Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Chemical Geology 384 (2014) 47–61

Contents lists available at ScienceDirect

Chemical Geology
journal homepage: www.elsevier.com/locate/chemgeo

Reaction rind formation in the Catalina Schist: Deciphering a history of


mechanical mixing and metasomatic alteration
Sarah C. Penniston-Dorland a,⁎, Julia K. Gorman a, Gray E. Bebout b, Philip M. Piccoli a, Richard J. Walker a
a
Department of Geology, University of Maryland, College Park, MD 20742, USA
b
Department of Earth and Environmental Sciences, Lehigh University, Bethlehem, PA 18015-3001, USA

a r t i c l e i n f o a b s t r a c t

Article history: In the subduction-related Catalina Schist, reaction zones or ‘rinds’ developed at contacts of metamorphosed mafic
Received 23 March 2014 blocks with chemically distinct mélange matrix. These rinds exhibit complex chemical compositions that defy
Received in revised form 22 June 2014 simple, single-process explanations. A comparison of high-grade and low-grade reaction rinds, including field
Accepted 24 June 2014
observations, bulk-rock geochemistry, thin section textural observations, mineral compositional analyses, and
Available online 2 July 2014
thermodynamic modeling provides constraints on the nature and timing of the rind-forming processes.
Editor: K. Mezger Bulk-rock variations in the highly siderophile elements (HSE, including Os, Ir, Ru, Pt, Pd and Re) and MgO, Cr,
and Ni provide insight into the physical processes of mixing of ultramafic material of the mélange matrix, rich
Keywords: in these elements, with mafic material derived from the blocks. Enrichments in the concentrations of these
Subduction zone elements are found in both low-grade and high-grade reaction rinds suggesting that the processes that enrich
Mélange these elements and create the reaction rinds occur over a wide range of P–T conditions in subduction zones.
Metamorphism Differences in the chemical composition of garnets in an amphibolite-grade metamafic block, compared with gar-
Highly siderophile elements nets in the associated reaction rind, suggest that at least some of the mixing occurred before the growth of garnet
Garnet
in the blocks, perhaps at P–T conditions different from those producing the garnet. Bulk-rock variations in
elements commonly considered fluid-mobile, such as K2O, Ba and Rb, are observed in all reaction rinds. At
high-grade, enrichments of K2O, Ba and Rb are associated with replacement of garnet and amphibole by phengite
and chlorite. Because phengite is the major host for K, Ba and Rb, the addition of these elements by fluids appears
to have been a post-garnet growth phenomenon. This study highlights the significance of processes of metasomatism
and mechanical mixing in mélange zones for influencing the geochemical evolution of the slab–mantle interface. Any
“fluids” (hydrous or silicate melts) emanating from subducting slabs and entering the mantle wedge and arc lava
source regions would necessarily reflect the hybrid compositions created by these processes.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction fluid–rock interactions (e.g., Bebout and Barton, 1989; Sorensen and
Grossman, 1989; Nelson, 1995; Bebout and Barton, 2002; King et al.,
The process of subduction juxtaposes rock types with drastically 2006, 2007). Chemical and isotopic tracers can be used to elucidate
different chemical and isotopic compositions, leading to great potential processes (e.g., mechanical mixing, melting, fluid infiltration, diffusion)
for metasomatic alteration (Bebout, 2013). The compositional diversity occurring within subduction zones, but careful petrologic analysis of
along the slab–mantle interface can be exploited to trace various metamorphic rocks provides an important context for understanding
subduction zone components from sources through metamorphic these processes (e.g., Sorensen and Barton, 1987; Bebout and Barton,
processes and ultimately into arc magmas (e.g., Gill, 1981; Hawkesworth 1989; Brunsmann et al., 2001; Breeding et al., 2004; Bebout, 2007;
et al., 1993; Plank and Langmuir, 1993; Davidson, 1996; Alves et al., John et al., 2008; Spandler et al., 2008; Marschall et al., 2009;
1999; Elliott, 2003; Morris and Ryan, 2003). Subduction-related metamor- Penniston-Dorland et al., 2010; Bebout, 2014).
phic rocks also show evidence for a variety of processes occurring within The tectonic mélange zones found in many subduction zone
the “subduction channel,” the dynamic interface between the subducting metamorphic complexes around the world are thought to represent the
slab and the overlying mantle wedge (Shreve and Cloos, 1986; Cloos and interface between the subducting slab and the overlying mantle wedge.
Shreve, 1988a,b; Bebout and Barton, 1989; Gerya et al., 2002; Angiboust Reaction zones, sometimes referred to as ‘rinds’, between different lithol-
et al., 2011), that act to hybridize these disparate crustal and mantle ogies, commonly at the margins of tectonic blocks, are found in mélange
lithologies, including tectonic mixing and metasomatism due to zones worldwide. These reaction zones vary in mineralogy, but are
commonly composed of hydrous minerals such as amphibole, epidote,
⁎ Corresponding author. Tel.: +1 301 405 6239; fax: +1 301 405 3597. phengite, talc, and chlorite (e.g., Evans et al., 1979; Sorensen and
E-mail address: sarahpd@umd.edu (S.C. Penniston-Dorland). Grossman, 1989; Sorensen et al., 1997; Fitzherbert et al., 2004;

http://dx.doi.org/10.1016/j.chemgeo.2014.06.024
0009-2541/© 2014 Elsevier B.V. All rights reserved.
48 S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61

Marschall et al., 2009), and in some localities are nearly monomineralic Sorensen and Grossman (1989) and Bebout and Barton (1989, 2002),
(e.g., Miller et al., 2009). Reaction zones are commonly thought to have provide a glimpse into the metasomatic/mixing processes operating at
formed due to fluid-assisted metasomatic alteration (e.g., Sorensen and higher temperatures (650–750 °C), albeit at depths at which such
Barton, 1987; Sorensen and Grossman, 1989; Altherr et al., 2004; Beane temperatures would not ordinarily be experienced in modern, steady-
and Liou, 2006; Miller et al., 2009), however, tectonic mixing has state subduction zones.
additionally been proposed as a mechanism for creating the chemical Major and trace-element data suggest that the protoliths for the
diversity observed in these rinds (e.g., Bebout and Barton, 2002; Catalina metamafic rocks were likely ocean-floor basalts (Sorensen,
Penniston-Dorland et al., 2012). 1986), and Sm–Nd isotope data indicate depleted mantle sources
This study investigates reaction rinds of the Catalina Schist, a (Barton et al., 1987; data in Bebout, 1989).
subduction-related metamorphic complex in California, combining The mineralogy of the metamafic blocks in these mélange zones of
petrology and geochemistry to unravel the mechanisms and relative the Catalina Schist varies with metamorphic grade. Lawsonite–albite
timing of the various metamorphic events experienced by these hybrid and lawsonite–blueschist facies metamafic blocks are dominantly
rock types. The contrasting behaviors of different groups of more composed of chlorite, albite, calcic–sodic amphibole, phengite and
fluid-mobile elements vs. elements that are relatively immobile in fluids lawsonite. Amphibolite facies metamafic blocks are dominantly
are exploited to elucidate the roles played by advective transport via composed of garnet, amphibole, rutile and titanite (Sorensen and
metamorphic fluids and by mechanical mixing of materials. The concen- Grossman, 1989; Bebout, 1997).
trations of highly siderophile elements (HSE) Os, Ir, Ru, Pt, Pd and Re Mélange matrix on Catalina consists of talc ± calcic amphibole at all
and Os isotopic composition record an ultramafic signature in reaction metamorphic grades with additional phases as follows: ± serpentine ±
rinds (Penniston-Dorland et al., 2012), providing a clearer picture of fuchsite in lawsonite–albite facies mélange; ± fuchsite ± sodic amphibole
the mechanical mixing component contributing to rind formation (see in lawsonite–blueschist facies mélange; and ± chlorite ± anthophyllite ±
also Bebout and Barton, 2002, which focused more on the evolution of ankerite ± enstatite in amphibolite facies mélange (Grove and Bebout,
the mélange matrix). This study extends the earlier work on HSE to 1995; King et al., 2006). Reaction rinds surrounding metamafic blocks in
reaction rinds from different metamorphic grades in the Catalina Schist, mélange have previously been described for garnet amphibolite blocks in
and combines these data with petrologic information from high- and the Catalina Schist (Sorensen and Grossman, 1989). The mineralogy
low-grade rinds. Petrology and thermodynamic modeling are used to of these rinds includes calcic amphibole, phengite, quartz, and
constrain the relative timing of the various mechanisms in the overall chlorite ± rutile ± talc. This manuscript contains the first detailed
history of the formation of these reaction rinds. A comparison of study of the low-grade rinds in the Catalina Schist.
low-grade (lawsonite–albite and lawsonite–blueschist facies) and
high-grade (amphibolite facies) rinds is made along with field and 3. Analytical methods
textural observations to construct a sequential history of reaction rind
formation, shedding light on complex metamorphic processes at Mineral assemblages were determined in thin section by optical
approximate depths of 15–45 km in subduction zones over temperatures methods and back-scattered electron (BSE) imaging using the JEOL
ranging from ≤200 °C to 750 °C. 8900 Electron Probe Microanalyzer (EPMA) at the University of
Maryland. Compositions of minerals were determined using
2. Geologic background wavelength-dispersive spectrometry, natural mineral standards, and a
ZAF correction scheme (Armstrong, 1988). Mineral modes were
Geochronology of the Catalina Schist indicates that metamorphism measured by counting N 650 points in thin section using BSE imaging.
occurred in the Cretaceous (K–Ar of lawsonite–blueschist facies Any uncertainty in the identification of a particular point was resolved
glaucophane, 98 ± 2 Ma and of amphibolite facies hornblende, 112 ± by obtaining an energy-dispersive X-ray spectrum (EDS).
3 Ma; Suppe and Armstrong, 1972, recalculated by Mattinson, 1986; Bulk-rock major and trace element compositions were analyzed
U–Pb of titanite, apatite, amphibole and garnet in amphibolite facies using X-ray fluorescence (XRF) at Franklin and Marshall College. Loss
metamafic blocks, 112–114 Ma; Mattinson, 1986; Lu–Hf of garnet in on ignition (LOI) was determined for each sample first, followed by
an amphibolite facies metamafic block, 114.5 Ma; Anczkiewicz XRF analysis on a fusion glass disk made from the resulting anhydrous
et al., 2004) with 40Ar/ 39 Ar cooling ages of white mica for most powder. Analysis totals for oxides should be near 100%. Repeat analyses
units (T between 200 and 300 °C) ranging from 90 to 100 Ma (Grove of whole-rock powders resulted in a relative standard deviation (RSD)
and Bebout, 1995). Upper bounds on depositional ages based on detrital of b 0.6% for major elements and b5% for trace elements with concentra-
zircons in metasedimentary rocks range from 122 Ma in amphibolite tions N 25 ppm (1σ) except Cr (12.7%), with elements of lower concen-
facies to 97 Ma in lawsonite–blueschist and lawsonite–albite facies tration falling below 35% RSD. Comparison of measured whole-rock
rocks (Grove et al., 2008). compositions of BHVO-1 and QLO-1 standards to accepted compositions
Mineralogical, thermobarometric, field, geochronological and indicates reproducibility of major element compositions within 2% RSD
detrital zircon constraints have led to the conclusion that the different (1σ) for all elements with concentrations above 0.6 wt.% and within 12%
metamorphic units of the Catalina Schist experienced differing prograde RSD for elements with lower concentrations except for Fe2O3 and Na2O
P–T paths, and that the highest grade amphibolite facies rocks may for BHVO-1 (11.95 wt.% with 2.3% RSD and 2.41 wt.% with 6.6% RSD).
represent underthrusting along one or more major faults in the forearc, Comparison of trace element compositions for trace elements with
significantly separated in space from the thrust fault along which the concentrations above 30 ppm (including Cr in BHVO-1) was within 10%
lower grade units were subducted and metamorphosed (Grove et al., RSD and trace elements with lower concentrations varied with most
2008). Peak P–T estimates of the Catalina Schist metamorphic units relative standard deviations falling below 35% except for Th (concentration
considered in this study are ~250 °C and ~0.5 GPa for lawsonite–albite 0.4 with 100% RSD) and Cr (concentration 1 ppm, 67% RSD).
facies rocks, ~ 350 °C and ~ 1.0 GPa for lawsonite–blueschist facies The chemical separation and measurement techniques used in this
rocks, and ~ 700 °C and ~ 1.0 GPa for amphibolite facies rocks (Platt, study for Os isotopic measurement and isotope dilution analysis of Re,
1975; Sorensen and Barton, 1987; Grove and Bebout, 1995). The Os Ir, Ru, Pt and Pd have been published in detail elsewhere (Shirey
lower-grade units of the Catalina Schist, including the lawsonite–albite and Walker, 1995; Rehkämper and Halliday, 1997; Becker et al., 2006;
and lawsonite–blueschist units investigated here (and by King et al., Schulte et al., 2009). Samples were sanded, crushed in an alumina
2006, 2007), experienced low-T prograde P–T paths similar to those mortar and pestle, and ground to a fine powder using an alumina
experienced by rocks subducting into modern subduction margins. ceramic container in a SPEX 8000M Mixer-Mill to avoid metal contam-
The amphibolite grade mélange rocks, investigated in detail by ination. Finely-ground sample powders were processed by inserting
S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61 49

~ 0.5–2.5 g of sample powder into Pyrex™ Carius tubes along with lower-Os samples. The Re, Ir, Ru, Pt, and Pd blanks for the procedure
individual 185Re and 190Os spikes, and a separate mixed spike consisting averaged 12 ± 17, 0.5 ± 0.7, 4 ± 2, 122 ± 33, and 17 ± 13 pg respectively
of 191Ir, 99Ru, 194Pt and 105Pd. Approximately 6 mL of concentrated (n = 3). For most samples the blanks for Ir, Pd, and Ru were b0.05%, 1%
HNO3 and 4 mL of chilled, concentrated HCl were also added to the and 1.3% respectively (except LA10-2A which were 5.4%, 12%, and 3.8%
tubes. The tubes were sealed, then heated at ~260 °C for at least 72 h. respectively). Otherwise blanks were b 6%, except for the Pt blank for
Osmium was purified using a carbon tetrachloride solvent extraction sample LA10-2A which was 33%. The higher uncertainties are for the
technique (Cohen and Waters, 1996) followed by microdistillation samples with the lowest concentrations and do not impact the conclusions
purification (Birck et al., 1997). Rhenium, Ir, Ru, Pt, and Pd were separated made from these data.
and purified from the residual acid using two successive anion exchange Osmium was analyzed via negative thermal ionization mass spec-
columns. The Os total processing blank averaged 1.2 ± 0.9 pg with 187Os/ trometry (Creaser et al., 1991). External reproducibility in 187Os/188Os
188
Os = 0.1517 ± 0.020 (1σ, n = 3). The blank was ≤0.5% for the two for Os-rich samples was about ± 0.2% based on analysis of the UB-N
high-Os samples and ranged up to 9% of the concentration of the standard (Puchtel et al., 2008). Rhenium, Ir, Ru, Pt and Pd analyses

Fig. 1. a) Field photo of lawsonite–blueschist facies block. Block core, block rind and mélange matrix are labeled. White box represents region of closeup photos in panels b and c. b) Closeup
photo of contact between rind and matrix. c) Interpretive sketch of photo in panel b. The interface between rind and matrix is complex and convoluted, with pieces of rind within the
matrix and pieces of matrix within the block rind. d) Drill core sample of lawsonite–albite facies block. Block rind is light colored material labeled “7d” with arrow in black ink. e) Inter-
pretive sketch of block in panel d showing complex nature of rind–matrix interface.
50 S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61

Table 1
Mineral modes (volume percent).

Facies Amphibolite Lawsonite–albite Lawsonite–blueschist

Core Rind Core Rind Core Rind Core Rindb Core Rind

Mineral A10-3D A10-3A A12-A4C A12-A4R LA10-3B LA10-3F LB10-1A LB10-1B LB12-E1 LB12-E3

nc 2063 2127 2299 2129 1062 1022 659 2232 1008 1007

Amphibole 72.71 55.15 78.90 83.70 33.0 0.6 17.03 46.0


Garnet 18.57 0.28 14.83
Phengite 1.31 24.31 0.91 1.78 16.4 4.4 8.06 30.9 57.0
Chlorite 1.45 5.41 1.48 7.84 26.9 62.8 54.8 28.27 3.7 35.8
Rutile 1.99 0.56 1.48 0.23
Apatite 0.92 0.33 0.65 0.5 0.6 0.3 0.13 0.1
Pyrite 0.34 tr 0.23 tr
Quartz 0.10 13.68 2.17 7.89 0.8 1.21 0.3 0.1
Titanite 0.10 0.28 0.04 3.9 2.8 4.9 1.52 1.5 2.4
Zircon 0.10 tr 0.05 tr tr
Lawsonite 0.09 0.05 27.6 0.9 0.09 4.3
Albite 23.1 26.1 38.62 10.8 4.3
Calcite 1.3 8.0 5.06 0.9 0.3
K-feldspar 1.3 0.1
Epidote 0.19 0.2 0.2
Fe-oxide 0.1
Monazite tr
Garneta inclusions 2.28 0.34
a
Garnet inclusions are chlorite, amphibole, albite, phengite, epidote, quartz, pyrite, rutile, and titanite, bphengite and chlorite interlayered on fine-scale, EDS spectra were difficult to
interpret but results were binned based on ratio of peak heights of K vs. Mg and Mg vs. Al. cn indicates the number of points used for determination of mineral modes.

were conducted using a Nu Plasma multi-collector inductively coupled facies. The mineral modes for representative block core and rind
plasma mass spectrometer coupled with a Cetac Aridus™ desolvating samples are shown in Table 1.
nebulizer. All analyses were obtained using static, multiple multipliers.
Measured concentrations of Re, Os, Ir, Ru, Pt and Pd in the UB-N
standard were 0.210, 3.61, 3.29, 6.12, 7.30, and 7.83 ppb respectively. 4.2.1. Amphibolite facies metamafic blocks and reaction rinds
Except for Pd, these values fall within the range of reported measure- Amphibolite facies metamafic blocks are composed of amphibole
ments for this standard (Puchtel et al., 2008). The Pd concentration and garnet with phengite, chlorite and rutile and minor amounts of
measured for UB-N for this study was about 15% higher than the reported apatite, pyrite, titanite, quartz and zircon. The associated rinds have
range. Other analyses of Pd done at the same time on the same standard calcic amphibole, phengite, quartz, and chlorite and minor amounts of
powder were within the accepted range, so an anomalous concentration rutile, apatite, titanite, and in rare cases garnet.
in the powder aliquot analyzed is suspected. Detailed petrography of rind samples A10-3A and A10-3B revealed a
number of features. Quartz veinlets within this sample are oriented par-
allel to the concentric foliation. There are small euhedral garnets within
4. Results
the quartz veins. The sample also contains resorbed garnets in the rest of
the rock that are partially pseudomorphed by phengite and chlorite
4.1. Field observations
(Fig. 2a). Rind amphiboles are also pseudomorphed by chlorite and
phengite (Fig. 2b). The replacement chlorite and phengite are randomly
Reaction rinds surrounding metamafic blocks in mélange have
oriented, not aligned parallel to the concentric foliation.
previously been described for garnet amphibolite blocks in the Catalina
Average major-element compositions of the main mineral phases in
Schist as well as for mafic eclogite and blueschist from other localities
the block core and rind of amphibolite facies samples are shown in
(e.g., Sorensen and Barton, 1987; Sorensen and Grossman, 1989;
Table 2. The phengite in the block core contains greater concentrations
Sorensen et al., 1997). Block and rind samples for this study were
of Si compared to the phengite in the rinds, which are more Al-rich (cf.
collected mostly along beaches and in valleys where they can be
Sorensen, 1988). The rind phengite and chlorite have an overlapping
found either as bedrock exposed by waves or runoff or as large float
range of Mg/(Mg + Fe) compared to phengite and chlorite from the
blocks (see Table A1 for location details). There is generally a sharp
block core. Amphiboles in the block core and rind are dominantly
contact between a metamafic block and its reaction rind. The rinds are
Mg-hornblende. Rind hornblendes have higher or similar average Mg/
concentrically foliated with amphiboles and sheet silicates oriented
(Mg + Fe) when compared to hornblendes from the associated
such that they wrap around the exterior of the semi-spherical
block core (0.79 compared to 0.73 for A10-5; 0.72 compared to
metamafic blocks. The low-grade rinds also typically have concentric
0.69 for A10-3).
foliation, however, the nature of the interface between blocks and the
Garnets in both the block core and rind in sample A10-3 are
encasing matrix in lower-grade rocks is complex (see King et al.,
generally almandine-rich (with X Fe ranging from approximately
2006). Where observed, the boundary is convoluted and pieces of
0.50 to 0.65). Rind garnets have higher Mg and lower Ca than core
reaction rind are observed apparently broken off from the block and
garnets (Table 2). X-ray maps reveal concentric growth zoning in
surrounded by mélange matrix. Mélange matrix appears between
garnets in the block core (Figs. 3 and 4). Core garnets have high
pieces of broken-off rind (Fig. 1).
concentrations of Ca and Mn in the centers, decreasing towards
rims, and low concentrations of Mg in garnet centers, increasing to-
4.2. Mineralogy and mineral chemistry wards rims. Rind garnets did not exhibit systematic zoning in any
of the elements analyzed. Neither Cr nor Fe displayed any systematic
Samples were obtained of metamafic blocks and associated reaction variation either from garnet centers to rims, or between garnets in
rinds from the lawsonite–albite, lawsonite–blueschist, and amphibolite block cores and rinds.
S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61 51

shown normalized to estimates for primitive mantle (McDonough and


Sun, 1995) in Fig. 5. The major- and trace-element compositions of mé-
lange matrix samples are reported in Bebout and Barton (2002), King
et al. (2006), and King et al. (2007). Rinds generally have higher concen-
trations of MgO, Ni and Cr (see Figs. 5–7), compared to metamafic block
cores. These elements are also enriched in mélange matrix, since there is
a large peridotitic contribution in the mélange matrix (Bebout and
Barton, 2002). Along with these enrichments, there are depletions in
CaO, Na2O, Al2O3, and TiO2 in rinds, compared to block cores, and
these elements are relatively depleted in peridotite. Concentrations of
large ion lithophile elements (LILE) Rb, Ba, and K2O as well as SiO2 are
variable. In some rinds they are elevated relative to the associated
metamafic block cores and in some rinds they are depleted (Fig. 6).

4.4. Whole-rock highly siderophile element geochemistry

Samples of amphibolite facies blocks and rinds were previously


analyzed for concentrations of HSE as well as 187Os/188Os for two

Fig. 2. a) Photomicrograph in plane-polarized light of amphibolite facies rind sample A10-


3A. Garnet (gar) in the center of the image is surrounded by and being replaced by chlorite
(chl) and phengite (ph). b) Photomicrograph in plane polarized light of amphibolite facies
rind sample A10-3B. Chlorite (chl) and phengite (ph) pseudomorphically replace amphibole.
Dashed outline shows original amphibole crystal habit. Phengite and chlorite are randomly
oriented, not aligned with general foliation (approximately N–S on thin section).

4.2.2. Lawsonite–albite and lawsonite–blueschist facies metamafic blocks


and reaction rinds
Lawsonite–albite and lawsonite–blueschist facies metamafic blocks are
generally composed of lawsonite, albite and phengite pseudomorphing
original plagioclase phenocrysts along with chlorite and minor titanite
and apatite. The associated rinds are composed of chlorite and actinolite,
along with phengite, titanite and apatite. Both blocks and rinds contain
calcite that generally occurs in veins. One lawsonite–blueschist facies
metamafic block (LB12-E) consists of glaucophane, albite, phengite and
lawsonite along with chlorite, titanite, quartz, apatite and iron oxide and
K-feldspar and calcite in veins. The associated rind is mostly phengite and
chlorite with lawsonite, titanite, and small amounts of quartz, and calcite.
Average major-element compositions of the main mineral phases in
the block core and rind of lawsonite–albite and lawsonite–blueschist
facies samples are shown in Table 2. Amphibole in most of the
lawsonite–albite and lawsonite–blueschist facies rinds is Mg-rich
actinolite (Mg/(Mg + Fe) = 0.64 to 0.87). Phengite in the rind is also
Mg-rich with an average Mg/(Mg + Fe) of 0.80 that is higher than that
of phengite in the block core (0.62). Average Mg/(Mg + Fe) of rind
chlorite is also higher (0.60) than chlorite in the block core (0.49).

4.3. Whole-rock major- and trace-element geochemistry


Fig. 3. a) Calcium and b) manganese X-ray maps of garnet from block core sample A10-3D.
High concentrations are indicated by hot colors and low concentrations by cool colors (red
Whole-rock XRF major- and trace-element concentrations for a suite N yellow N blue). Euhedral outlines of zones within garnets suggest that original growth
of high- and low-grade block cores and rinds are reported in Table 3 and concentrations are retained.
52 S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61

Table 2
Electron probe microanalyzer major-element compositions of minerals.

Phengitea Chloritea

Facies Amphibolite Lawsonite–blueschist Amphibolite Lawsonite–blueschist

Core Rind Core Rind Core Rind Core Rind

A10-3D A10-3A LB12-E1 LB12-E3 A10-3D A10-3A LB12-E1 LB12-E3

SiO2 53.73 49.28 52.07 52.33 26.39 27.70 28.89 28.26


TiO2 0.13 0.50 0.10 0.04 0.10 0.04 0.07 0.03
Al2O3 23.17 29.27 26.70 26.66 23.05 20.68 19.13 20.72
FeO 3.83 2.80 3.35 2.22 22.41 21.13 25.64 21.45
MnO 0.03 0.04 0.05 0.03 0.20 0.42 0.30 0.34
MgO 5.01 3.47 3.41 4.91 16.39 17.70 14.03 17.83
CaO 0.04 0.03 0.03 0.02 0.02 0.02 0.11 0.06
BaO 0.11 0.72 0.86 0.46 0.04 0.05 0.00 0.03
Na2O 0.03 0.26 0.18 0.03 0.01 0.02 0.24 0.01
K2O 10.25 9.87 9.78 9.76 0.01 0.01 0.03 0.02
Oxide sum 96.32 96.24 96.53 96.46 88.64 87.78 88.43 88.75

Si 3.56 3.28 3.45 3.448 2.69 2.84 3.00 2.86


Al 1.81 2.29 2.08 2.06 2.78 2.50 2.34 2.48
Ti 0.01 0.03 0.01 0.00 0.01 0.00 0.01 0.00
Feb 0.21 0.16 0.19 0.12 1.92 1.82 2.23 1.82
Mn 0.00 0.00 0.00 0.00 0.02 0.04 0.03 0.03
Mg 0.49 0.34 0.34 0.48 2.49 2.70 2.17 2.69
Ca 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.01
Ba 0.00 0.02 0.02 0.01 0.00 0.00 0.00 0.00
Na 0.00 0.03 0.02 0.00 0.00 0.00 0.05 0.00
K 0.87 0.84 0.83 0.82 0.00 0.00 0.00 0.00
ne 10 18 16 14 18 10 6 14
Mg/Mg + Fe 0.70 0.69 0.62 0.80 0.56 0.60 0.49 0.60
(Range) 0.65–0.74 0.64–0.74 0.44–0.78 0.78–0.84 0.48–0.68 0.53–0.73 0.48–0.51 0.50–0.71

Amphibolea

Facies Amphibolite Lawsonite–albite Lawsonite–blueschist

Core Rind Core Rind Rind Rind Core

A10-3D A10-3A A10-5B A10-5A LA10-3F LB10-1B LB12-E1

actc hblc Na-poor Na-rich

SiO2 47.32 48.04 47.17 55.41 49.24 57.56 52.68 56.98 56.89
TiO2 0.50 0.47 0.37 0.02 0.35 0.03 1.29 0.04 0.07
Al2O3 11.70 11.62 13.35 2.04 11.06 0.10 2.92 0.66 11.75
FeO 11.54 10.65 9.76 7.66 7.81 5.66 10.49 6.77 15.21
MnO 0.06 0.27 0.06 0.08 0.08 0.16 0.25 0.16 0.13
MgO 14.35 15.07 14.66 19.53 16.55 21.73 17.94 20.34 6.55
CaO 11.14 10.64 11.27 12.79 11.28 13.45 9.34 12.57 0.40
Na2O 1.51 1.92 1.86 0.64 1.83 0.24 3.12 0.78 6.88
K2O 0.18 0.15 0.28 0.07 0.30 0.04 0.43 0.06 0.28
Oxide sum 98.29 98.83 98.78 98.25 98.49 98.96 98.47 98.38 98.19

Si 6.77 6.81 6.67 7.75 6.91 7.91 7.50 7.91 7.95


Al 1.97 1.94 2.23 0.34 1.83 0.02 0.50 0.11 1.94
Ti 0.05 0.05 0.04 0.00 0.04 0.00 0.14 0.00 0.01
Feb 1.38 1.26 1.16 0.90 0.92 0.65 1.26 0.79 1.78
Mn 0.01 0.03 0.01 0.01 0.01 0.02 0.03 0.02 0.02
Mg 3.06 3.18 3.09 4.07 3.46 4.46 3.80 4.21 1.37
Ca 1.71 1.61 1.71 1.92 1.70 1.98 1.43 1.87 0.06
Na 0.42 0.53 0.51 0.17 0.50 0.06 0.86 0.21 1.86
K 0.03 0.03 0.05 0.01 0.05 0.01 0.08 0.01 0.05
n 16 11 19 3 14 3 10 15 16
Mg/Mg + Fe 0.69 0.72 0.73 0.82 0.79 0.87 0.79 0.84 0.44
(Range) 0.67–0.72 0.69–0.77 0.68–0.78 0.81–0.83 0.77–0.81 0.87 0.64–0.84 0.82–0.86 0.35–0.55

Garneta

Facies Amphibolite

Core Rind

A10-3C A10-3D A10-3A

Garnet 8 Garnet 9 Garnet 1 Garnet 3

Center Rim Center Rim

SiO2 37.51 37.91 37.69 38.10 38.03 37.98


TiO2 0.06 0.13 0.14 0.14 0.07 0.07
Al2O3 21.34 21.51 21.16 21.43 21.26 21.28
FeO 25.94 25.92 25.17 25.07 25.21 25.44
MnO 1.58 0.71 1.15 0.42 1.70 1.97
S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61 53

Table 2 (continued)
Garneta

Facies Amphibolite

Core Rind

A10-3C A10-3D A10-3A

Garnet 8 Garnet 9 Garnet 1 Garnet 3

Center Rim Center Rim

MgO 4.56 5.68 3.81 5.77 6.91 7.22


CaO 8.05 7.58 10.26 8.48 5.69 5.10
Oxide sum 99.05 99.42 99.38 99.42 98.88 99.07

Ca 0.68 0.64 0.87 0.71 0.48 0.43


Feb 1.72 1.70 1.67 1.64 1.66 1.67
Mg 0.54 0.66 0.45 0.67 0.81 0.85
Mn 0.11 0.05 0.08 0.03 0.11 0.13
Ti 0.00 0.01 0.01 0.01 0.00 0.00
Al 1.99 1.99 1.97 1.98 1.97 1.97
Si 2.98 2.98 2.98 2.98 2.99 2.98
n 4 11 5 15 10 7
Mg/Mg + Fe 0.24 0.28 0.21 0.29 0.33 0.34
(Range) 0.23–0.24 0.24–0.31 0.21 0.25–0.32 0.32–0.34 0.33–0.34
a
Number of oxygens used in calculations of mineral formulae (per formula unit) — phengite = 11, chlorite = 14, amphibole = 23, garnet = 12.
b
Calculations assume Fe+2.
c
act = actinolite, hbl = hornblende. en indicates number of analyses for each sample.

Fig. 4. Major-element compositions of garnet (in atoms per formula unit — apfu) measured by electron probe microanalyzer across traverses within garnets. Gray lines are compositions of
block core garnets, black lines are compositions of block rind garnets. X-axis on all graphs is analysis point number, a proxy for distance across garnet. The spacing between points ranged
from 30 to 100 μm for both core and rind garnets. a. Ca, b. Mg, c. Mn.
54 S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61

Table 3
Whole-rock major and trace element chemistry.

Block cores Block rinds

Amphibolite Lawsonite–albite Lawsonite– Amphibolite Lawsonite–albite Lawsonite–


blueschist blueschist

A10-3D A10-5B A12-A4Ca A10-2A4 LA10-2A LA10-3B LB10-1A LB12-E1 A10-3A A10-5A A12-A4Ra A10-2A1 LA10-2C LA10-3F LB10-1B LB12-E3

(Wt.%)
SiO2 43.06 41.26 44.23 36.97 45.13 45.39 38.07 54.36 56.03 50.20 52.46 55.42 56.35 40.12 51.78 46.47
TiO2 2.87 4.71 2.49 5.10 3.48 2.14 1.41 0.65 0.99 0.61 0.89 0.09 0.03 1.77 1.03 0.88
Al2O3 13.79 14.95 14.79 17.63 17.62 24.01 17.61 18.21 10.84 10.84 11.46 5.50 2.43 12.81 14.81 21.88
Fe2O3T 16.65 15.10 15.92 17.90 13.83 11.54 11.86 10.02 10.06 8.83 11.11 7.07 7.03 13.49 7.44 9.24
MnO 0.21 0.30 0.33 0.35 0.18 0.17 0.16 0.12 0.25 0.09 0.18 0.12 0.12 0.27 0.13 0.18
MgO 11.32 11.13 10.95 5.64 9.14 4.61 9.08 5.26 11.18 16.77 13.17 20.44 25.48 25.40 12.32 13.10
CaO 10.15 10.65 9.32 12.26 3.90 6.75 16.50 3.13 7.42 10.31 8.84 10.28 7.58 5.37 6.93 2.32
Na2O 1.11 1.10 1.44 1.26 4.66 2.49 2.68 4.49 0.94 1.40 1.56 0.93 0.22 0.25 4.24 1.33
K2O 0.44 0.22 0.26 0.19 0.41 2.47 1.97 3.75 2.13 1.03 0.38 0.05 0.01 0.02 1.02 4.84
P2O5 0.34 0.79 0.27 2.76 0.57 0.33 0.27 0.15 0.16 0.01 0.02 0.01 0.01 0.29 0.20 0.04
Total 99.91 100.09 99.98 99.91 98.92 99.90 99.61 100.14 100.00 100.09 100.09 99.91 99.26 99.79 99.90 100.28
LOIb 2.27 1.50 1.77 2.07 6.26 7.21 13.08 4.74 2.98 2.07 2.15 1.37 4.13 9.17 6.52 8.55
FeO 12.12 8.54 10.10 13.56 11.49 8.67 8.56 7.04 7.92 7.05 8.65 5.42 5.42 10.22 5.89 7.05
Fe2O3 3.18 5.61 4.69 2.83 1.06 1.90 2.35 2.20 1.26 0.99 1.49 1.05 1.01 2.13 0.89 1.40

(ppm)
Rb 11.2 5.0 3.5 5.9 19.6 53.8 41.6 87.5 50.2 24.9 5.5 2.4 0.8 1.7 26.6 132.4
Sr 40 72 57 226 208 290 662 371 45 65 46 46 39 27 268 185
Y 63.2 47.3 59.5 89.9 16.2 38.5 34.2 24.9 31.8 8.8 46.2 2.4 1.0 27.2 24.4 35.4
Zr 145 195 113 738 141 150 92 51 99 81 155 12 15 31 54 53
V 574 429 495 243 264 240 319 288 232 272 230 31 51 223 220 453
Ni 115 188 96 73 137 106 95 53 348 617 335 1330 1418 769 304 162
Cr 376 155 200 73 133 225 237 75 658 1013 863 1595 2592 1381 410 101
Nb 7.1 48.7 5.2 67.8 26.0 30.4 6.4 1.3 4.4 2.9 1.8 b 0.5 b0.5 20.1 16.3 2.1
Ga 19.3 15.7 17.2 9.9 19.5 21.9 15.9 16.0 13.4 17.1 14 6.2 5.9 13.6 10.3 16.5
Cu 82 23 33 27 9 25 62 97 62 25 52 6 10 40 49 191
Zn 158 139 145 94 141 156 87 88 105 149 117 135 58 80 48 83
Co 69 62 64 39 41 43 33 36 42 57 47 52 71 89 43 49
Ba 90 81 66 112 78 827 691 1261 793 285 142 11 8 21 337 1076
La 11 20 7 28 6 20 7 10 8 3 7 4 3 3 8 11
Ce 32 97 23 162 13 50 10 14 11 5 16 5 4 12 19 27
U 1.6 2.4 b0.5 5.9 1.4 1.5 1.2 b0.5 2.5 2.4 b0.5 1.4 b0.5 0.5 1.6 b0.5
Th 2.5 10.1 3.0 11.5 1.2 5.5 1.1 2.4 2.8 0.6 b0.5 1.0 b0.5 2.0 1.4 3.4
Sc 59 40 56 29 38 30 54 31 31 25 33 9 8 22 27 21
Pb 1 2 1 1 b1 b1 1 5 b1 1 b1 b1 b1 1 b1 2
a
Concentrations are the average of 6 (A12-A4C) or 7 analyses (A12-A4R).
b
LOI = loss on ignition.

block–rind pairs and one sample of mélange matrix from the Catalina 5.1. Peridotite vs. mafic component in rinds
Schist (Penniston-Dorland et al., 2012). The HSE concentrations and iso-
topic ratios for three additional block–rind pairs of lawsonite–albite Reaction rinds generally have elevated concentrations of MgO,
(LA10-2, LA10-3) and lawsonite–blueschist (LB10-1) facies samples Ni, Cr, Os, Ir and Ru compared to metamafic block cores (see
are reported in Table 4. HSE patterns are shown in Fig. 8 normalized Tables 3 and 4 and Figs. 6 to 9). This suite of elements is also
to primitive upper mantle values (PUM) of Becker et al. (2006). enriched in rocks of the mélange matrix and has previously been
The low-grade samples share some HSE characteristics with the considered to reflect a peridotitic component in the matrix and
higher-grade rocks. Both high- and low-grade block cores have rind samples (Bebout and Barton, 2002; Penniston-Dorland et al.,
HSE concentrations and 187 Os/ 188 Os consistent with basalt. Both 2012). Correlated with these enrichments are depletions in rinds
high- and low-grade rinds have higher concentrations of Os, Ir in elements that are found in low concentration in peridotite and
and Ru, and lower initial 187Os/188Os, compared to cores (Figs. 6, comparatively high concentration in mafic rocks, such as CaO,
8 and 9). Na 2 O, Al2 O 3, and TiO2 . The 187 Os/ 188 Os of rinds is lower than
metamafic block cores, consistent with the low, long-term Re/Os
that is characteristic of peridotites, compared with metamafic rocks
5. Discussion (Fig. 9).
The bulk rock difference in chemical composition is also reflected in
Reaction rinds found on the outer parts of metamafic blocks of the the chemical composition of the minerals in the rinds and metamafic
Catalina Schist have distinctly different mineralogy and chemistry blocks. The Fe–Mg-bearing minerals including phengite, chlorite, and
when compared to both the metamafic blocks and the mélange ma- amphibole are generally more Mg-rich in the rinds than in the
trix. The chemical composition of the rinds can be divided into three metamafic block cores.
types of elements: those that reflect a peridotite component, those Elevated concentrations of HSE in high-grade block rinds from the
that reflect a component of the metamafic blocks and those that do Catalina Schist and other subduction zone mélanges have been
not reflect either metamafic block or peridotite. Below we explore interpreted as the product of mechanical mixing of material derived
these types of chemical differences and consider the possible mech- from metamafic blocks and peridotitic material (Penniston-Dorland
anisms that created the chemical variations observed in the reaction et al., 2012). Penniston-Dorland et al. (2012) reported that Os, Ir, and
rinds. Ru concentrations in high-grade rinds were only modestly lower than
S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61 55

Fig. 5. Bulk rock compositions of block cores and rinds normalized to primitive mantle (McDonough and Sun, 1995). a) Major-element composition of amphibolite facies block cores (open
symbols) and rinds (filled symbols). b) Trace-element composition of amphibolite facies block cores and rinds. c) Major-element compositions of lawsonite–albite (squares) and
lawsonite–blueschist (triangles) facies block cores and rinds. d) Trace-element composition of lawsonite–albite and lawsonite–blueschist facies block cores and rinds.

normal mantle peridotite, and were little fractionated from one another, mobilized. A mechanical mixing mechanism has been suggested for
relative to typical mantle peridotites. Similar results are observed for the rind formation based on simple mixing calculations using Os concentra-
lawsonite–albite and lawsonite–blueschist facies rinds reported here tions and 187Os/188Os of amphibolite-facies blocks and a typical mantle
(Fig. 8). As with the high-grade blocks, the rinds are well-discriminated peridotite endmember (Penniston-Dorland et al., 2012). Simple mixing
from the block cores for elements enriched in peridotite such as MgO, Cr, calculations were performed using the measured Os concentrations and
187
Ni, and the HSE, as well as Os isotopes (Fig. 9). Os/188Os of two high-grade block cores as one endmember and
These characteristics are difficult to explain via fluid transport, as the assuming mixing with a typical mantle peridotite bearing an Os concen-
enriched elements, particularly the HSE, are not efficiently fluid tration of 3.9 ppb and 187Os/188Os of 0.125. The block core endmembers

Fig. 6. Bulk-rock chemical compositions of rinds relative to associated metamafic block cores. Elements are in order of decreasing average ratio. Elements with consistently higher concen-
trations in rinds compared to metamafic block cores include Ir, Ru, Os, Cr, Ni, and MgO. Elements with variable relative abundances include Ba, Rb, K2O, and SiO2. Elements with consis-
tently lower concentrations in rinds compared to metamafic block core include CaO, Na2O, Al2O3, and TiO2.
56 S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61

Fig. 8. Primitive upper mantle (PUM) normalized HSE abundances for low-grade samples
of the Catalina Schist. PUM values from Becker et al. (2006). For comparison, the gray field
outlines the range of compositions reported for abyssal peridotites (Rehkämper and
Halliday, 1997; Luguet et al., 2003; Liu et al., 2009; Schulte et al., 2009) and the stippled
pattern outlines the range of compositions reported for mafic rocks (Rehkämper and
Halliday, 1997; Bézos et al., 2005; Dale et al., 2008, 2009; Schulte et al., 2009).

Complex, CA, and the Samana Metamorphic Complex, Dominican Repub-


lic (Penniston-Dorland et al., 2012). The low-grade rind compositions fall
near mixing curves defined by the high-grade block cores and peridotite.
The percentage of peridotite required ranges from b1% to 39% based on
Os concentrations (see Table 5).

5.2. Fluid-mobile element variations in rinds

Reaction rinds have variable concentrations of K2O, SiO2, Ba, and Rb


compared to metamafic block cores (see Sorensen, 1988; Sorensen and
Grossman, 1989). This suite of elements is not enriched in mélange
matrix, except SiO2, and tends to be somewhat depleted in matrix
relative to high-grade rind material (see Bebout and Barton, 2002).
While the high-grade rinds are enriched in these elements, the
Fig. 7. Differences in bulk-rock chemical composition of high- and low-grade block cores
low-grade rinds are depleted in these elements compared to metamafic
and rinds. a. Ni vs. Cr, b. MgO vs. Al2O3. Rinds from all three units have correlated concen- block cores (see Fig. 6). The concentrations of these relatively
trations of Cr, Ni, and MgO and lower Al2O3 when compared to block cores. Data for high- fluid-mobile elements are likely controlled by exchange between infiltrat-
grade samples from Penniston-Dorland et al. (2012). ing metamorphic fluids and the rinds (Bebout and Barton, 1989; Bebout,
1997; Bebout and Barton, 2002; see also Sorensen and Grossman, 1989;
Sorensen et al., 1997). The observed high-grade rind enrichments and
that were chosen have the highest and lowest 187Os/188Os measured in low-grade rind depletions in these elements compared to metamafic
metamafic rocks to constrain the upper and lower bounds of mixing. block cores suggest that there may be a difference in fluid composition
The mass percentage of each component that is required to create the between metamorphic grades. This difference may be due to differences
measured 187Os/188Os of rinds was calculated and the resulting mixing in fluid source (e.g. more or less of a sedimentary component in some
curves are illustrated in Fig. 10 along with the Os concentrations and iso- fluids) or difference in element or mineral solubility under the different
topic compositions of low-grade blocks and rinds from this study, as well P–T conditions in each metamorphic facies, suggesting a regional control
as data for high-grade blocks of the Catalina Schist, the Franciscan on the source and transport of these elements. The chemical differences

Table 4
Highly siderophile element concentrations, 187Os/188Os, 187Re/188Os, calculated initial 187Os/188Osi, for bulk samples. HSE concentrations are in ng/g.
187
Sample Type Os Ir Ru Pt Pd Re Os/188Os 187
Re/188Os 187
Os/188Osia

Lawsonite–albite facies
LA10-2A Core 0.017 0.010 0.057 0.145 0.073 0.066 0.2686 19 0.2374
LA10-2C Rind 1.541 2.29 3.21 1.68 1.26 0.038 0.1244 0.118 0.1242
LA10-3B Core 0.012 0.011 0.035 0.103 0.109 0.009 0.3352 3.7 0.3292
LA10-3F Rind 0.250 0.171 0.331 1.79 6.47 0.014 0.1322 0.27 0.1318

Lawsonite–blueschist facies
LB10-1A Core 0.017 0.018 0.055 0.184 0.359 0.014 0.3374 4.0 0.3309
LB10-1B Rind 0.045 0.038 0.083 0.337 0.380 0.134 0.3407 15 0.3166
a
114.5 Ma, the age of peak metamorphism of the Catalina Schist (Anczkiewicz et al., 2004), was used for calculation of initial 187Os/188Os ratios from measured 187Re/188Os.
S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61 57

Fig. 10. Mixing calculations (curves) compared to Os concentrations and 187Os/188Os for
low-grade samples from this study, indicated by squares, plotted along with samples
from Penniston-Dorland et al. (2012) indicated by diamonds, for amphibolite, eclogite
and blueschist blocks and rinds from the Franciscan Complex (CA) and the Samana Meta-
morphic Complex (Dominican Republic) and the Catalina Schist (larger diamonds) along
with Catalina Schist melange matrix (gray diamond).

are reflected in the modal mineralogy of the reaction rinds. There is a


greater abundance of phengite in the high-grade rinds which show overall
addition of K2O, Ba and Rb (see discussion in the next section).

5.3. Timing of chemical changes

The presence of garnets in the garnet amphibolite block cores, and


also in one of the amphibolite facies rinds, allows for the determination
of the timing of chemical changes relative to the timing of garnet
growth. As described above, the addition of K2O, Ba, and Rb to
high-grade rinds is attributed to fluid infiltration. The replacement of
both garnet and amphibole by chlorite and phengite in the amphibolite
facies indicates a late-stage alteration of garnet and amphibole in the
rinds, and since phengite is the main host of K (and likely Ba and Rb)
in these rocks, it appears that these elements were added post-garnet
growth. Therefore the fluid infiltration event that added K2O, Ba, and
Rb likely occurred after garnet growth, relatively late in the sequence
of events that affected the rinds. The fact that these elements are
enriched only in block rinds and not in the mélange matrix suggests
that there is also a mineralogical control on their distribution. The
phengite that hosts these elements is a common mineral found in
high-grade reaction rinds, but not in the amphibolite facies mélange
matrix. Fluids had to travel through the matrix in order to reach the
rinds, so it may be puzzling that the matrix is not similarly enriched in
K2O, Ba and Rb. The reason for this may be found in considering the
relative proportions of Al and Mg in mélange matrix compared to
reaction rinds (Fe contents of matrix and rinds are similar). The average
concentration of Al2O3 in mélange matrix is ~ 6 wt.% and MgO is ~ 26
wt.% (Bebout and Barton, 2002). The average Al2O3 concentration of
the rinds reported in this study is ~ 11 wt.% and MgO is ~17 wt.%. The
Al/Mg ratio of the sheet silicates increases from chlorite to biotite to
Fig. 9. Correlation between bulk-rock contents of block cores, rinds and matrix of low- (“L”)
phengite. It is likely that the relatively low concentrations of Al in the
and high-grade (“H”) samples, and mélange matrix of the Catalina Schist. a) Os, Ir, Ru with matrix stabilized chlorite, but were not high enough to stabilize
MgO, b) Os, Ir, Ru with Ni, c) Os, Ir, Ru with Cr, and d) 187Os/188Os with MgO. Data for phengite or biotite. This may explain why K2O, Ba, and Rb were not
high-grade samples from Penniston-Dorland et al. (2012). enriched in the matrix since there was not an appropriate mineral
58 S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61

Table 5
Calculated percentage of peridotite component in rinds.

Os isotopic compositions

Sample 0.124 0.123 Maximum% (up to 0.129) [Os] [Ir] [Ru] [Cr] [Ni]
a
LA10-2A/C 70.9% 28.8% n/a 39.2% 65.5% 45.4% 98.6% 71.0%
LA10-3B/F 7.0% 6.3% 17.5% 6.1% 4.6% 4.2% 48.2% 35.8%
LB10-1A/B 0.03% 0.03% 0.03% 0.72% 0.57% 0.40% 7.2% 11.2%
a
LA10-2C rind has 187Os/188Os = 0.1242 so peridotite composition involved in mixing for this sample must be less than or equal to 0.1242.

host. By contrast, the rinds have a higher Al/Mg — sufficient to stabilize This clear difference in garnet chemistry between core and rind,
the phengite which hosted the K2O, Ba and Rb enrichments found in especially MgO, suggests that the chemical changes in these elements
some of the rinds. associated with rind formation occurred prior to the growth of garnets,
In comparison, the chemical changes in elements that exhibit suggesting that mechanical mixing occurred in the amphibolite facies
variations consistent with mechanical mixing (e.g., MgO, Ni, Cr, HSE, rinds prior to the growth of garnet. Early mechanical mixing in the
Al2O3, CaO, Na2O, TiO2) likely occurred earlier in the sequence of history of rind formation is consistent with the evidence for mechanical
rind-forming events. If the garnets in the block core and rind grew mixing (geochemical evidence and the brecciated nature of the
before the chemical change associated with the mechanical mixing block–matrix contacts) in the lawsonite–albite and lawsonite–blueschist
aspect of rind formation occurred, then garnet compositions in core facies rocks, which can be thought of as proxies for early stages in the rind
and rind would be the same, however if garnets grew after the chemical formation process. It is envisioned that these regions, in which pieces of
change, then the chemical composition of the garnets could potentially metamafic block comingle with pieces of more ultramafic-rich matrix,
be different in the core and the rind. A third possibility is that the could continually evolve and recrystallize as the mélange zone materials
garnets grew during rind formation, in which case the centers of the move deeper into the subduction zone. This model is an extension of that
garnets might reflect similar compositions to those of the block core, proposed by Bebout and Barton (2002) and King et al. (2006) in which
but the evolving whole-rock composition of the rind would be reflected hydration of materials in the mélange is envisioned to result in a change
in different garnet compositions radially outward in the rind garnets in rheology of the rind material (forming more readily deformable sheet
compared to the core garnets. silicate minerals) which would have led to a focusing of deformation
The concentric growth-zoning present in X-ray maps of the block along reaction zones. Subsequent breaking off of rind material created
core garnets suggests that the compositions of the original garnet the mélange matrix which would have been more permeable to fluid
growth are preserved. Rind garnets have higher Mg and Mn contents, flow thereby enhancing subsequent reaction zone formation. Deforma-
whereas core garnets have higher Ca contents (Fig. 4). The Fe contents tion fabrics observed in reaction rinds (see Fig. 1 of King et al., 2006)
of the two types of garnets overlap. The range of concentrations of Ca, document the syn-tectonic nature of reaction rind formation.
Mg and Mn in the rind garnets does not overlap with the range of If the chemical change to a more ultramafic-rich composition due to
compositions of those elements in the core garnets. Even if the rind mixing occurred prior to garnet growth in the rind, then the growth of
garnets selected for analysis were not sectioned through the true garnet in the rinds should be predicted by thermodynamic modeling
centers of the garnets, i.e., if they represented some part of the garnets of the bulk composition of the rind. Thermodynamic modeling was
with compositions of those observed in the cores, there would be some performed using Theriak Domino software (de Capitani and
overlap in composition. Here, however, there is little overlap observed. Petrakakis, 2010) with the thermodynamic database JUN92 (Berman,
1988). Thermodynamic data for Mn-endmembers were added by
applying exchange enthalpies, entropies and molar volumes based on
Fe–Mg–Mn endmembers in the Holland and Powell (1998 updated
2002) database. The bulk rock composition used for the modeling was
that of the rind sample in which garnets were found, sample A10-3A
(with K2O reduced to reflect rind composition before reaction with
later K2O-rich fluids). The modeling predicts the growth of garnet, at
temperatures consistent with the estimated P–T conditions of the am-
phibolite facies (see Fig. 11), above 600 °C and at pressures above
0.7 GPa.
It is inferred, therefore, that the low-grade rinds represent an earlier
stage of the continuous process by which the high-grade rinds formed.
The combination of information obtained from both low-grade and
high-grade rinds, thus, may provide information about the sequence
of events that took place within the subduction channel, creating the
reaction rinds.

5.4. Overall sequence of events

Combining evidence from both high- and low-grade rinds, a general-


ized sequence of events can be inferred for the development of reaction
rinds (Fig. 12). The process is most likely not a single event occurring at
a single set of P–T conditions, but rather a continuous phenomenon that
Fig. 11. Pressure–temperature diagram showing calculated P–T conditions for garnet may occur over a range of P–T conditions, if these materials are subducted
stability for rind sample A10-3A. Results are from calculations using Theriak Domino and
from shallow levels to greater depths within the subduction zone.
are overlain on estimates of P–T conditions for the metamorphic facies of the Catalina Schist
(after Bebout, 2007). LA = lawsonite–albite, LBS = lawsonite–blueschist, GS = greenschist, Mafic and sedimentary material from the subducting slab breaks off
EBS = epidote blueschist, EA = epidote amphibolite, AM = amphibolite. Arag = aragonite, and mingles with ultramafic material, possibly from the overlying
Cal = calcite, Jd = jadeite, Qtz = quartz, Sil = sillimanite, And = andalusite. mantle wedge, beginning at relatively shallow depths. Evidence for
S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61 59

Fig. 12. Schematic sequence of events during rind formation. In the expanded sequence of stages 1 through 4, the color black is used to indicate mafic oceanic crust and gray is used to
indicate peridotitic material. The light gray hexagons indicate garnets. In stage 4 the white wavy arrows indicate fluid flow and white mineral outlines suggest mineralogical alteration
mediated by fluid flow. The relative positions of stages 1 through 4 are indicated on master subduction cross-section on the right of the figure.

the mechanical nature of the rind-forming process is found at the happen at relatively shallow depths and that blocks can continue to
lawsonite–albite and lawsonite–blueschist facies conditions of the move into the subduction zone, after this mixing process started, to
Catalina Schist which are thought to record P ~ 0.5 to 1.0 GPa (depths experience higher-grade metamorphic conditions.
of ~ 15–30 km). The shearing nature of the plate interface over time These observations focus on the mixing of mafic and ultramafic
creates semi-rounded blocks out of the relatively rigid basaltic rock materials since these are the two lithologies dominant in the amphibolite
(and some metasedimentary and ultramafic rocks) within a matrix facies of the Catalina Schist, which is the location of the metamafic block
derived from varying proportions of all of these lithologies, which has a that was the focus of this detailed study. However similar concepts can
chemical composition intermediate between these lithologies (e.g., apply equally to mélange containing more sedimentary material, such
Bebout and Barton, 2002; King et al., 2006). At the edges of the blocks, as the lower-grade rocks of the Catalina Schist.
the outermost part of the block becomes hydrated, sheared and breaks There is likely infiltration of fluids at various points during the history
off, as witnessed by broken pieces of block found at the block–matrix in- of the rind, however the only discrete evidence for this in the rinds
terface (see Fig. 1). Some blocks exhibit zonation in rinds with inner rinds studied is found in the late garnet and amphibole replacement textures
showing less enrichment in Mg, Cr, and Ni and outer rinds with more en- (Fig. 2). Both garnets and amphiboles are replaced by randomly
richment (cf. Sorensen and Grossman, 1989) which possibly demonstrate oriented phengite and chlorite, suggesting that this late-stage fluid
the progressive nature of this process. With time, some or all of the mixed infiltration takes place in a more static environment. The higher concen-
material may disperse into the matrix, creating the hybrid compositions tration of K2O, Ba and Rb in some rinds is likely related to the late-stage
observed in mélange matrix (see Bebout and Barton, 2002; King et al., infiltration of fluids through the rind, perhaps in part during cooling
2006, 2007). The block pieces and matrix still at the edge of the block be- and exhumation.
come intermixed and this material recrystallizes to minerals stable at in-
creasing P–T conditions. This process is likely an iterative process with 6. Implications
new material breaking off and mixing with more matrix over time. The
concentric foliation in the rind is evidence for the dynamic nature of the Processes of mechanical mixing occur continuously throughout the
mechanical mixing processes involved in rind formation. In cases where history of subducted material. The evidence presented here suggests
all of the mixed material disperses into the matrix, reaction rinds would that metamafic blocks were surrounded by mélange matrix starting at
not form on blocks. This could explain why there are many examples in relatively low temperatures and, therefore, at shallow depths
the Catalina Schist of blocks without reaction rinds. where T ≤ 200 °C, based on the fact that lawsonite–albite facies rocks
During subduction the metamafic block and associated rind of the have reaction rinds that show evidence for mechanical mixing and
amphibolite facies sample examined in detail reach the stability field where T b ~ 600 °C, based on the initial temperature at which garnet
for garnet and garnet growth occurs in both. The garnets that crystallize appears in the thermodynamic models. Therefore, deformation and
in the rind are more Mg-rich and Ca-poor than the garnets that crystallize mechanical mixing likely occurred over a range of P and T, throughout
in the block core, reflecting the difference in bulk-rock composition that the history of the subducted rocks. Such mechanical mixing of peridotitic
has already developed at this point, due to mixing of mafic and ultramafic and basaltic materials, with or without sediment, results in an
materials. These observations illustrate that mechanical mixing can intermediate or hybrid rock type with compositions between those
60 S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61

of the endmember materials. Measurements of HSE in arc volcanic Anczkiewicz, R., Platt, J.P., Thirlwall, M.F., Wakabayashi, J., 2004. Franciscan subduction off
to a slow start: evidence from high-precision Lu–Hf garnet ages on high grade blocks.
rocks (Alves et al., 1999, 2002) have shown ranges of concentrations Earth Planet. Sci. Lett. 225, 147–161.
and isotopic compositions that are consistent with variable amounts Angiboust, S., Agard, P., Raimbourg, H., Yamato, P., Huet, B., 2011. Subduction interface
of mixing of peridotitic mantle with radiogenic Os components, processes recorded by eclogite-facies shear zones (Monviso, W. Alps). Lithos 127,
222–238.
reflecting different proportions of subducted oceanic crust and Armstrong, J.T., 1988. Quantitative analysis of silicate and oxide minerals: comparison of
sediments in the arc lavas. The variable amounts of mixing of peridotite Monte Carlo, ZAF and phi-rho-z procedures. In: Newbury, D.E. (Ed.), Microbeam
suggested by the HSE concentrations and isotopic compositions in Analysis — 1988. San Francisco Press, San Francisco, CA, pp. 239–246.
Barton, M.D., Bebout, G.E., Sorensen, S.S., 1987. Isotopic constraints on the geochemical
reaction rinds of the Catalina Schist demonstrate that mechanical
evolution of an ultramafic subduction zone mélange: Catalina Schist terrane, California
mixing of disparate materials at the slab–mantle wedge interface is a [abstr.]. EOS Trans. Am. Geophys. Union 68, 1525.
plausible mechanism to produce the HSE variability observed in arc Beane, R.J., Liou, J.G., 2006. Metasomatism in serpentinite mélange rocks from the
high-pressure Maksyutov Complex, Southern Ural Mountains, Russia. Int. Geol. Rev.
volcanic rocks.
47, 24–40.
Mélange mixing has been proposed as a source of other geochemical Bebout, G.E., 1989. Geological and Geochemical Investigations of Fluid Flow and Mass
signals observed in arc volcanic rocks (King et al., 2006, 2007; Marschall Transfer During Subduction-Zone Metamorphism(Ph.D. Thesis) Univ. Calif., Los
and Schumacher, 2012). Both sets of authors demonstrated the genera- Angeles, CA (370 pp.).
Bebout, G.E., 1997. Nitrogen isotope tracers of high-temperature fluid–rock interactions:
tion of trace element and isotopic signatures overlapping with those case study of the Catalina Schist, California. Earth Planet. Sci. Lett. 151, 77–90.
observed in arc volcanic rocks. Transport of mélange material through Bebout, G.E., 2007. Metamorphic chemical geodynamics of subduction zones. Earth
the mantle wedge to underneath arc volcanoes by diapiric rise (in a Planet. Sci. Lett. 260, 373–393.
Bebout, G.E., 2013. Metasomatism in subduction zones of subducted oceanic slabs, mantle
“cold plume”; e.g. Gerya and Yuen, 2003; Castro et al., 2010; Marschall wedges, and the slab–mantle interface. In: Harlov, D.E., Austrheim, H. (Eds.), Metaso-
and Schumacher, 2012; Tumiati et al., 2013), could provide a way to matism and the Chemical Transformation of Rock. Springer-Verlag, Berlin, pp.
preserve the mixed signals of mélange zone materials, including the 289–349.
Bebout, G.E., 2014. Chemical and Isotopic Cycling in Subduction Zones, In: Holland, H.D.,
HSE concentrations and isotopic compositions, that are then directly Turekian, K.K. (Eds.), Second ed. Treatise on Geochemistry, 4, pp. 703–747.
melted to produce arc volcanic rocks. We suggest that, whether or not Bebout, G.E., Barton, M.D., 1989. Fluid-flow and metasomatism in a subduction zone
mélange is conveyed into the mantle wedge in diapirs, it is capable of hydrothermal system — Catalina-Schist Terrane, California. Geology 17, 976–980.
Bebout, G.E., Barton, M.D., 2002. Tectonic and metasomatic mixing in a high-T,
strongly influencing the transfer of “slab signatures” into arc source
subduction-zone mélange — insights into the geochemical evolution of the slab–mantle
regions. Fluids (aqueous fluids or siliceous melts) infiltrating, and interface. Chem. Geol. 187, 79–106.
produced in, such zones would bear hybridized chemical and isotopic Becker, H., Horan, M.F., Walker, R.J., Gao, S., Lorand, J.-P., Rudnick, R.L., 2006. Highly
siderophile element composition of the Earth's primitive upper mantle: constraints
compositions inherited from the disparate lithologies in the subducting
from new data on peridotite massifs and xenoliths. Geochim. Cosmochim. Acta 70,
oceanic lithosphere and overlying sediment. 4528–4550.
Berman, R.G., 1988. Internally-consistent thermodynamic data for minerals in the system
Acknowledgments Na2O–K2O–MgO–FeO–Fe2O3–Al2O3–SiO2–TiO2–H2O–CO2. J. Petrol. 29, 445–522.
Bézos, A., Lorand, J.-P., Humler, E., Gros, M., 2005. Platinum-group element systematics in
Mid-Oceanic Ridge basaltic glasses from the Pacific, Atlantic, and Indian Oceans.
We thank the NSF for support (grant EAR-1119111). We thank Horst Geochim. Cosmochim. Acta 69, 2613–2627.
Marschall and an anonymous reviewer for their constructive reviews Birck, J.L., Barman, M.R., Capmas, F., 1997. Re–Os isotopic measurements at the femtomole
level in natural samples. Geostand. Newslett. 21, 19–27.
and Klaus Mezger for his editorial handling of the paper. We acknowl- Breeding, C.M., Ague, J.J., Brocker, M., 2004. Fluid–metasedimentary rock interactions in
edge the support of the Maryland NanoCenter and its NISPLab. We subduction-zone mélange: implications for the chemical composition of arc magmas.
thank Matt Kohn for assistance in thermodynamic modeling and Igor Geology 32, 1041–1044.
Brunsmann, A., Franz, G., Erzinger, J., 2001. REE mobilization during small-scale high-
Puchtel for assistance with the HSE analyses. The Catalina Island Conser- pressure fluid–rock interaction and zoisite/fluid partitioning of La to Eu. Geochim.
vancy is acknowledged for logistics and support of sample collection of Cosmochim. Acta 65, 559–570.
the Catalina Schist. Castro, A., Gerya, T., García-Casco, A., Fernández, C., Díaz-Alvarado, J., Moreno-Ventas, I., Löw,
I., 2010. Melting relations of MORB-sediment mélanges in underplated mantle wedge
plumes: implications for the origin of Cordilleran-type batholiths. J. Petrol. 5, 1267–1295.
Appendix A Cloos, M., Shreve, R.L., 1988a. Subduction-channel model of prism accretion, mélange
formation, sediment subduction, and subduction erosion at convergent plate
Table A1 margins: 1. Background and description. Pure Appl. Geophys. 128, 455–500.
Sample locations. Cloos, M., Shreve, R.L., 1988b. Subduction-channel model of prism accretion, mélange
formation, sediment subduction, and subduction erosion at convergent plate
Sample Location Latitude, longitude margins: 2. Implications and discussion. Pure Appl. Geophys. 128, 501–545.
A10-3A, B, C, D Ripper's Cove beach N 33°25′28.6″ W 118°25′58.7″ 46 m elev. Cohen, A.S., Waters, F.J., 1996. Separation of osmium from geological materials by solvent
extraction for analysis by thermal ionization mass spectrometry. Anal. Chim. Acta.
A12A-4C, R Valley of Ollas N 33°25′31.3″ W 118°25′58.6″ 33 m elev.
332, 269–275.
A10-2A1, 4 Mama Melt N 33°24′57.5″ W 118°26′24.0″ 265 m elev.
Creaser, R.A., Papanastassiou, D.A., Wasserburg, G.J., 1991. Negative thermal ion
A10-5A, B Valley of Ollas N 33°25′28.6″ W 118°25′58.7″ 46 m elev. mass-spectrometry of osmium, rhenium, and iridium. Geochim. Cosmochim. Acta
LA10-2A, C Starlight Beach N 33°28′28.1″ W 118°34′52.8″ 0 m elev. 55, 397–401.
LA10-3B, F Starlight Beach N 33°28′27.6″ W 118°35′11.5″ 0 m elev. Dale, C.W., Luguet, A., Macpherson, C.G., Pearson, D.G., Hickey-Vargas, R., 2008. Extreme
LB10-1A, B Crab Cove, SW N 33°22′26.3″ W 118°28′56.3″ 0 m elev. platinum-group element fractionation and variable Os isotope compositions in
of Little Harbor Philippine Sea Plate basalts: tracing mantle source heterogeneity. Chem. Geol. 248,
LB12-E1, 3 W of Little Harbor N 33°23′12.2″ W 118°28′32.0″ 24 m elev. 213–238.
Dale, C.W., Burton, K.W., Pearson, D.G., Gannoun, A., Alard, O., Argles, T.W., Parkinson, I.J.,
2009. Highly siderophile element behaviour accompanying subduction of oceanic
crust: whole rock and mineral-scale insights from a high-pressure terrain. Geochim.
Cosmochim. Acta 73, 1394–1416.
Davidson, J.P., 1996. Deciphering Mantle and Crustal Signatures in Subduction Zone
Magmatism. In: Bebout, G.E., Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction
References Top to Bottom, 96. American Geophysical Union Geophysical Monograph, Washington,
DC, pp. 251–262.
de Capitani, C., Petrakakis, K., 2010. The computation of equilibrium assemblage diagrams
Altherr, R., Topuz, G., Marschall, H., Zack, T., Ludwig, T., 2004. Evolution of a tourmaline- with Theriak/Domino software. Am. Mineral. 95, 1006–1016.
bearing lawsonite eclogite from the Elekdag area (Central Pontides, N. Turkey): Elliott, T., 2003. Tracers of the Slab. In: Eiler, J.M. (Ed.), Inside the Subduction Factory,
evidence for infiltration of slab-derived B-rich fluids during exhumation. Contrib. 138. American Geophysical Union Geophysical Monograph, Washington, DC, pp.
Mineral. Petrol. 148, 409–425. 23–45.
Alves, S., Schiano, P., Allegre, C.J., 1999. Rhenium–osmium isotopic investigation of Java Evans, B.W., Trommsdorff, V., Richter, W., 1979. Petrology of an eclogite-metarodingite
subduction zone lavas. Earth Planet. Sci. Lett. 168, 65–77. suite at Cima di Gagnone, Ticino. Switzerland. American Mineralogist 64, 15–31.
Alves, S., Schiano, P., Capmas, F., Allegre, C.J., 2002. Osmium isotope binary mixing arrays Fitzherbert, J.A., Clarke, G.L., Marmo, B., Powell, R., 2004. The origin and P-T evolution of
in arc volcanism. Earth Planet. Sci. Lett. 198, 355–369. peridotites and serpentinites of NE New Caledonia: prograde interaction between
S.C. Penniston-Dorland et al. / Chemical Geology 384 (2014) 47–61 61

continental margin and the mantle wedge. Journal of Metamorphic Geology 22, Penniston-Dorland, S.C., Sorensen, S.S., Ash, R.D., Khadke, S.V., 2010. Lithium isotopes as a
327–344. tracer of fluids in a subduction zone mélange: Franciscan Complex, CA. Earth Planet.
Gerya, T.V., Yuen, D.A., 2003. Rayleigh–Taylor instabilities from hydration and melting Sci. Lett. 292, 181–190.
propel ‘cold plumes’ at subduction zones. Earth Planet. Sci. Lett. 212, 47–62. Penniston-Dorland, S.C., Walker, R.J., Pitcher, L., Sorensen, S.S., 2012. Mantle–crust interac-
Gerya, T.V., Stöckhert, B., Perchuk, A.L., 2002. Exhumation of high-pressure metamorphic tions in a paleosubduction zone: evidence from highly siderophile element systematics
rocks in a subduction channel: a numerical simulation. Tectonics 21. http://dx.doi. of eclogite and related rocks. Earth Planet. Sci. Lett. 319, 295–306.
org/10.1029/2002TC001406. Plank, T., Langmuir, C.H., 1993. Tracing trace elements from sediment input to volcanic
Gill, J.B., 1981. Orogenic Andesites and Plate Tectonics. Springer Verlag, Berlin (390 pp.). output at subduction zones. Nature 362, 739–743.
Grove, M., Bebout, G.E., 1995. Cretaceous tectonic evolution of coastal southern California: Platt, J.P., 1975. Metamorphic and deformational processes in the Franciscan Complex,
insights from the Catalina Schist. Tectonics 14, 1290–1308. California: some insights from the Catalina Schist terrane. Geol. Soc. Am. Bull. 86,
Grove, M., Bebout, G.E., Jacobson, C.E., Barth, A.P., Kimbrough, D.L., King, R.L., Zou, H., 1337–1347.
Lovera, O.M., Mahoney, B.J., Gehrels, G.E., 2008. The Catalina Schist: evidence for Puchtel, I.S., Walker, R.J., James, O.B., Kring, D.A., 2008. Osmium isotope and highly
middle Cretaceous subduction erosion of southwestern North America. Geol. Soc. siderophile element systematics of lunar impact melt breccias: implications for the
Am. Spec. Pap. 436, 335–361. late accretion history of the Moon and Earth. Geochim. Cosmochim. Acta 72,
Hawkesworth, C.J., Gallagher, J.M., Hergt, J.M., McDermott, F., 1993. Mantle and slab 3022–3042.
contributions in arc magma. Ann. Rev. Earth Planet. Sci. 21, 175–204. Rehkämper, M., Halliday, A.N., 1997. Development and application of new ion exchange
Holland, T.J.B., Powell, R., 1998. An internally consistent thermodynamic data set for techniques for the separation of the platinum group and other siderophile elements
phases of petrological interest. J. Metamorph. Geol. 1998 (16), 309–343. from geological samples. Talanta 44, 663–672.
John, T., Klemd, R., Gao, J., Garbe-Schonberg, C.-D., 2008. Trace-element mobilization in Schulte, R.F., Schilling, M., Anma, R., Farquhar, J., Horan, M.F., Komiya, T., Piccoli, P., Pitcher,
slabs due to non steady-state fluid–rock interaction: constraints from an eclogite- L., Walker, R.J., 2009. Chemical and chronologic complexity in the convecting upper
facies transport vein in blueschist (Tianshan, China). Lithos 103, 1–24. mantle: evidence from the Taitao ophiolite, southern Chile. Geochim. Cosmochim.
King, R.L., Bebout, G.E., Moriguti, T., Nakamura, E., 2006. Elemental mixing systematics Acta 73, 5793–5819.
and Sr–Nd isotope geochemistry of mélange formation: obstacles to identification Shirey, S.B., Walker, R.J., 1995. Carius tube digestion for low-blank rhenium–osmium
of fluid sources to arc volcanics. Earth Planet. Sci. Lett. 246, 288–304. analysis. Anal. Chem. 67, 2136–2141.
King, R.L., Bebout, G.E., Grove, M., Moriguti, T., Nakamura, E., 2007. Boron and lead isotope Shreve, R.L., Cloos, M., 1986. Dynamics of sediment subduction, mélange formation, and
signatures of subduction-zone mélange formation: hybridization and fractionation prism accretion. J. Geophys. Res. 91, 10229–10245.
along the slab–mantle interface beneath volcanic arcs. Chem. Geol. 239, 305–322. Sorensen, S.S., 1986. Petrologic and geochemical comparison of the blueschist and
Liu, C.Z., Snow, J.E., Brügmann, G., Hillebrand, E., Hofmann, A.W., 2009. Non-chondritic greenschist units of the Catalina Schist terrane, southern California. Geol. Soc. Am.
HSE budget in Earth's upper mantle evidenced by abyssal peridotites from Gakkel Mem. 164, 59–75.
ridge (Arctic Ocean). Earth Planet. Sci. Lett. 283, 122–132. Sorensen, S.S., 1988. Petrology of amphibolite-facies mafic and ultramafic rocks from the
Luguet, A., Lorand, J.-P., Seyler, M., 2003. Sulfide petrology and highly siderophile element Catalina Schist, Southern California; metasomatism and migmatization in a subduc-
geochemistry of abyssal peridotites: a coupled study of samples from the Kane tion zone metamorphic setting. J. Metamorph. Geol. 6, 405–435.
Fracture Zone (45°W 23°20 N, MARK Area, Atlantic Ocean). Geochim. Cosmochim. Sorensen, S.S., Barton, M.D., 1987. Metasomatism and partial melting in a subduction
Acta 67, 1553–1570. complex: Catalina Schist, southern California. Geology 15, 115–118.
Marschall, H.R., Schumacher, J.C., 2012. Arc magmas sourced from mélange diapirs in Sorensen, S.S., Grossman, J.N., 1989. Enrichment of trace elements in garnet amphibolites
subduction zones. Nat. Geosci. 5, 862–867. from a paleo-subduction zone: Catalina Schist, southern California. Geochim.
Marschall, H.R., Altherr, R., Gmeling, K., Kasztovszky, Z., 2009. Lithium, boron and chlorine Cosmochim. Acta 53, 3155–3178.
as tracers for metasomatism in high-pressure metamorphic rocks: a case study from Sorensen, S.S., Grossman, J.N., Perfit, M.R., 1997. Phengite-hosted LILE enrichment in
Syros (Greece). Mineral. Petrol. 95, 291–302. eclogite and related rocks: implications for fluid-mediated mass transfer in subduction
Mattinson, J.M., 1986. Geochronology of high-pressure–low-temperature Franciscan zones and arc magma genesis. J. Petrol. 38, 3–34.
metabasites: a new approach using the U–Pb system. Geol. Soc. Am. Mem. 164, 95–105. Spandler, C., Hermann, J., Faure, K., Mavrogenes, J.A., Arculus, R.J., 2008. The importance of
McDonough, W.F., Sun, S.-s., 1995. The composition of the Earth. Chem. Geol. 120, 223–253. talc and chlorite “hybrid” rocks for volatile recycling through subduction zones:
Miller, D.P., Marschall, H.R., Schumacher, J.C., 2009. Metasomatic formation and petrology evidence from the high-pressure subduction mélange of New Caledonia. Contrib.
of blueschist-facies hybrid rocks from Syros (Greece), implications for reactions at the Mineral. Petrol. 155, 181–198.
slab–mantle interface. Lithos 107, 53–67. Suppe, J., Armstrong, R.L., 1972. Potassium-argon dating of Franciscan metamorphic rocks.
Morris, J.D., Ryan, J.G., 2003. Subduction zone processes and implications for changing American Journal of Science 272, 217–233.
composition of the upper and lower mantle. In: Carlson, R. (Ed.), Treatise on Tumiati, S., Fumagalli, P., Tiraboschi, C., Poli, S., 2013. An experimental study on
Geochemistry, 2. The Mantle and Core. Elsevier, New York, pp. 451–470. COH-bearing peridotite up to 3.2 GPa and implications for crust–mantle recycling.
Nelson, B.K., 1995. Fluid flow in subduction zones: evidence from Nd- and Sr-isotope var- J. Petrol. 54, 453–479.
iations in metabasalts of the Franciscan Complex California. Contrib. Mineral. Petrol.
119, 247–262.

You might also like