1,138
Views
7
CrossRef citations to date
0
Altmetric
South Island

Intraplate volcanism on the Zealandia Eocene-Early Oligocene continental shelf: the Waiareka-Deborah Volcanic Field, North Otago

ORCID Icon, ORCID Icon & ORCID Icon
Pages 450-468 | Received 11 May 2020, Accepted 18 Jun 2020, Published online: 09 Jul 2020

ABSTRACT

Volcaniclastic deposits, pillow lavas, dikes and sills of the intraplate Waiareka-Deborah Volcanic Field in North Otago were emplaced into and onto the Zealandia Eocene-Early Oligocene continental shelf. The on-land extent is ∼890 km2 but offshore volcanic rocks occurring over an additional ∼3500 km2 may be related. Examination of the on-land volcaniclastic deposits indicates volcanism was dominated by short-lived clustered surtseyan-style eruptions. Pyroclast glass and sill and lava bulk rock chemistries show that the magmas were mainly sub-alkaline basalt to basaltic andesite. Minor alkaline centres are best represented by basanitic-melanephelinitic volcaniclastic deposits at Kakanui, which also contain an array of megacrysts plus mantle and crustal xenoliths. The sub-alkaline and alkaline source reservoirs had similar Sr and Nd isotope ratios (87Sr/86Sr34 Ma = 0.70346 ± 40, 143Nd/144Nd34 Ma = 0.51282 ± 4) but the sub-alkaline rocks tend to be slightly less radiogenic in 206Pb/204Pb or 208Pb/204Pb versus 207Pb/204Pb space. The dominant sub-alkaline nature and the isotopic compositions distinguish this volcanic field from the nearby Dunedin Volcanic Group and Alpine Dike Swarm. As the Waiareka-Deborah isotopic compositions are poorly represented in the Otago mantle lithosphere, the magmas may have been derived from the asthenosphere. There were multiple modes of Cenozoic intraplate volcanism in Otago.

Introduction

The Waiareka-Deborah Volcanic Field near Oamaru in the South Island () displays spectacular evidence for submarine volcanic processes on the mid-Cenozoic Zealandia continental shelf. This intraplate volcanic field has long been known to contain locally thick piles of volcaniclastic sediments, pillow lavas, diatremes, sills and dikes (e.g. Park Citation1918; Uttley Citation1918; Benson Citation1943, Citation1944; Gage Citation1957; Coombs et al. Citation1986), including the renowned xenocryst- and xenolith-rich Kakanui Mineral Breccia (e.g. Thomson Citation1907; Mason Citation1968; Dickey Citation1968a; Reay and Sipiera Citation1987; Fulmer et al. Citation2010). The geochemistry of the province, however, has had little attention beyond the few analyses of Benson (Citation1943), Coombs et al. (Citation1986) and Hoernle et al. (Citation2006), and regional syntheses compiled by Benson (Citation1943, Citation1944) and Coombs et al. (Citation1986) are out of date; for example, some of the alkaline intrusions formerly grouped in this province are now known to be part of the younger Dunedin Volcanic Group (Coombs et al. Citation2008).

Figure 1. Geological map showing the distribution of the Waiareka-Deborah Volcanic Field. The map is compiled from data presented in Benson (Citation1943), Coombs et al. (Citation1986), Forsyth (Citation2001) and personal observations. Ages shown are Ar-Ar dates from Hoernle et al. (Citation2006). Inset shows location of geological map (bold box), as well as the extent of the Alpine Dike Swarm and Dunedin Volcanic Group, which are other intraplate volcanic provinces in Otago.

Figure 1. Geological map showing the distribution of the Waiareka-Deborah Volcanic Field. The map is compiled from data presented in Benson (Citation1943), Coombs et al. (Citation1986), Forsyth (Citation2001) and personal observations. Ages shown are Ar-Ar dates from Hoernle et al. (Citation2006). Inset shows location of geological map (bold box), as well as the extent of the Alpine Dike Swarm and Dunedin Volcanic Group, which are other intraplate volcanic provinces in Otago.

We provide a synthesis of the physical volcanology and the geochemistry of the Waiareka-Deborah Volcanic Field. Phreatomagmatic deposits are spectacularly exposed in cliffs bordering the Pacific Ocean and have been well-studied in the past two decades, primarily by University of Otago students. The sparse existing crystalline rock geochemical data are supplemented with new major and trace element and Sr-Nd-Pb isotopic analyses to give a regional perspective of the magmatism and its mantle sources. These data enable this intraplate province to be compared with the spatially overlapping but younger (25-9 Ma) Dunedin Volcanic Group as well as the younger and spatially distinct (∼25-20 Ma) Alpine Dike Swarm, both of which are also intraplate volcanic fields in the Otago region.

Geological setting

After breakaway from Gondwana at 84 Ma, New Zealand’s sedimentary record shows this continental mass became largely submerged. The peak of marine transgression occurred in the Early Oligocene (e.g. Forsyth Citation2001), with later marine regression and uplift of the North and South Islands initiated by mid-Oligocene propagation of the Australia-Pacific plate boundary through the centre of the continent. Zealandia has throughout its history been punctuated by volcanism. While the most prominent volcanism, forming the volcanoes of the Taupo Volcanic Zone, is arc-related, Zealandia has a long history of intraplate volcanism (e.g. Mortimer and Scott Citation2020) and one product of this history is the topic of this paper: mid-Cenozoic volcanic rocks near Oamaru in North Otago in the South Island. The volcanic rocks around Oamaru were historically considered to comprise two separate formations with the older Waiareka Volcanic Formation separated from the younger Deborah Volcanic Formation by the Totara Limestone (e.g. Park Citation1918; Uttley Citation1918, Citation1920; Benson Citation1943; Coombs and Dickey Citation1965). This limestone, subsequently renamed as Ototara, was deposited during and between times of volcanic activity but the inference that there is an older and a younger member separated by a hiatus involving regional carbonate formation is a major oversimplification. Instead, there were many local periods of non-volcanic deposition that produced limestone and marine mud- and silt-stone units (Coombs et al. Citation1986; Edwards Citation1991; Andrews Citation2003; Maicher Citation2003). The two volcanic formations are now grouped as part of the Alma Group (Forsyth Citation2001) and following Coombs et al. (Citation1986) we refer to the volcanic rocks collectively as the Waiareka-Deborah Volcanic Field.

The minimum extent of the Waiareka-Deborah Volcanic Field is ∼890 km2 (). However, erosion has likely removed some inland occurrences and there is evidence in seismic lines and specimens from test wells that indicates intrusions and tephra offshore (Wilding and Sweetman Citation1971; Field et al. Citation1989; Bischoff et al. Citation2020). If the offshore volcanism is included within the Waiareka-Deborah Volcanic Field (∼3,500 km; Bischoff et al. Citation2020), then the entire volcanic field occurs over an area probably closer to ∼4400 km2. Biostratigraphic ages from units bounding tuffs from Endeavour-1 Well indicate, however, that this areal extent also includes Paleocene volcanic rocks (Wilding and Sweetman Citation1971), and so the extent of offshore volcanism correlating with the on-land Waiareka-Deborah Volcanic Field remains unknown.

The crustal ‘basement’ underlying the Waiareka-Deborah Volcanic Field is the dominantly quartzofeldspathic Otago Schist, which was metamorphosed to greenschist facies conditions in the Late Jurassic – Early Cretaceous and exhumed by ∼ 100 Ma (Mortimer Citation2000; Forsyth Citation2001). However, felsic granulite facies xenoliths in the Kakanui Mineral Breccia indicate that the middle and lower crust, at least in part, comprises rocks metamorphosed at ∼92 Ma under extreme high to ultra-high temperature conditions (∼900°C; Jacob et al. Citation2017). The Otago Schist is unconformably overlain in the study area by terrestrial Cretaceous quartzose conglomerate, sandstone and mudstone belonging to the Matakea Group, and/or Late Cretaceous to Early Oligocene Onekakara Group sediments that evolve from non-marine to marine (Forsyth Citation2001). The volcanic rocks of the Alma Group – the topic of our paper – were emplaced during the Eocene-Late Oligocene, very late in the accumulation of region-wide Onekakara Group sandstone and mudstone (e.g. Thompson et al. Citation2014). A regional unconformity, the so-called ‘Marshall Paraconformity’, separates the Alma Group and Onekakara Group from Late Oligocene to Miocene Kekendon Group sediments. It was during Kekendon Group sedimentation that a significant portion of Zealandia emerged above sea-level and this was accompanied by Dunedin Volcanic Group magmatism, which is an extensive alkaline volcanic province that in part overlaps spatially with the Waiareka-Deborah volcanism () (Coombs et al. Citation1986, Citation2008; Scott et al. Citation2020). Emplacement of the intraplate alkaline-carbonatitic Alpine Dike Swarm in West Otago also occurred in the Oligocene to Early Miocene (e.g. Cooper Citation2020).

Volcaniclastic rocks

Volcaniclastic rocks are the most areally extensive component of the on-land Waiareka-Deborah Volcanic Field (), although they are very poorly exposed except in coastal sections. Detailed sedimentological descriptions of these deposits have been made, from north to south, from Cape Wanbrow (Uttley Citation1918; Coombs et al. Citation1986; Moorhouse et al. Citation2015; Moorhouse and White Citation2016), the areas around Maheno and Clarks Mill (Hicks Citation2014), Kakanui (Dickey Citation1968a; Corcoran and Moore Citation2008), Aorere Point-Bridge Point (Cas et al. Citation1989; Hicks Citation2014), Lookout Bluff (Maicher Citation2003) and Moeraki (Andrews Citation2003) (). These sequences are more than 150 m thick in some places (Coombs et al. Citation1986; Andrews Citation2003; Corcoran and Moore Citation2008; Moorhouse et al. Citation2015). A road cutting now largely concreted over once exposed a few-metre thick section of mantle xenolith-bearing tuff breccia and lapilli tuff overlying pillow basalt at Alma (Reay et al. Citation2002).

All of the studied sections comprise sequences of tuff breccia, lapilli tuff and tuff (A) that are in places interbedded with pillow lava and limestone, mudstone or marl (B). The contact relationships among different volcaniclastic deposits, often featuring sharp but local erosion surfaces (C) and/or cross-bedding, indicate that they formed primarily as deposits of submarine density currents from proximal volcanic edifices, emplaced both as primary deposits during eruptions, and through subsequent post-eruptive redistribution (Cas et al. Citation1989; Andrews Citation2003; Maicher Citation2003; Corcoran and Moore Citation2008; Moorhouse et al. Citation2015). Volcaniclasts are mostly now-altered vesicular basaltic glass, or blocks fragmented from dikes associated with the eruptions (D). Anomalously, Kakanui deposits also contain a variety of crust and mantle fragments (described below). Palaeontology indicates that water depths were generally shallow (∼ <80 m) during the volcanic field's duration (Lee et al. Citation1997; Hicks Citation2014) and it is probable that volcanic edifices breached the sea surface, with those of entirely pyroclastic construction being rapidly planed off to normal wave base (e.g. Thorarinsson Citation1967). The volcanoes eroded rapidly, with evidence for calcareous algae, bryozoans and serpulids encrusting basalt boulders and cobbles (Lee et al. Citation1997) indicating the volcanoes formed sites favourable for rhodolith, bryozoan, brachiopod, echinoderm, bivalve and gastropod colonies (Utlley 1918; Cas et al. Citation1989; Lee et al. Citation1997; Corcoran and Moore Citation2008; Hicks Citation2014).

Figure 2. Examples of Waiareka-Deborah Volcanic Field phreatomagmatic rocks. A, View of fine-grained tephra beds that are extensive around Cape Wanbrow. The cliffs are ∼30 m high. B, A spectacular sequence of tephras and thin calcareous tuffs, tuffaceous limestone and limestone horizons is exposed at Boatmans Harbour on the northern side of Cape Wanbrow. The occurrence of limestone beds, and elsewhere limestone or mudstone, indicates that the volcanic field had numerous hiatuses between intermittent eruptions. Photo courtesy of Marco Brenna. C, Truncated bedforms are very common at Kakanui and the erosion surfaces are of only local significance. D, Breccia components typical comprise basalt, in places appearing to be made from fragmented dikes. Outcrop is at Lookout Bluff. Pocket knife is ∼8 cm long.

Figure 2. Examples of Waiareka-Deborah Volcanic Field phreatomagmatic rocks. A, View of fine-grained tephra beds that are extensive around Cape Wanbrow. The cliffs are ∼30 m high. B, A spectacular sequence of tephras and thin calcareous tuffs, tuffaceous limestone and limestone horizons is exposed at Boatmans Harbour on the northern side of Cape Wanbrow. The occurrence of limestone beds, and elsewhere limestone or mudstone, indicates that the volcanic field had numerous hiatuses between intermittent eruptions. Photo courtesy of Marco Brenna. C, Truncated bedforms are very common at Kakanui and the erosion surfaces are of only local significance. D, Breccia components typical comprise basalt, in places appearing to be made from fragmented dikes. Outcrop is at Lookout Bluff. Pocket knife is ∼8 cm long.

It is widely accepted that the Waiareka-Deborah volcaniclastic deposits are products of phreatomagmatic interaction of magma with seawater, which drove surtseyan-style eruptions on the continental shelf () (Coombs et al. Citation1986; Cas et al. Citation1989; Andrews Citation2003; Maicher Citation2003; Corcoran and Moore Citation2008; Moorhouse et al. Citation2015). The volcaniclastic sedimentary record shows that in multiple sites deposits from multiple volcanoes, formed at different times, overlap in the stratigraphy (Maicher Citation2003; Moorhouse et al. Citation2015). The presence of limestone, mudstone and glauconitic siltstone between volcaniclastic deposits shows, however, that there were significant hiatuses between the eruptions that formed different small volcanoes in the field, even where they are spatially clustered (Andrews Citation2003; Maicher Citation2003; Corcoran and Moore Citation2008). Formation of tephra deposits in the basin away from immediate volcano flanks was probably by some combination of (1) distal eruption-fed density currents (e.g. Verolino et al. Citation2018); (2) vertical density currents driven by accumulation of ash near the water surface (e.g. Bradley Citation1965; Jacobs et al. Citation2015), and; (3) settling of ash through the water column (e.g. Kutterolf et al. Citation2018).

Figure 3. Cartoon showing the inferred shallow-level development of the Waiareka-Deborah Volcanic Field.

Figure 3. Cartoon showing the inferred shallow-level development of the Waiareka-Deborah Volcanic Field.

Original pyroclasts preserved in the volcaniclastic rocks are commonly extensively palagonitised and only plagioclase crystals survive with any regularity (e.g. Moorhouse et al. Citation2015). Nonetheless, sideromelane with small crystals of olivine, plagioclase and clinopyroxene has been found dispersed through the rocks (Coombs et al. Citation1986; Andrews Citation2003), with the glass chemistry indicating mostly transitional alkaline to sub-alkaline compositions (Coombs et al. Citation1986). Evidence for alkaline eruptions is recorded from a basanitic ash horizon at Cape Wanbrow and sideromelane pyroclasts at Kakanui (Dickey Citation1968a; Coombs et al. Citation1986). The volcaniclastic record is therefore interpreted to show that were there numerous small surtseyan-style eruptions in the field () with different magma batches for each volcano.

Lavas, sills and dikes

Pillow lavas, sills and dikes make up a smaller part of the outcrop area than do volcaniclastic rocks in the on-land portion of the volcanic field (), and as non-fragmental deposits they provide information about non-explosive eruptions, or parts of eruptions. The most comprehensive descriptions of these coherent rocks, although reconnaissance in nature, remain those made by Uttley (Citation1918), Benson (Citation1943) and Nakamura and Coombs (Citation1973). Sills and intrusions have also been mapped in offshore seismic sections (Bischoff et al. Citation2020), but none have been sampled or dated.

The best exposures of pillow lava occur at Boatmans Harbour, where a 30 m thick horizon of basaltic pillows set within Ototara Limestone matrix is spectacularly exposed (Park Citation1918; Uttley Citation1918; Kawachi and Pringle Citation1988) (A, B). The pillow rinds are unaltered sideromelane enclosing fresh magmatic labradorite and olivine, with clinopyroxene present in the pillow interiors (Coombs and Dickey Citation1965; Coombs et al. Citation1986; Kawachi and Pringle Citation1988). Single pillows commonly have more than one rind. This feature may indicate fragmentation of initial rinds followed by water invasion and formation of new ‘interior’ rinds, either as pillows partly imploded during emplacement (Kawachi and Pringle Citation1988) or as rinds were broken and partly detached during pillow expansion (Walker Citation1992). Slightly higher in the sequence, and just above an intervening thick limestone horizon, a thick tuff breccia represents a lava-fed delta, containing broken pillows and other glassy fragments (B). Lava in Awamoa Creek reportedly comprises pillows up to ‘18 feet in diameter’ (Benson Citation1943, p. 120). Very altered pillow lava occurs at the Alma road cut (Reay et al. Citation2002) and Oamaru Creek (Uttley Citation1918) but these have not been examined in much detail.

Figure 4. Flows, sills and dikes of the volcanic field. A, Pillow lava overlain by tuffaceous material at Boatmans Harbour (Cape Wanbrow; sequence also illustrated in B). The beach to the top of the cliffs on the right-hand side of 4A are about 15 m high. B, The pillows have glassy margins and have interstitial bryozoan limestone. C, The Tokorahi Sill, here exposed on Dip Hill Road, has sub-vertical columnar jointing. The basal contact with Eocene sandstone is just to the left of this photo. Although not obviously differentiated at outcrop scale, the whole rock geochemistry shows come variation in composition across the body. D, View of the Mt Charles Sill, which intrudes concretion-bearing mudstone. The dolerite is extremely friable on this cliff, but fresher (although less well exposed) material occurs at the southern end of the sill. This sill is chemically differentiated vertically, from olivine dolerite at the base upwards to quartz dolerite (Benson Citation1943). View is to the northeast from State Highway 1. The exposed sill is ∼20 m thick. E, The Tawhiroko Sill, which is also geochemically differentiated (Benson Citation1943), on the Moeraki Peninsula has horizons rich in schist clasts that are derived from the underlying Otago Schist. Photos courtesy of M Brenna. F, Features of mafic dikes on Moeraki Peninsula. The dikes appear to be made of multiple injections and are packed, in places, with calcite-filled vesicles. Inset shows the ∼6 cm wide white porcellanite margin to one dike. Photo courtesy of AF Cooper.

Figure 4. Flows, sills and dikes of the volcanic field. A, Pillow lava overlain by tuffaceous material at Boatmans Harbour (Cape Wanbrow; sequence also illustrated in Figure 2B). The beach to the top of the cliffs on the right-hand side of 4A are about 15 m high. B, The pillows have glassy margins and have interstitial bryozoan limestone. C, The Tokorahi Sill, here exposed on Dip Hill Road, has sub-vertical columnar jointing. The basal contact with Eocene sandstone is just to the left of this photo. Although not obviously differentiated at outcrop scale, the whole rock geochemistry shows come variation in composition across the body. D, View of the Mt Charles Sill, which intrudes concretion-bearing mudstone. The dolerite is extremely friable on this cliff, but fresher (although less well exposed) material occurs at the southern end of the sill. This sill is chemically differentiated vertically, from olivine dolerite at the base upwards to quartz dolerite (Benson Citation1943). View is to the northeast from State Highway 1. The exposed sill is ∼20 m thick. E, The Tawhiroko Sill, which is also geochemically differentiated (Benson Citation1943), on the Moeraki Peninsula has horizons rich in schist clasts that are derived from the underlying Otago Schist. Photos courtesy of M Brenna. F, Features of mafic dikes on Moeraki Peninsula. The dikes appear to be made of multiple injections and are packed, in places, with calcite-filled vesicles. Inset shows the ∼6 cm wide white porcellanite margin to one dike. Photo courtesy of AF Cooper.

Doleritic sills, emplaced at very shallow levels, are a widespread feature of the volcanic field ( and ). The sills intrude Eocene sandstone at Tokarahi near Maerewhenua, limestone at Clarks Mill near Maheno, mudstone at Mt Charles, and siltstone and mudstone on the Moeraki Peninsula (Benson Citation1943, Citation1944) (). Dolerite at Round Hill may also be a sill (Benson Citation1943), although this has not been confirmed. The most regionally extensive of these sills are the ∼20 m thick columnar-jointed intrusion at Tokarahi (C), which may have been emplaced at the same time as Basalt Hill to the north (), and the ∼ 50 m thick Mt Charles Sill (D). Each of these sills is estimated to have an extent of over ∼25 km2 and volumes on the order of 1–2 km3 (Coombs et al. Citation1986). The Mt Charles Sill is underlain by flaggy concretion-bearing mudstone with the sill itself grading upwards from (weathered) olivine dolerite into quartz dolerite (Benson Citation1943; Coombs et al. Citation1986). The best studied sills, however, are those on the Moeraki Peninsula. Here, the Tawhiroko Sill is at least 50 m thick (the top having been lost to erosion) and has olivine dolerite at the base but a pegmatitic quartz dolerite core (Benson Citation1943, Citation1944). It also contains horizons remarkably rich in schist xenoliths, locally size-graded in bands within the sill (E) despite Otago Schist being 10–100 m below the sill. Nakamura and Coombs (Citation1973) interpreted chemical zoning in clinopyroxene grains in this sill to be due to progressive crystallisation of the magma. Benson (Citation1943) attributed the chemical differentiation in the Waiareka-Deborah sills to gravitational settling, but the differentiation could also be due to emplacement of variably evolved magmas injected as different batches. This would better explain the occurrence of schist-rich horizons at Moeraki. Coombs and Roedder (Citation1994) reported the occurrence of CO2 inclusions in plagioclase microphenocrysts from some Moeraki Peninsula sill rocks and inferred these to represent immiscible CO2 droplets in the melt at ∼ 10 km depth.

Doleritic dikes are widespread but exposures are concentrated on Moeraki Peninsula and around Enfield (Benson Citation1943; Coombs et al. Citation1986; Andrews Citation2003). Dikes on Moeraki Peninsula are mostly < 1 m wide, but range up to several metres width for those extending in outcrop for many 10s of metres. Dikes characteristically show evidence for multiple phases of injection (F). Marginal porcellanite from one Moeraki Peninsula dike was apparently quarried for tools by Māori prior to the arrival of Europeans (Benson Citation1943). A complication is that some dikes cropping out in the area mostly occupied by the Waiareka-Deborah Volcanic Field may belong to the younger alkaline Dunedin Volcanic Group (Coombs et al. Citation2008; Scott et al. Citation2020); chemical and isotopic analysis is required to test dike heritage.

The mantle under the Waiareka-Deborah Volcanic Field

Mantle xenolithic material occurs in five locations (Scott Citation2020). Very small (∼1 cm) spinel peridotite fragments occur in tuff beds above the intact pillow lava at Boatmans Harbour (Coombs et al. Citation1986; Moorhouse Citation2015) (A). The second occurrence is in lapilli tuff ∼ 2.5 km east of Round Hill, which is packed with moderately altered spinel peridotite xenoliths up to ∼ 8 cm in diameter (Scott et al. Citation2014b). A third occurrence is in thin lapilli tuff and tuff breccia beds directly beneath Ototara Limestone at Alma. These beds contain spinel peridotite and garnet pyroxenite xenoliths, as well as garnet, kaersutite, anorthoclase and augite megacrysts (Reay et al. Citation2002; Scott et al. Citation2014b). The peridotite clasts from Alma are extremely altered, with the mantle orthopyroxene and olivine having been completely converted to clay (Scott et al. Citation2014b). It is not known whether spinel lherzolite-bearing basaltic boulders in the river at Five Forks originate from the Dunedin Volcanic Group or Waiareka-Deborah-Volcanic Field.

In contrast to the aforementioned xenolith occurrences, the intra-vent lapilli tuff breccia enclosed within bedded surtseyan lapilli tuff on a marine platform at Kakanui (A) contains a spectacular array of xenocrystic and xenolithic material set in a calcite and zeolite-cemented primary volcaniclastic matrix (B) (Mantell Citation1850; Dickey Citation1968a, Dickey Citation1968b; Reay and Sipiera Citation1987; White and Houghton, Citation2006). The most xenolith-rich location, South Head, is the centre of a single surtseyan volcano (Corcoran and Moore Citation2008). Here, the Kakanui Mineral Breccia contains spinel lherzolite and harzburgite, cut in places by amphibole-phlogopite-bearing veins, as xenoliths up to about 20 cm in diameter (but commonly much smaller and usually heavily replaced by carbonate) within thin-rinded bombs of melanephelinite (C, D) (Dickey Citation1968a; Reay and Sipiera Citation1987; Klemme Citation2004; McCoy-West et al. Citation2013; Scott et al. Citation2014b). These xenoliths are fragments of the mantle lithosphere from under the volcanic field. Olivine (Mg# = 100*Mg/(Mg + Fe) = 89.3–91.0) and spinel (Cr# = 100*Cr/(Cr + Al) = 8.8–44.4) compositions from lherzolites and harzburgites indicate that this mantle is fairly to moderately fertile (Reay and Sipiera Citation1987; McCoy-West et al. Citation2013; Scott et al. Citation2014b; Moorhouse Citation2015). A single whole rock Os analysis (187Os/188Os = 0.11953; McCoy-West et al. Citation2013) and three clinopyroxene Hf analyses (εHf = +100 to +14; Scott et al. Citation2014b) indicate depletion ages that range back to Proterozoic time; the geological significance of the old ages remains debated, with suggestions that the underlying mantle is either a vast block of ancient lithosphere (McCoy-West et al. Citation2013) or that it is young lithosphere with embedded ancient fragments (Liu et al. Citation2015; Scott et al. Citation2019). Enriched clinopyroxene trace element data show that the underlying mantle was metasomatized after depletion (Scott et al. Citation2014b; McCoy-West et al. Citation2015), and this is supported by the occurrence of apatite crystals and crystallised fluoride melt in one xenolith (Klemme Citation2004) and the amphibole-phlogopite-bearing veins (D).

Figure 5. Kakanui Mineral Breccia outcrops and hand specimens. A, Aerial view of the Kakanui Mineral Breccia, located near the centre of a submarine diatreme. The breccia is overlain by loess and obscured by the Kakanui township. Image from Google Earth. White arcuate lines indicate bedding orientations. B, The tuff breccia facies contains xenoliths of lherzolite (commonly carbonated), garnet pyroxenite, nephelinite with amphibole megacrysts/xenocrysts, all set within a calcite-zeolite cement. Fine-grained dark clasts are basaltic/melanephelinite fragments. C, Rare peridotite xenoliths up to about 20 cm in size have a thin nephelinite annulus. Photo courtesy of DG Pearson. D, Peridotite xenoliths have cross-cutting amphibole veinlets. E, Megacrysts form an important component to the mineral breccia. The most common are amphibole, followed by anorthoclase. Garnet is rare. Many grains have a rounded nature with a polished surface, possibly due to abrasion during entrainment and emplacement. F, View of granulite xenoliths. The xenoliths provide insight into the middle-lower crust under the region. Felsic granulites are most common, and have been found to preserve evidence for Late Cretaceous ultra-high temperature metamorphism of the lower Otago crust.

Figure 5. Kakanui Mineral Breccia outcrops and hand specimens. A, Aerial view of the Kakanui Mineral Breccia, located near the centre of a submarine diatreme. The breccia is overlain by loess and obscured by the Kakanui township. Image from Google Earth. White arcuate lines indicate bedding orientations. B, The tuff breccia facies contains xenoliths of lherzolite (commonly carbonated), garnet pyroxenite, nephelinite with amphibole megacrysts/xenocrysts, all set within a calcite-zeolite cement. Fine-grained dark clasts are basaltic/melanephelinite fragments. C, Rare peridotite xenoliths up to about 20 cm in size have a thin nephelinite annulus. Photo courtesy of DG Pearson. D, Peridotite xenoliths have cross-cutting amphibole veinlets. E, Megacrysts form an important component to the mineral breccia. The most common are amphibole, followed by anorthoclase. Garnet is rare. Many grains have a rounded nature with a polished surface, possibly due to abrasion during entrainment and emplacement. F, View of granulite xenoliths. The xenoliths provide insight into the middle-lower crust under the region. Felsic granulites are most common, and have been found to preserve evidence for Late Cretaceous ultra-high temperature metamorphism of the lower Otago crust.

Clinopyroxenite and garnet pyroxenite are common in the Kakanui Mineral Breccia (B) (Mason Citation1968; White et al. Citation1972; Reay and Sipiera Citation1987; Zack et al. Citation1997; Sun Citation2018). The clinopyroxenites have attracted little attention. However, the garnet pyroxenite xenoliths, which were considered to be eclogite but lack sufficiently aluminous clinopyroxene for this classification, are mainly picro-basalt to basaltic in composition (). Metamorphic equilibration temperatures are in excess of 950°C (Zack et al. Citation1997), and some record Cretaceous Lu-Hf metamorphic ages (Scott, unpublished data). This latter result means that the garnet pyroxenites are not related to Eocene-Oligocene magmatism in the field.

Figure 6. Total alkali versus SiO2 diagram showing data from Cenozoic intraplate volcanic provinces in Otago. Waiareka-Deborah lava and sill data are presented in ; the Dunedin Volcanic Group data are from the compilation presented by Scott et al. (Citation2020); Alpine Dike Swarm data are from the compilation of Cooper (Citation2020); and the Kakanui garnet pyroxenite xenolith data are unpublished.

Figure 6. Total alkali versus SiO2 diagram showing data from Cenozoic intraplate volcanic provinces in Otago. Waiareka-Deborah lava and sill data are presented in Table 1; the Dunedin Volcanic Group data are from the compilation presented by Scott et al. (Citation2020); Alpine Dike Swarm data are from the compilation of Cooper (Citation2020); and the Kakanui garnet pyroxenite xenolith data are unpublished.

The tuff breccia at Kakanui contains a remarkable array of xenocrysts. These are, in order of estimated decreasing occurrence: kaersutitic to pargasitic amphibole, anorthoclase, augite, pyrope (E), as well as biotite, ilmenite and apatite (Dickey Citation1968a; Reay and Wood Citation1974; Merrill and Wyllie Citation1975; Wallace Citation1977; Reay and Sipiera Citation1987; Reay et al. Citation1989, Citation1993; Fulmer et al. Citation2010; Urosevic et al. Citation2018). Many of the xenocryst margins are curved and shiny in outcrop, and many are enclosed in melanephelinite lapilli. Several large crystals of kaersutite, pyrope, anorthoclase and augite have been distributed globally as chemical micro-analytical standards and are available from the Smithsonian Institute upon request. Whilst it is reported that some megacrysts reach 20 cm in length (Reay and Sipiera Citation1987), exposed grains of this size have long been collected and only the amphiboles, anorthoclase and augite of up to about 5 cm diameter can now be readily found. Lu-Hf isotope data indicate that pyrope was in equilibrium with the melanephelinite, whereas the amphibole megacrysts were not (Fulmer et al. Citation2010); we have no comparable information for other megacryst species.

An additional feature of the Kakanui Mineral Breccia is that it contains abundant granulite xenoliths. These are geodynamically significant for Otago because they preserve evidence for an ultra-high-temperature metamorphic event (∼900°C) that affected the base of the Otago crust at ∼92 Ma (Jacob et al. Citation2017), an event for which there is no known evidence at the Otago surface. Among the granulite xenoliths, weakly foliated felsic ones are common (F) and mafic garnet-pyroxene granulites are rare. Jacob et al. (Citation2017) interpreted the granulites to have formed by bottom-up heating of the Otago crust caused by Late Cretaceous slab rollback coupled with concomitant regional crustal extension.

Age

Age determinations for Waiareka-Deborah volcanism have historically been established by paleontological means because deposits are enclosed in strata containing molluscs, brachiopods and foraminifera. These indicate that volcanism mostly occurred in the New Zealand’s Eocene Runangan (36.4–34.6 Ma) and Late Eocene to Early Oligocene Whaingaroan (34.6–27.6 Ma) stages (Benson Citation1943; Gage Citation1957; Coombs et al. Citation1986; Hicks Citation2014). Published radiometric dates are restricted to 40Ar/39Ar analyses of 33.6 ± 1.8 Ma (Round Hill), 34.0 ± 0.6 Ma (Maheno) and 34.2 ± 0.4 Ma and 34.3 ± 0.9 Ma (Boatmans Harbour) (Hoernle et al. Citation2006) (), all of which confirm Runangan to Early Whaingaroan emplacement. Two basaltic clasts in tuff breccia at Bridge Point gave ages of 39.5 ± 1.8 Ma (Eocene Bortonian or Kaiatan) and 34.3 ± 0.5 Ma (Runangan) (Hoernle et al. Citation2006); the older age incorporates 69% of the 39Ar and is therefore probably significant despite immediately overlying fossiliferous volcaniclastic deposits being Runangan in age (Lee et al. Citation1997; Hicks Citation2014). Two fractions of the same megacrystic kaersutite from Kakanui, KK1, prepared by Reay et al. (Citation1989), yielded Runangan-Whaingaroan ages of 33.7 ± 0.3 and 34.1 ± 0.1 Ma (Hoernle et al. Citation2006). Fulmer et al. (Citation2010) report an approximate isochron age of 34 Ma for garnet, amphibole and whole-rock melanephelinite from Kakanui. In summary, palaeontological and radiometric data confirm that most Waiareka-Deborah volcanism occurred over a geologically short duration that definitely initiated in the Eocene and probably extended into the Early Oligocene.

Geochemistry

There have been few whole-rock geochemical studies of the Waiareka-Deborah Volcanic Field because most volcaniclastic deposits are altered (Coombs et al. Citation1986; Moorhouse et al. Citation2015). We supplement the existing major and trace data of Hoernle et al. (Citation2006; n = 3), Hoke et al. (Citation2000; n = 1), Klemme (Citation2004; n = 1), Reay and Sipiera (Citation1987; who reported the average of six clasts), and trace elements for one sample (Fulmer et al. Citation2010), with a further 9 analyses from sills and flows sampled across the volcanic field (). The whole-rock data were obtained from ALS Minerals in Brisbane on fused Li-borate disks, following the method reported by Scott et al. (Citation2020). Although these data enable characterisation of the sampled coherent volcanic rocks, their volume is far exceeded by fragmental rocks; results may not be representative for the field as a whole. The samples are also commonly slightly altered, as indicated by LOI values reaching up to 5.5 wt%; there has probably been loss of some mobile elements.

Table 1. New and published whole rock geochemical and isotopic data. Oxides are in weight %; trace elements are in ppm; locations are latitude, longitude.

Using the total alkali versus silicate classification (TAS: Le Maitre Citation2002), the Waiareka-Deborah whole-rock compositions of coherent volcanic rocks (lavas, dikes or sills) are sub-alkaline basalt to basaltic andesite, except for the alkaline melanephelinite and basanitic clasts in the Kakanui Mineral Breccia (). The occurrence of both alkaline and sub-alkaline components is consistent with the chemistry of glass preserved in volcaniclastic deposits (Coombs et al. Citation1986). The Mg# (=100*Mg/(Mg + Fe)) of the sub-alkaline volcanic rocks ranges from 42 to 60, which is very similar to the Kakanui melanephelinite (57-56) but higher than the Kakanui basanites (40-33) (). In samples not belonging to the sills that are clearly chemically differentiated, the high MgO (>7 wt%), Cr (∼300 ppm) and Ni (>250 ppm) indicates that these rocks are fairly primitive. Trace element data from the two Kakanui analyses show the rock is slightly to significantly more enriched incompatible element abundances than are the sub-alkaline rocks (A, B). La/Lu(N) values for the sub-alkaline components (5.1–15.8) are less than for the alkaline Kakanui components (20.1–57.7). Some sub-alkaline samples display negative K and Rb anomalies, and all sub-alkaline rocks display positive Sr and P anomalies.

Figure 7. A, Trace element data for the Waiareka-Deborah Volcanic Field, compared to summaries of mafic magmas in the nearby Dunedin Volcanic Group (Scott et al. Citation2020) and Alpine Dike Swarm (Cooper Citation2020). B, REE data for Waiareka-Deborah sub-alkaline and alkaline magmas. Normalising values in both diagrams are from Sun and McDonough (Citation1989). Otago Schist greyschist composition is from Scanlan et al. (Citation2020).

Figure 7. A, Trace element data for the Waiareka-Deborah Volcanic Field, compared to summaries of mafic magmas in the nearby Dunedin Volcanic Group (Scott et al. Citation2020) and Alpine Dike Swarm (Cooper Citation2020). B, REE data for Waiareka-Deborah sub-alkaline and alkaline magmas. Normalising values in both diagrams are from Sun and McDonough (Citation1989). Otago Schist greyschist composition is from Scanlan et al. (Citation2020).

Sr-Nd-Pb isotopes

The mantle source for the Waiareka-Deborah volcanic field has not been investigated in any detail prior to our work. The data of Hoernle et al. (Citation2006; Sr-Nd-Pb; n = 3) and Fulmer et al. (Citation2010; Hf, n = 1), are here supplemented by Sr-Nd-Pb for 9 samples (). The new isotope data were obtained at the University of Cape Town in South Africa following the method reported by Scott et al. (Citation2020).

The 87Sr/86Sr(m; measured) values for sub-alkaline Waiareka-Deborah volcanic rocks cluster between 0.70321 and 0.70387, overlapping those for the Kakanui melanephelinite (0.70342). Low Rb/Sr means that age correction to 34 Ma makes little or no difference to the 87Sr/86Sr values (87Sr/86Sr(i) sub-alkaline = 0.70310–0.70385; alkaline = 0.70339) (A). The Tawhiroko and Moeraki sills contain locally abundant pyrometamorphosed Otago schist xenoliths in which Rb-rich micas appear to have been assimilated (E), but the isotopic similarity between these samples and other sub-alkaline flows indicates that the assimilation process was localised and the isotopic compositions of the analysed samples were not significantly changed. Alternatively, lack of a geochemical signature indicating crustal contamination may indicate that schist-rich layers formed as late injections into an inflating sill (Marsh Citation1996).

Figure 8. Bulk rock isotopic data for the Waiareka-Deborah Volcanic Field compared to other volcanic fields in Otago and clinopyroxene in mantle xenoliths of those fields. A, 87Sr/86Sr(i; initial) versus 143Nd/144Nd(i) shows the Waiareka-Deborah Volcanic Field tends to be slightly more radiogenic in Sr than the bulk of the other intraplate provinces. B, 207Pb/204Pb versus 208Pb/204Pb(m; measured) highlights the distinct nature of the Waiareka-Deborah Volcanic Field in 207Pb/204Pb isotopes to the other intraplate provinces and the mantle lithosphere, except for basalts in the Maniototo in the Dunedin Volcanic Group. C, 206Pb/204Pb versus 208Pb/204Pb(m) illustrates the mostly less radiogenic nature of the mantle source to the Waiareka-Deborah Volcanic Group compared to other Otago intraplate provinces. Data sources are: Waiareka-Deborah Volcanic Field rocks (this study; Hoernle et al. Citation2006) and xenoliths (McCoy-West et al. Citation2016); Dunedin Volcanic Group rocks (Price et al. Citation2003; Hoernle et al. Citation2006; Sprung et al. Citation2007; Scanlan et al. Citation2020; Scott et al. Citation2020) and its peridotite xenoliths (Scott et al. Citation2014a; Scott et al. Citation2014b; McCoy-West et al. Citation2016; Dalton et al. Citation2017); Alpine Dike Swarm rocks (Barreiro and Cooper Citation1987) and peridotite xenoliths (Scott et al. Citation2014b, Citation2016); crustal Sr-Nd-Pb data are from Scanlan et al. (Citation2018, 2020) and Scott et al. (Citation2020).

Figure 8. Bulk rock isotopic data for the Waiareka-Deborah Volcanic Field compared to other volcanic fields in Otago and clinopyroxene in mantle xenoliths of those fields. A, 87Sr/86Sr(i; initial) versus 143Nd/144Nd(i) shows the Waiareka-Deborah Volcanic Field tends to be slightly more radiogenic in Sr than the bulk of the other intraplate provinces. B, 207Pb/204Pb versus 208Pb/204Pb(m; measured) highlights the distinct nature of the Waiareka-Deborah Volcanic Field in 207Pb/204Pb isotopes to the other intraplate provinces and the mantle lithosphere, except for basalts in the Maniototo in the Dunedin Volcanic Group. C, 206Pb/204Pb versus 208Pb/204Pb(m) illustrates the mostly less radiogenic nature of the mantle source to the Waiareka-Deborah Volcanic Group compared to other Otago intraplate provinces. Data sources are: Waiareka-Deborah Volcanic Field rocks (this study; Hoernle et al. Citation2006) and xenoliths (McCoy-West et al. Citation2016); Dunedin Volcanic Group rocks (Price et al. Citation2003; Hoernle et al. Citation2006; Sprung et al. Citation2007; Scanlan et al. Citation2020; Scott et al. Citation2020) and its peridotite xenoliths (Scott et al. Citation2014a; Scott et al. Citation2014b; McCoy-West et al. Citation2016; Dalton et al. Citation2017); Alpine Dike Swarm rocks (Barreiro and Cooper Citation1987) and peridotite xenoliths (Scott et al. Citation2014b, Citation2016); crustal Sr-Nd-Pb data are from Scanlan et al. (Citation2018, 2020) and Scott et al. (Citation2020).

The sub-alkaline 143Nd/144Nd(m) data have a tight moderately radiogenic cluster (0.51284–0.51288; εNd = +4.7 to +3.9), with the melanephelinite plotting at the more radiogenic end of this dataset (0.51290; εNd = +5.1) (A). Low Sm/Nd means that age correction to 34 Ma for all these rocks makes little difference (0.51281–0.51288; εNd = +4.7 to +3.3).

The Pb isotopes show little variation across the volcanic field but do indicate a subtle distinction between sub-alkaline rocks and the melanephelinite (). There are, however, insufficient high-quality U-Th-Pb trace element data to apply an age correction, so only uncorrected data are reported. 206Pb/204Pb of the sub-alkaline rocks (19.042–19.247) is less radiogenic than in the Kakanui melanephelinite (19.313), as are 208Pb/204Pb ratios (sub-alkaline = 38.763–38.921; melanephelinite = 38.948). The sub-alkaline 207Pb/204Pb data (15.617–15.649) overlap with those from melanephelinite (15.622), yet are distinct in 207Pb/204Pb versus 206Pb/204Pb space (B, C).

Clinopyroxene separates have been analysed for Sr, Nd and Pb from 8 peridotite xenoliths (Scott et al. Citation2014b, n = 7; McCoy-West et al. Citation2016, n = 1). This suite comprises two xenoliths from Round Hill, two from Alma and four from Kakanui. 87Sr/86Sr(i) in the peridotite xenoliths (0.70242-0.70316) is less radiogenic than solidified host magmas, whereas 143Nd/144Nd(i) data (0.51360–0.51285) overlap. The extremely unradiogenic Sr and radiogenic Nd of several samples testify to portions of this mantle lithosphere having been isotopically isolated for long periods of time, although most of the samples appear to preserve a metasomatic isotopic composition that has been shown to be widespread in the Otago mantle (Scott et al. Citation2014a; Scott et al. Citation2014b; McCoy-West et al. Citation2016; Dalton et al. Citation2017; Scott et al. Citation2020). Peridotite clinopyroxene Pb isotope data extend from extremely unradiogenic through to radiogenic values (206Pb/204Pb = 17.77–20.2, 207Pb/204Pb = 15.41–15.65 and 208Pb/204Pb = 37.47–39.46) and only the Kakanui melanephelinite overlaps with this array. In detail, however, the xenolith Pb array is pulled to unradiogenic composition by an extremely unradiogenic single sample; exclusion of this would mean that no volcanic component so far analysed in the Waiareka-Deborah Volcanic Field overlaps with the Pb isotopic composition of analysed North Otago mantle xenoliths.

Discussion

Shallow-marine volcanism and shallowly emplaced sills

The dominance of volcaniclastic deposits formed by largely shallow-submarine eruptions is atypical of most studied volcanic fields in intraplate continental settings. Volcanologically analogous submarine volcanic fields formed on the ocean floor, such as the Vestmann Islands in Iceland (Jakobsson Citation1968, Citation1979) or North Arch in Hawaii (e.g. Frey et al. Citation2000; Davis and Clague Citation2006), are difficult to investigate in detail. In terms of volcano morphology and the general setting of the Waiareka-Deborah volcanic field, the Vestmann Islands in Iceland (Jakobsson Citation1968, Citation1979) offer a partial analogue. The depth of the Vestmann Island shelf, ∼100-150 metres, is very similar to depths inferred for the Waiareka-Deborah (mostly ‘inner shelf’, but up to ∼200 m; Hicks Citation2014). In both fields, the shallowest depths would have been at the coastlines of temporary islands. Observations during the eruption of Surtsey, and of other surtseyan volcanoes, show quick growth to sea level and above to form islands, with rapid, often repeated, planation during or soon after eruption (Hoffmeister et al. Citation1929; Thorarinsson Citation1967; Johnson and Tuni Citation1987). After formation, the Waiareka-Deborah volcanoes would have been reduced to shoals at fair-weather wavebase or below, over the course of months or years.

A significant feature of this field is the extent and volume of shallowly emplaced sills. Individual sills have been mapped across tens of km, whereas the extent of significant deposits from single volcanic centres appears to only be a few km or less. Coombs et al. (Citation1986) inferred that the ‘ … total volume of all the currently visible intrusives … may have reached 10 km3 .’ This is a volume that may be half that of known Waiareka-Deborah volcaniclastic deposits. The sills were emplaced into a sedimentary succession that was on the order of only ∼ 100 to perhaps several hundred metres thick (Thompson et al. Citation2014), which means the ∼ 50 m thickness of several sills is very large relative to the overlying host rocks. For example, the Tawhiroko sill at Moeraki is closely associated with deformed submarine pyroclastic deposits, quite possibly co-genetic, from which it is separated by tens of metres or less of marine mudstone. The thinness of the sills is in stark contrast to those in thick basinal deposits of the North Sea, for example (Hospers and Ediriweera Citation1991; Thomson and Schofield Citation2008); it is not clear whether this is simply the result of smaller volumes of magma delivered to the Waiareka-Deboorah sills, or reflects an additional control by the thin basinal host deposits. The thinness of the Oamaru region's sedimentary succession reflects a long period of slow deposition and occasional hiatuses during which hardgrounds formed (Thompson et al. Citation2014). Features that could have promoted sill emplacement include readily deformable near-surface deposits (Galland et al. Citation2018), or rheological differences among the sedimentary units (Kavanagh et al. Citation2006). Understanding why this thin sedimentary sequence captured such substantial proportions of the magma sent toward the surface during Waiareka-Deborah time is a topic for further research.

Mantle sources for Waiareka-Deborah magmatism

Together the new major element () and trace element () data emphasise that the majority of analysed Debora-Waiareka Volcanic Formation is distinct from the slightly younger Alpine Dike Swarm and Dunedin Volcanic Group, which are Cenozoic intraplate volcanic provinces in west and east Otago respectively ( inset). The lower incompatible element abundances of the Waiareka-Deborah suite cannot easily be explained as a result of crustal contamination because, as has already been argued for New Zealand intraplate rocks (Sprung et al. Citation2007), the continental crust is enriched in many of these elements (e.g. Otago Schist composition in ) and assimilation should lead to magmas with higher incompatible element abundances. Furthermore, there is no clear mixing array between crust and mantle in Pb isotopic space (C). While the Waiareka-Deborah data are similar to the cluster of transitional alkaline–sub-alkaline Maniototo basalts in the NW corner of Dunedin Volcanic Group (), the Maniototo basalts lack positive Sr and P anomalies and have ubiquitous negative K anomalies (A) (Scott et al. Citation2020). While the positive Sr anomalies in the Waiareka-Deborah Volcanic Field (A) could be due to plagioclase accumulation, this seems unlikely because Eu, an element that is also preferentially taken up in feldspar, shows only a very weak positive anomaly (B). There is also no correlation between Sr and Mg# (not shown), which means that the Sr and P anomalies are not tied to crystal fractionation. A weak positive correlation of Sr and P (not shown) could be due to apatite in the mantle source. The positive Sr and P, the high SiO2 content, low incompatible elements and absence of ubiquitous K anomaly compared to other Otago primitive intraplate basalts indicates the Waiareka-Deborah Volcanic Field magmas may have different mantle sources.

A distinct mantle source to the Waiareka-Deborah Volcanic Field compared to other Otago intraplate provinces is supported by radiogenic isotope data. 87Sr/86Sr data show that Waiareka-Deborah rocks are more radiogenic than most of the Dunedin Volcanic Group and Alpine Dike Swarm (). Although the higher 87Sr/86Sr of this province could be a result of seawater interaction during submarine emplacement, this could not account for the distinct Pb isotopes (A, B) and in any case interaction with seawater should give more extreme Sr isotope ratios due to the elevated Eocene seawater Sr isotope composition (e.g. Nelson et al. Citation2004). If interaction with seawater controlled the isotopic composition, it would mean that the sills were contaminated despite not breaching the surface and would require interaction with marine pore water to an equivalent degree across the entire volcanic field. Another option is that the elevated Sr isotope ratios are due to Otago Schist assimilation by an alkaline magma; however, this would have required large amounts of Otago Schist (since it has a much lower Sr content than the volcanic rocks) (Scanlan et al. Citation2020) and, in any case, this process would not deplete the incompatible element abundances such as Rb, Ba, Th, U, K etc. (). Furthermore, although the Waiareka-Deborah 207Pb/204Pb versus 206Pb/204Pb data trend towards continental crust (B), the 206Pb/204Pb versus 208Pb/204Pb data plot along the Northern Hemisphere Reference Line (C). The mantle sources to this volcanic province are also distinct from almost all the fragments of peridotitic mantle that have been exhumed from beneath the West, East and North Otago, which represent snapshots of the composition of the lithospheric mantle. Therefore, unlike the magmas of the Dunedin Volcanic Group and Alpine Dike Swarm, which have Sr, Nd and Pb isotopic compositions overlapping the mantle lithosphere, the isotopic properties of the Waiareka-Deborah Volcanic Field magmas appear to require derivation from a lithospheric mantle component that is either very rare, or derivation from the asthenosphere and rapid ascent without interaction with the lithospheric mantle.

Peridotite geothermobarometry and comparison with geotherms indicates that the Otago New Zealand lithosphere reaches about ∼70 km depth (Scott et al. Citation2014a; Scott et al. Citation2014b). If this is the case, then partial melting in the asthenosphere should involve partitioning of elements into refractory garnet and this should be chemically detectable. Sprung et al. (Citation2007) pointed out that Lu/Hf fractionates more than Sm/Nd during melting in the presence of garnet. The very small variation in Sm/Nd (0.22–0.27) but large variation in Lu/Hf (0.045–0.09) in the Waiareka-Deborah rocks is therefore consistent with garnet having participated during melting. Furthermore, the alkaline basalts in the Otago area have near-ubiquitous negative K and Ti anomalies that represent either derivation from a K, Ti-depleted mantle (e.g. Timm et al. Citation2010) or amphibole ± phlogopite were residual during melting in the mantle source (Panter et al. Citation2006; Sprung et al. Citation2007; Pilet et al. Citation2008); however, these hydrous phases are thermally stable at lithospheric mantle temperatures of less than ∼ 1200°C. The Waiareka-Deborah Volcanic Field sub-alkaline basalts lack ubiquitous K or Ti anomalies (A) and this observation coupled with the distinct isotopes suggests that they lack the lithospheric mantle signature apparently prevalent in the other Otago intraplate provinces.

A further feature relevant to the petrogenesis of the Waiareka-Deborah Volcanic Field is that the alkaline Kakanui Mineral Breccia melanephelinite (1) plots in the same (or very similar in the case of Pb) isotopic space as the sub-alkaline rocks and is (2) distinct from almost all the other alkaline basalts in Otago (). This observation, coupled with overlapping Ar-Ar age, implies that the alkaline magmatism has the same (or very similar) isotopic reservoir to the sub-alkaline components and indicates that the petrogenesis of both types is likely related. Very low degrees (several percent) partial melting of NZ metasomatised peridotite has been modelled to be capable of giving rise to the distinctive trace element composition of alkaline basalt (Scott et al. Citation2016). With increasing degrees of melting, the incompatible trace element abundances of successive magmas are lowered and this results in a trace element composition similar to that of a sub-alkaline basalt. However, while the Kakanui Mineral Breccia magma has a prominent positive P anomaly, it lacks a positive Sr anomaly and also has prominent depletions in K, Zr, Hf and Ti. These latter element properties, which typically are interpreted to reflect source characteristics of the magmas (e.g. Panter et al. Citation2006; Sprung et al. Citation2007), imply that it is unlikely that both types of magma could be simply derived from different degrees of melting of a single mantle source and requires that the alkaline magmas have a distinct component that was melted at the same time.

Although sub-alkaline magmas have been experimentally generated by melting of mantle pyroxenite (e.g. Pertermann and Hirschmann Citation2003), the high Cr (∼300 ppm) in Waiareka-Deborah volcanic rocks (excluding those of the differentiated sills) indicates that peridotite has played an important role in the melting process. Furthermore, although garnet pyroxenite must reside in the mantle beneath the Deborah-Waiareka Volcanic Field (B), their Sr, Nd and Pb data are more similar to the metasomatised mantle peridotites and alkaline volcanic rocks elsewhere in Otago than sub-alkaline volcanic components (Scott, unpublished data). In any case, there is no pyroxenite melting model that would alone account for the formation of SiO2-undersaturated alkaline magma (e.g. Pilet Citation2015). The Kakanui Mineral Breccia melanephelinite could represent a hybrid magma that originated from the asthenosphere but assimilated metasomatised lithospheric mantle (e.g. Sprung et al. Citation2007). However, if this were the case, then the high Sr content of the melanephelinite (1379 ppm) compared to the sub-alkaline rocks (average of 456 ppm) requires that the mantle lithosphere contaminant had (1) a similar 87Sr/86Sr composition to the sub-alkaline source because the much higher abundance Sr abundance would lead to a large isotopic contribution, and (2) had sufficient residual amphibole and/or phlogopite to generate the negative K and Ti anomalies. The petrogenesis of the alkaline components in this dominantly sub-alkaline volcanic province is unresolved.

Conclusions

The intraplate Waiareka-Deborah Volcanic Field surrounding Oamaru in the South Island preserves a superb record of Eocene-Oligocene submarine volcanism on the Zealandia continental shelf. The on-land extent is ∼890 km2, and the offshore extent could, potentially, be nearly four times that area. The volcaniclastic beds, which are extensive and well-exposed along the coastline, show that the volcanic field formed from small volume, short duration eruptions, with volcanoes built from tephra deposited by eruption-fed density currents, and more widespread tephras emplaced from ash initially transported in subaerial eruption plumes (). The magma also crystallised as sills, pillow lavas and dikes. The sills are extensive, reaching up to ∼25 km2, were emplaced at shallow levels in the thin sedimentary veneer that covered Otago at the time of formation, and are commonly chemically differentiated. Volcaniclastic glass chemistry and bulk rock composition of sills and pillow lavas show that the volcanic field mainly erupted sub-alkaline magmas, although there were minor coeval alkaline components, the best known of which is the xenocryst- and xenolith-rich Kakanui melanephelinite. Radiogenic isotopic properties of the sills and dikes and one melanephelinite clast indicate that the mantle source for the magmas was distinct from the nearby intraplate alkaline Oligocene Alpine Dike Swarm and the adjacent and spatially overlapping Oligocene-Miocene Dunedin Volcanic Group. Intraplate volcanism in Otago therefore has multiple mantle sources.

Acknowledgments

Alan Bischoff, Alan Cooper, Alessio Pontesilli, Daphne Lee and Marco Brenna are thanked for comments on various drafts. Stephen Read helped with . Jenni Hopkins and John Smellie are thanked for reviews.

Data availability statement

All data related to this paper are summarised in , or in the cited papers.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Andrews B. 2003. Eruptive and depositional mechanisms of an Eocene shallow submarine volcano, Moeraki Peninsula, New Zealand. Washington DC American Geophysical Union Geophysical Monograph Series. 140:179–188.
  • Barreiro BA, Cooper AF. 1987. A Sr, Nd, and Pb isotope study of alkaline lamprophyres and related rocks from Westland and Otago, South Island, New Zealand. Geological Society of America Special Paper. 215:115–126. doi: 10.1130/SPE215-p115
  • Benson WN. 1943. The basic igneous rocks of eastern Otago and their tectonic environment. Part IV – The mid-Tertiary basalts, tholeiites, and dolerites of north-eastern Otago. Transactions of the Royal Society of New Zealand. 73:116–138.
  • Benson WN. 1944. The basic igneous rocks of eastern Otago and their tectonic environment. Part IV - The mid-Tertiary basalts, tholeiites, and dolerites of north-eastern Otago. Section B – Petrology, with special reference to the crystallisation of pyroxene. Transactions of the Royal Society of New Zealand. 74:71–123.
  • Bischoff A, Barrier A, Beggs M, Nicol A, Cole J, Sahoo T. 2020. Magmatic and tectonic interactions revealed by buried volcanoes in Te Riu-a-Māui/Zealandia sedimentary basins. New Zealand Journal of Geology and Geophysics.
  • Bradley WH. 1965. Vertical density currents. Science. 150:1423–1428. doi: 10.1126/science.150.3702.1423
  • Cas RAF, Landis CA, Fordyce RE. 1989. A monogenetic, Surtla-type, Surtseyan volcano from the Eocene-Oligocene Waiareka-Deborah volcanics, Otago, New Zealand: a model. Bulletin of Volcanology. 51:281–298. doi: 10.1007/BF01073517
  • Coombs DS, Adams CJ, Roser BP, Reay A. 2008. Geochronology and geochemistry of the Dunedin Volcanic Group, eastern Otago, New Zealand. New Zealand Journal of Geology and Geophysics. 51:195–218. doi: 10.1080/00288300809509860
  • Coombs DS, Cas RA, Kawachi Y, Landis CA, McDonough WF, Reay A. 1986. Cenozoic volcanism in North, East and Central Otago. Royal Society of New Zealand Bulletin. 23:278–312.
  • Coombs DS, Dickey JS. 1965. The early Tertiary petrographic province of north-east Otago: Waiareka and Deborah Volcanic formations. NZ Dept. Scientific and Industrial Research, Information Series. 51:38–54.
  • Coombs DS, Roedder E. 1994. On the significance of CO2 inclusions in plagioclase microphenocrysts in tholeiite from Moeraki, New Zealand. Bulletin of Volcanology. 56:23–28. doi: 10.1007/BF00279726
  • Cooper AF. 2020. Petrology and petrogenesis of an intraplate alkaline lamprophyre-phonolite-carbonatite association in the Alpine Dyke Swarm, New Zealand. New Zealand Journal of Geology and Geophysics.
  • Corcoran PL, Moore LN. 2008. Subaqueous eruption and shallow-water reworking of a small-volume Surtseyan edifice at Kakanui, New Zealand. Canadian Journal of Earth Sciences. 45:1469–1485. doi: 10.1139/E08-068
  • Dalton HB, Scott JM, Liu J, Waight TE, Pearson DG, Brenna M, le Roux PJ, Palin JM. 2017. Diffusion-zoned pyroxenes in an isotopically heterogeneous mantle lithosphere beneath the Dunedin Volcanic Group, New Zealand, and their implications for intraplate alkaline magma sources. Lithosphere. 9:463–475. doi: 10.1130/L631.1
  • Davis AS, Clague DA. 2006. Volcaniclastic deposits from the North Arch volcanic field, Hawaii: explosive fragmentation of alkalic lava at abyssal depths. Bulletin of Volcanology. 68(3):294–307. doi: 10.1007/s00445-005-0008-7
  • Dickey JS. 1968a. Eclogitic and other inclusions in the mineral breccia member of the Deborah volcanic formation at Kakanui, New Zealand. American Mineralogist. 53:1304–1319.
  • Dickey JS. 1968b. Observations on the Deborah volcanic formation near Kakanui, New Zealand. New Zealand Journal of Geology and Geophysics. 11:1159–1162. doi: 10.1080/00288306.1968.10420244
  • Edwards AR. 1991. The Oamaru Diatomite. New Zealand Geological Survey Paleontological Bulletin, 64 p.
  • Field BD, Browne GH, Davy BW, Herzer RH, Hoskins RH, Raine JI, Wilson GJ, Sewell RJ, Smale D, Watters WA. 1989. Cretaceous and Cenozoic sedimentary basins and geological evolution of the Canterbury region, South Island, New Zealand. Lower Hutt: New Zealand Geological Survey. New Zealand Geological Survey Basin Studies. 2:94.
  • Forsyth PJ. 2001. Geology of the Waitaki area. Institute of Geological & Nuclear Sciences 1:250 000 Geological Map 19. 1 sheet + 64 p. Lower Hutt: Institute of Geological & Nuclear Sciences Ltd.
  • Frey FA, Clague D, Mahoney JJ, Sinton JM. 2000. Volcanism at the edge of the Hawaiian plume: petrogenesis of submarine alkalic lavas from the North Arch volcanic field. Journal of Petrology. 41:667–691. doi: 10.1093/petrology/41.5.667
  • Fulmer EC, Nebel O, van Westrenen W. 2010. High-precision high field strength element partitioning between garnet, amphibole and alkaline melt from Kakanui, New Zealand. Geochimica et Cosmochimica Acta. 74:2741–2759. doi: 10.1016/j.gca.2010.02.020
  • Gage M. 1957. The geology of Waitaki subdivision. Issue 55. New Zealand Department of Scientific and Industrial Research.
  • Galland O, Bertelsen HS, Eide CH, Guldstrand F, Haug ØT, Leanza HA, Mair K, Palma O, Planke S, Rabbel O, et al. 2018. Chapter 5 - Storage and transport of magma in the layered crust—formation of sills and related flat-lying intrusions. In: Burchardt S, editor. Volcanic and igneous plumbing systems. Amsterdam: Elsevier; p. 113–138.
  • Hicks S. 2014. Paleoecology and sedimentology of the volcanically active Late Eocene continental shelf, Northeast Otago, New Zealand. PhD Thesis, University of Otago. 413p.
  • Hoernle K, White JDL, van den Bogaard P, Hauff F, Coombs DS, Werner R, Timm C, Garbe-Schönberg D, Reay A, Cooper AF. 2006. Cenozoic intraplate volcanism on New Zealand: Upwelling induced by lithospheric removal. Earth and Planetary Science Letters. 248:350–367. doi: 10.1016/j.epsl.2006.06.001
  • Hoffmeister JE, Ladd HS, Alling HL. 1929. Falcon Island. American Journal of Science. 18:461–471. doi: 10.2475/ajs.s5-18.108.461
  • Hoke L, Poreda R, Reay A, Weaver SD. 2000. The subcontinental mantle beneath southern New Zealand, characterized by helium isotopes in intraplate basalts and gas-rich springs. Geochimica et Cosmochimica Acta. 64:2489–2507. doi: 10.1016/S0016-7037(00)00346-X
  • Hospers J, Ediriweera KK. 1991. Depth and configuration of the crystalline basement in the Viking Graben area, Northern North Sea. Journal of the Geological Society. 148(2):261–265. doi: 10.1144/gsjgs.148.2.0261
  • Jacob JB, Scott JM, Turnbull RE, Tarling MS, Sagar MW. 2017. High-to ultrahigh-temperature metamorphism in the lower crust: An example resulting from Hikurangi Plateau collision and slab rollback in New Zealand. Journal of Metamorphic Geology. 35:831–853. doi: 10.1111/jmg.12257
  • Jacobs CT, Goldin TJ, Collins GS, Piggott MD, Kramer SC, Melosh HJ, Wilson CRG, Allison PA. 2015. An improved quantitative measure of the tendency for volcanic ash plumes to form in water: implications for the deposition of marine ash beds. Journal of Volcanology and Geothermal Research. 290:114–124. doi: 10.1016/j.jvolgeores.2014.10.015
  • Jakobsson SP. 1968. The geology and petrology of the Vestmann Islands. A preliminary report. Surtsey Research Progress Report. 4:113–129.
  • Jakobsson SP. 1979. Petrology of recent basalts of the Eastern Volcanic Zone, Iceland. Acta Naturalia Islandica. 26:1–103.
  • Johnson RW, Tuni D. 1987. Kavachi, an active forearc volcano in the western Solomon Islands: reported eruptions between 1959 and 1982. Circum-Pacific Council on Energy and Mineral Research, Earth Science Series. 7:89–112.
  • Kavanagh JL, Menand T, Sparks RSJ. 2006. An experimental investigation of sill formation and propagation in layered elastic media. Earth and Planetary Science Letters. 245(3-4):799–813. doi: 10.1016/j.epsl.2006.03.025
  • Kawachi Y, Pringle IJ. 1988. Multiple-rind structure in pillow lava as an indicator of shallow water. Bulletin of Volcanology. 50:161–168. doi: 10.1007/BF01079680
  • Klemme S. 2004. Evidence for fluoride melts in Earth's mantle formed by liquid immiscibility. Geology. 32:441–444. doi: 10.1130/G20328.1
  • Kutterolf S, Schindlbeck JC, Robertson AHF, Avery A, Baxter AT, Petronotis K, Wang KL. 2018. Tephrostratigraphy and provenance from IODP Expedition 352, Izu-Bonin arc: tracing tephra sources and volumes from the Oligocene to the Recent. Geochemistry, Geophysics, Geosystems. doi 10.1002/2017GC007100.
  • Lee DE, Scholz J, Gordon DP. 1997. Paleoecology of a late Eocene mobile rockground biota from North Otago, New Zealand. Palaios. 12:568–581. doi: 10.2307/3515412
  • Le Maitre RW. 2002. Igneous rocks: a classification and glossary of terms: recommendations of the IUGS, Subcommission on the Systematics of Igneous rocks. University Press.
  • Liu J, Scott JM, Martin CE, Pearson DG. 2015. The longevity of Archean mantle residues in the convecting upper mantle and their role in young continent formation. Earth and Planetary Science Letters. 424:109–118. doi: 10.1016/j.epsl.2015.05.027
  • Maicher D. 2003. A cluster of surtseyan volcanoes at Lookout Bluff, North Otago, New Zealand: Aspects of edifice spacing and time. Geophysical Monograph-American Geophysical Union. 140:167–178.
  • Mantell GA. 1850. Notice of the remains of the Dinornis and other birds, and of fossils and rock-specimens, recently collected by Mr. Walter Mantell in the middle Island of New Zealand; with additional notes on the Northern Island. Quarterly Journal of the Geological Society. 6:319–342. doi: 10.1144/GSL.JGS.1850.006.01-02.30
  • Marsh B. D. 1996. Solidification fronts and magmatic evolution. Mineralogical Magazine. 60(398):5–40. doi: 10.1180/minmag.1996.060.398.03
  • Mason B. 1968. Eclogitic xenoliths from volcanic breccia at Kakanui, New Zealand. Contributions to Mineralogy and Petrology. 19:316–327. doi: 10.1007/BF00389414
  • McCoy-West AJ, Bennett VC, Amelin Y. 2016. Rapid Cenozoic ingrowth of isotopic signatures simulating “HIMU” in ancient lithospheric mantle: distinguishing source from process. Geochimica et Cosmochimica Acta. 187:79–101. doi: 10.1016/j.gca.2016.05.013
  • McCoy-West AJ, Bennett VC, O’Neill HSC, Hermann J, Puchtel IS. 2015. The interplay between melting, refertilization and carbonatite metasomatism in off-cratonic lithospheric mantle under Zealandia: an integrated major, trace and platinum group element study. Journal of Petrology. 56:563–604. doi: 10.1093/petrology/egv011
  • McCoy-West AJ, Bennett VC, Puchtel IS, Walker RJ. 2013. Extreme persistence of cratonic lithosphere in the southwest Pacific: Paleoproterozoic Os isotopic signatures in Zealandia. Geology. 41:231–234. doi: 10.1130/G33626.1
  • Merrill RB, Wyllie PJ. 1975. Kaersutite and kaersutite eclogite from Kakanui, New Zealand—water-excess and water-deficient melting to 30 kilobars. Geological Society of America Bulletin. 86:555–570. doi: 10.1130/0016-7606(1975)86<555:KAKEFK>2.0.CO;2
  • Moorhouse BL. 2015. Cape Wanbrow: A stack of Surtseyan-style volcanoes built over millions of years in the Waiareka-Deborah Volcanic Field, New Zealand. PhD Thesis, 203p.
  • Moorhouse BL, White JD. 2016. Interpreting ambiguous bedforms to distinguish subaerial base surge from subaqueous density current deposits. The Depositional Record. 2:173–195. doi: 10.1002/dep2.20
  • Moorhouse BL, White JDL, Scott JM. 2015. Cape Wanbrow: A stack of Surtseyan-style volcanoes built over millions of years in the Waiareka–deborah volcanic field, New Zealand. Journal of Volcanology and Geothermal Research. 298:27–46. doi: 10.1016/j.jvolgeores.2015.03.019
  • Mortimer N. 2000. Metamorphic discontinuities in orogenic belts: example of the garnet–biotite–albite zone in the Otago Schist, New Zealand. International Journal of Earth Sciences. 89:295–306. doi: 10.1007/s005310000086
  • Nakamura Y, Coombs DS. 1973. Clinopyroxenes in the Tawhiroko tholeiitic dolerite at Moeraki, north-eastern Otago, New Zealand. Contributions to Mineralogy and Petrology. 42:213–228. doi: 10.1007/BF00371586
  • Nelson CS, Lee D, Maxwell P, Maas R, Kamp PJ, Cooke S. 2004. Strontium isotope dating of the New Zealand Oligocene. New Zealand Journal of Geology and Geophysics. 47:719–730. doi: 10.1080/00288306.2004.9515085
  • Panter KS, Blusztajn J, Hart SR, Kyle PR, Esser R, McIntosh WC. 2006. The origin of HIMU in the SW Pacific: evidence from intraplate volcanism in southern New Zealand and subantarctic islands. Journal of Petrology. 47(9):1673–1704. doi: 10.1093/petrology/egl024
  • Park J. 1918. The geology of the Oamaru district, North Otago (eastern Otago division). Government Printer, 20 p.
  • Pertermann M, Hirschmann MM. 2003. Partial melting experiments on a MORB-like pyroxenite between 2 and 3 GPa: Constraints on the presence of pyroxenite in basalt source regions from solidus location and melting rate. Journal of Geophysical Research:Solid Earth. 108(B2. doi: 10.1029/2000JB000118
  • Pilet S. 2015. Generation of low-silica alkaline lavas: Petrological constraints, models, and thermal implications. The Interdisciplinary Earth: A volume in Honor of Don L. Anderson: Geological Society of America Special Paper. 514:281–304. doi: 10.1130/2015.2514(17)
  • Pilet S, Baker MB, Stolper EM. 2008. Metasomatized lithosphere and the origin of alkaline lavas. Science. 320(5878):916–919. doi: 10.1126/science.1156563
  • Price RC, Cooper AF, Woodhead JD, Cartwright I. 2003. Phonolitic diatremes within the Dunedin volcano, South Island, New Zealand. Journal of Petrology. 44:2053–2080. doi: 10.1093/petrology/egg070
  • Reay A, Chappell D, Garden B. 2002. A new garnet-bearing mineral breccia from North Otago, New Zealand. New Zealand Journal of Geology and Geophysics. 45:461–466. doi: 10.1080/00288306.2002.9514985
  • Reay A, Johnstone RD, Kawachi Y. 1989. Kaersutite, a possible international microprobe standard. Geostandards Newsletter. 13:187–190. doi: 10.1111/j.1751-908X.1989.tb00471.x
  • Reay A, Johnstone RD, Kawachi Y. 1993. Anorthoclase, a second microprobe standard from Kakanui, New Zealand. Geostandards Newsletter. 17:135–136. doi: 10.1111/j.1751-908X.1993.tb00129.x
  • Reay A, Sipiera PP. 1987. Mantle xenoliths from the New Zealand region. In: Nixon PH, editor. Mantle xenoliths. Chichester: John Wiley and Sons; p. 347–358.
  • Reay A, Wood CP. 1974. Ilmenites from Kakanui, New Zealand. Mineralogical Magazine. 39:721–722. doi: 10.1180/minmag.1974.039.306.14
  • Scanlan EJ, Scott JM, le Roux PJ. 2020. Pyrometamorphosed Otago Schist xenoliths cause minor contamination of Dunedin Volcanic Group basanite. NZJGG.
  • Scanlan EJ, Scott JM, Wilson VJ, Stirling CH, Reid MR, le Roux PJ. 2018. In-situ 87Sr/86Sr of scheelite and calcite reveals proximal and distal fluid-rock interaction during orogenic W-Au mineralization, Otago Schist, New Zealand. Economic Geology. 113:1571–1586. doi: 10.5382/econgeo.2018.4603
  • Scott JM. 2020. An updated catalogue of New Zealand’s mantle peridotite and serpentinite. New Zealand Journal of Geology and Geophysics.
  • Scott JM, Brenna M, Crase JA, Waight TE, van der Meer QHA, Cooper AF, Palin JM, le Roux PJ, Münker C. 2016. Peridotitic lithosphere metasomatised by volatile-bearing melts, and its association with intraplate alkaline HIMU-like magmatism. Journal of Petrology. 57:2053–2078. doi: 10.1093/petrology/egw069
  • Scott JM, Hodgkinson A, Palin JM, Waight TE, van der Meer QHA, Cooper AF. 2014a. Ancient melt depletion overprinted by young carbonatitic metasomatism in the New Zealand lithospheric mantle. Contributions to Mineralogy and Petrology. 167(963):20.
  • Scott JM, Liu J, Pearson DG, Harris GA, Czertowicz TA, Woodland SJ, Riches AJV, Luth RW. 2019. Continent stabilisation by lateral accretion of subduction zone-processed depleted mantle residues; insights from Zealandia. Earth and Planetary Science Letters. 507:175–186. doi: 10.1016/j.epsl.2018.11.039
  • Scott JM, Pontesilli A, Brenna M, White JDL, Giacalone E, Palin JM, le Roux PJ. 2020. The Dunedin Volcanic Group and a revised model for Zealandia's alkaline intraplate volcanism. New Zealand Journal of Geology and Geophysics.
  • Scott JM, Waight TE, van der Meer QHA, Palin JM, Cooper AF, Münker C. 2014b. Metasomatized ancient lithospheric mantle beneath the young Zealandia microcontinent and its role in HIMU-like intraplate magmatism. Geochemistry, Geophysics, Geosystems. 15:3477–3501. doi: 10.1002/2014GC005300
  • Sprung P, Schuth S, Münker C, Hoke L. 2007. Intraplate volcanism in New Zealand: the role of fossil plume material and variable lithospheric properties. Contributions to Mineralogy and Petrology. 153:669–687. doi: 10.1007/s00410-006-0169-1
  • Sun J. 2018. Origin of garnet pyroxenites in the Kakanui Mineral Breccia. MSc Thesis, University of Otago. 73p.
  • Sun S-S, McDonough WF. 1989. Chemical and isotopic systematics of oceanic basalts: implications for mantle composition and processes. Geological Society, London, Special Publications. 42:313–345. doi: 10.1144/GSL.SP.1989.042.01.19
  • Thompson NK, Bassett KN, Reid CM. 2014. The effect of volcanism on cool-water carbonate facies during maximum inundation of Zealandia in the Waitaki–Oamaru region. New Zealand Journal of Geology and Geophysics. 57:149–169. doi: 10.1080/00288306.2014.904385
  • Thomson JA. 1907. III.—Inclusions in some volcanic rocks. Geological Magazine. 4:490–500. doi: 10.1017/S0016756800133941
  • Thomson K, Schofield N. 2008. Lithological and structural controls on the emplacement and morphology of sills in sedimentary basins. Geological Society, London, Special Publications. 302(1):31–44. doi: 10.1144/SP302.3
  • Thorarinsson S. 1967. Surtsey. The New Island in the North Atlantic. New York: The Viking Press. 47 p.
  • Timm C, Hoernle K, Werner R, Hauff F, den Bogaard PV, White J, Mortimer N, Garbe-Schönberg D. 2010. Temporal and geochemical evolution of the Cenozoic intraplate volcanism of Zealandia. Earth Science Reviews. 98:38–64. doi: 10.1016/j.earscirev.2009.10.002
  • Urosevic M, Nebel O, Padrón-Navarta JA, Rubatto D. 2018. Assessment of O and Fe isotope heterogeneity in garnet from Kakanui (New Zealand) and Erongo (Namibia). European Journal of Mineralogy. 30:695–710. doi: 10.1127/ejm/2018/0030-2755
  • Uttley GH. 1918. The volcanic rocks of Oamaru, with special reference to their position in the stratigraphic series. Transactions of the Royal Society of New Zealand. 50:106–117.
  • Uttley GH. 1920. Remarks on Bulletin No. 20 (New Series) of the New Zealand Geological Survey. Transactions of the Royal Society of New Zealand. 52:169–182.
  • Verolino A, White JDL, Brenna M. 2018. Eruption dynamics at Pahvant Butte volcano, Utah, western USA: insights from ash-sheet dispersal, grain size, and geochemical data. Bulletin of Volcanology. 80(11):81. DOI:10.1007/s00445-018-1256-7.
  • Walker GP. 1992. Morphometric study of pillow-size spectrum among pillow lavas. Bulletin of Volcanology. 54(6):459–474. doi: 10.1007/BF00301392
  • Wallace RC. 1977. Anorthoclase-calcite rodding within a kaersutite xenocryst from the Kakanui mineral breccia, New Zealand. American Mineralogist. 62:1038–1041.
  • White AJR, Chappell BW, Jakeš P. 1972. Coexisting clinopyroxene, garnet and amphibole from an “eclogite”, Kakanui, New Zealand. Contributions to Mineralogy and Petrology. 34:185–191. doi: 10.1007/BF00373290
  • White JDL, Houghton BF. 2006. Primary volcaniclastic rocks. Geology. 34(8):677–680. doi: 10.1130/G22346.1
  • Wilding A, Sweetman LAD. 1971. Endeavour-1 Well Completion Report. Prepared by BP Shell Aquitaine Todd Petroleum Developments for the Ministry of Economic Development, Crown Minerals, Petroleum Report Series No. PR303, Wellington.
  • Zack T, Foley SF, Jenner GA. 1997. A consistent partition coefficient set for clinopyroxene, amphibole and garnet from laser ablation microprobe analysis of garnet pyroxenites from Kakanui, New Zealand. Neues Jahrbuch fur Mineralogie, Abhandlungen. 172:23–41.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.