0% found this document useful (0 votes)
44 views

Rheology and Tectonic Signifi Cance of Serpentinite: Greg Hirth and Stéphane Guillot

1) Serpentinites play an important role in tectonic processes like continental rifting, oceanic spreading, and subduction due to their unique rheological properties. 2) Experimental studies show that serpentinite can deform through both brittle and ductile processes depending on pressure and temperature conditions. At low temperatures, dislocation glide allows ductile flow while microcracking leads to semibrittle behavior at higher strains. 3) The tendency for serpentinite to undergo localized strain during semibrittle deformation has implications for understanding fault stability during earthquakes and the extrapolation of lab data to natural settings.

Uploaded by

Rio Cendrajaya
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
44 views

Rheology and Tectonic Signifi Cance of Serpentinite: Greg Hirth and Stéphane Guillot

1) Serpentinites play an important role in tectonic processes like continental rifting, oceanic spreading, and subduction due to their unique rheological properties. 2) Experimental studies show that serpentinite can deform through both brittle and ductile processes depending on pressure and temperature conditions. At low temperatures, dislocation glide allows ductile flow while microcracking leads to semibrittle behavior at higher strains. 3) The tendency for serpentinite to undergo localized strain during semibrittle deformation has implications for understanding fault stability during earthquakes and the extrapolation of lab data to natural settings.

Uploaded by

Rio Cendrajaya
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 8

Rheology and Tectonic

Significance of Serpentinite Sheared serpentinite,


Oman

Greg Hirth1 and Stéphane Guillot2

1811-5209/13/0009-107$2.50 DOI: 10.2113/gselements.9.2.107

S
erpentinites occur in many active geologic settings and control the studies provide fundamental data
rheology of the lithosphere where aqueous fluids interact with ultra- for understanding fracture and
plastic flow, friction, and the
mafic rocks. The crystal structure of serpentine-group minerals results in role of dehydration reactions on
diagnostic physical properties that are important for interpreting a wide range fault rheology.
of geophysical data and impart unique rheological behaviors. Serpentinites
Fracture and Plastic Flow
play an important role during continental rifting and oceanic spreading, in
Both brittle and crystal plastic
strain localization along lithospheric strike-slip faults, and in subduction zone de for mat ion pro cesses a re
processes. The rheology of serpentine is key for understanding the nucleation strongly controlled by the aniso-
and propagation of earthquakes, and the relative weakness of serpentinite tropic properties of serpentine.
Dislocation glide (a deformation
can significantly affect geodynamic processes at tectonic plate boundaries. process where strain is accommo-
KEYWORDS : seismicity, ocean–continent transition, seafloor, strike-slip fault, dated by stress-driven motion of
subduction zone, serpentinization crystal defects) is easy along the
basal plane of serpentine minerals,
leading to ductility of serpentinite
at low temperatures. Nonetheless,
INTRODUCTION
experimental studies have produced apparently contra-
Plate tectonics is controlled by a combination of ductile dictory results regarding high-temperature flow behavior.
flow and frictional resistance in fault zones. At plate bound- Several studies indicate that serpentinites deform by macro-
aries, where deformation is localized along centimeter- to scopically ductile processes at high pressure (Hilairet et
kilometer-scale shear zones, the influence of serpentinite al. 2007 and references therein). In contrast, Chernak
on tectonic processes is linked to its unique rheological and Hirth (2010) observed localized faulting under high-
properties. While the rheological weakness of serpentinite temperature/high-pressure conditions, although these
relative to other rocks has been appreciated through field authors document macroscopically ductile flow at strains
observations in various geological settings, such as along less than ~0.1 to 0.2. These results can be reconciled by
slow-spreading oceanic ridges, along strike-slip faults, and considering the von Mises strain compatibility criterion,
in subduction zones, the quantification of its influence which states that five independent slip systems are required
on tectonic processes requires an understanding of the to accommodate homogeneous flow of a polycrystalline
factors that control ductile flow and frictional resistance. material. Dislocation glide on the basal plane and kinking
Experimental observations constrain these properties and can accommodate a limited amount of deformation, but
highlight the rheological behaviors that are important for to achieve higher strains additional processes are required.
understanding the role of serpentinization in strain local- Experimental studies on serpentinite (Chernak and Hirth
ization in the lithosphere, its role in both fault creep and 2010) indicate that microcracking initiates at moderate
dynamic fault rupture, and the influence of the dehydra- strains, leading to a semibrittle rheology for which strength
tion of serpentinite on the seismicity of subducting slabs remains pressure sensitive. Thus, while anisotropy in
and the exhumation of high- to ultrahigh-pressure rocks. plasticity facilitates dislocation glide at low temperatures,
this process is insufficient to promote fully crystal plastic
RHEOLOGICAL PROPERTIES deformation at high strain.
OF SERPENTINITE
Flow laws for antigorite (a high-pressure type of serpentine)
Since the pioneering work of Raleigh and Paterson (1965), are reported by Hilairet et al. (2007). Based on analysis
who documented how dehydration reactions in serpenti- of the stress exponent (n) and the ratio of strength to
nite result in weakening and strain localization, numerous confi ning pressure [(σ1 – σ3 )/σ3 ], they concluded that
experimental studies have been conducted to investigate samples at 4 GPa deformed by dislocation creep (where
the physical properties of the serpentine minerals. These the motion of crystal defects is thermally controlled by
diffusive processes) while samples at 1 GPa deformed by
dislocation glide. These data represent results of pioneering
high-pressure deformation experiments, though the strains
1 Department of Geological Sciences, Brown University in the experiments of Hilairet et al. are below those for
Providence, RI 02912, USA which semibrittle processes were observed in the study by
E-mail: [email protected] Chernak and Hirth (2010). The observation of a power-law
2 CNRS, Institut des Sciences de la Terre, Université de Grenoble relationship between stress and strain rate (i.e. ė ∝ σn with
1381 rue de la Piscine, 38041 Grenoble cedex, France n ≈ 3) suggests that deformation is limited by dislocation
E-mail: [email protected]

E LEMENTS , V OL . 9, PP. 107–113 107 A PR IL 2013


climb (the diffusive process that generally controls disloca- A
tion creep), which requires rapid diffusion of all compo-
nents of the crystal structure (i.e. Si, Mg, O, and H for
serpentine). However, this seems unlikely at the tempera-
ture (down to 200 oC) of these experiments. Nonetheless,
the high strain-rate sensitivity of serpentine deformation
has important implications for geologic processes.
The tendency for strain localization during semibrittle flow
(where deformation is accommodated by a combination of
brittle and ductile processes) of serpentine has important
implications for understanding the stability of fault slip
when an earthquake rupture propagates into a serpen-
tinized region. Constraining the conditions where strain
localization occurs is also important for extrapolating lab
data to natural conditions, owing to the trade-offs between
effective viscosity and shear zone width (e.g. Wada et al.
2008). Escartín et al. (1997) documented a transition from
localized to distributed deformation with an increase in
confi ning pressure from 200 to ~400 MPa at room temper-
ature. They inferred that the transition occurred when
the bulk sample strength became less than the frictional
B
strength. However, with increasing temperature at constant
pressure, a transition back to localized semibrittle deforma-
tion is observed (Chernak and Hirth 2010). The transition
back to localized semibrittle behavior at high temperature
suggests that the temperature dependence of the friction
coefficient is greater than the temperature dependence of
the bulk sample strength.
While the details of semibrittle flow remain unresolved, C
the observation of a deformation-induced alignment of
minerals [i.e. lattice-preferred orientation (LPO)] in both
experimental and natural samples of serpentinite (e.g.
Katayama et al. 2009; Bezacier et al. 2010) has important
implications for interpretating seismic anisotropy (FIG. 1).
Serpentine has a high compressibility normal to the basal
plane (parallel to the c-axis), which leads to a low isotropic
P-wave velocity (Vp) (e.g. Bezacier et al. 2010). Similarly,
the shear modulus (ratio of shear stress to shear strain)
parallel to the basal plane is very low. Thus, where a strong FIGURE 1 (A) Photomicrograph (crossed polarizers) of an antig-
LPO arises owing to deformation, the elastic anisotropy orite sample deformed in the lab at 300 oC and 1 GPa.
(B) Contoured, lower-hemisphere stereographic projections
leads to strong seismic anisotropy (Katayama et al. 2009). showing the lattice-preferred orientation (LPO) of antigorite with a
Nonetheless, the limited knowledge of elastic proper- [001] maximum oriented perpendicular to the shear plane (the
ties makes interpretation of seismic data nonunique. For shear plane pole is horizontal in the pole figure) and a maximum of
example, minor amounts of lizardite/chrysotile and/or LPO poles to (100) oriented parallel to the slip direction. (C) Anisotropic
velocity imparted by this LPO: (LEFT) contours of P-wave velocity;
can strongly impact the Vp /Vs ratio and the magnitude of ( MIDDLE) contours of % anisotropy of S-wave velocity; ( RIGHT) polar-
shear wave splitting (e.g. Bezacier et al. 2010). In addition, ization direction of shear waves. After Katayama et al. (2009)
the magnitude of shear wave splitting depends on both the
strength of LPO and the width of the anisotropic region
that waves pass through. Thus, if strain is highly local- general, theories for rock friction suggest that µ should
ized under geologic conditions, shear wave splitting will be relatively insensitive to temperature (e.g. Scholz 1990).
be negligible. In adhesion theory, µ is related to the ratio of the shear
strength of asperity contacts to the indentation hardness
Friction of the asperities (FIG. 2). A lack of temperature dependence
Examining the frictional behavior of fault materials for µ is inferred based on the notion that the processes that
over a range of sliding velocities is essential for under- control the shear strength of asperities are the same as
standing the dynamics of stress evolution and slip during those that control indentation hardness (e.g. Scholz 1990).
earthquakes. Phyllosilicate minerals are often inferred to However, when considering a strongly anisotropic material,
control the behavior of creeping sections of faults (e.g. another possibility arises. Assume that the basal planes of
Moore and Rymer 2007). Friction experiments on serpen- phyllosilicate grains at asperity contacts are dominantly
tinite document velocity-strengthening behavior (which aligned parallel to the fault surface. In this scenario, the
promotes aseismic fault slip) at tectonic displacement rates shear strength (e.g. controlled by dislocation glide) may
(e.g. Reinen et al. 1994), consistent with this hypothesis. be significantly lower than the indentation hardness (e.g.
Experiments also indicate a decrease in the friction coeffi- controlled by semibrittle flow). The temperature depen-
cient (µ) of antigorite with increasing temperature. Reinen dence of basal slip would be expected to be greater than
et al. (1994) measured µ = 0.5–0.85 for antigorite at room that for indentation hardness, resulting in temperature
temperature; Moore et al. (1996) determined µ in the dependence and a lower value for µ.
range of 0.4–0.6 for antigorite gouge at temperatures of High-velocity friction experiments on serpentinite
25–194 °C. Chernak and Hirth (2010) determined values document dramatic dynamic weakening and dehydration
of µ from 0.1 to 0.35 at temperatures of 400–550 °C. In at slip velocities above ~0.1 m/s (see Kohli et al. 2011 and

E LEMENTS 108 A PR IL 2013


(Jung et al. 2004) suggest that antigorite dehydration can
at least generate microseismicity. However, Chernak and
Hirth (2010) observed macroscopically ductile flow and
A velocity-strengthening behavior in dehydrated antigorite at
1.5 GPa and 700 °C, challenging the hypothesis that inter-
mediate-depth earthquakes result directly from dehydra-
tion embrittlement. Motivated by these results, Chernak
and Hirth (2011) conducted tests in which temperature
was ramped across the dehydration boundary, reasoning
B C that earthquake-like instabilities would be most likely if
samples dehydrated at high load under conditions where
strain was already localized. No frictional instabilities were
observed; rather, stress relaxed stably over several minutes,
providing experimental documentation of stable fault slip
during dehydration reactions. Similar interpretations were
made based on the lack of detectable acoustic emissions
during dehydration experiments on antigorite in a multi-
anvil device (Gasc et al. 2011). These observations suggest
Schematics of a fault surface (shown perpendicular that a mechanism other than dehydration may be respon-
FIGURE 2
to the fault plane) in serpentinite at different scales. sible for intermediate-depth earthquakes (see discussion
(A) The fault surface is rough, making contact only at asperities. and references in Chernak and Hirth 2011). One possibility
(B) The basal planes of serpentine grains (dashed lines) are aligned
parallel with the fault surface at asperities. (C) The shear deforma- is that the migration of fluids produced via dehydration
tion of asperities is accommodated by dislocation glide (as repre- promotes seismogenic failure in the surrounding unaltered
sented by symbols parallel to dashed lines). Grains are much rocks. Alternatively, earthquakes may nucleate via
stronger for deformation normal to the fault zone, resulting in a thermally induced viscous flow instabilities. In this case,
low, temperature-dependent coefficient of friction.
the fi ne-grained and polycrystalline products of dehydra-
tion reactions may promote the initial strain localization
references therein). These results indicate that a material required for the thermally induced viscous instability.
exhibiting velocity-strengthening behavior at tectonic
displacement velocities (~10 -9 m/s) may become dynami- GEOLOGICAL OBSERVATIONS
cally unstable at near-seismic slip velocities after only a
Since their discovery on the seafloor of the central Atlantic
few millimeters of displacement (FIG. 3). Such behavior
in the 1970s, serpentinites have been recognized worldwide
may be particularly important for understanding processes
in different active tectonic settings. Due to their unique
that control the dynamics of great subduction earthquakes,
physical properties, described above, serpentinites play a
which likely propagate across patches that creep during
major role in strain localization.
interseismic periods. When the asperity-contact lifetime
during frictional slip is short compared to the time it
The Ocean–Continent Transition
takes to diffuse heat away from asperities, the resulting
“flash heating” can produce dramatic weakening. Kohli and Slow-Spreading Ridges
et al.’s experiments indicate that weakening occurs at a Continental rifting and subsequent ocean-ridge spreading
critical weakening velocity (Vw), at which dehydration involve extensional faulting, exhumation of subconti-
of serpentine to talc occurs at asperities (500–700 o C). nental mantle, and magmatism, processes that reflect the
Based on extrapolation of Kohli et al.’s data, Vw is about tendency for localized deformation in the lithosphere.
25 mm/s at an ambient T = 300 °C. This analysis suggests
a possible explanation for the spectrum of fault-slip behav-
iors observed along fracture zones and subduction zones,
where the alteration of mantle peridotite produces serpen-
tinites. The effectiveness of dynamic weakening depends
strongly on the width of the actively deforming zone. Thus,
an understanding of how deformation localizes in serpen-
tinite during interseismic deformation remains critical for
constraining the significance of dynamic weakening.

Dehydration Embrittlement
Dehydration embrittlement (where fluids resulting from
dehydration reactions induce brittle deformation) has
become the accepted mechanism invoked to explain
intermediate-depth seismicity (e.g. Hacker et al. 2003).
Several issues are important for the application of these
results to higher pressures. First, the canonical experi-
ments were conducted at confi ning pressures where the
Clapeyron slope of the dehydration reaction is positive
(where dehydration results in a volume increase, promoting
microcracking) and the thermal stability of serpentine
is relatively modest (inhibiting plastic flow processes).
Second, the experiments did not show evidence for stick–
slip behavior (or velocity weakening), which is required to FIGURE 3 Influence of sliding velocity on the frictional behavior
of antigorite (after Kohli et al. 2011). At tectonic slip
nucleate earthquakes. Recent studies have extended these
rates, the behavior is velocity strengthening, resulting in creep.
results to higher pressure. Acoustic emissions during defor- However at seismic slip rates, samples exhibit dramatic dynamic
mation experiments at P = 1–6 GPa and T = 550–820 °C weakening, which could facilitate earthquake rupture through aseis-
mically creeping patches.

E LEMENTS 109 A PR IL 2013


The denudation of serpentinized mantle along the ocean– complexes (OCCs) (FIG. 5). The domes exhibit corrugated
continent transition (OCT) and by seafloor spreading elicits surfaces that are interpreted to be caused by exposed
the following questions: when and how is lithosphere detachment faults. OCCs are usually younger than 2
serpentinized and what role does serpentinization play Ma and, when active, show hydrothermal activity and
during rifting processes and ongoing slow-spreading-ridge moderate seismicity down to 3–4 km depth along the main
activity? detachment fault. OCCs cover areas of ~200 to 500 km 2,
extending ~25 km along the ridge axis and ~15 km perpen-
Along the western Iberia margin, the OCT forms a 100
dicular to the ridge axis. The corrugated surfaces include
km wide basin dominated by completely (95–100%) to
serpentinized peridotite and altered gabbro; some basalt
moderately (50%) serpentinized peridotite (e.g. Afilhado et
is observed close to the axial ridge (FIG. 5). Seismic studies
al. 2008). Gravimetric and seismic profiles indicate that the
show relatively high velocities (Vp > 7 km/s) within a 2 to
serpentinized mantle is about 5 km thick. Locally, tectonic
3 km thick zone parallel to the surface, suggesting that
breccia and semibrittle gouge composed of gabbro, amphib-
this zone is dominantly composed of a mixture of gabbro
olite, and serpentinite clasts in a matrix of calcite- and
and partially serpentinized peridotite (Karson et al. 2006).
chlorite-rich cataclasite are recovered during drilling. The
geochemical signatures of primary minerals indicate that Deformation is localized in a ~100 to 200 m thick shear
the serpentinized mantle is of depleted, subcontinental zone at the footwall of the detachment zone, and the rest
type. The oxygen isotope signature and the absence of of the massif is mostly undeformed (Karson et al. 2006). Far
antigorite in these rocks suggest temperatures less than from the shear zone, serpentinites are massive and sparsely
300 °C during fluid infi ltration (Skelton and Valley 2000). crosscut by joints. Partially serpentinized peridotites retain
Later infi ltration of cooler seawater (<100 °C) along brittle porphyroclastic textures acquired during high-tempera-
normal faults accompanied the seafloor exhumation of ture mantle deformation. Mesh-textured serpentinites
the mantle. Mohn et al. (2012) proposed the following show variable densities of crosscutting serpentine veins.
scenario of mantle exhumation: Crustal thinning is fi rst Within the detachment faults, the mineralogy is more
accommodated by pure shear along conjugate, crustal-scale complex; the presence of talc, amphibole, and chlorite
shear zones and by ductile flow of the middle crust (FIG. 4A). indicates the addition of silica, calcium, and aluminum
The upper mantle is progressively thinned and replaced from hydrothermal fluids derived through seawater–
by deeper, hotter mantle impregnated by mafic magmas gabbro interactions. Serpentinization along the detach-
(FIG. 4B). Faults cut from the surface into the mantle, and ment faults involves at least several stages of alteration
seawater infi ltrates down to the upper mantle, initiating and veining (e.g. Andreani et al. 2007). Early serpentine
moderate-temperature serpentinization
in the lizardite stability field (FIG. 4 C ).
Serpentinization weakens the upper A
mantle layer, promoting strain localiza-
tion and allowing the normal faults in
the distal margin to root at low angle. The
ongoing extension leads to buoyant flow
of partially serpentinized mantle along
low-angle detachment faults, resulting in
the horizontal stretching of the subconti-
nental mantle at the seafloor (FIG. 4D).
At slow- to ultraslow-spreading ridges B
(<40 mm/y), 20 to 25% of the seafloor is
composed of serpentinite (Cannat et al.
2009). These areas of outcropping serpen-
tinite are associated with domes, up to
500 meters high, known as oceanic core

C
FIGURE 4
Dynamic evolution of a rifted continental
margin up to the formation of the ocean–
continent transition (modified after Mohn
et al. 2012). UCC = upper continental crust;
MCC = middle continental crust; LCC = lower
continental crust; OCT = ocean–continent
transition; LT = low-temperature; HT = high-
temperature; dark green = lower oceanic
crust; blue to green = partially serpentinized
upper mantle; yellow = marine sediments

E LEMENTS 110 A PR IL 2013


mation in a closed chemical system. The
last generation of veins show a granular
texture formed by micron-scale lamellae
of Al-rich lizardite. These compositions
and microtextures require a change in
ambient conditions, with significant water
circulation at shallow depths (<2 km) and
temperatures of <200 °C.

Strike-Slip Faults
Large-scale strike-slip faults can either
slip by aseismic creep or move abruptly
during earthquakes. Weak fault behavior
can be caused by high fluid pressures,
locally high geothermal gradients, or
the presence of weak materials such as
serpentinite. The San Andreas Fault (SAF),
one of the most active strike-slip faults
on Earth, extends at least 15 km into the
crust and is marked at the surface by a
complex network of crushed and fractured
rocks a few hundred meters to several
(B OTTOM) Schematic 3-D block diagram of an oceanic kilometers wide (FIG. 6). Just north of the
FIGURE 5
core complex (OCC) (after Cannat et al. 2009). Parkfield segment, the SAF creeps at a rate of 28 mm/y.
(UPPER LEFT) Corrugated surface of an OCC (area shown is 10 × 5 km). Serpentinized ultramafic rocks derived from the Franciscan
(U PPER RIGHT) Photomicrograph (crossed polarizers) of a weakly Complex mélange have been associated with fault creep,
serpentinized sample of troctolite from IODP Hole U1309D
(Bar scale: 200 µm). PHOTO IODP, EXPEDITION 305 as well as low fault strength, in this region (e.g. Moore
and Rymer 2007). The SAFOD project (San Andreas Fault
Observatory at Depth) sampled and instrumented the SAF
veins are irregular and thin (<0.5 mm) and exploit preex- 9 km northwest of Parkfield. At about 3 km depth, the drill
isting cracks in olivine. The local presence of antigorite hole intersects the Great Valley Formation, a sedimentary
and Fe-rich brucite suggests that early serpentinization unit rich in serpentinites associated with the Coast Range
occurred at about 300–400 °C at a low water/rock ratio. ophiolite (e.g. Moore and Rymer 2007). Microstructural
Second-generation veins, composed dominantly of lizardite observations from core samples show evidence of defor-
or chrysotile and magnetite, are usually interconnected and mation across the damaged zone. The deforming zones
a few centimeters in length, and they form the primary contain serpentinite clasts and highly sheared siltstones.
network of serpentine veins. These veins are often oriented Moore and Rymer (2007) also reported talc in the cuttings.
parallel to the detachment fault, suggesting that they This discovery was noteworthy, as the frictional strength
opened during tectonic unroofi ng. The third generation of talc at elevated temperatures is low enough to meet the
of veins contain oblique synkinematic fibers that record constraints on the shear strength of the fault, and its stable
a shear component; the fibers are typically chrysotile and sliding behavior is consistent with fault creep.
polygonal serpentine and document incremental defor-

A B FIGURE 6 (A) Distribution


of serpentinite
outcrops in central California
and localization of major faults
with creeping (aseismic)
segments. SF, San Francisco;
SB, Santa Barbara; SJ, San Jose;
M, Monterey; P, Parkfield; C,
Cholame; O, SAFOD; ID, Idria;
SAF, San Andreas Fault; HF,
Hayward Fault; CF, Calaveras
Fault; SYF, Santa Ynez Fault;
LPF, Little Pine Fault.
(B) Geology of the SAFOD
Borehole with location of
cuttings (yellow circle)
containing serpentinite-derived
clay minerals. Red lines indicate
active fault strands.
(C) Scanning electron micro-
C scope image of gouge
showing microshear zones
in serpentinite-derived clay
minerals (SAFOD Photography
Atlas). Width of field of view =
2 mm

E LEMENTS 111 A PR IL 2013


Recent investigations on the main 200 m thick damage Indirect petrological evidence for serpentinization at
zone at ca. 3 km depth at SAFOD show that the creeping subduction zones is provided by the intimate association of
area consists of clasts of serpentinite and sedimentary rock serpentinites with eclogites (high- and ultrahigh-pressure
dispersed in a matrix of saponite (Mg-rich smectite clay), rocks). Guillot et al. (2009) estimated that almost 30%
which is one of the weakest phyllosilicates. Saponite is a of Phanerozoic eclogitic massifs worldwide are associated
low-temperature (<150 °C) product of metasomatic reactions with serpentinites. Evidence for prograde metamorphism
between the quartzofeldspathic wall rocks and serpenti- of serpentinite includes the transformation of lizardite to
nite blocks in the fault (e.g. Lockner et al. 2011). Thus, antigorite (FIG. 7) and the presence of metamorphic olivine
serpentinite-derived clay minerals alone appear sufficient and enstatite associated with Ti clinohumite, which reflects
to explain the low fault strength and creeping behavior of partial dehydration above 450 °C. The rheological proper-
the San Andreas Fault. However, considering the tempera- ties of serpentine may help explain the mechanical evolu-
ture stability field of saponite and the geothermal gradient tion of subduction zones. Single-sided subduction appears
within the drill hole, it is unlikely that this very weak to require a weak hydrated slab interface and high slab
mineral is present deeper than 4 km along the fault zone. strength (Gerya et al. 2008). Tectonic settings that favor
Thus, stable creep and low strength in the SAF down to serpentinization of strong mantle rocks near the surface
a minimum depth of 10 km must reflect the presence of (such as the OCT and fracture zones) are thus good candi-
other low-strength minerals (such as talc), elevated fluid dates for the location of subduction initiation. Serpentinites
pressure, or other deformation mechanisms (e.g. pressure can lubricate the nascent interplate fault zone, facilitating
solution) in serpentinites or quartzofeldspathic rocks. asymmetric plate movement. Similarly, during exhumation
of high-pressure or ultrahigh-pressure rocks, the opposite
Subduction Zones trajectories of exhumation and subduction require a decou-
Decoupling between the mantle wedge and the subducting pling zone within the subducting slab. A serpentinized
slab is inferred from modeling of heat flow data down to layer such as the OCT prior to subduction may become a
75 ± 15 km (Wada et al. 2008). The decoupled zone corre- decoupling zone between the oceanic crust and underlying
sponds to the stability of phyllosilicates such as serpentine, lithosphere. Moreover, the buoyancy of serpentinite likely
whose weakness and/or ductility can explain a rheological contributes to eclogite exhumation because the average
weakening of the interface (Hilairet et al. 2007; Wada et density of eclogite and serpentinite mélange is lower than
al. 2008). Hydrous silicates (e.g. serpentine, amphibole, that of anhydrous mantle peridotite.
talc, chlorite, lawsonite) destabilize at increasing pressure
and temperature in subduction zones and release water CONCLUSION
at different depths (Hacker et al. 2003). With ~13 wt% Our understanding of the rheological properties of
water in its structure, serpentine is an efficient carrier and serpentinite has advanced significantly during the last
source of water at depth during subduction (see Evans et al. ten years, and improved geophysical observations allow
2013 this issue). Dehydration of serpentine has also been detection of serpentinite bodies in active geological
hypothesized to play an important role in the origin of
double seismic zones, observed in some subduction zones,
and in the origin of intermediate and deep seismicity
along the slab (e.g. Yamasaki and Seno 2003). However, as
noted above, the exact mechanism by which dehydration
promotes seismicity remains a matter of debate. In addition,
high-resolution seismic tomography indicates that seismic
velocities in the lower seismicity plane are better explained
by seismic anisotropy of anhydrous peridotite than by the
presence of serpentinite (Reynard et al. 2010).
A low-velocity zone above the high-velocity anomalies of
a cold subducting slab has been imaged by seismic tomog-
raphy below the Cascadia forearc and is interpreted as
partially serpentinized mantle wedge (Bostock et al. 2002).
The hydrated subduction interface, called the serpentinite
channel, is more difficult to detect as it is probably thinner
than a few kilometers. Receiver functions show a slow layer
forming on top of the subducting plate below Japan at
depths between 80 and 140 km, a structure interpreted as
evidence for a serpentinite layer (Kawakatsu and Watada
2007). Bezacier et al. (2010) argued that serpentinite layers
with kilometer-size folds could result in an orientation-
independent anisotropy, with bulk anisotropy close to the
mean value of 0.05–0.06 s km−1. In that case, the observed
S-wave delay times (up to 1 s or more) would require a
20 km thick serpentinite layer.
Serpentinite mud volcanoes containing clasts of serpen-
tinized mantle peridotite in the Mariana forearc provide
direct evidence of serpentinites in subduction zones (e.g.
Fryer et al. 1999). The faults in the upper plate link mud
volcanism to the main decollement fault and provide a FIGURE 7 Schematic diagram illustrating the geological context
route for buoyant serpentine-rich mud. Both the uncon- during subduction-related serpentinization (COURTESY
OF F. D ESCHAMPS) and associated photomicrographs (COURTESY OF S.
solidated serpentine and the serpentinized peridotite clasts SCHWARTZ). With increasing depth, low-grade chryzotile-lizardite (Lz)
are dominated by chrysotile ± lizardite and antigorite that transforms into high-grade antigorite (Atg). Below 50 km depth,
formed at 350–450 °C and 600–700 MPa. antigorite breakdown releases fluids and metamorphic olivine
(Ol) crystallizes.

E LEMENTS 112 A PR IL 2013


settings. In spite of emerging consistency among laboratory, along crustal-scale strike-slip faults, in subduction zones,
geophysical, and geological observations, uncertainties and in oceanic environments.
remain concerning the rheological laws used to fit the
data and the extrapolation of the laws to natural strain ACKNOWLEDGMENTS
rates. Future work will focus on comparing experimental
We acknowledge funding from the National Science
deformation mechanisms with natural examples and on
Foundation (USA) and Labex OSUG @2020Carbon.
investigating the processes that control the brittle–ductile
Comments and suggestions by the editors and M. Cannat,
transition and strain localization. These properties are
D. Cluzel, G. Mohn, and J. Nakajima were helpful in
key for understanding the nucleation and propagation of
improving this article.
earthquakes and the significance of localized deformation

REFERENCES Funiciello F (eds) Subduction Zone Moore DE, Rymer MJ (2007) Talc-bearing
Dynamics. Springer-Verlag, Berlin, serpentinite and the creeping section
Afi lhado A, Matias L, Shiobara H, Hirn A, pp 175-204 of the San Andreas fault. Nature 448:
Mendes-Victor L, Shimamura H (2008) 795-797
From unthinned continent to ocean: Hacker BR, Abers GA, Peacock SM (2003)
The deep structure of the West Iberia Subduction factory: 1. Theoretical Moore DE, Lockner DA, Summers R,
passive continental margin at 38°N. mineralogy, densities, seismic waves Shengli MA, Byerlee JD (1996) Strength
Tectonophysics 458: 9-50 speed, and H 2O contents. Journal of chrysotile-serpentinite gouge
of Geophysical Research B 108: under hydrothermal conditions: Can
Andreani M, Mével C, Boullier A-M, doi:10.1029/2001JB001127 it explain a weak San Andreas fault?
Escartín J (2007) Dynamic control on Geology 24: 1041-1044
serpentine crystallization in veins: Hilairet N, Reynard B, Wang Y, Daniel
Constraints on hydration processes I, Merkel S, Nishiyama N, Petitgirard S Raleigh CB, Paterson MS (1965)
in oceanic peridotites. Geochemistry (2007) High-pressure creep of serpen- Experimental deformation of serpen-
Geophysics Geosystems 8: doi: tine, interseismic deformation, and tinite and its tectonic implications.
10.1029/2006GC001373, 24 pp initiation of subduction. Science 318: Journal of Geophysical Research 70:
1910-1913 3965-3985
Bezacier L, Reynard B, Bass JD, Sanchez-
Valle C, Van de Moortèle B (2010) Jung H, Green HW II, Dobrzhinetskaya LF Reinen LA, Weeks JD, Tullis TE (1994)
Elasticity of antigorite, seismic detec- (2004) Intermediate-depth earthquake The frictional behavior of lizardite and
tion of serpentinites, and anisotropy in faulting by dehydration embrittlement antigorite serpentinites: Experiments,
subduction zones. Earth and Planetary with negative volume change. Nature constitutive models, and implications
Science Letters 289: 198-208 428: 545-549 for natural faults. Pure and Applied
Geophysics 143: 317-358
Bostock MG, Hyndman RD, Rondenay S, Karson JA, Früh-Green GL, Kelley DS,
Peacock SM (2002) An inverted conti- Williams EA, Yoerger DR, Jakuba M Reynard B, Nakajima J, Kawakatsu H
nental Moho and serpentinization of (2006) Detachment shear zone of (2010) Earthquakes and plastic defor-
the forearc mantle. Nature 417: 536-538 the Atlantis Massif core complex, mation of anhydrous slab mantle
Mid-Atlantic Ridge, 30°N. Geochemistry in double Wadati-Benioff zones.
Cannat M, Sauter D, Escartín J, Lavier Geophysics Geosystems 7: doi: Geophysical Research Letters 37: doi:
L, Picazo L (2009) Oceanic corrugated 10.1029/2005GC001109 10.1029/2010GL045494
surfaces and the strength of the axial
lithosphere at slow spreading ridges. Katayama I, Hirauchi K-I, Michibayashi K, Scholz CH (1990) The Mechanics of
Earth and Planetary Science Letters Ando JI (2009) Trench-parallel anisot- Earthquakes and Faulting. Cambridge
288: 174-183 ropy produced by serpentine deforma- University Press, Cambridge, 471 pp
tion in the hydrated mantle wedge.
Chernak LJ, Hirth G (2010) Deformation Skelton ADL, Valley JW (2000) The
Nature 461: 1114-1117
of antigorite serpentinite at high relative timing of serpentinisation and
temperature and pressure. Earth and Kawakatsu H, Watada S (2007) Seismic mantle exhumation at the ocean–conti-
Planetary Science Letters 296: 23-33 evidence for deep-water transportation nent transition, Iberia: constraints from
in the mantle. Science 316: 1468-1471 oxygen isotopes. Earth and Planetary
Chernak LJ, Hirth G (2011) Syndeforma- Science Letters 178: 327-338
tional antigorite dehydration produces Kohli AH, Goldsby DL, Hirth G, Tullis
stable fault slip. Geology 39: 847-850 T (2011) Flash weakening of serpenti- Wada I, Wang K, He J, Hyndman RD
nite at near-seismic slip rates. Journal (2008) Weakening of the subduction
Escartín J, Hirth G, Evans B (1997) of Geophysical Research B 116: doi: interface and its effects on surface heat
Nondilatant brittle deformation 10.1029/2010JB007833 flow, slab dehydration, and mantle
of serpentinites: Implications for wedge serpentinization. Journal of
Mohr-Coulomb theory and the strength Lockner DA, Morrow C, Moore D,
Geophysical Research B 113: doi:
of faults. Journal of Geophysical Hickman S (2011) Low strength of deep
10.1029/2007JB005190
Research B 102: 2897-2913 San Andreas fault gouge from SAFOD
core. Nature 472: 82-85 Yamasaki T, Seno T (2003)
Evans BW, Hattori K, Baronnet A (2013) Double seismic zone and dehydration
Serpentinites: What, why, where? Mohn G, Manatschal G, Beltrando M,
embrittlement of the subducting slab.
Elements 9: 99-106 Masini E, Kusznir N (2012) Necking
Journal of Geophysical Research B 108,
of continental crust in magma-poor
Fryer P, Wheat CG, Mottl MJ (1999) doi: 10.1029/2002JB001918
rifted margins: Evidence from the
Mariana blueschist mud volcanism: fossil Alpine Tethys
Implications for conditions within the margins. Tectonics
subduction zone. Geology 27: 103-106 31: doi: 10.1029/
Gasc J, Schnubnel A, Brunet F, Guillon 2011TC002961
S, Mueller H-J, Lathe C (2011)
Simultaneous acoustic emissions
monitoring and synchrotron X-ray
diffraction at high pressure and temper-
ature: Calibration and application to
PFA LABWARE
serpentinite dehydration. Physics of Essential for
the Earth and Planetary Interiors 189: ultra trace analysis
121-133 in geochemistry
Gerya TV, Connolly JAD, Yuen DA (2008)
Why is terrestrial subduction one-sided?
Geology 36: 43-46
Guillot S, Hattori KH, Agard P, Schwartz
S, Vidal O (2009) Exhumation processes
in oceanic and continental subduction
contexts: a review. In: Lallemand S,
OUR EXPERIENCE … YOUR PROFIT! www.ahf.de :: [email protected]

E LEMENTS 113 A PR IL 2013

You might also like