spaced. The period of activity is set to 7.5 Myr.
Fig. 4. Main results of the friction heat production model. All graphs are transient states 7.5 Myr after the onset of deformation, except Graph D (gives the thermal evolution through
time) and E (reaches permanent regime). Graphs A to E are 900 km-long proles across the reverse faults network (Central Indian Basin model, 1 cm/yr of total shortening). Graph F is
a 1000 km crossing vertical strike-slip faults (Wharton Basin model, 2 cm/yr of total shearing). A) Faults spacing variation; B) Brittle depth variation; C) Dip-angle variation; D)
Evolution through time; E) Brittle depth variation (permanent regime); F) Strike-slip faults spacing variation.
144 M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
Fig. 4 shows the sensibility of the frictional model to some of the
main parameters. Varying the fault spacing (Graph A) does not change
the overall level of the thermal anomaly (5.5 mW/m
2
) since the
integrated amount of shortening remains identical. Given the over-
estimation that we made, it is readily apparent that friction cannot
explain the observed heat-ow anomaly.
Some other straightforward and useful information can be extracted
from the model. Shearing on individual faults could eventually be
detected through surface heat-ow measurements if their spacing
would be greater than 40 km (Graph A). By varying the thickness of the
brittle zone (Graph B), we learn that the heat source must be shallower
than 10 km to clearly individualize 30 km spaced faults: whatever the
strength and mechanism, if the heat source is linked to the faults, heat
conduction implies that we have virtually no chancegiven the number
of faults to measure a non-anomalous heat ow.
The lower the dip angle, the greater the frictional surface. However,
this effect is over compensated (Graph C) by the fact that the slip-rate
grows for larger dip angles, and also that the shear stress grows when
the dip angle is larger than π
41
2atan(µ). This is however not sufcient to
reach heat-ow anomalies signicantly greater than 5 mW/m
2
. The next
graphs show the evolution through time before permanent regime is
reached.A period of 7.5Myr of heating is far from the permanent regime
if the source is deep (40 km in Graph D). A permanent regime is almost
reached if the source depth is shallower than 10 km (compare Graph E
and Graph B), but then the thermal anomaly is small.
We performed the same calculations for the Wharton Basin to
evaluate frictional heating along a 1000 km-long EWprole normal to
theactiveNS strike-slip faults (Graph F). The total rate of shear was
assumed to be 2 cm/yr, which is probably maximized (Delescluse and
Chamot-Rooke, 2007). Our calculations lead to a maximum heat-ow
anomaly ranging from a 1mW/m
2
100 0 km-long plat eau in the case
of 10 km spaced strike slip faults to narrow 14 mW/m
2
spikes if the
shear is distributed onto 4 faults only. If shear heating is present in the
Wharton Basin, the chance to resolve it with surface heat-ow survey is
low unless measurements are performed close to the reactivated
fracture zones.
3.2. Exothermic serpentinization
Some serpentinization reactions are known to be signicantly
exothermic (Fyfe, 1974; Lowell and Rona, 2002; Fruh-Green et al.,
2003; Emmanuel and Berkowitz, 2006). Such heating seems to play a
major role in the Lost-City vent eld near the Mid Atlantic Ridge,
where mantle peridotites are exhumed at shallow depth (Fruh-Green
et al., 2003). It may be held responsible for sustained convection in
this off-ridge area (Kelley et al., 2001; Lowell and Rona, 2002;
Emmanuel and Berkowitz, 2006). Exothermic serpentinization is thus
an appealing mechanism to generate heat at rather shallow depth, and
we will review pieces of evidence that it may be at work in the Indian
Ocean in Section 4. Beyond the crude estimates found in Verzhbitsky
and Lobkovsky (1993), we describe here the thermal model that we
built to quantitatively calculate the amount of heat that may be
available in the deforming portion of the Indian Ocean.
3.2.1. Thermal model and kinetics of the serpentinization reaction
The hydration reaction is possible up to temperatures of 400 °C
(Emmanuel and Berkowitz, 2006) to 500 °C (Francis,1981). The reaction
is most efcient around 250300 °C (MacDonald and Fyfe,1985; Fruh-
Green et al., 2003; Emmanuel and Berkowitz, 2006) where lizardite is
the stable mineral while greater temperatures and pressures lead to a
stable antigorite mineral (Christensen, 2004). The reaction becomes
extremely slowaround and below 100 °C. The typical rock temperature
at Moho depth is 120 °C for a Cretaceous oceanic lithosphere, which
potentially allows for serpentinization in the lower oceanic crust and in
the upper mantle down to 25 km. We consider that the mantle is
virtually anhydrous atthese depths in the Central Indian Basin, sinceany
serpentinization process that would have occur at the very early
formation of the oceanic lithosphere would have been active at shallow
levels only due to the high thermal state at the ridge axis.
We use here the reaction kinetics described in Emmanuel and
Berkowitz (2006), i.e.:
Aρf
At¼Krρfð3Þ
where ρ
f
is the volumic mass of forsterite in the mantle. K
r
is the rate
coefcient, expressed as K
r
=Ae
b(Tc)
2
, this last formulation being
empirically constructed from kinetic data (Emmanuel and Berkowitz,
2006). A,band care different kinetic coefcients compiled in Table 1.
Serpentinization produces H=290 kJ/kg of forsterite (MacDonald and
Fyfe, 1985; Emmanuel and Berkowitz, 2006).
We also use the rate of heat generation per unit volume dened in
Emmanuel and Berkowitz (2006) as:
Q¼HAρf
At:ð4Þ
We nally assume that there is 75% of forsterite in the mantle which
leads to ρ
f
(t=0) =2400 kg m
3
for a mantle density of ρ
m
=3200kgm
3
.
A2Dnite difference algorithm is used to solve the transient heat
diffusion equation with spatially and temporally variable heat pro-
duction. We do not take into account volume changes for simplica-
tion, but we will discuss the matter in Section 4. Simple lithosphere
cooling is rst solved in 1D until t
start
=72.5 Ma, which will represent
theapproximativethermalstateattheonsetofdeformation
(56 mW/m
2
). The 2D nite difference grid is then chosen to be
30 km wide and 125 km deep with a cell dimension of about
150 m× 150 m. 45° dipping faults are disposed regularly, with periodic
thermal boundary condition, which is a good way to get rid of side
effects, and is appropriate for the fault network that we study. Heat
production then follows serpentinization fronts moving horizontally
away from the fault planes (Fig. 5). The reaction in the heating fronts is
forced to stop when half of the forsterite is serpentinized. At time
t
stop
=80 Ma, we thus obtain an average 50% serpentinization which
leads to a density of 2900 kg m
3
. The rationale is to keep the mantle
density close to the value obtained by Louden (1995) at the base of the
crust.
The amount of heat produced and the temperature eld evolution
throughtime are mainlycontrolled by three parameters: (1)the reaction
rate coefcient A, which is a measure of the efciency of the serpen-
tinization reaction; (2) the serpentinization front velocity S, which will
relate to the mechanical strength around the fault; and (3) the fault
spacing, i.e. the initial surface of serpentinization. Fault spacing was set
to 10 km in most of the runs, which is the distance between the main
faults observed at the ODP Leg116 site. The rsttwo parameters are very
difcult to set in a deep environment. One of the rare studies dealing
with serpentinization in relatively deep (5 km) oceanic environment
(MacDonald and Fyfe,1985) quotes velocity of 1000 m/Myr at 300 °C for
water propagation in t he host rock from an open crack. The same value is
used by Ranero et al. (2003) for serpentinization that may be active
around the bending related normal faults at subduction bulges.
The kinetics of the serpentinization is even less constrained. It is
fast when operated in the laboratory (a few days, A10
6
s
1
), but it is
most probably much slower in situ because of a smaller reaction
surface (Emmanuel and Berkowitz, 2006). An additional limiting factor
is the water supply (MacDonald and Fyfe,1985) rather than the kinetics
Table 1
Parameters for the kinetics of serpentinization
Rate coefcient magnitude A0.5 to 8 ×10
12
s
1
Peak reaction temperature c540 K
Kinetic coefcient b2.5×10
4
K
2
Serpentinization front velocity S3002000 m/Myr
145M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
itself. We choose a rate coefcient magnitude (A) spanning two orders of
magnitude and somewhat smaller than what is used by Emmanuel and
Berkowitz (2006) (i.e. around 10
12
s
1
instead of 10
10
s
1
), assuming
that the reaction surface at depth is less compared to the shallow
environment of Lost-City. Half-life of the reaction is then around
20,000 yrs at peak temperature, the coefcient bbeing chosen so that
the reaction is extremely slow for temperatures below 100 °C and also
virtually stopped above 400 °C (Emmanuel and Berkowitz, 2006).
3.2.2. Results of the serpentinization model
The goal is to reach at least a 20 mW/m
2
surface heat-ow
anomaly, which would be a signicant part of what is observed today.
The second constraint is timing, since a fundamental difference of the
serpentinization model compared to the frictional heating model is
that the source decays with time: by essence, the obtained anomaly is
thus transient and vanishes with time. Fig. 6 shows one typical run
and summarizes the sensitivity of the results to the main parameters.
The test run is for a 10 km faults spacing and a typical front velocity of
1000 m/Myr. The associated temperature anomaly in the upper
mantle does not exceed 80 K and lasts a few Myr only. In the case of a
shorter fault spacing, not shown here, the temperature eld is more
homogeneous and does not exceed 50 K. In all cases, the temperature
anomaly is limited to depths between 5 and 25 km. A lowtemperature
anomaly and a thermally quasi-normal lithosphere were actually two
requirements quoted in the early study of Stein and Weissel (1990).
In this simple test run, the peak surface anomaly (22 mW/m
2
)is
reached about 5 Myr after the onset of serpentinization and remains
above 10 mW/m
2
the remaining 2 to 3 Myr. Faster front propagation
(S=200 0 m/Myr) produces a sharper peak anomaly of 30 mW/m
2
which
does not last long enough, while slow propagation (S=500 m/Myr) just
reaches a low maximum of 12 mW/m
2
at 80 Ma. Changing the kinetics
results in marginal variations: increasing Ato 10
11
s
1
does not
produce a higher heat ow for any of the assumed front velocities. The
reason is thatfaster reaction rates result in a saturated heat output, since
the velocity of the serpentinization front is thetrue limiting factor. When
the front propagates to the next discrete 150 m wide cell of the model,
the previously hydrated volume has already released all of its potential
heat. In other words, in our model, the reaction is water limited, and is
equivalent to the successive serpentinization of spatially limited massifs
modeled or idealized by the serpentinization front. Finally,
decreasing the fault spacing to 7.5 km or even to 3 km has little effects,
provided that the front propagation velocity is downgraded (respec-
tively to S=750 m/Myr and S= 300 m/Myr). This is not surprising, since a
constant fault spacing to front propagation ratio results in a similar
history of serpentinization (i.e. 50% serpentinization completed after
5Myr).
Several sets of parameters thus lead to the required surface heat
ow. The remaining difculty is that the peak anomaly generally
comes too early in the thermal history. One possibility is that
serpentinization was delayed with respect to the initiation of de-
formation. In the 3 km spaced faults case, a linear progression of the
serpentinization front velocity from 150 m/Myr to 500 m/Myr would
correspond to an onset of serpentinization at 6.25 Ma. Equivalent
scenarios can be reached with 7.5 km and 10 km faults spacings (Fig. 6,
C-curves 1). New seismic data published in Delescluse et al. (2008)
show massive reactivation of the 3 km spaced oceanic fabric normal
faults at 9 Ma. After a 2 Myr period, faults are selectively abandoned
and only the present-day active faults remain. Our model is thus in
better agreement with the onset of this second phase of deformation
starting at 7 Ma under the effect of a severe weakening of fault planes
Delescluse et al. (2008).
4. Discussion
4.1. Clues to serpentinization in the CIB
Arguments in favor of past and present serpentinization in the
Indian Ocean mainly come from various seismic data. Refraction data
in the CIB southwest of the ODP site revealed a low velocity zone in
the lower crust, interpreted as partially serpentinized (35%,
ρ=2950 kg/m
3
) clasts of peridotites in the gabbro Louden (1995).
One seismic reection prole shows injection into one reverse fault that
has been interpreted as a serpentinite diapir based on gravity modeling
(Krishna et al., 2002). Seismic refraction data from Russian surveys also
Fig. 5. Sketch for the serpentinization model. Serpentinization fronts propagate horizontally from the faults; Serpentinization ends in cells where 400 °C is reached. Periodic thermal
boundary conditions are assumed on both sides. Grey shading indicates the degree of serpentinization.
146 M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
suggest velocities up to 7.3 km/s in the shallow crust, which could be
related to such diapirism (MacDonald and Fyfe, 1985; Levchenko and
Verzhbitsky, 2003; Verzhbitsky and Neprochnov, 2005).
We provide new pieces of evidence from our reprocessing of one
Phèdre seismic prole, where we detect enigmatic, bright and oblique
reectors deep in the mantle (Fig. 7). These reectors are disposed at
the termination of a much less visible reector which connects at the
basement interface with a clear reverse fault. Their approximate depth
(8 to 15 km) coincides with the maximum value of the serpentiniza-
tion reaction rate. We thus interpret these very bright reectors as
serpentinization fronts either active or fossil near an active thrust
fault used as uid path. Deeper reectors may be present, since the
serpentinization rate drops 20 km below the top of crust (Figs. 6,7).
Finally, serpentinization of peridotites appears as a quite efcient
heating mechanism: the transient heat-ow anomaly it produces is
much larger than what friction can provide (20 mW/m
2
versus
5 mW/m
2
), recalling that we maximized the friction effect in our
calculations. However, extreme surface heat-ow values cannot be
modeled without including uid ow: the exothermic serpentiniza-
tion provides a strong enough basal heat ow at depth, which is then
redistributed by convective uid circulation in the sediments
(Ormond et al., 1995), but also possibly in the crust (Louden, 1995).
Our thermal model shows that serpentinization is compatible with a
wide range of present-day surface heat ow (including close to
normal values), taking into account the transient nature of the thermal
anomaly. Serpentinization requires a signicant uid input that we
did not discuss nor take into account in our conductive model. The
question of the long term access of uids to fresh mantle also has to be
addressed.
4.2. Downward ow of water required by serpentinization: amount and
thermal effect
First, the required water input rate is not unrealistic. Serpentiniza-
tion consumes 300 L of water per cubic meter of olivine (MacDonald
and Fyfe, 1985; Fruh-Green et al., 2003). Our model involves half
serpentinization of, say 20 km (depth)× 30 km (width) × 1 km
(arbitrary since our model is 2D)= 600 km
3
of olivine, which means
that we have to bring a total of 0.5×60 300 × 10
9
=10
14
L of water
during the 5 Myr reaction period. This is equivalent to a ow of
0.02 µL s
1
.m
2
, or 0.6 L s
1
through the 30 km×1 km arbitrary
surface at the seaoor. With a conservative value of one single fault
used as uid path every 30 km, and a uid path width of 10 m, the
downward uid velocity is around 6× 10
8
ms
1
. This value is similar
to the 7 × 10
8
ms
1
upward uid velocity derived from non-linear
temperature gradients by (Geller et al., 1983).
Fig. 6. A) Snapshots of the anomalous thermal eld through time for the exothermic serpentinization model after 1 Myr, 2.5 Myr, 4.5 Myr and 6.5 Myr of deformation; B) Evolution of
the surface heat ow through time since the onset of deformation for three different serpentinization front velocities (500, 1000 and 2000 m/Myr) and variable kinetic coefcients.
Dotted curve is for Wharton Basin with vertical faults having 30 km spacing and a serpentinization front velocityof 1000 m/Myr. C) Evolution of the surface heat ow with the xaxis
chosen so that all models still reach 20 mW/m
2
at the origin: 1) Three overlapping curves show equivalent thermal history for different fault spacings (10 km, 7.5 km, and 3 km) but
scaled serpentinization front speeds (respectively 1000 m/Myr, 750 m/Myr and 300 m/Myr); 2) Test for a linearly increasing serpentinization front velocity with time (from 100 m/
Myr to 500 m/Myr with 3 km fault-spacing); 3) Test for a high serpentinization front velocity 2000 m/Myr for comparison.
147M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
Fig. 7. Bottom: Phèdre Leg 1 reprocessed prole showing deep and bright reectors at sub-Moho depth. Top: Same prole with sediments removed, using trace time-shift and mute to start the display at the top of the oceanic crust. The two-
way travel time is converted to depth using two different values for the mean velocity (6.5 km/s and 7.5 km/s). Right: Reaction rate coefcient for serpentinization as a function of depth, at the time of initiation of deformation and 7.5 Myr later.
Bright enigmatic reectors coincide with the predicted depth of maximum rate of serpentinization and are interpreted as serpentinization fronts.
148 M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
The nature and extent of recharge zones are however unknown and
the calculated velocity above is likely to be overestimated. Thesearch for
uid recharge zones was central to the debate on the ODP Leg 116 heat-
ow prole interpretation due to the absence of lower than normal
heat-ow measurements in the CIB. Williams (1990) proposed a uid
advection process to explain the84 mW m
2
average value, the modeled
section (15 km)being considered asa discharge zone. Noadditional heat
source is necessary but this approach requires a distant recharge zone
where downward ow of water should cool the sediments. Such a zone
where the heat ow would be lower than normal has never been
observed. The probability that all measurements missed recharge zones
is small, and this leaded previous authors (Geller et al., 1983; Stein and
Weissel, 1990; Ormond et al., 1995) to conclude to an anomalously hot
basement. Ormond et al. (1995) set up a 3D model including anomalous
basal heat ow, and showed that uids are more likely to convect in the
fault planes themselves rather than between two faults through a deep
conduit. In this case, along-strike variations of heat ow should be
observed and recharge zones are those where heat ow equal the
theoretical values. Although we cannot totally rule out that the linear
temperature gradients at ODP sites 717 and 719 showing normal heat
ow (Fig. 3) do represent the conductive background heat ow, they
may equally be these areas of diffuse recharge.
If exothermic serpentinization is responsible for the high heat-ow
anomaly, then the required deep downward ow of water (i.e. water
needed to feed the serpentinization reaction) must be negligible
compared to the surface uid convection input, otherwise low heat-
ow values should be observed. Ormond et al. need 2× 10
8
Lyr
1
.km
2
of water input in their model while serpentinization only needs
2×10
7
Lyr
1
/(30 km
2
)=7×10
5
Lyr
1
.km
2
, which is more than two
orders of magnitude less. We thus conclude that the deep downward
ow of water probably has a limited cooling effect if recharge zones
are diffuse.
4.3. Water progression in the oceanic mantle
Serpentinization tends to reduce rock porosity to zero (MacDonald
and Fyfe, 1985), so that the uid paths toward the mantle may become
rapidly closed. In a tectonically active setting, the easiest paths for
uids are then faults and cracks, which need to be permanently
refreshed(Ranero et al., 2003). Our model suggests that a more or
less continuous band of serpentinization is needed to reach the
required surface heat ow, since isolated massifs of serpentinites
would not release enough heat. Connection between propagating
serpentinization fronts could be achieved through the dense reverse
faults network. The details of the faults population indicate two
subsets of faults in the Indian Ocean (Delescluse et al., 2008): (1) small
offset faults (b100 m) as close as 2 to 3 km from one another, well
expressed in the sediments, in particular on high resolution seismic
proles (VanOrman et al., 1995; Delescluse et al., 2008), and (2) large
offset faults, reaching hundreds of meters to a kilometer, generally
associated with a deep bright reector (Chamot-Rooke et al., 1993;
Jestin,1994), with a much larger spacing (about 20 km). If the presence
of reectors is related to water, only the second population of large
offset faults could be used as uid path. Small offset faults are inactive,
but they are numerous and the fact that the lithosphere has been
highly fractured could help the propagation of water between large
faults. Ranero et al. (2003) also notice that some reectors in the
mantle near trench outer rises are not specically linked to an
observed fault at the surface. If as they point in their study, deep
reectors are caused by serpentinization or free uids, then the 5 to
6 km wide massifs observed in Fig. 7 would suggest that water can
propagate several kilometers away from the main fault. The low
frequency Phèdre seismic (Ziolkowsky et al., 2003) does not reach the
resolution of Ranero et al. seismic, but broadening of faults as they
deepen (Ranero et al., 2003) and/or undetected small faults at depth
may explain the lateral propagation of uids.
4.4. Expansion during serpentinization
Expansion during the serpentinization reaction may reach 53%
(MacDonald and Fyfe,1985; O'Hanley,1992)in volume. On onehand, the
difculty is space accommodation of this additional volume, specically
in a compressional tectonic setting, but on the other hand damages
resulting from thevolume expansion itselfcould allow uid progression.
At large degree of serpentinization, expansion should show at the
seaoor as uplift, and we should see a shallower than normal seaoor
depth. O'Hanley (1992), while acknowledging that eld evidences of an
increase in volume has not been recognized, mentions that realistic
increase in volume is between 25 and 45%. Serpentinization is restricted
to a depth of 20 km due to the temperature threshold. 50% serpen-
tinization affecting a 15 km thick column of mantle (with 75% of
forsterite) should thus produce a 50%× 25%× 75%×15,000=1400 m
uplift. This is a clear issue of the large-scale serpentinization hypothesis.
The bathymetric anomaly is actually opposite to the one that is
predicted by expansion. In details, the unloaded basement depth
(Fig. 8) shows that the 300 km buckling wavelength is superimposed
on a 800 km-long topographic depression which coincides with the
deepest geoid low on earth (Ćadek and Fleitout, 2003; Crosby et al.,
2006). This low was interpreted as the dynamic effect of deep mantle
sources, such as viscosity variations (Ćadek and Fleitout, 2003). The
reference topography in the CIB may thus be signicantly deeper than
the age prediction. Apart from partially canceling this very long
wavelength, the extravolume may also participate to the formation of
the buckling wavelength that has grown to a kilometer since the onset
of deformation (Gerbault, 2000).
We nally cannot rule out that other hydration reactions are also
exothermic. Fyfe (1974) cites some very exothermic hydration
reactions in the crust which would double the heat production in
our case. Less than 50% of mantle serpentinization would then be
needed, leading to a lesser expansion. Unfortunately, we know little
about the kinetics of these shallower reactions.
4.5. Serpentinization elsewhere
Fault spacing may explainwhy, while strain rate is at least as large in
the Wharton Basin than in the CIB, no clear heat-ow anomaly is
detected there. The two regions exhibit contrasting style of deformation
(Deplus et al., 1998; Delescluse and Chamot-Rooke, 2007). The CIB is an
area of ubiquitous reactivation of a dense network of normal faults
acquired at the ridge axis (Bull and Scrutton, 1990a, 1992; Jestin, 1994;
Delescluse et al., 2008), while in the Wharton Basin deformation is
Fig. 8. From top to bottom, bathymetry, unloaded basement depth (black), and
basement depth of the entire 1600+ km-long Phèdre Leg 1 seismic prole. The prole
starts at 14°S (55 Ma old oceanic lithosphere) and ends around 1°N (90 Ma old oceanic
lithosphere). Theoretical basement depth obtained from a plate cooling model is shown
by the dotted line (parameters slightly adapted from (Crosby et al., 2006): plate
thickness of 95 km and ridge depth at age zero of 2750 m). The dashed line shows a
hypothetical reference basement depth in the case of a dynamic topography following
the geoidal low of southern India.
149M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
accommodated by left-lateral, widely-spaced strike slip faults. Serpen-
tinization may be limited to short distances around them in the absence
of a dense active network of secondary faults as found in the CIB. Fig. 6
shows that even if propagation is effective, one vertical fault plane every
30 km leads to a heat-ow anomaly smaller than 10 mW/m
-2
.One
alternative is that heat has already passed in the WB, either because the
serpentinization delay may have been shorter or the deformation
started earlier there. Mantle dipping reectors and reduced seismic
velocities have been taken as strong evidences for serpentinization of
the oceanic mantle at the outer rise of the Middle America Trench,
possibly down to 2030 km (Ranero et al., 2003; Grevemeyer et al.,
2007), and at the Central Chile trench outer rise (Grevemeyer et al.,
2005; Contreras-Reyes et al., 2007). In both areas, the surface heat ow
shows a negative anomaly if any (Grevemeyer et al., 2005). The
exothermic nature of the reaction is not discussed and the reduced
heat ow is interpreted as the cooling effect of the downward ow of
water (Ranero et al., 2003; Grevemeyer et al., 2005; Emmanuel and
Berkowitz, 2006) that overcomes the nascent thermal heating at depth.
The contradiction with our results is only apparent. At fast subduction
rate (8 cm/yr in this case), bending faults do not reside for a long time in
the hydrationwindow (0.3 to 0.7 Myr Ranero et al., 2003), and heat
will distribute within the cold wedge once the lithosphere has reached
the trench. More data is needed at slow subductions to test this
hypothesis.
5. Conclusion
Among the physical mechanisms that may be responsible for the
high heat ow in the actively deforming Central Indian Basin, our
models indicate that exothermic serpentinization is the only process
that leads to a signicant amount of heat release. In contrast to
friction-related release of heat, the exothermic serpentinization
model is by essence transient and delivers a pulse of extra heat that
will rapidly decay with time once serpentinization is completed. We
quantitatively show that a pulse of serpentinization in the Central
Indian Basin that would have started with the onset of the intraplate
deformation there is compatible with the present-day surface heat ow
which is 2530 mW/m
2
above normal, provided a million year delay for
the establishment of a mature hydrothermal circulation and a high
enough thermal efciency of the hydration reactions. The moderate
transient temperature anomaly is followed by a denitive weakening of
the CIB fault network due to the low friction coefcient of partially
serpentinized peridotites (Escartin et al., 1997, 2001). In the Wharton
Basin, the thermal pulse has already passed, in relation to a totally
different geometry of the fault network (widely-spaced vertical strike-
slip faultsrather that closely spaced reverse faults). The enigmatic bright
reectors that we imaged in the CIB on multichannel seismic proles at
depth of 8 to 15 km below the top of the crust coincide with the location
of the maximum reaction rate coefcient of serpentinization in our
model. They are likely to result from serpentinization fronts propagation
at mantle depth. Our results suggest that pulses of serpentinization
within the oceanic lithosphere result in transient rheology that breaks
the age-dependent mechanicalstrength law: this mechanism may have
been responsible in the past for sudden plate reorganizations at global
scale, andit may also explain theabnormally low rigidity foundtoday for
some of the oldest oceanic plates (Bry and White, 2007; Korenaga,
2007).
Acknowledgements
We thank three anonymous reviewers for their constructive
reviews and the editor Claude Jaupart for his stimulating comments.
Phèdre multichannel lines were reprocessed with the GeoCluster
seismic software of CGGVeritas in the framework of the Enabling Grids
for E-sciencE project, co-funded by the European Commission through
the Sixth Framework Programme (http://www.eu-egee.org).
References
Bry, M., White, N., 2007. Reappraising elastic thickness variation at oceanic trenches.
J. Geophys. Res. 112, B08414. doi:10.1029/2005JB004190.
Bull, J.M., Scrutton, R.A., 1990a. Sediment velocities and deep structure from wide angle
reection data around Leg 116 sites. Proc. Ocean Drill. Prog. Sci. Results 116, 311316.
Bull, J.M., Scrutton, R.A., 1990b. Fault reactivation in the central Indian Ocean and the
rheology of oceanic lithosphere. Nature 344, 855858.
Bull, J.M., Scrutton, R.A., 1992. Seismic reection images of intraplate deformation,
Central Indian Basin, and their tectonic signicance. J. Geol. Soc. 149, 955966.
Ćadek, O., Fleitout, L., 2003. Effect of lateral viscosity variations in the top 300 km on the
geoid and dynamic topography. Geophys. J. Int. 152, 566580.
Carslaw, H.S., Jaeger, J.C., 1959. Conduction of Heat in Solids. Oxford University Press,
New York. 510 pp.
Chamot-Rooke, N., Jestin, F., De Voogd, B., the Phèdre Working Group,1993. Intraplate
shortening in the central Indian Ocean determined from 2100-km-long north
south deep seismic reection prole. Geology 21, 10431046.
Christensen, N.I., 2004. Serpentinites, peridotites, and seismology. Int. Geol. Rev. 46,
795816.
Cochran, J.R., et al., 1989. Intraplate deformation and Bengal Fan sedimentation:
background and objectives. Proc. Ocean Drill. Program, Initial Rep. 116, 311.
Contreras-Reyes, E., Grevemeyer, I., Flueh, E.R., Scherwath, M., 2007. Alteration of the
subducting oceanic lithosphere at the southern central Chile trench-outer rise.
Geochem. Geophys. Geosys. 8, Q07003. doi:10.1029/2007GC001632.
Crosby, A.G., McKenzie, D., Sclater,J.G., 2006. The relationship between depth, age and gravity
in the oceans. Geophys. J. Int. 166, 553573. do i: 10 .1111/j .13 65 -2 4 6X .2 0 06. 03 015. x.
D'Alessio, M.A., Williams, C.F., Burgmann, R., 2006. Frictional strength heterogeneity
and surface heat ow: implications for the strength of the creeping San Andreas
fault. J. Geophys. Res. 111, B05410. doi:10.1029/2005JB003780.
Delescluse, M., Chamot-Rooke, N., 2007. Instantaneous deformation and kinematics of the
IndiaAustralia plate.Geophys. J. Int. 168,818842. doi:10.1111/j.1365-246X.2006.03181.x.
Delescluse, M., Montési, L.G.J., Chamot-Rooke, N., 2008. Fault reactivation and selective
abandonment in the oceanic lithosphere. Geophys. Res. Lett. 35, L16312.
doi:10.1029/2008GL035066.
Deplus, C., Diament, M., Hébert, H., et al., 1998. Direct evidence of activedeformation in
the eastern Indian oceanic plate. Geology 26 (2), 131134.
Emmanuel, S., Berkowitz, B., 2006. Suppression and stimulation of seaoor hydrothermal
convection by exothermic mineral hydration. Earth Planet. Sci. Lett. 243, 657668.
Escartin, J., Hirth, G., Evans, B., 1997. Effects of serpentinization on the lithospheric
strength and the style of normal faulting at slow spreading ridges. Earth Planet. Sci.
Lett. 151, 181189.
Escartin, J., Hirth, G., Evans, B., 2001. Strength of slightly serpentinized peridotites:
implication for the tectonics of oceanic lithosphere. Geology 29 (11), 10231026.
Francis, T.J.G.,1981. Serpentinization faults and their role in tectonics of slow spreading
ridges. J. Geophys. Res. 86, 1161611622.
Fruh-Green, G.L., Kelley,D.S., Bernasconi, S.M., Karson, J.A., Ludwig, K.A ., Buttereld, D.A.,
Boschi, C., Proskurowski, G., 2003. 30,000 years of hydrothermal activity at the lost
city vent eld. Science 301, 495498.
Fyfe, W.S., 1974. Heats of chemical reactions and submarine heat production. Geophys.
J. R. Astron. Soc. 37, 213215.
Geller, C.A., Weissel, J.K., Anderson, R.N.,1983. Heat transfer and intraplate deformation
in the Central Indian Ocean. J. Geophys. Res. 88, 10181032.
Gerbault, M., 2000. At what stress level is the central Indian Ocean lithosphere
buckling? Earth Planet. Sci. Lett. 178, 165181.
Gordon, R.G., DeMets, C., Argus, D.F., 1990. Kinematic constraints on distributed
lithospheric deformation in the equatorial Indian-ocean from present motion
between the Australian and Indian plates. Tectonics 9 (3), 409422.
Grevemeyer, I., Kaul, N., Diaz-Naveas, J.L., Villinger, H.W., Ranero, C.R., Reichert, C., 2005.
Heat ow and bending-related faulting at subduction trenches: case studies of
Nicaragua and Central Chile. Earth Planet. Sci. Lett. 236, 238248.
Grevemeyer, I., Ranero, C.R., Flueh, E.R., Klaschen, D., Bialas, J., 2007. Passive and active
seismological study of bending-related faulting and mantle serpentinization at the
Middle America trench. Earth Pla net. Sci. Lett. 258, 528542. doi:10.1016/j.
epsl.2007.04.013.
Jestin, F., (1994). Cinématique rigide et déformations dans la jonction triple Afar et dans
le Bassin Indien Central. PhD thesis, Université Paris VI.
Kelley, D.S., Karson, J.A., Blackman, D.K., Fruh-Green, G.L., 2001. An off-axis hydrothermal
vent eld near the Mid Atlantic Ridge at 30 degrees N. Nature 412,145149.
Korenaga, J., 2007. Thermal cracking and the deep hydration of oceanic lithosphere: a
key to the generation of plate tectonics? J. Geophys. Res. 112, B05408. doi:10.1029/
2006JB004502.
Krishna, K.S., Bull, J.M., Scrutton, R.A., 2001. Evidence for multiphase folding of the
Central Indian Ocean lithosphere. Geology 29 (8), 715718.
Krishna, K.S., Gopala Rao, D., Neprochnov, Y.P., 2002. Formation of diapiric structure in
the deformation zone, Central Indian Ocean: a model from gravity and seismic
reection data. Proc. Indian Acad. Sci. 111, 1728.
Lachenbruch, A.H., Sass, J.H., 1980. Heat ow and energetics of the San Andreas fault
zone. J. Geophys. Res. 85, 61856222.
Lachenbruch, A.H., Sass, J.H., 1992. Heat ow from Cajon Pass, fault strength, and
tectonic implications. J. Geophys. Res. 97, 49955015.
Levchenko, 0.V., Verzhbitsky, E.V., 2003. Structural aspects of the intraplate deforma-
tions of the oceanic crust north of the Afanasy Nikitin rise (Indian Ocean.
Oceanology 43 (6), 840851.
Louden, K.E., 1995. Variation in crustal structure related to intraplate deformation:
evidence from seismic refraction and gravity proles in the Central Indian Basin.
Geophys. J. Int. 120, 375392.
150 M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
Lowell, R.P., Rona, P.A., 2002. Seaoor hydrothermal systems driven by the serpenti-
nization of peridotite. Geophys. Res. Lett. 29 (26), 14.
MacDonald, A.H., Fyfe, W.S., 1985. Rate of serpentinization in seaoor environments.
Tectonophysics 116, 123135.
O'Hanley, D.S., 1992. Solution to the volume problem in serpentinization. Geology 20,
705708.
Okal, E.A., 1983. Oceanic intraplate seismicity. Annu. Rev. Earth Planet. Sci.11,195214.
Ormond, A., Boulègue, J., Genthon, P., 1995. A thermoconvective interpretation of heat
ow data in the area of Ocean Drilling Program Leg 116 in a distal part of the Bengal
Fan. J. Geophys. Res. 100, 80838095.
Pollack, H.N., Hurter, S.J., Johnson, J.R.,1993. Heat ow from the earth's interior: analysis
of the global data set. Rev. Geophys. 31 (3), 267280.
Ranero, C.R., Phipps Morgan, J., McIntosh, K., Relchert, C., 2003. Bending-related faulting
and mantle serpentinization at the Middle America trench. Nature 425, 367373.
Sandwell, D.T., Smith, W.H.F., 1997. Marine Gravity from Geosat and ERS-1 Altimetry. J.
Geophys. Res. 102, 1003910054.
Scholz, C.H., 2000. Evidence for a strong San Andreas fault. Geology 28 (2), 163166.
Shipboard Scientic Party,1989. Site 718: Bengal Fan.In: Cochran, J.R., Stow, D.A.V., et al.
(Eds.), 1988 Proc. ODP, Init. Repts., College Station, TX (Ocean Drilling Program),
vol. 116, pp. 91154.
Stein, C.A., Weissel, J.K., 1990. Constraints on the Central Indian Basin thermal structure
from heat ow, seismicity and bathymetry. Tectonophysics 176, 315332.
Stein, C.A., Hobart, M.A., Abbott, D.A., 1988. Has the Wharton Basin's heat ow been
perturbed by the formation of a diffuse plate boundary in the Indian Ocean?
Geophys. Res. Lett. 15, 455458.
VanOrman, J., Cochran, J.R., Weissel, J.K., et al., 1995. Distribution of shortening between
the Indian and Australian plates in the Central Indian Ocean. Earth Planet. Sci. Lett.
133, 3546.
Verzhbitsky, E.V., Lobkovsky, L.I., 1993. On the mechanism of heating-up of the Indo-
Australian plate. J. Geodyn. 17, 2738.
Verzhbitsky, E.V., Neprochnov, Y.P., 2005. Deep structure of the central Indian Ocean
inferred from geophysical data. Geotectonics 39 (3), 213223.
Wang, K., Mulder,T., Rogers, G.C., Hyndman, R.D., 1995. Case for verylow coupling stress
on the Cascadia subduction fault. J. Geophys. Res. 100, 1290712918.
Weissel, J.K., Anderson, R.N., Geller, C.A.,1980. Deformation of the Indo-Australian plate.
Nature 287, 284291.
Williams, C.F., 1990. Hydrothermal circulation and intraplate deformation: constraints
and predictions from in-situ measurements and mathematical models. In: Cochran,
J.R., Stow, D.A.V., et al. (Eds.), Proc. ODP, Sci. Results, 116:College Station, TX (Ocean
Drilling Program), pp. 345359.
Wiens, D.A., Stein, S., DeMets, C., Gordon, R.G., Stein, C., 1986. Plate tectonic models for
Indian-Ocean intraplate deformation. Tectonophysics 132, 3748.
Ziolkowsky, A., Hanssen, P., Gatliff, R., Jakubowicz,H., Dobson, A., Hampson,G., Li, X.Y., Liu,
E., 2003. Use of low frequencies for sub-basalt imaging. Geophys. Prospect. 51 (3),
16918 2.
151M. Delescluse, N. Chamot-Rooke / Earth and Planetary Science Letters 276 (2008) 140151
... The progress of the serpentinization reaction is given by dξ hyd = k hyd (1 − ξ hyd )dt, where the rate constant k hyd is obtained from Delescluse & Chamot-Rooke (2008;see Appendix B) and (1 − ξ hyd ) accounts for the progressive loss of reactants as the reaction proceeds in any one location. Heating due to the exothermic hydration reaction is included using (dξ hyd /dt)|ΔH hyd |, where ΔH hyd is the enthalpy of the serpentinization reaction (see Appendix B). ...
... and a h = 8 × 10 −12 s −1 , b h = 2.5 × 10 −4 K −2 , and c h = 400 K. These constants are from Delescluse & Chamot-Rooke (2008), with the exception that we revised c h downward from their original value of 540 K. With this adjustment the peak in the rate constant occurs at 400 K (120°C). ...
Article
Full-text available
We present oxygen isotopic analyses of fragments of the near-Earth C b -type asteroid Ryugu returned by the Hayabusa2 spacecraft that reinforce the close correspondence between Ryugu and CI chondrites. Small differences between Ryugu samples and CI chondrites in Δ ′ 17 O can be explained at least in part by contamination of the latter by terrestrial water. The discovery that a randomly sampled C-complex asteroid is composed of CI-chondrite-like rock, combined with thermal models for formation prior to significant decay of the short-lived radioisotope ²⁶ Al, suggests that if lithified at the time of alteration, the parent body was small (≪50 km radius). If the parent planetesimal was large (>50 km in radius), it was likely composed of high-permeability, poorly lithified sediment rather than consolidated rock.
... In addition to origination of fluids (Dmitriev et al., 1999), serpentinization is accompanied by the formation of a new magnetic layer due to crystallization of magnetite and change of the primary magnetic anomalies pattern (Astafurova et al., 1996;Oufi et al., 2002), the volumetric extension of serpentinized rock and decreasing of its density to 20% (Physical …, 1984), and the enhanced heat flow due to exothermic effect (Delescluse and Chamot-Rooke, 2008). The scarce measurements of heat flow in the deep-sea polar area (Khutorskoy et al., 2013) do not allows correct estimation of serpentinization heating contribution to its value. ...
Article
Full-text available
Flat spot anomalies in the Quaternary part of the section of the Nansen Basin are imaged in seismic records and are interpreted to be related to gas-rich fluid accumulations. The flat spots are mainly located above basement highs between magnetic spreading anomalies C20 (~43 Ma) and C12 (~33 Ma). The complex morphometric analysis of flat spots show that serpentinization processes identified from modelling of gravity anomalies could be original gas source. This process also makes smoothing of the basement highs amplitudes. The depth of the top of the flat spots below the seafloor has an almost constant value of ~390 m indicating the ascent of gases from variable basement depths to a common subsurface fluid trap. The depth of the anomalies below the seafloor corresponds to a theoretical thickness of gas hydrate stability zone in the studied region. Gravity modeling along the Arktika-2011-03 section showed the position of the upper mantle blocks with lower (to 2.95 g/cm3) density within the highs of the acoustic basement. The flat spot anomalies occur above basement highs, below which blocks with lower density typical of serpentinized rocks are modelled. Thus, the serpentinization of the upper mantle ultramafic rocks is considered a main geochemical process, which can explain generation and accumulation of gas in oceanic abyss at a 1–3 km thick sedimentary cover, as well as small vertical movements of the basement blocks due to density reduction and expansion of serpentinized rock.
... In addition to the formation of fluids (Dmitriev et al., 1999), serpentinization is accompanied by the appearance of a new magnetoactive layer due to the formation of magnetite, dilatation of the rock subjected to serpentinization, and simultaneous decrease of its density down to 20% (Fizicheskie …, 1984), as well as manifold increase heat flow due to the exothermic effect of this process (Delescluse and Chamot-Rooke, 2008). The formation of chemogenic magnetite undoubtedly affects the primary linear image of magnetic anomalies by imposing an additional component that creates a mosaic pattern of the anomalous mag-LITHOLOGY AND MINERAL RESOURCES Vol. ...
Article
Full-text available
The paper analyzes the "bright spot" and "flat spot" anomalies of seismic data on the southeastern flank of the Knipovich Ridge associated with the accumulation of free gas in the sedimentary cover above the oceanic basement. The identified anomalies are associated spatially with negative values of the residual Bou-guer anomaly and positive magnetic field (ΔТа) anomalies. This fact indicates the existence of decompaction zones in the crust and upper mantle related to serpentinization that can also provoke the superimposed, probably modern, chemogenic magnetization and distortion of the primary linear pattern of magnetic anomalies in the oceanic basement in the study area. Serpentinization was also responsible for vertical displacements of the crustal and upper mantle blocks on the flanks, leading to deformations of the sedimentary cover with the rock dilation. Off-axis seismicity indicates tectonic disruptions on flanks of the ridge with a higher access of water necessary for the serpentinization and the subsequent change in the physical properties of rocks reflected in geophysical fields. The eastern flank of the Knipovich Ridge underwent tectonic activation along the basement structures representing the northern extension of the Senja fracture zone, resulting in accumulations of free gas in the sedimentary cover.
... The upper mantle at the footwalls of the two inferred detachments shows strong reflections, which may result from the physical property changes associated with serpentinization caused by the fluid input along deep-rooted fault planes, resembling the upper mantle alteration near deep-rooted faults in other areas (e.g. Delescluse and Chamot-Rooke, 2008;Bayrakci et al., 2016;Ferrand et al., 2017;Ferrand, 2020;Prigent et al., 2020). ...
Article
The 85°E Ridge is a prominent linear structure in the northeastern Indian Ocean. Its nature and origin are still controversial due to the coverage of thick Bengal Fan sediments. Here we present a newly collected 400-km-long 2-D seismic reflection profile which crosses the 85°E Ridge near 11°N and shows the basement morphology and internal structures of the ridge. The ridge has a thickened crust characterized by two basement highs rising up to ~2–3 km above the surrounding oceanic crust and a Moho depression with amplitude of ~3–4 km. Typical volcanic structures formed by hotspot volcanism are identified in the ridge. Deep-rooted faults played important roles in the formation of the 85°E Ridge. They acted as the preferred conduit for the upwelling of hotspot melts, but their syn- and post-volcanic reactivations induced the collapse of volcanic sequences and the relative uplift of the East Ridge. These observations support that the 85°E Ridge near 11°N having formed by hotspot magmatism drained along a leaky fracture zone. Combined with previous results, a new single-hotspot formation model of the 85°E Ridge is proposed, in which the entire 85°E Ridge was formed by a weak and pulsating mantle plume acting on the northward drifting Indian Plate. Its along-strike variations in location, morphology, and structure in different segments are interpreted to reflect different elastic strength of the oceanic lithosphere and the existence of a pre-existing fracture zone which laterally redistributed the shallow melts.
... However, the gravity model ( Fig. 12) indicates that also the area south of 470 km model distance is underlain by material with lower densities. Delescluse and Chamot-Rooke (2008) interpret bright reflectors in MCS data at a Moho depth level (8-15 km below the seafloor) as serpentinization fronts in the Central Indian Basin. Deep faults cut through the entire oceanic crust and offset the Moho leading to sea water infiltration into the uppermost mantle and an exothermic serpentinization reaction. ...
Article
Full-text available
We present results derived from a seismic refraction experiment and gravity measurements about the crustal structure of southern Sri Lanka and the adjacent Indian Ocean. A P-wave velocity model was derived using forward modelling of the observed travel times along a 509 km long, N-S trending profile at 81°E longitude. Our results show that the continental crust below southern Sri Lanka is up to 38 km thick. A ~ 65 km wide transition zone, which thins seawards to ~7 km thickness, divides stretched continental from oceanic crust. The adjacent, 4.7 to 7 km thick normal oceanic crust is covered by up to 4 km thick sediments. The oceanic crust is characterized by intra-crustal reflections and displays P-wave velocity variations, especially in oceanic layer 2, along our profile. In the central part of the profile, the uppermost mantle layer is characterized by normal P-wave mantle velocities of 8.0–8.1 km/s. At the southern end of the profile, unusual low upper mantle seismic velocities, ranging from 7.5 to 7.6 km/s only, characterize the uppermost mantle layer. These low upper mantle velocities are probably caused by partially serpentinized upper mantle. At even greater depths the upper mantle layer is characterized by velocities of 8.3 km/s on average. The type of margin along our profile is difficult to identify, since it is characterized by features typical for different types of margins.
Article
Several sources of natural hydrogen are known or postulated but the process of serpentinization, the action of water on ultramafic rocks, is shown to be the most effective. Studies indicate that the rates and volumes generated by high-temperature serpentinization, (i.e., in the temperature range of 200-320°C), could feed a focused hydrogen system potentially capable of sealing and trapping gas-phase hydrogen in commercially-sized accumulations. Natural hydrogen is generated by serpentinization wherever ultramafic rocks can be penetrated by aqueous fluids. This includes diverse geotectonic settings ranging from divergent and convergent plate margins to intra-plate orogenic belts and Precambrian cratons. The ‘hydrogen system’ describes the generation, migration and sealing/trapping of hydrogen. There are two parts to the ‘generic hydrogen system’: the ‘source-generation sub-system’ requires an ultramafic protolith, usually in basement, and a supply of water penetrating basement rocks. In the ‘migration-retention sub-system’ migration, sealing and entrapment of gas-phase hydrogen behaves the same as for hydrocarbon gases. The hydrogen system by serpentinization is used to develop play models to guide exploration in the accessible and exploitable geotectonic settings of continental cratons, ophiolites and convergent margins. Thematic collection: This article is part of the Hydrogen as a future energy source collection available at: https://www.lyellcollection.org/topic/collections/hydrogen
Article
Full-text available
El presente trabajo es un estudio de la litosfera continental y oceánica en el Margen Continental Argentino. Estas investigaciones están basadas en datos geofísicos existentes en la zona correspondiente a la plataforma, talud, emersión y llanura abisal de Argentina, comprendida entre los 50° a 66° de longitud oeste y 36° a 50° de latitud sur. El mismo consiste en la obtención de un modelo de compensación isostática de la zona y su correlación con modelos existentes. Con el doble objetivo de evaluar el com­portamiento isostático de la región y buscar mecanismos que permitan explicar su desarrollo, se consideran modelos de compensación local (mecanismo de Airy) y compensación regional (mecanismo Flexural). Siendo correlacionados con los modelos de a) Profundidad del Basamento y Espesores Sedimentarios, b) Profundidad de la Discontinuidad corteza-manto superior y c) Profundidad del Punto de Curie (CPD, Curie Point Depth), para analizar el estado isostático de la zona de cuencas offshore, adelgazamiento cortical y niveles astenosféricos, entre otras cosas. El análisis comparativo de estos elementos permite determinar el grado de compensación isostática de esta zona, la geometría de las cuencas offshore en su relación con el adelgazamiento cortical y ascenso astenosférico, ubicar los sectores donde el manto superior estaría magnetizado, modelar la asimetría lístrica de la cuenca del Colorado, determinar por primera vez los límites y geometría de los altos estructurales de Belén y San Javier y discutir las causas que dan origen a la gran profundidad del basamento en la Cuenca Argentina, entre otras como zonas de sepentinización y la zona de falla de transformación de Belén. De la misma manera se abren interrogantes que requieren nuevos análisis y estudios, como los relacionados al CPD en las cuencas de Valdés, sectorizar la zona modelada para buscar un mejor ajuste de densidades, profundizar las causas de la gran batimetría de la Cuenca Argentina, entre otros.
Article
Full-text available
The recycling of water into the Earth’s mantle via hydrated oceanic lithosphere is believed to have an important role in subduction zone seismicity at intermediate depths. Hydration of oceanic lithosphere has been shown to drive double planes of intermediate-depth, Wadati-Benioff zone seismicity at subduction zones. However, observations from trenches show that pervasive normal faulting causes hydration ~25 km into the lithosphere and can explain neither locations where separations of 25–40 km between Wadati-Benioff zone planes are observed nor the spatial variability of the lower plane in these locations, which suggests that an additional mechanism of hydration exists. We suggest that intraplate deformation of >50-m.y.-old lithosphere, an uncommon and localized process, drives deeper hydration. To test this, we relocated the 25 November 2018 6.0 MW Providencia, Colombia, earthquake mainshock and 575 associated fore- and aftershocks within the interior of the Caribbean oceanic plate and compared these with receiver functions (RF) that sampled the fault at its intersection with the Mohorovičić discontinuity. We examined possible effects of velocity model, initial locations of the earthquakes, and seismic-phase arrival uncertainty to identify robust features for comparison with the RF results. We found that the lithosphere ruptured from its surface to a depth of ~40 km along a vertical fault and an intersecting, reactivated normal fault. We also found RF evidence for hydration of the mantle affected by this fault. Deeply penetrating deformation of lithosphere like that we observe in the Providencia region provides fluid pathways necessary to hydrate oceanic lithosphere to depths consistent with the lower plane of Wadati-Benioff zones.
Thesis
Full-text available
Some of the world’s most prolific petroleum systems are located in salt-bearing passive margins. Due to its rheological characteristics, salt largely affects the sedimentary basins. In the Western Mediterranean Sea, salt was deposited during the Messinian Salinity Crisis (MSC, 5.96-5.32 Ma), mainly in the deep sub-basins and lower slope.Using an extensive seismic dataset, a regional description and categorization of salt structures is presented. The salt structures change morphology at the boundary between different crustal natures. The compilation of crustal segmentation and salt morphologies in different salt-bearing margins, such as the Santos (Brazil), Angolan, Gulf of Mexico, Morocco and Nova Scotia margins, seems to confirm this correspondence. In the Western Mediterranean, the constant initial salt thickness, the horizontal salt base in the deep basin, the homogenous multikilometer pre-Messinian sequence, the early deformation above transitional and oceanic crust and a weak sedimentary cover in the slope (except for the Gulf of Lion) seem to call into question the classical mechanisms of salt tectonics, which hardly explain the crustal segmentation – salt morphologies coincidence. The hypothesis of a thermal influence on deformation is therefore evaluated. The thermal regime was investigated by analysing and interpreting surface heat flow (HF) data recently acquired during the WestMedFlux1-2 surveys (2016-2018) with the complement of a thermo-kinematic model. Among the small-wavelength HF influenced by local phenomena, a regional thermal segmentation is observed. The role of temperature on salt deformation remains the only plausible hypothesis for the explanation of the observed correspondence and deserves further investigation.
Thesis
Full-text available
Plate tectonics usually describes rigid plates limited by narrow boundaries. It is howevernot always true and diffuse boundaries are quite frequent (Figure 1), particularly on thecontinents. Among the known diffuse deformation zones, the central Indian ocean clearly appearsas a singularity because of its exceptional size and because deformation occurs in theoceanic lithosphere (Figure 1). Furhter, the thick sedimentation from the Bengal fan allowsthe unravelling of the history of deformation since the lateMiocene through seismic profiles.Intraplate diffuse deformation between India and Australia is thus the ideal place to study thelocalization processes of a nascent deformation and the role of preexistant structures, sincethe structural heritage is much more simple in the oceanic domain.In a first chapter,we present the formation of the Indian Ocean starting with the GondwanaBreakup.We focus on the India-Asutralia plate and discuss the different steps leading tothe formation of the Central Indian Basin (CIB), the NinetyEast Ridge (NyR) and the WhartonBasin (WB). We then discuss the concept of diffuse boundary between India and Australia,detailing the different styles of deformation from seismicity and new seismic profiles.In a second chapter, we will focus on the age of the onset of deformation using seismic profilesand ODP holes. The evolution of deformation is also discussed in a paper published inGeophysical Journal International (Delescluse et Chamot-Rooke, 2007). This paper presentsthe instantaneous deformation field and the associated velocity field derived from GPS andfocal mechanisms. The instantaneous kinematics is in good agreement with a continuum ofdeformation since lateMiocene.The next chapter deals with the localization processes in the CIB. First, we study the origin ofthe heat-flow anomaly which seems correlated to reverse faults. Heat-flow is used as a proxyfor the rheology of the lithosphere in our best kinematic model but its origin remains unknown.In a paper published in Earth and Planetary Science Letters (Delescluse et Chamot-Rooke, 2008), we propose that the thermal anomaly is a result of the exothermic serpentinizationof the oceanic mantle. We show that the thermal anomaly is a consequence of deformationand is not preexistant. The equatorial localization of deformation is best explained bythe reactivation of paleo-normal faults from the oceanic fabric when large intraplate stressesare correctly oriented. Using high resolution seismic profiles and a finite element model, weshow in a paper published in Geophysical Research Letters (Delescluse et al., 2008) that followingfull reactivation, faults are selectively abandoned because of the weakening of some ofthe fault planes.In a last chapter, we show that the Aceh (2004) and Nias (2005) megathrust earthquakes actedas an intraplate seismicity "accelerator" in the near-trench oceanic plate (Wharton Basin).Using Coulomb stress variations, we show that these subduction earthquakes favor theoccurence of left-lateral strike-slip along N-S paleo-transform faults rather than neoformedNE-SWreverse faults.
Article
Full-text available
Large areas of passive ocean margins are covered by basalts which conceal sediments that may contain large accumulations of hydrocarbons. These basalts are extremely heterogeneous and scatter the seismic energy of the conventional seismic reflection system. We propose to modify the system to emphasize the low frequencies, using much larger air guns, and towing the source and receivers at about 20 m depth. The rationale for this approach is supported by synthetic seismograms over a realistic 1-D earth model. In the summer of 2001 we plan to obtain data over basalt in the North-East Atlantic using a suitablymodified system.
Article
Full-text available
The reduction of six wide-angle reflection profiles shot within the two fault blocks visited by Ocean Drilling Program (ODP) Leg 116 in combination with the ODP sonic logs has produced a velocity-depth structure for this area. The sediment velocity increases from 1.6-1.7 km/s in the near surface to 3.4-3.5 km/s immediately above basement with a velocity gradient of 0.75/s. A depth converted seismic reflection profile suggests that the pre-deformational basement surface was similar to the abyssal hill topography developed in the Pacific Ocean. A velocity for the top of oceanic layer 2 of 4.1 km/s was identified as layer 2A. Assuming a velocity gradient of 0.7/s, an estimate of layer 2 thickness was obtained of 1.5 km. It is possible to interpret residual depth anomalies in terms of a layer 3 that may be thinner than for normal oceanic crust.
Article
Full-text available
Multichannel seismic reflection profiles collected from the intraplate deformation area in the Central Indian Ocean Basin are used to describe brittle structures produced under compressive stress. Reflectors within oceanic basement are divided into four types: north-dipping and south-dipping reverse faults that are interpreted as reactivations of structures formed at the spreading-centre as outward and inward dipping faults respectively; lower-angle, north-dipping reflectors that probably represent new faults or faults just initiating; and sub-horizontal reflectors within the uppermost crust that are interpreted as hydrothermal alteration fronts. Within the sedimentary cover upwards fault propagation shows the faults steepening from c. 40-degrees just above basement to near vertical and is preceded by sediment folding. A fractal analysis of faulting suggests two fault populations possibly reflecting different criteria for brittle failure. Measurements of north-south crustal shortening indicate a shortening rate of 2.5 (+/- 0.9) mm a-1, which is at the lower end of predictions from plate motions, but significant enough to recognize this area as a diffuse plate boundary. The formation of long-wavelength basement undulations and the reactivation of fracture zones and ridge-parallel fault fabrics are linked in a unified tectonic model driven by the high level of intraplate compressive stress in the area. There is little evidence from the seismic profiles for intraplate deformation starting before the widespread unconformity dated as 7 Ma.
Article
Full-text available
From a data set comprising 110 spreading rates, 46 transform azimuths, and 151 earthquake slip vectors from the Indian Ocean and Gulf of Aden, a new rigid plate model describing the motion since 3Ma between India and Australia is determined. The Euler vector (ω=0.313°/Ma ~5°S, 78°E) lies near the middle of the equatorial, diffuse plate boundary dividing the Indian from the Australian plate and predicts a rate of north-south shortening along 85°E of 4±3mm/yr. The new model also predicts north-south extension of 6±2mm/yr (at 68°E) along the western segment of the diffuse plate boundary. Using data only along the Carlsberg and Central Indian ridges and no other plate boundaries, it is shown that plate motion data cannot be fit by a single Euler vector. However, the data are well fit by two Euler vectors when an east-west striking India-Australia plate boundary is assumed to intersect the Central Indian Ridge near the equator. -from Authors
Article
On the basis of the tectonic interpretation of the continuous seismic profiling data with account for the data of deep seismic sounding, we assessed the structure of the test area Sh-22-I of the Labrador Current in the region of the recent intraplate deformations of the Indian Ocean lithosphere northeast of the Afansii Nikitin Rise (Mid-Indian Basin). The possibility of the existence of wide diffuse sinistral shears over ancient transform faults was revealed, which proves their recent tectonic reactivation. The concept of the significant structure-forming role of the processes of serpentinization of the ultramafic rocks of the lower crust and upper mantle and diagonal shears is developed. We show the complicated tectonic interrelations between the young (recent tectonic) compressional structural pattern and that of the ancient Late Cretaceous spreading in the Mid-Indian Basin. The structural relations between the deep crustal inhomogeneities and their manifestation in the recent compressional structures of the sedimentary cover are illustrated.
Article
The integrated analysis of data on the heat flow and deep structure of the Earth's crust in the central Indian Ocean reveals additional heat sources (10-20 m W/m2) beneath zones of tectonic deformations in the Central, Wharton, and Arabian basins that are inconsistent with the geothermal model of lithosphere cooling. These sources may be related to serpentinization of ultramafic rocks. This is confirmed by the seismic Layer 3B with velocities of 7.2-7.6 km/s characteristic of serpentinite is traceable in the lower crust. In contrast to these basins, the heat-flow distribution in the Crozet Basin is consistent with the theoretical geothermal curve. The lack of an additional heat source beneath this basin implies that serpentinites are absent here, as is evident from DSS records (the lack of Layer 3B), and is in agreement with the low seismicity of the basin. Despite the appreciable difference in the geodynamic regime of the respective plates, the spreading rate decreased approximately twofold about 50 Ma ago in the Central and Arabian basins (Indian Plate) and in the Crozet Basin (Antarctic Plate).
Article
During ODP Leg 116, three sites were drilled on a transect across two tilted fault blocks in the region of central Indian Ocean intraplate deformation. Well logs and core analyses from these sites, when combined with site survey data, provide detailed information about the physical, chemical, hydraulic, and thermal properties of the 2 km thick sedimentary sequence. Within the underformed sediments, heat flow matches the 52 mW/m2 predicted by cooling plate models for 78 Ma crust. Within 5 km of a buried reversed fault, heat flow averages 84 mW/m2 and reaches values as high as 166 mW/m2. Evidence from downhole temperatures, interstitial water chemistry, and sediment diagenesis establishes the presence of active hydrothermal circulation near the buried fault. A mathematical model has been developed to study the nature and magnitude of this hydrothermal flow. -from Author
Article
Two major families of faults dominate the tectonics of slow spreading ridges: the transform faults paralleling the spreading direction and the 'GLORIA' faults, observed by large-scale side-scan sonar, normal to it. The observation of microearthquakes to 8 km depth beneath the Mid-Atlantic Ridge axis implies that not only the transform faults but the GLORIA faults as well penetrate into the mantle. Since large horizontal temperature gradients are often associated with them, both types of fault provide the means of circulating large volumes of sea water through the oceanic mantle. Under these conditions, serpentinization of upper mantle rocks will occur in a predictable way. Its influence on the tectonic development of slow spreading ridges is discussed.
Article
Kernel pattern consists of a rim of completely serpentinized peridotite surrounding a core of partly serpentinized, or unserpentinized, peridotite. It is found adjacent to shear zones in some serpentinites. Kernels are commonly rectangular, with length/width ratios of 1 to 3. The boundary between the completely serpentinized rim and the unserpentinized core is commonly cut at a high angle by cross-fractures. The cross-fractures, which contain either picrolite or chrysotile asbestos, taper out in the core and widen in the rim. They are almost always perpendicular to the core-rim boundary and point into the core of the kernel. In this sense they are radially distributed in the rim. The kernel pattern can be observed in thin section, in outcrop, and on the megascopic scale of tens of metres. Cross-fractures represent expansion fractures in serpentinite in the rim of the kernel caused by an increase in volume acompanying serpentinization of the peridotite in the core of the kernel. As the serpentinization front penetrates into the kernel, the serpentinized peridotite expands. However, the serpentinite in the rim cannot expand and must fracture to accommodate the increase in volume. As serpentinization proceeds, more expansion occurs, extending the fractures farther into the core and requiring the fractures in the existing serpentinite to widen. These fractures are subsequently filled by serpentine minerals. The resulting pattern consists of a rim of serpentinite, cut by serpentine veins, that surrounds a core of peridotite. Measured densities of cross-fractures in the rims of kernels are consistent with those expected from the magnitude of the volume change accompanying serpentinization of peridotite.
Article
Understanding the physical properties of ultramafic rocks is important for evaluating the wide variety of petrologic models for the Earth's upper mantle and lower oceanic crust. At comparable temperatures and pressures, velocities of compressional and shear waves in ultramafic rocks decrease with increasing serpentinization. A major factor affecting these velocities is the variety of serpentine present. Antigorite, the serpentine species stable at high temperatures, has higher velocities and a lower Poisson's ratio than the serpentine polymorphs lizardite and chrysotile. In addition, seismic properties of ultramafic rocks vary with their proportion of olivine to pyroxene and abundances of accessory minerals formed during serpentinization, such as brucite, magnetite, magnesite, tremolite, and talc. Seismic anisotropy is an important property of relatively unaltered peridotites. Large, well-exposed ultramafic massifs provide the best information on the nature of upper mantle compressional wave anisotropy and shear wave splitting. Average compressional wave anisotropy in these massifs is approximately 5%. Shear wave splitting magnitudes vary significantly with propagation direction. The major lithologies present in much of the lower ocean crust are metadiabase and gabbro. Lizardite-chrysotile-bearing serpentinites are abundant, however, in regions that have allowed penetration of sea water into the upper mantle. Dehydration of subducting slabs and the rising of released fluids have resulted in hydrothermal alteration and a lowering of velocities in forearc mantle wedges. In addition to serpentinization, high pore pressures and metamorphism producing chlorite are required to explain the seismic properties of the forearc upper mantle.