<<

J OURNAL OF Journal of Petrology, 2018, Vol. 59, No. 3, 517–558 doi: 10.1093/petrology/egy036 P ETROLOGY Advance Access Publication Date: 10 April 2018 Original Article

Melt Origin across a Rifted Continental Margin: a Case for Subduction-related Metasomatic Agents in the Lithospheric Source of Alkaline , NW , Kurt S. Panter1*, Paterno Castillo2, Susan Krans3, Chad Deering4, William McIntosh5, John W. Valley6, Kouki Kitajima6, Philip Kyle7, Stan Hart8 and Jerzy Blusztajn8

1Department of Geology, Bowling Green State University, Bowling Green, OH 43403, USA; 2Scripps Institute of Oceanography, University of California San Diego, La Jolla, CA 92093, USA; 3Department of Earth and Environmental Sciences, Michigan State University, East Lansing, MI 48824, USA; 4Department of Geology & Mining Engineering & Sciences, Michigan Tech University, Houghton, MI 49931, USA; 5New Mexico Bureau of Geology & Mineral Resources, Socorro, NM 87801, USA; 6Department of Geoscience, University of Wisconsin, Madison, WI 53706, USA; 7Department of Earth & Environmental Science, New Mexico Tech, Socorro, NM 87801, USA; 8Woods Hole Oceanographic Institution, Woods Hole, MA 02543, USA

*Corresponding author. Telephone: 01 419 372 7337. E-mail: [email protected]

Received August 4, 2017; Accepted March 23, 2018

ABSTRACT Alkaline magmatism associated with the West Antarctic rift system in the NW Ross Sea (NWRS) includes a north–south chain of complexes extending 260 km along the coast of Northern (NVL), numerous small volcanic seamounts located on the continental shelf and hundreds more within an 35 000 km2 area of the oceanic Adare Basin. New 40Ar/39Ar age dat- ing and geochemistry confirm that the seamounts are of –Pleistocene age and petrogeneti- cally akin to the mostly middle to late volcanism on the continent, as well as to a much broader region of diffuse alkaline volcanism that encompasses areas of , Zealandia and eastern Australia. All of these continental regions were contiguous prior to the late-stage breakup of Gondwana at 100 Ma, suggesting that the magmatism is interrelated, yet the mantle source and cause of melting remain controversial. The NWRS provides a rare opportunity to study cogenetic volcanism across the transition from continent to ocean and consequently offers a unique perspective from which to evaluate mantle processes and the roles of lithospheric and sub- lithospheric sources for mafic alkaline magmas. Mafic alkaline magmas with > 6 wt % MgO (, , , and tephrite) erupted across the transition from continent to ocean in the

NWRS show a remarkable systematic increase in silica-undersaturation, P2O5, Sr, Zr, Nb and light rare earth element (LREE) concentrations, as well as LREE/HREE (heavy REE) and Nb/Y ratios. Radiogenic isotopes also vary, with Nd and Pb isotopic compositions increasing and Sr isotopic compositions decreasing oceanward. These variations cannot be explained by shallow-level - al contamination or by changes in the degree of mantle partial melting, but are considered to be a function of the thickness and age of the mantle lithosphere. We propose that the isotopic signature of the most silica-undersaturated and incompatible element enriched best represent the 87 86 18 composition of the sub-lithospheric magma source with low Sr/ Sr (07030) and d Oolivine (50&), and high 143Nd/144Nd (05130) and 206Pb/204Pb (20). The isotopic ‘endmember’ signa- ture of the sub-lithospheric source is derived from recycled subducted materials and was trans- ferred to the lithospheric mantle by small-degree melts (carbonate-rich silicate liquids) to form amphibole-rich metasomes. Later melting of the metasomes produced silica-undersaturated

VC The Author(s) 2018. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: [email protected] 517

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 518 Journal of Petrology, 2018, Vol. 59, No. 3

liquids that reacted with the surrounding peridotite. This reaction occurred to a greater extent as the melt traversed through thicker and older lithosphere continentward. Ancient and/or more re- cent (550–100 Ma) subduction along the Pan-Pacific margin of Gondwana supplied the recycled subduction-related material to the asthenosphere. Melting and carbonate metasomatism were trig- gered during major episodes of extension beginning in the Late , but alkaline magma- tism was very limited in its extent. A significant delay of 30 to 20 Myr between extension and magmatism was probably controlled by conductive heating and the rate of thermal migration at the base of the lithosphere. Heating was facilitated by regional mantle upwelling, possibly driven by slab detachment and sinking into the lower mantle and/or by edge-driven mantle flow estab- lished at the boundary between the thinned lithosphere of the West Antarctic rift and the thick East Antarctic craton.

Key words: 40Ar/39Ar ages; geochemistry; Sr–Nd–Pb–O isotopes; West Antarctic rift system

INTRODUCTION sediment subducted and recycled deep within the man- The source and processes that generate alkaline mag- tle (Hofmann & White, 1982; Zindler & Hart, 1986; Weaver, 1991; Chauvel et al., 1992; Willbold & Stracke, mas in intraplate continental settings, particularly in 2010). Models proposed for West Antarctica include a young rifts such as East Africa and West Antarctica, re- mantle plume beneath the active Erebus volcano (Kyle main enigmatic. The debate revolves broadly around et al., 1992), a large plume that may have helped initiate whether melt is produced within the asthenospheric or and control the position of continental break-up in the lithospheric mantle, whether the melting and ultimately Late Cretaceous (Lanyon et al., 1993; Weaver et al., eruption is triggered by passive extension, lithospheric 1994; Storey et al., 1999) and older ‘fossil’ plumes pro- drip or upwelling plumes, and is further perpetuated by posed to have underplated and accreted to the base of the uncertainty in how melt is modified as it rises the Gondwana lithosphere prior to break-up (Rocholl through the continental plate. The Late mag- et al., 1995; Hart et al., 1997; Panter et al., 2000; Kipf matism found within the West Antarctic Rift System et al., 2014). For non-plume sources the enrichment has (WARS, Fig. 1a) has been explained by a variety of mod- been ascribed to mantle metasomatism caused either els that attempt to reconcile the petrology, geochemis- by small degrees of melting during the early phase of try and isotopic signatures of mafic alkaline rocks rifting prior to continental break-up (Rocchi et al., 2002; within the context of the last several hundred million Nardini et al., 2009) or by slab-derived fluids related to a years of Antarctica’s tectonic history, specifically the prolonged period of subduction that occurred prior to tectonic events leading up to and including the breakup rifting (Finn et al., 2005; Panter et al., 2006; Martin et al., of the proto-Pacific margin of Gondwana, which sepa- 2013, 2014; Aviado et al., 2015). rated Zealandia from West Antarctica in the Late In this study we offer a unique perspective on this Cretaceous. debate through interpretation of geochemical, isotopic A unifying model for WARS magmatism is war- (Sr, Nd, Pb and O), and age (40Ar/39Ar) data for basalts ranted given the relative uniformity of mafic alkaline erupted across the transition from oceanic lithosphere compositions emplaced within a very broad to thinned continental lithosphere adjacent to the East (1500 2500 km) region of thinned continental litho- Antarctic craton. The Neogene alkaline magmatism of sphere over the last 48 Myr. The mafic compositions are the NW Ross Sea (NWRS) include deposits on land as similar to ocean island basalts (OIB) and have HIMU-like well as numerous seamounts on the continental shelf 206 204 87 86 isotopic affinities ( Pb/ Pb > 195, Sr/ Sr < 07035, and hundreds more located within the oceanic Adare 143 144 Nd/ Nd > 05128). Furthermore, the magmatism Basin (Fig. 1b). Oceanic basalt from the NWRS have has been linked compositionally and tectonically to a been shown to be of the same age and petrogenetically much broader region of long-lived, diffuse, alkaline vol- akin to basalt from the rest of the WARS (Panter & canism that encompasses West Antarctica, Zealandia Castillo, 2007). Apart from the Cameroon volcanic line and eastern Australia (Finn et al., 2005). Existing models in West Africa (Fitton & Dunlop, 1985; Njome & de Wit, for WARS magmatism and beyond, however, are varied 2014) we know of no other place on Earth where con- but can be distilled into two fundamental types: those temporaneous, small-volume, alkaline volcanism that involve mantle plumes and those that do not. Both occurs across a rifted continental margin, and in par- types require mantle sources for the alkaline magmas ticular, a region that has experienced a relatively recent that have been enriched in incompatible minor and and rapid transition (110–80 Ma) from subduction to trace elements relative to primitive mantle to explain extension to break-up without voluminous magmatism their geochemical and radiogenic isotope composi- (i.e. no flood basalts; Finn et al., 2005). Thus the geo- tions. For the mantle plume source, enrichments are chronology and petrology of NWRS basalts allow for a considered to be via addition of oceanic crust and novel assessment of lithospheric (continental and

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 519

Fig. 1. Regional and sample location maps. (a) Subglacial surface elevation map (BEDMAP2; Fretwell et al., 2013) with the approxi- mate boundaries of the West Antarctic rift system delimited by red dashed lines. Also shown are locations of oceanic volcanic fields (seamounts and islands) that are compared in this study. MBS, Marie Byrd Seamounts. (b) Regional map of the NW Ross Sea (NWRS) showing the extent of Late Cenozoic volcanism above and below sea level. Faults and volcanic seamounts in the Adare Basin were mapped by Granot et al. (2010) using a combination of seismic and multibeam bathymetry data. The locations of the basalts examined in this study (solid color symbols) and from other studies (open symbols) are shown with symbol shapes corre- sponding to their major element compositional name as defined in Fig. 3. Volcanic deposits along the northern coastline of Victoria Land and islands are part of the Hallett volcanic province (Kyle, 1990). Here we defined NWRS basalts as those that are included in the Hallett volcanic province, the Malta Plateau of the Melbourne volcanic province (Kyle, 1990), as well as seamounts located on the continental shelf and within the Adare Basin. The bold dashed line delimits the approximate boundary between the East Antarctic craton and the West Antarctic rift. The bold dotted line at 1500 mbsl delimits the approximate boundary between rifted continental lithosphere and oceanic lithosphere. AP, ; PI, ; HP, Hallett Peninsula; DP, Daniel Peninsula; Cl, Coulman Island; MP, Malta Plateau.

oceanic) and sub-lithospheric sources for alkaline mag- Ordovician times (Borg & Stump, 1987; Boger & Miller, matism, as well as the potential overprint of these 2004; Di Vincenzo et al., 2016). Devonian and domains by subduction-related processes. Carboniferous calc-alkaline intrusions in NVL (Admiralty suite, 370–350 Ma; Stump, 1995) indicate younger subduction-related magmatism. All of this was TECTONIC AND MAGMATIC HISTORY part of the more extensive Terra Australis orogen that is The pre-Cenozoic fabric of Northern Victoria Land characterized by convergent deformation along the en- (NVL), which includes the continental portion of the tire 18 000 km pre-dispersal length of the Pan-Pacific NWRS (Fig. 1b), is expressed by three -bounded margin of the Gondwana supercontinent from the lithotectonic blocks (Wilson, Bowers and Robertson Bay Neoproterozoic to the end of the Paleozoic (Cawood, terranes) that were amalgamated via subduction during 2005). The Terra Australis orogen terminated with the the Ross orogeny in late Neoproterozoic to late assembly of Pangaea and heralded the beginning of the

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 520 Journal of Petrology, 2018, Vol. 59, No. 3

late Carboniferous to Triassic Gondwanide orogeny WARS. The Meander Intrusive Group consists of (Cawood, 2005), which is manifested in NVL by Eocene– (48–23 Ma) plutonic and subvolcanic deformed siliciclastic strata (Vaughan & Pankhurst, alkaline rocks found in NVL (Mu¨ ller et al., 1991; Tonarini 2008). Following a period of erosion that produced the et al., 1997). Rocchi et al. (2002, 2005) and Nardini et al. Kukri peneplain (Isbell, 1999), early to middle Jurassic (2009) proposed a multi-stage model for the origin of rifting and magmatism marked the initial stages of these rocks that calls upon upper mantle enrichment by Gondwana break-up and the separation of Africa and small-degree melts during Late Cretaceous rifting South America from the Dronning Maud Land margin (amagmatic phase). This was followed more than of Antarctica. Flood basalts and intrusions of the Ferrar 30 Myr later by melting of this enriched mantle by a large igneous province were emplaced in a narrow lin- combination of warm asthenosphere flow beneath the ear band over a distance of 3500 km within a very short rift towards the East Antarctic craton and localized de- time interval at 183 Ma (Elliot & Fleming, 2004; compression in response to the reactivation of NW–SE Burgess et al., 2015; Ivanov et al., 2017). The origin of trans-lithospheric faults by far-field plate-tectonic Ferrar magmatism remains controversial; some stresses (Rocchi et al., 2002). researchers relate it to mantle plume activity and long- By the middle Cenozoic WARS extension was distance transport of magma within the crust (Storey & focused in the Victoria Land Basin, which lies along the Kyle, 1997; Elliot et al., 1999; Vaughan & Storey, 2007) central and southern margin of the western Ross Sea, whereas others call for mantle melting fluxed by slab and the Adare Basin located in the north western Ross fluids from subduction along the length of the Sea (NWRS; Fig. 1b). The Adare Basin represents a fos- Gondwana margin (Hergt et al., 1991; Ivanov et al., sil spreading center, the slowest arm (ultraslow; 2017). 12 mm a–1 full-spreading rate) of a three-ridge junction The subduction of the Phoenix oceanic plate beneath between the Australia, East Antarctic and West the eastern margin of Gondwana ceased, abruptly, in Antarctic plates, that was active from 43 to 26 Ma the Cretaceous (Bradshaw, 1989; Mukasa & Dalziel, (chrons C20–C8) and resulted in 140–170 km of ENE– 2000; Tulloch et al., 2009) and the extension that formed WSW plate separation (Cande et al., 2000; Cande & the present-day WARS began. The sudden change from Stock, 2006; Granot et al., 2013). Post-spreading events a mostly compressional tectonic regime to a tensional were minor and occurred at 24 Ma and 17 Ma, result- one has been explained by the oblique subduction of ing in an additional 7 km of extension and the creation the Pacific–Phoenix spreading center (Bradshaw, 1989), of the Adare Trough by normal faulting (Granot et al., possibly with the aid of a mantle plume that weakened 2010, 2013). These faults display an en echelon NW–SE the lithosphere and controlled the location of rifting be- structural trend (Fig. 1b). A strong kinematic relation- tween Zealandia and the margin of ship between sea floor spreading and the WARS is sup- Antarctica (Weaver et al., 1994). The shutting down of ported by magnetic, seismic and structural trends that subduction has also been explained by the mechanism show continuity across the shelf break into the of slab capture (Luyendyk, 1995) and also by a collision Northern Basin (Cande & Stock, 2006; Davey et al., with a young and thermally buoyant Hikurangi oceanic 2006; Damaske et al., 2007; Ferraccioli et al., 2009; plateau (Davey et al., 2008). Whatever the cause, rifting Selvans et al., 2012). This indicates a continuity in crust- led to the final disintegration of Gondwana and the iso- al type that is most probably oceanic (Selvans et al., lation of Antarctica by 83 Ma (magnetic isochron C34; 2012, 2014; Granot et al., 2013). Granot et al. (2013) sug- Veevers, 2012, and references therein). It has been esti- gested that the shallow bathymetry of the Northern mated that this early rifting phase resulted in 500– Basin (Fig 1b) may be accommodated by thick deposits 1000 km of crustal extension across the entire Ross Sea of glacio-marine sediments on top of oceanic crust that between Marie Byrd Land and the western boundary of have filled the basin. the East Antarctica craton (DiVenere et al., 1994; Neogene alkaline volcanism along the northern Luyendyk et al., 1996). During the Late Cretaceous and coastline of Victoria Land and offshore islands com- early Cenozoic, north–south-oriented basins, including prises the Hallett volcanic province, which consists of the Northern Basin (Fig. 1b) and Victoria Land Basin, roughly north–south elongated shield volcanoes that and ridges (e.g. Coulman High) were formed (Cooper range in age from 14 Ma to 2Ma(Kyle, 1990; Mu¨ ller et al., 1987). et al., 1991; Armienti et al., 2003; Nardini et al., 2003; The early phase of broadly distributed extension in Mortimer et al., 2007; Smellie et al., 2011). Volcanism the WARS transformed into a more focused phase of on the adjacent continental shelf and Adare Basin was extension in the Paleogene (Huerta & Harry, 2007, and first identified through ocean bottom bathymetry references therein). This phase of rifting in the western (Panter & Castillo, 2007) and consists of hundreds of Ross Sea led to 300 km of extension between East and relatively small volcanic centers distributed over an West Antarctica between 80 and 40 Ma (Molnar et al., area of 35 000 km2. The trend of volcanism overlaps 1975; Cande et al., 2000), and was accompanied by with the roughly north–south pattern of normal faulting, rapid exhumation and uplift of the adjacent particularly the major bounding faults of the Adare Transantarctic Mountains (Fitzgerald et al., 1986) and Trough, and is generally aligned with volcanism on the oldest known igneous activity associated with the land in the Hallett volcanic province (Fig. 1b). The age of

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 521

seamount volcanism was first constrained by seismic phenocrysts (, clinopyroxene, amphibole and stratigraphic relationships. Granot et al. (2010) noted plagioclase) from each sample were mounted and sin- that all of the volcanic features are exposed above the gly polished on 1 cm diameter epoxy rounds. These seafloor surface, penetrating nearly the entire strati- were then secured into 1 inch brass mounts (three graphic sequence (to the late Pliocene) and that buried rounds per mount) and carbon coated for analysis by intrusions, which might indicate an older period of mag- electron microprobe. Major and minor element chemis- matism, are not observed along the 3200 km length of try was measured on a Cameca SX-100 Electron the multi-channel seismic data collected in the Adare Microprobe Analyzer at the University of Michigan Basin. The spatial distribution, relative age and prelim- EMAL laboratory using a 15 kV accelerating voltage inary petrological data all indicate that the volcanic ac- beam, 15 nA beam current, and 10–12 lm spot size. tivity in the oceanic Adare Basin is an extension of A total of 311 unknowns were analyzed: 174 olivine, 76 WARS volcanism. Based on this it was proposed that clinopyroxene, 23 amphibole, and 38 plagioclase both the continental and oceanic portions of the WARS grains. The following elemental routines were per- in the NWRS share a common origin (Panter & Castillo, formed for each mineral: Si, Cr, Fe, Mn, Mg, Ca, P, Ni 2007, 2008; Panter et al., 2016). for olivine; Si, Al, Ti, Cr, Fe, Mn, Mg, Ca, Na, K, P, Ni, Cl for clinopyroxene; Mg, Al, K, Ca, Ti, Cr, Mn, Fe, Na, P, Si, Sr, Ba for plagioclase; Mg, Al, K, Ca, Ti, Cr, Mn, Fe, SAMPLE DETAILS AND METHODS Na, P, Si, Cl, F for amphibole. Measurements (a total of This study is based on rock samples collected from sea- 475 unknowns) were taken primarily within the cores of floor dredging, rock and powder samples acquired from minerals (or presumed cores on fragments), and also at personal collections and rock samples provided by the rim and intermediate grain locations when compos- U.S. Polar Rock Repository, Byrd Polar and Climate itional or textural variations were observed (Krans, Research Center (http://research.bpcrc.osu.edu/rr/). 2013). Dredge samples were collected during a 38 day, National Science Foundation funded, marine geophys- Major and trace elements ical survey cruise (NBP0701) from December 2006 to Concentrations of major oxides and some of the trace January 2007 onboard the icebreaker RV/IB Nathaniel elements in whole-rock powders were determined by B. Palmer. The dredging was accomplished at a relative- X-ray fluorescence spectrometry (XRF) at the ly high level of precision by using the Palmer’s automat- GeoAnalytical Laboratory of Washington State ic dynamic global positioning system, depth sounding University (Castillo et al., 2010) and at New Mexico using a pinger attached to the wire at 300 m above the Tech (Hallett & Kyle, 1993). For some samples concen- dredge, and detailed seafloor bathymetry. trations of rare earth elements (REE) and other trace ele- The primary objective of this study is to evaluate ments (Rb, Sr, Y, Ba, Pb, Th, Zr, Nb, and Hf) on whole- mantle sources and processes responsible for alkaline rocks were determined by high-resolution inductively volcanism across the continental–oceanic transition in coupled plasma mass spectrometry (ICP-MS) using a the NWRS. To accomplish this we selected 30 relatively Finnigan Element 2 system at the Scripps Institution of unfractionated basalts (MgO > 6 wt %). Petrographic Oceanography (SIO) Analytical Facility, following the data along with mineral chemistry, major and trace ele- method of Janney & Castillo (1996) with some modifica- ments, Sr–Nd–Pb–O isotopes, and 40Ar/39Ar geochron- tions. Prior to ICP analysis, rock samples were crushed ology were obtained on subsets of these samples in an alumina ceramic jaw crusher. The resultant rock (Table 1). The location of the samples is indicated in chips were ultrasonically washed in deionized water for Fig. 1b and their whole-rock geochemistry and isotopic 30 min, dried in an oven overnight at 110C, and then compositions are presented in Tables 2 and 3. Twelve fresh-looking pieces were hand-picked under a binocu- of the 30 samples are from other studies (see Table 2 lar microscope. The selected chips were powdered in for details and references) and of those 12, four (MA- an alumina ceramic grinder. For each sample, about 009a, MA-117, P74794, P74833) were analyzed in this 25 mg of powder was digested with an ultrapure 2:1

study for mineral chemistry and oxygen isotopes on concentrated HF–HNO3 solution in a Teflon beaker, and olivine separates (Table 3) and two of those (MA-009a, the mixture was placed on a hot plate (60C) and dried

MA-117) were also analyzed for Sr–Nd–Pb isotopes on under a heat lamp. About 2 ml of ultrapure 12N HNO3 whole-rock powders (Table 2). An additional sample, was twice added to the digested sample and evapo- SAX20 (Nardini et al., 2009), is not from the NWRS but rated to dryness. After drying, the digested sample was

from a location to the south within the Melbourne vol- diluted 4000-fold with 2% HNO3 solution containing canic province (Kyle, 1990) and is included for 1 ppb In as an internal standard. The instrument drift comparison. was monitored and corrected by measuring an in- house rock standard as an unknown within the run. For Mineral chemistry selected samples some trace elements for whole-rocks Rocks were crushed and sieved to separate phenocrysts were analyzed by instrumental neutron activation at between 250 lm and 2 mm in diameter. Six to eight New Mexico Tech (Hallett & Kyle, 1993).

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 on 04 September 2018 by University ofWisconsin-Madison Libraries user Downloaded fromhttps://academic.oup.com/petrology/article-abstract/59/3/517/4966906 522

Table 1: Summary of 40 Ar/39 Ar results and analytical methods

40 36 39 39 Sample Locality Latitude Longitude Rock type Irrad. # Material Age analysis Integrated 2r n Isochron 2r n MSWD Ar/ Ar 2r Plateau 2r n MSWD Ar ArK K/Ca 2r Preferred Method age (Ma) age (Ma) age (Ma) (%) age (Ma)

Adare seamounts D2-1 trough W scarp 69861 171844 hawaiite NM-217 GM conc. furnace step-heat 1593 076 10 494011 1593 6 076 integrated D3-1 trough W scarp 70000 171990 hawaiite NM-217 GM conc. furnace step-heat 493 066 10 456 027 10 163 296223461 019 10 152 1000 1000028 036 461 6 019 plateau D4-1 trough center 70070 172316 basanite NM-251 GM conc. laser step-heat 344 008 11 376 013 11 403 293115335 011 4 037 599 1379053 102 335 6 711 plateau D4-3 trough center 70070 172316 basanite NM-251 GM conc. laser step-heat 245 002 11 251 009 8 794 3110160261 002 5 165 5036711203251 6 109 isochron D7-1 continental shelf 71820 171900 tephrite NM-217 GM conc. furnace step-heat 047 073 10 029 015 10 614 295935032 023 7 695 917957035 036 032 6 523 plateau D9-1 S basin 71628 172523 tephrite NM-217 GM conc. furnace step-heat 291 030 10 289 016 10 1582 293168286 014 10 1468 1000 16792 0925 115 286 6 714 plateau D12-1 S basin 71631 172634 basanite NM-217 GM conc. furnace step-heat 352 050 10 287 007 9 805 298010289 006 9 627 998 1495035 039 289 6 906 plateau D15-1 S basin 71579 172733 tephrite NM-217 GM conc. furnace step-heat 312 022 9 282 004 2 082 553 1875030 047 312 6 722 integrated D16-1 S basin 71597 172821 tephrite NM-217 GM conc. furnace step-heat 297 056 11 266 019 11 125 296919276 013 9 144 992 1582027 031 276 6 813 plateau D17-1 S basin 71499 173486 NM-217 GM conc. furnace step-heat 020 007 10 014 001 9 181 295240014 002 9 158 999 2682118149014 6 802 plateau Possession Islands A227B McCormick Is. 71842 170976 hawaiite NM-251 GM conc. laser step-heat 0168 0031 11 0058 0029 8 205 2955160053 0031 8 254 956 1632030 034 0053 6 3031 plateau HC-SPI-4 (3873)* Possession 71885 171167 tephrite NM-251 GM conc. laser step-heat 010 004 11 017 002 11 698 294321016 002 5 054 776 1075030 031 016 6 702 plateau NV-3C (5166)* Possession 71885 171167 benmorite NM-251 GM conc. laser step-heat 140 003 10 141 001 8 064 298260142 008019 856 29342321142 6 301 plateau NV-4C (5171)* Foyn 71950 171083 hawaiite NM-251 GM conc. laser step-heat 025 014 11 049 032 10 212 294639033 018071 983 1026021 022 033 6 210 plateau A225A Foyn 71957 171093 basanite NM-251 GM conc. laser step-heat 026 003 11 0076 0042 10 300 301324024 003 3 085 665828016 030 024 6 003 plateau

*Sample provided by the Polar Rock Repository—PRR-# (http://research.bpcrc.osu.edu/rr/). MSWD, mean square weighted deviation. Integrated age calculated by summing isotopic measurements of all steps. Integrated age error calculated by quadratically combining errors of iso- topic measurements of all steps. Plateau age is inverse-variance-weighted mean of selected steps. Plateau age error is inverse-variance-weighted mean error (Taylor, 1982) times root MSWD where MSWD > 1. Decay constants and isotopic abundances after Steiger & Ja¨ ger (1977). Electron multiplier sensitivity for MAP 215-50 averaged 854e – 17 moles pA–1 for furnace analyses and 437e – 17 for laser analyses. Laser analyses sensitivity on Argus VI averaged 984e – 17 moles fA–1. MAP 215-50 total system blank and background for groundmass concentrate analyses averaged 5850, 710, 521, 224, 204 10–18 moles at masses 40, 39, 38, 37 and 36, respectively. MAP 215-50 total system blank and background for K-feldspar analyses averaged 409, 252, 1 –18 10, 249, 775 10 moles at masses 40, 39, 38, 37 and 36, respectively. Argus VI total system blank and background for groundmass concentrate analyses averaged 358, 28, 019, 050, 1 Petrology of Journal –18 3 10 moles on masses 40, 39, 38, 37 and 36, respectively. J-factors determined by CO2 laser-fusion of six single crystals from each of 10 radial positions around the irradiation tray. 40 39 36 37 Correction factors for interfering nuclear reactions were determined using K-glass and CaF2 and are as follows: ( Ar/ Ar)K ¼ 00000 6 00004; ( Ar/ Ar)Ca ¼ 000028 6 000001; 39 37 ( Ar/ Ar)Ca ¼ 000068 6 000005. 08 o.5,N.3 No. 59, Vol. 2018, , Journal of Petrology, 2018, Vol. 59, No. 3 523

Table 2: Geochemical and radiogenic isotope (Sr, Nd, Pb) analyses of NWRS basalts

Sample no.: SAX20 MA009a MA-117 MP24 A223 MP8 HC-DP-3 MP34 MP32 P74794 P74833 A210B PRR-#: 3839 Dec. latitude S: –73770 –73040 –73020 –73063 –72872 –72830 –72667 –73490 –73490 –71668 –71653 –72023 Dec. longitude E: 165950 167470 167630 167750 168970 169098 169633 169580 169580 170080 170120 170160 Geographical Greene Malta Malta Malta Wof Wof Daniell Coulman Coulman Robertson Robertson SW of location: Point, Plateau Plateau Plateau Daniell Daniell Peninsula Island Island Bay Bay Cape MVP Peninsula Pennisula Roget

Rock name (TAS): TEPH AKB AKB BAS AKB AKB AKB BAS BAS AKB AKB AKB

Major elements (wt %) SiO2 4161 4989 4748 4379 4631 4468 4516 4320 4432 4771 4895 4565 TiO2 369 245 266 283 224 274 319 311 342 249 210 235 Al2O3 1205 1360 1438 1402 1448 1506 1412 1492 1587 1388 1615 1342 FeOt 1479 1115 1158 1047 1186 1291 1168 1178 1206 1132 1017 1060 MnO 029 016 019 018 019 020 019 019 020 017 016 019 MgO 952 935 849 1264 949 963 911 1124 844 941 804 1308 CaO 942 996 1098 1115 1122 1103 1294 1109 1052 1126 1023 983 Na2O542 223 266 302 275 255 225 283 336 245 283 316 K2O157 089 116 132 095 080 085 104 112 094 100 117 P2O5 163 032 043 057 050 039 052 061 068 038 037 055 LOI 090 082 044 084 073 244 058 060 054 -048 Total 10000 10082 10045 10000 9999 10000 10001 10000 10000 10055 10000 10000 Mg# 534599567683588571582630555597585688 CIPW norm minerals (wt %) Ne 23724125505550103921186 Lc Di 265187228249230205286212188240163212 Hy 177 15 Ol 23796199248225245184251210210190277 Trace elements (ppm) Sc 17417 3233353282632125 V 130 254 302 317 259 180 Cr 369 272 349 7 465 389 713 382 477 304 479 Ni 86 93 116 150 99 Cu 22 42 81 69 28 Zn 108 108 83 88 74 Ga 17 20 21 Rb 72 21 30 38 39 20 21 24 32 20522926 Sr 1494 461 531 640 608 552 657 667 867 477 585 592 Y 49243323409242682530239218264 Zr 500 172 200 194 225 158 217 204 236 167 160 397 Nb 18322852706424403576675853413401677 Cs 031 015 012 054 Ba 830 229 337 431 256 251 273 329 399 236 252 418 La 12512436 3271 4053277 293392398518298294360 Ce 5096681 86 6905 65 780 81 109 6115838299 Pr 28 639 8 102932 8 10 13 754 701 1090 Nd 10432592 3201 39 3609 33 4004035123032783998 Sm 189547 708 73759 69779966581 754 Eu 57177 213 25246 242632223 195 239 Gd 177509 676 77738310 616 536 Tb 224 082 1 101 116 101 106 125 094 082 110 Dy 118454 567 51634 555565522 467 519 Ho 193 0811089 116 099 101 118 098 087 090 Er 496 207 306 225 338 258 258 312 243 222 264 Tm 061 029 04029 035 033 039 032 029 Yb 376 173 255 187 277 205 2124185 174 211 Lu 057 037 037 026 027 03029 033 027 027 026 Hf 128462 486 469 499 56432 399 748 Ta 1025 417 270 257 391 478 267 256 390 Pb 5 22201 171421243 187 213 245 Th 1665 589 501 343 50522 605 378 402 507 U493 163 091 102 26146 182 096 092 168 Radiogenic isotopes (measured values) 87Sr/86Sr 070284 0705035 0703786 070316 0703424 0703357 0702854 0703458 143Nd/144Nd 0512979 0512748 0512824 0512876 0512883 0512902 0512891 206Pb/204Pb 195683 19222 1932 196465 19703 196407 197314 19363 207Pb/204Pb 155399 15625 1562 155665 15599 155169 155955 15615 208Pb/204Pb 389781 39327 3927 392801 39387 392011 392857 39202 (continued)

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 524 Journal of Petrology, 2018, Vol. 59, No. 3

Table 2: Continued

Sample no.: A240B-1 A232B NV-6E VC-21 A227B AW82205 HC-FI-2A NV-4 NV-4C NV-4B A225A HC-SPI-2 PRR-#: 3842 5169 5171 3872 Dec. latitude S: –71302 –71703 –71619 –71694 –71842 –71950 –71950 –71950 –71950 –71955 –71957 –71885 Dec. longitude E: 170210 170298 170556 170557 170976 171083 171083 171083 171083 171092 171093 171167 Geographical location: Adare Adare McCormick Foyn Foyn Foyn Foyn Foyn Foyn Possession Adare Peninsula Peninsula Peninsula Island Island Island Island Island Island Island Island

Rock name (TAS): BAS HAW BAS BAS HAW AKB AKB AKB HAW AKB BAS BAS

Major elements (wt %) SiO2 4253 4687 4272 4494 4692 4557 4546 4573 4564 4506 4546 4148 TiO2 380 287 386 324 283 296 301 297 313 280 315 353 Al2O3 1454 1543 1367 1456 1573 1474 1458 1455 1539 1342 1550 1308 FeOt 1266 1154 1305 1286 1059 1143 1171 1139 1132 1173 1122 1220 MnO 021 021 017 018 021 019 019 019 019 019 020 022 MgO 872 714 1002 707 736 923 933 945 784 1123 774 956 CaO 1195 984 1098 1114 992 1071 1062 1065 1055 1089 1058 1279 Na2O347 375 384 376 430 350 345 355 395 313 406 459 K2O142 161 094 154 154 106 105 093 131 099 137 141 P2O5 069 073 074 070 060 061 061 060 067 057 071 115 LOI 082 038 270 049 056 070 Total 9999 9999 10000 10000 10000 10000 10001 10001 9999 10000 9999 10001 Mg# 551525578495553590587597553631552583 CIPW norm minerals (wt %) Ne 1598015012711491898711390123210 Lc 65 Di 286194265269216227225228227250229319 Hy Ol 183184218169168207213210178241174184 Trace elements (ppm) Sc 30 25 22 25 25 25 24 23 27 V 312 235 235 240 278 251 250 253 209 263 255 Cr 305 260 225 218 230 431 367 388 269 395 238 332 Ni 115 99 435 373 177 174 173 119 411 117 156 Cu 72 57 226 206 60 55 55 54 101 58 65 Zn 95 100 75 80 91 97 92 96 78 94 109 Ga 21 21 21 19 19 21 21 20 Rb 32 32 21 75 37 26 24 17 33 16 36 44 Sr 800 787 800 695 812 684 672 642 761 592 831 1034 Y350350303030027526229027330337 Zr 263 276 225 311 223 224 227 269 203 280 331 Nb 830700586871064162177853820 1076 Cs 040 069 047 029 033 Ba 422 616 323 419 527 319 311 300 379 269 396 564 La 5250 5180 43 49 6400 4220 432379503395420 739 Ce 10360 10520 78 84 12520 8500 846802 1014 70 10780 1389 Pr 103106 88 Nd 5000 4590 44 44 5110 4050 401408454364640 613 Sm 995 966 9791997 818 76918 Eu 304 344 2928311 258 23282 Gd 9 9 7 Tb 121 113 1212116 101 1 108 Dy 5957 51 Ho 1 1 08 Er 2625 23 Tm Yb 232 233 2424263 205 21219 Lu 033 030 0304035 026 03034 Hf 673 644 5653764 541 46647 Ta 525 424 4451603 451 38519 Pb 200 500 1719430 200 0911311 200 29 Th 611 553 4664777 437 53556646664 91 U170 200 1 23211 090 12062209204 26 Radiogenic isotopes (measured values) 87Sr/86Sr 0702944 0703741 0702900 0702900 0702780 0702970 0702870 143Nd/144Nd 0512930 0512874 0512937 0512945 0513000 0512935 0513000 206Pb/204Pb 20026 19432 20180 20184 20044 207Pb/204Pb 15633 15642 15680 15653 15587 208Pb/204Pb 39504 39285 39794 39703 39438 (continued)

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 525

Table 2: Continued

Sample no.: AW82214 NV-3B D2-1 D4-1 D4-3 D9-1 D12-1 PRR-#: Dec. latitude S: –71900 –71900 –69861 –70070 –70070 –71628 –71631 Dec. longitude E: 171167 171200 171844 172316 172316 172523 172634 Geographical Possession Possession Adare Adare Adare S Adare S Adare location: Island Island Trough, W scarp Trough, center Trough, center Basin Basin

Rock name (TAS): BAS BAS HAW BAS BAS TEPH BAS

Major elements (wt %) SiO2 4164 4478 4675 4307 4198 4526 4460 TiO2 347 360 263 299 475 318 306 Al2O3 1307 1582 1718 1292 1359 1518 1334 FeOt 1223 1293 785 1020 1335 1148 1147 MnO 022 022 015 020 027 026 021 MgO 978 645 883 1332 986 639 1070 CaO 1260 1037 1043 1084 1041 942 1002 Na2O429 365 343 364 315 558 412 K2O152 121 173 188 174 216 154 P2O5 117 096 101 094 090 109 093 LOI 368 179 122 173 117 Total 9999 10000 9999 10000 10000 10000 9999 Mg# 588471667700568498625 CIPW norm minerals (wt %) Ne 1978574167130211148 Lc 7044 Di 321183152276229245247 Hy Ol 188186181246223148226 Trace elements (ppm) Sc 27 25 27 30 16 22 V 296 215 197 233 325 185 214 Cr 375 35 277 540 244 135 355 Ni 167 296 162 338 139 71 205 Cu 30 101 40 46 42 25 46 Zn 101 87 71 81 115 136 102 Ga 21 17 16 23 21 19 Rb 47 20 34 31 37 40 35 Sr 1034 925 901 861 1131 1216 1007 Y38028350265355337303 Zr 332 217 256 292 394 553 347 Nb 118061751936962 1550894 Cs 054 Ba 626 371 604 467 595 558 401 La 7380 47 5906 5101 7088858 5648 Ce 14260 83 10527 9553 1524 16712 11197 Pr 1071280 1029 1706 1205 Nd 6150 46 4644 4058 6946393 4843 Sm 1201 96845 723 1077 911 Eu 368 31280 245 335 290 Gd 9 1475 2444 1803 Tb 139 12109 111 145 130 Dy 56610 490 638 574 Ho 09109 090 114 103 Er 24305 191 250 214 Tm 042 031 041 034 Yb 237 22287 172 235 194 Lu 031 03039 028 034 028 Hf 757 52 Ta 689 47 Pb 200 16283 252 130344 201 Th 833 55726 533 90677 481 U280 1 175 141 47220 133 Radiogenic isotopes (measured values) 87Sr/86Sr 0702884 070343 0702870 0702900 0702779 143Nd/144Nd 0512945 051282 0513000 0512960 0513001 206Pb/204Pb 19851 20028 19237 20234 20111 207Pb/204Pb 15620 15658 15583 15653 15625 208Pb/204Pb 39354 39916 38812 39715 39551

Sample numbers in bold are from this study. PRR, samples on loan from the Polar Rock Respository (http://research.bpcrc.osu.edu/ rr/). Rock name based on total alkali vs silica diagram (TAS) shown in Fig. 2. FeOt is total iron expressed as FeO, normalized to 100%. LOI, loss on ignition. Mg# ¼ 100Mg/(Mg þ Fe2þ). Major elements (wt %) by XRF and trace elements (ppm) by XRF and ICP- MS. CIPW normative (wt %) calculated following Irving & Baragar (1971). Samples SAX20, MP24, MP8, MP34, MP32 are from Nardini et al. (2009). Samples MA009a and MA-117 are from Rocholl et al. (1995) with Sr, Nd and Pb isotopes from this study itali- cized. Samples P74794 and P74833 are from Mortimer et al. (2007). Samples NV-6E, NV-21, NV-4B and NV-3B from Aviado et al. (2015). Samples A223, A210B, A240B-1, A232B, A227B and A225A were collected and described by Hamilton (1972) and re-analyzed in this study.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 526 Journal of Petrology, 2018, Vol. 59, No. 3

Table 3: Oxygen isotope data from olivine phenocrysts

1 2 Sample Location Fo SIMS 2SD SEM na Core* 2SD SEM na LF 2SD MgO (wr) d18O(&) d18O(&) d18O(&) (wt %)

HC-DP-3 (3839) Daniel Pen. 84 6 2489 022 003 5 911 HC-FI-2A (3842) Foyn Is. 81 6 5506 023 008 7 488 020 933 HC-SPI-2 (3872) Possession Is. 75 6 9523 028 007 7 490 012 956 HC-DP-1 (3873) Daniel Pen. 81 6 5527 024 007 6 490 451 NV-4 (5169) Foyn Is. 83 6 2482 018 007 7 487 027 006 7 486 022 945 NV-4C (5171) Foyn Is. 82 6 5471 022 006 7 462 021 005 6 784 A214B W of Daniel P. 76 6 7510 016 005 4 410 A223 W of Daniel P. 87 6 3525 023 012 5 503 949 A225A Foyn Is. 81 6 3515 024 008 8 556 015 003 6 774 A227B McCormick Is. 82 6 5508 023 004 9 736 A232B Adare Pen. 81 6 4530 020 008 6 506 714 A240B-1 Cape Adare 82 6 5520 021 007 6 872 D12-1 S Adare Basin 84 6 2535 039 007 9 1070 D2-1 Adare Trough 87 6 1525 041 005 4 883 D4-1 Adare Trough 90 6 1504 013 010 5 525 1332 D4-3 Adare Trough 79 6 1543 023 011 4 550 026 005 7 517 014 986 D9-1 S Adare Basin 79 6 1505 039 010 5 639 MA-009a Malta Plateau 78 6 7544 021 008 8 573 026 006 6 935 MA-117 Malta Plateau 79 6 7539 023 011 7 578 018 004 6 526 012 849 P74794 Robertson Bay 80 6 7533 041 003 6 941 P74833 Robertson Bay 70 6 7515 024 011 7 540 013 003 6 804

1Numbers in parenthesis are PPR-# of samples on loan from Polar Rock Repository (http://research.bpcrc.osu.edu/rr/). 2Fo (%) ¼ 100Mg/(Mg þ Fe2þ) with 6 1r. SIMS data are averages of multiple spot analyses (na) on 2–3 grains. Precision to 2 standard deviations (2SD) on standard San Carlos olivine is 6028&. Core* is the average calculated for multiple SIMS spot analyses (na) within the core of a single olivine grain. Laser fluorination (LF) data are based on 4–5 olivine grains. Standard deviation precision (2SD) based on repeat analysis. Precision on standard UWG-2 (Core Mountain Garnet) is 6010& (2r). SEM, standard error of mean. (wr), whole-rock MgO content (wt %).

Radiogenic isotopes (Sr, Nd, Pb) chemistry for whole-rock powders was determined with Radiogenic isotopes were analyzed at the Scripps Sr-Spec and Ln-Spec resin, respectively. Lead was sep-

Institution of Oceanography (SIO) and the Woods Hole arated following the HBr–HNO3 procedure of Oceanographic Institution (WHOI). At SIO whole-rock Abouchami et al. (1999) using a single column pass. Sr, powders were dissolved following the dissolution pro- Nd and Pb analyses were carried out on the NEPTUNE cedure described above for the ICP-MS analyses. Lead multi-collector ICP-MS system. For Sr and Nd, the in- was first separated from the dissolved samples using ternal precision is 10–20 ppm (2r); external precision, small ion exchange columns in an HBr medium. after adjusting to 0710240 and 0511847 for the Residues from Pb extraction were collected and then SRM987 and La Jolla Nd standards respectively, is esti- passed through ion-exchange columns using HCl as mated to be 15–25 ppm (2r). Pb analyses carry internal eluent to collect Sr and REE. Finally, Nd was separated precisions on 206, 207, 208/204 ratios of 15–50 ppm; by passing the REE cuts through small ion exchange SRM997 Tl was used as an internal standard, and exter- columns using alpha hydroxyisobutyric acid as eluent. nal reproducibility (including full chemistry) ranges The Pb, Sr and Nd isotopes were measured by thermal from 20 ppm (2r) for 207Pb/206Pb, to 120 ppm (2r) for ionization mass spectrometry (TIMS) using a nine-col- 208Pb/204Pb. lector, Micromass Sector 54 system at SIO. Lead iso- topes were fractionation corrected using the isotope Oxygen isotopes values of NBS 981 relative to those of Todt et al. (1996). Oxygen isotopes on olivine separates were analyzed by Strontium isotopes were fractionation corrected to traditional laser fluorination/gas source mass spectrom- 86Sr/88Sr ¼ 01194 and are reported relative to NBS 987 etry (LF) and also in situ by secondary ionization mass 87Sr/86Sr ¼ 070258. Neodymium isotopes were meas- spectrometry (SIMS) at the University of Wisconsin– ured in oxide forms, fractionation corrected to Madison (WiscSIMS) Laboratory (Table 3). In both 146NdO/144NdO ¼ 072225 (146Nd/144Nd ¼ 07219) and cases, the composition of olivine was first determined are reported relative to 143Nd/144Nd ¼ 0511850 for the by electron microprobe analysis (EMPA), specifically La Jolla Nd standard. Analytical uncertainties based on the forsterite content, which was needed to produce a repeated measurements of standards are 60010 for calibration curve to correct for SIMS bias in measured 206Pb/204Pb and 207Pb/204Pb, 60024 for 208Pb/204Pb, oxygen values relative to SCOL (see Valley & Kita, 60000018 for 87Sr/86Sr, and 60000014 for 143Nd/144Nd. 2009). chosen for analysis by SIMS contain Routine analytical blanks are generally <35 pg for Sr, varying percentages of melt and oxide inclusions, <10 pg for Nd and <60 pg for Pb. At WHOI Sr and Nd which unlike LF, are easily avoided by the SIMS

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 527

technique (10 lm diameter spot size). Furthermore, calculated from 3–4 analyses of the running standard the small number of olivine grains separated owing to (San Carlos olivine) before and after analyses of the limited amount of material for some of the samples unknowns to yield an average standard precision of required the use of single crystal SIMS analysis. 6028& (2SD). Eight unknowns representing the most 18 For LF, olivine was analyzed in 16–26 mg aliquots extreme d Ool values were also selected and an add- (4–5 grains) using a CO2 laser probe system (BrF5 re- itional 5–6 analyses were taken from cores to test intra- agent) attached to a dual-inlet five-collector Finnigan/ grain reproducibility. The Core* value reported in MAT 251 mass spectrometer. In cases where a given Table 3 is the average of multiple spots in the core of a sample yielded two size populations of phenocrysts, ali- single olivine grain in an individual sample. quots were grouped comparatively by size. Measurements for unknowns were corrected by the 40Ar/39Ar dating average difference between measured and accepted Selected basalt samples dredged from seamounts and values for UWG-2 Gore Mountain garnet standard a basalt from the Possession Islands were crushed, 18 (d OVSMOW ¼ 580&, Valley et al., 1995) measured on sieved, leached with dilute HCl, washed in distilled the same day. Daily standard deviation for UWG-2 was water and hand-picked to remove phenocrysts and any 6008–010& (2SD, n ¼ 8). Duplicates were attempted altered material to produce a groundmass concentrate. for all runs but were obtained on only five of the nine The samples were irradiated in two batches. Samples samples analyzed (Table 3) owing to sample size and for batch NM-217 were loaded into a machined Al disc number of highest quality olivines available. Duplicate and irradiated for 7 h in the D-3 position, Nuclear measurements of unknowns had an average reproduci- Science Center, College Station, TX. Samples for batch bility of 60 16& for olivine (2SD). NM-251 were irradiated for 20 h in the C.T. position, US The best olivine grains (i.e. pristine cores with few to Geological Survey TRIGA, Denver, CO. Fish Canyon no inclusions and no alteration) from each sample were sanidine (FC-2), with an assigned age of 2802 Ma also selected for SIMS. Mineral compositions deter- (Renne et al., 1998), was used as the neutron flux moni- mined by EMPA are within 15 mm of SIMS analyses, but tor in both batches. within 5 mm of pits. Olivine grains were sorted based on Samples in batch NM-217 were measured using a relative size and cast in three 25 cm mounts. Unknowns Mass Analyzer Products 215-50 mass spectrometer on and grains of the San Carlos olivine standard were line with automated all-metal extraction system. placed within 5 mm of the center of each mount. Groundmass concentrates were step-heated using a Mounts were ground and polished to minimize grain Mo double-vacuum resistance furnace. Reactive gases topography to less than 1 mm(Kita et al., 2009). were removed during a 12 min reaction with two SAES Reflected light and back-scattered electron images were GP-50 getters, one operated at 450C and the other at taken of each grain prior to analysis to assess compos- itional zoning and select locations on individual grains 20 C. Gas was also exposed to a W filament operated at to be measured in situ by electron microprobe and 2000 C and a cold finger operated at –140 C. Samples SIMS. Over 100 olivine phenocrysts from 21 samples in batch NM-251 were measured using an Argus VI (2–3 grains per sample; Table 3) were measured follow- mass spectrometer on line with an automated all-metal ing the procedures of Kita et al. (2009) and Valley & Kita extraction system. Groundmass concentrates were (2009). A CAMECA IMS-1280 large-radius, multi- step-heated by a 50 W Synrad CO2 laser. Reactive gases collection SIMS system was used with a primary 133Csþ were removed during a 2 min reaction with two SAES beam intensity of 15–18 nA (spot size 10 lm). The GP-50 getters. Getters, W filament and cold finger were secondary ion beams for 16O, 16O1H, and 18O were col- operated at the same temperatures as batch NM-217. lected with Faraday cup detectors at MRP 2500 (mass- Electron multiplier sensitivities, blanks and other analyt- resolving power). The OH/O ratio is a relative measure ical parameters for both batches are reported in of the amount of hydrogen in the volume of olivine ana- Table 1. lyzed. Ratios of OH/O are background corrected by sub- tracting the OH/O in bracketing analyses of SCOL, to RESULTS remove variability owing to sample and vacuum condi- tions and provide a sensitive check against cryptic hy- Field data—Adare Basin seamounts drous alteration or contamination (Wang et al., 2014). The NBP0701 dredging campaign produced a total of Given the wide range of olivine compositions (Fo-70 to 1000 lb of rock from 17 sites within the Ross Sea Fo-90; Table 3) a calibration curve based on multiple (Supplementary Data Table A1 and Figs A1–A9). Ice- olivine standards with Mg# from 100 to 1 (HN-OL, rafted debris (IRD) was recovered from all dredge sites, SCOL, OR-OL, Fayalite 50278; Supplementary Data but a total of 48 in situ mostly basaltic samples Table A3; supplementary data are available for down- were collected from the western flank of the Adare loading at http://www.petrology.oxfordjournals.org) Trough and individual seamounts within the basin and was used to correct for bias in the oxygen isotopic val- on the continental shelf. The lava samples were initially ues of the unknowns relative to the San Carlos olivine distinguished shipboard from IRD by their angularity, standard (Valley & Kita, 2009). A bracketed average was absence of manganese coatings and glassy texture.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 528 Journal of Petrology, 2018, Vol. 59, No. 3

Further examination in thin section revealed a predom- young. Radial intergrowths (i.e. variolitic texture) of inance of textures indicative of rapid undercooling plagioclase and/or in the glassy groundmass (described below) as would be expected for lava in many of the indicate rapid undercooling, as erupted on the seafloor. The main strategy was to col- would be expected with eruption on the seafloor. In this lect lava from these seamounts with increasing distance study, basalt from eight seamounts and basalt from the from the continent. Individual seamounts are steep- western scarp of the Adare Trough have been dated by sided, conical-shaped edifices that range in height and the 40Ar/39Ar method to determine the age of volcanism basal diameter from 100 m to >700 m and from <1km and to provide further temporal constraints on the to 4 km, respectively, with minimum exposed vol- Neogene history of the Adare Basin. Furthermore, the 3 umes between 003 and 3 km . In some areas, closely age determinations are compared with the age of near- spaced seamounts are merged on elevated platforms by onshore deposits to resolve the progression of mag- (Supplementary Data Figs A7 and A8) that were most matism in the NWRS region of the WARS. probably constructed by the coalescence of lava flows. The morphological characteristics of the seamounts are 40Ar/39Ar chronology consistent with monogenetic volcanoes. New 40Ar/39Ar age dating results for four basalt samples Four sites were dredged in the central portion of the dredged from the Adare Trough, five basalts from five Adare Trough. Samples from three of the sites (D1, D2 different seamounts in the southern Adare Basin, and and D3) on the western flank at 1000 to < 2000 m one basalt sample from a seamount located on the con- below sea level (mbsl) include basaltic boulders and tinental shelf, provide the first direct dating of oceanic IRD such as dolerite and metamorphic rocks; all of the volcanism in the NWRS. Additionally, five age determi- samples are coated with a thick layer of manganese nations on basalt collected from the Possession Islands oxide. Dredge D3 sampled a seamount exposed in the (two from Possession, two from Foyn and one from east-facing upper western scarp that appears to have McCormick) provide the first age data for this island been cut by the bounding normal fault of the trough group. A drop stone recovered at D17 was also (Supplementary Data Fig. A2). The other dredge lines dated to determine the age and provenance of the pyro- on the western flank (D1 and D2) were positioned on clastic activity that produced it. The results and analytic- the upper scarp facing away from any obvious sea- al details are provided in Table 1 and representative age mounts in an attempt to sample oceanic crust release and Ca/K spectra are shown in Fig. 2. The re- (Supplementary Data Fig. A1). The fourth dredging site mainder of the spectra and isotope correlation (D4) is a volcano duplet located in the middle of the (36Ar/40Ar vs 39Ar/40Ar) diagrams are provided in Adare Trough at > 2000 mbsl (Supplementary Data Fig. Supplementary Data Fig. A10 and all 40Ar/39Ar analytical A3) and the samples acquired were mainly angular frag- data are provided in Supplementary Data Table A2. ments of vesicular basalt and some pebbles and cob- The new age data confirm the relatively young his- bles of IRD. The manganese oxide coating on these tory of seamount volcanism in the NWRS. The ages samples is very thin when compared with the samples range from 15 93 6 0 76 Ma to 0 14 6 0 02 Ma (Table 1), from the western flank. A cluster of more than 100 sea- mounts was found in the southern portion of the Adare with a median value of 288 Ma. All of the seamounts Basin and on the shelf. Thirteen of these seamounts dated are younger than 5 Ma, verifying the Pliocene age were dredged. The shallowest ones (D5 to D7) on the of the volcanism that was previously based on seismic continental shelf, 15 nautical miles away from Cape stratigraphy (Granot et al., 2010), but further reveal Adare, are 400 to 600 mbsl (Supplementary Data Figs Pleistocene activity (26 Ma) at three seamounts (D4-3, A4 and A5) and heavily covered with coralline materi- D7-1 and D17-1). The middle Miocene age (16 Ma) als. The farthest seamount (D17), some 65 nautical determined on basalt sample D2-1 collected on the miles from Cape Adare, is 2000 to 1400 mbsl western scarp face of the Adare Trough is too young to (Supplementary Data Fig. A9), is free of coralline represent oceanic crust, which in this area has ages be- materials. tween 335 and 309 Ma (bounded by magnetic anoma- The seamounts were erupted through a thick (up to lies 12o and 13o; Granot et al., 2010). Furthermore, its 2000 m) sequence of sediments that has been accu- petrological characteristics are not those of oceanic mulating in the basin since its opening at 43 Ma (Cande crust formed at slow- to fast-spreading centers (also et al., 2000); however, based on seismic stratigraphy known as mid-ocean ridge basalt; MORB) but are the that has been correlated to Deep Sea Drilling Project same as those of the seamounts (discussed below). The (DSDP) sites 273 and 274 (Granot et al., 2010, and refer- next oldest age determination at 461 6 019 Ma (D3-1) ences therein) the sediments should be much younger is also from the western flank of the Adare Trough, but (Early Oligocene to Pliocene). None of the seamounts unlike D2-1 it was collected from a prominent sea- sampled appear to be volcanically active, but the fact mount. It is interesting to note that this seamount is cut that a significant proportion of the lavas are glassy and by the bounding fault, and if the age is representative of lack manganese oxide coatings, except for those recov- the seamount on the whole, then it would suggest fault ered from the western scarp of the Adare Trough, sug- activity younger than 46 Ma. Granot et al. (2010) con- gests that the volcanic activity in this region is relatively cluded that uplift of the western flank of the Adare

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 529

Fig. 2. Representative 40Ar/39Ar ages for NWRS basalts. Age release, radiogenic argon release (40Ar*) and Ca/K spectra on ground- mass concentrates are shown. Thickness of boxes for apparent age represent 2r error.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 530 Journal of Petrology, 2018, Vol. 59, No. 3

The Possession Islands are the closest land deposits to the Adare Basin and are located adjacent to the Adare Peninsula (Fig. 1b). Age determinations for five basalts collected from three islands range from 142 6 001 Ma to 0053 6 0031 Ma (Table 1) and have a me- dian value of 024 Ma. The new ages represent the youngest volcanic activity yet recorded within the Hallett volcanic province. Smellie et al. (2011) provided 40Ar/39Ar age data as well as a compilation of ages from previous dating campaigns (40Ar/39Ar and K–Ar ages) that range from 138Ma to 23 Ma. The youngest date, 227 6 05 Ma (K–Ar age; McIntosh & Kyle, 1990), is from located at the southern end of the Adare Peninsula and within 15 km of Foyn Island (Possession Islands). Previous studies indicate that there is no overall mi- gration of activity within the Hallett volcanic province; however, three overlapping volcanic centers on the Daniell Peninsula (Fig. 1b) show a northward-younging age progression over a period of 5 Myr between 10 Ma and 5Ma(Smellie et al., 2011). New age deter- minations in this study show that there has been a very broad northeasterly shift in the location and volume of eruptive activity in the NWRS from continental shield volcanism along the coastline in the middle to late Miocene (14 to > 5 Ma) to monogenetic island and seamount volcanism on the continental shelf and Adare Basin in the Pliocene to Pleistocene (< 5Mato< 100 ka). Evidence for this broad shift in activity is underscored Fig. 3. Total alkalis vs SiO2 (wt %) classification of basalts (Le when taking into account the occurrence of Eocene– Bas et al., 1986). Thirty basalts from the NWRS (sample loca- Oligocene (48–23 Ma) plutonic and subvolcanic alkaline tions are shown in Fig. 1b and compositions in Table 2) are marked by solid color symbols (this study, n ¼ 18), and open rocks (Meander Intrusive Group), which are found with- color symbols (n ¼ 12) are analyses from Rocholl et al. (1995), in the Melbourne volcanic province (Kyle, 1990) imme- Mortimer et al. (2007), Nardini et al. (2009) and Aviado et al. diately adjacent to and to the SW of the Hallett volcanic (2015). Southern Ross Sea (SRS) seamounts and islands (open black symbols) are from Aviado et al. (2015) and the Hubert province, and are considered the earliest expression of Miller seamount (Marie Byrd seamount group) and Peter I WARS magmatism (Rocchi et al., 2002). Island (blue crosses) are from Hart et al. (1995) and Kipf et al. (2014). One sample (violet triangle, olivine nephelinite, SAX20, Nardini et al., 2009) is from the Melbourne volcanic province Petrography and mineral chemistry (MVP), which lies adjacent and to the south of the NWRS area. Twenty of the 30 basalts analyzed for whole-rock chem- All basalts plotted have MgO > 6 wt % and are normalized to istry and isotopes for the NWRS were examined petro- 100% volatile-free with total Fe as FeO. Over 400 ocean island graphically. Their phenocrysts were analyzed by Krans basalts (OIB) from the Saint Helena, Austral–Cook and Society Islands (GEOROC database, MgO > 7 wt %) are shown for com- (2013). parison and include those derived from the melting of enriched Basalts from the NWRS are porphyritic and ground- (EM) and high U/Pb (HIMU) mantle sources. mass ranges from holocrystalline to holohyaline with pilo- taxitic to trachytic textures. The glassy groundmasses are Trough occurred much earlier (17 Ma) but that vertical most prominent in seamount samples. The samples are normal faulting may be younger and, in places, still ac- variably vesicular and some basalts from seamounts dis- tive. The age determined for the western flank sea- play amygdaloidal textures, with vesicles partially filled mount is 1 Myr older than the seamount dated in the with calcite and clay material. Olivine and clinopyroxene center of the Trough (D4). The two age determinations are the dominant phenocryst phases, with subordinate from this volcano (251 6 009 Ma and 335 6 011 Ma) plagioclase and amphibole. Plagioclase and magnetite bracket the range determined for four of the five sea- are common groundmass phases, with lesser amounts of mounts dated in the southern Adare Basin (Table 1). clinopyroxene, olivine and amphibole. In a few samples The two youngest seamount samples, one from the plagioclase is absent altogether. continental shelf (D7-1, 032 6 023 Ma) and the other Olivine is euhedral to subhedral. It often contains from the easternmost seamount dredged within the inclusions of opaque oxide and melt inclusions that basin (D17-1, 014 6 002 Ma), reveal significantly make up to 2% by volume. In seamount samples skel- younger, Pleistocene activity. etal crystals of olivine are common and are the result of

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 531

quench crystallization owing to rapid undercooling. Ross Sea (Terror Rift basin; Aviado et al., 2015) and in Also in seamount samples some olivine grains show the Amundsen Sea off the Marie Byrd Land coastline minor alteration to iddingsite along cleavage fractures. (Hubert Miller seamount and Peter I Island: Hart et al., The forsterite (Fo) content [100Mg/(Mg þ Fe2þ)] of oliv- 1995; Kipf et al., 2014)(Fig. 1a). The purpose is to com- ine ranges from Fo70 to Fo90 with an average of Fo81. pare basalts derived from melting beneath regions that Most olivines display weak to strong normal zoning (de- have minimum crustal thicknesses and from oceanic crease of 1–17 mol % Fo), with rare reverse (up to settings that are considered to be related to the WARS. 3 mol % increase in Fo) and oscillatory zoning. Also shown are a selection of mafic alkaline ocean is- Clinopyroxene is euhedral to subhedral with up to 3 land basalts (OIB) from the Atlantic and southern Pacific vol. % opaque oxide and melt inclusions. In seamount oceans whose geochemical characteristics (and isotop- samples microphenocrysts of clinopyroxene often ic signatures discussed below) have been repeatedly occur as radial intergrowths (variolitic texture) with likened to basalts from the WARS. plagioclase that, like the occurrence of skeletal olivine, The composition of NWRS basalts ranges from indicate a high degree of undercooling. Clinopyroxene strongly silica-undersaturated (Ne-normative > 8to phenocrysts are bimodal in their size distribution. The 21 wt %) basanite and tephrite to more moderately larger phenocrysts (05 mm) often exhibit disequilib- silica-undersaturated (Ne-normative 11 wt %) hawai- rium textures with spongy cores and prominent zoning, ite and alkali basalt, two of which are silica-saturated whereas smaller phenocrysts (< 05 mm) are euhedral (Hy-normative) (Table 2). In addition to their higher sil- to subhedral and are only occasionally zoned. The clino- ica and lower total alkali (K2O þ Na2O) contents, the al- are classified as diopside and core composi- kali basalt also has overall lower TiO2 and P2O5 (Fig. 4) tions are in the range of Wo 37–52, En 31–52 and Fs and lower concentrations of highly incompatible trace 7–21, with Mg#cpx varying from 70 to 90. Diopside phe- elements (e.g. La, Ce, Sr and Zr) relative to basanite and nocrysts generally display weak to strong normal zon- tephrite (Fig. 5). The overall compositional spectrum of ing (up to 3% decrease in Mg# from core to rim) and cpx the NWRS basalt matches that of WARS basalt from the less common, reverse zoning (up to 3% increase in southern Ross Sea (Aviado et al., 2015) and Amundsen Mg# from core to rim). cpx Sea (Hart et al., 1995; Kipf et al., 2014) as well as OIB, Plagioclase phenocrysts are present in only six of the with the notable exception of Yb, which is much lower 20 NWRS basalts examined. Euhedral plagioclase phe- in concentration for Peter I Island (Fig. 5). There are no nocrysts have core compositions that range from An42 distinct compositional trends on plots of major and to An69; most are subhedral to anhedral with sieve tex- trace elements versus MgO wt % with the exception of tures and reaction rims. In addition, they are reversely Al O and Cr (Figs 4 and 5). The variation in concentra- zoned with anorthite contents that show an increase 2 3 tion of these elements with decreasing MgO content from core to rim by up to 22 mol %. The evidence for indicates an overall control by clinopyroxene during disequilibrium in the majority of plagioclase suggests that they are antecrysts (i.e. crystals that formed earlier magmatic differentiation. Nickel contents (not shown) in the differentiation process and have been reincorpo- decrease with MgO in samples with Ni < 250 ppm, indi- rated into the magma). cating olivine control. Amphibole phenocrysts are less common, observed Primitive mantle normalized trace element patterns in five of the 20 basalts examined, and typically exhibit (Fig. 6a–c) of basalts from the NWRS, along with other weak to extensive reaction rims. The amphibole is ferri- WARS basalts, display enriched patterns and well- kaersutite, following the classification scheme and no- developed K and Pb negative anomalies for nearly all samples. Alkali basalts are again distinguished from menclature of Hawthorne et al. (2012), with Mg#amp 66–72. Kaersutitic amphibole is often found in alkaline basanite (and tephrite) by having overall lower elemen- basalts and within the NWRS and the adjacent tal concentrations, less pronounced K and Pb anomalies Melbourne Volcanic Province (Kyle, 1986; Rocchi et al., and small negative P anomalies (Fig. 6b). All of the 2002; Coltorti et al., 2004; Perinelli et al., 2006, 2011; basalts are enriched in light rare earth elements (LREE) Armienti & Perinelli, 2010; Perinelli et al., 2017). relative to heavy REE (HREE), which is typically linked to garnet control during small degrees of mantle partial Major and trace element characteristics melting. Alkali basalts have lower LaN/YbN ratios ( 15) The whole-rock composition of the 30 selected NWRS compared with and tephrites (LaN/YbN 12– basalts includes alkali basalt, basanite, hawaiite and 27) in NWRS samples, contributing to their slightly flat- tephrite (Table 2, Fig. 3). All were selected to represent ter patterns. Chondrite-normalized REE plots of NWRS relatively unfractionated compositions (MgO 6wt %) and other WARS basalts lack Eu anomalies (Eu/Eu* be- and include deposits located on the continent, continen- tween 09 and 11) indicating that the magmas were tal shelf and oceanic Adare Basin (Fig. 1b). The major not influenced by the fractionation or accumulation of and trace element compositions of NWRS basalts are plagioclase. All of the WARS basalts fall within the com- plotted in Figs 3–5 with other basalts (MgO 6wt %) positional range of OIB and display very similar trace from seamounts and islands located in the southern element distribution patterns.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 532 Journal of Petrology, 2018, Vol. 59, No. 3

Fig. 4. Major elements (wt %) vs MgO (wt %). References, symbols and data criteria are given in Fig. 3 and Table 2.

Sr–Nd–Pb isotopes encompass a moderate range in measured 87Sr/86Sr Radiogenic isotope ratios for NWRS basalts are (070278–070379) and a narrow range in measured reported in Table 2 and plotted with other WARS 143Nd/144Nd (051282–051300). Pb isotope composi- basalts and alkaline OIB in Fig. 7. Apart from one sam- tions measured for NWRS basalts also show a ple from the Malta Plateau (MA009a), NWRS basalts restricted range in 207Pb/204Pb (1552–1568) and a

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 533

Fig. 5. Selected trace elements (ppm) vs MgO (wt %). Symbols and data criteria are given in Fig. 3 and Table 2.

moderately wide range in 208Pb/204Pb (3881–3992) and source (Hart et al., 1995; Kipf et al., 2014). Seven NWRS 206Pb/204Pb (1922–2023). Values for other WARS basalts have high 206Pb/204Pb (> 20), six of which also basalts fall within the same range, except for Peter I have low 87Sr/86Sr (< 07030) and high 143Nd/144Nd Island, which has a much higher 207Pb/204Pb (>157) (> 05129) (Table 2). The isotopic signatures coupled that is explained by derivation from a localized EMII-like with their OIB-like compositional characteristics trend

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 534 Journal of Petrology, 2018, Vol. 59, No. 3

during recycling through the mantle (Stracke et al., 2005). Alternatively, PREMA, which is very similar, if not identical, to the so-called focal zone (FOZO; Hart et al., 1992) or the common component (‘C’; Hanan & Graham, 1996) in the mantle sources of oceanic basalts, is the lithospheric mantle portion of the subducting slab (Castillo, 2015, 2016). Accordingly, the DM–PREMA array may simply be the long-time (some billions of years) integrated record of geochemical depletion of the upper mantle with time, from Bulk Silicate Earth (BSE) to modern DMM (Depleted MORB Mantle).

Oxygen isotopes Oxygen isotopic compositions determined by conven- tional laser fluorination (LF) and by secondary ionization mass spectrometry (SIMS) for NWRS basalts are listed in Table 3 and plotted against olivine Fo content and whole- rock MgO in Fig. 8. Oxygen isotopes measured by LF for nine NWRS basalts range from 486 to 526&. Overall, the average LF values for NWRS olivine in this study (503 6 016&, 2SD) are statistically indistinguishable from average olivine LF values (d18O ¼ 529 6 013&, 2SD) reported for other Northern Victoria Land (NVL) basalts (Nardini et al., 2009), as well as olivine LF values (d18O ¼ 529 6 018&, 2SD) from xenoliths hosted in NVL basalts, which are considered to represent cumulates from near-primary mantle melts (Perinelli et al.,2011). Olivine measured by SIMS, including core and rim analyses, and multiple analyses on single grain cores (Core*), vary from 44to59& (range ¼ 15&) and aver-

age 521 6 026& (2SD, number of samples ns ¼ 21, number of analyses na ¼ 176). Intra-crystal variations in 18 d Ool are as much as 06& lighter to 05& heavier at rims than cores (average Dcore–rim ¼ 002&) and are not statistically significant with respect to instrument preci- sion based on the SCOL standard (6 028& at 2SD). The 18 lowest single spot d Ool values are from samples NV-4 (45&) and NV-4C (44&), with average values being

slightly higher at 482 6 018& (na ¼ 7) and 466 6 025& 18 (na ¼ 13), respectively. The highest single spot d Ool values are from samples MA-009a and MA-117 (59&), Fig. 6. Primitive mantle normalized (Sun & McDonough, 1989) with averages at 554 6 022& (n ¼ 14) and trace element diagrams. References, symbols and data criteria a are the same as in previous figures. 557 6 026& (na ¼ 13), respectively. The difference be- 18 tween single spot analysis of high and low d Ool values is resolvable beyond analytical precision. 18 towards what has been referred to as HIMU [i.e. derived SIMS d Ool average values are not statistically dif- from melting of a mantle source with high- ferent from LF average values for individual samples 238 204 l ¼ ( U/ Pb)t¼0](Rocholl et al., 1995; Hart et al., 1997; (Table 3) or for the dataset as a whole (Fig. 8). Multiple Panter et al., 1997, 2000, 2006; Rocchi et al., 2002; core analyses on single grains measured by SIMS Nardini et al., 2009; Martin et al., 2013; Kipf et al., 2014; (Core*) are either indistinguishable (e.g. NV-4) or heav- Aviado et al., 2015). Nevertheless, most of the basalts in ier (e.g. D4-3 and MA-117) relative to LF (Table 3). It is this study fall within the range of OIB compositions interesting to note that out of the seven samples meas- assigned to a PREMA (PREvalent MAntle; Zindler & ured using the Core* technique, five show heavier val- Hart, 1986) mantle source; specifically falling within the ues (A225A, D4-3, MA-009a, MA-117 and P74833) than DM (Depleted Mantle)–PREMA array as defined by standard SIMS measurements. It should be recalled Stracke (2012, and references therein). The develop- that the standard measurement is an average of mul- ment of the DM–PREMA array is attributed to the con- tiple spots on two to three grains (mostly cores but tinuous subduction and aging of MORB-type crust includes some rims) and that Core* is the average of

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 535

Fig. 7. Measured Radiogenic isotope (Sr, Nd, Pb) compositions of Antarctic basalts compared with ocean island basalts. Mantle source endmembers PREMA (FOZO), HIMU and EM are taken from Stracke (2012). References, symbols and data criteria are given in Fig. 3 and Table 3. Basalts MA009a and MA-117 from the Malta Plateau (Rocholl et al., 1995) were measured for Pb isotopes in this study and the results are presented in Table 2.

18 multiple spots in the core of one grain only. It is pos- correlation between Fo content and d Ool nor a correl- sible that there is some sampling bias contributing to ation with respect to whole-rock composition (i.e. bas- the difference between the two techniques but, if so, the anite, alkali basalt, etc.). There is a very weak positive differences fall within analytical precision. correlation, however, when considering only LF analy- 18 All d Ool analyses by SIMS and LF (this study and ses (this study and Nardini et al., 2009). In Fig. 8b, 18 Nardini et al., 2009) are plotted against Fo % and pre- whole-rock MgO (wt %) is plotted versus average d Ool sented in Fig. 8a. Also plotted is a range of mantle oliv- LF and SIMS data for each sample. Again, there is a 18 ine values from peridotite (LF; Mattey et al., 1994) and very weak positive correlation (i.e. lighter d Ool with olivine data from basalts at several oceanic locations decreasing MgO content) mainly exhibited by LF analy- that include the Canary Islands (LF and SIMS; Gurenko ses, but when considering both SIMS and LF data for et al., 2006, 2011), Reykjanes Peninsula, Iceland (LF; basanite and alkali basalt there is no correlation; in fact, 18 Peate et al., 2009) and the Azores Islands (LF; Grenske nearly the full range of d Ool values for NWRS basalt et al., 2013), all of which have also been interpreted to falls between 9 and 10 wt % MgO (Fig. 8b). 18 represent mantle values. The LF data for NWRS olivine d Ool analyses by SIMS and LF averaged for each have nearly the same range in values as measured for sample are plotted against measured Sr, Nd and Pb iso- the Canary and Azores islands. Individual SIMS analy- topic compositions, as shown in Fig. 9. There is a broad ses for NWRS olivine have a greater range, encompass- correlation between oxygen and radiogenic isotopes, 18 18 87 86 ing these islands and most d Ool values for OIB where lower d Ool corresponds to lower Sr/ Sr and measured by LF (Eiler, 2001). However, the average higher 143Nd/144Nd and 206Pb/204Pb ratios (as well as 207 18 & 208 204 value of all d Ool analyses in this study (520 6 025&) Pb/ Pb, not shown). Significantly, a similar rela- 18 is similar to average mantle d Ool (518 6 028&; tionship had been generally observed in OIB samples Mattey et al., 1994). There is neither an overall (Eiler et al., 1996).

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 536 Journal of Petrology, 2018, Vol. 59, No. 3

from land to sea. Oceanward, the basalts show an in- crease in total alkalis, silica-undersaturation (Ne þ Lc

normative, not shown), P2O5 and incompatible trace elements Sr, Zr, Nb, LREE (not shown), and LREE/HREE ratios. Also towards the ocean, the 87Sr/86Sr ratios de- crease and 143Nd/144Nd and 206Pb/204Pb ratios increase. Trends in 207Pb/204Pb and 208Pb/204Pb isotopes (not shown), although poorly defined, also generally in- crease oceanward. Variations in oxygen isotopes are less well defined but when samples classified as alkali basalt are considered (Figs 1b and 3) there is a general decrease (i.e. values become lighter) towards the ocean (Fig. 10). The systematic compositional variation requires the evaluation of gradational changes in man- tle source, melting processes and/or contamination of the melt as it rises through the lithosphere.

DISCUSSION The main objective of this study is to better understand the mantle sources and processes responsible for the production of alkaline magmas within the WARS. The unique contribution of this research relative to past efforts is the location of our basaltic sample suite, which encompasses both oceanic and continental domains (Fig 1). Given their similar composition and roughly equivalent age, the oceanic and continental basalts must share a common origin and therefore provide an effective assessment of the role of continental litho- sphere in magma genesis. In addition, the NWRS was located adjacent to the former pan-Pacific margin of Gondwana, where subduction dominated the tectonic regime from the Neoproterozoic to the Late Cretaceous 18 Fig. 8. Relationships between d Ool and (a) average Fo % and ( 450 Myr); thus, the influence of subduction-related (b) whole-rock MgO wt %. Oxygen isotope compositions of olivine were measured by SIMS (single spot) and LF (4–5 processes on mantle source domains is also examined. grains) analysis for NWRS samples (this study) and LF analy- ses of olivine from other NVL basalts, including those in the Contamination by crust? NWRS and MVP (Mortimer et al., 2007; Nardini et al., 2009). (a) 18 The increase in silica-saturation, and correlation of d Ool obtained by SIMS and Fo% by EMPA include both cores and rims of individual olivine grains, comprising a total of 21 higher 87Sr/86Sr and lower 143Nd/144Nd with higher samples (Table 3). The line marking the average d18O value of 18 d Ool values in NWRS basalts erupted on the Antarctic NWRS olivine (520 6 025&, 2SD) is based on 190 analyses, including nine samples by LF (Table 3). The Fo content for sam- continent might suggest that the magmas have assimi- ples analyzed by LF is the average of multiple grains deter- lated crust. Contrary to this presumption is the fact that mined by EMPA in same sample. The dashed field for mantle our ‘continental’ samples are also distinguished by hav- olivine represents an average value of 518 6 028& at 2SD (LF; Mattey et al., 1994) and Fo % of 87–96 (Deer et al., 1992). For ing lower incompatible element concentrations relative comparison, values for olivine from the Canary Islands (LF and to their oceanic counterparts (Fig. 10). Progressive as- SIMS; Gurenko et al., 2006, 2011), Reykjanes Peninsula, Iceland similation of silicic crust coupled with fractional crystal- (LF; Peate et al., 2009) and the Azores Islands (LF; Genske et al., 18 lization (AFC; DePaolo, 1981) should produce the 2013) are shown. (b) SIMS d Ool are the average of all spot analyses within a single sample (cores and rims on several opposite relationship. In Fig. 11 we have modeled the grains). Tie-lines connect individual samples that have been effects of AFC processes using a variety of crustal rock 18 analyzed by multiple methods. Core* is the average d Ool types found within NVL. The plots clearly show that the value of 5–6 spots within the core of a single olivine grain. assimilation of these rocks by basaltic magmas will lead to lower Nd and higher Sr isotopic signatures, but Geochemical and isotopic variation across the will also produce higher concentrations of incompatible ocean–continent transition elements such as Rb (as well as Nb, Zr, Ba, LREE) even The 30 basalts from the NWRS (Table 2 and Fig. 1b) are if they are significantly lower in the assimilant (e.g. gab- plotted versus decimal degree longitude in Fig. 10 and bro). Furthermore, there is no systematic increase or

display systematic variations in major and trace ele- decrease in MgO, Al2O3 and CaO or weakly incompat- ments as well as isotopic ratios across the transition ible to compatible trace element (e.g. Y, Yb, Cr) contents

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 537

continental rocks by mantle-derived magmas (Harmon & Hoefs, 1995; Eiler, 2001, and references therein). In cases where anomalously light oxygen isotope signa- 18 tures for olivine (i.e. low d Ool) occur in OIB, they have been attributed to the assimilation of hydrothermally altered oceanic lower crust (Hart et al., 1999; Skovgaard et al., 2001; Genske et al., 2013) or cannibalization of hydrothermally altered lavas within the volcanic edifice itself (Wang & Eiler, 2008; Garcia et al., 2008). A rough 18 positive correlation between d Ool and Fo %, first reported by Nardini et al. (2009) and weakly supported by additional LF analyses in this study (Fig. 8a), led the authors to conclude that the trend towards lighter 18 d Ool values found in NVL basalts is the result of canni- balization of altered volcanic rocks. To more fully assess the possible effects of assimilation we have selected, in addition to NVL crustal rocks, an altered mugearite lava from Mount Sidley, Marie Byrd Land, which has a very low d18O whole-rock value of 18&; a result of alteration by highly 18O-depleted Antarctic meteoric water (Panter et al., 1997). The interaction between basalts and poten- 18 tial contaminants is evaluated on a plot of d Ool versus whole-rock Sr concentration shown in Fig. 12. The trend 18 in d Ool of the alkali basalts with longitude (Fig. 10i) and the strong correlation with Sr concentration (Fig. 12) may be explained by (1) AFC involving ‘normal’ to slightly 18O-enriched basaltic magma interacting with low d18O volcanic rocks and the contamination increas- ing oceanward, (2) mixing between low d18O(< 5&) basaltic magma and mafic crust (e.g. gabbro, d18O 7&) with the proportion of crust increasing con- tinentward, or (3) reaction between low d18O melt and

peridotite (e.g. Opx þ Liq0 $ Ol þ Liq1; Yaxley & Green, 1998; Pilet et al., 2008; Lambart et al., 2012) that is enhanced continentward as melt migrates though thicker mantle lithosphere. The first scenario (cannibal- ization) is difficult to defend given the trend of the alkali basalts towards lower 87Sr/86Sr and higher 143Nd/144Nd ratios oceanward (Fig. 10g and h). The Sr and Nd iso- 18 Fig. 9. d Ool vs measured Sr, Nd and Pb isotope compositions. topic ratios should remain relatively constant if the con- The oxygen isotopic values for olivine phenocrysts from taminant was altered by Antarctic meteoric water. If the NWRS basalts are LF (open symbols) and SIMS data averaged 18 for multiple spots on 2–3 grains in each sample (Table 3). Error low d O signature of the contaminant was a result of al- bars are standard error of the mean (SEM). Symbols for the teration by seawater at high temperatures ( 250C; basalts are the same as those given in Fig. 8. Radiogenic iso- Muehlenbachs, 1986; Gao et al., 2012) it would also pos- tope ratios are measured values on whole-rocks. Vertical line marks the average d18O value of NWRS olivine (520&) calcu- sess an elevated Sr isotope signature, and therefore as- lated for all analyses (n ¼ 190). similation would lead to a magma that has higher 87Sr/86Sr along with a low d18O value, which is not the from ocean to continent, which would be very difficult case here. Simple mixing provides a better fit for the to explain if the continental basalts were preferentially geochemical and isotopic trends exhibited by the alkali contaminated by crust. However, simple mixing of basalts (Figs 11 and 12) but again requires a significant magma with mafic rocks such as gabbro or eclogite proportion of mafic crust (i.e. gabbro/eclogite) and a could predict many of the geochemical and Sr–Nd iso- source for low d18O magma. There is a general positive 18 topic variations but require a mixture that consists of correlation between d Ool and Sr concentration for 40% of those crust types. Achieving this proportion, basanite, hawaiite and tephrite lavas, which can be especially at shallow depths with relatively small vol- mostly enveloped by AFC model curves constructed umes of magma, is unlikely. based on the interaction between a low-18O magma Oxygen isotopes have long been used to assess con- (e.g. NV-4C) and gabbroic crust, but only if the propor- tamination of magma by crust and are a particularly tion of assimilation to crystal fractionation is allowed to strong tool for detecting the assimilation of 18O-rich vary significantly (r 1to03; Fig. 12). In addition, the

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 538 Journal of Petrology, 2018, Vol. 59, No. 3

Fig. 10. Variation in major and trace elements and measured Sr–Nd–Pb–O isotopes with decimal degree longitude for NWRS basalts. References, symbols and data criteria are the same as in Fig. 3, and for oxygen isotopes, Fig. 8. The geographical location of each sample is shown in Fig. 1b. Ocean and continental sectors are defined in Fig. 1b and the ocean to continent transition zone (OCTZ) is demarcated as the region between the base of the continental shelf (1500 mbsl) and the inferred boundary between the East Antarctica craton and extended lithosphere comprising the WARS (bold dashed line in Fig. 1b). Lines of regression, which in- 2 18 clude all data, have r that range from 04to06 and Spearman rank correlation coefficients that range from 06to08. For d Ool only alkali basalt is regressed. Oxygen values are SIMS analyses except for three NWRS samples (open symbols) measured by LF from Nardini et al. (2009). Error bars represent the standard error of the mean (SEM). Sample La/Yb ratios are normalized to primi- tive mantle (Sun & McDonough, 1989).

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 539

Fig. 11. Measured Sr and Nd isotopic compositions vs Rb concentrations (ppm). (a, b) Selected basalt and crustal rock compositions used for mixing (Langmuir et al, 1977) and assimilation–fractional crystallization (AFC; DePaolo, 1981) models. Basalt sample D12-1 was chosen to be representative of a parental magma composition based on its oceanic location and relatively unfractionated com- position (MgO ¼ 107wt%,Cr¼ 355 ppm, Ni ¼ 205 ppm; Table 2). Also selected is sample SAX20 (Table 2), which is overall the most incompatible element enriched sample (Fig. 6) and has been equated to a primary (metasomatic) melt by Coltorti et al. (2004) and Perinelli et al. (2006). Crustal rocks from the NVL consist of gabbro, diorite and granodiorite (Di Vincenzo & Rocchi, 1999; Dallai et al., 2003), metasediment (Henjes-Kunst & Schussler, 2003; Di Vincenzo et al.,2014) and eclogite (Di Vincenzo et al.,1997; Ghiribelli, 2000). Gray fields encompass AFC model curves [shown in detail in (c) and (d)] using DRb ¼ 001, DSr ¼ 01, DNd ¼ 04 and proportion of as- similation to crystallization (r) ¼ 08. Mixing curves between basalt, eclogite and gabbro are also shown. Numbered tick marks are the percentage of crust in each mixture. (c, d) Detail showing AFC model curves using basalts D12-1 (Table 2)andSAX20(Nardini et al., 2009) and different contaminants (eclogite is sample G3; gabbro is DR13 and AC6; diorite is sample AF16; metasediment is argillitic). Numbers along each curve represents the fraction of liquid remaining (F). Gray dashed arrow shows the trend of increasing SiO2 (wt %) in basalt samples. References, symbols and data criteria for the basalts are the same as those given in Fig. 3.

lower Sr concentrations in alkali basalts relative to Mantle sources 18 basanite and tephrite cannot be explained by these The wide range in d Ool exhibited by NWRS basalts models. (variation of 15&; Fig. 8a) is comparable with the In summary, processes involving shallow-level con- range observed in OIB, which has been explained by tamination by crust do not adequately account for the mantle source heterogeneity; a result of variably variation in major and trace elements and isotopes altered, recycled oceanic crust (6 sediments) within the shown by the NWRS basalts. Furthermore, none of the convecting mantle (Eiler et al., 1996; Eiler, 2001; Day scenarios discussed above are able to emulate the et al., 2010; Gurenko et al., 2011). A recent comprehen- well-defined trends in major and trace elements and sive review of petrological and geochemical constraints isotopes between land and sea (Fig. 10). Consequently, on the origin of mafic alkaline rocks (primarily OIB) has the third scenario, which discounts the involvement of been provided by Pilet (2015) and details how the direct crust altogether, will be evaluated along with melting melting of pyroxenite (recycled upper or lower oceanic processes and mantle source(s) in the following crust ¼ MORB or gabbro, respectively), with or without section. added sediment, cannot reproduce the major element

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 540 Journal of Petrology, 2018, Vol. 59, No. 3

CO2 (Dasgupta & Hirschman, 2007); (2) pyroxenite melt reacts with peridotite (dissolution/precipitation reac- tion ¼ olivine þ liquid $ orthopyroxene þ liquid; precipi- tation of either orthopyroxene or olivine may also include clinopyroxene and garnet) under dry conditions (Lambart et al., 2012; Mallik & Dasgupta, 2014) or in the

presence of CO2 (Mallik & Dasgupta, 2014) prior to reaching the surface; (3) a multi-stage process that begins with a low-degree melt from the asthenosphere that penetrates and forms amphibole-bearing cumu- lates within the lithospheric mantle. Later melting of these relatively small-volume, trace element and volatile-rich cumulates (metasomatic veins) at high degrees, and reaction of this liquid [as in (2)] with the surrounding mantle on its way to the surface, can repro- duce the major and trace element characteristics of the mafic alkaline magma (Pilet et al., 2008). The third source type, metasomatized lithosphere, has also been proposed to explain the petrogenesis of alkaline basalts erupted in continental settings (Stein et al., 1997; Jung et al., 2005; Panter et al., 2006; Ma et al., 2011; Mayer et al., 2014; Rooney et al., 2014, 2017) including those within the WARS (Rocchi et al., 2002; Panter et al., 2003; Nardini et al., 2009; Perinelli et al., 2011; Martin et al., 2013; Aviado et al., 2015). In the con- tinental NVL, Perinelli et al. (2011) considered the vari- 18 Fig. 12. Variations in d Ool with whole-rock Sr concentrations ability in olivine oxygen isotope data for mantle for NWRS basalts modeled by AFC using an altered basaltic composition (mugearite MB32.11; Panter et al., 1997) and mix- cumulate xenoliths (500–572&), together with olivine ing and AFC using gabbro from NVL (Di Vincenzo & Rocchi, from alkaline basalts (492–553&, Nardini et al., 2009) 1999; Dallai et al., 2003). Sample MA009a was selected to rep- and peridotites (480–581&, Perinelli et al., 2006) to sig- resent a slightly 18O-enriched basaltic magma and sample NV- 18 18 nify extensive O compositional heterogeneity in the 4C was selected to represent a low d O basaltic magma 18 lithospheric mantle. Our new d Ool values collected (Tables 2 and 3). AFC models use DSr ¼ 01, oxygen isotope bulk mineral–melt fractionation ¼ 04& (Eiler, 2001), and the from continental and oceanic NWRS basalts extend that proportion of assimilation to crystallization (r) ¼ 08, 05 and range (444–592&; Fig. 8a). Perinelli et al. (2011) con- 03. Numbered tick marks on AFC curves represent the amount cluded that mantle cumulates were modally and cryp- of liquid remaining and tick marks on the mixing curve repre- 18 sent the proportion of gabbro in the mixture. All but two of the tically metasomatized by low d O melts, which in turn basalts have MgO > 6wt%(Table 3) and the symbols used are originated by the melting of earlier formed metasomes the same as in Fig. 8. The method for mixing and AFC are the emplaced within the continental lithosphere during rift- same as in Fig. 11. ing in the Late Cretaceous. The ultimate origin of the low d18O mantle signature found in NVL and the NWRS compositions of low-silica alkaline basalts, which in- will be discussed below. clude basanite, tephrite and nephelinite (Fig. 3). The The major element compositions of continental and melting experiments for silica-deficient pyroxenite in oceanic basalts from West Antarctica reported in the presence of CO2 (Dasgupta, 2006; Gerbode & this study are consistent with their being derived Dasgupta, 2010), however, can produce liquids that are from mantle sources that are predominantly silica- closer in SiO2 and total alkalis content but have Al2O3/ undersaturated and contain amphibole-rich lithologies TiO2 and Na2O/K2O ratios that vary significantly, which (Fig. 13) and align with the metasomatized lithospheric is in contrast to the relatively low and constant ratios hypothesis. The presence of amphibole in the source is observed in natural low-silica alkaline basalts (Pilet, also supported by negative K anomalies and high Nb– 2015), including basalts from the NWRS (505 6 106 Ta contents displayed on multi-element, primitive and 275 6 047, respectively). Therefore, to account for mantle-normalized patterns (Fig. 6). Potassium occurs the trace element and radiogenic isotope signatures in in stoichiometric proportions in the mantle minerals mafic alkaline rocks, which are ultimately attributed to amphibole (pargasite, kaersutite) and mica (phlogopite). recycled materials, a pyroxenitic melt must be further If these minerals are not completely consumed during conditioned and homogenized before it reaches the sur- mantle partial melting, then K will be retained in the face. Pilet (2015) described three possible scenarios for source and the melt will be depleted in K relative to this: (1) pyroxenite melt infiltrates and enriches (meta- neighboring trace elements (e.g. Nb, Ta, La and Ce), somatizes) peridotite in the asthenosphere, which is which will behave incompatibly (LaTourrette et al., then melted at low degrees (1–5%) in the presence of 1995; Ionov et al., 2002; Tiepolo et al., 2007, and

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 541

Fig. 13. Plot of normalized (CIPW) compositions calculated from whole-rock major element data. Silica-undersaturated composi- tions plot within the Ol–Ne þ Lc–Di triangle (truncated at 50 wt % Ne þ Lc) and silica-saturated compositions plot within the Ol–Hy– Di triangle (up to 50 wt % Hy). Symbols and data criteria are the same as in previous figures. Fields for basaltic compositions derived from melting experiments include carbonated pyroxenite (29 GPa, 1275–1475C; Gerbode & Dasgupta, 2010), carbonated eclogite (35 GPa, 1250–1400C; Kiseeva et al., 2012), hornblendite (15 GPa, 1150–1350C; Pilet et al., 2008), clinopyroxene hornblen- dite (1 GPa, 1250–1400C; Pilet et al., 2008), amphibole wehrlite (10 GPa, 1225–1350C; Me´ dard et al., 2006), garnet clinopyroxenite (2–25 GPa, 1340–1500C; Hirschmann et al., 2003; Keshav et al., 2004). Also included is melt derived from hornblendite plus perido- tite sandwich experiments (15 GPa, 1225–1325C; Pilet et al., 2008). Dashed gray arrow represents increasing proportion of melt fraction (F) determined for garnet clinopyroxenite in experiments by Keshav et al. (2004).

references therein). High Nb and Ta concentrations Morning, which is close to the global average for carbo- ( 100 primitive mantle; Sun & McDonough, 1989) natites (35; Chakhmouradian, 2006), as well as having a have been measured on vein amphibole in peridotite Nb concentration (547 ppm) that is mostly higher than (Ionov & Hofmann, 1995; Ionov et al., 2002) and for both phlogopite (and amphibole) from mantle xenoliths that vein and disseminated amphibole found in peridotite have been attributed to mantle lithosphere enriched by from NVL (Coltorti et al., 2004; Perinelli et al., 2006). carbonate-rich silicate liquids (Ionov & Hofmann, 1995; Niobium enrichment of the lithospheric mantle has also Ionov et al., 2002). Experimental results indicate that been related to metasomatism by carbonatite, which carbonate phases can coexist with amphibole and can have concentrations > 200–1000 primitive mantle phlogopite in spinel and garnet peridotite at high pres- (Pfa¨ nder et al., 2012, and references therein). Martin sures (Brey et al., 1983; Olafsson & Eggler, 1983). et al. (2013) ascribed variations in minor and trace elem- Sources with elevated Nb and Ta concentrations may ent ratios (e.g. elevated Nb/Ta as established by also reside in rutile-bearing peridotite (Kalfoun et al., Pfa¨ nder et al., 2012) to signify carbonate metasomatism 2002) and eclogite (Rudnick et al., 2000; John et al., in the lithospheric mantle sources of alkaline basaltic 2004). An eclogitic contribution to the metasomatic en- rocks from Mount Morning in Southern Victoria Land richment of the sub-continental lithosphere beneath (SVL, Fig. 1a). This is also consistent with the Nb/Ta NVL has been proposed by Melchiorre et al. (2011) ratio (39) measured on mantle phlogopite from Mount based on radiogenic values of Os and Hf isotopes

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 542 Journal of Petrology, 2018, Vol. 59, No. 3

measured on sulfide and clinopyroxene (respectively) from amphibole-bearing and dry peridotite xenoliths. Here we also emphasize that the radiogenic isotope signatures in mafic alkaline rocks are ultimately sourced from materials recycled by subduction. The trace elem- ent and Sr, Nd and Pb isotopic signatures of the alkaline rocks fall within the compositional range of OIB. The Sr, Nd and Pb isotopic composition of OIB, however, comes from recycled subducted materials (Hofmann, 2014; White, 2015, and references therein). Specifically, the Pb isotope compositions of some of the alkaline rocks are highly radiogenic, similar to HIMU OIB, whereas those of the bulk of the rocks are less radio- genic, akin to PREMA (or FOZO)-type OIB, which are also sometimes called young HIMU (Thirlwall, 1997; Stracke et al., 2005). Although it is widely accepted that the highly radiogenic Pb signature of HIMU is due to recycling of variably altered basaltic crust (e.g. Hofmann, 2014; White, 2015, and references therein), a recycled marine carbonate source for the radiogenic Pb isotope composition of HIMU has recently been pro- posed (Castillo, 2015, 2016). This alternative proposal posits that recycled marine carbonates act as dirty Pb isotopic ‘spike’ that variably contaminates (i.e. from PREMA to HIMU) the mantle sources of OIB and other intraplate lavas. Additionally, the Nb–Ta enrichment but Pb depletion that are pervasive features of OIB (e.g. Hart & Gaetani, 2006; Jackson et al., 2008; Hofmann, 2014; Peters & Day, 2014) is strongly complemented by the Nb–Ta depletion but Pb enrichment in subduction- related lavas (Castillo, 2015). Notably, the K depletion of the alkaline rocks is also an additional, distinctive fea- ture of HIMU-type OIB (Castillo, 2015; Weiss et al., 2016). Thus, we propose that a compositional connec- tion exists between the mafic alkaline rocks and previ- ously subducted carbonate-rich materials. This does not conflict with our proposal for metasomatized litho- sphere given that the metasomatizing fluids that form the amphibole-rich metasomes are considered to be themselves derived from melting of a sublithospheric Fig. 14. La/YbN, Zr/Nb and Ti/Zr vs Nb (ppm). Basalts are com- pared with experimental melts by Pilet et al. (2008) on amphi- source containing recycled subducted materials (dis- bole-bearing lithologies at 15 GPa and temperatures 1200– cussed below). 1400C. La/Yb ratio is normalized to primitive mantle from Sun & McDonough (1989). Symbols and data criteria for the basalts are the same as those given in Fig. 3. Curves represent modal Mantle melting and non-modal batch melting using four different source com- The variation in major and trace elements and isotopes positions. Model batch melt of eclogite (see Pfa¨ nder et al., displayed by NWRS basalts in Fig. 10 has been sug- 2007) uses an average eclogite composition (including Group III type) from John et al. (2004), source and melt mode of 080 gested to be the result of the tapping of a heteroge- clinopyroxene, 020 garnet, and partition coefficients from neous, metasomatically enriched, mantle lithosphere at Pertermann et al. (2004; experiment A343). Non-modal batch melting of garnet lherzolite uses the source composition for primitive mantle (PM) from Hofmann (1988), excepting Nb con- bearing) from Hartmann & Wedepohl (1990) with source and centration, which has been reset to 0491 ppm (Mu¨ nker et al., melt mode reactions and partition coefficients from Pfa¨ nder 2003), and a source derived from Hobbs Coast basalts in Marie et al. (2012, and references therein). Titanium is compatible in phl/melt Byrd Land by Hart et al. (1997). Source mode and melt mode phlogopite (KD ¼ 3, Pilet et al., 2011, and references reactions follow Salters (1996) where the source mode is 053 therein) and therefore non-modal batch melting of phlogopite- olivine, 004 orthopyroxene, 038 clinopyroxene and 005 gar- bearing peridotite (H-group) in (c) was modeled using a source net, and the melt mode is 005 olivine, –049 orthopyroxene, mode of 055 olivine, 022 orthopyroxene, 015 clinopyroxene, 131 clinopyroxene and 013 garnet. Partition coefficients are 003 spinel, 005 phlogopite and a melt mode of from Pilet et al. (2011, and references therein). The fourth 005:005:065:005:02, respectively (see Ersoy et al., 2010). source is spinel peridotite that has experienced metasomatic Numbers along the curves represent the degrees of melting in enrichment (average ‘H-group’, high-potassium, phlogopite- per cent.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 543

progressively smaller melt fractions oceanward (Panter the continent to ocean boundary differs substantially. et al., 2011). In Fig. 14, basalts from the NWRS along Estimates of crustal thicknesses predict Moho depths with the other oceanic WARS basalts highlighted in this beneath the eastern flank of Transantarctic Mountains study are compared with model curves for modal and to be between 25 and 35 km (Behrendt, 1999; Salimbeni non-modal batch melting that involve three potential et al., 2010) and in the Northern Basin (Fig. 1) between mantle sources that are relevant to what has been dis- 20 and 15 km (Busetti et al., 1999). A Moho depth of cussed above. They include a mantle composition cal- 5–6 km is estimated at the shelf break between the culated as the source of the basaltic volcanism erupted Northern Basin and the Adare Basin (Selvans et al., near the Hobbs Coast in Marie Byrd Land (Hart et al., 2014) and that depth extends north into the Adare 1997), an average of Zambian eclogites interpreted to Trough (Mu¨ ller et al., 2005; Selvans et al., 2014). The represent fossil subducted slab material (John et al., depth to the base of the lithosphere is broadly resolved 2004) and an average composition for metasomatized through regional tomographic studies that indicate (high-K, phlogopite-bearing) spinel peridotite xenoliths thicknesses of > 150 km beneath the East Antarctic cra- from NW Germany (H-group; Hartmann & Wedepohl, ton, 100–70 km within the West Antarctic rift and 1990). Melting of a primitive mantle (PM) composition < 80 km beneath oceanic lithosphere (Danesi & Morelli, (Hofmann, 1988) is also shown. Additionally, for com- 2001; Ritzwoller et al., 2001; Heeszel et al., 2016). If man- parison, the results of high-pressure and high- tle source compositions and mantle potential tempera- temperature melting experiments on natural hornblen- tures near the base of the lithosphere are similar across dite and clinopyroxene hornblendite from Pilet et al. the region, then partial melting should occur at much (2008) are plotted. shallower depths beneath the Adare Basin relative to Models for PM, Hobbs and H-group sources produce the continent. This scenario would result in higher steeper positive slopes with decreasing melt fractions degrees of partial melting away from the continent, in Nb versus La/YbN relative to the data arrays for basalt which is contrary to the interpretation based solely on and melting experiments (Fig. 14a), whereas eclogite geochemistry. provides a very good fit to the data. In Nb versus Zr/Nb The most Si-undersaturated and incompatible elem- (Fig. 14b), melting of metasomatized peridotite (H- ent enriched basalts (basanite and tephrite) in the group) provides the best fit, but eclogite is also reason- NWRS were also interpreted by Panter et al. (2011) to able at low melt fractions. All four models produce low represent small-degree melts that scavenged the most Ti/Zr ratios (Fig. 14c), but at lower melt fractions (1–3%) easily fusible material in the lithosphere (i.e. amphi- ratios for eclogite are similar to those of the basalts. In bole). Higher Nb/Y ratios in WARS basalts correlate general, the results demonstrate that the overall trend with the higher values anticipated for melting of carbo- of increasing Nb concentration from alkali basalt to bas- nated peridotite sources (Herzberg & Asimow, 2008) anite to tephrite could be explained by melting to a and the higher Ne þ Lc normative contents, and are smaller degree of an enriched mantle source, which in similar to experimental results from high-degree melt- the case of the NWRS occurs oceanward, as indicated ing of hornblendite (Figs 13 and 15a and b). As by the more common occurrence of strongly silica- mentioned above, Nb can reach high concentrations in undersaturated compositions (i.e. basanite and tephrite) amphibole and in mantle that has experienced carbona- away from the continent (Figs 1b and 10). There are two fundamental short-falls with this hy- tite metasomatism. Furthermore, the compatibility of Nb in amphibole is mostly less than unity whereas Y is pothesis. First is the difficulty in matching compositions amph/melt (both major and trace elements) derived by low-degree compatible [D ¼ 053 and 148, respectively; melting of eclogitic and peridotitic sources to those of average values calculated by Pilet et al. (2011) using data from Tiepolo et al. (2000a, 2000b, 2007)]. Thus we low-Si alkaline basalts (< 45 wt % SiO2) as emphasized by Pilet (2015). Although models of low degrees of melt- interpret the higher Nb/Y ratios in basanite and tephrite ing ( 3%) of eclogite produce similar high field as indicating greater contributions from amphibole-rich strength element (HFSE; e.g. Nb, Zr, Ti) concentrations, (6 CO2) metasomes within the lithospheric mantle. In they also generate significantly higher large ion litho- the case of the NWRS, the signature becomes more phile element (LILE; e.g. Rb, Ba, and K) and Pb concen- prominent oceanward (Fig. 15c). trations and lower HREE (Yb and Lu) and Y concentrations compared with the basalts. The melting Melt genesis between ocean and continent of metasomatized peridotite (H-group) yields a better We propose that the geochemical signature of mafic al- match for LILE and HFSE, but the Ti concentrations pro- kaline basalts in the NWRS is a consequence of high- duced are too low (< 1 wt %) relative to the basalts (20– degree melting of amphibole-bearing metasomatic 48wt%,Table 2) and thus give much lower Ti/Zr ratios veins in both oceanic and continental lithosphere and (Fig. 14c). Second is the difficulty in explaining the sys- that the progressive interaction of the melt with the sur- tematic variation in major and trace elements of the rounding mantle (see Pilet et al., 2008), occurring to a NWRS basalts across the continent to ocean transition greater extent continentward, provides the best explan- (Fig. 10) by changes in the degree of partial melting. ation for the systematic, gradational (i.e. non-stepwise) This is problematic as the lithospheric thickness across variation in major and trace element and isotopic data

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 544 Journal of Petrology, 2018, Vol. 59, No. 3

Fig. 15. Compositional parameters and geographical locations (NWRS) of basalts plotted vs Nb/Y ratios. Symbols and data criteria for the basalts are the same as those given in Fig. 3. (a) Volatilized peridotite index [¼ CaO – (2318SiO2 –93626)] defined by Herzberg & Asimow (2008) as a line on a CaO vs SiO2 diagram that separates magmas derived from carbonated peridotite (Dasgupta et al., 2007), values > 0, and magma derived from carbonate-free peridotite (Walter, 1998; Herzberg, 2004, 2006), values < 0. Experimental melts of amphibole-rich sources (Pilet et al., 2008) are also shown. Line of regression excludes tephrite sample D9-1 (solid green triangle). (b) Normalized (CIPW) values of basalts and experimental melts. Line of regression excludes sample MA009a. (c) Longitude of NWRS basalts vs Nb/Y (refer to Fig. 8 for criteria for geographical sectors). Line of regression excludes tephrite D9-1 and a basanite from the continental sector (MP24, Nardini et al., 2009). (d, e) Nd and Sr isotope compositions vs Nb/Y. Tephrite D9-1 is excluded from the regression in both plots and MA009a is also excluded in (e).

(Figs 10 and 15c). As briefly outlined above, the multi- element composition of amphibole cumulates is not de- step process to produce low-Si alkaline basalts pro- pendent on the composition of the underlying astheno- posed by Pilet et al. (2004, 2005, 2008, 2010, 2011) and sphere. The major element composition, including K Pilet (2015) requires lithospheric enrichment by small- and Ti, of amphibole-bearing cumulates is controlled by degree melts from sub-lithospheric mantle sources, fol- the stoichiometry of the minerals, whereas the trace lowed by high-degree melting of the resultant element content is controlled mostly by the exchange min/melt amphibole-rich metasomes, which in turn react with the equilibrium between minerals and liquid (KD ). surrounding peridotite during melt migration. There are A second vital feature of their model is that unlike major several important aspects of the Pilet et al. model that and trace elements, the isotopic composition of the cu- need to be emphasized and aligned with what is being mulate is inherited from the underlying asthenosphere. proposed here. First, their model demonstrates (theor- Therefore it is suggested that the isotopic signature of etically and experimentally) that the major and trace the most Si-undersaturated and incompatible element

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 545

enriched basalts with the highest Nb/Y ratios may best isotopic compositions may be displaced from sub- represent the composition of the sub-lithospheric lithospheric sources by acquisition of lithospheric man- source with low 87Sr/86Sr ( 07030), high 143Nd/144Nd tle signatures. The longer residence time required for a ( 05130) and high 206Pb/204Pb ( 20) (Figs 7, 10, 12 and melt, which is generated from amphibole-bearing veins 15d, e). This isotopic ‘endmember’ is characteristic of near the base of the lithosphere, to migrate upward sources for Cenozoic alkaline magmas in West through progressively thicker and older lithospheric Antarctica, as well as alkaline basalts erupted on widely mantle will advance melt–peridotite reactions. scattered continental fragments of East Gondwana Experimental studies show that the reaction between (Zealandia and eastern Australia). As noted above, the Si-undersaturated (i.e. Ne-normative) liquid and perido- signature has been deemed ‘young HIMU’ or more re- tite will cause dissolution of orthopyroxene and the li- cently ‘CarboHIMU’ (McCoy-West et al., 2016) to distin- quid will become enriched in silica while crystallizing guish it from endmember HIMU from oceanic settings olivine (Shaw et al., 1998; Shaw, 1999; Lundstrom et al., (e.g. St Helena) and plots within the DMM–PREMA field 2000). Average olivine–liquid equilibrium temperatures (Fig. 7) defined by Stracke (2012). Indeed, most calculated for the NWRS basalts (Krans, 2013) are researchers agree that the potential mantle sources for 1244 6 40C for anhydrous conditions (Beattie, 1993), this diffuse alkaline magmatic province (DAMP; Finn 1211 6 37C for hydrous conditions (Putirka et al., 2007) et al., 2005) reside in metasomatized continental litho- and 1145 6 29 C for CO2-rich conditions (Sisson & spheric domains, although the timing and cause of Grove, 1993), and do not vary systematically between metasomatism as well as the ultimate source of the iso- continent and ocean. Layered hornblendite and perido- topic signatures remain controversial (Rocchi et al., tite melting experiments performed at 15–25 GPa and 2002; Panter et al., 2006; Timm et al., 2010; Scott et al., 1200–1325C on natural samples by Pilet et al. (2008) 2014; McCoy-West et al., 2010, 2016; van der Meer demonstrate that this reaction will evolve liquids with et al., 2017). The results presented here and by Kipf lower alkalinity, TiO2, CaO, FeO and incompatible trace et al. (2013) confirm that PREMA (i.e. young HIMU or element contents. It will also produce slightly higher CarboHIMU) isotopic signatures are also present in MgO and Al O contents, while maintaining relatively basalts erupted through oceanic lithosphere. 2 3 constant K O/Na O and Al O /TiO ratios. These fea- As proposed earlier, the trace element and radiogen- 2 2 2 3 2 tures are consistent with the major and trace element ic isotopic signatures of PREMA and HIMU originate trends from nephelinite/basanite to alkali basalt (41– from previously subducted slab materials. Specifically, 50 wt % SiO ; Fig. 3) that are documented in this study. the endmember HIMU may owe its highly radiogenic 2 Alternatively, increasing the degree of partial melting of Pb, low 87Sr/86Sr, and distinctive trace element signa- a common mantle source has been used to explain this tures to Archaean marine carbonates (Castillo, 2015, compositional continuum observed in both oceanic and 2016; Weiss et al., 2016). The variable and less radio- continental lavas (Frey et al., 1978; Caroff et al., 1997; genic Pb plus slightly higher 87Sr/86Sr of PREMA, on the Beccaluva et al., 2007; Bosch et al., 2014). However, fol- other hand, come from subducted lithospheric mantle lowing the arguments that we have presented above, that was carbonated, or ‘spiked’ by a lesser amount of Archaean or possibly younger marine carbonates. we consider the reaction of Si-undersaturated liquids Notably, our proposal is consistent with recent results (i.e. nephelinitic/basanitic) with surrounding peridotite indicating that alkaline and carbonatitic rocks from the in the lithosphere to be a more plausible explanation for northern and eastern regions near the edges of the the systematic geochemical and isotopic variation in North China Craton acquire some of their distinctive the NWRS basalts from ocean to continent. compositional features from subducted marine carbo- Further support for this assertion is provided by the nates (Chen et al., 2016; Li et al., 2017, and references olivine oxygen isotope data. It should be recalled 18 therein). These regions are tectonically similar to WARS that d Ool values for NWRS basalts are highly variable in that they were previous convergent margins. (D 15&; Fig. 8a) and are not adequately explained by 18 In the NWRS, measured 87Sr/86Sr increases whereas crustal contamination (Fig. 12). High d Ool values cor- 143Nd/144Nd and 206Pb/204Pb decrease in basalts erupted relate with the lower Nb/Y ratios of corresponding with decreasing distance from the continent (Fig. 10). whole-rock compositions (Fig. 16a) and trend towards The overall age of the volcanism also increases conti- mineral values for orthopyroxene separated from peri- nentward; however, applying an age-correction for dotite xenoliths hosted by NVL basalts (Perinelli et al., radiogenic in-growth cannot explain this variation (i.e. 2006). The correlation is consistent with reaction be- 18 corrections are < 5 10–5 for all three isotopic systems). tween low d O liquid generated by melting of What is changing is the thickness and age of the litho- amphibole-rich metasomes (i.e. high Nb/Y source) and sphere, both of which increase from the Adare Basin to- peridotite, where dissolution of orthopyroxene and wards the East Antarctic craton. Additionally, as crystallization of olivine evolve the liquid towards more inferred from the correlation between isotopic and Nb/Y Si-saturated (Fig. 16b) and 18O–enriched compositions. ratios (Fig. 15d and e), the basalts erupted closer to the Moreover, variations in trace elements, including Th

continent are interpreted to have less contribution (i.e. and La, with SiO2 content (Fig. 16c and d) parallel the lower Nb/Y ratios) from metasomes and thus their trends produced by the melting and reaction

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 546 Journal of Petrology, 2018, Vol. 59, No. 3

Fig. 16. (a) Nb/Y ratio vs d18O(&). Oxygen isotopes for olivine phenocrysts from NWRS basalts were measured by LF and SIMS analysis (Table 3). Only samples with whole-rock MgO values greater than 6 wt % are plotted. Open symbols represent LF analyses of olivine from Nardini et al. (2009). In this study multiple grains of olivine in the same sample were analyzed by SIMS and averaged (solid colored symbols). Also shown are the average of multiple SIMS spot analyses (Core*) within the cores of individual olivine grains for several samples (symbols filled with cross). The Nb/Y ratios represent whole-rock values of the basalts that host the oliv- ine. Orthopyroxene and vein amphibole d18O and Nb/Y ratios measured on peridotite xenoliths from NVL are from Perinelli et al. (2006). Whole-rock gabbro values provided by Dallai et al. (2003) for rocks from NVL. Error bars represent standard error of mean (SEM). (b) SiO2 (wt %) vs Nb/Y ratios (symbols and data criteria of basalts are the same as in Fig. 3) are compared with the mineral chemistry of orthopyroxene and amphibole from peridotite xenoliths hosted in alkaline basalts from NVL (Coltorti et al., 2004; Perinelli et al., 2006). Whole-rock eclogite from NVL provided by Di Vincenzo et al. (1997) and average eclogite composition for whole-rocks from Zambia are from John et al. (2004). Line of regression (r2 ¼ 052, Spearman rank ¼ 068) for Antarctic basalts excludes sample D9-1 (solid green triangle). SiO2 (wt %) vs Th (c) and La (d) ppm of basalts are compared with natural peridotite, eclogite and gabbro as in (b), along with results of melting experiments conducted by Pilet et al. (2008). The experimental data are interpreted to represent changes in degree of partial melting or source heterogeneity (solid gray arrow) and reaction between basa- nitic melt and peridotite (dashed black arrow). These trends are emulated by basanite and tephrite/nephelinite (red arrow) and alkali basalt (blue arrow) in this study.

experiments of Pilet et al. (2008), providing additional for mixing must be able to emulate the systematic geo- support for this hypothesis. chemical variation from land to sea in the NWRS, which An alternative explanation, however, is that the would require the proportion of pyroxenite to increase trends may be a result of mixing between Si- towards the continent. Here again we favor the melt– undersaturated melt and a liquid generated by the par- peridotite reaction scenario in that it is better con- tial melting of recycled pyroxenite (pyroxene þ garnet). strained by the evidence and provides a much more Mixing proportions require > 20–80% of gabbro/eclog- straightforward model (compare Occam’s razor) for ite to account for the compositional range of the basalts NWRS volcanism, yet some contribution from recycled (Figs 11, 12 and 16a, b). Based on experimental and pyroxenite may be warranted. Melchiorre et al. (2011) thermodynamic constraints, partial melting of volatile- called upon the melting of a sub-lithospheric source free pyroxenite cannot explain the major element com- containing between 15 and 60% eclogite to explain the

positions of the low-SiO2 alkaline basalts (Lambart radiogenic Os and Hf isotope values measured in meta- et al., 2013; Pilet, 2015, and references therein), al- somatized peridotite xenoliths from the NVL. Evidence though homogenization with volatile-rich Si-under- from xenoliths, seismic and geodynamic investigations saturated melt may have some merit. Even so, a model indicate that recycled volatiles and eclogitic material

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 547

are likely to have been introduced into the sub- upon partial melting of metasomatized mantle perido- lithospheric mantle beneath West Antarctica during tite that contains 10% recycled pyroxenite/eclogite, it- Paleozoic–Mesozoic Gondwana subduction (Finn et al., self a product of high-temperature alteration of oceanic 2005; Sutherland et al., 2010; Melchiorre et al., 2011; crust and mantle, to account for the correlation between 18 206 204 Emry et al., 2015; Martin et al., 2015; Broadley et al., low d Ool (487 6 018&) and high Pb/ Pb (values 2016). up to 20) in basalts from La Palma, Canary Islands. 18 d Ool values for NWRS basalt display a weak correl- Subduction-related agents in metasomes ation with 206Pb/204Pb, as well as Sr and Nd isotopes 18 The lower oxygen isotope values measured on olivine (Fig. 9) and indicate that the low d O signature is asso- 87 86 143 144 (d18O 5&) in the NWRS basalts are most probably a ciated with the low Sr/ Sr, high Nd/ Nd and high 206 204 signature derived from hydrothermally altered sub- Pb/ Pb. As discussed above, the origin of radiogen- ducted materials. Metasomatism of the NVL lithosphere ic Pb signatures (> 195) in mafic alkaline rocks found by seawater-derived volatiles is indicated by heavy throughout West Antarctica and continental fragments halogen (Br, I) and noble gas (Ar, Kr, Xe) compositions of Gondwana (DAMP; Finn et al., 2005) is controversial liberated from fluid inclusions in olivine and pyroxene and has been related to both lithospheric and sub- 206 204 in peridotite xenoliths (Broadley et al., 2016). Broadley lithospheric sources. The highest Pb/ Pb ratios et al. (2016) suggested that the volatiles were released, (> 207) measured in mafic alkaline rocks are from con- possibly at depths up to 200 km (Sumino et al., 2010; tinental basalts in Marie Byrd Land and Chatham Island Kendrick et al., 2013), during subduction and were (Panter et al., 2000, 2006) and values greater than 21 incorporated into the overlying mantle wedge and sub- have been measured on spinel peridotite xenoliths continental lithospheric mantle beneath the Gondwana from Chatham Island and the Waitaha domain of south- ern New Zealand (McCoy-West et al., 2016). For basalts continental arc during the Paleozoic. Their investigation within the NWRS the highest 206Pb/204Pb ratios (> 20) along with those of Coltorti et al. (2004), Perinelli et al. occur oceanward (Fig. 10) and, as discussed above, we (2006, 2011) and Melchiorre et al. (2011) provide strong consider that this signature, along with radiogenic Nd support for metasomatism of the lithosphere by sub- and depleted Sr and O isotopic values, was ‘originally’ duction zone processes and contributions from recycled derived from recycled crustal materials and transferred subducted material to the source of the alkaline mag- by small-degree melts to the lithosphere and stored in mas in the NVL, including the NWRS. amphibole-rich cumulates (i.e. metasomes). The higher Mafic eclogites formed by past subduction of ocean- 206Pb/204Pb ratios found in some continental basalts ic lithosphere have highly variable oxygen isotopic and xenoliths scattered across the DAMP may be compositions that range from extremely heavy explained by compositional heterogeneity of the (d18O > þ10&) to extremely light (d18O < –10&) values recycled material in the sub-lithospheric mantle, melt- for mineral and whole-rock samples (Zheng et al., 1998, ing of higher proportions of the recycled material (see 2003; Putlitz et al., 2000; Fru¨ h-Green et al., 2001). The 18 Day et al., 2010), and/or may be due to rapid in-growth variability of d O is explained by the exchange be- facilitated by high U/Pb ratios (McCoy-West et al., 2016). tween fluids (meteoric or seawater d18O 0&) and rock 18 The fractionation required to produce high U/Pb has (d Obasalt/gabbro/peridotite > 5–7&) over a wide range of been explained by compatibility differences between temperatures and fluid/rock ratios in low- to high-pres- these elements during carbonatite metasomatism sure environments during the subduction cycle. (McCoy-West et al., 2016) or during post-metasomatic Generally, lower temperature hydrothermal alteration partial dehydration (Panter et al., 2006). within the upper oceanic crust will enrich rocks in Importantly, the fractionation of oxygen isotopes 18 16 O/ O whereas higher temperature ( 250 C) ex- (18O/16O) between silicate liquids and silicate minerals change within the lower crust and upper mantle will de- at asthenospheric temperatures ( 1350C) or tempera- 18 16 plete rocks in O/ O(Muehlenbachs, 1986). Low tures in the lithospheric mantle [> 900–1100Cat 18 oxygen isotope values (d O 5&) measured on whole- 09–17 GPa; spinel peridotite stability determined for rocks and clinopyroxene in veins and eclogitized meta- continental NVL by Armienti & Perinelli (2010)] would gabbro from the Erro–Tobbio peridotite, Italian Alps, be negligible. Therefore if the increase in 206Pb/204Pb were explained by Fru¨ h-Green et al. (2001) to be a con- observed in the NWRS basalts oceanward (Fig. 10)isa sequence of high-temperature alteration of lower result of radiogenic in-growth, then the correlation with oceanic crust and upper mantle by seawater that has d18O(Fig. 9) would not be expected (i.e. decoupled sys- been preserved (i.e. lack of oxygen isotopic re- tems). Once again, we regard the deviation in oxygen equilibration with ambient mantle) through eclogitiza- and radiogenic isotopes from the ocean ‘endmember’ tion and limited fluid circulation within a relatively of the NWRS to be a consequence of the reaction be- closed system at high pressures ( 2 GPa). tween silica-undersaturated melts derived from meta- In HIMU ocean island basalts, low oxygen values somes and surrounding peridotite, increasing as melt 18 (d Ool 47–52&) are also interpreted to be a conse- migrates through thicker mantle lithosphere continent- quence of long-term recycling of subducted materials ward. The isotopic signatures of the metasomes, on the (Eiler et al., 1996; Eiler, 2001). Day et al. (2010) called other hand, were ultimately formed from recycled

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 548 Journal of Petrology, 2018, Vol. 59, No. 3

oceanic crustal and mantle materials that were trans- and is in accordance with the time offset between rifting ferred to the asthenosphere through carbonatitic sili- and magmatism discussed above. Rocchi et al. (2002, cate melts derived from the down-going slab (Castillo, 2005) and Nardini et al. (2009) suggested that during the 2015, 2016). initial Late Cretaceous phase of WARS extension small- volume melts from the asthenosphere produced Relative timing of extension, metasomatism and amphibole-rich metasomatic veins that become a litho- magmatism spheric mantle source for subsequent alkaline magma- There are significant time offsets between major epi- tism from the middle Eocene to Present. The timing is sodes of extension and alkaline magmatism in West probably controlled by the thermal gradient in the litho- Antarctica. The earliest magmatism, the Meander sphere, which must reach temperatures of 1150–1175C Intrusive Group (48 to 23 Ma), in NVL followed to allow melting of amphibole-rich cumulates (Pilet Gondwana break-up and a broad phase of rifting within et al., 2008). Thermobarometric calculations for perido- the WARS (105 to 80 Ma) by 30 Myr. A focused phase tite xenoliths (Armienti & Perinelli, 2010; Perinelli et al., of continental extension within the WARS (80 to 40 Ma) 2011) reveal a change in the geothermal gradient from occurred 25 Myr prior to the activity that formed the 05to3Ckm–1 in NVL lithosphere. The thermal evolu- larger shield volcanoes along the continental coastline tion is estimated to have taken 10 Myr and is credited in the NWRS (14 to > 5 Ma). Monogenetic island and to the development of edge-driven mantle convection seamount volcanism on the continental shelf and Adare along the boundary between the thinned lithosphere of Basin (< 5to< 100 ka) followed extension in the the WARS and the thick East Antarctic craton estab- Northern Basin and seafloor spreading in the Adare lished by the Eocene (Faccenna et al., 2008). Basin (43 to 26 Ma) by 20 Myr. Although the estimates Here we extend the model for metasomatism to in- are loosely constrained, it appears that progressively clude the oceanic portion of the NWRS and call upon shorter time periods of offset between extension and the influence of younger, mostly amagmatic rifting to magmatism (i.e. 30 ! 25 ! 20 Myr) are spatially coin- generate the low-degree sub-lithospheric melts that cident with the broad shift in the location and volume of have enriched the lithosphere. The ultraslow spreading volcanism, as well as with a decrease in post- –1 in the Adare Basin (12 mm a ; Cande et al., 2000) and extensional lithospheric thickness to the northeast with- its crustal characteristics, which are consistent with in the NWRS. oceanic core complexes observed at ultraslow- The timing of metasomatism of the mantle litho- spreading ridges (Selvans et al., 2014), probably facili- sphere in West Antarctica and New Zealand has been tated the repose between the melting that formed estimated based on model ages from isotopes. Lead amphibole-rich veins and the melting of these meta- model ages evaluate the in-growth of high 206Pb/204Pb somes to produce alkaline volcanism. Studies of ultra- ratios ( 20) starting from a PREMA–FOZO composition slow seafloor spreading (Dick et al., 2003; Michael et al., ( 195) with variable U/Pb ratios. Estimates range from 2003; Cannat et al., 2006) reveal that mantle melting 500 to 200 Ma (Hart et al., 1997; Panter et al., 2000) and and the resulting volcanism is not a simple function of 150 to 50 Ma (Nardini et al., 2009). Model ages calcu- spreading rate but that mantle temperatures and mantle lated for spinel peridotites from southern New Zealand lithology/composition are also controlling factors. reveal metasomatic enrichments of 180 Ma (McCoy- West et al., 2016), including extremely rapid in-growth Ultraslow spreading will lead to cooler temperatures over 35 to 13 Ma for some samples with very high with depth and therefore thicker lithosphere (Reid & 238U/204Pb ratios (m ¼ 63–466). Melchiorre et al. (2011) Jackson, 1981). Moreover, temperatures for sub- presented Os model ages (time of Re depletion) for lithospheric mantle beneath ultraslow-spreading ridges interstitial sulfides in metasomatized peridotite xeno- (70–80 km depth) may be colder by 180 C relative to liths from NVL that span the evolutionary history of mantle beneath fast-spreading ridges (Husson et al., Antarctica from the Archean–Proterozoic up to the 2015). We suggest that the temperature at the base of Cretaceous, but speculated that the high proportions of the oceanic lithosphere beneath the Adare Basin was eclogite required to explain the radiogenic 187Os/186Os cool enough to ‘freeze-in’ low-degree melts, and the could have been introduced into the sub-lithospheric metasomatic cumulates that they produced, during or mantle by subduction during the Ross Orogeny (550– soon after seafloor spreading (43–26 Ma). Furthermore, 600 Ma). They concluded that the ancient eclogitic sig- we propose that the delay in melting of the metasomes natures were probably incorporated into partial melts leading to seamount volcanism in the Adare Basin (<5 during the initial phase of WARS extension. Ma to <200 ka) was controlled by the rate of conductive According to Pilet (2015), the difference between heating from the base of the lithosphere. Basal heating how alkaline magmas are generated in continental ver- was enhanced by regional mantle upwelling related to sus oceanic settings is that in continental settings the subduction death and the sinking of the Pacific slab into episode of metasomatic enrichment may be decoupled the lower mantle (Finn et al., 2005; Sutherland et al., from the events that cause melting of the metasomes. 2010) and possibly more recent (Neogene) edge-driven This scenario has been proposed in continental NVL flow (Faccenna et al., 2008).

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 549

(a)(Pre-break-up subduction (> 100 Ma) d) Adare/Northern Basin extension (43 to 26 Ma)

WEarc volcanism W Northern Victoria Land Northern Basin Adare Basin E TAM 0 oceanicoceanic ccrustrust 0 ccontinentalontinental ccrustrust oceanicoceanic crustcrust oceanic OCTZOCTZ ccontinentalontinental ccrustrust lithospheric mantle 50 20 asthenosphere 40 100

continental ) melt-peridotite Depth (km) lithospheric mantle flux melting 60 reaction km 150 from ( dehydrating slab 80 850°C pth

change e D in scale 100 lithospheric mantle 200 carbonate-rich asthenosphere recycled 120 slab material 50°C detached slab 11 140 upwelling 300 km 30 km 500 mantle 160

(b)(Initiation of WARS extension, phase I (100 to 80 Ma) e) Post Adare Extension I (26 - 5 Ma)

WEWENorthern Victoria Land Northern Basin Adare Basin Adare trough 0 0 EEastast AntarcticaAntarctica WWestest AAntarcticantarctica ccontinentalontinental ccrustrust ccontinentalontinental crustcrust OCTZOCTZ 20 50

m) 40 metasomes lithospheric mantle ) 100 850°C 60 km ( Depth (k

h °C t 80 850 1150°C 150 ep D lithospheric mantle low degree melting of 100 carbonate-rich asthenosphere asthenosphere 200 recycled slab material 120 C 1150° 140 300 km 30 km 250 160

(c)(WARS extension phase II (80 to 40 Ma) f) Post Adare Extension II (< 5 Ma)

WEWENorthern Victoria Land Northern Basin Adare Basin TAM Meander Intrusive group 0 0 EEastast AntarcticaAntarctica ccontinentalontinental ccrustrust WWestest AAntarcticantarctica ccontinentalontinental ccrustrust OCTZOCTZ 20 50 lithospheric mantle high degree melting 40 of metasomes 100 850°C 60 pth (km) km) (

De 0°C C 115 h 850° t 80 150

low degree melting of Dep lithospheric mantle carbonate-rich recycled 100 slab material asthenosphere asthenosphere 120 200 C 1150° 140 300 km 30 km 250 160

Fig. 17. Schematic illustration of Late Cretaceous to recent tectonism and the genesis of alkaline magmatism in the NWRS. (a) Pre- breakup continental arc. The subduction tectonic regime was nearly continuous from the late Neoproterozoic (550 Ma) to Late Cretaceous (100 Ma) along the Paleo-Pacific margin of Gondwana (Bradshaw, 1989; Cawood, 2005) and provided slab material to the asthenosphere either directly or recycled from depth in mantle upwellings. Deeper mantle upwelling may have also delivered more ancient subducted material to the asthenosphere (Castillo, 2016). (b) Broad regional extension (100 to 80 Ma) of Antarctic continental lithosphere (DiVenere et al., 1994; Luyendyk et al., 1996) caused decompression melting of slab material to a small de- gree. The resultant carbonate-rich silicate liquids rose and froze within the cooler lithospheric mantle to produce metasomes (amphibole-rich veins; see Pilet et al., 2008). The isotherms shown in this panel and in what follows are approximations that are in part constrained by the geothermal gradient estimated for NVL continental lithosphere by Armienti & Perinelli (2010). (c) Focused extension (80 to 40 Ma) between East and West Antarctica caused incipient lithospheric necking (Huerta & Harry, 2007) and uplift of the Transantarctic Mountains (TAM; Fitzgerald et al., 1986). Heating of the lithosphere at its base was intensified by mantle up- welling and the higher temperatures required to melt the metasomes (>1150C) to a high degree occurred by about 50 Ma, which initiated the oldest know alkaline igneous activity (Meander Intrusive Group, 48–23 Ma; Rocchi et al., 2002) associated with the WARS. The silica-undersaturated liquid that was produced reacted with the surrounding peridotite as it traversed upward through the lithosphere [see (d)]. (d) (note the change in scale). Ultraslow seafloor spreading formed the oceanic lithosphere of the Adare Basin from 43 to 26 Ma (Cande et al., 2000). The continuity of magnetic, seismic and structural trends from the Adare Basin into the Northern Basin (Cande & Stock, 2006; Davey et al., 2006; Damaske et al., 2007; Ferraccioli et al., 2009; Selvans et al., 2012) indicates that the oceanic crust extends across the continental shelf break (Granot et al., 2013; Selvans et al., 2016) and thus the transition in lithospheric type may be gradational from ocean to continent (OCTZ). Edge-driven convective flow was established at the boundary between thinned lithosphere and the thick East Antarctic craton (Faccenna et al., 2008). Metasomes were formed in the cooler OCTZ and oceanic lithosphere while the earlier-formed metasomes in the warmed continental lithosphere continued to melt sup- plying magma for the Meander Intrusive Group until 23 Ma. (e) Minor post-spreading extensional events in the Adare Basin occurred at 24 and 17 Ma and formation of the Adare Trough occurred (Granot et al., 2010, 2013). Alkaline magmatism resumed after an 10 Myr hiatus to form shield volcano complexes (14 to > 5 Ma) located along rift-shoulder fault systems. (f) Thermal evo- lution of the lithosphere oceanward reached the melting temperature of metasomes by 5 Ma to produce small-volume volcanic seamounts on the continental shelf and in the Adare Basin.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 550 Journal of Petrology, 2018, Vol. 59, No. 3

Integrated model for magmatism boundary between continental and oceanic lithosphere, The model we use to explain the origin of alkaline mag- and is further constrained by the dynamic thermal–tec- matism in the NWRS is illustrated in Fig. 17. Our model tonic evolution of the region. The contribution of litho- is intimately connected to what is known of the tectonic spheric mantle, both continental and oceanic, as a evolution of the region beginning at a time prior to the source and a contaminant of Si-undersaturated melts is break-up of the proto-Pacific margin of Gondwana important to understand and should be carefully con- (Fig. 17a). The protracted history of subduction along sidered in studies of alkaline basalts worldwide. this margin is considered to have supplied carbonate- rich slab material directly or by longer-term recycling into the upper asthenosphere; however, our model ACKNOWLEDGEMENTS does not preclude the contribution from more ancient We would like to thank Anne Grunow from the US Polar subduction-related sources. A series of extensional Rock Repository (supported by NSF OPP 1141906) and phases occurred in the development of the WARS, Nick Mortimer (GNS Science) for providing some of the which caused lithospheric thinning and decompression land-based samples used in this study. We are especial- melting of this recycled material in the asthenosphere, ly grateful to the Captain and crew of RV/IB Nathanial B. probably aided by heating from passive mantle upwell- Palmer for the successful dredging campaign in the ing and edge-driven flow (Fig. 17b–e). The carbonate- Adare Basin. We thank Mike Spicuzza for analysis of rich silicate melt produced by these small degrees of oxygen isotope ratios at University of Wisconsin– melting did not reach the crust but crystallized in the Madison by laser fluorination and gas-source mass mantle lithosphere to form amphibole-rich cumulates spectrometry. Gordon Moore is thanked for his guid- (metasomes) that later became the source of alkaline ance with electron microprobe analysis at the magmatism. The metasomatic enrichment of the litho- University of Michigan. We are grateful for the thought- sphere is invoked in our model not only to explain ful and constructive reviews provided by Joshua some of the geochemical and isotopic characteristics of Schwartz, Sebastien Pilet and Karsten Haase, as well as the basalts and xenoliths described in this and other the editorial handling by Georg Zellmer. Also, we are studies, but also to explain why major periods of exten- appreciative of the helpful comments provided by sion are not concurrent with, or immediately followed Tyrone Rooney and Adam Martin during the writing of by, magmatism. Other geodynamic evidence for litho- this paper. spheric mantle sources for alkaline magmatism found on former pieces of Gondwana (DAMP; Finn et al., 2005) has been provided by Panter et al. (2006). They FUNDING documented the occurrence of relatively uniform mafic This study was supported by National Science alkaline compositions on Chatham Island that were Foundation grants ANT 0943503, ANT 0943274 and OPP derived from an amphibole-bearing mantle source and 05-38374. P.R.K. acknowledges support from NSF grant erupted intermittently over an 80 Myr period (85 to ANT 1141534. WiscSIMS is supported by NSF EAR 5 Ma) as the continental lithosphere of Zealandia rifted 1355590 and the University of Wisconsin–Madison. away from West Antarctica and drifted north over a dis- tance > 3000 km. Specific to the NWRS portion of the DAMP is that this region represents a recently rifted SUPPLEMENTARY DATA continental margin across which the lithosphere varies Supplementary data for this paper are available at in both thickness and age (Fig. 17). We have shown that Journal of Petrology online. this is coincident with gradational changes in basalt geochemistry, Sr–Nd–Pb–O isotopes (Figs 10 and 15c) REFERENCES and broadly with eruption age. Our model explains these features by invoking an oceanward progression Abouchami, W., Galer, S. J. G. & Koschinsky, A. (1999). Pb and in melting of the metasomes facilitated by the thermal Nd isotopes in NE Atlantic Fe–Mn crusts: proxies for trace metal paleosources and paleocean circulation. Geochimica evolution of the lithosphere and coupled to a decrease et Cosmochimica Acta 63, 1489–1505. in the extent of reaction between melt and peridotite Armienti, P. & Perinelli, C. (2010). Cenozoic thermal evolution of prior to eruption, which, in turn, is controlled by litho- lithospheric mantle in northern Victoria Land (Antarctica): spheric thickness (Fig. 17c–f). evidences from mantle xenoliths. Tectonophysics 486, This multi-stage model for alkaline volcanism in the 28–35. NW Ross Sea is relevant to interpretations on the origin Armienti, P., Francalanci, L., Landi, P. & Vita, G. (2003). Age and and cause of alkaline magmatism within the rest of the geochemistry of volcanic rocks from Daniell Peninsula and West Antarctic rift system, as well as other continental Coulman Island, Hallett Volcanic Province, Antarctica. In: Tessensohn, F. & Ricci, C. A. (eds) Aspects of a Suture Zone. areas of the southwestern Pacific (i.e. DAMP). Geologisches Jahrbuch, Polar Issue 9, 409–445. Importantly, our model is constructed using contempor- Aviado, K. B., Rilling-Hall, S., Bryce, J. G. & Mukasa, S. B. aneous, petrogenetically related, alkaline magmas that (2015). Submarine and subaerial lavas in the West Antarctic show a transition (compositionally) across the Rift System: temporal record of shifting magma source

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 551

components from the lithosphere and asthenosphere. Castillo, P. R., Clague, D. A., Davis, A. S. & Lonsdale, P. F. Geochemistry, Geophysics, Geosystems 16, 4344–4361. (2010). Petrogenesis of Davidson Seamount lavas and its Beattie, P. (1993). Olivine–melt and orthopyroxene–melt equili- implications for fossil spreading center and intraplate mag- bria. Contributions to Mineralogy and Petrology 115, matism in the eastern Pacific. Geochemistry, Geophysics, 103–111. Geosystems 11, doi:10.1029/2009GC002992. Beccaluva, L., Bianchini, G., Bonadiman, C., Coltorti, M., Milani, Cawood, P. A. (2005). Terra Australis Orogen: Rodinia breakup L., Salvini, L., Siena, F. & Tassinari, R. (2007). Intraplate litho- and development of the Pacific and Iapetus margins of spheric and sublithospheric components in the Adriatic do- Gondwana during the Neoproterozoic and Paleozoic. main: Nephelinite to tholeiite magma generation in the Earth-Science Reviews 69, 249–279. Paleogene Veneto volcanic province, Southern Alps. In: Chakhmouradian, A. R. (2006). High-field-strength elements in Beccaluva, L., Bianchini, G., & Wilson, W. (eds) Cenozoic carbonatitic rocks: geochemistry, crystal chemistry and sig- Volcanism in the Mediterranean Area. Geological Society of nificance for constraining the sources of carbonatites. America, Special Papers 418, 131–152. Chemical Geology 235, 138–160. Behrendt, J. C. (1999). Crustal and lithospheric structure of the Chauvel, C., Hofmann, A. W. & Vidal, P. (1992). HIMU-EM: the West Antarctic Rift System from geophysical investiga- French Polynesian connection. Earth and Planetary Science tions—a review. Global and Planetary Change 23, 25–44. Letters 110, 99–119. Boger, S. D. & Miller, J. M. (2004). Terminal suturing of Chen, C., Liu, Y., Foley, S. F., Ducea, M. N., He, D., Hu, Z., Chen, Gondwana and the onset of the Ross–Delamerian Orogeny: W. & Zong, K. (2016). Paleo-Asian oceanic slab under the the cause and effect of an Early Cambrian reconfiguration of North China craton revealed by carbonatites derived from plate motions. Earth and Planetary Science Letters 219, subducted limestones. Geology 44, 1039–1042. 35–48. Coltorti, M., Beccaluva, L., Bonadiman, C., Faccini, B., Ntaflos, T. Borg, S. G. & Stump, E. (1987). Paleozoic magmatism and asso- & Siena, F. (2004). Amphibole genesis via metasomatic reac- ciated tectonic problems of Northern Victoria Land, tion with clinopyroxene in mantle xenoliths from Victoria Antarctica. In: McKenzie, G. D. (ed.) Gondwana Six: Land, Antarctica. Lithos 75, 115–139. Cooper, A. K., Davey, F. J. & Behrendt, J. C. (1987). Seismic stra- Structure, Tectonics, and Geophysics. Geophysical tigraphy and structure of the Victoria Land Basin, Western Monograph, American Geophysical Union 40, 67–75. Ross Sea, Antarctica. In: Cooper, A. K. & Davey, F. J. (eds) Bosch, D., Maury, R. C., El Azzouzi, M., Bollinger, C., Bellon, H. The Antarctic Continental Margin: Geology and Geophysics & Verdoux, P. (2014). Lithospheric origin for of the Western Ross Sea, CPCEMR Earth Science Series, v. Neogene–Quaternary Middle Atlas lavas (Morocco): clues 5B. Circum-Pacific Council for Energy and Mineral from trace elements and Sr–Nd–Pb–Hf isotopes. Lithos 205, Resources, p. 27. 247–265. Dallai, L., Ghezzo, C. & Sharp, Z. (2003). Oxygen isotope evi- Bradshaw, J. D. (1989). Cretaceous geotectonic patterns in the dence for crustal assimilation and magma mixing in the New Zealand region. Tectonics 8, 803–820. Granite Harbour Intrusives, Northern Victoria Land, Brey, G., Brice, W. R., Ellis, D. J., Green, D. H., Harris, K. L. & Antarctica. Lithos 67, 135–151. Ryabchikov, I. D. (1983). Pyroxene–carbonate reactions in Damaske, D., La¨ ufer, A. L., Goldmann, F., Mo¨ ller, H. D. & Lisker, the upper mantle. Earth and Planetary Science Letters 62, F. (2007). Magnetic anomalies north-east of Cape Adare, 63–74. northern Victoria Land (Antarctica), and their relation to on- Broadley, M. W., Ballentine, C. J., Chavrit, D., Dallai, L. & shore structures. In: Cooper, A. K. & Raymond, C. R., et al. Burgess, R. (2016). Sedimentary halogens and noble gases (eds) Proceedings of the 10th International Symposium on within Western Antarctic xenoliths: implications of exten- Antarctic Earth Sciences. US Geological Survey Open-File sive volatile recycling to the subcontinental lithospheric Report 2007-1047, Short Research Paper 016, p. 5. mantle. Geochimica et Cosmochimica Acta 176, 139–156. Danesi, S. & Morelli, A. (2001). Structure of the upper mantle Busetti, M., Spadini, G., Van der Wateren, F. M., Cloetingh, S. & under the Antarctic Plate from surface wave tomography. Zanolla, C. (1999). Kinematic modelling of the west Antarctic Geophysical Research Letters 28, 4395–4398. rift system, Ross Sea, Antarctica. Global and Planetary Dasgupta, R., Hirschmann, M. M. & Stalker, K. (2006). Change 23, 79–103. Immiscible transition from carbonate-rich to silicate-rich Cande, S. C. & Stock, J. M. (2006). Constraints on the timing of melts in the 3 GPa melting interval of eclogite þ CO2 and extension in the Northern Basin, Ross Sea. In: Futterer, D. K., genesis of silica-undersaturated ocean island lavas. Journal Damaske, D., Kleinschmidt, G., Miller, H. & Tessensohn, F. of Petrology 47, 647–671. (eds) Antarctica: Contributions to Global Earth Sciences. Dasgupta, R., Hirschmann, M. M. & Smith, N. D. (2007). Partial

Springer, pp. 319–326. melting experiments on peridotite þ CO2 at 3 GPa and gen- Cande, S. C., Stock, J. M., Mu¨ ller, R. D. & Ishihara, T. (2000). esis of alkalic ocean island basalts. Journal of Petrology 48, Cenozoic motion between east and west Antarctica. Nature 2093–2124. 404, 145–150. Davey, F. J., Cande, S. C. & Stock, J. M. (2006). Extension in the Cannat, M., Sauter, D., Mendel, V., Ruellan, E., Okino, K., western Ross Sea region—links between Adare Basin and Escartin, J., Combier, V. & Baala, M. (2006). Modes of sea- Victoria Land Basin. Geophysical Research Letters 33, doi: floor generation at a melt-poor ultraslow-spreading ridge. 10.1029/2006GL027383. Geology 34, 605–608. Davy, B., Hoernle, K. & Werner, R. (2008). Hikurangi Plateau: Caroff, M., Maury, R. C., Guille, G. & Cotten, J. (1997). Partial crustal structure, rifted formation, and Gondwana subduc- melting below Tubuai (Austral Islands, French Polynesia). tion history. Geochemistry, Geophysics, Geosystems 9, doi: Contributions to Mineralogy and Petrology 127, 369–382. 10.1029/2007GC001855. Castillo, P. R. (2015). The recycling of marine carbonates and Day, J. M. D., Pearson, D. G., Macpherson, C. G., Lowry, D. & sources of HIMU and FOZO ocean island basalts. Lithos Carracedo, J. C. (2010). Evidence for distinct proportions of 216–217, 254–263. subducted oceanic crust and lithosphere in HIMU-type man- Castillo, P. R. (2016). A proposed new approach and unified so- tle beneath El Hierro and La Palma, Canary Islands. lution to old Pb paradoxes. Lithos 252–253, 32–40. Geochimica et Cosmochimica Acta 74, 6565–6589.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 552 Journal of Petrology, 2018, Vol. 59, No. 3

Deer, W. A., Howie, R. A. & Zussman, J. (1992). An Introduction Pacific without rift or plume origin. Geochemistry, to the Rock-Forming Minerals, 2nd edn. Longman, 696 pp. Geophysics, Geosystems 6, doi:10.1029/2004GC000723. DePaolo, D. J. (1981). Trace element and isotopic effects of Fitton, J. G. & Dunlop, H. M. (1985). The Cameroon line, West combined wallrock assimilation and fractional crystalliza- Africa, and its bearing on the origin of oceanic and continen- tion. Earth and Planetary Science Letters 53, 189–202. tal alkali basalt. Earth and Planetary Science Letters 72, Dick, H. J., Lin, J. & Schouten, H. (2003). An ultraslow-spreading 23–38. class of ocean ridge. Nature 426, 405–412. Fitzgerald, P. G., Sandiford, M., Barrett, P. J. & Gleadow, A. J. DiVenere, V. J., Kent, D. V. & Dalziel, I. W. D. (1994). (1986). Asymmetric extension associated with uplift and Mid-Cretaceous paleomagnetic results from Marie Byrd subsidence in the Transantarctic Mountains and Ross Land, West Antarctica: a test of post-100 Ma relative motion Embayment. Earth and Planetary Science Letters 81, 67–78. between East and West Antarctica. Journal of Geophysical Fretwell, P., Pritchard, H. D., Vaughan, D. G., et al. (2013). Research 99, 15115. Bedmap2: improved ice bed, surface and thickness datasets Di Vincenzo, G. & Rocchi, S. (1999). Origin and interaction of for Antarctica. Cryosphere 7, 375–393. mafic and felsic magmas in an evolving late orogenic set- Frey, F. A., Green, D. H. & Roy, S. D. (1978). Integrated models ting: the Early Paleozoic Terra Nova Intrusive Complex, of basalt petrogenesis: a study of quartz tholeiites to olivine Antarctica. Contributions to Mineralogy and Petrology 137, melilitites from south eastern Australia utilizing geochem- 15–35. ical and experimental petrological data. Journal of Di Vincenzo, G., Palmeri, R., Talarico, F., Andriessen, P. A. M. & Petrology 19, 463–513. Ricci, G. A. (1997). Petrology and geochronology of eclogites Fru¨ h-Green, G. L., Scambelluri, M. & Vallis, F. (2001). O–H iso- from the Lanterman Range, Antarctica. Journal of Petrology tope ratios of high pressure ultramafic rocks: implications 38, 1391–1417. for fluid sources and mobility in the subducted hydrous Di Vincenzo, G., Grande, A. & Rossetti, F. (2014). Paleozoic silici- mantle. Contributions to Mineralogy and Petrology 141, clastic rocks from northern Victoria Land (Antarctica): prov- 145–159. enance, timing of deformation, and implications for the Gao, Y., Vils, F., Cooper, K. M., Banerjee, N., Harris, M., Hoefs, Antarctica–Australia connection. Geological Society of J., Teagle, D. A. H., Casey, J. F., Elliott, T., Laverne, C., Alt, J. America Bulletin 126, 1416–1438. C. & Muehlenbachs, K. (2012). Downhole variation of lithium Di Vincenzo, G., Horton, F. & Palmeri, R. (2016). Protracted and oxygen isotopic compositions of oceanic crust at East (30 Ma) eclogite-facies metamorphism in northern Victoria Pacific Rise, ODP Site 1256. Geochemistry, Geophysics, Land (Antarctica): implications for the geodynamics of the Geosystems 13, doi:10.1029/2012GC004207. Ross/Delamerian Orogen. Gondwana Research 40, 91–106. Garcia, M. O., Ito, E. & Eiler, J. M. (2007). Oxygen isotope evi- Eiler, J. M. (2001). Oxygen isotope variations of basaltic lavas dence for chemical interaction of Kilauea historical magmas and upper mantle rocks. In: Valley, J. W. & Cole, D. (eds) with basement rocks. Journal of Petrology 49, 757–769. Stable Isotope Geochemistry. Mineralogical Society of Genske, F. S., Beier, C., Haase, K. M., Turner, S. P., Krumm, S. & America and Geochemical Society, Reviews in Mineralogy Brandl, P. A. (2013). Oxygen isotopes in the Azores islands: and Geochemistry 43, 319–364. crustal assimilation recorded in olivine. Geology 41, Eiler, J. M., Farley, K. A., Valley, J. W., Hofmann, A. W. & 491–494. Stolper, E. M. (1996). Oxygen isotope constraints on the Gerbode, C. & Dasgupta, R. (2010). Carbonate-fluxed melting of sources of Hawaiian volcanism. Earth and Planetary Science MORB-like pyroxenite at 2–9 GPa and genesis of HIMU Letters 144, 453–468. ocean island basalts. Journal of Petrology 51, 2067–2088. Elliot, D. H. & Fleming, T. H. (2004). Occurrence and dispersal of Ghiribelli, B. (2000). Evoluzione tettonica e metamorfica del magmas in the Jurassic Ferrar large igneous province, margine attivio del Gondeana: Lanterman e Salamander Antarctica. Gondwana Research 7, 223–237. Ranges (Antartide). PhD thesis, Universita` degli Studi di Elliot, D. H., Fleming, T. H., Kyle, P. R. & Foland, K. A. (1999). Siena, Siena, 127 pp. Long-distance transport of magmas in the Jurassic Ferrar Granot, R., Cande, S. C., Stock, J. M., Davey, F. J. & Clayton, R. Large Igneous Province, Antarctica. Earth and Planetary W. (2010). Postspreading rifting in the Adare Basin, Science Letters 167, 89–104. Antarctica: regional tectonic consequences. Geochemistry, Emry, E. L., Nyblade, A. A., Julia` , J., Anandakrishnan, S., Aster, Geophysics, Geosystems 11, doi:10.1029/2010GC003105. R. C., Wiens, D. A., Huerta, A. D. & Wilson, T. J. (2015). The Granot, R., Cande, S. C., Stock, J. M. & Damaske, D. (2013). mantle transition zone beneath West Antarctica: seismic evi- Revised Eocene–Oligocene kinematics for the West dence for hydration and thermal upwellings. Geochemistry, Antarctic rift system. Geophysical Research Letters 40, Geophysics, Geosystems 16, 40–58. 279–284. Ersoy, E. Y., Helvacı, C. & Palmer, M. R. (2010). Mantle source Gurenko, A. A., Hoernle, K. A., Hauff, F., Schmincke, H.-U., Han, characteristics and melting models for the early–middle D., Miura, Y. N. & Kaneoka, I. (2006). Major, trace element Miocene mafic volcanism in Western Anatolia: implications and Nd–Sr–Pb–O–He–Ar isotope signatures of shield stage for enrichment processes of mantle lithosphere and origin lavas from the central and western Canary Islands: insights of K-rich volcanism in post-collisional settings. Journal of into mantle and crustal processes. Chemical Geology 233, Volcanology and Geothermal Research 198, 112–128. 75–112. Faccenna, C., Rossetti, F., Becker, T. W., Danesi, S. & Morelli, A. Gurenko, A. A., Bindeman, I. N. & Chaussidon, M. (2011). (2008). Recent extension driven by mantle upwelling be- Oxygen isotope heterogeneity of the mantle beneath the neath the (East Antarctica). Tectonics Canary Islands: insights from olivine phenocrysts. 27, doi:10.1029/2007TC002197. Contributions to Mineralogy and Petrology 162, 349–363. Ferraccioli, F., Armadillo, E., Jordan, T., Bozzo, E. & Corr, H. Hallett, R. B. & Kyle, P. R. (1993). XRF and INAA determinations (2009). Aeromagnetic exploration over the East Antarctic Ice of major and trace elements in Geological Survey of Japan Sheet: a new view of the Wilkes Subglacial Basin. igneous and sedimentary rock standards. Geostandards Tectonophysics 478, 62–77. and Geoanalytical Research 17, 127–133. Finn, C. A., Mu¨ ller, R. D. & Panter, K. S. (2005). A Cenozoic dif- Hamilton, W. (1972). The Hallett Volcanic Province. US fuse alkaline magmatic province (DAMP) in the southwest Geological Survey Professional Paper 456C, 61 pp.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 553

Harmon, R. S. & Hoefs, J. (1995). Oxygen isotope heterogeneity Ionov, D. A. & Hofmann, A. W. (1995). Nb–Ta-rich mantle of the mantle deduced from global 18O systematics of amphiboles and micas: implications for subduction-related basalts from different geotectonic settings. Contributions to metasomatic trace element fractionations. Earth and Mineralogy and Petrology 120, 95–114. Planetary Science Letters 131, 341–356. Hart, S. R. & Gaetani, G. A. (2006). Mantle Pb paradoxes: the sul- Irvine, T. N. & Baragar, W. R. A. (1971). A guide to the chemical fide solution. Contributions to Mineralogy and Petrology classification of the common volcanic rocks. Canadian 152, 295–308. Journal of Earth Sciences 8, 523–548. Hart, S. R., Hauri, E. H., Oschmann, L. A. & Whitehead, J. A. Isbell, J. L. (1999). The Kukri Erosion Surface; a reassessment of (1992). Mantle plumes and entrainment: isotopic evidence. its relationship to rocks of the Beacon Supergroup in the Science 256, 517–520. central Transantarctic Mountains, Antarctica. Antarctic Hart, S. R., Blijsztajn, J. & Craddock, C. (1995). Cenozoic volcan- Science 11, 228–239. ism in Antarctica; Jones Mountains and Peter I Island. Ivanov, A. V., Meffre, S., Thompson, J., Corfu, F., Kamenetsky, Geochimica et Cosmochimica Acta 59, 3379–3388. V. S., Kamenetsky, M. B. & Demonterova, E. I. (2017). Timing Hart, S. R., Blusztajn, J., LeMasurier, W. E. & Rex, D. C. (1997). and genesis of the Karoo–Ferrar large igneous province: Hobbs Coast Cenozoic volcanism: implications for the West new high precision U–Pb data for Tasmania confirm short Antarctic rift system. Chemical Geology 139, 223–248. duration of the major magmatic pulse. Chemical Geology Hart, S. R., Blusztajn, J., Dick, H. J. B., Meyer, P. S. & 455, 32–43. Muehlenbachs, K. (1999). The fingerprint of seawater circu- Jackson, M. G., Hart, S. R., Saal, A. E., Shimizu, N., Kurz, M. D., lation in a 500-meter section of ocean crust gabbros. Blusztajn, J. S. & Skovgaard, A. C. (2008). Globally elevated Geochimica et Cosmochimica Acta 63, 4059–4080. titanium, tantalum, and niobium (TITAN) in ocean island Hartmann, G. & Wedepohl, K. H. (1990). Metasomatically basalts with high 3He/4He. Geochemistry, Geophysics, altered peridotite xenoliths from the Hessian Depression Geosystems 9, doi:10.1029/2007GC001876. (Northwest Germany). Geochimica et Cosmochimica Acta Janney, P. E. & Castillo, P. R. (1996). Basalts from the Central 54, 71–86. Pacific Basin: evidence for the origin of Cretaceous igneous Hawthorne, F. C., Oberti, R., Harlow, G. E., Maresch, W. V., complexes in the Jurassic western Pacific. Journal of Martin, R. F., Schumacher, J. C. & Welch, M. D. (2012). IMA Geophysical Research: Solid Earth 101, 2875–2893. report: nomenclature of the amphibole supergroup. John, T., Scherer, E. E., Haase, K. & Schenk, V. (2004). Trace American Mineralogist 97, 2031–2048. element fractionation during fluid-induced eclogitization in Heeszel, D. S., Wiens, D. A., Anandakrishnan, S., Aster, R. C., a subducting slab: trace element and Lu–Hf–Sm–Nd isotope Dalziel, I. W. D., Huerta, A. D., Nyblade, A. A., Wilson, T. J. & systematics. Earth and Planetary Science Letters 227, Winberry, J. P. (2016). Upper mantle structure of central and 441–456. West Antarctica from array analysis of Rayleigh wave phase Jung, S., Pfa¨ nder, J. A., Bru¨ gmann, G. & Stracke, A. (2005). velocities: mantle structure of West Antarctica. Journal of Sources of primitive alkaline volcanic rocks from the Central Geophysical Research: Solid Earth 121, 1758–1775. European Volcanic Province (Rho¨ n, Germany) inferred from Henjes-Kunst, F. & Schussler, U. (2003). Metasedimentary units Hf, Os and Pb isotopes. Contributions to Mineralogy and of the Cambro-Ordovician Ross Orogen in northern Victoria Petrology 150, 546–559. Land and : implications for their provenance and Kalfoun, F., Ionov, D. & Merlet, C. (2002). HFSE residence and geotectonic setting from geochemical and Nd–Sr isotope Nb/Ta ratios in metasomatised, rutile-bearing mantle peri- data. Terra Antartica 10, 105–128. dotites. Earth and Planetary Science Letters 199, 49–65. Hergt, J. M., Peate, D. W. & Hawkesworth, C. J. (1991). The Kendrick, M. A., Honda, M., Pettke, T., Scambelluri, M., Phillips, petrogenesis of Mesozoic Gondwana low-Ti flood basalts. D. & Giuliani, A. (2013). Subduction zone fluxes of halogens Earth and Planetary Science Letters 105, 134–148. and noble gases in seafloor and forearc serpentinites. Earth Herzberg, C. (2006). Petrology and thermal structure of the and Planetary Science Letters 365, 86–96. Hawaiian plume from Mauna Kea volcano. Nature 444, Keshav, S., Gudfinnsson, G. H., Sen, G. & Fei, Y. (2004). 605–609. High-pressure melting experiments on garnet clinopyroxen- Herzberg, C. & Asimow, P. D. (2008). Petrology of some oceanic ite and the alkalic to tholeiitic transition in ocean-island island basalts: PRIMELT2.XLS software for primary magma basalts. Earth and Planetary Science Letters 223, 365–379. calculation. Geochemistry, Geophysics, Geosystems 9, doi: Kipf, A., Hauff, F., Werner, R., Gohl, K., van den Bogaard, P., 10.1029/2008GC002057. Hoernle, K., Maicher, D. & Klu¨ gel, A. (2014). Seamounts off Hirschmann, M. M., Kogiso, T., Baker, M. B. & Stolper, E. M. the West Antarctic margin: a case for non-hotspot driven (2003). Alkalic magmas generated by partial melting of gar- intraplate volcanism. Gondwana Research 25, 1660–1679. net pyroxenite. Geology 31, 481–484. Kita, N. T., Ushikubo, T., Fu, B. & Valley, J. W. (2009). High preci- Hofmann, A. W. (1988). Chemical differentiation of the Earth: sion SIMS oxygen isotope analysis and the effect of sample the relationship between mantle, continental crust, and topography. Chemical Geology 264, 43–57. oceanic crust. Earth and Planetary Science Letters 90, Krans, S. R. (2013). New mineral chemistry and oxygen iso- 297–314. topes from alkaline basalts in the Northwest Ross Sea, Hofmann, A. W. (2014). Sampling Mantle Heterogeneity Antarctica: insights on magma genesis across rifted contin- through Oceanic Basalts: Isotopes and Trace Elements, Vol. ental and oceanic lithosphere. MS thesis, Bowling Green 3, 2nd edn. Treatise on Geochemistry, Elsevier. pp. 67–101. State University, Bowling Green, OH. Huerta, A. D. & Harry, D. L. (2007). The transition from diffuse to Kyle, P. R. (1986). Mineral chemistry of Late Cenozoic McMurdo focused extension: modeled evolution of the West Antarctic Volcanic Group rocks from the Pleiades, Northern Victoria Rift system. Earth and Planetary Science Letters 255, Land. In: Stump, E. (ed.) Geological Investigations in 133–147. Northern Victoria Land. Antarctic Research Series, Husson, L., Yamato, P. & Be´ zos, A. (2015). Ultraslow, slow, or American Geophysical Union 46, 305–337. fast spreading ridges: arm wrestling between mantle con- Kyle, P. R. (1990). A. McMurdo Volcanic Group Western Ross vection and far-field tectonics. Earth and Planetary Science Embayment. In: LeMasurier, W. E. & Thomson, J. W. (eds) Letters 429, 205–215. Volcanoes of the Antarctic Plate and Southern Oceans.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 554 Journal of Petrology, 2018, Vol. 59, No. 3

Antarctic Research Series, American Geophysical Union 48, geochemistry, thermobarometry and oxybarometry of peri- 18–145. dotite and pyroxenite xenoliths from the Mount Morning Kyle, P. R. & Muncy, H. L. (1989). The geochemistry of eruptive centre. Journal of Petrology 56, 193–226. McMurdo Volcanic Group rocks. Antarctic Journal of the Mattey, D., Lowry, D. & Macpherson, C. (1994). Oxygen isotope United States 1, 14–16. composition of mantle peridotite. Earth and Planetary Kyle, P. R., Moore, J. A. & Thirlwall, M. F. (1992). Petrologic evo- Science Letters 128, 231–241. lution of anorthoclase lavas at Mount Erebus, Mayer, B., Jung, S., Romer, R. L., Pfa¨ nder, J. A., Klu¨ gel, A., Pack, Ross Island, Antarctica. Journal of Petrology 33, 849–875. A. & Gro¨ ner, E. (2014). Amphibole in alkaline basalts from Lambart, S., Laporte, D., Provost, A. & Schiano, P. (2012). Fate intraplate settings: implications for the petrogenesis of alka- of pyroxenite-derived melts in the peridotitic mantle: line lavas from the metasomatised lithospheric mantle. thermodynamic and experimental constraints. Journal of Contributions to Mineralogy and Petrology 167, 2095–2123. Petrology 53, 451–476. McCoy-West, A. J., Baker, J. A., Faure, K. & Wysoczanski, R. Lambart, S., Laporte, D. & Schiano, P. (2013). Markers of the (2010). Petrogenesis and origins of mid-Cretaceous contin- pyroxenite contribution in the major-element compositions ental intraplate volcanism in Marlborough, New Zealand: of oceanic basalts: review of the experimental constraints. implications for the long-lived HIMU magmatic Lithos 160–161, 14–36. mega-province of the SW Pacific. Journal of Petrology 51, Lanyon, R., Varne, R. & Crawford, A. J. (1993). Tasmanian 2003–2045. Tertiary basalts, the Balleny plume, and opening of the McCoy-West, A. J., Bennett, V. C. & Amelin, Y. (2016). Rapid Tasman Sea (southwest Pacific Ocean). Geology 21, Cenozoic ingrowth of isotopic signatures simulating ‘HIMU’ 555–558. in ancient lithospheric mantle: distinguishing source from LaTourrette, T., Hervig, R. L. & Holloway, J. R. (1995). Trace process. Geochimica et Cosmochimica Acta 187, 79–101. element partitioning between amphibole, phlogopite, and McIntosh, W. C. & Kyle, P. R. (1990). Adare Peninsula. In: basanite melt. Earth and Planetary Science Letters 135, LeMasurier, W. E. & Thomson, J. W. (eds) Volcanoes of the 13–30. Antarctic Plate and Southern Oceans, Antarctic Research Le Bas, M. J. L., Le Maitre, R. W., Streckeisen, A. & Zanettin, B. Series, American Geophysical Union 48, 32–35. (1986). A chemical classification of volcanic rocks based on Me´ dard, E., Schmidt, M. W., Schiano, P. & Ottolini, L. (2006). the total alkali–silica diagram. Journal of Petrology 27, Melting of amphibole-bearing wehrlites: an experimental 745–750. study on the origin of ultra-calcic nepheline-normative Li, H., Xu, Y., Ryan, J. G. & Whattam, S. A. (2017). Evolution of melts. Journal of Petrology 47, 481–504. the mantle beneath the eastern North China Craton during Melchiorre, M., Coltorti, M., Bonadiman, C., Faccini, B., O’Reilly, the Cenozoic: linking geochemical and geophysical observa- S. Y. & Pearson, N. J. (2011). The role of eclogite in the tions. Journal of Geophysical Research: Solid Earth 122, rift-related metasomatism and Cenozoic magmatism of 224–246. Northern Victoria Land, Antarctica. Lithos 124, 319–330. Lundstrom, C. C., Gill, J. & Williams, Q. (2000). A geochemically Michael, P. J., Langmuir, C. H., Dick, H. J. B., Snow, J. E., consistant hypothesis for MORB generation. Chemical Goldstein, S. L., Graham, D. W., Lehnert, K., Kurras, G., Geology 162, 105–126. Jokat, W., Mu¨ he, R. & Edmonds, H. N. (2003). Magmatic and Luyendyk, B. P. (1995). Hypothesis for Cretaceous rifting of east amagmatic seafloor generation at the ultraslow-spreading Gondwana caused by subducted slab capture. Geology 23, Gakkel ridge, Arctic Ocean. Nature 423, 956–961. 373–376. Molnar, P., Atwater, T., Mammerickx, J. & Smith, S. M. (1975). Luyendyk, B., Cisowski, S., Smith, C., Richard, S. & Kimbrough, Magnetic anomalies, bathymetry and the tectonic evolution D. (1996). Paleomagnetic study of the northern Ford Ranges, western Marie Byrd Land, West Antarctica: motion between of the South Pacific since the Late Cretaceous. Geophysical West and East Antarctica. Tectonics 15, 122–141. Journal International 40, 383–420. Ma, G. S.-K., Malpas, J., Xenophontos, C. & Chan, G. H.-N. Mortimer, N., Dunlap, W. J., Isaac, M. J., Sutherland, R. P. & (2011). Petrogenesis of latest Miocene–Quaternary contin- Faure, K. (2007). Basal Adare volcanics. Robertson Bay, ental intraplate volcanism along the northern Dead Sea fault North Victoria Land, Antarctica: late Miocene intraplate system (Al Ghab–Homs ), Western Syria: evi- basalts of subaqueous origin. In: Cooper, A. K. & Raymond, dence for lithosphere–asthenosphere interaction. Journal of C. R., et al. (eds) Proceedings of the 10th International Petrology 52, 401–430. Symposium on Antarctic Earth Sciences. US Geological Survey Open-File Report 2007-1047, Short Research Paper Mallik, A. & Dasgupta, R. (2014). Effect of variable CO2 on eclogite-derived andesite and lherzolite reaction at 045, 7 pp. 3 GPa—implications for mantle source characteristics of al- Muehlenbachs, K. (1986). Alteration of the oceanic crust and 18 kalic ocean island basalts. Geochemistry, Geophysics, the O history of seawater. In: Valley, J. W., Taylor, H. P., Jr Geosystems 15, 1533–1557. & O’Neil, J. R. (eds) Stable Isotopes in HighTemperature Martin, A. P., Cooper, A. F. & Price, R. C. (2013). Petrogenesis of Geological Processes. Mineralogical Society of America, Cenozoic, alkalic volcanic lineages at Mount Morning, West Reviews in Mineralogy 16, 425–444. Antarctica and their entrained lithospheric mantle xenoliths: Mukasa, S. B. & Dalziel, I. W. D. (2000). Marie Byrd Land, West lithospheric versus asthenospheric mantle sources. Antarctica: evolution of Gondwana’s Pacific margin con- Geochimica et Cosmochimica Acta 122, 127–152. strained by zircon U–Pb geochronology and feldspar Martin, A. P., Cooper, A. F. & Price, R. C. (2014). Increased man- common-Pb isotopic compositions. Geological Society of tle heat flow with on-going rifting of the West Antarctic rift America Bulletin 112, 611–627. system inferred from characterisation of plagioclase perido- Mu¨ ller, P., Schmidt-Thome´ , M., Kreuzer, H., Tessensohn, H. & & tite in the shallow Antarctic mantle. Lithos 190–191, Vetter, U. (1991). Cenozoic peralkaline magmatism at the 173–190. western margin of the Ross Sea, Antarctica. Memorie della Martin, A. P., Price, R. C., Cooper, A. F. & McCammon, C. A. Societa` Geologica Italiana 46, 315–336. (2015). Petrogenesis of the rifted southern Victoria Land Mu¨ ller, R. D., Cande, S. C., Stock, J. M. & Keller, W. R. (2005). lithospheric mantle, Antarctica, inferred from petrography, Crustal structure and rift flank uplift of the Adare Trough,

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 555

Antarctica. Geochemistry, Geophysics, Geosystems 6, doi: Peate, D. W., Baker, J. A., Jakobsson, S. P., Waight, T. E., Kent, 10.1029/2005GC001027. A. J. R., Grassineau, N. V. & Skovgaard, A. C. (2009). Historic Mu¨ nker, C., Pfa¨ nder, J. A., Weyer, S., Bu¨ chl, A., Kleine, T. & magmatism on the Reykjanes Peninsula, Iceland: a Mezger, K. (2003). Evolution of planetary cores and the snap-shot of melt generation at a ridge segment. Earth–Moon system from Nb/Ta systematics. Science 301, Contributions to Mineralogy and Petrology 157, 359–382. 84–87. Perinelli, C., Armienti, P. & Dallai, L. (2006). Geochemical and Nardini, I., Armienti, P., Rocchi, S. & Burgess, R. (2003). O-isotope constraints on the evolution of lithospheric man- 40Ar–39Ar chronology and petrology of the Miocene tle in the Ross Sea rift area (Antarctica). Contributions to rift-related volcanism of Daniell Peninsula (Northern Victoria Mineralogy and Petrology 151, 245–266. Land, Antarctica). Terra Antartica 10, 39–62. Perinelli, C., Armienti, P. & Dallai, L. (2011). Thermal evolution Nardini, I., Armienti, P., Rocchi, S., Dallai, L. & Harrison, D. of the lithosphere in a rift environment as inferred from the (2009). Sr–Nd–Pb–He–O isotope and geochemical con- geochemistry of mantle cumulates, Northern Victoria Land, straints on the genesis of Cenozoic magmas from the West Antarctica. Journal of Petrology 52, 665–690. Antarctic Rift. Journal of Petrology 50, 1359–1375. Perinelli, C., Gaeta, M. & Armienti, P. (2017). Cumulate xenoliths Njome, M. S. & de Wit, M. J. (2014). The Cameroon Line: ana- from Mt. Overlord, northern Victoria Land, Antarctica: a win- lysis of an intraplate magmatic province transecting both dow into high pressure storage and differentiation of oceanic and continental lithospheres: constraints, contro- mantle-derived basalts. Lithos 268–271, 225–239. versies and models. Earth-Science Reviews 139, 168–194. Pertermann, M., Hirschmann, M. M., Hametner, K., Gu¨ nther, D. Olafsson, M. & Eggler, D. H. (1983). Phase relations of amphi- & Schmidt, M. W. (2004). Experimental determination of bole, amphibole–carbonate, and phlogopite–carbonate peri- trace element partitioning between garnet and silica-rich li- dotite: petrologic constraints on the asthenosphere. Earth quid during anhydrous partial melting of MORB-like eclog- and Planetary Science Letters 64, 305–315. ite. Geochemistry, Geophysics, Geosystems 5, doi: Panter, K. S. & Castillo, P. (2007). Petrogenesis and source of 10.1029/2003GC000638. lavas from seamounts in the Adare Basin, Western Ross Peters, B. J. & Day, J. M. D. (2014). Assessment of relative Ti, Sea: implications for the origin of Cenozoic magmatism in Ta, and Nb (TITAN) enrichments in ocean island basalts. Antarctica. In: Cooper, A. K. & Raymond, C. R., et al. (eds) Geochemistry, Geophysics, Geosystems 15, 4424–4444. Proceedings of the 10th International Symposium on Pfa¨ nder, J. A., Mu¨ nker, C., Stracke, A. & Mezger, K. (2007). Nb/Ta and Zr/Hf in ocean island basalts—implications for Antarctic Earth Sciences. US Geological Survey Open-File crust–mantle differentiation and the fate of niobium. Earth Report 2007-1047, Extended Abstract 069, 4 pp. Panter, K. S. & Castillo, P. (2008). Petrology and source of lavas and Planetary Science Letters 254, 158–172. Pfa¨ nder, J. A., Jung, S., Mu¨ nker, C., Stracke, A. & Mezger, K. from seamounts in the Adare Basin, western Ross Sea: (2012). A possible high Nb/Ta reservoir in the continental implications for the origin of Cenozoic magmatism in lithospheric mantle and consequences on the global Nb Antarctica. EOS Transactions, American Geophysical Union budget—evidence from continental basalts from Central 89, Fall Meeting Abstract V11F-05. Germany. Geochimica et Cosmochimica Acta 77, 232–251. Panter, K. S., Kyle, P. R. & Smellie, J. L. (1997). Petrogenesis of a Pilet, S. (2015). Generation of low-silica alkaline lavas: petrolog- phonolite–trachyte succession at Mount Sidley, Marie Byrd ical constraints, models, and thermal implications. In: Land, Antarctica. Journal of Petrology 38, 1225–1253. Foulger, G.R., Lustrino, M. & King, S. D. (eds) The Panter, K. S., Hart, S. R., Kyle, P., Blusztanjn, J. & Wilch, T. Interdisciplinary Earth: In Honor of Don L. Anderson. (2000). Geochemistry of late Cenozoic basalts from the Geological Society of America, Special Papers 514, 514–517. Crary Mountains: characterization of mantle sources in Pilet, S., Hernandez, J., Bussy, F. & Sylvester, P. J. (2004). Marie Byrd Land, Antarctica. Chemical Geology 165, Short-term metasomatic control of Nb/Th ratios in the man- 215–241. tle sources of intraplate basalts. Geology 32, 113. Panter, K. S., Blusztajn, J., Wingrove, D., Hart, S. & Mattey, D. Pilet, S., Hernandez, J., Sylvester, P. & Poujol, M. (2005). The (2003). Sr, Nd, Pb, Os, O isotope, major and trace element metasomatic alternative for ocean island basalt chemical data from basalts, South Victoria Land, Antarctica: evidence heterogeneity. Earth and Planetary Science Letters 236, for open-system processes in the evolution of mafic alkaline 148–166. magmas. European Geophysical Society, Geophysical Pilet, S., Baker, M. B. & Stolper, E. M. (2008). Metasomatized Research Abstracts 5, 07583. lithosphere and the origin of alkaline lavas. Science 320, Panter, K. S., Blusztajn, J., Hart, S., Kyle, P., Esser, R. & 916–919. McIntosh, W. (2006). The origin of HIMU in the SW Pacific: Pilet, S., Ulmer, P. & Villiger, S. (2010). Liquid line of descent of evidence from intraplate volcanism in Southern New a basanitic liquid at 15 GPa: constraints on the formation of Zealand and Subantarctic Islands. Journal of Petrology 47, metasomatic veins. Contributions to Mineralogy and 1673–1704. Petrology 159, 621–643. Panter, K. S., Castillo, P. & McIntosh, W. C. (2011). Geochemical Pilet, S., Baker, M. B., Muntener, O. & Stolper, E. M. (2011). variations in alkaline magmas of HIMU affinity from contin- Monte Carlo simulations of metasomatic enrichment in the ent to ocean, northwestern Ross Sea, lithosphere and implications for the source of alkaline Antarctica—constraints on mantle source, flow and melting basalts. Journal of Petrology 52, 1415–1442. for volcanism. International Symposium on Antarctic Earth Putirka, K. D., Perfit, M., Ryerson, F. J. & Jackson, M. G. (2007). Science (ISAES) XI, 10–15 July, Edinburgh, Scotland, Ambient and excess mantle temperatures, olivine thermom- Abstract PS5.9. etry, and active vs. passive upwelling. Chemical Geology Panter, K., Castillo, P., Krans, S. & McIntosh, W. (2016). Melt ori- 241, 177–206. gin across a rifted continental edge: 40Ar/39Ar geochron- Putlitz, B., Matthews, A. & Valley, J. W. (2000). Oxygen and ology, geochemistry and Sr–Nd–O isotope systematics of hydrogen isotope study of high-pressure metagabbros and alkaline basalts, northwestern Ross Sea. XXXIV SCAR Open metabasalts (Cyclades, Greece): implications for the sub- Science Conference, 22–26 August, Kuala Lumpur, duction of oceanic crust. Contributions to Mineralogy and Malaysia. Petrology 138, 114–126.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 556 Journal of Petrology, 2018, Vol. 59, No. 3

Reid, I. & Jackson, H. R. (1981). Oceanic spreading rate and Skovgaard, A. C., Storey, M., Baker, J., Blusztajn, J. & Hart, S. R. crustal thickness. Marine Geophysical Researches 5, (2001). Osmium–oxygen isotopic evidence for a recycled 165–172. and strongly depleted component in the Iceland mantle Renne, P. R., Swisher, C. C., Deino, A. L., Karner, D. B., Owens, plume. Earth and Planetary Science Letters 194, 259–275. T. L. & DePaolo, D. J. (1998). Intercalibration of standards, Smellie, J. L., Rocchi, S., Gemelli, M., Di Vincenzo, G. & absolute ages and uncertainties in 40Ar/39Ar dating. Armienti, P. (2011). A thin predominantly cold-based Late Chemical Geology 145, 117–152. Miocene East Antarctic inferred from glaciovol- Ritzwoller, M. H., Shapiro, N. M., Levshin, A. L. & Leahy, G. M. canic sequences in northern Victoria Land, Antarctica. (2001). Crustal and upper mantle structure beneath Palaeogeography, Palaeoclimatology, Palaeoecology 307, Antarctica and surrounding oceans. Journal of Geophysical 129–149. Research: Solid Earth 106, 30645–30670. Stein, M., Navon, O. & Kessel, R. (1997). Chromatographic Rocchi, S., Armienti, P., D’Orazio, M., Tonarini, S., Wijbrans, J. metasomatism of the Arabian–Nubian lithosphere. Earth R. & Di Vincenzo, G. (2002). Cenozoic magmatism in the and Planetary Science Letters 152, 75–91. western Ross Embayment: role of mantle plume versus Storey, B. C. & Kyle, P. R. (1997). An active mantle mechanism plate dynamics in the development of the West Antarctic for Gondwana breakup. South African Journal of Geology Rift System. Journal of Geophysical Research: Solid Earth 100, 283–290. 107, ECV 5-1–ECV 5-22. Storey, B. C., Leat, P. T., Weaver, S. D., Pankhurst, R. J., Rocchi, S., Armienti, P. & Di Vincenzo, G. (2005). No plume, no Bradshaw, J. D. & Kelley, S. (1999). Mantle plumes and rift magmatism in the West Antarctic Rift. Plates, plumes Antarctica–New Zealand rifting: evidence from and paradigms. In: Geological Society of America, Special mid-Cretaceous mafic dykes. Journal of the Geological Papers 388, 435–447. Society, London 156, 659–671. Rocholl, A., Stein, M., Molzahn, M., Hart, S. R. & Wo¨ rner, G. Stracke, A. (2012). Earth’s heterogeneous mantle: a product of (1995). Geochemical evolution of rift magmas by progres- convection-driven interaction between crust and mantle. sive tapping of a stratified mantle source beneath the Ross Chemical Geology 330–331, 274–299. Sea Rift, Northern Victoria Land, Antarctica. Earth and Stracke, A., Hofmann, A. W. & Hart, S. R. (2005). FOZO, HIMU, Planetary Science Letters 131, 207–224. and the rest of the mantle zoo. Geochemistry, Geophysics, Rooney, T. O., Nelson, W. R., Dosso, L., Furman, T. & Hanan, B. Geosystems 6, Q05007, doi:10.1029/2004GC000824. (2014). The role of continental lithosphere metasomes in the Stump, E. (1995) The Ross Orogen of the Transantarctic production of HIMU-like magmatism on the northeast Mountains. Cambridge University Press, 284 pp. African and Arabian plates. Geology 42, 419–422. Sumino, H., Burgess, R., Mizukami, T., Wallis, S. R., Holland, G. Rooney, T. O., Nelson, W. R., Ayalew, D., Hanan, B., Yirgu, G. & & Ballentine, C. J. (2010). Seawater-derived noble gases and Kappelman, J. (2017). Melting the lithosphere: metasomes halogens preserved in exhumed mantle wedge peridotite. as a source for mantle-derived magmas. Earth and Earth and Planetary Science Letters 294, 163–172. Planetary Science Letters 461, 105–118. Sun, S -S. & McDonough, W. F. (1989). Chemical and isotopic Rudnick, R. L., Barth, M., Horn, I. & McDonough, W. F. (2000). systematics of oceanic basalts: implications for mantle com- Rutile-bearing refactory eclogites: missing link between position and processes. In: Saunders, A. D. & Norry, M. J. continents and depleted mantle. Science 287, 278–281. (eds) Magmatism in the Ocean Basins. Geological Society, Salimbeni, S., Pondrelli, S., Danesi, S. & Morelli, A. (2010). London, Special Publications 42, 313–345. Seismic anisotropy of the Victoria Land region, Antarctica: Sutherland, R., Spasojevic, S. & Gurnis, M. (2010). Mantle up- seismic anisotropy beneath West Antarctica. Geophysical welling after Gondwana subduction death explains anomal- Journal International 182, 421–432. ous topography and subsidence histories of eastern New Scott, J. M., Waight, T. E., van der Meer, Q. H. A., Palin, J. M., Zealand and West Antarctica. Geology 38, 155–158. Cooper, A. F. & Mu¨ nker, C. (2014). Metasomatized ancient Thirlwall, M. F. (1997). Pb isotopic and elemental evidence for lithospheric mantle beneath the young Zealandia microcon- OIB derivation from young HIMU mantle. Chemical Geology tinent and its role in HIMU-like intraplate magmatism. 139, 51–74. Geochemistry, Geophysics, Geosystems 15, 3477–3501. Tiepolo, M., Oberti, R., Zanetti, A., Vannucci, R., Foley, S. F. Selvans, M. M., Clayton, R. W., Stock, J. M. & Granot, R. (2012). (2007). Trace-element partitioning between amphibole and Using overlapping sonobuoy data from the Ross Sea to con- silicate melt. In: Hawthorne, F. C., Oberti, R., Della Ventura, struct a 2D deep crustal velocity model. Marine Geophysical G. & Mottana, A. (eds) Amphiboles: Crystal Chemistry, Research 33, 17–32. Occurrence and Health Issues. Mineralogical Society of Selvans, M. M., Stock, J. M., Clayton, R. W., Cande, S. & Granot, America and Geochemical Society, Reviews in Mineralogy R. (2014). Deep crustal structure of the Adare and Northern and Geochemistry 67, 417–452. Basins, Ross Sea, Antarctica, from sonobuoy data. Earth Timm, C., Hoernle, K., Werner, R., Hauff, F., van den Bogaard, and Planetary Science Letters 405, 220–230. P., White, J., Mortimer, N. & Garbe-Scho¨ nberg, D. (2010). Shaw, C. S. (1999). Dissolution of orthopyroxene in basanitic Temporal and geochemical evolution of the Cenozoic intra- magma between 04 and 2 GPa: further implications for the plate volcanism of Zealandia. Earth-Science Reviews 98, origin of Si-rich alkaline glass inclusions in mantle xenoliths. 38–64. Contributions to Mineralogy and Petrology 135, 114–132. Todt, W., Cliff, R. A., Hanser, A. & Hoffman, A. W. (1996). Shaw, C. S., Thibault, Y., Edgar, A. D. & Lloyd, F. E. (1998). Evaluation of a 202Pb–205Pb double spike for high-precision Mechanisms of orthopyroxene dissolution in lead isotope analysis. In: Basu, A. & Hart, S. (eds) Earth silica-undersaturated melts at 1 atmosphere and implica- Processes: Reading the Isotopic Code. Geophysical tions for the origin of silica-rich glass in mantle xenoliths. Monograph, American Geophysical Union 95, 429–437. Contributions to Mineralogy and Petrology 132, 354–370. Tonarini, S., Rocchi, S., Armienti, P. & Innocenti, F. (1997). Sisson, T. W. & Grove, T. L. (1993). Experimental investigations Constraints on timing of Ross Sea rifting inferred from

of the role of H2O in calc-alkaline differentiation and subduc- Cainozoic intrusions from northern Victoria Land, tion zone magmatism. Contributions to Mineralogy and Antarctica. In: Ricci, C. A. (ed.) The Antarctic Region: Petrology 113, 143–166. Geological Evolution and Processes, Proceedings of the 7th

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Journal of Petrology, 2018, Vol. 59, No. 3 557

International Symposium on Antarctic Earth Sciences. Terra from in situ measurement of oxygen isotopes, laser Raman, Antarctica, pp. 511–521. U–Pb ages, and trace elements. Chemical Geology 389, Tulloch, A. J., Ramezani, J., Mortimer, N., Mortensen, J., van 122–136. den Bogaard, P. & Maas, R. (2009). Cretaceous felsic volcan- Wang, Z. & Eiler, J. (2008). Insights into the origin of low-d18O ism in New Zealand and Lord Howe Rise (Zealandia) as a basaltic magmas in Hawaii revealed from in situ measure- precursor to final Gondwana break-up. In: Ring, U. & ments of oxygen isotope compositions of olivines. Earth Wernicke, B. (eds) Extending a Continent: Architecture, and Planetary Science Letters 269, 377–387. Rheology and Heat Budget. Geological Society, London, Weaver, B. L. (1991). The origin of ocean island basalt Special Publications 321, 89–118. end-member compositions: trace element and isotopic con- Valley, J. W. & Kita, N. T. (2009). In situ oxygen isotope geo- straints. Earth and Planetary Science Letters 104, 381–397. chemistry by ion microprobe. In: Fayek, M. (ed) Secondary Weaver, S. D., Storey, B. C., Pankhurst, R. J., Mukasa, S. B., Ion Mass Spectrometry in the Earth Sciences: Gleaning the DiVenere, V. J. & Bradshaw, J. D. (1994). Antarctica–New Big Picture from a Small Spot. Mineralogical Association of Zealand rifting and Marie Byrd Land lithospheric magma- Canada, Short Course Series 41, 19–63. tism linked to ridge subduction and mantle plume activity. Valley,J.W.,Kitchen,N.E.,Kohn,M.J.,Niendorf,C.R.&Spicuzza, Geology 22, 811–814. M. J. (1995). UWG-2, a garnet standard for oxygen isotope ratio: Weiss, Y., Class, C., Goldstein, S. L. & Hanyu, T. (2016). Key new strategies for high precision and accuracy with laser heating. pieces of the HIMU puzzle from olivines and diamond inclu- Geochimica et Cosmochimica Acta 59, 5223–5231. sions. Nature 537, 666–670. van der Meer, Q. H. A., Waight, T. E., Scott, J. M. & Mu¨ nker, C. White, W. M. (2015). Probing the earth’s deep interior through (2017). Variable sources for Cretaceous to recent HIMU and geochemistry. Geochemical Perspectives 4, 251. HIMU-like intraplate magmatism in New Zealand. Earth and Willbold, M. & Stracke, A. (2010). Formation of enriched mantle Planetary Science Letters 469, 27–41. components by recycling of upper and lower continental Vaughan, A. P. & Storey, B. C. (2007). A new supercontinent crust. Chemical Geology 276, 188–197. self-destruct mechanism: evidence from the Late Yaxley, G. M. & Green, D. H. (1998). Reactions between eclogite Triassic–Early Jurassic. Journal of the Geological Society, London 164, 383–392. and peridotite: mantle refertilisation by subduction of Vaughan, A. P. M. & Pankhurst, R. J. (2008). Tectonic overview oceanic crust. Schweizerische Mineralogische und of the West Gondwana margin. Gondwana Research 13, Petrographische Mitteilungen 78, 243–255. 150–162. Zheng, Y.-F., Fu, B., Li, Y., Xiao, Y. & Li, S. (1998). Oxygen and Veevers, J. J. (2012). Reconstructions before rifting and drifting hydrogen isotope geochemistry of ultrahigh-pressure eclo- reveal the geological connections between Antarctica and gites from the Dabie Mountains and the Sulu terrane. Earth its conjugates in Gondwanaland. Earth-Science Reviews and Planetary Science Letters 155, 113–129. 111, 249–318. Zheng, Y.-F., Fu, B., Gong, B. & Li, L. (2003). Stable isotope geo- Walter, M. J. (1998). Melting of garnet peridotite and the origin chemistry of ultrahigh pressure metamorphic rocks from of komatiite and depleted lithosphere. Journal of Petrology the Dabie–Sulu orogen in China: implications for geody- 39, 29–60. namics and fluid regime. Earth-Science Reviews 62, Wang, X.-L., Coble, M. A., Valley, J. W., Shu, X.-J., Kitajima, K., 105–161. Spicuzza, M. J. & Sun, T. (2014). Influence of radiation dam- Zindler, A. & Hart, S. (1986). Chemical geodynamics. Annual age on late Jurassic zircon from southern China: evidence Review of Earth and Planetary Sciences 14, 493–571.

Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018 Downloaded from https://academic.oup.com/petrology/article-abstract/59/3/517/4966906 by University of Wisconsin-Madison Libraries user on 04 September 2018