01.05.2013 Views

Silicic Segregations of the Ferrar Dolerite Sills, Antarctica

Silicic Segregations of the Ferrar Dolerite Sills, Antarctica

Silicic Segregations of the Ferrar Dolerite Sills, Antarctica

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 PAGES1927^1964 2011 doi:10.1093/petrology/egr035<br />

<strong>Silicic</strong> <strong>Segregations</strong> <strong>of</strong> <strong>the</strong> <strong>Ferrar</strong> <strong>Dolerite</strong> <strong>Sills</strong>,<br />

<strong>Antarctica</strong><br />

KARINA ZAVALA 1 *, ALISON M. LEITCH 2 ANDGEORGEW.FISHER 3<br />

1 BRUNEAU CENTRE FOR RESEARCH AND INNOVATION, MEMORIAL UNIVERSITY OF NEWFOUNDLAND,<br />

ST. JOHN’S, NL A1C 5S7, CANADA<br />

2<br />

DEPARTMENT OF EARTH SCIENCES, MEMORIAL UNIVERSITY OF NEWFOUNDLAND, ST. JOHN’S, NL A1B 3X5,<br />

CANADA<br />

3<br />

DEPARTMENT OF EARTH AND PLANETARY SCIENCES, JOHNS HOPKINS UNIVERSITY, BALTIMORE, MD 21218, USA<br />

RECEIVED JULY 21, 2007; ACCEPTED JUNE 28, 2011<br />

The upper parts <strong>of</strong> <strong>the</strong> 300 m thick Basement Sill, one <strong>of</strong> <strong>the</strong> <strong>Ferrar</strong><br />

<strong>Dolerite</strong> <strong>Sills</strong> in <strong>Antarctica</strong>, contain sharply defined, coarse-grained,<br />

leucocratic segregations that are temporally, spatially and chemically<br />

related to <strong>the</strong> host dolerite. The morphologies <strong>of</strong> <strong>the</strong> segregations<br />

vary gradationally with distance from <strong>the</strong> upper contact <strong>of</strong> <strong>the</strong><br />

Basement Sill from thin (1cm) streaks, pods and stringers close to<br />

<strong>the</strong> upper contact, (0·03^0·2 m) pipes and veins about 15 m below<br />

<strong>the</strong> contact, and (0·5^2 m) sub-horizontal, anastomosing lenses<br />

about 30 m beneath <strong>the</strong> contact.Their compositions range from diorite<br />

to granodiorite; <strong>the</strong>y are enriched in P 2O 5,TiO 2,FeO T, and to a<br />

lesser extent in alkalis compared with <strong>the</strong> host dolerite, and <strong>the</strong>y display<br />

linear trends in Harker variation diagrams that are compatible<br />

with an origin as fractionated liquids formed after simultaneous<br />

crystallization <strong>of</strong> pyroxene and plagioclase from <strong>the</strong> host sill<br />

magma. The segregations show systematic variations in mineralogy<br />

and texture with composition, with <strong>the</strong> lowest SiO 2 segregations<br />

having a similar mineralogy to <strong>the</strong> host and ophitic to sub-ophitic<br />

textures; <strong>the</strong> highest SiO 2 segregations are granodioritic and exhibit<br />

a granophyric texture. In segregations with highly siliceous compositions,<br />

plagioclase crystals have resorbed edges and pyroxene forms<br />

blebby intergrowths with plagioclase and Fe^Ti oxides. Most segregations<br />

are interpreted to have formed when gravitational instability<br />

caused <strong>the</strong> upper solidification front <strong>of</strong> <strong>the</strong> sill to tear, and mixtures<br />

<strong>of</strong> crystals and residual liquid entered <strong>the</strong> tears by porous flow from<br />

<strong>the</strong> surrounding crystal mush and by channelling from lower levels<br />

in <strong>the</strong> sill. Modeling using <strong>the</strong> <strong>the</strong>rmodynamic program MELTS<br />

shows that <strong>the</strong> bulk compositions <strong>of</strong> <strong>the</strong> segregations are consistent<br />

with wet (0·5wt % H 2O) fractional crystallization <strong>of</strong> <strong>the</strong> host<br />

dolerite magma at 100 MPa, although <strong>the</strong> data are not well matched<br />

by a single parent magma composition. We present a physical model<br />

*Corresponding author: E-mail: kzavala@mun.ca<br />

<strong>of</strong> segregation formation involving multiple episodes <strong>of</strong> tearing triggered<br />

by fresh influxes <strong>of</strong> magma into <strong>the</strong> sill.<br />

KEY WORDS: <strong>Ferrar</strong> <strong>Dolerite</strong>s; segregations; mush instability; differentiation;<br />

lenses<br />

INTRODUCTION<br />

The role <strong>of</strong> fractional crystallization in producing <strong>the</strong> diversity<br />

<strong>of</strong> igneous rocks has proven to be one <strong>of</strong> <strong>the</strong> most<br />

persistent problems in igneous petrology (e.g. Harker,<br />

1909; Bowen, 1928; Holmes, 1931; Wager & Deer, 1939;<br />

Yoder, 1973; McBirney, 1975). Lenses <strong>of</strong> coarse-grained, leucocratic<br />

material in <strong>the</strong> central and upper parts <strong>of</strong> thick<br />

(4100 m) basaltic sills, lava lakes, and flood basalts are<br />

among <strong>the</strong> richest sources <strong>of</strong> information on how basaltic<br />

magmas have differentiated in nature. In this study we<br />

refer to <strong>the</strong>se leucocratic lenses as segregations, following<br />

<strong>the</strong> usage <strong>of</strong> Tomkeieff (1928), Richter & Moore (1966),<br />

Moore & Evans (1967), Wright et al. (1976), Wright &<br />

Okamura, 1977; Helz (1980, 1987), Helz et al. (1989),Marsh<br />

(1996, 2002, 2004) and Boudreau & Simon (2007). These<br />

features have elsewhere been described as pegmatites<br />

(Cornwall, 1951; Walker, 1953; Ragland & Arthur, 1985;<br />

Greenough & Dostal, 1992; Steiner et al., 1992; Puffer &<br />

Horter, 1993; Puffer & Volkert, 2001; Philpotts et al., 1996)<br />

and as pegmatoids (Lindsley et al., 1971; Bailey, 1989;<br />

Larsen & Brooks, 1994; Be¤ dard et al., 2007).<br />

ß The Author 2011. Published by Oxford University Press. All<br />

rights reserved. For Permissions, please e-mail: journals.permissions@<br />

oup.com<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

The Jurassic age <strong>Ferrar</strong> <strong>Dolerite</strong> <strong>Sills</strong> (FDS) <strong>of</strong> <strong>the</strong> Dry<br />

Valleys region, Victoria Land, <strong>Antarctica</strong>, contain unusually<br />

well-exposed examples <strong>of</strong> such segregations<br />

(Marsh, 2004). Most are sub-horizontal 0·5^2 m thick<br />

lenses in <strong>the</strong> upper 100 m <strong>of</strong> <strong>the</strong> sills, containing crystals<br />

one to two orders <strong>of</strong> magnitude larger than those <strong>of</strong> <strong>the</strong><br />

host dolerite, and have sharp upper and lower contacts. A<br />

few form smaller sub-vertical veins and amorphous pods.<br />

Some have gradational contacts with <strong>the</strong> host dolerite.<br />

The compositions <strong>of</strong> <strong>the</strong>se segregations span <strong>the</strong> full range<br />

from <strong>the</strong> host dolerite to granite, and so <strong>of</strong>fer an unusually<br />

detailed window on differentiation mechanisms in basaltic<br />

magma systems.<br />

In this study we develop a new petrogenetic model for<br />

<strong>the</strong> formation <strong>of</strong> <strong>the</strong> segregations based on detailed field<br />

observations made by K. Zavala and B. D. Marsh in 2001,<br />

toge<strong>the</strong>r with petrographic data, mineral chemistry and<br />

bulk-rock major element geochemistry combined with<br />

MELTS modelling.<br />

GEOLOGICAL SETTING<br />

The FDS are well exposed in <strong>the</strong> McMurdo Dry Valleys<br />

region (Fig. 1) c. 20 km inland from <strong>the</strong> Ross Sea between<br />

<strong>the</strong> McKay Glacier to <strong>the</strong> north (c. 10 km north <strong>of</strong> Fig. 1b)<br />

and <strong>the</strong> <strong>Ferrar</strong> Glacier to <strong>the</strong> south. The basement in this<br />

region consists <strong>of</strong> Precambrian to Devonian metamorphic<br />

rocks and interbedded marbles, hornfels, and schists <strong>of</strong><br />

<strong>the</strong> Koettlitz group (Gunn & Warren, 1962; McKelvey &<br />

Webb, 1962) and undeformed granitic plutons <strong>of</strong> <strong>the</strong><br />

Granite Harbour Intrusive Complex (Gunn & Warren,<br />

1962). These units are overlain by gently dipping<br />

Devonian^Triassic sedimentary rocks known as <strong>the</strong><br />

Beacon Supergroup. These two main groups are separated<br />

by <strong>the</strong> Kukri Peneplain unconformity, which formed after<br />

<strong>the</strong> emplacement <strong>of</strong> <strong>the</strong> plutons and corresponds to a<br />

period <strong>of</strong> uplift and erosion. Both <strong>the</strong> basement and <strong>the</strong><br />

cover sequence are intruded by Jurassic ( 180 Ma) <strong>Ferrar</strong><br />

dolerites, which are exposed over an area <strong>of</strong> 10 5 km 2<br />

and have an estimated volume <strong>of</strong> 10 5 km 3 (Kyle, 1980).<br />

The dolerites form part <strong>of</strong> <strong>the</strong> <strong>Ferrar</strong> Large Igneous<br />

Province (FLIP), which extends more than 4000 km from<br />

<strong>the</strong> Weddell Sea to sou<strong>the</strong>astern Australasia. The FLIP<br />

occurs as an elongate belt along <strong>the</strong> Transantarctic<br />

Mountains and has been interpreted as forming in a<br />

failed rift tectonic setting (Elliot & Fleming, 2004) associated<br />

with <strong>the</strong> break-up <strong>of</strong> Gondwanaland (Brewer et al.,<br />

1992). The FDS represent by far <strong>the</strong> largest preserved<br />

volume <strong>of</strong> rock in <strong>the</strong> FLIP, which also includes <strong>the</strong> Dufek<br />

layered mafic intrusion, pyroclastic rocks and <strong>the</strong><br />

Kirkpatrick basalt flows (Elliot & Fleming, 2008). Based<br />

on 40 Ar/ 39 Ar geochronology, <strong>the</strong> duration <strong>of</strong> <strong>the</strong> magmatic<br />

activity was <strong>of</strong> <strong>the</strong> order <strong>of</strong> 1 Myr (Heimann et al., 1994;<br />

Fleming et al., 1997). Two sills and <strong>the</strong> Dufek intrusion<br />

have been dated at 183·6 1·8 Ma, based on U^Pb<br />

1928<br />

geochronology <strong>of</strong> zircon and baddeleyite, <strong>the</strong> same<br />

age as <strong>the</strong> Karoo igneous province in sou<strong>the</strong>rn Africa,<br />

demonstrating <strong>the</strong> vast size <strong>of</strong> <strong>the</strong> igneous province<br />

(Encarnacion et al., 1996; Minor & Mukasa, 1997).<br />

Emplacement paths for <strong>the</strong> <strong>Ferrar</strong> magmas are unknown:<br />

dyke swarms are not seen and dykes cutting basement<br />

rocks are uncommon. Elliot & Fleming (2004) advocated<br />

long distance (thousands <strong>of</strong> kilometres) crustal transport<br />

<strong>of</strong> magmas from a source suggested to be in <strong>the</strong> Weddell<br />

triple junction region through deep basement rocks, followed<br />

by localized vertical transport to shallower levels<br />

and <strong>the</strong>n lateral transport in flows and sills over scales <strong>of</strong><br />

a few hundred kilometres. The Dufek Massif is located at<br />

<strong>the</strong> edge <strong>of</strong> <strong>the</strong> Weddell Sea about 2000 km distant from<br />

<strong>the</strong> Dry Valleys. Though clearly part <strong>of</strong> <strong>the</strong> FLIP, Elliot &<br />

Fleming (2004) argued that, based on its stratigraphic<br />

placement above <strong>the</strong> basement rocks, <strong>the</strong> Dufek Massif is<br />

unlikely to be a feeder region for <strong>the</strong> FDS.<br />

Fieldwork by Gunn (1962), Gunn & Warren (1962),<br />

Hamilton (1965) and Marsh (2004) in <strong>the</strong> Taylor and<br />

Wright valleys identified a stack <strong>of</strong> four massive sills over<br />

a 4 km vertical extent (Fig. 1b and c) that are named,<br />

from top to bottom, <strong>the</strong> Mt. Fleming Sill (100 m thick),<br />

<strong>the</strong> Asgard Sill (250 m), <strong>the</strong> Peneplain Sill (PPS) (350 m),<br />

and <strong>the</strong> Basement Sill (BS) (350 m). Marsh (2004) proposed<br />

that each sill occurs as a series <strong>of</strong> lobes fed from a<br />

central location near <strong>the</strong> Dais, which is <strong>the</strong> lowest topographic<br />

exposure <strong>of</strong> <strong>the</strong> BS (Fig. 1b), and that at <strong>the</strong> time<br />

<strong>of</strong> emplacement <strong>the</strong>re was a stack <strong>of</strong> vertically interconnected<br />

sills and magma chambers within <strong>the</strong> crust forming<br />

a magmatic mush column. The upper three sills consist <strong>of</strong><br />

homogeneous fine- to medium-grained dolerite, whereas<br />

<strong>the</strong> lowermost BS has a central 50^250 m thick cumulate<br />

layer or ‘tongue’ consisting <strong>of</strong> large (1^20 mm) orthopyroxene<br />

and smaller (0·5^1mm) plagioclase crystals, enclosed<br />

by fine- to medium-grained dolerite like that <strong>of</strong> <strong>the</strong> o<strong>the</strong>r<br />

sills. <strong>Silicic</strong> segregations are common in <strong>the</strong> upper 100 m<br />

<strong>of</strong> <strong>the</strong> BS and PPS. This study focuses on <strong>the</strong> petrography,<br />

chemical composition and formation <strong>of</strong> <strong>the</strong> segregations<br />

in <strong>the</strong> BS. Studies by Be¤ dard et al. (2007) and Boudreau &<br />

Simon (2007) suggest that Fe-rich pegmatoids (here<br />

referred to as segregations) were formed from an evolved<br />

and degassed silicate liquid that was expelled from a compacting<br />

crystal mush below and that vapour-saturated<br />

zones, along with late interstitial liquid from <strong>the</strong> lower<br />

and central parts <strong>of</strong> <strong>the</strong> sill, may have ascended into localized<br />

regions in <strong>the</strong> upper zone to form <strong>the</strong>se pegmatoids.<br />

FIELD AND ANALYTICAL<br />

METHODS<br />

Sampling locations<br />

Samples <strong>of</strong> <strong>the</strong> segregations were collected from <strong>the</strong><br />

Taylor Valley region in 1993 by Marsh and Wheelock<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 1. (a) Map showing location <strong>of</strong> <strong>the</strong> Dry Valleys, <strong>Antarctica</strong>. Modified from Be¤ dard et al. (2007). (b) Geological map <strong>of</strong> <strong>the</strong> Dry<br />

Valleys Area. Modified from B. D. Marsh & M. J. Zieg (unpublished, with permission). White stars indicate <strong>the</strong> locations where detailed field<br />

observations were made. WBP, West Bull Pass; EBP, East Bull Pass; Dais, Dais Intrusion. Black stars indicate locations where analyzed samples<br />

were collected: SV, Simmons Valley; PV, Pearse Valley; SR, Solitary Rocks; McT, McMurdo Tip; MP, Mount Peleus. (c) Cross-section AB<br />

through Bull Pass. Elevation is relative to sea level. Simplified from McKelvey & Webb (1962).<br />

1929<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Table 1: Basement Sill (BS) segregations and adjacent dolerites<br />

Seg pr<strong>of</strong>ile Seg group Location Type Sample Sub-lenses D (m) Comments<br />

A S5 PV lens 93-BS-78 1-A 12 lens with medium- to<br />

S 3 1-B 12 coarse-grained<br />

S 5 2-A 12 sub-lenses<br />

S 3 2-B 12<br />

A-1 S 4 SV lens 93-BS-80 60·3 coarse-grained<br />

S3 lens 93-BS-81 60·6 medium-grained<br />

S4 lens 93-BS-82 60·82 coarse-grained<br />

S1 dolerite 93-BS-84 60 fine-grained above 80<br />

B S5 SV lens 93-BS-88 52·7 homogeneous<br />

S 4 lens 93-BS-89 A-1 52·7 seg with sub-lenses<br />

S 4 A-2 52·7<br />

S 4 A-3 52·7<br />

S 4 lens 93-BS-90 A 52·4 above and adjacent to 89<br />

S1 dolerite B 52·4<br />

S1 dolerite 93-BS-91 52·7 above and adjacent to 89<br />

S4 lens 93-BS-92 A-1 52·68 lens cut into<br />

S4 A-2 52·71 alternating mafic and<br />

S 4 A-3 52·76 felsic sub-lenses<br />

S 4 B-1 52·83<br />

S 4 B-2 52·91<br />

S 5 1-A-1 52·68<br />

S3 1-A-2 52·74<br />

S2 1-B-1-A 52·86<br />

S1 1-B-1-B 52·96<br />

S4 1-B-2-A 53·09<br />

S 1 1-B-2-B 53·25<br />

C S 1 SR lens 93-BS-95 24·2 homogeneous<br />

S 1 dolerite 93-BS-96 A-1 24·1 between two segs<br />

S 1 A-2 24·1<br />

S1 B-1 24·1<br />

S1 B-2 24·1<br />

D S4 SR lens 93-BS-97 23<br />

S1 dolerite 93-BS-99 A 24<br />

B 24·45 dolerite<br />

E * SR lens 93-BS-107 A-1 20 seg lens beneath<br />

A-2 20·5 grey felsic rock<br />

B 20·9<br />

F-1 S5 SR lens 93-BS-141 30·3 upper part <strong>of</strong> <strong>the</strong> BS<br />

F S4 SV lens 94-BS-150 A 93·8 lens<br />

S4 lens 94-BS-151 A 93·8 adjacent to 150-A<br />

S1 B 93·75<br />

S 5 lens 94-BS-152 93·8 same as 150-A<br />

S 1 lens 94-BS-153 93·42 0·4 m above 150-A<br />

S 2 lens 94-BS-154 A 93·8 lateral end <strong>of</strong> 150<br />

S 1 B 93·8<br />

1930<br />

(continued)<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/<br />

by guest on April 30, 2013


Table 1: Continued<br />

ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Seg pr<strong>of</strong>ile Seg group Location Type Sample Sub-lenses D (m) Comments<br />

G S1 SV lens 94-BS-155 60·3 top <strong>of</strong> seg lens<br />

S4 lens 94-BS-157 A-1 60·32 seg across all<br />

S3 A-2 60·36 lenses <strong>of</strong> 155<br />

S1 A-3 60·43<br />

S2 A-4 60·52<br />

S4 A-5 60·61<br />

S1 A-6 60·76<br />

S1 dolerite 94-BS-158 59·94 0·36 m above 155<br />

S1 dolerite 94-BS-159 59·98 0·32 above 155<br />

G-1 S1 SV lens 94-BS-160 A-1-1 60·38 spans seg 155, and 157<br />

S1 A-1-2 60·51<br />

S4 A-2-1 60·71<br />

S1 A-2-2 60·96<br />

S4 A-3-1 61·26<br />

S4 A-3-2 61·59<br />

S1 A-4-1 61·97<br />

S4 A-4-2 62·42<br />

S3 B-1-1 62·96<br />

S4 B-1-2 63·56<br />

S4 B-2-1 64·24<br />

S4 B-2-2 65·04<br />

I S1 SV lens 94-BS-210 A 38·96 shows directional growth<br />

S1 B 38·96<br />

S4 lens 94-BS-211 38·96 next to 210<br />

S5 lens 94-BS-212 38·96 next to 211<br />

J S4 SV lens 94-BS-219 A 23·7 adjacent to dolerite<br />

S1 B 23·7<br />

K S4 SV lens 94-BS-223 A 48·1 adjacent to dolerite<br />

S1 B 48·1<br />

S4 lens 94-BS-224 48·1 coarse-grained top part<br />

S4 lens 94-BS-225 47·9 med-grained lower part<br />

L S3 SV vein 94-BS-231 A 39 top <strong>of</strong> vertical 0·5 m vein<br />

S1 B 39·1<br />

S4 vein 94-BS-232 39·1 1 m vein next to 231<br />

S4 vein 94-BS-233 A 39·07 bottom <strong>of</strong> 232<br />

BS McT dolerite 96-A-26 0 chilled margin<br />

dolerite MP dolerite 96-A-33 120 dolerite midway in BS<br />

D, distance from <strong>the</strong> upper contact <strong>of</strong> BS.<br />

*Thin sections only.<br />

(Wheelock & Marsh, 1993) and in 2001 by Zavala and<br />

Marsh (Zavala & Marsh, 2001). Detailed sample descriptions<br />

are given inTable 1 and major element chemical analyses<br />

in Table 2. Sample locations are indicated by black<br />

stars in Fig. 1b. Twenty-nine <strong>of</strong> <strong>the</strong>se samples were collected<br />

in <strong>the</strong> Simmons Valley (SV) and one was collected in <strong>the</strong><br />

Pearse Valley (PV), both <strong>of</strong>fshoots <strong>of</strong> <strong>the</strong> Taylor Valley. Six<br />

1931<br />

samples were collected at Solitary Rocks (SR), a promontory<br />

at <strong>the</strong> junction <strong>of</strong> <strong>the</strong> Taylor and <strong>Ferrar</strong> glaciers.<br />

During <strong>the</strong> summer field season <strong>of</strong> 2003^2004 fur<strong>the</strong>r<br />

sampling and detailed field observations were made at different<br />

heights and locations along <strong>the</strong> Wright Valley where<br />

<strong>the</strong> upper and lower contacts <strong>of</strong> <strong>the</strong> BS are well exposed,<br />

particularly near West Bull Pass (WBP), East Bull Pass<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


Table 2: Representative major element compositions <strong>of</strong> host dolerites <strong>of</strong> groups S1^S5<br />

Pr<strong>of</strong>ile: SegA-1 Seg B Seg C Seg D Seg F Seg G Seg G-1 Seg G-1 Seg I Seg J Seg K Seg L av. min. max.<br />

Sample: 84 90-B 96-B-1 99-A 154-B 157-A-3 160-A1-1 160-A1-2 210-A 219-B 223-B 231-B<br />

Group S 1<br />

SiO 2 53·05 51·89 51·10 54·84 53·25 53·35 53·48 54·35 55·16 53·03 52·13 53·16 53·23 51·10 55·16<br />

TiO 2 0·53 0·38 0·36 0·69 0·49 0·62 0·62 0·75 0·71 0·46 0·47 0·48<br />

Al 2O 3 14·86 13·81 17·86 14·45 13·53 15·68 14·37 16·60 14·83 16·93 19·29 18·59<br />

FeOT 9·31 10·57 8·85 10·62 8·31 10·42 10·93 9·75 10·29 9·01 7·65 8·48<br />

Fe2O3 1·55 1·76 1·48 1·77 1·39 1·74 1·82 1·63 1·72 1·50 1·28 1·41<br />

FeO 7·91 8·98 7·52 9·03 7·06 8·86 9·29 8·29 8·75 7·66 6·50 7·21<br />

MnO 0·17 0·19 0·15 0·17 0·18 0·17 0·18 0·15 0·16 0·16 0·13 0·13<br />

MgO 8·16 9·33 6·63 5·61 9·10 5·31 6·00 4·43 4·93 5·84 4·83 4·79<br />

CaO 11·85 11·77 12·59 10·15 11·60 10·79 10·76 9·80 8·95 11·66 11·33 11·22<br />

Na 2O 1·45 1·33 1·62 1·81 1·43 2·05 1·72 2·22 2·14 2·07 2·16 2·07<br />

K 2O 0·57 0·42 0·44 1·01 0·57 0·81 0·88 0·99 1·00 0·72 0·72 0·73<br />

P2O5 0·06 0·04 0·05 0·10 0·07 0·09 0·08 0·12 0·12 0·07 0·07 0·07<br />

LOI 0·58 0·46 0·41 0·67 0·66 0·65 1·66 1·30 0·41 0·62 1·08 1·23<br />

Total 100·74 100·37 100·20 100·18 99·27 100·03 100·91 100·42 100·36 100·71 99·99 101·10<br />

Mg-no. 64·75 64·91 61·09 52·54 69·65 51·65 53·50 48·78 50·11 57·60 56·96 54·21 57·15 48·78 69·65<br />

Alkalis 2·02 1·75 2·06 2·82 2·00 2·86 2·60 3·21 3·14 2·79 2·88 2·80 2·58 1·75 3·21<br />

FI 1·49 1·36 1·65 1·89 1·47 2·11 1·79 2·30 2·22 2·12 2·21 2·12 1·89 1·36 2·30<br />

MI 0·53 0·53 0·57 0·65 0·48 0·66 0·65 0·69 0·68 0·61 0·61 0·64 0·61 0·48 0·69<br />

Pr<strong>of</strong>ile: Seg B Seg F Seg G av. min. max.<br />

Sample: 92-1-B-1-A 154-A 157-A-4<br />

Group S 2<br />

SiO 2 54·68 52·84 54·72 54·08 52·84 54·72<br />

TiO 2 0·96 0·43 0·88<br />

Al 2O 3 12·43 7·96 10·46<br />

FeOT 12·69 12·06 13·50<br />

Fe2O3 2·12 2·01 2·25<br />

FeO 10·79 10·25 11·48<br />

MnO 0·20 0·23 0·23<br />

MgO 6·47 13·02 6·81<br />

CaO 8·85 11·71 10·08<br />

Na 2O 1·75 0·82 1·46<br />

K 2O 0·83 0·41 0·87<br />

P2O5 0·12 0·05 0·09<br />

LOI 0·73 0·60 1·45<br />

Total 99·80 100·27 100·64<br />

JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Mg-no. 51·66 69·35 51·39 57·47 51·39 69·35<br />

Alkalis 2·58 1·23 2·33 2·05 1·23 2·58<br />

FI 1·82 0·85 1·53 1·40 0·85 1·82<br />

MI 0·66 0·48 0·66 0·60 0·48 0·66<br />

1932<br />

(continued)<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/<br />

by guest on April 30, 2013


Table 2: Continued<br />

Pr<strong>of</strong>ile: Seg A Seg A SegA-1 Seg B Seg G Seg G-1 Seg L av. min. max.<br />

Sample: 78-1-B 78-2-B 81 92-1-A-2 157-A-2 160-B1-1 231-A<br />

Group S 3<br />

SiO 2 55·56 55·11 56·48 55·21 56·03 55·14 57·33 55·72 55·14 56·48<br />

TiO 2 0·84 0·84 0·94 0·89 1·00 0·87 0·79<br />

Al 2O 3 14·34 14·38 13·75 14·46 14·78 14·49 13·93<br />

FeOT 10·62 11·05 10·79 11·05 11·76 11·08 9·96<br />

Fe2O3 1·77 1·84 1·80 1·84 1·96 1·85 1·66<br />

FeO 9·03 9·39 9·17 9·39 10·00 9·42 8·47<br />

MnO 0·16 0·16 0·17 0·17 0·16 0·16 0·16<br />

MgO 4·65 4·61 4·19 4·38 3·18 4·09 4·16<br />

CaO 9·21 9·32 9·11 9·38 7·96 8·89 8·84<br />

Na 2O 1·95 1·94 2·26 1·95 2·30 2·14 2·06<br />

K 2O 1·21 1·13 1·18 0·97 1·26 0·95 1·10<br />

P2O5 0·13 0·12 0·12 0·13 0·16 0·12 0·14<br />

LOI 1·78 1·50 1·10 1·39 1·12 0·80 0·69<br />

Total 100·53 100·29 100·15 100·07 99·81 99·67 99·32<br />

Mg-no. 47·86 46·65 44·87 45·38 36·18 43·62 46·68 42·51 36·18 45·38<br />

Alkalis 3·16 3·07 3·44 2·92 3·56 3·09 3·16 3·25 2·92 3·56<br />

FI 2·05 2·03 2·35 2·03 2·41 2·22 2·15 2·25 2·03 2·41<br />

MI 0·70 0·71 0·72 0·72 0·79 0·73 0·71 0·74 0·72 0·79<br />

Pr<strong>of</strong>ile: SegA-1 SegA-1 Seg B Seg D Seg F Seg G Seg G-1 Seg G-1 Seg I Seg J SegK SegL Seg L av. min. max.<br />

Sample: 80 82 90-A 97 151-A 157-A-5 160-B1-2 160-B2-2 211 219-A 225 232 233-A<br />

Group S 4<br />

ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

SiO 2 58·52 58·18 62·92 57·92 60·37 56·78 60·77 54·43 57·46 58·47 57·20 55·31 61·76 58·19 54·43 62·92<br />

TiO 2 1·30 1·10 1·26 1·39 1·31 0·87 1·15 1·42 1·78 0·97 1·10 0·81 1·10<br />

Al 2O 3 12·56 13·55 12·19 12·30 12·68 12·33 11·91 11·13 13·03 14·14 14·02 10·17 12·67<br />

FeOT 12·01 11·01 10·31 13·20 10·82 11·47 11·35 15·14 13·66 10·99 10·86 12·39 10·07<br />

Fe2O3 2·00 1·84 1·72 2·20 1·80 1·91 1·89 2·52 2·28 1·83 1·81 2·07 1·68<br />

FeO 10·21 9·36 8·76 11·22 9·20 9·75 9·65 12·87 11·61 9·34 9·23 10·53 8·56<br />

MnO 0·17 0·17 0·15 0·19 0·15 0·19 0·16 0·23 0·18 0·16 0·15 0·23 0·15<br />

MgO 3·12 3·21 2·42 2·83 2·89 5·17 2·70 4·55 2·51 3·25 3·24 7·22 3·10<br />

CaO 7·18 7·68 5·76 6·90 6·30 8·90 6·19 8·92 7·31 7·66 7·44 10·09 6·02<br />

Na 2O 2·41 2·44 2·48 2·34 2·36 1·88 2·38 1·82 2·35 2·41 2·28 1·44 2·31<br />

K 2O 1·32 1·40 1·62 1·51 1·59 1·12 1·51 0·93 1·29 1·34 1·25 0·87 1·42<br />

P2O5 0·18 0·16 0·20 0·20 0·21 0·13 0·20 0·12 0·21 0·17 0·18 0·07 0·19<br />

LOI 1·32 1·13 1·12 1·16 1·13 1·55 1·41 1·39 0·71 1·02 0·98 1·44 1·97<br />

Total 100·29 100·21 100·60 100·15 99·99 100·58 99·92 99·81 100·72 100·76 98·88 99·24 100·92<br />

Mg-no. 35·25 37·93 32·98 31·00 35·89 48·58 33·27 38·65 27·81 38·27 38·47 54·98 39·22 37·76 27·81 54·98<br />

Alkalis 3·73 3·84 4·10 3·85 3·95 3·00 3·89 2·75 3·64 3·75 3·53 2·31 3·73 3·53 2·31 4·10<br />

FI 2·53 2·56 2·64 2·48 2·52 1·97 2·53 1·90 2·47 2·53 2·39 1·51 2·46 2·34 1·51 2·64<br />

MI 0·79 0·77 0·81 0·82 0·79 0·69 0·81 0·77 0·84 0·77 0·77 0·63 0·76 0·77 0·63 0·84<br />

1933<br />

(continued)<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/<br />

by guest on April 30, 2013


Table 2: Continued<br />

Pr<strong>of</strong>ile: Seg A Seg A Seg B Seg B Seg F-1 Seg F Seg I av. min. max.<br />

Sample: 78-1-A 78-2-A 88 92-1-A-1 141 152 212<br />

Group S 5<br />

JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

SiO 2 66·95 64·52 64·07 59·00 59·28 59·09 64·78 62·53 59·00 66·95<br />

TiO 2 1·01 1·28 1·54 1·90 1·44 1·32 1·35<br />

Al 2O 3 11·60 12·09 11·51 12·55 13·08 13·49 11·55<br />

FeOT 6·88 8·40 10·98 12·60 10·61 10·55 10·66<br />

Fe2O3 1·15 1·40 1·83 2·10 1·77 1·76 1·78<br />

FeO 5·85 7·14 9·33 10·71 9·02 8·97 9·06<br />

MnO 0·11 0·13 0·14 0·18 0·16 0·16 0·13<br />

MgO 1·94 2·28 1·67 2·20 2·99 3·06 0·99<br />

CaO 3·50 4·40 4·28 5·76 6·07 6·38 3·61<br />

Na 2O 3·52 3·26 2·98 2·39 2·62 2·40 3·06<br />

K 2O 1·40 1·45 1·43 1·36 2·18 1·62 2·49<br />

P2O5 0·25 0·29 0·27 0·24 0·19 0·20 0·35<br />

LOI 1·82 2·24 1·66 1·12 1·14 1·45 0·97<br />

Total 99·09 100·47 100·71 99·51 99·94 99·90 100·11<br />

Mg-no. 37·15 36·26 24·17 26·79 37·13 37·81 16·29 30·80 16·29 37·81<br />

Alkalis 4·92 4·71 4·41 3·75 4·80 4·02 5·55 4·59 3·75 5·55<br />

FI 3·69 3·42 3·14 2·53 2·82 2·56 3·33 3·07 2·53 3·69<br />

MI 0·78 0·79 0·87 0·85 0·78 0·78 0·92 0·82 0·78 0·92<br />

FI ¼ (Na2O þ K2O)/(Na2O þ K2O þ CaO); MI ¼ [FeOT/(FeOT þ MgO)].<br />

(EBP), and 12 km west <strong>of</strong> Bull Pass area at <strong>the</strong> Dais.<br />

These locations are indicated by white stars in Fig. 1b.<br />

Although <strong>the</strong>re are as yet no chemical data for this set <strong>of</strong><br />

rocks, <strong>the</strong> detailed field observations made during this<br />

field season have provided valuable new information<br />

about <strong>the</strong> morphology <strong>of</strong> <strong>the</strong> segregations in <strong>the</strong> sill.<br />

Sampling methods and nomenclature<br />

Sample collection was started at <strong>the</strong> upper contact <strong>of</strong> <strong>the</strong><br />

BS and continued down section about 300 m (Table 1).<br />

Sample size varied between tens <strong>of</strong> centimetres to 2 cm, depending<br />

on <strong>the</strong> thickness and texture <strong>of</strong> <strong>the</strong> segregation.<br />

Some samples consist <strong>of</strong> several layers <strong>of</strong> coarse-grained<br />

felsic rock alternating with medium- to fine-grained mafic<br />

rock (Fig. 2a), whereas o<strong>the</strong>rs comprise homogeneous<br />

coarse-grained felsic rock (Fig. 2b).<br />

We use <strong>the</strong> term ‘segregation pr<strong>of</strong>ile’ for a vertical sequence<br />

<strong>of</strong> samples from a single locality. For example, segregation<br />

pr<strong>of</strong>ile A-1 consists <strong>of</strong> segregation samples<br />

93-BS-80, 93-BS-81 and 93-BS-82 and <strong>the</strong> overlying dolerite<br />

sample 93-BS-84. Most segregations have textural variations;<br />

<strong>the</strong> letters A and B indicate layers with different<br />

textures within <strong>the</strong> pr<strong>of</strong>ile. Trailing numbers after <strong>the</strong><br />

letter refer to any additional textural variations within a<br />

layer. For example, segregation sample 94-BS-160 was cut<br />

1934<br />

into six layers, 94-BS-160-A-1 to A-4, and 94-BS-160-B-1<br />

and B-2 (Fig. 2a). The layer described by 94-BS-160-A-1<br />

was fur<strong>the</strong>r subdivided into two sub-layers 94-BS-160-A-<br />

1-1 and 94-BS-160-A-1-2, based on observed textural<br />

variations.<br />

Analytical methods<br />

Samples are typically fresh and non-vesicular, although<br />

some have micro-fractures with minor alteration. Samples<br />

were cut taking into account <strong>the</strong> textural and contact relationships<br />

observed in <strong>the</strong> field and in hand specimen.<br />

Thin sections and chemical analyses were made <strong>of</strong> each<br />

layer in each segregation and <strong>of</strong> <strong>the</strong> adjacent dolerite. For<br />

chemical analyses, samples were crushed to a mesh size <strong>of</strong><br />

80 mm, using a ball mill and a nickel shatter box. The powders<br />

were prepared for major element analyses by X-ray<br />

fluorescence (XRF) at Franklin and Marshall College following<br />

<strong>the</strong> procedure <strong>of</strong> Boyd & Mertzman (1987). The<br />

procedure involved melting <strong>the</strong> powders using a cadmium^nickel<br />

charge, <strong>the</strong>n quenching <strong>the</strong> liquid into a<br />

solid disc. Each analysis was run three times and <strong>the</strong> average<br />

value was used. Relative analytical error is between<br />

0·11 and 0·26% for all major elements, and accuracy<br />

is 0·2%, estimated by analyzing accepted standards<br />

(Abbey, 1983; Govindaraju, 1994). FeO is reported as FeO T,<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/<br />

by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 2. (a) Segregation sample 94-BS-160 (pr<strong>of</strong>ile G-1) has internal<br />

stratification consisting <strong>of</strong> multiple felsic coarse-grained horizons<br />

(160: A-2, A-3, and B-1) and less felsic medium-grained (160: A-1, A-4,<br />

and B-2). (b) Coarse-grained segregation with no internal stratification.<br />

<strong>Segregations</strong> have sharp contacts with <strong>the</strong> host dolerite and are<br />

composed <strong>of</strong> coarse pyroxene and plagioclase crystals with minor<br />

opaque minerals and variable amounts <strong>of</strong> granophyric intergrowths<br />

<strong>of</strong> quartz and K-feldspar.<br />

1935<br />

Table 3: <strong>Segregations</strong> and dolerite modes obtained by <strong>the</strong><br />

SEM^MLA method (see text for details)<br />

Sample: 93-BS-96-B-1 94-BS-154 93-BS-78-1-B 93-BS-88 93-BS-78-1-A<br />

Group: S1 S2 S3 S5 S5<br />

augite 21·1 24·5 16·1 1·8 2<br />

pigeonite 12·1 10·7 1·6 0·1 0·2<br />

orthopyroxene 4·0 6·4 0·2 0 0·1<br />

plagioclase 49·3 46·3 40·3 14·2 14·3<br />

K-feldspar 2·0 2·4 5 8·1 7·1<br />

quartz 5·0 5·3 17·1 59·7 63·1<br />

Ca-amphibole 6·0 4 18·1 12·5 11·2<br />

Fe–Ti oxides 0·4 0·3 1·4 3·5 2<br />

and an FeO/FeOT ratio <strong>of</strong> 0·85 is assumed (Ragland, 1989,<br />

p. 36).<br />

Mineral compositions <strong>of</strong> <strong>the</strong> major phases in segregation<br />

pr<strong>of</strong>ile A-1 (samples 93-BS-80, 81, 82 and adjacent dolerite<br />

93-BS-84) were obtained using <strong>the</strong> JEOL 8600 electron<br />

microprobe at Johns Hopkins University. The microprobe<br />

was operated at 15 kV with a beam current <strong>of</strong> 20 nA. The<br />

beam size was 5 mm for augite, diopside and pigeonite,<br />

and 10 mm for plagioclase and quartz and K-feldspar intergrowths.<br />

Counting times were 60 s at <strong>the</strong> peak and 30 s<br />

for background. Data were reduced using <strong>the</strong> CITZAF<br />

program <strong>of</strong> Armstrong (1989). Analytical relative errors<br />

for major elements are considered to vary between 0·5<br />

and 1%. Pyroxene and plagioclase crystals were probed<br />

in transverse sections to check for compositional zoning.<br />

High-resolution imaging and automated quantitative estimates<br />

<strong>of</strong> <strong>the</strong> modal mineralogy <strong>of</strong> selected thin sections<br />

was carried out at Memorial University using an FEI<br />

Quanta 400 scanning electron microprobe (SEM)^mineral<br />

liberation analyzer (MLA). The data are reported in<br />

Table 3.<br />

Sample grouping<br />

If segregations are <strong>the</strong> result <strong>of</strong> fractional crystallization <strong>of</strong><br />

<strong>the</strong> host dolerite, <strong>the</strong>n mass balance requires that <strong>the</strong> composition<br />

<strong>of</strong> <strong>the</strong> segregations is equal to <strong>the</strong> composition <strong>of</strong><br />

an initial, fractionated liquid plus or minus <strong>the</strong> crystallized<br />

phases entrained into or lost from this liquid. A set <strong>of</strong><br />

MELTS modeled liquids and crystals fractionating between<br />

12008C and 9008C were linearly regressed to<br />

<strong>the</strong> final bulk composition <strong>of</strong> <strong>the</strong> segregations, using <strong>the</strong><br />

singular value decomposition technique (Zavala, 2005).<br />

This technique has previously been used by Fisher (1989)<br />

to analyse metamorphic assemblages. The best-fit model<br />

to a given segregation provided <strong>the</strong> temperature at which<br />

<strong>the</strong> residual liquid separated from <strong>the</strong> modeled liquids<br />

and <strong>the</strong> proportions <strong>of</strong> minerals subsequently lost through<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Table 4: Definitions <strong>of</strong> segregation groups<br />

Group MML Plagioclase Augite Pigeonite<br />

S1 þ þ þ þ<br />

S 2 þ – þ þ<br />

S 3 þ þ þ –<br />

S 4 þ – þ –<br />

S 5 þ – – –<br />

MML (melts modeled liquids); þ, added; , subtracted.<br />

crystallization and/or gained through entrainment. When<br />

<strong>the</strong> segregations are grouped according to whe<strong>the</strong>r <strong>the</strong>y<br />

had lost or gained <strong>the</strong> major mineral phases plagioclase,<br />

augite and pigeonite <strong>the</strong>y fall into five categories, given in<br />

Table 4. Group S 1 corresponds to <strong>the</strong> host dolerite: that is,<br />

compositions are best fitted by a MELTS model liquid<br />

(MML) plus all three crystallizing phases. The segregations<br />

S 2^S 5 correspond to more fractionated rocks, in<br />

which one or more <strong>of</strong> <strong>the</strong> crystallizing phases has segregated<br />

from <strong>the</strong> MML. GroupS 5 is <strong>the</strong> most fractionated,<br />

corresponding to segregations that are best modelled as<br />

pure residual liquids. More detailed description <strong>of</strong> <strong>the</strong><br />

technique has been given by Zavala (2005). The designated<br />

group is given for each sample in Table 1, and samples are<br />

arranged by group inTable 2.<br />

LOCATION AND MORPHOLOGY<br />

OF SEGREGATIONS IN THE<br />

BASEMENT SILL<br />

Sets <strong>of</strong> sub-horizontal, anastomosing, 0·1^2 m thick lenses<br />

are common in <strong>the</strong> upper 100 m <strong>of</strong> <strong>the</strong> BS and make up<br />

about 15% by volume <strong>of</strong> <strong>the</strong> sill. The most common types<br />

<strong>of</strong> segregation in both Taylor Valley and Wright Valley are<br />

sub-horizontal lenses (Fig. 3c and d) that extend for tens<br />

<strong>of</strong> metres along strike and are connected to o<strong>the</strong>r lenses<br />

by smaller sub-vertical veins. Small ( 5^10 cm) felsic<br />

pods (Fig. 3a) and thin ( 2^10 cm) felsic stringers<br />

(Fig. 3b), which occur above and adjacent to <strong>the</strong>se lenses,<br />

are irregular features that are usually located a few<br />

metres ( 5^15 m) below <strong>the</strong> upper contact. Overall segregation<br />

textures, morphologies, and field relationships are<br />

similar throughout <strong>the</strong> BS; however, at <strong>the</strong> Dais, which is<br />

<strong>the</strong> lowest topographic exposure <strong>of</strong> <strong>the</strong> BS and consists <strong>of</strong><br />

an orthopyroxene- and plagioclase-rich cumulate body,<br />

<strong>the</strong>re is a larger variation in segregation morphology that<br />

varies systematically with distance from <strong>the</strong> upper contact<br />

(Fig. 4).<br />

1936<br />

Dais segregations<br />

In <strong>the</strong> interval 5^7 m below <strong>the</strong> upper contact <strong>of</strong> <strong>the</strong> BS at<br />

<strong>the</strong> Dais, thin (1^5 cm) sub-horizontal felsic streaks<br />

(4^10 cm long) are ubiquitous (Fig. 4a). These give way<br />

10 and 15 m below <strong>the</strong> upper contact to sparse 1^5 cm<br />

diameter felsic, pod-shaped, segregations (Fig. 4b) oriented<br />

parallel to <strong>the</strong> upper contact <strong>of</strong> <strong>the</strong> sill. Some pods are connected<br />

by 1^5 cm thick and 10^15 cm long undulating<br />

stringers. Both pods and stringers have sharp contacts<br />

with <strong>the</strong> host-rock. Between 15 and 20 m below <strong>the</strong> upper<br />

contact, segregations form 3^10 cm diameter sub-vertical<br />

pipes and 15^20 cm wide by 20^50 cm long sub-vertical<br />

veins that swell and pinch out laterally forming a<br />

saw-tooth pattern (Fig. 4c). Finally, about 30 m below <strong>the</strong><br />

upper contact, <strong>the</strong>re is a thick ( 300 cm), anastomosing,<br />

segregation lens, which extends laterally for more than<br />

20 m (Fig. 4d) and has undulating sharp upper and lower<br />

contacts with <strong>the</strong> host-rock. The upper contact <strong>of</strong> this massive<br />

layer with <strong>the</strong> host-rock is convex downward (white<br />

dashed line in Fig. 4d), and texturally <strong>the</strong> layer shows internal<br />

stratification, with a coarse-grained horizon in <strong>the</strong><br />

middle that becomes progressively finer-grained towards<br />

<strong>the</strong> margins.<br />

PETROGRAPHY OF THE SILICIC<br />

SEGREGATIONS<br />

As described above, <strong>the</strong> dolerites and segregations have<br />

been subdivided into five groups (S 1^S 5) based on<br />

mass-balance relationships. Petrographic study <strong>of</strong> samples<br />

from each group indicates that <strong>the</strong>re are distinct mineralogical<br />

and textural features associated with each group.<br />

Table 3 gives <strong>the</strong> mineral modes <strong>of</strong> representative samples<br />

obtained by <strong>the</strong> SEM^MLA method and Table 5 provides<br />

a summary <strong>of</strong> <strong>the</strong> petrography <strong>of</strong> each group.<br />

S 1 rocks (dolerites) are characterized by fine- to<br />

medium-grained ophitic textures (Fig. 5a) and contain<br />

prismatic (0·5^4 mm) euhedral to subhedral moderately<br />

embayed augite phenocrysts, smaller (0·1^3 mm) equant,<br />

subhedral^anhedral, fractured orthopyroxene phenocrysts,<br />

all enclosed by a network <strong>of</strong> (0·05^1mm) plagioclase<br />

laths. There are minor amounts <strong>of</strong> interstitial quartz and<br />

alkali feldspar in granophyric intergrowths and a few<br />

(0·01^0·1mm) grains <strong>of</strong> Fe^Ti oxide (ilmenite and<br />

magnetite). A representative modal analysis <strong>of</strong> <strong>the</strong> dolerite<br />

(sample 93-BS- 96-B-1) is given in Table 3. Intergrowths <strong>of</strong><br />

quartz and alkali feldspar are commonly hard to distinguish<br />

optically but <strong>the</strong>y are readily identified with <strong>the</strong><br />

SEM. Ca-amphibole occurs as an alteration product<br />

around pyroxene crystals ra<strong>the</strong>r than as a primary phase.<br />

The S2 segregations have fine- to medium-grained ophitic<br />

to sub-ophitic textures with sub-equal amounts <strong>of</strong> pyroxene<br />

(low- and high-Ca pyroxene) and plagioclase<br />

(Fig. 5b). Low-Ca pyroxene crystals are 1^5 mm long,<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/<br />

by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 3. Field photographs showing that segregations can occur as small ( 5^10 cm) felsic pods (a) and thin irregular ( 5 cm) stringers subparallel<br />

to <strong>the</strong> upper contact <strong>of</strong> <strong>the</strong> BS (b). Between 30 and 100 m below <strong>the</strong> upper contact segregations form (0·5^2 m) thick, anastomosing, climbing<br />

felsic lenses (c, d) that are <strong>of</strong>ten interconnected to o<strong>the</strong>r smaller lenses above by thin sub-vertical veins. Stratigraphic section shows <strong>the</strong><br />

locations <strong>of</strong> <strong>the</strong> photographs.<br />

equant and partly embayed, with fractures and reaction<br />

rims consisting <strong>of</strong> dark alteration minerals, and contain<br />

fine exsolution lamellae <strong>of</strong> pigeonite. O<strong>the</strong>r low-Ca pyroxene<br />

crystals are subhedral, moderately embayed and contain<br />

abundant fine exsolution lamellae <strong>of</strong> augite parallel<br />

to (001) and inclined to (100). Augite crystals are 2^5 mm<br />

long, euhedral to subhedral, prismatic and less embayed<br />

and compositionally twinned. They commonly have exsolution<br />

lamellae <strong>of</strong> pigeonite parallel to (001), as confirmed<br />

by <strong>the</strong> different spectral patterns in <strong>the</strong> SEM^<br />

MLA, and some have inverted pigeonite cores with complex<br />

twinning patterns. Plagioclase crystals are 0·4<br />

1mm long, euhedral to subhedral, commonly bladed<br />

and normally zoned. A representative modal analysis <strong>of</strong><br />

this rock (sample 94-BS-154) is given in Table 3. This rock,<br />

unlike <strong>the</strong> o<strong>the</strong>rs, has sub-rounded orthopyroxene crystals<br />

with reaction rims and fractures filled with alteration<br />

1937<br />

minerals; <strong>the</strong>se resemble <strong>the</strong> orthopyroxene cumulate crystals<br />

found in <strong>the</strong> opx tongue.<br />

The S 3 segregations have sub-ophitic and poikilophitic<br />

textures consisting <strong>of</strong> approximately sub-equal proportions<br />

<strong>of</strong> plagioclase and augite (plus minor pigeonite), and between<br />

20 and 30% vermicular intergrowths <strong>of</strong> quartz and<br />

alkali feldspar (Fig. 5c). The rims <strong>of</strong> <strong>the</strong> plagioclase crystals<br />

are commonly optically continuous with <strong>the</strong> granophyric<br />

intergrowths. In <strong>the</strong> poikilophitic textures<br />

plagioclase and augite crystals have similar size: <strong>the</strong> poikilophitic<br />

domains consist <strong>of</strong> medium- to coarse-grained<br />

prismatic augite interlocking with tabular plagioclase<br />

crystals; granophyric intergrowths fill <strong>the</strong> interstices.<br />

The sub-ophitic domains consist <strong>of</strong> large ( 0·5mm)<br />

augite crystals embedded with smaller (0·1^1mm) plagioclase<br />

laths. In addition to <strong>the</strong>ir occurrence in granophyric<br />

textures, quartz and alkali feldspar form anhedral,<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 4. At <strong>the</strong> Dais in <strong>the</strong>Wright Valley, segregations vary in morphology with increasing distance from <strong>the</strong> upper contact. (a) Close (5^7 m) to<br />

<strong>the</strong> upper contact segregations occur as (1^5 cm) thin felsic streaks; (b) about 10^15 m below <strong>the</strong> upper contact segregations are ( 5 cm)<br />

pod-shaped. (c) Between 15 and 20 m below <strong>the</strong> upper contact segregations form pipes and (15^20 cm) sub-vertical anastomosing lenses.<br />

(d) Below 30 m, an unusually ( 3·5 m) thick felsic lens with variable grain size forms a sharp upper wavy contact with <strong>the</strong> host-rock. All segregation<br />

types have sharp contacts with <strong>the</strong> host-rock at this location. Stratigraphic section shows <strong>the</strong> locations <strong>of</strong> <strong>the</strong> photographs.<br />

sub-rounded grains from 0·1to0·5 mm in size that appear<br />

to be recrystallized. Pyroxene crystals are commonly<br />

rimmed with Ca-amphibole and biotite and <strong>the</strong> size and<br />

amount <strong>of</strong> Fe^Ti oxide grains is higher than in S 1 and S 2.<br />

Figure 5c illustrates <strong>the</strong> poikilophitic texture within a<br />

coarser felsic segregation (94-BS-231-A) in <strong>the</strong> upper left<br />

corner, and <strong>the</strong> sub-ophitic texture associated with <strong>the</strong><br />

dolerite (94-BS-231-B) in <strong>the</strong> lower right corner. Between<br />

<strong>the</strong>se two textural zones <strong>the</strong>re is a contact zone consisting<br />

predominantly <strong>of</strong> fine-grained plagioclase laths.<br />

<strong>Segregations</strong> in <strong>the</strong> S 3 group have higher modal amounts<br />

<strong>of</strong> quartz and alkali feldspar in granophyric intergrowths<br />

than <strong>the</strong> previous two groups (e.g. see Table 4). This group<br />

1938<br />

also has more alteration rims <strong>of</strong> Ca-amphibole and biotite<br />

around subhedral pyroxene crystals.<br />

The S 4 group is composed mostly <strong>of</strong> poikilophitic domains<br />

with some sub-ophitic domains; <strong>the</strong> amount <strong>of</strong><br />

granophyric quartz and K-feldspar intergrowths and alteration<br />

is considerably greater than in all previous groups<br />

(Fig. 5d). The poikilophitic domains consist <strong>of</strong> interlocking,<br />

1^8 mm, subhedral to anhedral, altered augite and<br />

0·5^7 mm euhedral to subhedral, tabular, fractured plagioclase<br />

surrounded by medium- to coarse-grained granophyric<br />

intergrowths. The augite and minor inverted<br />

pigeonite crystals, which are subhedral to anhedral and<br />

partially to completely replaced by Ca-amphibole, biotite,<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Table 5: Summary <strong>of</strong> petrographic observations <strong>of</strong> <strong>the</strong> segregations and adjacent dolerites<br />

Group Rock type Texture Mineral size Characteristics<br />

S1 dolerite ophitic with minor sub-ophitic<br />

domains<br />

S2 segregation sub-ophitic with minor ophitic<br />

domains<br />

S3 segregation sub-ophitic with poikilophitic<br />

domains<br />

S4 segregation Poikilophitic with granophyric<br />

domains<br />

S5 segregation Granophyric with few crystalline<br />

domains<br />

epidote, actinolite and chlorite, commonly form bleb-like<br />

intergrowths with fractured and altered plagioclase crystals.<br />

The Fe^Ti oxides grains (mostly ilmenite and magnetite)<br />

are larger (0·5 mm) and are always associated with<br />

<strong>the</strong> altered pyroxene and with <strong>the</strong> granophyric intergrowths.<br />

The plagioclase crystals contain some sericite,<br />

minor titanite and epidote alteration, and are commonly<br />

fractured. Whereas some crystals are euhedral and not resorbed,<br />

o<strong>the</strong>rs have inclusions <strong>of</strong> granophyric material<br />

through <strong>the</strong>ir cores and margins that are optically continuous<br />

with <strong>the</strong> granophyric intergrowths. Some crystals<br />

have kinked twins and are bent or broken (Fig. 6). The<br />

coarse-grained granophyric intergrowths are more pervasive,<br />

with fewer crystalline domains forming interstitially,<br />

and <strong>the</strong> K-feldspar in <strong>the</strong>m is commonly altered to sericite,<br />

imparting a brown color, whereas <strong>the</strong> quartz remains unaltered.<br />

There is no modal analysis available for this<br />

sample.<br />

The S5 group is distinguished by having more than 50%<br />

<strong>of</strong> <strong>the</strong> mode composed <strong>of</strong> granophyre (Fig. 5e and f), consisting<br />

<strong>of</strong> both coarse-grained vermicular intergrowths <strong>of</strong><br />

K-feldspar and quartz and large (0·5^10 mm) sub-rounded<br />

crystals <strong>of</strong> <strong>the</strong>se phases. Typically a few, long (5^20 mm)<br />

fine to medium euhedral–subhedral (0·5–4 mm) augite<br />

subhedral, equant (0·1–3 mm) pigeonite<br />

euhedral bladed (0·05–1 mm) plagioclase<br />

minor interstitial granophyric textures and opaque minerals<br />

fine to medium euhedral–equant (1–5 mm) orthopyroxene and<br />

pigeonite<br />

euhedral, tabular (2–5 mm) augite<br />

euhedral bladed and tabular (0·4–1 mm) plagioclase<br />

minor opaque minerals<br />

fine to medium euhedral–subhedral, prismatic, (0·5–5 mm) augite<br />

euhedral–subhedral equant (2–5 mm) pigeonite<br />

euhedral bladed and tabular (0·1–5 mm) plagioclase<br />

interstitial alkali feldspar and quartz intergrowths<br />

minor opaque minerals<br />

medium to coarse altered, subhedral, prismatic (1–8 mm) augite<br />

euhedral–<br />

subhedral, tabular (0·5–7 mm) plagioclase<br />

medium- to coarse-grained granophyric textures<br />

minor opaque minerals (0·5–1 mm)<br />

medium to coarse euhedral–subhedral, resorbed (5–20 mm) plagioclase<br />

anhedral, altered, intergrowths <strong>of</strong> augite and opaque<br />

minerals<br />

quartz and K-feldspars ( 5 mm)<br />

1939<br />

euhedral, moderately fractured plagioclase crystals are<br />

present, some <strong>of</strong> which may contain inclusions and have<br />

rims partially replaced by granophyric intergrowths. The<br />

pyroxene crystals are completely replaced by intergrowths<br />

<strong>of</strong> Ca-amphibole, biotite, minor phlogopite and chlorite,<br />

and form intergrowths with <strong>the</strong> plagioclase crystals and<br />

<strong>the</strong> granophyric textures. Fe^Ti oxides (ilmenite and magnetite)<br />

are also coarse-grained (up to 5 mm) and are<br />

associated with <strong>the</strong> replaced pyroxenes. The measured<br />

modal compositions <strong>of</strong> two representative samples <strong>of</strong> this<br />

group (93-BS-88 and 93-BS-78-1-A) are given in Table 3.<br />

The S 5 group is distinguished from <strong>the</strong> S 4 group by<br />

having higher amounts <strong>of</strong> quartz and alkali feldspar intergrowths<br />

and fewer crystalline domains, and from groups<br />

S 1 to S 3 by having larger crystal sizes, significantly less pyroxene,<br />

more alteration minerals and more granophyric<br />

intergrowths.<br />

Summary<br />

Overall <strong>the</strong>re is a gradual increase in <strong>the</strong> grain size and<br />

amount <strong>of</strong> granophyre and alteration minerals from <strong>the</strong> S 1<br />

to <strong>the</strong> S 5 group. This is particularly evident in groups S 3<br />

to S 5, which progressively contain more regions with large<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 5. Representative photomicrographs <strong>of</strong> petrographic groups S 1^S 5.(a)S 1 sample 93-BS-96-B-1; (b) S 2 sample 94-BS-154-A; (c) S 3 sample<br />

94-BS-231-A; (d) S 4 sample 93-BS-97; (e) S 5 sample 93-BS-88; (f) S 5 sample 94-BS-212. Groups S 1 and S 2 have fine- to medium-grained ophitic<br />

and subophitic textures with sub-equal amounts <strong>of</strong> plagioclase and pyroxenes and minor quartz and K-feldspar intergrowths (a, b). Groups S3<br />

and S4 have medium- to coarse-grained poikilophitic textures with minor subophitic domains and increasing amounts <strong>of</strong> quartz and alkali feldspar<br />

granophyric intergrowths and alteration minerals (c, d). Group S5 consists mostly <strong>of</strong> granophyric intergrowths with few crystalline domains<br />

(e, f) and greater amount <strong>of</strong> alteration and opaque minerals. Common phases are plagioclase (plg), augite (aug), pigeonite (pig),<br />

orthopyroxene (opx) and granophyric quartz^K-feldspar intergrowths (gr). Minor phases are epidote (ep), biotite (bt), and opaque minerals<br />

(opq).<br />

1940<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 6. (a, b) Representative photomicrographs <strong>of</strong> pyroxene crystals with curved twin planes: (a) sample 93-BS-107-A-1; (b) sample 93-BS-97<br />

(S 4). (c, d) Broken and partially resorbed plagioclase crystals: (c) sample 94-BS-157-A-2 (group S 3); (d) sample 93-BS-92-1-B-2-B (S 1). (e^h)<br />

O<strong>the</strong>r plagioclase crystals are bent, have kinked twins, and have inclusion <strong>of</strong> granophyric intergrowths: (e) sample 93-BS-92-A-2 (S 4);<br />

(f) sample 94-BS-157-A-2 (S3); (g) sample 94-BS-160-B-1-1 (S3); (h) sample 94-BS-211 (S4).<br />

1941<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/<br />

by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

interlocking fractured plagioclase and altered pyroxene<br />

crystals enclosing or enclosed by quartz and alkali feldspar<br />

in granophyric intergrowths. The coarse-grained plagioclase<br />

and augite crystals in groups S3 and S4, <strong>the</strong> large<br />

( 2 cm) euhedral plagioclase crystals and increasing<br />

amount <strong>of</strong> granophyric intergrowths in group S5, and <strong>the</strong><br />

increasing proportion <strong>of</strong> alteration minerals in all three<br />

groups S3 to S5, suggest that minor amounts <strong>of</strong> volatiles<br />

were present during <strong>the</strong> formation <strong>of</strong> <strong>the</strong>se segregations.<br />

In groups S 1^S 4 some augite crystals have curved twins<br />

(Fig. 6a and b) and some plagioclase crystals are bent<br />

(Fig. 6f) or broken (Fig. 6c, d and h), and have kinked<br />

twins (Fig. 6c^h). The bent and deformed plagioclase crystals<br />

contain recrystallized pockets <strong>of</strong> granophyric material<br />

(see Fig. 6e and h) and most <strong>of</strong> <strong>the</strong>se deformed crystals<br />

are enclosed by granophyric intergrowths.<br />

The compositional progression from dolerite to <strong>the</strong> most<br />

felsic segregations, toge<strong>the</strong>r with <strong>the</strong> observed range <strong>of</strong> textural<br />

relationships between early (plagioclase and pyroxene)<br />

and late (granophyre) phases and <strong>the</strong> increasing<br />

amounts <strong>of</strong> Ca-amphibole and Fe^Ti oxides (magnetite<br />

and ilmenite) as crystallizing phases suggest that <strong>the</strong> segregations<br />

formed at different stages <strong>of</strong> crystallization <strong>of</strong> <strong>the</strong><br />

dolerite magma.<br />

Table 6: Anorthite content <strong>of</strong> plagioclase crystals from<br />

Pr<strong>of</strong>ile A-1<br />

Sample Group n Core Rim Average<br />

93-BS-84 S1 4 50–84 40–77 64<br />

93-BS-80 S4 5 54–63 38–60 53<br />

93-BS-81 S3 3 60–70 50–59 61<br />

93-BS-82 S4 5 33–63 37–61 55<br />

Detailed analyses are given in Supplementary Data Tables<br />

6–1 to 6–4.<br />

Table 7: Representative compositions <strong>of</strong> pyroxene crystals from Pr<strong>of</strong>ile A-1<br />

MINERAL CHEMISTRY<br />

Chemical compositions <strong>of</strong> plagioclase and pyroxene were<br />

determined within Pr<strong>of</strong>ile A-1, which consists <strong>of</strong> three consecutive<br />

segregations (93-BS-80, 93-BS-81 and 93-BS-82)<br />

situated below dolerite (93-BS-84). Plagioclase compositions<br />

are shown schematically in Fig. 7. <strong>Segregations</strong><br />

93-BS-80 and 82 are coarse-grained and have a patchy texture,<br />

whereas 93-BS-81, which is situated between <strong>the</strong>m, is<br />

medium- to coarse-grained and less patchy. The compositions<br />

<strong>of</strong> <strong>the</strong> minerals change systematically with bulk-rock<br />

composition and texture. The two most felsic and differentiated<br />

segregations, 93-BS-80 and 82, are similar in composition<br />

and texture and belong to <strong>the</strong> S4 group; <strong>the</strong> less<br />

felsic and differentiated segregation, 93-BS-81, belongs to<br />

<strong>the</strong> S 3 group, and <strong>the</strong> dolerite, 93-BS-84, which has <strong>the</strong><br />

most primitive composition and an ophitic texture, belongs<br />

to <strong>the</strong> S 1 group. On average five plagioclase and pyroxene<br />

crystals per thin section in segregation pr<strong>of</strong>ile A-1 were<br />

analyzed to determine compositional variations within a<br />

given sample and across samples (Table 6 and Table 7,<br />

respectively).<br />

Plagioclase<br />

Plagioclase crystals in dolerite sample 93-BS-84 are small<br />

(0·01^2 mm), euhedral laths commonly embedded in<br />

larger pyroxene oikocrysts (Fig. 7a). Most crystals are normally<br />

zoned, although <strong>the</strong>re are a few, 0·5^1mm, subhedral,<br />

tabular grains that show complex zoning. The most<br />

felsic segregations, 93-BS-80 and 82, have 1^20 mm long,<br />

euhedral to subhedral tabular and blocky crystals that are<br />

commonly fractured and altered, and have resorbed edges<br />

and cores partially replaced with granophyric intergrowths<br />

(Fig. 7b and d). Segregation 93-BS-81, which is<br />

less felsic, has 1^8 mm long, euhedral to subhedral, tabular<br />

and bladed crystals that are less fractured and altered<br />

than those <strong>of</strong> <strong>the</strong> o<strong>the</strong>r two segregations (Fig. 7c).<br />

Table 6 summarizes anorthite contents <strong>of</strong> plagioclase<br />

crystals (cores, rims and averages) from <strong>the</strong> four samples<br />

<strong>of</strong> Pr<strong>of</strong>ile A-1. Details <strong>of</strong> <strong>the</strong> electron microprobe analyses<br />

Sample Group n Cores Rims Average<br />

93-BS-84 S 1 3 En(46–68)Fs(13–21)Wo(11–36) En(34–44)Fs(23–28)Wo(30–42) En 46Fs 21Wo 34<br />

93-BS-80 S 4 5 En(32–37)Fs(20–26)Wo(40–43) En(25–32)Fs(24–51)Wo(25–45) En 32Fs 27Wo 41<br />

93-BS-81 S 3 3 En(33–41)Fs(18–27)Wo(39–42) En(26–35)Fs(24–31)Wo(40–43) En 34Fs 25Wo 42<br />

93-BS-82 S4 3 En(28–41)Fs(16–30)Wo(41–43) En(26–38)Fs(19–31)Wo(41–43) En32Fs25Wo42<br />

Detailed analyses are given in Supplementary Data Tables 7-1 to 7-4. n, number <strong>of</strong> crystals analyzed in each sample.<br />

Average is <strong>of</strong> (core and rim) analyses. One crystal in group S1 has a core composition <strong>of</strong> Wo11 and two crystals have a<br />

core composition <strong>of</strong> Wo35 and Wo36.<br />

1942<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/<br />

by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 7. Photomicrographs <strong>of</strong> segregation pr<strong>of</strong>ile A-1, which consists <strong>of</strong> (a) dolerite (93-BS-84) overlying three consecutive segregations:<br />

(b) 93-BS-80; (c) 93-BS-81; (d) 93-BS-82. The white circles in <strong>the</strong> photomicrographs indicate micro-probed plagioclase crystals. The compositions<br />

<strong>of</strong> <strong>the</strong>ir cores, rims and intermediate zones are shown in <strong>the</strong> feldspar ternary diagram below (e). Panels are arranged according to <strong>the</strong><br />

order in which samples occur in <strong>the</strong> field, from top to bottom.<br />

1943<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 8. Compositional pr<strong>of</strong>iles <strong>of</strong> plagioclase crystals in segregation pr<strong>of</strong>ile A-1. Left panels shows <strong>the</strong> variation <strong>of</strong> anorthite content (An wt %)<br />

in crystals taken from one edge <strong>of</strong> <strong>the</strong> crystal to <strong>the</strong> o<strong>the</strong>r, except those labeled ‘C’ where <strong>the</strong> pr<strong>of</strong>ile terminates in <strong>the</strong> core <strong>of</strong> <strong>the</strong> crystal.<br />

Right panels show representative electron backscatter images indicating where <strong>the</strong> pr<strong>of</strong>iles were taken in <strong>the</strong> crystals. (a) <strong>Dolerite</strong> 93-BS-84<br />

shows irregular concave-down pr<strong>of</strong>iles; (b) segregation 93-BS-80 has smooth concave pr<strong>of</strong>iles; (c) segregation 93-BS-81 (least felsic) has little<br />

compositional variation from rim to core resulting in flat pr<strong>of</strong>iles; (d) segregation 93-BS-82 (most altered) has irregular slightly concave pr<strong>of</strong>iles.<br />

The main phases are plagioclase (plg), pyroxene (pyx), granophyre (gr), and opaque minerals (opq). Panels are arranged according to<br />

stratigraphic sequence in <strong>the</strong> field from top to bottom (see Fig. 7).<br />

are given in Electronic Appendix Tables 6.1^6.4 (all <strong>of</strong> <strong>the</strong><br />

supplementary material is available for downloading at<br />

http://www.petrology.oxfordjournals.org). In accord with<br />

trends expected if <strong>the</strong> segregations are related to <strong>the</strong> dolerites<br />

by crystal^liquid differentiation, <strong>the</strong> dolerite 93-BS-84<br />

has <strong>the</strong> most Ca-rich cores but Na-rich rims similar to <strong>the</strong><br />

most felsic segregations (93-BS-80 and 82), whereas <strong>the</strong><br />

1944<br />

most mafic <strong>of</strong> <strong>the</strong> three segregation (93-BS-81) has more<br />

Ca-rich cores and less Na-rich rims than <strong>the</strong> o<strong>the</strong>r two<br />

felsic segregations (Fig. 7e). Plagioclase varies in size and<br />

composition with changes in <strong>the</strong> bulk composition <strong>of</strong> <strong>the</strong><br />

host-rocks.<br />

Compositional pr<strong>of</strong>iles through plagioclase grains are<br />

shown in Fig. 8. Plagioclase in <strong>the</strong> dolerite (93-BS-84,<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

group S 1) has cores with <strong>the</strong> highest anorthite concentrations<br />

and albite-rich rim compositions. The compositional<br />

pr<strong>of</strong>iles in this rock are irregular, although exhibiting<br />

an overall concave-downward shape from rim to core,<br />

and <strong>the</strong>y display significant compositional variations<br />

both between crystals and from one edge <strong>of</strong> a crystal to <strong>the</strong><br />

o<strong>the</strong>r (Fig. 8a). One <strong>of</strong> <strong>the</strong> smallest (0·15 mm) crystals analysed<br />

(plg-01) in <strong>the</strong> dolerite has a relatively uniform and<br />

low-anorthite core region and <strong>the</strong> smallest rim to core compositional<br />

variation (An40^62) when compared with <strong>the</strong><br />

o<strong>the</strong>r crystals in this dolerite sample (Fig. 8a).<br />

Plagioclase in <strong>the</strong> S 4 segregations (93-BS-80 and 82)<br />

displays generally smooth, concave-down compositional<br />

pr<strong>of</strong>iles; however, <strong>the</strong>re are compositional variations<br />

between crystals in <strong>the</strong>se samples (Fig. 8b and d). In<br />

sample 93-BS-82 (Fig. 8d), plagioclase crystals 82-02<br />

and 82-04 exhibit irregular, jagged compositional pr<strong>of</strong>iles<br />

with a localized decrease in An (wt %) in <strong>the</strong> intermediate<br />

region between cores and rim, suggesting<br />

that <strong>the</strong>se crystals may have partially resorbed intermediate<br />

regions. The downward dip in <strong>the</strong> compositional pr<strong>of</strong>ile<br />

plg-82-01 corresponds to a fracture going across <strong>the</strong> crystal.<br />

Plagioclase in <strong>the</strong> S 3 segregation (93-BS-81) has <strong>the</strong><br />

smallest compositional variation both between and within<br />

analyzed crystals, resulting in almost horizontal pr<strong>of</strong>iles<br />

(Fig. 8c). The observation that <strong>the</strong> compositional pr<strong>of</strong>iles<br />

for <strong>the</strong> small (51mm) and <strong>the</strong> larger (41·5 mm) crystals<br />

are almost identical suggests that <strong>the</strong> larger crystals may<br />

have completely recrystallized when <strong>the</strong>y came in contact<br />

with <strong>the</strong> residual liquid, or <strong>the</strong> larger crystals may have<br />

grown in <strong>the</strong> same environment as <strong>the</strong> smaller ones.<br />

Pyroxenes<br />

Pyroxene crystals in segregation pr<strong>of</strong>ile A-1 vary in size,<br />

shape and composition as well as modal abundance. In<br />

<strong>the</strong> dolerite sample 93-BS-84, <strong>the</strong> pyroxene crystals consist<br />

<strong>of</strong> both augite and pigeonite, whereas in <strong>the</strong> segregations<br />

93-BS-80, 81 and 82 <strong>the</strong> primary pyroxenes are augite in<br />

composition with pigeonite exsolution lamellae parallel to<br />

(001) and less commonly (100). Augite crystals are commonly<br />

twinned. In general, <strong>the</strong> pyroxene crystals show<br />

a regular zoning from more Mg-rich cores to more<br />

Fe-rich rims. Detailed mineral compositions <strong>of</strong> cores<br />

and rims are given in Electronic Appendix Tables 7.1^7.4.<br />

Representative compositions are given in Table 7; <strong>the</strong>se<br />

compositions are plotted in Fig. 9 as not all pyroxene analyses<br />

yielded good totals.<br />

All analysed pyroxene crystals in <strong>the</strong> dolerite, 93-BS-84,<br />

have distinct core and rim compositions, and <strong>the</strong>y are<br />

characterized by more enstatite-rich and less Ca-rich compositions<br />

than <strong>the</strong> segregations (Table 7, Fig. 9a^d). One<br />

out <strong>of</strong> three augite crystals analyzed (84-03) has a pigeonite<br />

core partially replaced by orthopyroxene lamellae as<br />

indicated by <strong>the</strong> anomalously En-rich data points in<br />

Fig. 9a.<br />

1945<br />

Pyroxenes in <strong>the</strong> more felsic, S 4 segregation samples<br />

show considerable compositional overlap between <strong>the</strong><br />

cores and rims. Crystals in sample 93-BS-82 contain <strong>the</strong><br />

most Fe-rich cores whereas those in sample 93-BS-80 contain<br />

<strong>the</strong> most Fe-rich and Ca-rich rims. Pyroxenes in <strong>the</strong><br />

less felsic S 3 segregation sample (93-BS-81) overall have<br />

compositions that are intermediate between <strong>the</strong> S 1 and S 4<br />

samples, but much closer to <strong>the</strong> S 4 samples.<br />

Representative enstatite (En wt %) compositional pr<strong>of</strong>iles<br />

illustrating some <strong>of</strong> <strong>the</strong>se features are shown in<br />

Fig. 9e. The pr<strong>of</strong>ile corresponding to <strong>the</strong> dolerite (84-01)<br />

has <strong>the</strong> highest enstatite concentrations and shows <strong>the</strong><br />

most rim to core variation compared with pr<strong>of</strong>iles from<br />

<strong>the</strong> segregations samples. All four pr<strong>of</strong>iles display asymmetric<br />

concave shapes with higher enstatite concentrations<br />

within <strong>the</strong> core <strong>of</strong> <strong>the</strong> crystal. Pr<strong>of</strong>ile 81-03, from a crystal<br />

in <strong>the</strong> S 3 segregation, displays a slightly higher peak enstatite<br />

concentration than <strong>the</strong> pr<strong>of</strong>iles from <strong>the</strong> S4 segregations.<br />

Pr<strong>of</strong>ile 82-04A (diamonds in Fig. 9e) shows little<br />

variation in composition from <strong>the</strong> center towards <strong>the</strong> rim<br />

on <strong>the</strong> right side: this could be because exsolution lamellae<br />

within that region <strong>of</strong> <strong>the</strong> crystal prevented accurate sampling<br />

<strong>of</strong> <strong>the</strong> average composition.<br />

Granophyre<br />

There is a systematic increase in <strong>the</strong> proportions <strong>of</strong> quartz<br />

and K-feldspar in <strong>the</strong> granophyric intergrowths with<br />

increasing silica content in <strong>the</strong> segregations that is particularly<br />

noticeable in <strong>the</strong> most differentiated samples represented<br />

by <strong>the</strong> group S 5. In <strong>the</strong> segregation pr<strong>of</strong>ile A-1, <strong>the</strong><br />

S 4 segregations 93-BS-80 and 82 have <strong>the</strong> highest amount<br />

<strong>of</strong> granophyric intergrowths and are also <strong>the</strong> most felsic<br />

and differentiated samples.<br />

In general, as <strong>the</strong> granophyric texture becomes more<br />

pervasive, plagioclase crystals with partially replaced<br />

cores and margins that are optically continuous with <strong>the</strong><br />

granophyric intergrowths become more common. In <strong>the</strong><br />

S 5 group plagioclase crystals are generally replaced by<br />

granophyric intergrowths and only a few large ( 1cm) euhedral<br />

isolated plagioclase crystals appear to be in equilibrium<br />

with <strong>the</strong> granophyre.<br />

SEM electron backscatter images were used to study<br />

granophyric intergrowths <strong>of</strong> alkali feldspar and quartz in<br />

two dolerite samples (93-BS-84 and 93-BS-96-B-1) and<br />

one segregation (94-BS-154-A) sample. The data are given<br />

in Electronic Appendix Tables 6.5 and 6.6, respectively. In<br />

<strong>the</strong> dolerite and segregation samples <strong>the</strong> granophyric intergrowths<br />

are characterized by light grey areas corresponding<br />

to K-feldspar and dark grey areas corresponding to<br />

quartz. The K-feldspar commonly surrounds euhedral to<br />

subhedral plagioclase crystals in small interstitial pockets,<br />

whereas in <strong>the</strong> segregation sample 94-BS-154-A it forms<br />

larger interstitial pockets. In all samples <strong>the</strong> composition<br />

<strong>of</strong> <strong>the</strong> K-feldspar ranges between 84 and 89 wt % orthoclase<br />

(Or) with an average <strong>of</strong> 85 wt % Or. In <strong>the</strong> dolerite<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 9. Clinopyroxene compositions in dolerite 93-BS-84 (a) and segregations (b) 93-BS-80, (c) 93-BS-81, and (d) 93-BS-82. (e) Representative<br />

enstatite (En wt %) compositional pr<strong>of</strong>iles <strong>of</strong> <strong>the</strong>se crystals. Right panels show <strong>the</strong> electron backscatter images <strong>of</strong> where <strong>the</strong> compositional pr<strong>of</strong>iles<br />

shown in (e) were made. Panels are arranged according to stratigraphic sequence in <strong>the</strong> field from top to bottom.<br />

samples, <strong>the</strong> quartz forms small intergrowths but in <strong>the</strong><br />

segregations <strong>the</strong> quartz intergrowths are larger and occasionally<br />

form isolated crystals. The quartz analyses yielded<br />

compositions close to 100% SiO 2 within about 0·5% measurement<br />

uncertainty.<br />

1946<br />

Summary<br />

Plagioclase crystals in segregation pr<strong>of</strong>ile A-1 show regular<br />

compositional variations between <strong>the</strong> dolerite and <strong>the</strong><br />

three adjacent segregations. Plagioclase in <strong>the</strong> dolerite has<br />

<strong>the</strong> largest compositional variation, both between crystals<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

and between core and rim, indicative <strong>of</strong> continued<br />

fractionation.<br />

The segregations all have progressively more albite-rich<br />

core and rim compositions (Table 6) as would be expected<br />

if <strong>the</strong>se rocks formed as fractionated liquids from <strong>the</strong> dolerite.<br />

In addition, <strong>the</strong> segregations have systematically<br />

larger crystal sizes and <strong>the</strong>y display flatter core^rim pr<strong>of</strong>iles<br />

than plagioclase in <strong>the</strong> host dolerite, possibly reflecting<br />

faster diffusion rates enhanced by a more volatile-rich environment<br />

that promoted growth and re-equilibration <strong>of</strong><br />

<strong>the</strong> crystals. Plagioclase in <strong>the</strong> segregations has an average<br />

composition between An 53 and An 61, which is within <strong>the</strong><br />

range <strong>of</strong> rim to core compositional variation (An41^62) <strong>of</strong><br />

crystal 84-01, which is one <strong>of</strong> <strong>the</strong> smaller crystals from <strong>the</strong><br />

dolerite sample 93-BS-84. This is consistent with <strong>the</strong> segregations<br />

forming from <strong>the</strong> fractionated liquid <strong>of</strong> a dolerite<br />

parent.<br />

The two most felsic segregations, which belong to group<br />

S 4, have comparable plagioclase compositions, but <strong>the</strong><br />

more altered segregation 93-BS-82 has a wider compositional<br />

range and more irregular plagioclase compositional<br />

pr<strong>of</strong>iles. Plagioclases within segregation 93-BS-81 (group<br />

S 3) have smaller ranges <strong>of</strong> compositional variation characterized<br />

by slightly more anorthite-rich cores and less<br />

albite-rich rims (Table 6) than <strong>the</strong> two S 4 segregations, as<br />

would be expected for this less differentiated segregation.<br />

This sample shows very little compositional variation between<br />

cores and rims within analyzed crystals.<br />

The pyroxenes in <strong>the</strong> segregations have more Fe-rich<br />

cores and rims, and higher Ca contents than <strong>the</strong> pyroxenes<br />

in dolerite 93-BS-84, which is consistent with <strong>the</strong> more<br />

evolved composition <strong>of</strong> <strong>the</strong> segregations. <strong>Segregations</strong><br />

93-BS-80 and 93-82, which have similar textures and<br />

major element compositions, have overlapping core and<br />

rim compositions with overall more Fe-rich cores and<br />

more Fe-rich rims than <strong>the</strong> less differentiated segregation<br />

93-BS-81 and <strong>the</strong> adjacent dolerite 93-BS-84 (Table 7).<br />

MAJOR-ELEMENT<br />

GEOCHEMISTRY<br />

In <strong>the</strong> Quartz^Alkali Feldspar^Plagioclase (QAP) ternary<br />

diagram (Streckeisen, 1973), <strong>the</strong> dolerites and felsic segregations<br />

define a trend from quartz diorite to granodiorite<br />

passing through <strong>the</strong> quartz monzodiorite field (Fig. 10a).<br />

The S 1 group rocks (black diamonds) have quartz diorite<br />

and quartz monzodiorite compositions, except for one<br />

sample, 94-BS-160-A-2-2, which plots in <strong>the</strong> granodiorite<br />

field. This sample is part <strong>of</strong> a large composite segregation<br />

that contains alternating felsic and mafic sub-lenses. The<br />

S2 segregations (grey filled squares) have granodiorite<br />

compositions except for one less differentiated sample<br />

94-BS-154-A, which corresponds to <strong>the</strong> less felsic component<br />

<strong>of</strong> segregation 94-BS-154. This sample is unusual in<br />

1947<br />

o<strong>the</strong>r ways and will be discussed fur<strong>the</strong>r below. The S 3, S 4<br />

and S 5 segregations plot in <strong>the</strong> granodiorite field with <strong>the</strong><br />

quartz component generally increasing with group<br />

number. Only two samples (93-BS-141 and 94-BS-212)<br />

from <strong>the</strong> S 5 group deviate from <strong>the</strong> main trend: <strong>the</strong>se samples<br />

have higher K 2O concentrations and are more altered<br />

than any o<strong>the</strong>r segregation.<br />

In an Na 2O þK 2O^FeO T^MgO (AFM) diagram<br />

(Fig. 10b), <strong>the</strong> S 1 and <strong>the</strong> S 2 groups plot closer to <strong>the</strong><br />

Mg-rich corner, and both display a wide range <strong>of</strong> FeO T;<br />

<strong>the</strong> S 2 group, which contains a higher modal abundance<br />

<strong>of</strong> orthopyroxene, plots closest to <strong>the</strong> FeO T^MgO sideline.<br />

The S3 and <strong>the</strong> S4 segregations show a systematic enrichment<br />

in FeO T, except that <strong>the</strong> majority <strong>of</strong> <strong>the</strong> S 3 samples<br />

have consistently lower FeO T concentrations and form a<br />

tighter cluster than <strong>the</strong> majority <strong>of</strong> <strong>the</strong> S 4 samples. The S 5<br />

segregations exhibit two arrays <strong>of</strong> data; one that is slightly<br />

higher in FeOTand one that is lower in FeOT with continued<br />

fractionation. Although <strong>the</strong>se rocks are continuously<br />

enriched in <strong>the</strong> alkalis, <strong>the</strong> varying FeO T trends in<br />

Fig. 10b could be due to <strong>the</strong> different modal amounts <strong>of</strong><br />

pyroxene and <strong>the</strong> varying amounts <strong>of</strong> alteration to Fe^Ti<br />

oxide and biotite.<br />

The S 1^S 5 segregation suite shows a wide bulk-rock compositional<br />

range <strong>of</strong> 51^67% SiO 2, 1^13 wt % MgO,<br />

7^15 wt % FeO T,8^20wt%Al 2O 3, 4^13 wt % CaO,<br />

0·04^0·35 wt % P 2O 5, 0·4^2·3wt % TiO 2, 1^4 wt %<br />

Na2O, and 0·4^2·5wt % K2O. The Mg-number [MgO/<br />

(FeO T þ MgO)] ranges from 16 to 70. The alkalis<br />

(Na 2O þK 2O) range from 1·2 to5·5wt % with <strong>the</strong> S 5<br />

group having <strong>the</strong> lowest Mg-number and <strong>the</strong> highest alkalis,<br />

and <strong>the</strong> S 1 group having <strong>the</strong> highest Mg-number<br />

and <strong>the</strong> second lowest alkalis. Bulk-rock compositions <strong>of</strong><br />

groups S 2, S 3 and S 5 are given in Table 2; only representative<br />

bulk-rock compositions <strong>of</strong> groups S 1 and S 4 are reported.<br />

The complete dataset for Table 2 is available as an<br />

Electronic Appendix. The bulk-rock compositions are<br />

plotted on Harker variation diagrams and shown in<br />

Fig. 11. In this figure, two <strong>of</strong> <strong>the</strong> three samples from group<br />

S 2 behave like highly differentiated members <strong>of</strong> group S 1:<br />

<strong>the</strong> outlying samples is 94-BS-154-A, which we will argue<br />

is not part <strong>of</strong> <strong>the</strong> differentiation sequence.<br />

There is a systematic increase in <strong>the</strong> less compatible<br />

elements (P 2O 5,K 2O and Na 2O) with increasing SiO 2<br />

across <strong>the</strong> sample suite, and a less linear but regular<br />

decrease in <strong>the</strong> compatible elements (MgO, CaO, Al 2O 3).<br />

The FeO T and TiO 2 concentrations exhibit more<br />

scatter. Overall, as SiO2 increases FeOT concentrations<br />

show an increase for group S 1 and a decrease for groups<br />

S 3^S 5.TiO 2 concentrations increase with SiO 2 in groups<br />

S 1^S 3, show a large scatter in group S 4, and decrease in<br />

group S 5.<br />

Figures 10 and 11 show that <strong>the</strong>re is a compositional continuum<br />

between <strong>the</strong> host dolerites and <strong>the</strong> most felsic<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 10. (a) Normative mineral compositions plotted in a Quartz^Alkali Feldspar^Plagioclase diagram <strong>of</strong> <strong>the</strong> dolerites described by group<br />

S1 (diamonds, black ellipse), and segregations described by groups S2 (grey squares, dashed grey ellipse), S3 ( , continuous-line grey ellipse),<br />

S4 (triangles, black ellipse) and S5 (circles, black ellipse). Numbered regions correspond to IUGS classification rock types: 3, granite; 4, granodiorite;<br />

5, tonalite; 9*, quartz monzodiorite^quartz monzogabbro;10*, quartz diorite; 9, monzodiorite^monzogabbro;10, diorite^gabbro^anorthosite.<br />

(b) AFM diagram for <strong>the</strong> same segregation^dolerite suite showing a tholeiitic trend characterized by iron enrichment followed by alkali<br />

enrichment with progressive differentiation, in particular in groups S 4 and S 5.<br />

segregations and this differentiation trend can be best<br />

observed in <strong>the</strong> systematic decrease <strong>of</strong> CaO with increasing<br />

silica in groups S 1 to S 5. The Harker diagram trends<br />

can be explained as <strong>the</strong> result <strong>of</strong> simultaneous fractional<br />

1948<br />

crystallization <strong>of</strong> pyroxene and plagioclase from <strong>the</strong> most<br />

mafic (dolerite) magma. The decrease in FeO T and TiO 2<br />

at higher SiO 2 is explained by <strong>the</strong> late crystallization <strong>of</strong><br />

Fe^Ti oxides.<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 11. Major element compositions <strong>of</strong> <strong>the</strong> segregation^dolerite suite vs wt % SiO 2. Trends form a continuum between <strong>the</strong> most primitive composition<br />

corresponding to <strong>the</strong> dolerite (S 1) and <strong>the</strong> most differentiated compositions correspond to <strong>the</strong> most felsic segregations (S 5).<br />

MELTS MODELING OF THE<br />

SILICIC SEGREGATIONS<br />

The whole-rock and mineral chemistry data suggest that<br />

<strong>the</strong> silicic segregations are compositionally related to <strong>the</strong><br />

host dolerite. Their compositions are broadly compatible<br />

with a recent model for <strong>the</strong> formation <strong>of</strong> felsic segregations<br />

1949<br />

during cooling <strong>of</strong> thick sills (Marsh, 2002), in which a crystal<br />

mush growing downward from <strong>the</strong> top contact <strong>of</strong> <strong>the</strong><br />

sill becomes gravitationally unstable when it is 50^70%<br />

crystalline and tears. These tears are <strong>the</strong>n filled with residual<br />

liquid filter pressed from directly below. According<br />

to this model, <strong>the</strong> segregations in <strong>the</strong> BS represent residual<br />

liquids after 50^70% crystallization <strong>of</strong> <strong>the</strong> host dolerite.<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

As such, <strong>the</strong>ir bulk compositions should follow <strong>the</strong> liquid<br />

lines <strong>of</strong> descent <strong>of</strong> <strong>the</strong>ir parent magma composition.<br />

In this study <strong>the</strong> program MELTS (Ghiorso & Sack,<br />

1995; Asimow & Ghiorso, 1998) was used to model <strong>the</strong><br />

crystallization <strong>of</strong> <strong>the</strong> segregations from a parent dolerite<br />

liquid. After determining <strong>the</strong> best parent melt composition,<br />

we model <strong>the</strong> formation <strong>of</strong> <strong>the</strong> segregations using different<br />

crystallization models to obtain <strong>the</strong> best fit with <strong>the</strong><br />

compositions <strong>of</strong> <strong>the</strong> segregations.<br />

MELTS methods<br />

The MELTS program combines <strong>the</strong>rmodynamic principles<br />

and optimization procedures to compute stable heterogeneous<br />

equilibria between solids and liquids at<br />

specified temperatures and pressures in magmatic systems.<br />

Given a bulk composition, equilibrium states can be calculated<br />

at chosen temperature or pressure intervals during<br />

crystallization, melting, assimilation or mixing processes.<br />

The proportions and compositions <strong>of</strong> all phases are given<br />

at each interval, and <strong>the</strong> program allows forward modeling<br />

<strong>of</strong> <strong>the</strong> system under equilibrium crystallization (EC)<br />

or fractional crystallization (FC) conditions. MELTS has<br />

been successfully used to model magmatic processes in<br />

peridotite, basalt and andesite systems (Ghiorso &<br />

Carmichael, 1985; Hirschmann et al., 1999; Asimow et al.,<br />

2001; Eason & Sinton, 2006). However, it has been shown<br />

that <strong>the</strong> MELTS model tends to become unstable at high<br />

degrees <strong>of</strong> differentiation <strong>of</strong> a basaltic parent magma<br />

(Ghiorso & Sack, 1995); consequently, in this study all<br />

models were run until a lower temperature <strong>of</strong> 9008C, or<br />

80% crystallization in FC and 90% in EC was achieved.<br />

Determining <strong>the</strong> crystallization conditions<br />

in <strong>the</strong> Basement Sill<br />

To model <strong>the</strong> evolution <strong>of</strong> <strong>the</strong> BS, an initial liquid composition<br />

is assumed and <strong>the</strong> temperature is decreased at regular<br />

intervals between <strong>the</strong> liquidus and <strong>the</strong> solidus at a<br />

constant pressure. The change in liquid composition is<br />

defined as <strong>the</strong> liquid line <strong>of</strong> descent (LLD), analogous to a<br />

differentiation trend. If segregations in <strong>the</strong> FDS are<br />

simple fractionated liquids from a parent basaltic magma,<br />

<strong>the</strong>n <strong>the</strong>y should lie on <strong>the</strong> LLD.<br />

Based on <strong>the</strong> mineralogy and geological setting <strong>of</strong> <strong>Ferrar</strong><br />

tholeiites from <strong>the</strong> sou<strong>the</strong>rn Prince Albert Mountains in<br />

south Victoria Land, Demarchi et al. (2001) inferred that<br />

crystallization occurred at a pressure between 100 and<br />

500 MPa in a system that was closed to oxygen exchange.<br />

This estimate for <strong>the</strong> depth <strong>of</strong> emplacement may be<br />

extended to <strong>the</strong> BS on <strong>the</strong> basis <strong>of</strong> <strong>the</strong> work <strong>of</strong> Hersum<br />

et al. (2007), who estimated a paleodepth <strong>of</strong> 3^4 km<br />

based on <strong>the</strong> thickness <strong>of</strong> <strong>the</strong> capping comagmatic<br />

Kirkpatrick flood basalts. For systems that are closed to<br />

oxygen exchange, MELTS calculates <strong>the</strong> oxygen fugacity<br />

(fO2) by defining <strong>the</strong> ferrous to ferric ratio according to<br />

<strong>the</strong> bulk composition at each temperature for a constant<br />

1950<br />

pressure, whereas for systems that are an open to oxygen<br />

exchange (which we also considered) MELTS requires<br />

that <strong>the</strong> f O2 <strong>of</strong> <strong>the</strong> system be specified along a known<br />

oxygen buffer, which for magmas <strong>of</strong> this composition is<br />

taken to be <strong>the</strong> FMQ (fayalite^magnetite^quartz) buffer.<br />

MELTS calculations were also done with a more oxidizing<br />

buffer (nickel^nickel oxide) but as <strong>the</strong>se results did not<br />

reveal any significant difference, <strong>the</strong>y are not shown.<br />

The mineralogy <strong>of</strong> <strong>the</strong> dolerite is predominantly anhydrous,<br />

arguing for a relatively dry magma. However, <strong>the</strong><br />

segregations are characterized by higher concentrations <strong>of</strong><br />

hydrous phases and hydrous alteration <strong>of</strong> previously crystallized<br />

phases, and are coarse-grained, all indicative <strong>of</strong><br />

higher H 2O concentrations in <strong>the</strong> residual liquids.<br />

Moreover, <strong>the</strong> quartz and K-feldspar granophyric intergrowths<br />

characteristic <strong>of</strong> <strong>the</strong> segregations have been interpreted<br />

to develop as a result <strong>of</strong> sudden H 2O loss (e.g.<br />

Winter, 2001). We elected to run <strong>the</strong> MELTS models for anhydrous<br />

conditions (0 wt % H 2O in <strong>the</strong> initial liquid composition)<br />

and hydrous conditions (0·5 wt % <strong>of</strong> H 2O).<br />

Models for <strong>the</strong> solidification <strong>of</strong> thick sills (Philpotts et al.,<br />

1996; Marsh, 2002) suggest that over a large part <strong>of</strong> <strong>the</strong><br />

crystallization history <strong>the</strong>re is little relative motion between<br />

crystals, which are connected toge<strong>the</strong>r in permeable<br />

networks, and <strong>the</strong> surrounding interstitial liquid. This<br />

might suggest that equilibrium crystallization (EC) would<br />

be <strong>the</strong> appropriate crystallization mode for <strong>the</strong> BS.<br />

However, conductive cooling models indicate that <strong>the</strong><br />

upper levels <strong>of</strong> <strong>the</strong> sill, where <strong>the</strong> segregations occur,<br />

would have solidified relatively quickly (within hundreds<br />

<strong>of</strong> years) so <strong>the</strong>re would not have been time for <strong>the</strong> centers<br />

<strong>of</strong> early formed crystals to equilibrate with <strong>the</strong> surrounding<br />

liquid by diffusion; thus we infer that <strong>the</strong> earlier-formed<br />

crystals were chemically removed from <strong>the</strong> system, as in<br />

<strong>the</strong> fractional crystallization (FC) mode. As a result, we<br />

chose to run <strong>the</strong> MELTS models for both EC and FC<br />

conditions.<br />

Determining <strong>the</strong> best parent magma<br />

compositions for <strong>the</strong> segregations<br />

The BS (excluding <strong>the</strong> orthopyroxene cumulate tongue) is<br />

a relatively uniform dolerite and so representative dolerite<br />

liquids were sought as <strong>the</strong> parent magma composition. To<br />

determine <strong>the</strong> best parent composition for <strong>the</strong> segregations,<br />

two representative dolerite samples were chosen<br />

from each <strong>of</strong> three locations within <strong>the</strong> BS; however, as<br />

<strong>the</strong> results for <strong>the</strong> two samples from each location were<br />

very similar, only one <strong>of</strong> each pair is shown in Fig. 12.<br />

Their locations in <strong>the</strong> BS and <strong>the</strong>ir bulk compositions are<br />

given in Table 8. The three samples shown are from <strong>the</strong><br />

upper chilled margin <strong>of</strong> BS (96-A-26), a dolerite adjacent<br />

to a segregation lens (93-BS-90-B), and a medium-grained<br />

dolerite located from <strong>the</strong> centre <strong>of</strong> <strong>the</strong> sill above <strong>the</strong> cumulate<br />

layer (96-A-33). These samples are referred to in <strong>the</strong><br />

discussion below as <strong>the</strong> CM (‘chilled margin’),<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 12. Representative MgO and CaO element modeling using three initial dolerite compositions: SA, 93-BS-90-B (continuous black line);<br />

CM, 96-A-26 (long-dashed line); CS, 96-A-33 (short-dashed line), in EC (left panel) and FC (right panel) MELTS models that are closed to<br />

oxygen exchange, at 100 MPa, for anhydrous conditions (Dry) and hydrous (0·5wt%H 2O) (Wet) conditions.<br />

SA (‘segregation-adjacent’) and CS (‘central sill’) samples,<br />

respectively.<br />

Although <strong>the</strong>se candidates for <strong>the</strong> parental liquid have<br />

broadly similar compositions, <strong>the</strong>re are significant differences<br />

in <strong>the</strong>ir major element compositions (Table 8).<br />

Interestingly, <strong>the</strong> chilled margin sample has <strong>the</strong> most<br />

evolved composition, suggesting that it does not represent<br />

1951<br />

<strong>the</strong> parental liquid for <strong>the</strong> entire sill and that <strong>the</strong> sill was<br />

built up by more than one magma pulse.<br />

The three candidate parent liquid compositions were<br />

used as starting materials in <strong>the</strong> MELTS models with <strong>the</strong><br />

following permutations: (1) EC or FC, and (2) anhydrous<br />

or hydrous (0·5wt % H 2O present in <strong>the</strong> initial liquid).<br />

All models were run at a constant pressure <strong>of</strong> 100 MPa<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Table 8: <strong>Ferrar</strong> dolerites initial liquid compositions<br />

Type D (m) Sample Location SiO2 TiO2 Al2O3 FeOT Fe2O3 MnO MgO CaO Na2O K2O P2O5 Total<br />

CM 0 96-A-26 McT 54·83 0·61 14·88 9·90 1·65 0·17 7·00 10·95 1·70 0·83 0·08 101·12<br />

SA 52 93-BS90-B Seg B 51·89 0·38 13·81 10·57 1·76 0·19 9·33 11·77 1·33 0·42 0·04 100·37<br />

CS 30 96-A-33 MP 53·84 0·64 13·83 9·20 1·81 0·17 7·56 10·30 1·58 0·81 0·08 98·20<br />

and, in <strong>the</strong> first instance, for systems closed to oxygen<br />

exchange.<br />

MELTS results<br />

The results from <strong>the</strong>se MELT models are shown in Fig. 12<br />

and Electronic Appendix Fig. 12-S. Figure 12 shows LLDs<br />

for <strong>the</strong> major elements MgO and CaO for <strong>the</strong> three dolerite<br />

samples compared with <strong>the</strong> compositions <strong>of</strong> <strong>the</strong> segregations<br />

(black diamonds). Each symbol along <strong>the</strong> LLDs<br />

corresponds to a 108C decrease in temperature, with a corresponding<br />

increase in silica. The left and right panels in<br />

<strong>the</strong> figure correspond to EC and FC conditions, respectively.<br />

Supplementary Data Fig. 12-S shows LLDs for major<br />

elements FeO and Al 2O 3 for <strong>the</strong> three dolerite initial<br />

liquid compositions compared with <strong>the</strong> segregations.<br />

As expected for liquids with a dolerite composition, <strong>the</strong><br />

models indicate that with increasing SiO 2 and decreasing<br />

temperature <strong>the</strong> concentrations <strong>of</strong> MgO, CaO and Al 2O 3<br />

in <strong>the</strong> residual liquids should decrease systematically<br />

whereas <strong>the</strong> concentrations <strong>of</strong> FeO first increase and <strong>the</strong>n<br />

decrease when <strong>the</strong> iron oxide phases crystallize.<br />

It is noteworthy that <strong>the</strong> segregation compositions are<br />

scattered over a range at least as wide as that between <strong>the</strong><br />

LLDs, suggesting that <strong>the</strong>re was more than one parent<br />

liquid. It can also be seen that <strong>the</strong> LLDs for <strong>the</strong> CM<br />

(long-dashed line in Fig. 12) <strong>of</strong>ten fall between those <strong>of</strong><br />

<strong>the</strong> SA and CS, and that for <strong>the</strong> best-fit cases many <strong>of</strong> <strong>the</strong><br />

segregation compositions also fall between <strong>the</strong>se two dolerite<br />

LLDs. Therefore, only <strong>the</strong> SA and CS limiting compositions<br />

are shown in <strong>the</strong> following MELTS models.<br />

For <strong>the</strong> present models, closed to oxygen exchange, <strong>the</strong><br />

best-fit LLDs differ for different oxides. MgO and CaO<br />

concentrations are best matched by a wet EC model, FeO<br />

concentrations by FC models, and Al 2O 3 by a dry FC<br />

model.<br />

Refining <strong>the</strong> crystallization model for<br />

<strong>the</strong> segregations<br />

The roles <strong>of</strong> starting composition and oxygen diffusion are<br />

evaluated in Figs 13 and 14 and Electronic Appendix<br />

Figs 13-S and 14-S, which illustrate LLDs <strong>of</strong> selected<br />

oxides for <strong>the</strong> SA and CS parent liquid compositions, for<br />

systems closed (dark grey lines) or open (light grey lines)<br />

to oxygen exchange.<br />

1952<br />

These figures show that <strong>the</strong> degree <strong>of</strong> fit between LLDs<br />

and <strong>the</strong> segregation data is not sensitive to oxygen diffusion<br />

for any <strong>of</strong> <strong>the</strong> oxides, except perhaps for FeO when<br />

using <strong>the</strong> wet FC model. The starting composition is a<br />

more important variable. For <strong>the</strong> FC models, <strong>the</strong> FeO concentrations<br />

<strong>of</strong> <strong>the</strong> segregations are bracketed by <strong>the</strong><br />

LLDs for <strong>the</strong> SA and CS dolerite starting compositions,<br />

with <strong>the</strong> hydrous FC model providing <strong>the</strong> best fit<br />

(Fig. 13d). The inflection points in <strong>the</strong>se LLDs correspond<br />

to <strong>the</strong> crystallization <strong>of</strong> magnetite, which occurs at<br />

55 wt % SiO 2 and 60 wt % SiO 2 for <strong>the</strong> SA and CS<br />

samples, respectively.<br />

In <strong>the</strong> segregations <strong>the</strong> measured CaO concentrations<br />

decrease rapidly with increasing SiO 2 up to 59 wt %<br />

SiO 2 after which <strong>the</strong> decrease is less rapid, whereas in <strong>the</strong><br />

MELTS models this change occurs at 55 wt % and<br />

60 wt % SiO 2 in <strong>the</strong> FC models using SA and CS dolerite<br />

compositions, respectively (Fig. 13f and h). In <strong>the</strong> EC<br />

models <strong>the</strong>re is little change in <strong>the</strong> slope regardless <strong>of</strong> <strong>the</strong><br />

initial composition used except at <strong>the</strong> lowest SiO 2<br />

(Fig. 13e and g). In <strong>the</strong> models, <strong>the</strong> initial rapid decrease is<br />

due to <strong>the</strong> crystallization <strong>of</strong> Ca-rich pyroxene and plagioclase<br />

but with continued crystallization <strong>the</strong> plagioclase becomes<br />

progressively more albite rich and thus consumes<br />

more sodium than calcium. Supplementary Data Fig. 13-S<br />

shows LLDs for MgO and Al2O3. The models that best approximate<br />

<strong>the</strong> compositions <strong>of</strong> <strong>the</strong> segregations are for<br />

MgO <strong>the</strong> EC models (Supplementary Data Fig. 13-Sa and<br />

c), and for Al 2O 3 <strong>the</strong> dry FC model (Supplementary Data<br />

Fig. 13-Sf).<br />

Modeling <strong>of</strong> <strong>the</strong> less compatible elements, Na2O and<br />

K 2O, generally shows an increase in concentration with<br />

increasing silica and decreasing temperature, as expected<br />

(Fig. 14). However, <strong>the</strong> concentrations in <strong>the</strong> segregations<br />

differ significantly from those predicted by <strong>the</strong> MELTS<br />

LLDs. Above 59 wt % SiO 2,K 2O in <strong>the</strong> segregations remains<br />

approximately constant at 1·5 wt %, whereas <strong>the</strong><br />

K 2O concentration in <strong>the</strong> model liquids continues to increase<br />

with increasing SiO2. With respect to Na2O, <strong>the</strong><br />

concentrations in <strong>the</strong> model liquids rise to a maximum<br />

before declining at high wt % SiO2, whereas concentrations<br />

in <strong>the</strong> segregations increase monotonically with<br />

increasing SiO 2. The reason for <strong>the</strong> discrepancy is not<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Fig. 13. Representative FeO and CaO element modeling using SA, 93-BS-90-B (continuous line) and CS, 96-A-33 (dashed line), in a EC model<br />

(left panel) and FC model (right panel) for systems that are closed to oxygen exchange (dark grey (black)), and systems that are open to<br />

oxygen exchange (grey) (light grey), with no H 2O present in <strong>the</strong> liquid, and 0·5wt % H 2O present in <strong>the</strong> initial liquid composition. The<br />

black diamonds show segregation compositions.<br />

clear; however, evidence for small amounts <strong>of</strong> deuteric and<br />

hydro<strong>the</strong>rmal alteration in <strong>the</strong> segregations manifest by<br />

<strong>the</strong> presence <strong>of</strong> amphibole and biotite replacing pyroxene,<br />

minor secondary white mica replacing biotite, and sericite<br />

replacing plagioclase may be significant in this regard.<br />

Also, <strong>the</strong>re is some K and Na ion exchange along <strong>the</strong><br />

1953<br />

regions where <strong>the</strong> granophyric intergrowths surround subhedral<br />

or partially resorbed plagioclase crystals that<br />

might account for <strong>the</strong> difference in <strong>the</strong> K and Na trends<br />

in <strong>the</strong> segregations. Although none <strong>of</strong> <strong>the</strong> models provides<br />

a good fit to <strong>the</strong> alkali concentrations in <strong>the</strong> segregations,<br />

<strong>the</strong> wet FC model is closest (Fig. 14d and h).<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 14. Representative K 2O and Na 2O element modeling using SA, 93-BS-90-B (continuous line) and CS, 96-A-33 (dashed line), in a EC<br />

model (left panel) and FC model (right panel) for systems that are closed to oxygen exchange (black) (dark grey), and systems that are open<br />

to oxygen exchange (grey) (light grey), with no H2O present in <strong>the</strong> liquid, and 0·5wt % H2O present in <strong>the</strong> initial liquid composition. The<br />

black diamonds show segregation compositions.<br />

The P 2O 5 concentrations in <strong>the</strong> segregations are well<br />

approximated by <strong>the</strong> LLDs for this element, in particular<br />

<strong>the</strong> FC model (Electronic Appendix Fig. 14-Sb and d),<br />

whereas <strong>the</strong> TiO 2 concentrations show too wide a scatter<br />

to be properly modeled by ei<strong>the</strong>r EC or FC models<br />

1954<br />

(Electronic Appendix Fig. 14-Se^h), probably owing to <strong>the</strong><br />

coarse-grained and heterogeneous textures <strong>of</strong> <strong>the</strong><br />

segregations.<br />

In summary, <strong>the</strong> MELTS model results show that none<br />

<strong>of</strong> <strong>the</strong> models provides good fits for all oxides as <strong>the</strong><br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

segregation compositions do not follow <strong>the</strong> simple LLDs<br />

expected for fractionated residual liquids. However, overall,<br />

considering both compatible and less compatible elements<br />

(Figs 12^14 and 12-S-14-S), <strong>the</strong> wet FC closed-system<br />

model provides <strong>the</strong> best fit to <strong>the</strong> segregation compositions<br />

for most elements. For MgO and CaO EC models provide<br />

better fits but principally at high silica values for <strong>the</strong> most<br />

differentiated <strong>of</strong> <strong>the</strong> segregations (group S 5).<br />

The differences for single elements indicate that no<br />

single, simple crystallization model can adequately explain<br />

all <strong>the</strong> data for such a dynamically evolving system, but<br />

overall we conclude that <strong>the</strong> most representative model<br />

for <strong>the</strong> segregations is an FC model that starts from a<br />

range <strong>of</strong> different initial magma conditions (between SA<br />

and CS), which crystallize close to <strong>the</strong> FMQ buffer<br />

(closed to oxygen exchange) and have at least 0·5wt %<br />

H 2O present in <strong>the</strong> initial liquid composition.<br />

MELTS crystallizing phases<br />

The MELTS models predict <strong>the</strong> sequence and temperature<br />

<strong>of</strong> crystallization and <strong>the</strong> per cent crystallization (f) <strong>of</strong><br />

<strong>the</strong> solid phases from <strong>the</strong> liquid. This information is summarized<br />

in Fig. 15 for liquids SA and CS, using FC hydrous<br />

models closed to oxygen exchange. The top panel displays<br />

<strong>the</strong> silica content <strong>of</strong> <strong>the</strong> residual liquid, and <strong>the</strong> crystals<br />

that form from it, as a function <strong>of</strong> crystallinity, and <strong>the</strong><br />

bottom panel shows temperature as a function <strong>of</strong> crystallinity.<br />

The models were run from <strong>the</strong> liquidus to a<br />

near-solidus temperature <strong>of</strong> 9008C, beyond which <strong>the</strong><br />

MELTS models became unstable.<br />

Figure 15a shows that <strong>the</strong> cotectic mineral assemblage<br />

pigeonite (pig), augite (aug) and plagioclase (plg) form at<br />

18% and 16% crystallization (f) for <strong>the</strong> SA and <strong>the</strong> CS<br />

liquids, respectively. The first mineral to crystallize is<br />

Mg-rich olivine in <strong>the</strong> SA liquid and Mg-rich orthopyroxene<br />

in <strong>the</strong> CS dolerite liquid, which has a higher initial<br />

SiO 2 concentration. It is important to note that <strong>the</strong> first<br />

mineral to crystallize in both cases is shortly replaced by<br />

pigeonite or augite and subsequently joined by plagioclase.<br />

Magnetite begins to crystallize at 10778C corresponding<br />

to f ¼ 59% in <strong>the</strong> SA liquid, and 10918C corresponding<br />

to f ¼ 42% in <strong>the</strong> CS liquid. When <strong>the</strong> temperature<br />

decreases to 9978C (f ¼75%) in <strong>the</strong> SA liquid and 9218C<br />

(f ¼72%) in <strong>the</strong> CS liquid, an Fe-rich olivine phase replaces<br />

pigeonite and continues to crystallize along with<br />

augite, plagioclase and magnetite. The composition <strong>of</strong> <strong>the</strong><br />

augite becomes more Fe- and Ca-rich, <strong>the</strong> composition <strong>of</strong><br />

<strong>the</strong> plagioclase becomes more Na-rich and <strong>the</strong> composition<br />

<strong>of</strong> <strong>the</strong> magnetite becomes more Ti-rich with decreasing<br />

temperature. Only in <strong>the</strong> dry FC model using dolerite<br />

liquid CS does ilmenite (FeTiO 3) crystallize along with<br />

olivine, augite and plagioclase at a temperature <strong>of</strong> 9278C.<br />

These crystallization sequences were also calculated for<br />

different models. For <strong>the</strong> two starting dolerite compositions,<br />

both EC and FC models have similar crystallization<br />

1955<br />

sequences involving two pyroxenes (pigeonite and augite),<br />

plagioclase and magnetite, and at very low crystallization<br />

temperatures and under varying H 2O conditions, ilmenite,<br />

quartz and an H2O phase form, but at different temperatures<br />

and degrees <strong>of</strong> crystallization. The main differences<br />

between <strong>the</strong> EC and FC models are that <strong>the</strong> mineral compositions<br />

are more evolved in <strong>the</strong> FC model at each temperature<br />

step, as would be expected for an FC model.<br />

Also, well after (1008C or so cooling) magnetite crystallizes,<br />

an Fe-rich orthopyroxene in <strong>the</strong> EC model and an<br />

Fe-rich olivine in <strong>the</strong> FC model replace pigeonite. The<br />

quartz phase crystallizing during <strong>the</strong> last stages <strong>of</strong> crystallization<br />

in <strong>the</strong> EC model liquids could react with <strong>the</strong><br />

Ca-poor pyroxene to form more Fe-rich olivine.<br />

Ano<strong>the</strong>r distinction between EC and FC models is that<br />

quartz crystallizes in <strong>the</strong> dry EC models. This occurs at a<br />

temperature <strong>of</strong> 9078C (f ¼92%) for <strong>the</strong> SA liquid and at<br />

9198C (f ¼82%) for <strong>the</strong> CS liquid. Water vapour exsolves<br />

in <strong>the</strong> wet EC model below 9178C (f487%) for <strong>the</strong> SA<br />

liquid only. These phases were not seen in <strong>the</strong> FC models<br />

probably because <strong>the</strong> upper limit on crystallization,<br />

beyond which <strong>the</strong> MELTS model became unstable, was<br />

only 80% for <strong>the</strong> FC models, whereas for <strong>the</strong> EC model<br />

it was 98%. By extrapolating from <strong>the</strong> EC models it is<br />

plausible that quartz would crystallize and H 2O vapour<br />

would exsolve if <strong>the</strong> FC models could reach similar degrees<br />

<strong>of</strong> crystallinity. The crystallization <strong>of</strong> quartz and exsolution<br />

<strong>of</strong> volatiles (H2O) at <strong>the</strong> last stages <strong>of</strong> crystallization is<br />

compatible with <strong>the</strong> granophyric textures observed in <strong>the</strong><br />

segregations.<br />

The LLDs in <strong>the</strong> FC models (Figs 12 and 13) show that<br />

during <strong>the</strong> first several tens <strong>of</strong> degrees <strong>of</strong> cooling, <strong>the</strong><br />

MgO content in <strong>the</strong> liquid decreases sharply and <strong>the</strong>re is<br />

a less marked decrease in CaO, whereas <strong>the</strong> FeO content<br />

<strong>of</strong> <strong>the</strong> liquid increases. These changes can be directly tied<br />

to <strong>the</strong> crystallization <strong>of</strong> Mg-rich pyroxene and Ca-rich<br />

plagioclase, which remove MgO and CaO while allowing<br />

FeO and <strong>the</strong> less compatible elements to build up in <strong>the</strong><br />

liquid. After magnetite joins <strong>the</strong> cotectic <strong>the</strong>re is an<br />

abrupt decrease in <strong>the</strong> FeO and an abrupt increase in<br />

SiO 2 concentrations in <strong>the</strong> liquid. With continued crystallization,<br />

<strong>the</strong> TiO 2 concentration in <strong>the</strong> liquid also decreases<br />

abruptly as <strong>the</strong> magnetite becomes more Ti-rich<br />

and ilmenite (FeTiO3) joins <strong>the</strong> cotectic.<br />

The addition <strong>of</strong> H 2O depresses <strong>the</strong> liquidus temperature<br />

and promotes early crystallization <strong>of</strong> magnetite in both<br />

EC and FC models, decreasing <strong>the</strong> FeO concentrations in<br />

<strong>the</strong> modeled liquids. However, as mentioned above, an<br />

H 2O vapour phase is observed only in one <strong>of</strong> <strong>the</strong> wet EC<br />

models. With continued fractionation and increasing volatiles<br />

<strong>the</strong> MELTS program becomes unstable for <strong>the</strong>se<br />

types <strong>of</strong> compositions, making it impossible to quantify<br />

processes such as <strong>the</strong> formation <strong>of</strong> <strong>the</strong> granophyric intergrowths<br />

and <strong>the</strong> deuteric alteration process in <strong>the</strong><br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 15. (a) SiO 2 (wt %) variation with per cent crystallization for two LLDs corresponding to SA, 93-BS-90-B (circles) and CS, 96-A-33<br />

(squares) dolerite liquids using a closed, hydrous, FC model. The highest silica content <strong>of</strong> <strong>the</strong> segregations (67%) is used to provide an upper<br />

bound on <strong>the</strong> crystallization interval over which <strong>the</strong> segregations formed. For <strong>the</strong> SA liquid <strong>the</strong> segregations crystallized up to 81% and for<br />

<strong>the</strong> CS liquid <strong>the</strong>y crystallized up to 67%. A lower bound <strong>of</strong> 33% crystallization (which is well above <strong>the</strong> cotectic) is used for both liquids.<br />

It should be noted that silica becomes significantly enriched only after magnetite (mt) crystallizes. (b) Temperature vs crystallinity for <strong>the</strong><br />

same two LLDs. The shaded regions represent <strong>the</strong> crystallization interval over which <strong>the</strong> segregations could have formed from <strong>the</strong> SA liquid<br />

(light grey) and <strong>the</strong> CS liquid (dark grey). For <strong>the</strong> SA liquid <strong>the</strong> corresponding crystallization temperatures are between 1138 and 9348C and<br />

for <strong>the</strong> CS liquid <strong>the</strong>y are between 1113 and 9788C.<br />

1956<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

minerals, all <strong>of</strong> which occur at <strong>the</strong> very last stages <strong>of</strong><br />

crystallization.<br />

Crystallinity and temperature during<br />

felsic segregation formation<br />

Using <strong>the</strong> two model LLDs for <strong>the</strong> closed, hydrous FC<br />

model, we can use <strong>the</strong> silica content <strong>of</strong> <strong>the</strong> segregations<br />

(52^67 wt %) to attempt to find bounds on <strong>the</strong> crystallinity<br />

interval over which <strong>the</strong> segregations formed, as illustrated<br />

by grey shaded region in Fig. 15a. The highest silica contents<br />

<strong>of</strong> 67 wt % correspond to 81% crystallization <strong>of</strong><br />

liquid SA and 67% crystallization <strong>of</strong> liquid CS. We should<br />

note, however, that <strong>the</strong> silica content <strong>of</strong> <strong>the</strong> liquid changes<br />

little until magnetite begins to crystallize, which occurs at<br />

59% crystallization for liquid SA and 42% crystallization<br />

for liquid CS. Thus if a segregation forms prior to <strong>the</strong>se<br />

crystallization intervals, it would not be particularly enriched<br />

in silica.<br />

Because <strong>the</strong>re is significant overlap in <strong>the</strong> SiO2 contents<br />

<strong>of</strong> <strong>the</strong> dolerite samples (50 54 wt %) and <strong>the</strong> low SiO 2<br />

segregations in group S 1 (51^55 wt %), it is not possible to<br />

constrain accurately <strong>the</strong> lower bounds on <strong>the</strong> crystallinity<br />

interval over which <strong>the</strong> segregations form. Philpotts &<br />

Dickson (2000) suggested that an interconnected crystalline<br />

mush capable <strong>of</strong> being torn would form at as low as<br />

33% crystallinity in a basaltic magma. From Fig. 15a we<br />

observe that multiple saturation <strong>of</strong> two pyroxenes and<br />

plagioclase is reached at 18% and 16% crystallinity for<br />

liquids SA and CS, respectively, so that an extensive interconnected<br />

network <strong>of</strong> crystals would exist at a crystallinity<br />

<strong>of</strong> 33%. Using SA as <strong>the</strong> initial composition, <strong>the</strong> segregations<br />

might have separated from a mush at between 33<br />

and 81% crystallization <strong>of</strong> <strong>the</strong> parent liquid (light grey<br />

shaded area Fig. 15b), which corresponds to a temperature<br />

<strong>of</strong> between 1138 and 9348C. For a CS initial composition,<br />

and a range <strong>of</strong> crystallinity from 33 to 67%, <strong>the</strong> temperature<br />

range is from 1113 to 9788C (dark grey shaded area<br />

in Fig. 15b).<br />

At crystallinity <strong>of</strong> 33% for <strong>the</strong> CS composition, <strong>the</strong> silica<br />

content <strong>of</strong> <strong>the</strong> residual liquid is about 56%, corresponding<br />

to <strong>the</strong> middle <strong>of</strong> <strong>the</strong> range for <strong>the</strong> group S 3 segregations<br />

and <strong>the</strong> lower end <strong>of</strong> <strong>the</strong> range for <strong>the</strong> S4 segregations<br />

(Fig. 11). This suggests that <strong>the</strong> low-SiO 2 segregations<br />

(groups S 1 and S 2) originated from a parent melt similar<br />

to SA. We note, however, that according <strong>the</strong> physical<br />

model we present below, it is possible that poorly differentiated<br />

or undifferentiated liquids and crystals might be<br />

channelled upward into a segregation lens from deeper in<br />

<strong>the</strong> sill ra<strong>the</strong>r than separating from <strong>the</strong> adjacent crystal<br />

mush.<br />

At temperatures as low as 9348C <strong>the</strong> liquid may be too<br />

viscous to segregate effectively, so we suggest that if <strong>the</strong> segregations<br />

have silica concentrations greater than, say,<br />

55^60 wt % <strong>the</strong>y were probably segregated from a parent<br />

more like CS than SA. A residual liquid <strong>of</strong> CS with 67%<br />

1957<br />

SiO 2 would have separated from a mush that was 67%<br />

crystalline at a temperature <strong>of</strong> 9788C. Alternatively, <strong>the</strong><br />

most silica-rich segregations might have formed from a<br />

parent liquid like SA at high crystallinity, if <strong>the</strong> increasing<br />

concentration <strong>of</strong> volatiles (H2O) significantly reduced <strong>the</strong><br />

viscosity <strong>of</strong> <strong>the</strong> residual liquid, making it possible to segregate<br />

it at <strong>the</strong>se low temperatures.<br />

Summary <strong>of</strong> <strong>the</strong> MELTS modeling<br />

Although none <strong>of</strong> <strong>the</strong> MELTS models carried out in this<br />

study provide good matches for all <strong>of</strong> <strong>the</strong> segregation<br />

data, <strong>the</strong> segregation compositions are best explained by<br />

fractional crystallization (FC) <strong>of</strong> a parent dolerite liquid<br />

at about 100 MPa, with up to 0·5wt % H2O. This model<br />

is in agreement with previous models <strong>of</strong> <strong>the</strong> FDS<br />

(Demarchi et al., 2001) and more recent models on <strong>the</strong> crystallization<br />

<strong>of</strong> <strong>the</strong> BS (Marsh, 2004; Be¤ dard et al., 2007;<br />

Boudreau & Simon, 2007).<br />

The FC model is arguably more appropriate than EC<br />

models owing to rapid crystallization relative to solid-state<br />

diffusion, such that that <strong>the</strong> cores <strong>of</strong> early formed crystals<br />

remain chemically isolated from <strong>the</strong> liquid. This effect<br />

may be enhanced as a result <strong>of</strong> multiple intrusions within<br />

<strong>the</strong> BS (see below).<br />

The segregation compositions do not follow <strong>the</strong> LLDs<br />

expected for residual liquids. A number <strong>of</strong> factors are expected<br />

to contribute to this discrepancy. First, <strong>the</strong> segregation<br />

compositions could represent mixtures <strong>of</strong> liquids and<br />

crystals entrained from <strong>the</strong> mush, as <strong>the</strong>y are filter pressed<br />

through <strong>the</strong> mush to form <strong>the</strong> segregations (Zavala, 2005;<br />

Fodor, 2008). The scatter <strong>of</strong> segregation compositions also<br />

suggests that <strong>the</strong>se segregations were not derived from a<br />

single parent liquid; many <strong>of</strong> <strong>the</strong> segregations fall between<br />

<strong>the</strong> liquid lines <strong>of</strong> descent <strong>of</strong> <strong>the</strong> two dolerites: 93-BS-90-B,<br />

adjacent to a segregation lens, and 96-A-33, from a deeper<br />

level in <strong>the</strong> BS sill.<br />

MODEL FOR FORMATION OF<br />

THE SEGREGATIONS<br />

The proposed physical model for formation <strong>of</strong> <strong>the</strong> segregations<br />

is illustrated in Fig. 16, which shows schematic snapshots<br />

<strong>of</strong> <strong>the</strong> sill at different times. The main solidification<br />

process is based on existing models for sheet-like magma<br />

bodies (Philpotts & Dickson, 2000; Marsh, 2002; Philpotts<br />

& Philpotts, 2005). In <strong>the</strong>se models solidification fronts<br />

propagate inwards with time from <strong>the</strong> upper and lower<br />

boundaries. These fronts are crystal networks with lower<br />

crystallinity and higher temperature than <strong>the</strong> more completely<br />

crystallized magma nearer <strong>the</strong> contact. Near <strong>the</strong><br />

‘capture front’ <strong>the</strong> crystal network becomes weak and very<br />

porous and most crystals are connected toge<strong>the</strong>r on a<br />

short length scale; below <strong>the</strong> front a few crystals exist as<br />

single suspended crystals. The boundary between mush<br />

and magma is a fea<strong>the</strong>ry, irregular region with a large<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


(a)<br />

(b)<br />

(c)<br />

JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

ro<strong>of</strong><br />

surface area providing many sites for nucleation <strong>of</strong> new<br />

crystals. The upper solidification front is inherently gravitationally<br />

unstable, and we expect that in any but <strong>the</strong> thinnest<br />

sills, clusters <strong>of</strong> loosely bound crystals<br />

(glomerocrysts) periodically detach from <strong>the</strong> capture<br />

front and settle onto <strong>the</strong> lower solidification front; some <strong>of</strong><br />

<strong>the</strong>se crystals may be resorbed if <strong>the</strong>y reach <strong>the</strong> hotter<br />

liquid below, but on <strong>the</strong>ir way <strong>the</strong>y stir <strong>the</strong> intervening<br />

magma (Tait & Jaupart, 1992). This stage is illustrated<br />

schematically in Fig. 16a. With time, as <strong>the</strong> solidification<br />

front becomes broader and heavier, larger and more crystalline<br />

clusters can detach (Fig. 16b).<br />

For sills 100 m thick, detachment can occur as propagating,<br />

plastic tears with sharp contacts at levels in <strong>the</strong><br />

front where crystallinities are as low as 30^35%<br />

(Philpotts & Dickson, 2000). The segregations in <strong>the</strong> BS<br />

usually have sharp upper and lower contacts. This suggests<br />

floor<br />

Fig. 16. Schematic model for <strong>the</strong> evolution <strong>of</strong> a thick (4100 m) sill. Not to scale. Panels indicate crystallinity and dynamics <strong>of</strong> a representative<br />

section at different times during solidification. Shades <strong>of</strong> grey indicate degree <strong>of</strong> crystallinity: black 70^100%, dark grey 40^70%, medium<br />

grey 25^40% crystals, lightest grey is crystal-poor magma, and white is segregated residual magma. Arrows indicate direction <strong>of</strong> motion,<br />

hatched pattern indicates settled cumulate crystals. (a^c) Evolution <strong>of</strong> a single pulse <strong>of</strong> crystal-poor magma. (d, e) Evolution after a second<br />

pulse <strong>of</strong> crystal-rich magma, entering from <strong>the</strong> left. The increased thickness <strong>of</strong> <strong>the</strong> section should be noted. (See text for explanation.)<br />

1958<br />

(d)<br />

(e)<br />

that when tearing occurred <strong>the</strong> mush was sufficiently crystalline<br />

to preserve both margins <strong>of</strong> <strong>the</strong> tears. Eventually<br />

<strong>the</strong> upper and lower solidification fronts will meet,<br />

although not necessarily at <strong>the</strong> same time in all locations<br />

(Fig. 16c), filling <strong>the</strong> centre <strong>of</strong> <strong>the</strong> sill with a weak, heterogeneous<br />

crystal network. Tearing and <strong>the</strong> formation <strong>of</strong> segregations<br />

may occur in <strong>the</strong> upper front before <strong>the</strong> fronts<br />

meetçor after <strong>the</strong> fronts meet as <strong>the</strong> weak network compacts.<br />

Tearing ceases as <strong>the</strong> upper front rests stably on <strong>the</strong><br />

lower front.<br />

The weakest, central part <strong>of</strong> <strong>the</strong> crystalline network may<br />

subsequently be disrupted if a new pulse <strong>of</strong> magma inflates<br />

<strong>the</strong> sill (Fig. 16d). The melt injection may rip <strong>of</strong>f and disrupt<br />

weak crystal clusters from <strong>the</strong> fronts and introduce<br />

suspended crystals. The development <strong>of</strong> a crystal network<br />

requires a low-energy environment. In <strong>the</strong> energetic,<br />

forced flow <strong>of</strong> a new magma pulse, any solid fraction will<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

consists <strong>of</strong> single crystals that can move relative to one ano<strong>the</strong>r<br />

and can be sorted by sedimentary processes (e.g.<br />

Middleton, 1970). Sedimentation <strong>of</strong> introduced crystals is<br />

probably responsible for <strong>the</strong> cumulate tongue <strong>of</strong> <strong>the</strong> BS<br />

(Fig. 16d).<br />

When <strong>the</strong> crystal network in <strong>the</strong> centre <strong>of</strong> <strong>the</strong> sill is<br />

replaced with a layer <strong>of</strong> liquid (Fig. 16c and d), <strong>the</strong> upper<br />

solidification front becomes newly gravitationally unstable.<br />

If <strong>the</strong> upper solidification front is thin, a fea<strong>the</strong>ry boundary<br />

region is re-established and solidification continues as<br />

shown in Fig. 16a and b. If, however, <strong>the</strong> solidification<br />

front is thick, <strong>the</strong>n a new magma influx can instigate a<br />

new episode <strong>of</strong> tearing. Where earlier tears are incompletely<br />

crystallized and constitute weak regions in <strong>the</strong> front,<br />

<strong>the</strong>y may be reactivated and extended, propagating sideways<br />

and vertically in subparallel interconnected fractures<br />

(Hersum et al., 2005) (Fig. 16d). Earthquakes and fluid<br />

shear stresses associated with fresh magma influxes would<br />

add to <strong>the</strong> gravitational stresses to cause reactivation and<br />

extension <strong>of</strong> old tears. The sets <strong>of</strong> sub-horizontal anastomosing,<br />

thick ( 0·5^2 m) lenses observed in <strong>the</strong> upper<br />

parts <strong>of</strong> <strong>the</strong> BS are thought to have formed by this process.<br />

Interpretation <strong>of</strong> mafic pegmatoids at <strong>the</strong> Dais suggests a<br />

similar formation process, in which liquids forming <strong>the</strong><br />

pegmatoids were filter pressed from <strong>the</strong> surrounding gabbronorite<br />

during compaction <strong>of</strong> <strong>the</strong> mush (Geist et al.,<br />

2005).<br />

Tearing would cease when <strong>the</strong> gravitational instability<br />

causing <strong>the</strong> tear was eliminated, ei<strong>the</strong>r because <strong>the</strong> mush<br />

at deeper levels beneath <strong>the</strong> tear delaminated, reducing<br />

<strong>the</strong> weight, or because <strong>the</strong> foundering underlying mush<br />

became partially supported by <strong>the</strong> lower solidification<br />

front (Fig. 16e).<br />

The most felsic segregations, felsic pods and stringers,<br />

seen particularly well at <strong>the</strong> Dais, could have formed<br />

during <strong>the</strong> last stages <strong>of</strong> crystallization when an H 2O-rich<br />

fluid phase segregated from <strong>the</strong> residual liquid by<br />

crystallization-induced degassing (i.e. second boiling). It<br />

has been proposed that exsolution <strong>of</strong> gas leads to an increase<br />

in pressure, which drives late-stage residual fluid<br />

from <strong>the</strong> crystal-rich matrix into crystal-poor zones under<br />

lower pressure (Anderson et al., 1984; G<strong>of</strong>f, 1996; Sisson &<br />

Bacon, 1999; Car<strong>of</strong>f et al., 2000; Boudreau & Simon, 2007).<br />

Philpotts et al. (1996) related centimetre-scale felsic blobs<br />

within mafic segregation sheets to exsolution <strong>of</strong> volatiles.<br />

A higher partial pressure <strong>of</strong> H 2O or cooling (e.g. Puffer<br />

& Horter, 1993) could cause fracturing in <strong>the</strong> highly crystalline<br />

region near <strong>the</strong> upper contact <strong>of</strong> <strong>the</strong> BS. Late residual<br />

liquid could collect in <strong>the</strong>se cooling and degassing<br />

fractures to form pods and streaks <strong>of</strong> felsic material<br />

(Fig. 16e).<br />

Multiple influxes <strong>of</strong> magma<br />

The present model differs from that <strong>of</strong> Marsh (2002) in<br />

that <strong>the</strong> tearing process caused by gravitational instability<br />

1959<br />

<strong>of</strong> <strong>the</strong> upper solidification front occurred at various times<br />

and places throughout <strong>the</strong> crystallization <strong>of</strong> <strong>the</strong> BS, triggered<br />

by multiple episodes <strong>of</strong> magma influx. Marsh<br />

(2002) suggested that segregation lenses ceased opening<br />

when <strong>the</strong>y froze, whereas we propose tearing ceased when<br />

<strong>the</strong> gravitational stress was removed.<br />

The existence <strong>of</strong> multiple pulses <strong>of</strong> slightly different<br />

input magma (Philpotts & Philpotts, 2005) would help to<br />

account for <strong>the</strong> range <strong>of</strong> parent liquids suggested from <strong>the</strong><br />

MELTS modelling (Figs 12^14). These pulses are inferred<br />

to have undergone varied degrees <strong>of</strong> fractional crystallization<br />

at deeper levels. The most dramatic evidence for multiple<br />

inputs is <strong>the</strong> opx-rich tongue midway in <strong>the</strong> BS. Its<br />

location indicates that <strong>the</strong> lower BS had previously crystallized<br />

(<strong>the</strong>reby preventing <strong>the</strong> orthopyroxene phenocrysts<br />

from accumulating at <strong>the</strong> lower margin <strong>of</strong> <strong>the</strong> BS). Recent<br />

work on <strong>the</strong> crystallization <strong>of</strong> <strong>the</strong> BS (Marsh, 2004;<br />

Hersum et al., 2007; Be¤ dard et al., 2007; Boudreau &<br />

Simon, 2007) suggests that <strong>the</strong> orthopyroxene-rich tongue<br />

in <strong>the</strong> BS was intruded as a crystal-rich mush after being<br />

texturally equilibrated at a much deeper level.<br />

There is not a strong correlation between <strong>the</strong> SiO 2 concentration<br />

in <strong>the</strong> segregations and <strong>the</strong>ir position in <strong>the</strong><br />

BS. Although high-SiO 2 segregations tend to occur closer<br />

to <strong>the</strong> top contact, low-SiO 2 segregations are found at all<br />

levels (Fig. 17). Additionally, many segregations show internal<br />

layering that may be related to multiple pulses into<br />

a given lens (Table 1). The anomalous group S 2 segregation<br />

sample 94-BS-154-A occurs at a level close to <strong>the</strong> top <strong>of</strong><br />

<strong>the</strong> opx tongue (square at D 100 m in Fig. 17). This location<br />

in combination with its unusual composition suggests<br />

that it is related to <strong>the</strong> opx tongue intrusion, that it may<br />

in fact be a part <strong>of</strong> that crystal-rich influx forced into a<br />

tear in <strong>the</strong> dolerite mush.<br />

Therefore, it is suggested that <strong>the</strong>re may have been multiple<br />

episodes <strong>of</strong> segregation formation triggered by fresh<br />

influxes <strong>of</strong> magma into <strong>the</strong> sill, and that old tears could<br />

have been reactivated at different times. This is consistent<br />

with <strong>the</strong> wide range <strong>of</strong> temperatures and predicted crystallinities<br />

for <strong>the</strong> melt segregations. Moreover, tears in <strong>the</strong><br />

mush may propagate by a system <strong>of</strong> fractures (Philpotts<br />

et al., 1996; Zavala, 2005) allowing liquid derived from <strong>the</strong><br />

crystal mush at one level to move to a higher level, so that<br />

SiO 2-poor segregations need not necessarily be found adjacent<br />

to <strong>the</strong>ir parent mush. For instance, segregations in<br />

<strong>the</strong> Kilauea Iki lava lake include material from <strong>the</strong> adjacent<br />

mush and a deeper source ( 77^105 m) in <strong>the</strong> lake<br />

(T. R. Helz, personal communication, 2007).<br />

Mixed liquids and crystals<br />

Previous workers have considered that felsic segregations<br />

are filter-pressed residual liquids, whereas here we suggest<br />

that <strong>the</strong>y are made up <strong>of</strong> mixtures <strong>of</strong> crystals and residual<br />

liquids that might come from deeper levels in <strong>the</strong> sill as<br />

well as locally. If tearing is an abrupt process triggered by<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Fig. 17. SiO 2 (wt %) content <strong>of</strong> rocks vs distance from <strong>the</strong> upper contact <strong>of</strong> <strong>the</strong> BS. S 1 dolerites (light grey diamonds), S 2 segregation (squares),<br />

S3 segregation ( ), S4 (triangle), S5 (circles), cumulates (þ), and <strong>Ferrar</strong> <strong>Dolerite</strong>s elsewhere (black diamonds). All segregations occur within<br />

<strong>the</strong> upper 110 m <strong>of</strong> <strong>the</strong> BS.<br />

influxes <strong>of</strong> magma, ra<strong>the</strong>r than a very gradual process,<br />

<strong>the</strong>n it is easy to envisage that crystals could be entrained<br />

from <strong>the</strong> mush, whereas o<strong>the</strong>rs were trapped in <strong>the</strong> mush<br />

owing to unfavourable aspect ratios. The spread in <strong>the</strong><br />

bulk chemical compositions <strong>of</strong> <strong>the</strong> segregations and <strong>the</strong><br />

way in which <strong>the</strong> segregations can be grouped according<br />

to <strong>the</strong> minerals added to or subtracted from a residual<br />

liquid support this idea. The regular but varied range in<br />

texture, as well as <strong>the</strong> presence <strong>of</strong> deformed pyroxene<br />

and plagioclase crystals, are additional lines <strong>of</strong> evidence.<br />

Helz et al., 1989 reported a narrow compositional range<br />

( 1^6 wt % CaO) for filter-pressed interstitial liquids<br />

found in <strong>the</strong> upper parts <strong>of</strong> <strong>the</strong> Kilauea Iki lava lake. The<br />

fact that <strong>the</strong> BS segregations show such a wide compositional<br />

range (52^67 wt % SiO 2 and 4^13 wt % CaO)<br />

could suggest that <strong>the</strong> segregations represent mixtures <strong>of</strong><br />

crystals and liquids.<br />

SUMMARY AND CONCLUSIONS<br />

Coarse-grained leucocratic segregations are a ubiquitous<br />

feature within <strong>the</strong> upper 100 m <strong>of</strong> <strong>the</strong> 300 m thick<br />

Basement Sill (BS), <strong>the</strong> lowermost <strong>of</strong> four related <strong>Ferrar</strong><br />

<strong>Dolerite</strong> <strong>Sills</strong> exposed in <strong>the</strong> McMurdo DryValleys region<br />

1960<br />

<strong>of</strong> South Victoria Land, <strong>Antarctica</strong> (Fig. 1). Based on field<br />

observations, petrography, major element geochemistry<br />

and <strong>the</strong>rmodynamic modelling, we suggest that <strong>the</strong> segregations<br />

were formed by tearing <strong>of</strong> <strong>the</strong> upper solidification<br />

front <strong>of</strong> <strong>the</strong> partially solidified BS as a result <strong>of</strong> gravitational<br />

instability, and that <strong>the</strong>ir range <strong>of</strong> compositions and textures<br />

can be explained by multiple episodes <strong>of</strong> tearing and<br />

infilling by liquids and crystals <strong>of</strong> various compositions<br />

during <strong>the</strong> crystallization <strong>of</strong> <strong>the</strong> BS.<br />

<strong>Segregations</strong> within <strong>the</strong> BS have bulk chemical compositions<br />

and mineralogy that suggest <strong>the</strong>y are composed <strong>of</strong><br />

varying proportions <strong>of</strong> interstitial liquid and crystals<br />

derived during fractional crystallization <strong>of</strong> <strong>the</strong> host dolerite.<br />

Their compositions form a continuum between <strong>the</strong><br />

host sill (52 wt % SiO 2) and <strong>the</strong> most fractionated segregations<br />

(67 wt % SiO 2), following a tholeiitic differentiation<br />

trend (Fig. 10b).<br />

There are systematic changes in mineralogy and texture<br />

with increasing silica concentration. <strong>Dolerite</strong>s have fineto<br />

medium-grained ophitic and subophitic textures, with<br />

minor amounts <strong>of</strong> interstitial quartz and alkali feldspar in<br />

granophyric intergrowths, opaque minerals and alteration<br />

minerals. As <strong>the</strong> silica content increases, <strong>the</strong> segregations<br />

become coarser-grained and develop sub-ophitic and<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

poikilophitic textures, and <strong>the</strong> amounts <strong>of</strong> secondary alteration,<br />

opaque minerals and granophyric intergrowths increase.<br />

These features are explained by separation <strong>of</strong><br />

residual liquid, plus or minus crystals, from <strong>the</strong> crystallizing<br />

magma over a range <strong>of</strong> crystallinities. In segregations<br />

with intermediate compositions and variable grain size,<br />

some augite crystals are bent and some plagioclase crystals<br />

are broken, bent and have kinked twins, suggesting <strong>the</strong>se<br />

crystals were entrained in a partially crystalline, deforming<br />

medium.<br />

Modelling using <strong>the</strong> <strong>the</strong>rmodynamic program MELTS<br />

demonstrates that <strong>the</strong> major element compositions <strong>of</strong> <strong>the</strong><br />

segregations are largely consistent with <strong>the</strong>ir being residual<br />

liquids formed at various stages during fractional crystallization<br />

<strong>of</strong> a basaltic magma. The segregations could not<br />

have been derived from a single parent liquid; however,<br />

three dolerite samples from different positions in <strong>the</strong> BS<br />

constrain a range <strong>of</strong> fractionation paths that could account<br />

for most <strong>of</strong> <strong>the</strong> felsic segregation compositions. This suggests<br />

that <strong>the</strong> BS was built up from more than one pulse<br />

<strong>of</strong> slightly heterogeneous magma. However, considering<br />

<strong>the</strong> petrographic evidence <strong>of</strong> resorbed and deformed crystals,<br />

it is likely that <strong>the</strong> scatter in segregation compositions<br />

is at least partly due to <strong>the</strong> presence <strong>of</strong> crystals entrained<br />

into <strong>the</strong> melt lenses with <strong>the</strong> residual liquid.<br />

The above observations indicate that <strong>the</strong> felsic segregations<br />

were derived from <strong>the</strong> host dolerite mushes at various<br />

stages <strong>of</strong> <strong>the</strong>ir crystallization. Most segregations, in particular<br />

<strong>the</strong> most common, sub-horizontal anastomosing<br />

lenses, appear to have formed by tearing <strong>of</strong> partially crystallized<br />

mush within <strong>the</strong> upper solidification front and infilling<br />

<strong>of</strong> <strong>the</strong> tears with a mixture <strong>of</strong> interstitial liquid and<br />

crystals from <strong>the</strong> adjacent mush and from deeper in <strong>the</strong><br />

sill.<br />

Based on <strong>the</strong> distribution <strong>of</strong> segregation compositions<br />

with location in <strong>the</strong> BS (Fig. 17), we suggest that multiple<br />

episodes <strong>of</strong> tearing occurred over <strong>the</strong> same range <strong>of</strong><br />

elevations within <strong>the</strong> upper solidification front at different<br />

times. The episodes could have been triggered by new<br />

influxes <strong>of</strong> magma into <strong>the</strong> sill, which swept away a weak<br />

crystal network that was stabilizing <strong>the</strong> upper front, and<br />

suspended it on a layer <strong>of</strong> magma, making it newly susceptible<br />

to tearing. A thick, coarse-grained cumulate ‘tongue’<br />

located midway in <strong>the</strong> sill indicates that <strong>the</strong>re was at least<br />

one major late-stage magma influx event into <strong>the</strong> BS.<br />

Collection <strong>of</strong> residual liquid within <strong>the</strong> segregations was<br />

initially by porous flow through a crystal mush adjacent<br />

to <strong>the</strong> tears, and <strong>the</strong>n by channel flow along propagating<br />

tears and fractures. Stratified segregations are thought to<br />

represent tears that were reactivated multiple times,<br />

whereas <strong>the</strong> uniformly coarse-grained segregations with<br />

gradational contacts are thought to represent residual liquids<br />

that were injected into a cooler part <strong>of</strong> <strong>the</strong> mush via<br />

fractures or channels.<br />

1961<br />

Smaller, irregular, sub-vertical veins, pods and stringers<br />

are located above <strong>the</strong> anastomosing lenses, where <strong>the</strong><br />

mush was presumably cooler and more crystalline.<br />

Vertical transport at this level could have been aided by<br />

tearing or weaknesses in <strong>the</strong> front that formed in response<br />

to cooling and contraction, exsolution <strong>of</strong> volatiles at high<br />

degrees <strong>of</strong> crystallinities effectively reducing <strong>the</strong> viscosity<br />

<strong>of</strong> <strong>the</strong> liquid, and by <strong>the</strong> positive buoyancy <strong>of</strong> collections<br />

<strong>of</strong> interstitial liquid derived from deeper levels in <strong>the</strong> front<br />

similar to <strong>the</strong> segregation veins in <strong>the</strong> Kilauea Iki lava<br />

lake (Helz, 1980, 1987; Helz et al., 1989).<br />

A key objective in igneous petrology is to understand <strong>the</strong><br />

range <strong>of</strong> physical processes by which rocks <strong>of</strong> diverse composition<br />

can be produced. Many such processes involve<br />

<strong>the</strong> physical separation <strong>of</strong> crystals and liquid in partially<br />

molten systems. The segregations at <strong>the</strong> BS have compositions<br />

intermediate between basaltic and rhyolitic rocks<br />

that are rarely found in large-scale basaltic systems (e.g.<br />

Hildreth, 1979; Ewart, 1982; Marsh et al., 1991), and studies<br />

like <strong>the</strong> present one can provide insight into <strong>the</strong> processes<br />

that operate at <strong>the</strong>se intermediate stages <strong>of</strong> differentiation.<br />

Our model, involving gravitational detachment <strong>of</strong> unstable<br />

crystal mushes, <strong>the</strong> propagation <strong>of</strong> plastic tears, and <strong>the</strong><br />

movement <strong>of</strong> interstitial liquids by porous flow, channel<br />

flow and diapirism, takes important elements from <strong>the</strong><br />

models <strong>of</strong> previous workers (e.g. Tait & Jaupart, 1992;<br />

Philpotts et al., 1996; Marsh, 2002; Bachmann & Bergantz,<br />

2004). However, we argue for <strong>the</strong> importance <strong>of</strong> multiple<br />

inputs <strong>of</strong> magma into thick sills in generating and preserving<br />

segregation features, and point out that <strong>the</strong> material<br />

entering <strong>the</strong> segregations is rarely pure interstitial liquid<br />

(Zavala, 2005).<br />

Deformation and disruption <strong>of</strong> <strong>the</strong> crystal mush in <strong>the</strong><br />

sill and particularly in <strong>the</strong> gravitationally unstable upper<br />

solidification front are key elements <strong>of</strong> our model. To fully<br />

understand <strong>the</strong> temporal and spatial relationships <strong>of</strong> segregations<br />

like those in <strong>the</strong> BS, a better physical understanding<br />

<strong>of</strong> <strong>the</strong> rheology <strong>of</strong> crystal mushes is needed. Fur<strong>the</strong>r<br />

field and geochemical studies, and ma<strong>the</strong>matical and experimental<br />

modeling <strong>of</strong> crystal^liquid separation will be<br />

essential for fully understanding <strong>the</strong> physical processes <strong>of</strong><br />

segregation formation.<br />

ACKNOWLEDGEMENTS<br />

K.Z. would like to thank Bruce Marsh for <strong>the</strong> opportunity<br />

to participate in his Dry Valleys <strong>Antarctica</strong> project, and<br />

for input in <strong>the</strong> early stages <strong>of</strong> this work. K.Z. also thanks<br />

Maya Wheelock for collecting <strong>the</strong> samples that were analyzed<br />

in this study. Toby Rivers, Jean Be¤ dard, Teal Riley,<br />

Rosalind Helz and an anonymous reviewer are thanked<br />

for <strong>the</strong>ir helpful reviews. Careful and constructive comments<br />

from Executive Editor Marjorie Wilson are greatly<br />

appreciated.<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


FUNDING<br />

This research was supported by NSF Research grant<br />

P420C652232.<br />

SUPPLEMENTARY DATA<br />

Supplementary data for this paper are available at Journal<br />

<strong>of</strong> Petrology online.<br />

REFERENCES<br />

JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Abbey, S. (1983). Studies in ‘Standard Samples’ <strong>of</strong> silicate rocks minerals<br />

1969^1982. Geological Survey <strong>of</strong> Canada Paper 83, 1^114.<br />

Anderson, T. A., Jr, Swihart, H. G., Artioli, G. & Geiger, A. C. (1984).<br />

Segregation vesicles, gas filter-pressing, and igneous differentiation.<br />

Journal <strong>of</strong> Geology 92, 55^72.<br />

Antonini, P., Piccirillo, M. E., Petrini, R., Civetta, L., D’Antonio, M.<br />

& Orsi, G. (1999). Enriched mantle-Dupal signature in <strong>the</strong> genesis<br />

<strong>of</strong> <strong>the</strong> Jurassic <strong>Ferrar</strong> tholeiites from Prince Albert Mountains<br />

(Victoria Land, <strong>Antarctica</strong>). Contributions to Mineralogy and Petrology<br />

136,1^19.<br />

Armstrong, J. T. (1989). CITZAF: Combined ZAF and Phi-rho (Z) Electron<br />

Beam Correction Programs. Pasadena, CA: California Institute <strong>of</strong><br />

Technology.<br />

Asimow, P. D. & Ghiorso, M. S. (1998). Algorithmic modifications extending<br />

MELTS to calculate subsolidus phase relations. American<br />

Mineralogist 83, 1127^1132.<br />

Asimow, P. D., Hirschmann, M. M. & Stolper, E. M. (2001).<br />

Calculation <strong>of</strong> peridotite melting from <strong>the</strong>rmodynamic models <strong>of</strong><br />

minerals and melts, IV. Adiabatic decompression and <strong>the</strong> composition<br />

and mean properties <strong>of</strong> mid-ocean ridge basalts. Journal <strong>of</strong><br />

Petrology 42, 963^998.<br />

Bachmann, O. & Bergantz, G. W. (2004). On <strong>the</strong> origin <strong>of</strong> crystal-poor<br />

rhyolites: extracted from batholithic mushes. Journal <strong>of</strong> Petrology 45,<br />

1565^1582.<br />

Bailey, M. M. (1989). Evidence for magma recharge and assimilation<br />

in <strong>the</strong> Picture Gorge Basalt Subgroup, Columbia River Basalt<br />

Group. In: Reidel, S. P. & Hooper, P. R. (eds) Volcanism and<br />

Tectonism in <strong>the</strong> Columbia River Flood-Basalt Province. Geological Society<br />

<strong>of</strong> America, Special Papers 239,343^355.<br />

Be¤ dard, J. H. J., Marsh, D. B., Hersum, G. T., Naslund, H. R. &<br />

Mukasa, B. S. (2007). Large-scale mechanical redistribution <strong>of</strong><br />

orthopyroxene and plagioclase in <strong>the</strong> Basement Sill, <strong>Ferrar</strong><br />

<strong>Dolerite</strong>s, McMurdo Dry Valleys, Antractica: petrological,<br />

mineral-chemical and field evidence for channelized movement <strong>of</strong><br />

crystals and melt. Journal <strong>of</strong> Petrology 48, 2289^2326.<br />

Boudreau, A. & Simon, A. (2007). Crystallization and degassing in <strong>the</strong><br />

Basement Sill, McMurdo Dry Valleys, <strong>Antarctica</strong>. Journal <strong>of</strong><br />

Petrology 48, 1369^1386.<br />

Bowen, N. L. (1928). The Evolution <strong>of</strong> <strong>the</strong> Igneous Rocks. NewYork: Dover.<br />

Boyd, F. R. & Mertzman, S. A. (1987). Composition and structure <strong>of</strong><br />

<strong>the</strong> Kaapvaal lithosphere, Sou<strong>the</strong>rn Africa. In: Mysen, B. O. (ed.)<br />

Magmatic Processes. Geochemical Society Special Publications 1,13^24.<br />

Brewer, T. S., Hergt, J. M., Hawkesworth, C. J., Rex, D. & Storey, B.<br />

C. (1992). Coats Land dolerites and <strong>the</strong> generation <strong>of</strong> Antarctic<br />

continental flood basalts. In: Storey, B. C., Alabaster, T. &<br />

Pankhurst, R. J. (eds) Magmatism and <strong>the</strong> Causes <strong>of</strong> Continental<br />

Break-up. Geological Society, London, Special Publications 68, 185^208.<br />

Car<strong>of</strong>f, M., Maury, C. M., Cotten, J. & Clement, J. P. (2000).<br />

Segregation structures in vapor differentiated basaltic flows.<br />

Bulletin <strong>of</strong> Volcanology 62, 171^187.<br />

1962<br />

Cornwall, R. H. (1951). Differentiation in lavas <strong>of</strong> <strong>the</strong> Keweenawan<br />

series and <strong>the</strong> origin <strong>of</strong> <strong>the</strong> copper deposits <strong>of</strong> Michigan. Geological<br />

Society <strong>of</strong> America Bulletin 62, 159^202.<br />

Demarchi, G., Antonini, P., Piccirillo, M., Orsi, G., Civetta, L. &<br />

D’Antonio, M (2001). Significance <strong>of</strong> orthopyroxene and major<br />

element constraints on <strong>the</strong> petrogenesis <strong>of</strong> <strong>Ferrar</strong> tholeiites from<br />

sou<strong>the</strong>rn Prince Albert Mountains, Victoria Land, <strong>Antarctica</strong>.<br />

Contributions to Mineralogy and Petrology 142(2),127^146.<br />

Eason, D. & Sinton, J. (2006). Origin <strong>of</strong> high-Al N-MORB by fractional<br />

crystallization in <strong>the</strong> upper mantle beneath <strong>the</strong> Galapagos<br />

Spreading Center. Earth and Planetary Science Letters 252,423^436.<br />

Elliot, D. H. & Fleming, T. H. (2004). Occurrence and dispersal <strong>of</strong><br />

magmas in <strong>the</strong> Jurassic <strong>Ferrar</strong> Large Igneous Province,<br />

<strong>Antarctica</strong>. Gondwana Research 7, 223^237.<br />

Elliot, D. H. & Fleming, T. H. (2008). Physical volcanology and geological<br />

relationships <strong>of</strong> <strong>the</strong> Jurassic <strong>Ferrar</strong> Large Igneous Province,<br />

<strong>Antarctica</strong>. Journal <strong>of</strong> Volcanology and Geo<strong>the</strong>rmal Research 172,20^37.<br />

Encarnacion, J., Fleming, H. T., Elliot, H. D. & Eales, V. H. (1996).<br />

Synchronous emplacement <strong>of</strong> <strong>Ferrar</strong> and Karoo dolerites and <strong>the</strong><br />

early breakup <strong>of</strong> Gondwana. Geology 24, 53^58.<br />

Ewart, A. (1982). The mineralogy and petrology <strong>of</strong> Tertiary^Recent<br />

orogenic volcanic rocks with special reference to <strong>the</strong> andesite^<br />

basalt compositional range. In: Thorpe, R. S. (ed.) Andesites;<br />

Orogenic Andesites and Related Rocks. Chichester: John Wiley,<br />

pp. 1^200.<br />

Fisher, G. W. (1989). Matrix analysis <strong>of</strong> metamorphic assemblages and<br />

reactions. Contributions to Mineralogy and Petrology 102,69^77.<br />

Fleming, T. H., Heimann, A., Foland, K. A. & Elliot, D. H. (1997).<br />

40 Ar/ 39 Ar geochronology <strong>of</strong> <strong>Ferrar</strong> <strong>Dolerite</strong> sills from <strong>the</strong><br />

Transantarctic Mountains, <strong>Antarctica</strong>; implications for <strong>the</strong> age<br />

and origin <strong>of</strong> <strong>the</strong> <strong>Ferrar</strong> magmatic province. Geological Society <strong>of</strong><br />

America Bulletin 109, 533^546.<br />

Fodor, V. R. (2008). Diorite segregations in gabbro: geochemical characteristics<br />

and conditions for origin assessed at diorite^gabbro contacts.<br />

Journal <strong>of</strong> Geology 117, 109^125.<br />

Geist, D., Boudreau, A., Garcia, M., Harpp, K., Ma<strong>the</strong>z, E. &<br />

Marsh, B. (2005). Origin <strong>of</strong> magic pegmatoids in <strong>the</strong> Dais<br />

Intrusion, Wright Valley, <strong>Antarctica</strong>, American Geophysical<br />

Union Meeting, San Francisco, abstract V14C-07.<br />

Ghiorso, M. S. & Carmichael, I. S. E. (1985). Chemical mass transfer<br />

in magmatic processes. II. Applications in equilibrium crystallization,<br />

fractional and assimilation. Contributions to Mineralogy and<br />

Petrology 90,121^141.<br />

Ghiorso, M. S. & Sack, R. O. (1995). Chemical mass transfer in<br />

magmatic processes IV. A revised and internally consistent<br />

<strong>the</strong>rmodynamic model for <strong>the</strong> interpolation and extrapolation<br />

<strong>of</strong> liquid^solid equilibria in magmatic systems at elevated temperatures<br />

and pressures. Contributions to Mineralogy and Petrology 119,<br />

197^212.<br />

G<strong>of</strong>f, F. (1996). Vesicle cylinders in vapor-differentiated basalt flows.<br />

Journal <strong>of</strong> Volcanology and Geo<strong>the</strong>rmal Research 97,167^185.<br />

Govindaraju, K. (1994). Compilation <strong>of</strong> Working Values and Sample<br />

Description for 383 Geostandards, Special Issue. Geostandards<br />

Newsletter 18,1^158.<br />

Greenough, D. J. & Dostal, J. (1992). Cooling history and differentiation<br />

<strong>of</strong> a thick North Mountain Basalt Flow <strong>of</strong> Nova Scotia,<br />

Canada. Bulletin <strong>of</strong> Volcanology 55,63^73.<br />

Gunn, B. M. (1962). Differentiation in <strong>Ferrar</strong> <strong>Dolerite</strong>s, <strong>Antarctica</strong>.<br />

New ZealandJournal <strong>of</strong> Geology and Geophysics 5, 820^863.<br />

Gunn, B. M. & Warren, G. (1962). Geology <strong>of</strong> Victoria Land between<br />

<strong>the</strong> Mawson and Mulock Glaciers, <strong>Antarctica</strong>. New Zealand<br />

Geological Survey Bulletin 71, 1^157.<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


ZAVALA et al. SILICIC SEGREGATIONS, FERRAR DOLERITE SILLS<br />

Hamilton, W. (1965). Diabase Sheets <strong>of</strong> <strong>the</strong> Taylor Glacier Region<br />

Victoria Land, <strong>Antarctica</strong>. US Geological Survey, Pr<strong>of</strong>essional Papers<br />

456-B,1^71.<br />

Harker, A. (1909). The Natural History <strong>of</strong> Igneous Rocks. London:<br />

Methuen.<br />

Heimann, A., Fleming, T. H., Foland, K. A. & Elliot, D. H. (1994). A<br />

short interval <strong>of</strong> continental flood basalt volcanism in <strong>Antarctica</strong><br />

as demonstrated by 40 Ar/ 39 Ar geochronology. Earth and Planetary<br />

Science Letters 121, 19^41.<br />

Helz, T. R. (1980). Crystallization history <strong>of</strong> Kilauea Iki Lava Lake as<br />

seen in drill core recovered in 1967^1979. Bulletin <strong>of</strong> Volcanology 43,<br />

675^701.<br />

Helz, T. R. (1987). Differentiation behavior <strong>of</strong> Kilauea Iki Lava Lake,<br />

Kilauea Volcano, Hawaii: An overview <strong>of</strong> past and current work.<br />

In: Mysen, B. O. (ed.) Magmatic Processes: Physiochemical Principles.<br />

Geochemical Society, Special Publications 1, 241^258.<br />

Helz, T. R., Kirschenbaum, H. & Marinenko, W. J. (1989). Diapiric<br />

transfer <strong>of</strong> melt in Kilauea Iki Lava Lake, Hawaii: A quick, efficient<br />

process <strong>of</strong> igneous differentiation. Geological Society <strong>of</strong> America<br />

Bulletin 101, 578^594.<br />

Hersum, G. T., Hilpert, M. & Marsh, B. D. (2005). Permeability and<br />

melt flow in simulated and natural partially molten basaltic<br />

magmas. Earth & Planetary Science Letters 237, 798^814.<br />

Hersum, G. T., Marsh, D. B. & Simon, C. A. (2007). Contact partial<br />

melting <strong>of</strong> granitic country rock, melt segregation, and re-injection<br />

as dikes into <strong>Ferrar</strong> <strong>Dolerite</strong> <strong>Sills</strong>, McMurdo Dry Valleys,<br />

<strong>Antarctica</strong>. Journal <strong>of</strong> Petrology 48, 2125^2148.<br />

Hildreth, W. (1971). The Bishop Tuff; Evidence for <strong>the</strong> origin <strong>of</strong> compositional<br />

zonation in silicic magma chambers. In: Chapin, C. E. &<br />

Elston, W. E. (eds) Ash-flow tuffs: Geological Society <strong>of</strong> America Special<br />

Paper, 180, 43^75.<br />

Hildreth, W. (1979). The Bishop Tuff: Evidence for <strong>the</strong> origin <strong>of</strong> compositional<br />

zonation in silicic magma chambers. Geological Society <strong>of</strong><br />

America, Special Papers 180, 43^75.<br />

Hirshmann, M. M., Asimow, P. D., Ghiorso, M. S. & Stolper, E. M.<br />

(1999). Calculation <strong>of</strong> peridotite partial melting from <strong>the</strong>rmodynamic<br />

models <strong>of</strong> minerals and melts. III. Controls on isobaric<br />

melt production and <strong>the</strong> effect <strong>of</strong> water on melt production.<br />

Journal <strong>of</strong> Petrology 40,831^851.<br />

Holmes, A. (1931). The problem <strong>of</strong> <strong>the</strong> association <strong>of</strong> acid and basic<br />

rocks in central complexes. Geological Magazine 68, 241^255.<br />

Kyle, P. (1980). Development <strong>of</strong> heterogeneities in <strong>the</strong> subcontinental<br />

mantle: evidence from <strong>the</strong> <strong>Ferrar</strong> Group, <strong>Antarctica</strong>. Contributions<br />

to Mineralogy and Petrology 73, 89^104.<br />

Larsen, B. R. & Brooks, C. K. (1994). Origin and evolution <strong>of</strong> gabbroic<br />

pegmatites in <strong>the</strong> Skaergaard intrusion, East Greenland. Journal <strong>of</strong><br />

Petrology 35, 1651^1679.<br />

Lindsley, D. H., Smith, D. & Haggerty, S. E. (1971). Petrography and<br />

mineral chemistry <strong>of</strong> a differentiated flow <strong>of</strong> Picture Gorge basalt<br />

near Spray, Oregon: geologic setting and mechanism <strong>of</strong> differentiation.<br />

Carnegie Institution <strong>of</strong> Washington Yearbook 69, 264^285.<br />

Marsh, B. D. (1996). Solidification fronts and magmatic evolution.<br />

Mineralogical Magazine 60, 5^40.<br />

Marsh, B. D. (2002). On bimodal differentiation by solidification front<br />

instability in basaltic magmas, part 1: basic mechanics. Geochimica<br />

et Cosmochimica Acta 66,2211^2229.<br />

Marsh, B. D. (2004). A magmatic mush column Rosetta Stone: <strong>the</strong><br />

McMurdo dry valleys <strong>of</strong>. Antractica. EOS Transactions, American<br />

Geophysical Union 85, 497^502.<br />

Marsh, B. D., Gunnarsson, B., Congdon, R. & Carmody, R. (1991).<br />

Hawaiian basalt and Icelandic rhyolite: indicators <strong>of</strong> differentiation<br />

and partial melting. Geologische Rundschau 80, 481^510.<br />

1963<br />

McBirney, A. R. (1975). Differentiation <strong>of</strong> <strong>the</strong> Skaergaard Intrusion.<br />

Nature 253, 691^694.<br />

McKelvey, B. C. & Webb, P. N. (1962). Geological investigations in<br />

sou<strong>the</strong>rn Victoria Land, <strong>Antarctica</strong>. New Zealand Journal <strong>of</strong><br />

Geophysical Research 5,143^162.<br />

Middleton, G. V. (1970). Experimental studies related to problems <strong>of</strong><br />

flysch sedimentation. In: Lajoie, J. (ed.) Flysch Sedimentation in North<br />

America. Geological Association <strong>of</strong> Canada, Special Papers 7,253^272.<br />

Minor, R. D. & Mukasa, B. S. (1997). Zircon U^Pb and hornblende<br />

40 Ar^ 39 Ar ages for <strong>the</strong> Dufek layered mafic intrusion, <strong>Antarctica</strong>:<br />

Implications for <strong>the</strong> age <strong>of</strong> <strong>the</strong> <strong>Ferrar</strong> large igneous province.<br />

Geochimica et Cosmochimica Acta 61, 2497^2504.<br />

Moore, J. G. & Evans, B. W. (1967). The role <strong>of</strong> olivine in <strong>the</strong> crystallization<br />

<strong>of</strong> <strong>the</strong> prehistoric Makaopuhi tholeiitic Lava Lake,<br />

Hawaii. Contributions to Mineralogy and Petrology 15, 202^223.<br />

Philpotts, A. R. & Dickson, L. D. (2000). The formation <strong>of</strong> plagioclase<br />

chains during convective transfer in basaltic magma. Nature 406,<br />

59^61.<br />

Philpotts, A. R. & Philpotts, D. E. (2005). Crystal-mush compaction<br />

in <strong>the</strong> Cohassett flood-basalt flow, Hanford, Washington. Journal <strong>of</strong><br />

Volcanology and Geo<strong>the</strong>rmal Research 145, 192^206.<br />

Philpotts, A. R., Carroll, M. & Hill, M. J. (1996). Crystal-mush compaction<br />

and <strong>the</strong> origin <strong>of</strong> pegmatitic segregation sheets in a thick<br />

flood-basalt flow in <strong>the</strong> Mesosozic Hartford Basin, Connecticut.<br />

Journal <strong>of</strong> Petrology 37, 811^836.<br />

Puffer, H. J. & Horter, L. D. (1993). Origin <strong>of</strong> pegmatitic segregation<br />

veins within flood basalts. Geological Society <strong>of</strong> America Bulletin 105,<br />

738^748.<br />

Puffer, H. J. & Volkert, A. R. (2001). Pegmatoid and gabbroid layers in<br />

Jurassic Preakness and Hook Mountain Basalts, Newark Basin,<br />

New Jersey. Journal <strong>of</strong> Geology 109, 585^601.<br />

Ragland, P. C. (1989). Basic Analytical Petrology. New York: Oxford<br />

University Press.<br />

Ragland, P. C. & Arthur, J. D. (1985). Petrology <strong>of</strong> <strong>the</strong> Boyds diabase<br />

sheet, nor<strong>the</strong>rn Culpeper Basin, Maryland. In: Robinson, G. R.,<br />

Jr & Froelich, A. J. (eds) Proceedings <strong>of</strong> <strong>the</strong> Second US Geological<br />

Survey Workshop on <strong>the</strong> Early Mesozoic Basins <strong>of</strong> <strong>the</strong> Eastern United<br />

States. US Geological Survey Circular 946, 91^99.<br />

Richter, D. H. & Moore, J. G. (1966). Petrology <strong>of</strong> <strong>the</strong> Kilauea Iki Lava<br />

Lake, Hawaii. US Geological Survey, Pr<strong>of</strong>essional Papers P 0537-B,<br />

B1^B26.<br />

Sisson, W. T. & Bacon, R. C. (1999). Gas-driven filter pressing in<br />

magmas. Geology 27, 613^616.<br />

Steiner, J. C., Walker, R. J., Warner, R. D. & Olson, T. R. (1992). A cumulus^transport^deposition<br />

model for <strong>the</strong> differentiation <strong>of</strong> <strong>the</strong><br />

Palisades sill. In: Puffer, J. H. & Ragland, P. C. (eds) Eastern North<br />

American Mesozoic Magmatism. Geological Society <strong>of</strong> America, Special<br />

Papers 268, 193^217.<br />

Streckeisen, L. A. (1973). Plutonic rocks classification and nomenclature<br />

recommended by <strong>the</strong> IUGS Subcommission on <strong>the</strong><br />

Systematics <strong>of</strong> Igneous Rocks. Geotimes 18, 26^30.<br />

Tait, S. & Jaupart, C. (1992). Compositional convection in a reactive<br />

crystalline mush and melt differentiation. Journal <strong>of</strong> Geophysical<br />

Research 97, 6735^6756.<br />

Tomkeieff, S. I. (1929). A contribution to <strong>the</strong> petrology <strong>of</strong> <strong>the</strong> Whin<br />

Sill. Journal <strong>of</strong> <strong>the</strong> Mineralogical Society 126, 100^120.<br />

Wager, L. R. & Deer, W. A. (1939). Geological investigations in East<br />

Greenland, Part III. The petrology <strong>of</strong> <strong>the</strong> Skaergaard intrusion,<br />

Kangerdlugssuaq, East Greenland. Meddelelser om GrÖnland 105,<br />

1^352.<br />

Walker, F. (1953). The pegmatitic differentiates <strong>of</strong> basic sheets.<br />

AmericanJournal <strong>of</strong> Science 251, 41^60.<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013


JOURNAL OF PETROLOGY VOLUME 52 NUMBER 10 OCTOBER 2011<br />

Wheelock, M. M. & Marsh, B. D. (1993). <strong>Silicic</strong> segregations in basaltic<br />

sills: bimodal differentiation. EOS Transactions, American<br />

Geophysical Union 74, 336^337.<br />

Winter,D.J.(2001).An Introduction to Igneous and Metamorphic Petrology.<br />

Upper Saddle River, NJ: Prentice Hall.<br />

Wright,L.T.&Okamura,T.R.(1977).Cooling and crystallization <strong>of</strong> tholeiitic<br />

basalt, 1965 Makaopuhi Lava Lake, Hawaii. US Geological Survey,<br />

Pr<strong>of</strong>essional Papers 1004,1^78.<br />

Wright,L.T.,Peck,L.D.&Shaw,R.H.(1976).KilaueaLavaLakes:<br />

Natural laboratories for study <strong>of</strong> cooling, crystallization, and<br />

1964<br />

differentiation <strong>of</strong> basaltic magma. In: The geophysics <strong>of</strong> <strong>the</strong> Pacific<br />

Ocean basin and its margin: American Geophysical Union 19, 375^390.<br />

Yoder, H. S. (1973). Contemporaneous basaltic and rhyolitic magmas.<br />

American Mineralogist 58, 153^171.<br />

Zavala, K. (2005). Differentiation <strong>of</strong> silicic segregations in <strong>the</strong> <strong>Ferrar</strong><br />

<strong>Dolerite</strong>s, <strong>Antarctica</strong>: A geochemical study, PhD <strong>the</strong>sis, Johns Hopkins<br />

University, Baltimore, MD.<br />

Zavala, K. & Marsh, B. D. (2001). On <strong>the</strong> filling process forming<br />

silicic segregations. American Geophysical Union Spring Meeting,<br />

Washington D.C, abstract V41A-05.<br />

Downloaded from<br />

http://petrology.oxfordjournals.org/ by guest on April 30, 2013

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!