• Keine Ergebnisse gefunden

Hikurangi Plateau subduction a trigger for Vitiaz arc splitting and Havre Trough opening (southwestern Pacific)

N/A
N/A
Protected

Academic year: 2022

Aktie "Hikurangi Plateau subduction a trigger for Vitiaz arc splitting and Havre Trough opening (southwestern Pacific)"

Copied!
5
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Manuscript received 15 May 2020 Revised manuscript received 12 November 2020 Manuscript accepted 15 November 2020

https://doi.org/10.1130/G48436.1

© 2021 The Authors. Gold Open Access: This paper is published under the terms of the CC-BY license.

CITATION: Hoernle, K., et al., 2021, Hikurangi Plateau subduction a trigger for Vitiaz arc splitting and Havre Trough opening (southwestern Pacific): Geology, v. 49, p. 536–540, https://doi.org/10.1130/G48436.1

Hikurangi Plateau subduction a trigger for Vitiaz arc splitting and Havre Trough opening (southwestern Pacific)

K. Hoernle1,2, J. Gill3, C. Timm1,4, F. Hauff1, R. Werner1, D. Garbe-Schönberg2 and M. Gutjahr1

1 GEOMAR Helmholtz Centre for Ocean Research Kiel, 24148 Kiel, Germany

2 Institute of Geosciences, Kiel University, 24118 Kiel, Germany

3 Department of Earth and Planetary Sciences, University of California, Santa Cruz, California 95064, USA

4 GNS Science, PO Box 30-368, Lower Hutt 5040, New Zealand

ABSTRACT

Splitting of the Vitiaz arc formed the Tonga-Kermadec and Lau-Colville Ridges (south- western Pacific Ocean), separated by the Lau Basin in the north and Havre Trough in the south. We present new trace element and Sr-Nd-Hf-Pb isotope geochemistry for the Kermadec and Colville Ridges extending ∼900 km north of New Zealand (36°S–28°S) in order to (1) compare the composition of the arc remnants with Quaternary Kermadec arc volcanism, (2) constrain spatial geochemical variations in the arc remnants, (3) evaluate the effect of Hikurangi igneous plateau subduction on the geochemistry of the older arc lavas, and (4) elucidate what may have caused arc splitting. Compared to the Kermadec Ridge, the Colville Ridge has higher more-incompatible to less-incompatible immobile element ratios and largely overlapping isotope ratios, consistent with an origin through lower degrees of melting of more enriched upper mantle in the Vitiaz rear arc. Between ca. 8 and 3 Ma, both halves of the arc (∼36°S–29°S) included a more enriched (EM1-type) composition (with lower 206Pb/204Pb and

207Pb/204Pb and higher Δ8/4 Pb [deviation of the measured 208Pb/204Pb ratio from a Northern Hemisphere basalt regression line] and 87Sr/86Sr) compared to older and younger arc lavas.

High-Ti basalts from the Manihiki Plateau, once joined to the Hikurangi Plateau, could serve as the enriched Vitiaz arc end member. We propose that the enriched plateau signature, seen only in the isotope ratios of mobile elements, was transported by hydrous fluids from the western margin of the subducting Hikurangi Plateau or a Hikurangi Plateau fragment into the overlying mantle wedge. Our results are consistent with plateau subduction triggering arc splitting and backarc opening.

INTRODUCTION

Volcanic arcs play a key role in the plate tectonic paradigm, being the surface expres- sion of plate convergence. Nevertheless, little is known about the long-term tectonic and geo- chemical evolution of submarine remnant arc systems formed by arc splitting and backarc basin opening, largely due to their inaccessi- bility. Contemporaneous Neogene volcanism on the Tonga-Kermadec and Lau-Colville Ridges (southwestern Pacific Ocean; Fig. 1) supports the idea that the subparallel ridges were once part of a single volcanic arc (termed the Vitiaz arc), which split at ca. 5.5–3 Ma to form the Lau Basin and Havre Trough (e.g., Gill, 1976; Dun-

can et al., 1985; Wright et al., 1996; Timm et al., 2019; Caratori Tontini et al., 2019). Mechanisms for triggering arc splitting, however, are con- troversial (Sdrolias and Müller, 2006; Wallace et al., 2009).

Subduction of young igneous oceanic pla- teaus is unlikely due to their buoyancy, as dem- onstrated by the Hikurangi Plateau when it col- lided with and accreted to the Chatham Rise at ca. 105 Ma, becoming part of the Zealandia microcontinent. The present-day Hikurangi Pla- teau represents a rare example of an oceanic plateau being subducted into Earth’s mantle beneath the North Island of New Zealand and the southern Kermadec arc (Fig. 1) (Reyners

et al., 2011). It formed at ca. 125 Ma as part of the Ontong Java–Manihiki–Hikurangi super- plateau, which broke apart shortly after forma- tion (e.g., Davy et al., 2008; Hoernle et al., 2010). The basement of the plateau fragments consists of two distinct geochemical types:

(1) low-Ti basalts (Kroenke and Kwaimbaita groups on Ontong Java) have isotopically inter- mediate compositions similar to that of primi- tive mantle, and (2) high-Ti basalts (Singallo group on Ontong Java) have enriched mantle 1 (EM1)–type basement with unradiogenic

206Pb/204Pb but radiogenic Sr isotope ratios (e.g., Tejada et al., 2004; Hoernle et al., 2010;

Timm et al., 2011; Golowin et al., 2018). Where stratigraphic information is available, the high- Ti basalts overlie the low-Ti basalts. Between ca. 117 and 79 Ma, spreading along the Osbourn Trough paleo–spreading center, now located at

∼25.5°S latitude, created ∼3000 km of seafloor between the Hikurangi and Manihiki Plateaus (e.g., Mortimer et al., 2019). The northern tip of the subducting Hikurangi Plateau is presently located at ∼36°S.

Here we present new trace element and Sr- Nd-Hf-Pb isotopic data from 40 locations on the Kermadec (KR) and Colville (CR) Ridges between ∼28°S and 36°S (Fig. S1 in the Supple- mental Material1), recovered primarily on the R/V Sonne cruise SO255. We show that geo- chemical variations along both ridges are nearly identical, confirming that they once formed a single arc, and that differences between the ridges reflect the KR having been the frontal arc and the CR the rear arc of the Neogene Vitiaz arc. Some of the late Neogene (8–3 Ma) Vitiaz arc had an enriched composition distinct from

1Supplemental Material. Supplemental information about the samples and analytical methods, including Fig. S1), Table S1 (geochemical data), and Table S2 (replicates and reference materials). Please visit https://doi .org/10.1130/GEOL.S.13377182 to access the supplemental material, and contact editing@geosociety.

org with any questions.

Published online 30 December 2020

(2)

that of the Quaternary Kermadec volcanic arc, consistent with subduction of the Hikurangi Pla- teau or a plateau fragment. The enriched lavas occur along a segment of the arc where part of the forearc is missing, consistent with removal by plateau subduction. Plateau collision and sub- duction is a possible mechanism for causing arc splitting and backarc basin opening.

RESULTS

We report analytical methods in the Supple- mental Material, trace element and isotope data in Table S1, and replicate materials and replicates in Table S2, for lavas of the KR and CR (together designated “KCR lavas”). Compared to the pri- marily tholeiitic KR samples, the primarily calc- alkaline CR samples (no adakites or boninites were found) have higher more-incompatible to less-incompatible immobile element ratios (Ta/

Yb, Nb/Yb, Th/Yb, La/Yb, and La/Sm). KCR lavas also have low Nb/Th (<3.4), low Ce/Pb (≤12, with two exceptions), and low Nb/U (<10 when loss on ignition [LOI] ≤4 wt%). Although the KR and CR lavas largely overlap in isotopic composition (Fig. 2), each can be divided into isotopically depleted and enriched groups. The depleted KCR lavas between ∼28°S and 37°S largely overlap with the Quaternary Kermadec arc and Havre Trough backarc lavas isotopi- cally (Fig. 2). Compared to the depleted KCR lavas, the enriched KCR lavas extend to higher K2O and incompatible element contents for a given MgO content, and to lower 206Pb/204Pb

(≤18.37) and 207Pb/204Pb (≤15.53) and higher

87Sr/86Sr (≥0.7047) at similar 208Pb/204Pb, Nd, and Hf isotope ratios, indicating an enriched mantle (EM1)–type component in the source of the KCR lavas (between ∼29.5°S and 36.5°S), not yet found in the Quaternary lavas.

Alkalic seamounts behind and late-stage cones on the CR (designated “intraplate CR”) have higher Nb/Th (4.3, 9.5–15.5), Ce/Pb (3–32), Nb/U (9–50, LOI <4 wt%) and 206Pb/204Pb, and lower 87Sr/86Sr, suggesting a similar intraplate source as for seamounts in the South Fiji Basin (Mortimer et al., 2007; Todd et al., 2011).

DISCUSSION AND CONCLUSIONS Origin of Enriched End Member

Published age data can be used to con- strain the duration of the enriched volca- nism (Fig. S1). Six 40Ar-39Ar ages published from enriched lavas range from 7.5 ± 2.0 to 2.63 ± 0.23 Ma (three from the KR, 4.4–

4.8 Ma; three from the CR, 7.5–2.6 Ma; Timm et al., 2019). Depleted lavas yielded 40Ar-39Ar ages of 3.40 ± 0.24 Ma from the KR (Timm et al., 2019) and 16.7 ± 0.1 Ma from the CR (Mortimer et al., 2010). Samples classified as depleted based on K content (0.17–0.29 wt%) from the KR have K-Ar ages of 1.25–2.04 Ma (n = 8, ∼32.3°S) and 7.84 ± 0.69 Ma (31.6°S) (Balance et al., 1999). In summary, volcanism on the KR (1.2–7.8 Ma) and CR (2.6–16.7 Ma) extends into the early Miocene, with enriched volcanism being restricted conservatively to

ca. 8–3 Ma and depleted volcanism perhaps continuously since the early Miocene.

Plots of isotope ratios versus latitude (with 1° latitude added to CR samples to compen- sate for northwest-southeast opening of the Havre Trough) show that isotopic variations are nearly identical along the KR and CR (Fig. 3), confirming that they once formed a single vol- canic arc (Gill, 1976; Timm et al., 2019; Cara- tori Tontini et al., 2019). The shift to higher more-incompatible to less-incompatible ele- ment ratios in the CR than the KR lavas at simi- lar isotopic composition suggests derivation of CR lavas through lower degrees of melting beneath the rear arc. KCR lavas with enriched compositions were found between 29°S and 37°S (corrected CR; Fig. 3) with the strongest enriched signal being located at ∼33°S, charac- terized by the lowest 206Pb/204Pb and 207Pb/204Pb and highest 87Sr/86Sr ratios. Therefore, the enriched arc signature appears to have been limited both temporally and spatially, although more geochronology is necessary to define its exact duration.

We now review possible origins of the enriched end member, beginning with the mantle wedge. Assuming corner flow, enriched intraplate mantle as found in South Fiji Basin seamounts and intraplate CR samples could have flowed from the backarc beneath the older arc. The South Fiji seamount and intra- plate CR source, however, has higher 206Pb/204Pb and lower 87Sr/86Sr isotope ratios than the KCR lavas and therefore cannot explain the enriched (EM1-type) signature (Fig. 2). There is also no evidence of a plume beneath the arc, because the enriched lavas show typical subduction zone incompatible element abundances, e.g., low Ce/Pb (2.0–11.2, n = 74) and Nb/U (0.7–7.6, n = 74) ratios, well below typical mantle values of 25 ± 5 and 47 ± 10, respectively, in global intraplate lavas (Hofmann et al., 1986). There is also no geophysical (Bassett et al., 2016) or geochemical evidence that continental litho- sphere underlies the KR and CR.

The enrichment of fluid-mobile elements relative to fluid-immobile elements is con- sistent with the enriched component coming from the subduction input. A decrease in Ce/

Pb from 25 (average for global dry oceanic mantle, after Hofmann et al. [1986]) to 6 (aver- age in enriched KCR lavas) represents a four- fold addition of fluid-mobile Pb compared to fluid-immobile Ce, implying enrichment of the mantle wedge primarily by hydrous fluids rather than melts. Enrichment by fluid addi- tion can also explain why the isotope ratios of fluid-mobile elements (Pb and Sr) have been affected more than those of fluid-immobile Nd and Hf (Turner and Hawkesworth, 1998).

There is, however, no evidence of an EM1-type hotspot track or seamounts being subducted along 900 km of this part of the arc that could Figure 1. (A) Map showing

location of the Tonga-Ker- madec arc-backarc system and Hikurangi Plateau.

Red box shows location of the map in Figure S1 (see footnote 1). Base map is from GEBCO_2014 Grid (version 20150318; http://

www.gebco.net).

(3)

explain the enriched composition (Castillo et al., 2009; Hoernle et al., 2010).

The Hikurangi Plateau currently subducts south of 36°S, but isotopic evidence suggests that it may underlie the present arc as far north as ∼32°S (Timm et al., 2014). The enriched KCR isotopic signature, however, cannot be explained by the composition of the low-Ti Hikurangi basement, sampled along ∼50 km of the Rapuhia scarp (Hoernle et al., 2010).

Detailed sampling and geochemical studies have also been conducted on the Manihiki Plateau, which was once joined to the north- ern part of the Hikurangi Plateau. The high-Ti lavas from the Manihiki North Plateau have appropriate Pb and Sr isotopic compositions (Timm et al., 2011; Golowin et al., 2018) to serve as the EM1 component in the enriched KCR lavas (Fig. 2). Although estimates of the maximum size of the Hikurangi Plateau based on super-plateau reconstructions extend the plateau to ∼32°S (Timm et al., 2014), it is possible that a relatively thin finger of plateau material or plateau fragment, separated from the western edge of the Manihiki Plateau, may

have extended as far north as 29°S. Alterna- tively, EM1 plateau material may have also been incorporated in the ocean lithosphere (crust and/or mantle) formed directly after plateau breakup at ca. 117 Ma (Fig. 4). We

propose that between ca. 8 and 3 Ma, high-Ti plateau basalts with an EM1-type composition, similar to Manihiki North Plateau lavas, sub- ducted beneath the southern Vitiaz arc possibly as far north as ∼29°S.

A

C

B

D

Figure 2. 206Pb/204Pb versus 208Pb/204Pb (A), 207Pb/204Pb (B), 87Sr/86Sr (C), and 143Nd/144Nd (D) isotope diagrams for depleted and enriched (EM1- type; ca. 8–3 Ma) Kermadec Ridge and Colville Ridge lavas, Quaternary Kermadec volcanic arc (QKVA) lavas, Havre Trough backarc (HTBA) lavas, sediments data, and Manihiki North Plateau samples (Golowin et al., 2018; Timm et al., 2019, and references therein; Hauff et al., 2021;

Gill et al., 2021). In C, the arrow labelled “Subducted sediments” points to the subducted sediment field, which plots above the range in the diagram. Pacific and Indian mid-oceanic ridge basalt (MORB) is from the PetDB (http://www.earthchem.org/petdb).

Figure 3. Plot of

206Pb/204Pb versus latitude, showing a peak in enrich- ment (lowest 206Pb/204Pb) at 33°S. One degree of latitude has been added to Colville Ridge samples to correct for the opening of the Havre Trough. Refer- ences are as for Figure 2.

(4)

Causes of Arc Splitting and Backarc Basin Opening

The causes of splitting of the Vitiaz arc into the KR and CR and the formation of the Havre Trough backarc basin are enigmatic.

Extension in the overriding plate can trig- ger backarc rifting and/or spreading (Karig, 1971). Rollback of the subducting slab causes trench retreat, resulting in extension in the overriding plate (Jurdy, 1979), but rollback is unlikely to be a major process during sub- duction of thickened plateau crust (e.g., as much as 35 km beneath the Hikurangi Plateau;

Reyners et al., 2011).

Another mechanism for generating exten- sion in the overriding plate is forearc rotation caused by collision of a buoyant indentor on

the incoming plate (e.g., seamount province, hotspot track, or oceanic plateau) with the subduction margin. The collision “pins” the subduction zone, resulting in trenchward rota- tion of the forearc about the pinned pivot point, causing extension, arc splitting, and backarc rifting and/or spreading (McCabe, 1984). An important criterion for causing backarc opening by a collision is that the “indentor enter[s] the subduction margin just prior to the initiation of the back-arc rifting event” (Wallace et al., 2009, p. 9). Based on the presently available age and geochemical data, the first evidence for plateau subduction was at ca. 8 Ma, pre- ceding the initial Havre Trough opening at ca.

5.5–3 Ma (Caratori Tontini et al., 2019), con- sistent with a connection between (1) plateau

subduction, and (2) Vitiaz arc splitting and Havre Trough opening.

Because there has been only minor clockwise rotation of the Kermadec forearc from 10 Ma to the present (Sdrolias and Müller, 2006), another possible mechanism is removal of forearc litho- sphere (Wolf and Huismans, 2019). Subduc- tion of a positive bathymetric anomaly on the downgoing plate, such as an aseismic ridge or plateau fragment, would enhance basal litho- spheric erosion of the overriding plate, result- ing in subsidence and extension of the margin (Clift et al., 2003). Between 29°S and 34°S, the lithosphere thickness beneath the forearc thins to ∼10–11 km but reaches thicknesses of 16–17 km to the north (25°S–26°S) (Stratford et al., 2015; Bassett et al., 2016), which may reflect basal forearc removal by plateau sub- duction. In order to explain the absence of the

∼110-km-wide Tonga Ridge (located ∼160 km from the trench at ∼25°S–26°S) in the forearc south of ∼30.5°S, Collot and Davy (1998) pro- posed frontal forearc removal. Removal of a large block of the forearc could have resulted in extension as the overriding plate moved trench- ward to compensate for forearc removal, and could explain migration of the arc westward away from the trench (Keppie et  al., 2009) beginning at ∼30°S and away from the KR into the eastern Havre Trough south of ∼32°S (Bas- sett et al., 2016). Thus, lithospheric removal by plateau collision and subduction could also be an important mechanism contributing to arc splitting and backarc basin opening.

Finally, the enriched plateau signal disap- pears in the KR and CR volcanism at ca. 3 Ma, and normal oceanic crust is presently subduct- ing in the region where the enriched signal was observed (29°S–36°S). Once the plateau frag- ment fully subducted and thinner “normal” Cre- taceous oceanic crust started subducting again, some slab rollback would have been likely. This could also have been an important mechanism causing or contributing to backarc rifting and/

or spreading. We need better constraints on the timing of Havre Trough opening and plateau subduction in order to constrain the exact mech- anism further.

As is clear from Ontong Java, plateau col- lision with a subduction zone will not always result in arc splitting and backarc rifting. Due to its larger size, greater thickness, and younger age at collision than the Hikurangi plateau mar- gin or plateau fragment when it collided with New Zealand, the Ontong Java Plateau was too buoyant and thick to subduct and clogged the subduction zone, resulting in subduction polar- ity reversal, so that normal ocean crust could again subduct. In summary, there appears to be a continuum from collision and subduc- tion of seamount clusters and hotspot tracks resulting in forearc rotation and backarc rift- ing (Wallace et al., 2009), to collision of older

A C

B D

E

Enriched oceanic lithosphere ?

Figure 4. Model to explain the presence of enriched signal in Kermadec Ridge and Colville Ridge lavas between ca. 8 and 3 Ma. (A) At ca. 120 Ma, Ontong Java rifts away from the Mani- hiki + Hikurangi plateau fragment. (B) At ca. 117–97 Ma, spreading along the Osbourn spreading center creates ∼3000 km of seafloor between Manihiki and Hikurangi (Mortimer et al., 2019).

Some oceanic lithosphere formed near rifted margins of plateau fragments may also have enriched plateau-like composition. At ca. 105 Ma, Hikurangi collides with the Gondwana sub- duction margin, which later becomes the Chatham Rise of Zealandia. (C) At ca. 10 Ma, the western margin of Hikurangi is just outboard of the Kermadec–North Island (New Zealand) trench. (D) From ca. 8 to 3 Ma, the western Hikurangi margin subducts beneath the Vitiaz arc, which splits into Colville Ridge (CR) and Kermadec Ridge (KR), forming the Havre Trough. (E) Present configuration. (Modified from Davy et al., 2008; Timm et al., 2014.)

(5)

plateau fragments causing lithospheric removal ( accompanied by arc splitting and backarc open- ing), to collision of large and thick plateaus over a long stretch of an arc that shuts down subduc- tion, causing polarity reversal in oceanic sub- duction zones.

ACKNOWLEDGMENTS

We thank the shipboard and scientific crews of R/V Sonne for a successful SO255 cruise; S. Hauff, K.

Junge, and U. Westernströer for analytical support;

O. Ishizuka, B. Jicha, and B. Stern for helpful formal reviews; I. Grevemeyer for helpful comments; and the German Federal Ministry of Education and Research (BMBF) (grant 03G0255A), GEOMAR Helmholtz Centre (Kiel, Germany), and GNS Science (New Zea- land) for funding this project.

REFERENCES CITED

Ballance, P.F., Ablaev, A.G., Pushchin, I.K., Pletnev, S.P., Birylina, M.G., Itaya, T., Follas, H.A., and Gibson, G.W., 1999, Morphology and history of the Kermadec trench–arc–backarc basin–

remnant arc system at 30 to 32°S: Geophysical profile, microfossil and K-Ar data: Marine Geol- ogy, v. 159, p. 35–62, https://doi .org/10.1016/

S0025-3227(98)00206-0.

Bassett, D., Kopp, H., Sutherland, R., Henrys, S., Watts, A.B., Timm, C., Scherwath, M., Greve- meyer, I., and de Ronde, C.E.J., 2016, Crustal structure of the Kermadec arc from MANGO seismic refraction profiles: Journal of Geophysi- cal Research: Solid Earth, v. 121, p. 7514–7546, https://doi .org/10.1002/2016JB013194.

Caratori Tontini, F., Bassett, D., de Ronde, C.E.J., Timm, C., and Wysoczanski, R., 2019, Early evolution of a young back-arc basin in the Havre Trough: Nature Geoscience, v. 12, p. 856–862, https://doi .org/10.1038/s41561-019-0439-y.

Castillo, P.R., Lonsdale, P.F., Moran, C.L., and Hawkins, J.W., 2009, Geochemistry of mid- Cretaceous Pacific crust being subducted along the Tonga-Kermadec Trench: Implications for the generation of arc lavas: Lithos, v. 112, p. 87–102, https://doi .org/10.1016/ j.lithos.2009.03.041.

Clift, P.D., Pecher, I., Kukowski, N., and Hampel, A., 2003, Tectonic erosion of the Peruvian forearc, Lima Basin, by subduction and Nazca ridge collision: Tectonics, v. 22, 1023, https://doi .org/10.1029/2002TC001386.

Collot, J.Y., and Davy, B., 1998, Forearc structures and tectonic regimes at the oblique subduction zone between the Hikurangi Plateau and the southern Kermadec margin: Journal of Geo- physical Research, v. 103, p. 623–650, https://

doi .org/10.1029/97JB02474.

Davy, B., Hoernle, K., and Werner, R., 2008, Hikurangi Plateau: Crustal structure, rifted formation, and Gondwana subduction history: Geochemistry Geophysics Geosystems, v. 9, Q07004, https://

doi .org/10.1029/2007GC001855.

Duncan, R.A., Vallier, T.L., and Falvey, D.A., 1985, Volcanic episodes at Eua, Tonga Islands, in Scholl, D.W., and Vallier, T.L., eds., Geology and Offshore Resources of Pacific Island Arcs—

Tonga Region: Houston, Texas, Circum-Pacific Council for Energy and Mineral Resources, Earth Science Series, v. 2, p. 281–290.

Gill, J.B., 1976, Composition and age of Lau Basin and Ridge volcanic rocks: Implica- tions for evolution of an interarc basin and remnant arc: Geological Society of Amer- ica Bulletin, v. 87, p. 1384–1395, https://doi .org/10.1130/0016-7606(1976)87<1384:CAA OLB>2.0.CO;2.

Gill, J., Hoernle, K., Todd, E., Hauff, F., Werner, R., Timm, C., Garbe-Schönberg, D., Gutjahr, M., 2021, Basalt geochemistry and mantle flow dur- ing early backarc basin evolution: Havre Trough and Kermadec Arc, southwest Pacific: Geo- chemistry Geophysics Geosystems, https://doi.

org/10.1029/2020GC009339 (in press).

Golowin, R., Portnyagin, M., Hoernle, K., Hauff, F., Werner, R., and Garbe-Schönberg, D., 2018, Geochemistry of deep Manihiki Plateau crust: Implications for compositional diver- sity of large igneous provinces in the Western Pacific and their genetic link: Chemical Geol- ogy, v. 493, p. 553–566, https://doi .org/10.1016/

j.chemgeo.2018.07.016.

Hauff, F., Hoernle, K., Gill, J., Werner, R., Timm, C., Garbe-Schönberg, D., Gutjahr, M., and Jung, S., 2021, R/V SONNE Cruise SO255 “VITIAZ”:

An integrated major element, trace element and Sr-Nd-Pb-Hf isotope data set of volcanic rocks from the Colville and Kermadec Ridges, the Qua- ternary Kermadec volcanic front and the Havre Trough backarc basin (version 1.0): Interdisci- plinary Earth Data Alliance (IEDA), https://doi .org/10.26022/IEDA/111723 (accessed October 2020).

Hoernle, K., Hauff, F., van den Bogaard, P., Werner, R., Mortimer, N., Geldmacher, J., Garbe-Schön- berg, D., and Davy, B., 2010, Age and geochem- istry of volcanic rocks from the Hikurangi and Manihiki oceanic plateaus: Geochimica et Cos- mochimica Acta, v. 74, p. 7196–7219, https://doi .org/10.1016/ j.gca.2010.09.030.

Hofmann, A.W., Jochum, K.P., Seufert, M., and White, W.M., 1986, Nb and Pb in oceanic basalts: New constraints on mantle evolution: Earth and Plan- etary Science Letters, v. 79, p. 33–45, https://doi .org/10.1016/0012-821X(86)90038-5.

Jurdy, D.M., 1979, Relative plate motions and the formation of marginal basin: Journal of Geo- physical Research, v. 84, p. 6796–6802, https://

doi .org/10.1029/JB084iB12p06796.

Karig, D.E., 1971, Origin and development of mar- ginal basins in the western Pacific: Journal of Geophysical Research, v. 76, p. 2542–2561, https://doi .org/10.1029/JB076i011p02542.

Keppie, D.F., Currie, C.A., and Warren, C., 2009, Subduction erosion modes: Comparing finite element numerical models with the geologi- cal record: Earth and Planetary Science Let- ters, v. 287, p. 241–254, https://doi .org/10.1016/

j.epsl.2009.08.009.

McCabe, R., 1984, Implications of paleomagnetic data on the collision related bending of island arcs: Tectonics, v. 3, p. 409–428, https://doi .org/10.1029/TC003i004p00409.

Mortimer, N., Herzer, R.H., Gans, P.B., Laporte- Magoni, C., Calvert, A.T., and Bosch, D., 2007, Oligocene–Miocene tectonic evolution of the South Fiji Basin and Northland Plateau, SW Pacific Ocean: Evidence from petrology and dating of dredged rocks: Marine Geol- ogy, v. 237, p. 1–24, https://doi .org/10.1016/

j.margeo.2006.10.033.

Mortimer, N., Gans, P.B., Palin, J.M., Mef- fre, S., Herzer, R.H., and Skinner, D.N.B., 2010, Location and migration of Miocene–

Quaternary volcanic arcs in the SW Pacific region: Journal of Volcanology and Geother- mal Research, v. 190, p. 1–10, https://doi .org/10.1016/ j.jvolgeores.2009.02.017.

Mortimer, N., van den Bogaard, P., Hoernle, K., Timm, C., Gans, P.B., Werner, R., and Riefstahl, F., 2019, Late Cretaceous oceanic plate reorganization and the breakup of Zealandia and Gondwana:

Gondwana Research, v. 65, p. 31–42, https://

doi .org/10.1016/ j.gr.2018.07.010.

Reyners, M., Eberhart-Phillips, D., and Bannister, S., 2011, Tracking repeated subduction of the Hikurangi Plateau beneath New Zealand: Earth and Planetary Science Letters, v. 311, p. 165–

171, https://doi .org/10.1016/ j.epsl.2011.09.011.

Sdrolias, M., and Müller, R.D., 2006, Controls on back-arc basin formation: Geochemistry Geo- physics Geosystems, v. 7, Q04016, https://doi .org/10.1029/2005GC001090.

Stratford, W., Peirce, C., Funnell, M., Paulatto, M., Watts, A.B., Grevemeyer, I., and Bassett, D., 2015, Effect of Seamount subduction on forearc morphology and seismic structure of the Tonga- Kermadec subduction zone: Geophysical Journal International, v. 200, p. 1503–1522, https://doi .org/10.1093/gji/ggu475.

Tejada, M.L.G., Mahoney, J.J., Castillo, P.R., Ingle, S.P., Sheth, H.C., and Weis, D., 2004, Pin-prick- ing the elephant: Evidence on the origin of the Ontong Java Plateau from Pb-Sr-Hf-Nd isotopic characteristics of ODP Leg 182 basalts, in Fit- ton, J.G., et al., eds., Origin and Evolution of the Ontong Java Plateau: Geological Society [Lon- don] Special Publication 229, p. 133–150, https://

doi .org/10.1144/GSL.SP.2004.229.01.09.

Timm, C., Hoernle, K., Werner, R., Hauff, F., van den Bogaard, P., Michael, P., and Coffin, M.F., and Koppers, A., 2011, Age and geochemistry of the oceanic Manihiki Plateau, SW Pacific: New evi- dence for a plume origin: Earth and Planetary Science Letters, v. 304, p. 135–146, https://doi .org/10.1016/ j.epsl.2011.01.025.

Timm, C., et al., 2014, Subduction of the oceanic Hikurangi Plateau and its impact on the Ker- madec arc: Nature Communications, v. 5, 4923, https://doi .org/10.1038/ncomms5923.

Timm, C., de Ronde, C.E.J., Hoernle, K., Cousens, B., Wartho, J.-A., Caratori Tontini, F., Wysoc- zanski, R., Hauff, F., and Handler, M., 2019, New age and geochemical data from the South- ern Colville and Kermadec Ridges, SW Pacific:

Insights into the recent geological history and petrogenesis of the Proto-Kermadec (Vitiaz) Arc:

Gondwana Research, v. 72, p. 169–193, https://

doi .org/10.1016/ j.gr.2019.02.008.

Todd, E., Gill, J.B., Wysoczanski, R.J., Hergt, J.M., Wright, I.C., Leybourne, M.I., and Mortimer, N., 2011, Hf isotopic evidence for small-scale hetero- geneity in the mode of mantle wedge enrichment:

Southern Havre Trough and South Fiji Basin back arcs: Geochemistry Geophysics Geosystems, v. 12, Q09011, https://doi .org/10.1029/2011GC003683.

Turner, S.P., and Hawkesworth, C.J., 1998, Using geo- chemistry to map mantle flow beneath the Lau Basin: Geology, v. 26, p. 1019–1022, https://doi .org/10.1130/0091-7613(1998)026<1019:UGT MMF>2.3.CO;2.

Wallace, L.M., Ellis, S., and Mann, P., 2009, Colli- sional model for rapid fore-arc block rotations, arc curvature, and episodic back-arc rifting in subduction settings: Geochemistry Geophys- ics Geosystems, v. 10, Q05001, https://doi .org/10.1029/2008GC002220.

Wolf, S.G., and Huismans, R.S., 2019, Mountain build- ing or backarc extension in ocean-continent sub- duction systems: A function of backarc lithospheric strength and absolute plate velocities: Journal of Geophysical Research: Solid Earth, v. 124, p. 7461–

7482, https://doi .org/10.1029/2018JB017171.

Wright, I.C., Parson, L.M., and Gamble, J.A., 1996, Evolution and interaction of migrating cross-arc volcanism and back-arc rifting: An example from the southern Havre Trough (35°20′–37°S): Jour- nal of Geophysical Research, v. 101, p. 22,071–

22,086, https://doi .org/10.1029/96JB01761.

Printed in USA

Abbildung

Figure 2.  206 Pb/ 204 Pb versus  208 Pb/ 204 Pb (A),  207 Pb/ 204 Pb (B),  87 Sr/ 86 Sr (C), and  143 Nd/ 144 Nd (D) isotope diagrams for depleted and enriched (EM1- (EM1-type; ca
Figure 4.  Model to explain the presence of enriched signal in Kermadec Ridge and Colville  Ridge lavas between ca

Referenzen

ÄHNLICHE DOKUMENTE

Lead isotope ratios of all peat samples analysed for trace element composition (n=41) were 200.. measured on the same quadrupole ICP-MS at Trinity

From these experiments, it is concluded that uptake of Ca via seawater endocytosis and transport by vesicles to the site of calcification during chamber formation can be excluded as

We therefore propose that the clinopyroxenite represents a lower crustal/upper mantle endmember beneath Rumble II East that has interacted with ascending melts. These

At the same time, glacial deep waters in the western Atlantic have become significantly less radiogenic in their Nd isotope composition due to enhanced northward flow of deep

Geochemical evidence for subduction related origin of the Bowers and Shirshov Ridges (Bering Sea, NW Pacific).. Maren Wanke 1,2 , Maxim Portnyagin 1 , Reinhard Werner 1 , Folkmar

Yangla copper deposit is hosted mainly by the gently dipping thrust faults near the Yangla granodiorite.. Five molybdenite samples yielded a well-constrained

S, Pb isotopic compositions of the Yangla copper deposit indicate that the ore-forming materials were derived from the 14.. mixture of upper crust and mantle,

Their ages may be bracketed using seismic stratigraphy, while some also show clear spatial associations with exposed vo1canic centres such as Auckland Islands