You are on page 1of 4

Quartz vein formation by local dehydration embrittlement along the

deep, tremorgenic subduction thrust interface


Åke Fagereng1,2, Johann F.A. Diener2, Francesca Meneghini3, Chris Harris2, and Ada Kvadsheim1
1
School of Earth and Ocean Sciences, Cardiff University, Cardiff CF10 3AT, UK
2
Department of Geological Sciences, University of Cape Town, Private Bag X3, Rondebosch 7701, South Africa
3
Earth Science Department, University of Pisa, Pisa 56100, Italy

ABSTRACT of the Khomas Sea underneath the Congo cra-


Hydrothermal quartz veins are ubiquitous in exhumed accretionary complexes, including ton from ca. 580 Ma to 540 Ma (Miller, 1983;
the Namibian Damara belt. Here, subduction-related deformation occurred at temperatures Meneghini et al., 2017). The high-T, low-P Cen-
≤550 °C, and vein geometry is consistent with plate interface shear, low effective normal tral Zone (CZ) is interpreted as the volcanic
stresses, and mixed-mode deformation. Quartz vein δ18O values relative to Standard Mean arc, whereas the medium-T, medium-P South-
Ocean Water (SMOW) range from 9.4‰ to 17.9‰ (n = 30), consistent with precipitation from ern Zone (SZ) comprises metaturbidites and
metamorphic fluids. A dominant subset of quartz veins away from long-lived high-strain zones minor metabasites of mid-oceanic ridge basalt
and basaltic slivers have δ18O values in a smaller range of 14.9‰ ± 1‰, requiring precipita- (MORB) affinity, interpreted as an accretion-
tion from a fluid with δ18O of 12‰ ± 1‰ at 470–550 °C. This uniform fluid isotope value is ary prism (Figs. 1A–1C) (Gray et al., 2007;
consistent with progressive local breakdown of chlorite allowing extensive hydrofracture at Meneghini et al., 2017). A foliation formed at
temperatures typical of the plastic regime. In active subduction zones, brittle deformation
within the plastic regime is inferred from observations of tectonic tremor, a noise-like seismic Congo
A B
signal including overlapping low- and very low-frequency earthquakes, which occurs below e craton
on

Ka
the seismogenic zone. Both tremor and hydrothermal veins correlate with zones of inferred N rn
Z a
ar

ok
rt he m

oB
high fluid pressure, could represent a mixture of shear and dilatant failure, and may therefore No Da

elt
e
be controlled by episodic hydrofracturing within a dominantly plastic and aseismic regime. 100 km on Kalahari
a lZ craton
ntr
Ce
Continental
INTRODUCTION The tremor signal includes low- and very low- basement SZ
Subducting sediments and oceanic crust expe- frequency earthquakes with focal mechanisms rocks D Z
SM
rience increasing pressure (P) and temperature indicating megathrust shear displacement (Ito et B’
(T), triggering metamorphic dehydration reac- al., 2007; Shelly et al., 2007). However, shear B (NW) Northern Central SMZ B’(SE)
Zone SZ
tions. It has been hypothesized that this dehydra- displacement in low-frequency earthquakes can- Zone
tion creates extreme overpressures at the base not always explain the full tremor signal (Frank Congo Kalahari
craton craton
of the subduction megathrust seismogenic zone et al., 2014), which is also comparable to acoustic Present day
(Saffer and Tobin, 2011). A consequence would emissions observed during laboratory dehydra- C NW Central SZ prism Incoming SMZ SE
Zone
be a localized zone of fluid-assisted deformation tion experiments (Burlini et al., 2009). Kalahari
Congo melts craton
and, depending on stress conditions, formation of Both tectonic veins and tectonic tremor are craton r e Underplated,
e
t h o sph dehydrating sediments =
a hydrothermal vein system (e.g., Yardley, 1983). hypothesized to result from fracture, healing li
anic veining and tremor
Oce ca. 550 Ma
Here, we test whether an intense regionally devel- by mineral precipitation, buildup of fluid pres- D Okahandja Lineament
oped vein system is consistent with fracture and sure, and near-periodic repetition of this cycle 18 19 MV
vein growth in a relatively small P-T range related (Yardley, 1983; Audet and Bürgmann, 2014). 17 16
15
to prograde metamorphism. Our hypothesis test Quartz veins can form incrementally at length
14
considers the geometry, microstructure, meta- scales consistent with small stress drops in low- ab
12 11 10 9 8 om ine
G L
morphic conditions, and quartz oxygen isotope frequency earthquakes (Fagereng et al., 2011), 13 te 7 iver
b oli R
ratios of syntectonic veins throughout the nearly and heal on time scales less than slow earth- h i
mp GB
80 km across-strike width of the exhumed accre- quake repeat times of weeks to months (Fisher s sA
h le Kalahari
tionary complex of the Namibian Damara belt. and Brantley, 2014). Hayman and Lavier (2014) tc 1-6 craton
Ma
A fossil vein system formed at the base of further calculated that local brittle failure within
the megathrust seismogenic zone should repre- the plastic regime, as tremor represents, results Figure 1. A: Simplified map of the inland branch
sent deformation recorded at similar conditions in periodic slow slip events as commonly asso- of the Damara belt, Namibia, after Miller (1983).
in active subduction zones, where geophysical ciated with tremor. We therefore hypothesize SZ—Southern Zone; SMZ—Southern Marginal
Zone. B: Schematic cross section along line
observations indicate elevated fluid pressures that syntectonic quartz veins within plastically
B-B′ shown in A, modified from Gray et al.
(e.g., Shelly et al., 2006). Tectonic tremor has been deformed rocks may be a record of tremor, and (2007). C: Reconstruction of margin during its
reported within such inferred fluid overpressured discuss implications of this inference for fluid- subduction phase (modified after Meneghini et
zones in most well-instrumented subduction mar- mechanical processes at the tremor source. al., 2017), where dotted yellow arrows refer to
gins (Beroza and Ide, 2011). Tremor is defined as particle paths in the prism (Moore et al., 2007).
D: Sample locations and boundaries referred
a weak, persistent, low-frequency (2–8 Hz) seis- GEOLOGICAL SETTING to in the text, depicted on Landsat satellite
mic signal that lasts for several days and repeats at The northeast-striking arm of the Damara belt imagery. Numbers refer to sample numbers
regular intervals of several months (Obara, 2002). formed during northwest-directed subduction starting with KV in Table DR1 (see footnote 1).

GEOLOGY, January 2018; v. 46; no. 1; p. 67–70  |  Data Repository item 2018015  | https://doi.org/10.1130/G39649.1 |  Published online 29 November 2017
GEOLOGY 
© |  VolumeGold
2017 The Authors. 46  |Open
  Number 1  | www.gsapubs.org
Access: This paper is published under the terms of the CC-BY license. 67
P ~1 GPa and 540 °C < T < 560 °C is consis- A NW B deformed veins, and foliation-oblique veins in
tent with top-to-the-southeast kinematics and mutually crosscutting relationship with the foli-
inferred to record peak subduction-related P-T ation-parallel veins (Fig. 2C), to have formed at
conditions (Cross et al., 2015). the same time as the tectonic, subduction-related
The Okahandja Lineament forms the north- foliation. Accepting the hypothesis that accre-
west boundary between the SZ and the CZ, 20 cm tionary prisms form by successive underplating
whereas the Gomab River Line separates the SZ C NW D NW of packages hundreds of meters in thickness
from the Southern Marginal Zone (SMZ) in the (Kimura and Ludden, 1995), these veins repre-
p
southwest (Fig. 1D). The Matchless Amphibo- sent top-to-the-southeast, dilatant shear within
o
lite (MA) comprises tectonically imbricated or close to the subduction thrust interface (Figs.
metamafic rocks and metaturbidites, present at p 1C, 2A, and 2C). Locally, e.g., along the Okah-
40 cm 0.5 m
an approximately constant tectonostratigraphic andja Lineament, veined rock assemblages are
level in the SZ for >350 km along strike (Miller, E F deformed by later faults related to exhumation
1983). Meneghini et al. (2017) described the MA within the prism.
in detail and interpreted it as an imbricated ridge
from the subducted Khomas Ocean. The SMZ is METHODS
a zone of imbricate metasediments, metamafic 200 µm 100 µm Analyses of quartz veins were performed
slivers, and calc-silicate horizons, interpreted as at the University of Cape Town (South Africa)
the passive margin sequence on the Kalahari cra- Figure 2. Photographs of quartz veins in using the laser fluorination method of Harris
ton, accreted before the transition from oceanic Kuiseb Schist (Damara belt, Namibia). White and Vogeli (2010), where 2–3 mg clean quartz
subduction to continental collision (Miller, 1983). arrows in panels A, C, and D show inferred chips were reacted with BrF5 and collected as O2.
sense of shear. A: Typical outcrop appearance
We consider a transect from the southern of folded and boudinaged, foliation-parallel Raw data were converted to δ-notation relative
edge of the SMZ through to the Okahandja Lin- veins. B: Isoclinal fold in vein parallel to, to standard mean ocean water (SMOW) based
eament (Fig. 1D): an 80 km horizontal distance and crosscut by, greenschist facies foliation. on an internal garnet standard (MON GT; δ18O =
through rocks dipping on average 45° northwest, C: Mutually crosscutting foliation-parallel 5.38‰). The long-term difference in δ18O values
(p) and foliation-oblique (o) veins; foliation
equivalent to a true thickness of ~55 km neglect- of two MON GT standards run with each batch
traces in white dashed lines. D: Boudinaged
ing any structural repetition. This represents the metamafic lens and quartz vein. E: Photomi- of 10 samples is 0.12‰ (n = 216), correspond-
final thickness of an accretionary prism, which crograph showing serrated vein quartz grain ing to a 2σ value of 0.15‰.
typically grows by successive underplating boundaries and inclusion trails across these Assuming burial along a profile shaped as
of subducted sediments and slices of oceanic boundaries (sample KV18). F: Photomicro- predicted by the analytical thermal model of
graph of vein quartz grain boundary with
crust that preserve structures formed along the recrystallized grains (white arrows) (sample Molnar and England (1990), and intersecting
megathrust (Kimura and Ludden, 1995; Moore GBLC). Both photomicrographs were taken in metamorphic P-T conditions reported for the
et al., 2007) (Fig. 1C). In the SZ, Meneghini et cross-polarized light. Damara accretionary prism (Cross et al., 2015),
al. (2017) recently demonstrated preservation of we assess the fluid release from an underthrust
subduction-related structures that escaped over- migration microstructures, with rare evidence package during subduction (Fig. 3). Fluid pro-
print by exhumation, which in the Damara belt for subgrain and bulging recrystallization, is typ- duction is calculated for a metapelite, the volu-
was largely accommodated on discrete struc- ical of natural quartz deformation at T > 500 °C metrically dominant lithology within the SZ and
tures such as the Okahandja Lineament. (Stipp et al., 2002). Solid and fluid inclusion SMZ (Fig. 3A), and the mafic component of the
bands are common, with variable spacing and MA (Fig. 3B). We use THERMOCALC version
HYDROTHERMAL VEINS orientation, although a dominant orientation can 3.45 (http://www.metamorph.geo.uni-mainz.de/
Ubiquitous hydrothermal quartz veins be determined locally (Figs. 2E and 2F). thermocalc/), with the thermodynamic data set
throughout the SZ and the SMZ are predomi- Following Fagereng et al. (2014), we interpret of Holland and Powell (2011; data set 6.2, cre-
nantly subparallel to the moderately northwest- the ubiquitous foliation-parallel and plastically ated 6 February 2012) and activity-composition
dipping foliation (Fig. 2A). The veins are com-
monly folded by tight to isoclinal, asymmetric
folds, with stretched, variably boudinaged limbs A Metapelite (+ mu + q + H2O) B Amphibolite (+ bi + ep + q + H2O) C 600
Temperature ( C) along burial path

and locally isolated hinges (Fig. 2B). Hinge sur-


g chl
g chl

g hb ru
ep ilm

g chl g bi chl ilm


1.2 1.2
8

ep sph g hb sph ru
faces are roughly foliation-parallel, and veins gl act
ep sp

bi ilm

peak T
12

chl sph 550


& vein growth

both crosscut and deflect the foliation (Figs. 2A


10
Pressure (GPa)

9
h ilm

Pressure (GPa)

Dehydration
g st

1.0 1.0
e
sph ru

hb sph ru
olit

and 2B). In places, foliation-parallel veins and


11
10
12

hib

chl ep 14.2
11

9–7

hb chl
sp 9

foliation-oblique veins are mutually crosscut- 500


h

sph 6
p

0.8 g bi chl 0.8


am
l
14

ch

ep ilm
ting and accommodate both shear and dilation hb sph
13

ilm
8 hb

14.5 12.3
l ab sp

pl
metapelite

(Fig. 2C). Quartz veins also occur in boudin bi 5 minor fluid


0.6 st 0.6 act chl sp
h 450
7

g ab sph pl production;
necks of plastically deformed, competent lenses
ch

hb
7
11–

minor vein
hb act

bi chl
6

g sill
of metachert, quartzite, or metabasite, and in 0.4 ep sph bi pl ilm 0.4 hb pl sph [ep] formation
400
places the veined boudin neck is itself boudi- 400 500 600 400 500 600 15 10 5
naged (Fig. 2D). Temperature ( C) ˚ Temperature ( C) ˚ Fluid content (vol %)

The veins are blocky at the mesoscale, and Figure 3. A,B: Pseudosections for metapelite (A) and amphibolite (B) compositions, contoured
at the microscale comprise quartz crystals of for volume percent fluid held in mineral assemblage (thin dashed lines). C: Fluid production
millimeters to centimeters in diameter (Fig. 2E). along subduction burial path, shown as thick dashed arrow in A and B, with very limited fluid
release below temperatures of 470 °C, followed by punctuated and steady fluid release to peak
Grain boundaries are sutured, and local sub- temperatures (T). Mineral abbreviations: ab—albite; act—actinolite; bi—biotite; chl—chlorite;
grains occur along grain boundaries (Figs. 2E ep—epidote; g—garnet; gl—glaucophane; hb—hornblende; ilm—ilmenite; mu—muscovite;
and 2F). The dominance of grain boundary pl—plagioclase; q—quartz; ru—rutile; sill—sillimanite; sph—titanite; st—staurolite.

|  Volume 46  |  Number 1  |  GEOLOGY


68 www.gsapubs.org 
relations of White et al. (2014) and Green et Data Repository1). Throughout the SZ, quartz values are also lower than in the SZ. This can
al. (2016). vein δ18O values are within a 2.0‰ range from partially be explained by local presence of
13.9‰ to 15.9‰ (n = 13). The MA veins show metabasalt; in addition, the range of passive
METAMORPHIC FLUID RELEASE lower quartz δ18O values, ranging from 9.4‰ margin sediments would also have provided a
Along the estimated P-T path, very little to 15.4‰ (n = 8); also, the SMZ has a larger range of fluid compositions. Higher quartz vein
fluid is released at T < 470 °C, as all low-grade range in quartz vein δ18O values, from 10.9‰ δ18O values in the Okahandja Lineament can be
hydrous minerals remain stable (Figs. 3A–3C). to 16.4‰ (n = 8). A sample from the Okahandja explained by larger ∆qtz-H O as deformation and
2
At 470 °C, biotite is introduced in the metapelite, Lineament has δ18O of 17.9‰, the highest value fluid flow persisted to lower temperatures in this
and hornblende becomes stable in the amphibo- in this study. The difference between samples of high-strain zone.
lite (Figs. 3A and 3B). In both cases, new min- mutually crosscutting veins oblique and subpar- In summary, the majority of quartz vein δ18O
eral growth consumes chlorite. Fluid production allel to schistose foliation is <1.0‰ (Table DR1). values are consistent with precipitation at 470–
in the metapelite is continuous but gradual as Vein δ18O is controlled by precipitation T 550 °C from locally released metamorphic fluids
chlorite is consumed with increasing T to the and the δ18O of the fluid from which quartz (Fig. 3). To be consistent in composition across
reported peak conditions of 550 °C, with ~2.2 precipitated. The difference between quartz the prism, this fluid release likely occurred
vol% fluid released over this section of the P-T and water δ18O values (∆qtz-H O) approximates within subducting sediments along the plate
2
path (Fig. 3C). The amphibolite initially experi- a × (106 T–2) – b, where a and b are empiri- interface, before incorporation into the prism
ences voluminous dehydration and fluid release, cal constants (determined by Matsuhisa et al., by underplating and downward migration of the
with ~2 vol% fluid produced at 470–475 °C as 1979). From the metamorphic constraints (Fig. megathrust (Fig. 1C).
actinolite and albite are exhausted (Fig. 3B), 3), we explore the T window from 470 °C to
before continued consumption of chlorite yields 550 °C, where the T-dependent ∆qtz-H O values are DEHYDRATION, DEFORMATION, AND
2
a further 3 vol% fluid as 550 °C is approached. 2.7‰ to 1.9‰, respectively (Fig. 4B). Within TECTONIC TREMOR
If the rocks are sufficiently impermeable to the SZ, the inferred fluid δ18O is 11‰–13‰, At the vein-forming conditions, metamor-
prevent flow along, into, or out of the fault zone, whereas measured quartz δ18O in the SMZ and phic fluids were released into low-permeability
the calculations predict few veins to form below MA require fluid δ18O as low as 8‰ and 6.5‰ rocks where fluid overpressure gradually devel-
470 °C, whereas numerous veins form as the respectively, if precipitation occurred at 470 °C oped until hydrofracture could occur (Yardley,
rocks are heated from 470 to 550 °C (Fig. 3C). (Fig. 4A). The elevated quartz δ18O in the Okah- 1983). That the dominant vein orientation is
This is consistent with the observation of plas- andja Lineament requires a fluid δ18O of 15‰– parallel to foliation (Figs. 2A–2D) implies the
tically deformed veins crosscutting a foliation 16‰ or precipitation at lower T. tensile strength along foliation was exceeded
defined by minerals stable at T > 470 °C (Fig. 2) before the tensile strength of intact rock. Excep-
VEIN-FORMING CONDITIONS tions to this are foliation-oblique veins associ-
VEIN AND FLUID OXYGEN ISOTOPE The value of ∆qtz-H O decreases with increas- ated with mixed-mode dilational shear failure,
2
COMPOSITIONS ing T (Fig. 4B); therefore, if seawater (δ18O = e.g., where foliation is deflected with a reverse
Vein quartz δ18O values vary from 9.4‰ to 0‰ ± 2‰) is trapped in subducting sediments sense of shear along a dilatant fracture (Fig. 2C).
17.9‰ (n = 30) (Fig. 4A; Table DR1 in the GSA and released gradually with depth, quartz precip- Thus, the vein system accommodated both ten-
itated from this pore fluid tracks a gradient from sile failure, requiring fluid pressure in excess
high to low values with increasing T of precipi- of the least compressive stress, and dilational
tation. O’Hara et al. (1997) found quartz δ18O shear, which could have produced space and
A SE NW
values decreasing from 17.8‰ to 3.4‰ with pressure drops for quartz precipitation without
Okahandja Lineament

16 increasing metamorphic grade along a transect requiring hydrofracture conditions (Lewis and
Gomab River Line

through the Altyn Tagh accretionary complex, Byrne, 2003). Foliation-parallel shear failure
δ18O (‰)

12
SMZ MA SZ China, and interpreted these veins as formed may also well have occurred, but would be near
8
δ18Oqtz progressively from pore fluid release. In the SZ, impossible to recognize given lack of marker
4 δ OH2O550˚C
18
such a gradient is not observed. Instead, calcu- horizons (Fagereng et al., 2014).
MA

δ18OH2O470˚C
lated fluid δ18O has a narrow range of 12‰ ± 1‰ In active subduction zones, tectonic tremor
0 10 20 30 40 50 60 70 80 (Fig. 4), similar to 13‰ ± 1‰ calculated for flu- events repeat within a particular depth range
Distance (km)
B 20 ids from which quartz precipitated across blue- (e.g., Beroza and Ide, 2011), which may relate
vein-forming

schist- to amphibolite-facies units in the Catalina to metamorphic dehydration reactions whose


∆qtz-H2O (‰)

Schist (Santa Catalina Island, California; Bebout, depth depends on the local geothermal gradient
Inferred
T range

10
1991). These values are consistent with precipi- (Fagereng and Diener, 2011). Local dehydration
2.74‰
1.89‰ tation from fluids in equilibrium with trench and at >>350 °C creates intensive brittle deforma-
0
100 200 300 400 500 600 ocean-floor metasediments (10‰–20‰; Savin tion in the plastic regime at a location depen-
˚
Temperature ( C) and Epstein, 1970), at a T where ∆qtz-H O is small.
2
dent on local thermal structure, not necessarily
Quartz vein δ18O values from the MA are coincident with the frictional-plastic transi-
Figure 4. A: Quartz vein δ18O values as function lower than in the metasedimentary SZ. Basalts tion zone (Gao and Wang, 2017), but critically
of across-strike distance from the structural have lower bulk-rock δ18O than shale (< 8‰; dependent on where dehydration and silicifica-
base of the Southern Marginal Zone, Damara Muehlenbachs, 1986); relatively low quartz-vein tion occur (Audet and Bürgmann, 2014; Hynd-
belt, Namibia (Fig. 1). Calculated fluid δ18O
values, also plotted against distance, are δ18O is therefore expected where fluid δ18O is man et al., 2015). We suggest that the Damara
based on fluid release temperatures of 470– buffered by host metabasites. In the SMZ, δ18O quartz veins record silicification of subducting
550 °C. B: Quartz-water fractionation factor, sediments, which may also have promoted their
after Matsuhisa et al. (1979) as a function of underplating and preservation by hardening of
temperature, with values used for calculations GSA Data Repository item 2018015, Table DR1
1 

in A indicated by gray shading. SMZ— (sample locations and oxygen isotope values), is avail- the underthrust plastically weak, fine-grained
Southern Marginal Zone; MA—Matchless able online at http://www.geosociety.org​/datarepository​ assemblage. If the vein-forming process is anal-
amphibolite; SZ—Southern Zone; qtz—quartz. /2018/ or on request from editing@geosociety.org. ogous to tremor, the noisy part of the tremor

GEOLOGY  |  Volume 46  |  Number 1  | www.gsapubs.org 69


signal may represent tensile hydrofracture and Guerrero, Mexico: Journal of Geophysical Re- Molnar, P., and England, P., 1990, Temperatures, heat
vein formation in intact rock, whereas low- and search: Solid Earth, v. 119, p. 7686–7700, https://​ flux, and frictional stress near major thrust faults:
doi​.org​/10.1002​/2014JB011457​. Journal of Geophysical Research, v. 95, p. 4833–
very low-frequency earthquakes are associated Gao, X., and Wang, K., 2017, Rheological separa- 4856, https://​doi​.org​/10​.1029​/JB095iB04p04833.
with low-effective-stress shear failure as occurs tion of the megathrust seismogenic zone and epi- Moore, J.C., Rowe, C., and Meneghini, F., 2007, How
along fluid-overpressured, vein-coated, weak sodic tremor and slip: Nature, v. 543, p. 416–419, accretionary prisms elucidate seismogenesis in
foliation planes (Fagereng et al., 2011). In this https://​doi​.org​/10​.1038​/nature21389. subduction zones, in Dixon, T.H., and Moore,
Gray, D.R., Foster, D.A., Maas, R., Spaggiari, C.V., J.C., eds., The Seismogenic Zone of Subduction
interpretation of the tremor source, the regular Gregory, R.T., Goscombe, B., and Hoffman, K.H., Thrust Faults: New York, Columbia Univer-
repeat time of episodic tremor and slow slip 2007, Continental growth and recycling by ac- sity Press, p. 288–315, https://​doi​.org​/10​.7312​
events may arise from a regular cycle of fluid cretion of deformed turbidite fans and remnant /dixo13866​-010.
pressure–driven failure—where a combination ocean basins: Examples from Neoproterozoic and Muehlenbachs, K., 1986, Alteration of oceanic crust
of fracture healing and fluid production rates Phanerozoic orogens, in Hatcher, R.D., Jr., et al., and the 18O history of seawater, in Valley, J.W.,
eds., 4-D Framework of Continental Crust: Geo- et al., eds., Stable Isotopes in High Temperature
determines repeat times (Audet and Bürgmann, logical Society of America Memoir 200, p. 63–92, Geological Processes: Washington, D.C., Miner-
2014; Fisher and Brantley, 2014). https://​doi​.org​/10​.1130​/2007​.1200​(05). alogical Society of America, Reviews in Mineral-
Green, E.C.R., White, R.W., Diener, J.F.A., Powell, ogy, v. 16, p. 425–444.
ACKNOWLEDGMENTS R., Holland, T.J.B., and Palin, R.M., 2016, Ac- Obara, K., 2002, Nonvolcanic deep tremor associ-
Fieldwork and isotope analyses were supported by Uni- tivity–composition relations for the calculation ated with subduction in southwest Japan: Sci-
versity of Cape Town Research Development grants to of partial melting equilibria in metabasic rocks: ence, v. 296, p. 1679–1681, https://​doi​.org​/10​
Fagereng and Diener, and a Claude Leon Foundation Journal of Metamorphic Geology, v. 34, p. 845– .1126​/science​.1070378.
grant to Meneghini at Stellenbosch University (South 869, https://​doi​.org​/10​.1111​/jmg​.12211. O’Hara, K.D., Yang, X.Y., Guoyuan, X., and Li, Z.,
Africa). Fagereng is supported by European Research Harris, C., and Vogeli, J., 2010, Oxygen isotope com- 1997, Regional δ18O gradients and fluid-rock in-
Council starting grant 715836 “MICA”. We thank J.C. position of garnet in the Peninsula Granite, Cape teraction in the Altay accretionary complex, north-
Lewis and N. Hayman for critical comments that sig- Granite Suite, South Africa: Constraints on melt- west China: Geology, v. 25, p. 443–446, https://​
nificantly improved the manuscript. ing and emplacement mechanisms: South African doi​.org​/10​.1130​/0091​-7613​(1997)025​<0443:​
Journal of Geology, v. 113, p. 401–412, https://​ ROGAFR>2​.3​.CO;2.
REFERENCES CITED doi​.org​/10​.2113​/gssajg​.113​.4​.401. Saffer, D.M., and Tobin, H.J., 2011, Hydrogeology
Audet, P., and Bürgmann, R., 2014, Possible control Hayman, N., and Lavier, L.L., 2014, The geological and mechanics of subduction zone forearcs: Fluid
of subduction zone slow-earthquake periodicity record of deep episodic tremor and slip: Geol- flow and pore pressure: Annual Review of Earth
by silica enrichment: Nature, v. 510, p. 389–392, ogy, v. 42, p. 195–198, https://​doi​.org​/10​.1130​ and Planetary Sciences, v. 39, p. 157–186, https://​
https://​doi​.org​/10​.1038​/nature13391. /G34990​.1. doi​.org​/10​.1146​/annurev​-earth​-040610​-133408.
Bebout, G.E., 1991, Field-based evidence for devola- Holland, T.J.B., and Powell, R., 2011, An improved Savin, S.M., and Epstein, S., 1970, The oxygen and
tilization in subduction zones: Implications for and extended internally consistent thermodynamic hydrogen isotope geochemistry of ocean sedi-
arc magmatism: Science, v. 251, p. 413–416, dataset for phases of petrological interest, involv- ments and shales: Geochimica et Cosmochim-
https://​doi​.org​/10​.1126​/science​.251​.4992​.413. ing a new equation of state for solids: Journal of ica Acta, v. 34, p. 43–63, https://​doi​.org​/10​.1016​
Beroza, G.C., and Ide, S., 2011, Slow earthquakes and Metamorphic Geology, v. 29, p. 333–383, https://​ /0016​-7037​(70)90150​-X.
nonvolcanic tremor: Annual Review of Earth and doi​.org​/10​.1111​/j​.1525​-1314​.2010​.00923.x. Shelly, D.R., Beroza, G.C., Ide, S., and Nakamula,
Planetary Sciences, v. 39, p. 271–296, https://​ Hyndman, R.D., McCrory, P.A., Wech, A., Kao, H., S., 2006, Low-frequency earthquakes in Shikoku,
doi​.org​/10​.1146​/annurev​-earth​-040809​-152531. and Ague, J., 2015, Cascadia subducting plate Japan, and their relationship to episodic tremor
Burlini, L., Di Toro, G., and Meredith, P., 2009, Seis- fluids channelled to fore‐arc mantle corner: ETS and slip: Nature, v. 442, p. 188–191, https://​doi​
mic tremor in subduction zones: Rock physics and silica deposition: Journal of Geophysical Re- .org​/10​.1038​/nature04931.
evidence: Geophysical Research Letters, v. 36, search: Solid Earth, v. 120, p. 4344–4358, https://​ Shelly, D.R., Beroza, G.C., and Ide, S., 2007, Non-
L08305, https://​doi​.org​/10​.1029​/2009GL037735. doi​.org​/10​.1002​/2015JB011920. volcanic tremor and low-frequency earthquake
Cross, C.B., Diener, J.F.A., and Fagereng, Å., 2015, Ito, Y., Obara, K., Shiomi, K., Sekine, S., and Hirose, swarms: Nature, v. 446, p. 305–307, https://​doi​
Metamorphic imprint of accretion and ridge H., 2007, Slow earthquakes coincident with .org​/10​.1038​/nature05666.
subduction in the Pan‐African Damara Belt, Na- episodic tremors and slow slip events: Science, Stipp, M., Stünitz, H., Heilbronner, R., and Schmid,
mibia: Journal of Metamorphic Geology, v. 33, v. 315, p. 503–506, https://​doi​.org​/10​.1126​/science​ S.M., 2002, Dynamic recrystallization of quartz:
p. 633–648, https://​doi​.org​/10​.1111​/jmg​.12139. .1134454. Correlation between natural and experimental
Fagereng, Å., and Diener, J.F.A., 2011, Non-volcanic Kimura, G., and Ludden, J., 1995, Peeling oceanic conditions, in De Meer, S., et al., eds., Defor-
tremor and discontinuous slab dehydration: Geo- crust in subduction zones: Geology, v. 23, p. 217– mation Mechanisms, Rheology and Tectonics:
physical Research Letters, v. 38, L15302, https://​ 220, https://​doi​.org​/10​.1130​/0091​-7613​(1995)​ Current Status and Future Perspectives: Geologi-
doi​.org​/10​.1029​/2011GL048214. 023​<0217:​POCISZ>2​.3​.CO;2. cal Society of London Special Publication 200,
Fagereng, Å., Remitti, F., and Sibson, R.H., 2011, In- Lewis, J.C., and Byrne, T.B., 2003, History of meta- p. 171–190, https://​doi​.org​/10​.1144​/GSL​.SP​.2001​
crementally developed slickenfibers—Geologi- morphic fluids along outcrop-scale faults in a Pa- .200​.01​.11.
cal record of repeating low stress-drop seismic leogene accretionary prism, SW Japan: Implica- White, R.W., Powell, R., and Johnson, T.E., 2014, The
events?: Tectonophysics, v. 510, p. 381–386, tions for prism-scale hydrology: Geochemistry effect of Mn on mineral stability in metapelites
https://​doi​.org​/10​.1016​/j​.tecto​.2011​.08​.015. Geophysics Geosystems, v. 4, 9007, https://​doi​ revisited: New a–x relations for manganese-bear-
Fagereng, Å., Hillary, G.W., and Diener, J.F., 2014, .org​/10​.1029​/2002GC000359. ing minerals: Journal of Metamorphic Geology,
Brittle‐viscous deformation, slow slip, and tremor: Matsuhisa, Y., Goldsmith, J.R., and Clayton, R.N., v. 32, p. 809–828, https://​doi​.org​/10​.1111​/jmg​
Geophysical Research Letters, v. 41, p. 4159– 1979, Oxygen isotopic fractionation in the sys- .12095.
4167, https://​doi​.org​/10​.1002​/2014GL060433. tem quartz-albite-anorthite-water: Geochimica et Yardley, B., 1983, Quartz veins and devolatilization
Fisher, D.M., and Brantley, S.L., 2014, The role of Cosmochimica Acta, v. 43, p. 1131–1140, https://​ during metamorphism: Journal of the Geologi-
silica redistribution in the evolution of slip insta- doi​.org​/10​.1016​/0016​-7037​(79)90099​-1. cal Society, v. 140, p. 657–663, https://​doi​.org​
bilities along subduction interfaces: Constraints Meneghini, F., Fagereng, Å., and Kisters, A., 2017, /10.1144​/gsjgs​.140​.4​.0657.
from the Kodiak accretionary complex, Alaska: The Matchless Amphibolite of the Damara belt,
Journal of Structural Geology, v. 69, p. 395–414, Namibia: Unique preservation of a late Neopro-
https://​doi​.org​/10​.1016​/j​.jsg​.2014​.03​.010. terozoic ophiolitic suture: Ofioliti, v. 42, p. 129–
Frank, W.B., Shapiro, N.M., Husker, A.L., Kostog­ 145, https://​doi​.org​/10​.4454​/ofioliti​.v42i2​.488. Manuscript received 23 August 2017
lo­dov, V., Romanenko, A., and Campillo, M., Miller, R.McG., 1983, Evolution of the Damara Oro- Revised manuscript received 19 October 2017
2014, Using systematically characterized gen of Southwest Africa/Namibia: Geological Manuscript accepted 23 October 2017
low-frequency earthquakes as a fault probe in Society of South Africa Special Publication 11,
515 p. Printed in USA

|  Volume 46  |  Number 1  |  GEOLOGY


70 www.gsapubs.org 

You might also like