1 s2.0 S1342937X23002617 Main
1 s2.0 S1342937X23002617 Main
1 s2.0 S1342937X23002617 Main
GR Focus Review
PII: S1342-937X(23)00261-7
DOI: https://doi.org/10.1016/j.gr.2023.09.013
Reference: GR 3142
Please cite this article as: S.H.A. van de Lagemaat, D.J.J. van Hinsbergen, Plate tectonic cross-roads:
Reconstructing the Panthalassa-Neotethys Junction Region from Philippine Sea Plate and Australasian oceans
and orogens, Gondwana Research (2023), doi: https://doi.org/10.1016/j.gr.2023.09.013
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
© 2023 The Author(s). Published by Elsevier B.V. on behalf of International Association for Gondwana Re-
search.
GR Focus Review
Highlights
Keywords: Kinematic reconstruction; Plate tectonics; GPlates; Junction Region; Philippine Sea
Plate
Abstract
The plate tectonic history of the Junction Region, which separated the Panthalassa and Tethys
realms, is notoriously challenging to reconstruct. The region has been dominated by intra-
oceanic subduction zones, which has led to a sparsely preserved geological record because not
only the down-going plates but also the overriding plates were lost to subduction. Even though
most lithosphere that was present in the Junction Region during the Mesozoic has been lost to
subduction, orogenic records preserve sparse geological data that provide information for a
plate tectonic reconstruction. Here we present a kinematic reconstruction of the Junction
Region back to the Jurassic, based on the present-day geological record of the circum-Philippine
Sea Plate and Australasian regions, and sparse paleomagnetic data. We provide a
comprehensive review of orogenic and oceanic architecture from Japan to the SW Pacific region
and use a systematic reconstruction protocol for a plate kinematic restoration back to the
Jurassic. Based on our reconstruction, we propose that the Molucca Sea Plate formed as an
Eocene back-arc basin behind a north-dipping subduction zone that consumed Australian
oceanic lithosphere. We find that the Jurassic oceanic lithosphere preserved in the Philippines
originated from the northern Australian margin when a back-arc basin formed. By placing our
reconstruction in mantle reference frames, we identify multiple cases of slab dragging and
suggest that the lithospheric collapse that led to Izu-Bonin Mariana forearc extension may have
been a trigger for the absolute plate motion change of the Pacific Plate that formed the Hawaii-
Emperor Bend. Finally, we show that there is no need for spontaneous subduction initiation at
the Izu-Bonin Mariana trench. Instead, subduction initiation was more likely forced through a
change in Pacific-Australia relative plate motion around 62 Ma. Subduction started along a pre-
existing Mesozoic subduction zone that had accommodated mostly transform motion since
about 85 Ma.
1. Introduction
Kinematic reconstructions of the modern oceans and orogens have revealed how since the
formation of the supercontinent Pangea in the late Paleozoic, the Earth’s plate tectonic system
has been organized in two main plate tectonic realms. In the Tethyan realm, enclosed by
Pangean continents, oceanic lithosphere formed and subducted along predominantly E-W
trending ridges and trenches, and in the Panthalassa realm, surrounding the Pangea continents,
subduction occurred predominantly radially away from the ocean, below Pangean continental
margins and marginal basins (e.g., Larson and Chase, 1972; Engebretson et al., 1985; Stampfli
and Borel, 2002; Seton et al., 2012; Torsvik and Cocks, 2017). In the Junction Region between
these two realms, located between Australia and Eurasia (Figure 1), these two plate systems
interacted, forming a complex plate boundary system with long- and short-lived subduction
zones and marginal basins (Seton and Müller, 2008; Hall, 2002, 2012; Zahirovic et al., 2014).
The interaction between these plate systems holds many clues for the understanding of the
drivers of plate tectonics and the formation and demise of oceans and subduction zones.
However, much of these subduction zones in the Junction Region were intra-oceanic, which has
a much lower propensity to leaving geological records of subduction than active continental
margins do, because not only the down-going, but also the overriding plates are eventually lost
to subduction.
To this end, we provide a comprehensive review of the modern oceanic basins and
orogenic architecture of the entire Junction Region, spanning from Japan to New Zealand
(Figure 1), following the approach of Van Hinsbergen et al. (2020a) and Boschman et al.
(2021a), and connected to the recent reconstruction of SE Asian orogenic and plate tectonic
evolution of Advokaat and Van Hinsbergen (2023), and Pacific Basin evolution as summarized
in Van de Lagemaat et al. (2023a). Based on this review, we systematically reconstruct the
Junction Region in the context of the entire plate tectonic system of the west Panthalassa and
eastern Neotethyan realms. Our reconstruction goes back to the oldest records of intra-oceanic
subduction that are preserved in the Junction Region, on the Philippines, i.e., back to the latest
Jurassic (Dimalanta et al., 2020). We will discuss the implications of our reconstruction for the
formation and destruction of plate boundaries and interpretation of modern mantle structure.
2. Reconstruction approach
To kinematically restore the Junction Region, we thoroughly review the ocean floor structure
and age of marginal basins and the surrounding orogenic architecture, of the Philippine Sea
Plate, New Guinea, and their surroundings. We review previously published geological and
geophysical data that provide kinematic information as input for our reconstruction. This
reconstruction is made using GPlates, a freely available plate reconstruction software
(www.gplates.org; Boyden et al., 2011; Müller et al., 2018). To ensure that our reconstruction is
reproducible, we use a systematic reconstruction hierarchy (Boschman et al., 2014; Van
Hinsbergen et al., 2020a), which also makes our reconstruction easily adaptable when new
kinematic data become available.
The first step in the reconstruction hierarchy aims at establishing the relative motions of
the major plates surrounding the Junction Region (Eurasia, Australia, Pacific) using the most
recent marine magnetic anomaly and fracture zone data. For the age of polarity chrons
(intervals of geologic time with a normal or reversed field) and magnetic field reversals, we use
the geomagnetic polarity time scale of Ogg (2020). Marine magnetic anomaly and fracture zone
data of marginal ocean basins within the Junction Region are reviewed, which provide
information about the timing, amount, and direction of their opening.
Our plate reconstruction has its root in Africa, of which the motion can be described
relative to an independent reference frame, i.e., the spin axis or the mantle. Prior to the
formation of the Pacific-Antarctic Ridge (~84 Ma), there is no plate circuit connection between
the Panthalassa and Indo-Atlantic realms. Following Boschman et al. (2019), we use the Pacific
mantle reference frame of Torsvik et al. (2019) combined with the slab reference frame of Van
der Meer et al. (2010) to reconstruct Panthalassa plate motions before 84 Ma.
Next, we review the architecture of accretionary orogens that contain relics of now-
subducted lithosphere and their overriding plates, and their subsequent deformation. We follow
the reconstruction philosophy of Van Hinsbergen and Schouten (2021), by reconstructing
upper/intraplate deformation and crustal accretion separately. Upper/intraplate deformation is
reconstructed based on estimates of displacement accommodated by crustal extension, strike-
slip faults, and tectonic shortening obtained from structural geology. This provides a plate
kinematic model that must be geometrically consistent, without any large over- or underlaps
when there is no geological evidence, and that follows the basic rules of plate tectonics, which
means that all plates are surrounded by plate boundaries that end in triple junctions (Cox and
Hart, 1986).
We then reconstruct the nature and geological history of the large portions of
lithosphere that have been lost to subduction, based on information preserved in accretionary
complexes and ophiolites. The analysis of the stratigraphy and metamorphic history of accreted
ocean-plate derived units (Ocean Plate Stratigraphy (OPS); Isozaki et al., 1990, Wakita and
Metcalfe, 2005; Wakita, 2015) allows us to determine the age and geological history of the
subducted oceanic lithosphere. Similarly, accreted Continental Plate Stratigraphy (CPS) allows
restoration of passive margins and microcontinents that were entrained in subduction zones
(Van Hinsbergen and Schouten, 2021). Key constraints for the history of oceanic basins also
comes from oceanic upper plate lithosphere preserved as ophiolites and overlying volcanic arcs.
Because ophiolites provide key information for the study of subduction initiation and cessation
(e.g., Guilmette et al., 2018; Stern and Gerya, 2018; Crameri et al., 2020; Lallemand and Arcay,
2021; Van Hinsbergen et al., 2021), we carefully include them in our reconstruction to provide
regional kinematic context for these processes. In addition, the geochemical composition of
ophiolites provides insight into the tectonic setting of formation of the oceanic lithosphere. We
review tectonic interpretations based on geochemistry, to facilitate the interpretation of the
reconstruction, but we do not use this information to build the reconstruction for which we only
use kinematic data. Finally, we test our reconstruction against a compilation of paleomagnetic
data from the Philippine Sea Plate (Van de Lagemaat et al., 2023b), using the paleomagnetic
reference frame of Vaes et al. (2023), and iterate where necessary, to ascertain that the
reconstruction is in accordance with paleomagnetic constraints.
To reconstruct the plate tectonic history of the Philippine Sea Plate and New Guinea region, we
first define and review the boundaries of the region of interest for our reconstruction (Figure 1
and 2; A0 versions of the geographic and tectonic maps are provided in the Supporting
Information as Figures S1 and S2). The Junction Region as we will use it throughout this study
encompasses the following region: The Philippine Sea and Caroline plates and their plate
boundaries, the Philippine archipelago and the northern Molucca Islands, New Guinea and
surrounding islands and intervening basins, the Melanesian borderlands, which includes the
Solomon Islands, Vanuatu, Fiji, and oceanic basins such as the Bismarck Sea and the Solomon
Sea, and the SW Pacific region, which comprises the series of ridges and oceanic basins to the
east of Australia (Figures 3 – 7, S1 and S2).
Surrounding the Junction Region are three major, mostly rigid tectonics plates: the
Eurasian, Australian, and Pacific plates that are connected in a plate circuit (Figure 2). To also
be able to use absolute plate motion frames, the plate circuit has its root in Africa, whose motion
relative to the Earth’s mantle or spin axis is available through e.g., hotspot and paleomagnetic
reference frames (O’Neill et al., 2005; Torsvik et al., 2008, 2012; Doubrovine et al., 2012; Vaes et
al., 2023). The plate circuit that we use in our reconstruction is Eurasia – North America – Africa
– Antarctica – Australia/Pacific, whereby we follow the circuit described in Vaes et al. (2023).
In addition to these rigid plates, there are also deformed regions that border the
Junction Region. These are the region west of the Philippine archipelago, i.e., the circum-South
China Sea region, as well as the Sulu, Celebes, and Banda Sea, and surrounding islands (Figure
3). This part of SE Asia contains continental fragments that were derived from Gondwana and
migrated to and collided with Eurasia during opening and closure of Tethyan ocean basins (Hall,
2012; Advokaat and Van Hinsbergen, 2023; Metcalf, 2013). We refer to this orogenic collage as
the SE Asian Tethysides. The south of the study region is the SW Pacific realm, hosting
extensional basins and fold-thrust belts east of Australia (Figure 4).
For the reconstruction of the rigid plates that are part of the global plate circuit and for
the deformed SE Asian and Zealandia regions, we use previously published reconstructions, as
reviewed below. The reconstruction of the circum-Junction Region provides a net area change
of the Junction Region through time, which serves as the plate kinematic boundary condition of
our new reconstruction of the Junction Region itself. Below, we first describe the plate circuit
and the kinematic data that describe the relative motions that are incorporated in our
reconstruction. Next, we give a short overview of the deformed regions adjacent to the Junction
Region, and, finally, we provide an extensive review the orogenic architecture of the Junction
Region itself.
The upper plate to the Junction Region is the Eurasian continent. The Eurasian Plate is currently
essentially rigid but is an amalgamation of formerly independently moving tectonic plates. The
North and South China blocks have moved in unison since the Triassic, but relative motion
between the China blocks and Siberia/Eurasia continued into the Cretaceous, until the closure
of the Mongol-Okhotsk oceanic basin (Klimetz, 1987; Kravchinsky et al., 2002; Cogné et al.,
2005; Van der Voo et al., 2015). For the motion of the China blocks relative to Eurasia before the
Cretaceous, we incorporate the plate model of Torsvik and Cocks (2017), which builds on the
reconstruction of Van der Voo et al. (2015).
Our reconstruction uses the global plate circuit used by Vaes et al. (2023) as basis for
their paleomagnetic reference frame. In this plate circuit, the Eurasian Plate is reconstructed
relative to the North American Plate, based on marine magnetic anomalies that formed in the
North Atlantic Ocean. We use the rotation poles of DeMets et al. (2015), Vissers and Meijer
(2012a, b), and Srivastava and Roest (1996). The North American Plate is reconstructed relative
to Africa, also based on marine magnetic anomalies in the Atlantic Ocean. In our reconstruction,
we use the rotation poles of DeMets et al. (2015), Müller et al. (1999), Gürer et al. (2022), and
Van Hinsbergen et al. (2020a).
The other two rigid plates, the Australian and Pacific plates, are both reconstructed
relative to Antarctica, which is in turn reconstructed relative to Africa. East Antarctica is
reconstructed relative to Africa based on rotation poles derived from marine magnetic
anomalies that formed along the Southwest Indian Ridge. We use the rotation poles of Bernard
et al. (2005), Cande et al. (2010), Mueller and Jokat (2019), and DeMets et al. (2021) for this
relative motion. The Australian Plate is reconstructed relative to East Antarctica based on
marine magnetic anomalies that formed through spreading along the Southeast Indian Ridge.
We incorporate the finite rotation poles of Cande and Stock (2004), Whittaker et al. (2007,
2013), and Williams et al. (2011). The Pacific Plate is reconstructed back to 83.7 Ma relative to
West Antarctica based on marine magnetic anomalies that formed along the Pacific-Antarctic
Ridge, for which we use the rotation poles of Croon et al. (2008) and Wright et al. (2015, 2016).
The motion of West Antarctica relative to East Antarctica is based on marine magnetic anomaly
constraints and continental extension estimates based on crustal thicknesses, for which we use
the rotation poles of Cande and Stock (2004), Granot et al. (2013), Granot and Dyment (2018),
and Van de Lagemaat et al. (2023a). Before the formation of the Pacific-Antarctic Ridge, the
Pacific Plate is disconnected from the plate circuit, and is instead reconstructed in a separate
Pacific absolute plate motion frame, for which we use the hotspot frame of Torsvik et al. (2019).
After its birth in the Panthalassa Ocean around 190 Ma, the Pacific Plate was surrounded
by mid-ocean ridges and was actively spreading with three oceanic plates: the Farallon Plate in
the east, the Phoenix Plate in the south, and the Izanami Plate and subsequently the Izanagi
Plate in the west (Engebretson et al., 1985; Nakanishi et al., 1992; Nakanishi and Winterer,
1998; Seton et al., 2012; Boschman and Van Hinsbergen, 2016; Boschman et al., 2021a). Marine
magnetic anomalies that record spreading between the Pacific and Izanami/Izanagi plates
between polarity chrons M5-M35 (127.5-160.9 Ma) are preserved on the Pacific Plate
(Nakanishi et al., 1992). Spreading between the Pacific and Phoenix plates is recorded by
marine magnetic anomalies that formed during chrons M1 and M29 (123.8 – 160.9 Ma;
Nakanishi et al., 1992). The Pacific oceanic crust that is currently subducting at the Izu-Bonin-
Mariana trench is oceanic crust that formed along the Pacific-Izanami/Izanagi ridges and is
younging northwards. The oldest crust that is currently subducting at the Izu-Bonin-Mariana
subduction zone in the south is Jurassic in age (c. 160 Ma), while Early Cretaceous (c. 130 Ma)
oceanic crust is currently subducting below the northernmost Izu-Bonin Mariana subduction
zone and below Japan. Pacific oceanic crust that formed through spreading with the Phoenix
Plate is in the west overlain by the Ontong Java Plateau, a Large Igneous Province (LIP) that
erupted at about 120 Ma (Mahoney et al., 1993; Larson, 1997; Chambers et al., 2004). The
emplacement of the Ontong Java LIP led to the break-up of the Phoenix Plate around 120 Ma
(Taylor, 2006; Chandler et al., 2012; Van de Lagemaat et al., 2023a). We incorporate the
evolution of the western Panthalassa oceanic basin using the reconstruction of Van de Lagemaat
et al. (2023a) for the Pacific and Phoenix plates, that of Boschman et al. (2021a, b) for the
Izanami and Izanagi plates, and of Vaes et al. (2019) for the deformation of the Eurasian margin
forming the Japan Sea between 23 and 15 Ma.
We here briefly summarize the tectonic evolution and architecture of the SE Asian Tethysides
that is overthrusted westwards by the Cretaceous and Cenozoic ophiolites of the Philippine
Mobile Belt (Rangin, 1991). We refer to Advokaat and Van Hinsbergen (2023) for a detailed
review of the SE Asian Tethysides and its tectonic reconstruction since the Mesozoic, which we
incorporate into our model.
The northernmost basin of this region is the South China Sea, which is an uppermost
Eocene to middle Miocene oceanic basin that formed after an Eocene phase of rifting that
separated a Cretaceous and older accretionary prism from the South China Block (Faure and
Ishida, 1990; Briais et al 1993, C. Li et al., 2014; Cao et al., 2021). The opening of the South
China Sea basin occurred while Borneo was converging with South China (Advokaat et al.,
2018), which was accommodated by southward subduction of a Cretaceous lithosphere that
became trapped between Borneo and South China in the Late Cretaceous: the so-called proto-
South China Sea (Van de Lagemaat et al., 2023c). On Palawan, this subduction zone was
associated with the formation of a supra-subduction zone ophiolite (the Central Palawan
Ophiolite; Figure 5) of which a plagiogranite was dated to 40.0±0.5 Ma using U-Pb zircon
geochronology (Dycoco et al., 2021) and basaltic flows were dated to 43.8±2.2 Ma using K-Ar
whole-rock dating (Fuller et al., 1991). 40Ar/39Ar and U-Pb dating of the metamorphic sole
associated with the ophiolite yielded ages of c. 34 Ma (Schlüter et al., 1996; Aurelio et al., 2014;
Keenan et al., 2016; Dycoco et al., 2021). The Central Palawan Ophiolite overlies an accretionary
prism that contains OPS of Cretaceous (c. 100 Ma) ocean island basalt (OIB) and island arc
basalts (IAT), referred to as the Southern Palawan or Calatuigas Ophiolite (Figure 5; Almasco et
al., 2000; Aurelio et al., 2014; Dycoco et al., 2021) and a fragment of the Cretaceous accretionary
prism that rifted off the South China margin, known as the Palawan Continental Terrane
(Rangin, 1991; Zamoras and Matusoka, 2001, 2004; Aurelio et al., 2014; Shao et al., 2017; Cao et
al., 2021; Advokaat and Van Hinsbergen, 2023).
The Palawan ophiolites are to the north of the Cagayan arc, and likely formed in its
forearc (Advokaat and Van Hinsbergen, 2023). The Cagayan arc is a mostly submerged
magmatic arc that formed on continental basement correlated to the SW Borneo mega-unit, that
is thought to have broken off Australia in the early-mid Mesozoic (Advokaat and Van
Hinsbergen, 2023, and references therein). The arc was active in Oligocene to Miocene time
during and likely related to the subduction of the Proto-South China Sea lithosphere during the
opening of the South China Sea (Bellon and Rangin, 1991; Silver and Rangin, 1991; Hutchison et
al., 2000). The Cagayan arc and its underlying continental basement share a passive margin with
the Sulu Sea oceanic basin to the south, a Miocene (~24-10 Ma) back-arc basin that formed by
N-S extension behind the Cagayan arc (Roeser, 1991; Schlüter et al., 1996). The Sulu Sea ocean
floor becomes younger southward, towards the Sulu Trench where a well-developed
accretionary prism separates the Sulu Sea ocean floor from the Sulu arc, which is built on
continental crust to the south (Schlüter et al., 1996; Advokaat and Van Hinsbergen, 2023, and
references therein). The conjugate of the sea floor of the Sulu Sea is interpreted to have
subducted at the Sulu Trench (Roeser, 1991). To the east, the Sulu Sea subducts eastward below
the Philippine Mobile Belt along the Negros Trench (Figure 3).
The Sulu arc is built on continental crust that is also thought to correlate to the SW
Borneo Mega-Unit (Advokaat and Van Hinsbergen, 2023). The available ages from the Sulu arc
give Miocene-Pliocene ages, ~18-3 Ma, and stratigraphic ages indicate volcanism may have
started in the late Oligocene (Bergman et al., 2000; Rangin et al., 1990). To the south, the Sulu
arc shares a passive margin with the Eocene Celebes Sea oceanic basin. Oceanic crust of the
Celebes Sea basin is being subducted towards the east below southwestern Mindanao at the
Cotabato Trench (Figure 3). The Celebes Sea is underlain by oceanic crust with magnetic
anomalies that young southward to the North Sulawesi Trench and were interpreted to have
46-36 Ma ages (Weissel, 1980; Beiersdorf et al., 1997; Gaina and Müller 2007). The conjugate
seafloor to the south is likely represented by ophiolites that are exposed on eastern and
southeastern Sulawesi and that underlie the volcanic arc of the Sulawesi North Arm (Monnier et
al., 1995; Advokaat et al., 2017). The ophiolites are underlain by units accreted from northward
subducted Australian plate-derived oceanic and continental units (Advokaat and Van
Hinsbergen, 2023, and references therein). The North Sulawesi Trench formed at approximately
8.5 Ma (Smith et al., 1990; Nichols and Hall, 1999) and may have reactivated the former
spreading ridge of the Celebes Sea (Advokaat and Van Hinsbergen, 2023).
The most important events in the kinematic reconstruction of the SE Asian Tethysides
for the Junction Region occurred prior to the Late Cretaceous, when the SW Borneo Block and
associated continental fragments were converging with Eurasia and diverging from Australia,
and since the Eocene, when the SE Asian tectonic collage was deformed, oceanic back-arc basins
opened, and the South China Sea formed (See Advokaat and Van Hinsbergen for details). During
the Eocene, at c. 45 Ma, the northward motion of the Australian Plate accelerated which resulted
in the interaction between the Australian Plate and southern Sundaland resulting in large-scale
counterclockwise rotation of Borneo in the Eocene and Miocene (Advokaat et al., 2018). This
rotation caused convergence between Borneo and South China that led to subduction of the
proto-South China Sea below northern Borneo and the Cagayan arc, while the South China Sea
opened in its wake (Rangin et al., 1990; Rangin and Silver, 1991; Hinz et al., 1994; Lee and
Lawver 1994, 1995; Hall 1996, 2002; Hall and Breitfeld, 2017). In the Mesozoic, the proto-South
China Sea region is thought to have been occupied by a paleo-Pacific Plate that subducted
northwards, westwards, and southwards below the South China, Indochina and northern
Borneo margins (e.g., Jahn et al., 1990; Lapierre et al., 1997; Hall and Breitfeld, 2017; Nong et al.,
2021). Evidence for this subduction zone is formed by arcs and accretionary prisms. Such
prisms formed during the Jurassic to Late Cretaceous (~85 Ma) in a northwest dipping
subduction zone below South China, exposed in Taiwan and on Palawan and the Calamian
Islands (Zamoras and Matsuoka, 2001, 2004; Yui et al., 2012). In addition, evidence for
extensive continental arc magmatism is found in South China, Vietnam, and SW Borneo (e.g., Li
et al., 2012; Liu et al., 2020; Nong et al., 2022; Breitfield et al., 2017; Batara and Xu, 2022).
Mesozoic subduction of the paleo-Pacific Plate is thought to have ceased around 85 Ma, based
on the end of magmatism (Breitfeld et al., 2017; Liu et al., 2020; Nong et al., 2022; Qian et al.,
2022), and the widespread deposition of upper Cretaceous-Eocene synrift sediments (Shao et
al., 2017; Breitfield et al., 2018; Conand et al., 2020; Cao et al., 2021, 2023). Subduction
continued below Japan (e.g., Isozaki et al., 1990; Vaes et al., 2019; Boschman et al., 2021a, b; Wu
et al., 2022) and the boundary between the ongoing and ceased subduction is the so-called
Qingdao Line (Wu et al., 2022). The Qingdao line represents the location of a hypothesized plate
boundary that was located between the northern and southern Ryukyu islands (around the
Kerama Gap; Figure 3) that separated ongoing Panthalassa oceanic plate subduction to the east
from a Proto-South China Sea surrounded by passive margins to the west during the late
Mesozoic and Early Cenozoic (Wu et al., 2022). It has been suggested that the end of subduction
in the proto-South China Sea embayment was related to the arrival of an oceanic plateau in the
subduction zone (e.g., Xu et al., 2022; Van de Lagemaat et al., 2023c), after which relative motion
between Pacific realm plates and Eurasia must have been accommodated at a plate boundary to
the east of the Proto-South China Sea, from the Qingdao Line southwards.
The northeastern and eastern Australian margin is deformed by extensional basins resulting in
a mostly submerged mosaic of basins and rises that consist of continental fragments separated
from Australia, as well as fragments of arc and LIP crust (Figure 4). Below the Coral Sea are
several oceanic plateaus separated by basins. The largest of these basins is the Coral Sea Basin,
where oceanic spreading was active between 62.5 and 52.9 Ma, during polarity chrons C27-C24
(Gaina et al., 1999). Opening of the Coral Sea Basin led to the separation of the Papuan and
Eastern plateaus from northeastern Australia (Figure 4). This opening may have been related to
opening of the Tasman Sea to the south, where oceanic spreading had been propagating
northwards since c. 84 Ma, after rifting had started c. 95 Ma (Gaina et al., 1998; Grobys et al.,
2008). Opening of these basins separated various plateaus from the Australian margin,
including the Eastern, Papuan, Louisiade, and Kenn plateaus and the Lord Howe, Pocklington,
and Mellish rises (Gaina et al., 1999; Collot et al., 2012; Van den Broek and Gaina, 2020;
Mortimer et al., 2017) (Figure 4).
The difference in spreading direction between the N-S opening Coral Sea and the E-W
opening Tasman Sea was accommodated by extension, crustal thinning, and in places formation
of oceanic crust in the northeastern part of the system. Short-lived oceanic spreading occurred
in the Louisiade Trough between the Louisiade Plateau and Mellish Rise between 62.5 and 59
Ma (Gaina et al., 1999). The northeast Australian margin also contains evidence for even older
extension: the Queensland Plateau, between the Coral Sea and Australia, rifted off Australia in
the Cretaceous or possibly the Late Jurassic, accommodated by extension in the Townsville
Basin (Falvey and Taylor, 1974; Struckmeyer and Symonds, 1996).
To the east of the Louisiade Plateau and Mellish Rise is an additional set of rises and
troughs. The nature of the bathymetric highs is mostly unknown. The Rennell Ridge, West
Torres Plateau, and Lapérouse rises (Figure 4) have been interpreted as continental fragments,
remnant island arcs, oceanic crust, or LIPs (Landmesser et al., 1973; Weissel and Watts, 1979;
Yan and Kroenke, 1993; Schellart et al., 2006; Seton et al., 2016). Seton et al. (2016) obtained an
40Ar/39Ar ages of 42.8±1.2 Ma from a dredged basalt sample from the Rennell Ridge, with an E-
MORB geochemical signature. In addition, a low-quality 38±5 Ma 40Ar/39Ar age was obtained
from a primitive arc tholeiitic basalt sample dredged from the western margin of the Rennell
Ridge (Mortimer et al., 2014a). Dredged basalt from the western margin of the West Torres
Plateau had N-MORB geochemistry and a 26.2±0.8 Ma 40Ar/39Ar age and a dredged basalt from
the East Lapérouse Rise yielded a 39.1±1.2 Ma 40Ar/39Ar age with E-MORB geochemistry (Seton
et al., 2016).
The Rennell Ridge is separated from the Louisiade Plateau by the Rennell Basin. Direct
age constraints for the Rennell Basin are lacking, but an Eocene to Miocene age was inferred
from seismic reflection lines (Récy et al., 1977). Additionally, based on the structure of the
basin, it was suggested that the basin formed an east-dipping trench of a subduction zone below
the Rennell Ridge (Récy et al., 1977; Weissel and Watts, 1979).
The West Torres Plateau is separated from the Rennell Ridge by the Santa Cruz Basin
(Figure 4). Seton et al. (2016) identified marine magnetic anomalies C20-C13 (43.5-33.7 Ma) in
the Santa Cruz Basin. They infer that the basin opened between 48 and 28 Ma, based on the
possible existence of marine magnetic anomaly C21 (47.8 Ma). The end of spreading is based on
29.3±1.6 Ma and 28±3 Ma 40Ar/39Ar plagioclase ages of dredged basalts from the ridge crest of
the South Rennell Trough (Mortimer et al., 2014), which is interpreted as the southern
continuation of the Santa Cruz Basin spreading center (Seton et al., 2016). Based on this
interpretation, spreading in the South Rennell Trough was also active between c. 48 and 28 Ma
(Mortimer et al., 2014a; Seton et al., 2016). The South Rennell Trough is flanked by the West
and East Lapérouse rises (Figure 4), and their elevated nature is thought to be the result of the
interplay between magma supply and extension (Seton et al., 2016).
The East Lapérouse Rise and Lord Howe Rise are separated from the West Torres
Plateau by the D’Entrecasteaux Basin. The age and geochemistry of the oceanic crust that
underlies the D’Entrecasteaux Basin remains unknown. Lapouille (1982) interpreted
Cretaceous marine magnetic anomalies, but this interpretation was rejected by Seton et al.
(2016), as the interpreted anomalies cross-cut structural trends of the basin. No new
interpretation of anomalies was made due to the lack of continuity of magnetic anomaly
patterns between magnetic profiles. Instead, Seton et al. (2016) suggested that the
D’Entrecasteaux Basin formed as part of the South Loyalty Basin, which is a hypothesized
Cretaceous oceanic basin that was lost to subduction at the New Caledonia subduction zone in
the Eocene-Oligocene (e.g., Cluzel et al., 2001).
To the east of the Lord Howe Rise are an additional set of ridges separated by oceanic
basins (Figure 4). The tectonic history of these basins was reviewed and reconstructed by Van
de Lagemaat et al. (2018a), and we incorporate their reconstruction with updates of Van de
Lagemaat et al. (2021, 2023a) into our model. We here briefly summarize the formation history
of these basins.
To the east of the Lord Howe Rise is the New Caledonia-Fairway-Aotea basin, which
separates the Norfolk Ridge from the Lord Howe Rise (Figure 4). The Norfolk Ridge forms the
easternmost part of Zealandia (Mortimer et al., 2017) and consists of Late Mesozoic and older
accretionary complexes overlain by foreland basin clastics that can be traced into New Zealand
and New Caledonia (Mortimer et al., 1998; Cluzel et al., 2012a; Maurizot et al., 2020a). The age
of opening of the New Caledonia-Fairway-Aotea basin remains uncertain (e.g., Lafoy et al., 2005;
Collot et al., 2009), but it likely opened in tandem with the Tasman Sea, between 85 and 56 Ma
(Lafoy et al., 2005; Van de Lagemaat et al., 2018a).
East of the Norfolk Ridge is the Norfolk Basin. Due to the absence of marine magnetic
anomalies, the age of the oceanic crust has been debated (Launay et al., 1982; Mortimer et al.,
1998; Sdrolias et al., 2003), but a late Oligocene to late Miocene age of formation was inferred
from a 23 ± 0.1 Ma 40Ar/39Ar age of a dredged seafloor tholeiite (Sdrolias et al., 2003). The
opening direction of the Norfolk Basin is constrained by the Cook and Vening Meinesz fracture
zones (Figure 4), which form the northern and southern limits of the basin (Herzer and Mascle,
1996; Sdrolias et al., 2004a). The Norfolk Basin is bounded in the east by the Three Kings Ridge,
which is offset from the Loyalty Ridge by the Cook Fracture Zone. The Three Kings Ridge is
interpreted as a remnant volcanic arc that formed during the Eocene-Oligocene above the east-
dipping New Caledonia subduction zone (Kroenke and Eade, 1982; Whattam et al., 2006, 2008).
Recently, latest Oligocene to earliest Miocene (25-22 Ma) and two Eocene (39-36 Ma) 40Ar/39Ar
ages were obtained from dredged lavas from the Loyalty and Three Kings Ridges (Gans et al.,
2023).
To the east of the Three Kings Ridge is the South Fiji Basin (Figure 4). This basin formed
in the Oligocene-Miocene as a back-arc basin above the Tonga-Kermadec subduction zone (e.g.,
Sdrolias et al., 2003). The age of formation of the basin is constrained by marine magnetic
anomalies, but different interpretations have been suggested (Watts et al., 1977; Malahoff et al.,
1982; Sdrolias et al., 2003; Mortimer et al., 2007). Using additional age constraints obtained
from 40Ar/39Ar dating, which yielded ages between 19.3 and 25.9 Ma, Mortimer et al. (2007)
suggested that the marine magnetic anomalies formed during chrons C9 and C6B (27.4-21.9
Ma), which is younger than previous interpretations. Spreading in the South Fiji Basin is
thought to have ceased around 15 Ma (Mortimer et al., 2007; Herzer et al., 2009, 2011).
The South Fiji Basin is bounded in the east by the Lau-Colville Ridge (Figure 4), which
formed part of the active arc above the Tonga-Kermadec subduction zone, until the arc split
around 7 Ma when the Lau Basin and Havre Trough started opening (Ruellan et al., 2003; Yan
and Kroenke, 1993). The Lau Basin-Havre Trough system is the currently active back-arc basin
above the Tonga-Kermadec subduction zone, where westward subduction of the Pacific Plate is
accommodated since the Eocene or Oligocene (Seton et al., 2012; Van de Lagemaat et al., 2018a,
2022; Sutherland et al., 2017)
The Philippine Sea Plate is an oceanic plate mostly surrounded by subduction zones, whose
oceanic crust formed at multiple spreading ridges since the Cretaceous (Figure 6; Hilde and Lee,
1984; Deschamps and Lallemand, 2002; Yamazaki et al., 2003; Sdrolias et al., 2004b; Hickey-
Vargas et al., 2013). In the east, the Philippine Sea Plate is in an upper plate position relative to
the Pacific Plate that is subducting along the Izu-Bonin-Mariana Trench (Figures 2 and 6). The
well-studied forearc above the Izu-Bonin-Mariana subduction zone reveals a magmatic
stratigraphy with geochemical compositions ranging from MORB-like to arc-like that is widely
interpreted as the product of catastrophic extension (for instance, extension rates in the Oman
ophiolite that is thought to have formed in a similar setting, were as high as 20 cm/a; Rioux et
al., 2013), associated with foundering of the nascent slab shortly after Pacific subduction
initiation below Izu-Bonin-Mariana (Stern and Bloomer, 1992; Stern, 2004; Stern et al., 2012).
The oldest age of this sequence is dated at ~52 Ma (Ishizuka et al., 2011a; Reagan et al., 2013,
2019), which is often interpreted as the age of inception of subduction at the Izu-Bonin-Mariana
trench (Stern et al., 2012; Arculus et al., 2015), although examples elsewhere showed that
similar catastrophic extension may postdate incipient subduction by some 10 Ma (Guilmette et
al., 2018). Close to the modern trench is the Omachi seamount that consists of serpentinite that
includes exhumed eclogite blocks. These eclogites have not been dated, but are overlain by
upper Eocene to lower Oligocene andesite, showing that by late Eocene, mature subduction was
underway (Ota and Kaneko, 2010).
To the west of the Izu-Bonin-Mariana trench lies the Izu-Bonin-Mariana magmatic arc.
This arc is in places emergent and exposes a stratigraphy that goes back to the Eocene (e.g., on
Guam and the Ogasawara Islands; Meijer et al., 1983; Reagan et al., 2008; Ishizuka et al., 2006).
To the west of the active arc are remnant arcs that form submarine ridges and that are
separated from the active arc by back-arc basins. Upper plate extension in the Izu-Bonin
segment in the north is restricted to the Ogasawara Trough (Nishimura, 2011). This basin
separates the Bonin Ridge, an uplifted section of the forearc, from the currently active Izu-Bonin
arc (Figure 3). Based on the interpretation that volcaniclastics, dredged from the eastern slope
of the Ogasawara Trough, were deposited before a c. 44-42 Ma episode of volcanism, Ishizuka et
al. (2006) suggested that the Ogasawara Trough opened during the Eocene, as a narrow back-
arc basin behind the Izu-Bonin-Mariana subduction zone. To the south, however, the Mariana
Trough is an active back-arc basin that separates the West Mariana Ridge, a remnant arc, from
the active Mariana arc (Figure 3). The Mariana Trough has been opening since magnetic
anomaly chron C3 (~5 Ma) that was identified in the northern half of the basin (Yamazaki et al.,
2003), consistent with the youngest volcanic rocks known from the West Mariana Ridge
remnant arc, which are 6-4 Ma based on biostratigrapy and 40Ar/39Ar whole-rock ages of
dredged basalts (Karig, 1971; Kroenke et al., 1981; Ishizuka et al., 2010).
To the west of the West Mariana Ridge and Izu-Bonin arc are the Oligocene to Miocene
Shikoku and Parece Vela basins (Figure 3). These basins separate the more recent Izu-Bonin
and Mariana arcs from the Kyushu-Palau Ridge to the west, which forms another remnant arc
(Figure 3). Sdrolias et al. (2004b) identified a conjugate set of marine magnetic anomalies in the
Shikoku Basin mirrored in a mid-basin high that is interpreted as an extinct spreading ridge.
These were interpreted in combination with radiometric ages of dredge samples as C7-C5B (24-
15 Ma) (Sdrolias et al., 2004b). In the western portion of the Parece Vela Basin, marine magnetic
anomalies interpreted as C9-C5D (27-15 Ma) were identified (Sdrolias et al., 2004b). Magnetic
anomalies were not identified elsewhere in this basin, but clear fracture zones provide
additional constraints for reconstructing its opening. These fracture zones are S-shaped and
reflect a change in spreading direction within the basin from E-W to NE-SW (Figure 3; Sdrolias
et al., 2004b). A 15.6 ± 0.1 Ma 40Ar/39Ar whole-rock age was obtained from a basalt lava that
was dredged from the Parece Vela extinct ridge (Ishizuka et al., 2010), providing a constraint for
the end of spreading.
In the south, the extinct mid-ocean ridge of the Parece Vela basin is continuous with the
Yap Trench (Figure 3). To the west of the Yap Trench lies oceanic crust that is contiguous with
the western Parece Vela Basin, younging eastwards towards the Yap Trench (Sdrolias et al.,
2004b). There is currently no active volcanism and little seismicity at the Yap trench (Sato et al.,
1997). The basement of Yap Island consists of a metamorphic igneous complex of greenschist to
amphibolite-facies rocks (Shiraki, 1971; Hawkins and Batiza, 1977). Ages of metamorphism
were obtained from dredged greenschists and amphibolites from south of the Yap Islands,
which returned a titanite U-Pb weighted mean age of 21.4 ± 0.9 Ma and 40Ar/39Ar amphibole
ages of 20.7 ± 0.1 Ma and 22.0 ± 2.1 Ma (Zhang and Zhang, 2020). Based on geochemistry, the
protoliths of the metamorphic rocks are interpreted to be of fore-arc basin basalt, ocean island
basalt, and island-arc tholeiite affinity, but the age of the protolith remains unknown (Zhang and
Zhang, 2020). The metamorphic complex is underthrusted by Miocene mélange, which contains
fragments of these metamorphic rocks as well as other volcanic and sedimentary rocks of
unknown age (Shiraki, 1971; Rytuba and Miller, 1990). The metamorphic complex and mélange
are both overlain by Miocene andesitic lava flows and tuffs, which is the youngest volcanism on
the Yap Islands (Shiraki, 1971; Rytuba and Miller, 1990). Collectively, these data show that
contractional deformation, burial, and metamorphism at the Yap Trench must have been
underway by ~21 Ma, i.e., while the Parece Vela Ridge to the north was still active. This suggests
that during the opening of the Parece Vela basin in the north, the southern part was
compressed, and the back-arc ridge inverted into a subduction zone that consumed the eastern
forearc plate.
The western boundary of the Shikoku and Parece Vela basins is the Kyushu-Palau Ridge,
another remnant arc built on lithosphere of the eastern West Philippine Basin (Figure 3).
Radiometric dating based on an extensive suite of 40Ar/39Ar ages of dredged and drilled samples
along the length of the ridge shows that it was active from at least 48 to about 25 Ma (Ishizuka
et al. 2011b), after which the Shikoku and Parece Vela Basins started opening. The
southernmost extension of the Kyushu-Palau Ridge forms the Palau arc, immediately west of the
Palau Trench. The arc is not associated with active volcanism, and subduction seismicity is
shallow and weak (Kobayashi et al., 1997). The Palau Islands consist mostly of basalt and
andesite lavas, flow breccias, and tuffs overlain by Mio-Pliocene lignite and limestones (Meijer
et al., 1983; Rytuba and Miller, 1990; Hawkins and Ishizuka, 2009). Reported K-Ar whole-rock
ages are mostly in the range of 30-34 Ma (Meijer et al., 1983; Cosca et al., 1998), with the oldest
reliable K-Ar whole-rock age being 37.7 ± 3.1 Ma (Haston et al., 1988). A late Eocene age was
also assigned based on foraminifera and nannofossils in interbedded tuffaceous marls and
limestones (Cole, 1950; Meijer et al., 1983). The youngest age derived from volcanic rocks is a K-
Ar whole-rock age of 20.1 ± 0.5 Ma (Meijer et al., 1983).
To the west of the Kyushu-Palau Ridge is the West Philippine Basin, the largest of the
Philippine Sea Plate basins (Figure 3). Marine magnetic anomalies interpreted to have formed
during polarity chrons C26 to C13 (~59-34 Ma) were identified by Hilde and Lee (1984).
Deschamps and Lallemand (2002) revised the model for the oldest part of the basin based on
improved bathymetric data and suggest spreading started at c. 54 Ma instead. Based on new
marine magnetic anomaly interpretations, Sasaki et al. (2014) suggested that spreading in the
basin already ceased around 36 Ma, and that formation of oceanic crust during the initial stages
of the basin (south of anomaly C21; 48 Ma) was disorganized, which resulted in a lack of
formation of clear marine magnetic anomalies. Nonetheless, the onset of spreading of the West
Philippine Basin is likely older than, and spreading occurred more or less perpendicular to, the
rapid extension of the Izu-Bonin-Mariana forearc and the associated formation of boninites at
52 Ma (Ishizuka et al., 2011a; Reagan et al., 2013, 2019).
Oceanic crust of the West Philippine Basin is overlain by the Benham Rise, Urdaneta
Plateau and the Oki-Daito Rise. The Benham Rise and Urdaneta Plateau are located south and
north of the extinct West Philippine Basin spreading center, roughly at equal distance. The Oki-
Daito Rise is located north of the Urdaneta Plateau. 40Ar/39Ar dating yielded ages of 35.6±0.4 Ma
and 36.2±0.5 Ma from the Benham Rise (Hickey-Vargas, 1998), and an average of 37.9±0.1 Ma
from the Urdaneta Plateau (Ishizuka et al., 2013). An older age of 40.5-44.4 Ma was obtained
from the Oki-Daito Rise, which is further away from the extinct spreading ridge, north of the
Urdaneta Plateau (Ishizuka et al., 2013). All three plateaus have OIB-like geochemistry, and
based on this and the age progression, it was suggested that a mantle plume (referred to as the
Oki-Daito Plume) triggered opening of the West Philippine Basin and that the West Philippine
Basin spreading center remained fixed on the plume for about 10 Ma (Ishizuka et al., 2013).
Similarly, Wu et al. (2016) suggested that the Cenozoic Philippine Sea Plate nucleated above the
Manus Plume and as Ishizuka et al. (2013) do not link the Oki-Daito Plume to a mantle
structure, their Oki-Daito Plume may reflect the Manus Plume.
To the north of the West Philippine Basin is a series of E-W trending ridges separated by
basins. These are, from south to north, the Oki-Daito Ridge (which is different from the Oki-
Daito Rise), the Minami-Daito Basin, the Daito Ridge, the Kita-Daito Basin, the Amami Plateau,
and the Kikai Basin. The Amami Plateau and the Kita-Daito Basin are bound to the east by the
Minami-Amami Escarpment that separates them from the Amami-Sankaku Basin (Figure 3). The
age of the basement of southernmost ridge, the Oki-Daito Ridge, remains unknown, but dredged
volcanics overlying the ridge yielded 40.5±0.3 – 48.4±0.1 40Ar/39Ar whole-rock ages (Ishizuka et
al., 2013). Dredged and drilled basalt samples showed that the Daito Ridge contains andesites,
which yielded much older, 116.9-118.9 Ma 40Ar/39Ar ages (Ishizuka et al., 2011b). Eocene
40Ar/39Ar ages from the Daito Ridge were obtained by Hickey-Vargas et al. (2013), with a 49.3 ±
0.5 plagioclase age obtained from a basalt clast and ages between 44.7 and 48.0 of hornblende
and biotite from volcaniclastic sediments. Basalts and tonalites dredged from the Amami
Plateau, again gave Cretaceous, 115.8-117.0 Ma 40Ar/39Ar ages (Hickey-Vargas, 2005). Based on
their geochemical signature, the Amami Plateau and Daito Ridge are interpreted as extinct
Mesozoic arc remnants (Hickey-Vargas, 2005; Ishizuka et al., 2011b, 2022; Hickey-Vargas et al.,
2013; Morishita et al., 2018). Even though the oldest volcanics and sediments recovered from
the Oki-Daito Ridge are of Eocene age, based on a similar stratigraphy, it was inferred that the
Oki-Daito Ridge is also a remnant Mesozoic arc (Ishizuka et al., 2022).
Dredged and drilled OIB-like basalt samples yielding 51.3-42.8 40Ar/39Ar ages were
recovered from the Minami-Daito Basin (Hickey-Vargas, 1998; Ishizuka et al., 2013). 45.8-41.0
Ma 40Ar/39Ar and 43 Ma U-Pb zircon ages were reported from andesites recovered from the
Kita-Daito Basin (Ishizuka et al., 2022). Similar 40Ar/39Ar whole-rock ages, between 49.3 and
46.8 were obtained from drilled lava flows of the Amami-Sankaku Basin (Ishizuka et al., 2018)
Collectively, these basins and ridges reveal that during the Eocene formation of the West
Philippine Basin, smaller basins also formed to the north, within an oceanic lithosphere that
contained ~118-116 Ma oceanic crust, or arc of that age built on even older crust. In other
words, the West Philippine Basin formed by breaking oceanic lithosphere that was at least 60
Ma old at the moment of inception of extension.
In the northwest of the Philippine Sea Plate, east of Taiwan, lies the Huatung Basin,
separated from the West Philippine Basin by the Gagua Ridge (Figure 3). Hilde and Lee (1984)
originally interpreted marine magnetic anomalies identified in the Huatung Basin as C19 to 16
(41 – 36 Ma), younging northward, an interpretation recently repeated by Doo et al. (2015).
However, Deschamps et al. (2000) showed that the ocean floor of the Huatung Basin has an
Early Cretaceous age: they obtained 114.7 ± 4.0 and 124.1 ± 2.5 Ma 40Ar/39Ar ages from
amphibole of dredged gabbros from the Huatung Basin. They consequently reinterpreted the
anomalies as M10-M1 (130.6–123 Ma), younging towards the south (Deschamps et al., 2000).
These Early Cretaceous ages were confirmed by a mean U-Pb age of 130.3 ± 1.0 Ma obtained
from 18 zircon grains from the same dredged gabbro sample (Huang et al., 2019). The
Cretaceous marine magnetic anomaly ages correspond well to Lower Cretaceous (Barremian)
radiolarian assemblages collected from the nearby Lanyu Island (Deschamps et al., 2000; Yeh
and Cheng, 2001). The original extent of the Huatung Basin, as well as the timing of onset and
cessation of spreading remains unknown, as no extinct spreading center has been identified.
Much of the originally surrounding lithosphere must have been consumed by subduction, e.g., at
the Ryukyu and East Luzon trenches (Deschamps et al., 2000). From the Gagua Ridge,
Cretaceous lavas were recovered (Qian et al., 2021). These lavas yielded 124.06 ± 0.27 and
123.99 ± 0.24 Ma 40Ar/39Ar ages of plagioclase and have a subduction-related arc geochemical
signature (Qian et al., 2021). The Gagua Ridge thus likely originally formed as a Cretaceous arc
that was used in the Eocene to form a transform plate boundary that separated the Cretaceous
Huatung Basin from the actively spreading West Philippine Basin.
The southern boundary of the West Philippine Basin is formed by the Mindanao
Fracture Zone, which separates it from the southern part of the Philippine Sea Plate known as
the Palau Basin (Figure 3). In this basin, Sasaki et al. (2014) identified short segments of N-S
trending marine magnetic anomalies, roughly perpendicular to the anomalies of the West
Philippine Basin. These anomalies were tentatively interpreted as the eastern flank of a
spreading center that formed during chrons C18-C15 (40-35 Ma) (Sasaki et al., 2014). The
correlation is based on a 40.4 Ma 40Ar/39Ar age of a dolerite sample from the Mindanao Fracture
Zone, located north of the magnetic survey lines (Sasaki et al., 2014; Ishizuka et al., 2015). This
N-S trending ridge must have ended against the Mindanao Fracture Zone. In the west, oceanic
crust of the Palau Basin is subducting below the Philippines, which consumed at least part of the
western conjugate lithosphere of the Palau Basin. The southeastern margin of the Palau Basin is
the ultra-slowly spreading Ayu Trough mid-oceanic ridge that separates the Philippine Sea Plate
from the Caroline Plate and that to the north transitions to the Palau Trench in the Caroline-
Philippine Sea Plate Euler pole of rotation (e.g., Weissel and Anderson, 1978; Fujiwara et al.,
1995). The total amount of rotation since the formation of the Ayu Trough is estimated at ~25°
based on the angle between the Ayu Trough rift margins (Figures 3 and 6). A crude estimate for
the age of onset of Ayu Trough spreading is 15-25 Ma based on the sediment thickness within
the basin and estimated sedimentation rates that are extrapolated from a nearby borehole
(Weissel and Anderson, 1978; Fujiwara et al., 1995). K-Ar whole rock dating of dredge samples
recovered from 62 km east of the Ayu Trough rift axis yielded 19.9 ± 0.7, 20.5 ± 1.5 and 25.2 ±
1.1 Ma ages (Kumagai et al., 1996). These dredge samples are high-alkali andesites with an
island arc geochemical signature (Kumagai et al., 1996), which suggest that the eastern margin
of the Ayu Trough was a subduction zone at least until c. 20 Ma.
The Ayu Trough forms the western boundary of the Caroline Plate, separating it from the
Philippine Sea Plate (Figures 3 and 6). The Caroline Plate is bounded to the Pacific Plate in the
north and east by the Sorol Trough and Mussau Trench. Whether the Caroline Plate is currently
moving relative to the Pacific Plate is uncertain due to its poorly understood plate boundaries
and scarce kinematic data (DeMets et al., 2010). The Mussau Trench may be the site of incipient
subduction of the Caroline Plate below the Pacific Plate (Hegarty et al., 1983; Gurnis et al.,
2004). The southern boundary of the Caroline Plate is formed by the New Guinea and Manus
trenches, which separate the Caroline Sea from the New Guinea orogen and from the Bismarck
Sea basin.
The Caroline Plate hosts an eastern and a western basin, both with E-W trending marine
magnetic anomalies (Gaina and Müller, 2007), separated by the Eauripik Rise (Figure 3). Gaina
and Müller (2007) interpreted marine magnetic anomalies in the eastern basin as chrons C15r
to C8n (c. 35-25 Ma) and in the western basin as chrons 16n to 8r (c. 36-25 Ma). While the
eastern basin contains a symmetric set of anomalies, mirrored in a single extinct spreading
ridge, the extinct spreading ridge in the western basin is not located in the center but well to the
north. Close to this extinct spreading ridge, a conjugate set of anomalies from chron C8 is
interpreted (Gaina and Müller, 2007). To the north of the extinct spreading ridge, the next
anomaly is interpreted to have formed during chron C15. To the south of the extinct spreading
ridge, two marine magnetic anomalies from chron C10 are interpreted, then two anomalies
from chron C11, and then the southern anomaly of chron C15, which is interpreted to be the
conjugate of the C15 anomaly in the north of the basin (Gaina and Müller, 2007). This marine
magnetic anomaly pattern is interpreted to be the result of several ridge-jumps during opening
of the western Caroline Basin (Gaina and Müller, 2007).
The Eocene-Oligocene oceanic crust of the Caroline Plate is juxtaposed in the north and
east with Jurassic oceanic crust of the Pacific Plate across the Sorol Trough and Mussau Trench
(Figure 3). The northern boundary between the Caroline and Pacific plates is overlain by the
West Caroline Rise and Caroline Islands Ridge, collectively known as the Caroline Ridge,
separated from each other by the Sorol Trough (Altis, 1999). The Caroline Ridge is interpreted
as a large igneous province, with an age-progressive seamount chain to the east, on the Pacific
Plate (Zhang et al., 2020). K-Ar and 40Ar/39Ar ages from different sites on the Caroline Ridge
yielded ages between 8.1±0.8 Ma and 23.9 Ma±1.2 Ma (Ridley et al., 1974; Zhang et al., 2020).
Based on the emplacement of the Caroline Ridge onto the Caroline and Pacific Plates, it was
inferred that there has only been minor motion between the two plates since at least the
Miocene (Wu et al., 2016).
Based on its morphology, the Sorol Trough was interpreted as a transtensional feature,
whereby extensional motion along the Sorol Trough was proposed to have split a once-
contiguous, plume-related Caroline Ridge into the West Caroline Rise and Caroline Islands
Ridge (Weissel and Anderson, 1978, Altis, 1999; Dong et al., 2018). The southern margin of the
Sorol Trough has large normal faults, but the northern margin is less distinctively faulted and
the crust within the trough is highly disrupted (Weissel and Anderson, 1978). Bracey and
Andrews (1974) suggested that the Sorol Trough formed as an interarc basin behind a north-
dipping subduction zone that was interpreted from a bathymetric trough along the southern
margin of the West Caroline Rise. If the Sorol Trough is an extensional feature, its modern width
shows that it would have accommodated less than 100 km of extension since the early Miocene
(Weissel and Anderson, 1978; Altis, 1999; Dong et al., 2018). A recent K-Ar age of 23.8±0.7 Ma
was obtained from basalt samples collected from a site that was interpreted to represent the
age of extension in the Sorol Trough (Yan et al., 2022). We find it more likely, however, that this
age represents volcanics from one of the Caroline ridges as the limited amount of extension that
occurred in the Sorol Trough is unlikely to have formed new oceanic crust. At present, the Sorol
Trough likely forms the current plate boundary between the Caroline and Pacific plates, with
very minor relative transform motion (Weissel and Anderson, 1978; Dong et al., 2018).
The Mussau Trench represents the present-day eastern plate boundary between the
Caroline and Pacific plates, where the Caroline Plate is starting to subduct below the Pacific
Plate (Hegarty et al., 1983; Gurnis et al., 2004). The Lyra Trough, to the east of the Mussau
Trench, however, may represent a former plate boundary (Hegarty et al., 1983; Gaina and
Müller, 2007; Wang et al., 2022) and is thought to have been a transform fault during opening of
the Caroline Sea basin (Hegarty and Weissel, 1988; Gaina and Müller, 2007).
Orogenic belts with accreted and deformed relics of now-subducted lithosphere, and remains of
forearcs are located along the Eurasian margin, on the Philippines, and in the Melanesian and
Polynesian regions. In this section, we review the orogenic architecture and the constraints they
provide for the tectonic and geological history of the lithosphere that was lost to subduction.
At the Boso triple junction, the Izu-Bonin-Mariana trench meets with the Japan and Nankai
trenches that accommodate subduction of the Pacific and Philippine Sea Plate below the islands
of Japan, respectively (Figure 3). The islands of Japan comprise a long-lived accretionary orogen,
where accretion of OPS occurred episodically during a subduction history of c. 500 million years
(Isozaki et al., 1990, 2010). The southwestern part of Japan, west of the Boso triple junction,
exposes foreland-younging thrust units that accreted in the late Permian the Middle-Late
Jurassic, the Late Cretaceous, the Eocene (around 50 Ma), and during the Miocene, alternating
with episodes of subduction without accretion, or with subduction erosion (Isozaki et al., 1990).
Throughout this history, arc magmatism remained active (Isozaki et al., 2010). This accretion
occurred at the continental margin of the South China Block, and within the accretionary
orogen, the Japan Sea back-arc basin opened in Early Miocene time (Jolivet et al., 1994; Martin,
2011; Van Horne et al., 2017). The accretionary orogen of southwest Japan is intersected by the
Median Tectonic Line, a major E-W striking fault, juxtaposing the older Inner Zone against the
younger Outer Zone. The Inner Zone comprises Permian-Jurassic accretionary complexes
intruded by Cretaceous granitoids, while the Outer Zone is dominated by Jurassic-Paleogene
accretionary complexes (e.g., Ito et al., 2009; Isozaki et al., 2010; Wallis et al., 2020) The Median
Tectonic Line is currently a right-lateral strike-slip fault (Okada, 1973; Sugiyama, 2012), but is
thought to have been the site of subduction during the Early to Late Cretaceous that may have
consumed a former back-arc basin (Boschman et al., 2021a, and references therein).
Towards the west, the Nankai Trench connects with Ryukyu Trench and the
accretionary complexes are traced from the main islands of Japan to the Ryukyu Islands (Kizaki,
1986; Wallis et al., 2020). There is a marked difference between the geology of the northern and
southern Ryukyu Islands, which are separated from each other by the Kerama Gap (Figure 3).
The youngest part of accretionary complex exposed on the northern Ryukyu Islands is exposed
along the southeast coast of Okinawa Island and is composed of trench-fill turbidites of Eocene
age (Kizaki, 1986; Miki, 1995; Ujiie, 1997, 2002; Hou et al., 2022). On the other hand, farther
northwards, on southern Kyushu, the youngest part of the accretionary complex, also mainly
composed of trench-fill turbidites, is of Early Miocene age (Taira et al., 1982; Kiminami et al.,
1994; Raimbourg et al., 2014). The Eocene accretionary complex on the Ryukyu Islands does
not expose OPS sequences, but Paleocene to early Eocene radiolarians were derived from blocks
of OPS within an Eocene mélange complex in Shikoku (Taira et al., 1988) and lower middle
Eocene mudstone associated with blocks of basalt are embedded in mid-middle Eocene
terrigenous trench-fill deposits on Kyushu (Saito, 2008). The southern Ryukyu Islands, to the
south of the Kerama Gap, on the other hand, have a basement of Late Paleozoic to Jurassic
accretionary complexes correlated to the Inner Zone, overlain by Eocene limestones,
sandstones, pyroclastics, and andesite lavas and Lower Miocene sandstones and limestones
(Kizaki, 1986). The deposition of Upper Miocene to Pleistocene siltstones, sandstones and tuff is
the first evidence of a shared geological history between the north and south Ryukyu Islands
(Kizaki, 1986). This has been used to suggest that the Izu-Bonin-Mariana trench, and therefore
the Boso trench-trench-trench triple junction, most likely migrated from a location in the
central Ryukyu Islands, at or near the Kerama Gap, towards the northeast since the Miocene
(e.g., Kimura et al., 2014).
The Ryukyu Islands are separated from continental crust of South China by the Okinawa
Trough (Figure 3). The Okinawa Trough is a young, currently active, extensional back-arc basin
underlain by thinned continental (Lee et al., 1980; Sibuet et al., 1987, 1995; Arai et al., 2017), or
possibly in places oceanic crust (Liu et al., 2016). Extension in the Okinawa Trough is thought to
have started in the late Miocene, based on the oldest sediments above an unconformity over
basement and may amount c. 100 km (Lee et al., 1980; Letouzey and Kimura, 1986; Sibuet et al.,
1987; Fabbri et al., 2004; Tanaka and Nomura, 2009; Gungor et al., 2012).
3.4.2. Taiwan
The Ryukyu Trench ends in a complex plate boundary zone that surrounds the Taiwan orogen,
where the polarity of subduction changes, and where continental crust of the South China
margin is underthrust below oceanic lithosphere of the Philippine Sea Plate (Figure 3). The
geology of Taiwan is divided into five roughly N-S trending tectonostratigraphic terranes,
separated by major east-dipping faults (Figure 5; e.g., Huang et al., 2006; Brown et al., 2011,
2017; Camanni and Ye, 2022). The three western terranes (the Coastal Plain, Western Foothills,
and the Hsuehshan Range) make-up the mostly non-metamorphic Taiwan fold-and-thrust belt,
and together with the metamorphic Central Range that lies adjacent to the suture with the
Philippine Sea Plate, forms a series of nappes that were accreted from the subducted South
China margin. The Coastal Range in the east comprises Miocene volcanic and volcaniclastic
rocks that are interpreted to be part of the Luzon arc built on the Philippine Sea Plate (Huang et
al., 2018; Brown et al., 2022).
The fold-and-thrust belt underthrusts the Central Range, comprising of the Tananao
metamorphic complex, which has been subdivided into the Tailuko and Yuli belts (e.g., Chen et
al., 2016, 2017). The Tailuko Belt comprises a schist unit, a marble unit associated with
metabasite, and a granitoid unit, which were metamorphosed to greenschist- to amphibolite-
facies (Yui et al., 2012). Based on the occurrence of metabasite, marble, and chert, the Tailuko
Belt is interpreted as metamorphosed OPS (Yui et al., 2012). Permian ages were obtained for the
marbles, based on forams, corals, and Sr-isotope data (Jahn et al., 1984, 1992). The age of
quartz-mica schist from the Tailuko Belt, interpreted as meta-foreland basin clastics, was
determined from dinoflagellate assemblies, which yielded Late Jurassic to Early Cretaceous ages
(~155-120 Ma; Chen, 1989 in Yui et al., 2012). More recently, U-Pb detrital zircon ages yielded
120-110 maximum depositional ages for the schist formation in the Tailuko Belt (Chen et al.,
2016). Based on these ages, as well as zircon-provenance data, it was inferred that the Tailuko
Belt represents a Late Jurassic to Late Cretaceous accretionary complex, where Upper Permian
oceanic crust was accreted at a northwest-dipping subduction zone below the South China
continental margin (Yui et al., 2012). The Tailuko Belt is locally intruded by Upper Cretaceous
granitic intrusions, which may belong to the youngest magmatic suite southwest of the Qingdao
Line/Kerama Gap (Wintsch et al., 2011; Yui et al., 2012; Wintsch and Li, 2014; Wu et al., 2022).
The Tailuko Belt is overlain by syn-rift sediments, which comprises Eocene shallow-marine
sandstones, conglomerates, and limestones and Miocene deep-marine argillite with thin layers
of sandstones, now metamorphosed at lower greenschist facies because of the late Neogene
Taiwan orogenesis. These are interpreted as passive margin sediments of the South China
margin that formed after the Late Cretaceous end of subduction zone and formation of the
Tailuko accretionary prism (Ho, 1986; Fisher et al., 2002; Conand et al., 2020). A similar
stratigraphic sequence is known from the northern margin of the South China Sea to the west
(Shao et al., 2017; Advokaat and Van Hinsbergen, 2023).
The Yuli Belt is to the east and structurally above the Tailuko Belt and comprises
greenschist facies, carbonaceous, quartz-micaschists, with allochthonous blocks of blueschist
facies metabasites (Zhang et al., 2020). Detrital zircon U-Pb geochronology showed that the Yuli
Belt comprises a mixture of host-rocks with ages between the Cretaceous and Miocene (Chen et
al., 2017). A middle Miocene maximum depositional age was inferred for the Yuli Belt from the
mean 15.6±0.3 Ma crystallization age of youngest zircons in blueschist samples (Chen et al.,
2017). The age of peak metamorphic conditions was determined for a retrogressed blueschist
sample to be 5.1±1.7 Ma based on a Lu-Hf garnet age (Sandmann et al., 2015). Based on these
constraints, the Yuli Belt was interpreted to have formed as the most distal part of the Miocene
South China margin, which was subsequently incorporated in a Miocene accretionary prism
(Sandmann et al., 2015; Chen et al., 2017).
To the east of and structurally above the Yuli Belt is the Lichi Mélange (Figure 5), the
unmetamorphosed equivalent of the Yuli Belt (e.g., Chen et al., 2017). The Lichi Mélange
comprises a Pliocene matrix with blocks of andesite volcanics derived from the Luzon arc to the
east, Miocene turbidites, and oceanic crust sequences with overlying deep-marine sediments
(Jahn, 1986; Lo et al., 2020). The blocks of oceanic crust are commonly referred to as the East
Taiwan Ophiolite (Chung and Sun, 1992). Zircons from radiolarian cherts and gabbros of the
East Taiwan Ophiolite yielded U-Pb ages of 14.7±0.2 Ma and 17.8±0.4 Ma, respectively (Lo et al.,
2020). U-Pb zircon ages of 16.7±0.2 Ma and 16.1±0.6 Ma were obtained from blocks of
pegmatitic gabbro and plagiogranite (Lin et al., 2019). A 14.8±1.2 Ma 40Ar/39Ar whole-rock age
was obtained from a pillow basalt sample (Lo et al., 2020). MORB geochemical signatures
further support that the oceanic crust of the East Taiwan Ophiolite formed part of the South
China Sea basin, which was subsequently accreted by nappe-stacking to the Luzon forearc
during subduction along the Manila trench (Lin et al., 2019). U-Pb zircon crystallization ages of
17.4–16.9 Ma were derived from metaplagiogranites in the Yuli Belt (Lo et al., 2022). These ages
were interpreted to be derived from the forearc crust that formed in a supra-subduction zone
environment, providing an estimate for the age of subduction initiation of the Eurasian Plate
below the Philippine Sea Plate (Lo et al., 2022).
The Luzon arc continues from Taiwan over Luzon southwards, associated with subduction of
oceanic crust of the South China Sea basin at the Manila Trench (Figure 3). The Babuyan and
Batan island groups, between Taiwan and Luzon, expose Late Miocene to recent volcanics
(Defant et al., 1989, 1990). The Luzon arc on the island of Luzon comprises 14 Ma to recent calc-
alkaline basalts and andesites, exposed in the Central Cordillera, the western mountain range of
northern Luzon to the north of the Philippine fault (Maleterre et al., 1988; Defant et al., 1990).
Motion on the 1200 km long Philippine Fault is thought to have started in the Pliocene,
around 4 Ma, contemporaneous with subduction initiation at the Philippine Trench (Aurelio et
al., 1991; Barrier et al., 1991; Aurelio, 2000). Estimates for total displacement on the fault vary
between 110 and 200 km (Mitchell et al., 1986; Cole et al., 1989).
The Luzon arc was built on oceanic crust. At the northern end of the Central Cordillera,
in the Ilocos Norte region (Figure 5), a mélange is exposed that contains blocks of radiolarian
cherts, peridotite, and subordinate metamorphic rocks in a serpentinite matrix (Queaño et al.,
2017a). Based on radiolarian biostratigraphy, the chert was assigned an uppermost Jurassic to
Lower Cretaceous age, the age of the matrix is unknown (Queaño et al., 2017a). Ultramafic and
mafic clasts within the mélange have an island arc and MORB geochemical signature and are
interpreted to have formed in a supra-subduction zone setting (Pasco et al., 2019). The mélange
is thrust over an Eocene volcaniclastic unit, with at its base interbeds of sandstones and
siltstones that conformably overly pillow basalts with an arc geochemical signature (Queaño et
al., 2017a, 2020). Other occurrences of oceanic basement are exposed in the southern Central
Cordillera, in the Baguio-Mankayan region (Figure 5), as pillow basalts and basaltic feeder
dikes, sporadically intercalated with undated radiolarian chert, and unconformably overlain by
epiclastic rocks, including volcaniclastics and turbidites (Ringenbach et al., 1990; Encarnación
et al., 1993; Queaño et al., 2008). The basement has not been dated directly, but the oldest
sedimentary formation overlying the basement in the Central Cordillera are volcaniclastics of
Eocene age (Ringenbach et al., 1990; Encarnación et al., 1993). The pillow basalts have a
transitional MORB-IAT geochemical affinity and are therefore interpreted to have formed in a
back-arc basin environment (Queaño et al., 2008).
The Central Cordillera is intruded by the Central Cordillera Diorite Complex, with active
magmatism since the late Oligocene (e.g., Hollings et al., 2011; Deng et al., 2020). This complex
intruded in several magmatic phases, of which a 26.8±0.4 U-Pb zircon age from a quartz diorite
(Encarnación et al., 1993) is the oldest reported age. Other ages from the Central Cordillera
Diorite Complex are between 23 and 0.5 Ma (MMAJ-JICA. 1977; Bellon and Yumul, 2000; Imai,
2001; Waters et al., 2011). Late Oligocene-early Miocene (c. 20-27 Ma) magmatic rocks with an
island-arc geochemical signature are also found to the east of the Central Cordillera Diorite
Complex (MMAJ-JICA, 1977; Knittel, 1983; Hollings et al., 2011).
Along the east coast of Luzon, several complete ophiolite complexes are exposed
(Figure 5), which include, from north to south, the Casiguran/Isabela, Dibut Bay/Baler,
Montalban, Camarines Norte/Calaguas Islands/Cadig, Lagonoy, and the Rapu-Rapu ophiolites
(e.g., Encarnación, 2004; Dimalanta et al., 2020). Major and trace element geochemistry
revealed transitional MORB to IAT geochemical signatures in these ophiolites and a back-arc
basin setting of formation is generally inferred (Geary, 1986; Geary et al., 1988; Geary and Kay,
1989; Billedo et al., 1996; Tamayo et al., 1998; Tejada and Castillo, 2002; Andal et al., 2005).
Radiolarian biostratigraphy from the chert sedimentary carapace of the northernmost,
Casiguran Ophiolite yielded Lower Cretaceous (upper Barremian-Albian) ages (Queaño et al.,
2013), i.e., similar as the ages of cherts on Lanyu Island (Deschamps et al., 2000; Yeh and Cheng,
2001), the ages of the ocean floor of the Huatung Basin (Huang et al., 2019) and the Gagua Ridge
arc lavas (Qian et al., 2021) to the north. For the other ophiolites, the sedimentary carapace is
either absent or has not been dated (Encarnación, 2004; Dimalanta et al., 2020). However,
radiometric ages have been reported for some of the ophiolites exposed in eastern Luzon, which
in all cases are minimum ages. A 92.0±0.5 40Ar/39Ar amphibole age was obtained from a foliated
amphibolite from the metamorphosed Dibut Bay Ophiolite, which is interpreted as the age of
metamorphism (Billedo et al., 1996). A similar metamorphic amphibole age of 99.9±7.0 Ma was
obtained from amphibolite of the structurally disrupted Camarines Norte ophiolite, using
40Ar/39Ar dating of a metagabbro (Geary et al., 1988). Ultramafic, metamorphic, and gabbroic
rocks of the Camarines Norte Ophiolite are juxtaposed on east-dipping thrust faults (Geary and
Kay, 1989). Estimates for the P-T conditions during formation of the amphibolites are 500-550
and 2-4 kbar (Geary and Kay, 1989). There are no interpretations about the cause of
metamorphism, so it remains unknown whether the metamorphism occurred as seafloor
metamorphism during extension of the ophiolite, in which case the metamorphic ages
essentially represent the age of formation of the oceanic crust, or whether metamorphism
occurred during a later phase of upper plate shortening, in which case the metamorphic ages
may be much younger than the formation of the oceanic crust.
In contrast to the Late Cretaceous metamorphic ages obtained from the Dibut Bay and
Camarines Nortes ophiolites, the Lagonoy Ophiolite yielded 40Ar/39Ar ages of 150.9±3.3 and
156.3±2.0 Ma, obtained from amphiboles from a metagabbro and metaleucodiabase sample
(Geary et al., 1988), i.e., as much as 25-30 Ma older than the crust of the Huatung Basin.
Whether these ages represent magmatic, metamorphic crystallization, or cooling ages remains
unknown, and they were thus regarded as the minimum age of the ophiolite (Geary et al., 1998).
No interpretations were made about possible causes of metamorphism. In addition, a 122.7±4.0
Ma K-Ar whole-rock age was obtained from a gabbro of the Lagonoy Ophiolite (David et al.,
1997), although this age may have suffered from Ar loss (Encarnación, 2004). Undated
metamorphic rocks, interpreted as a metamorphic sole, are exposed on easternmost Rapu-Rapu
island, below the (undated) Rapu-Rapu Ophiolite towards the west (Yumul et al., 2006). To the
southwest, on the island of Masbate to the west of the Philippine Fault, lies another Early
Cretaceous ophiolite: the Balud Ophiolite contains mafic magmatic rocks, including pillow lavas
with subordinate isotropic gabbros and diabase dikes (Manalo et al., 2015), although no
ultramafic rocks have been reported. The basalts are overlain by Lower-Upper Cretaceous
(Aptian-Cenomanian) radiolarian chert and has a transitional MORB-IAT geochemistry (Manalo
et al., 2015)
The ophiolites of southeastern Luzon and neighboring Calaguas and Rapu-Rapu islands
are unconformably overlain by Upper Cretaceous sedimentary sequences. Unconformably
overlying the Lagonoy Ophiolite are Upper Cretaceous volcanic sequences interpreted to be arc-
related, consisting of andesitic volcanics including pillow basalts, volcaniclastics, and
hemipelagic limestone interbedded with radiolarian chert (David et al., 1997). A 91.1±0.5 Ma
40Ar/39Ar was obtained from amphibole separates of a basalt flow (David et al., 1997). Finally, a
diorite pluton that intruded into the harzburgites of the Rapu-Rapu Ophiolite, yielded a
77.1±4.6 Ma K-Ar whole-rock age and may represent a plutonic equivalent of the Upper
Cretaceous arc sequence (David et al., 1997). The Camarines Norte Ophiolite to the north of the
Lagonoy Ophiolite is unconformably overlain by a sedimentary sequence of graywacke, arkose,
and mudstone, with minor spilite, which is possibly a distal equivalent of the Upper Cretaceous
arc sequence overlying the Lagonoy Ophiolite (Encarnación, 2004).
The Upper Cretaceous arc sequences of the East Luzon ophiolites are overlain by middle
to upper Eocene volcanics, volcaniclastics, limestones, and turbidite sequences, interpreted as
an arc sequence that is separate from the Cretaceous sequence (David et al., 1997; Billedo et al.,
1996; Encarnación, 2004; Queaño et al., 2020). The Eocene volcaniclastics are the equivalent of
volcaniclastics exposed in the northernmost part of Central Cordillera in the Ilocos Norte region
(Queaño et al., 2020). The Sierra Madre range of East Luzon is intruded by Eocene (49-43 Ma)
and Oligocene-earliest Miocene batholiths (33-22 Ma; MMAJ, 1977, 1987; Billedo et al., 1996;
Encarnación, 2004; Hollings et al., 2011). Also, the Balud Ophiolite of Central Luzon is overlain
by Eocene volcanics that consist of basaltic and andesitic flows with minor sandstones and
mudstones (MMAJ-JICA, 1986; Manalo et al., 2015). The Late Cretaceous and the Eocene arc
sequences of the Bicol region, to the south of the Sierra Madre range, are intruded by Oligocene
plutons which yielded K-Ar and 40Ar/39Ar ages between 36 and 30 Ma (David et al., 1997;
Encarnación, 2004).
In contrast to northern and eastern Luzon, western Luzon exposes Eocene ophiolites (Figure 5).
These ophiolites are in the hanging wall of the eastward subduction zone that consumes the
South China Sea crust, and the SE Asian tectonic mosaic including the Palawan accretionary
prism and ophiolite. The northernmost of these ophiolites is the Zambales Ophiolite, which is
subdivided into the Acoje and Coto blocks, which have transitional MORB-IAT and IAT
geochemical signatures, respectively (Hawkins and Evans, 1983; Yumul, 1989), showing they
were formed in or adjacent to the Eocene arc (e.g., Hawkins and Evans, 1983; Encarnación et al.,
1993; Yumul et al., 2000a). The Acoje Block yielded a 44.1±3 Ma K-Ar whole rock on sill cutting
pillow lavas (Fuller et al., 1989) and 44.2±0.9 Ma U/Pb zircon age on plagiogranite (Encarnación
et al., 1993), whereas the Coto block gave a whole-rock K-Ar age of 46.6±5.1 Ma from a diabase
dike intruding gabbros (Fuller et al., 1989) and a plagiogranite 45.1±0.5 Ma U/Pb zircon age
(Encarnación et al., 1993). The Coto Block is overlain by Eocene pelagic limestone interbedded
with tuffaceous turbidites (Garrison et al., 1979; Schweller et al., 1983). Emergence of the
Zambales Ophiolite occurred prior to the Early Miocene, based on the presence of ophiolite
clasts in a conglomerate and sandstone formation unconformably overlying the ophiolite,
interpreted to result from activity of the Manila Trench (Yumul et al., 2020). Miocene
sedimentary rocks overlying the Acoje Block rework Upper Jurassic to Lower Cretaceous
radiolarian cherts (Queaño et al., 2017b).
To the west of the Zambales Ophiolite are isolated outcrops of a mélange with a matrix
of sheared serpentinite containing blocks of Lower Cretaceous radiolarian chert and quartz-
sericite-chlorite schist, referred to as the West Luzon Shear Zone (Karig, 1983). This was
interpreted as a major left-lateral strike-slip fault zone that became inactive before the end of
the Oligocene, based on the lack of shearing in unconformably overlying upper Oligocene
sediments (Karig, 1983).
The incomplete and structurally disrupted Angat Ophiolite, to the southeast of the
Zambales Ophiolite, yielded a 48.1±0.5 zircon U-Pb age from a plagiogranite (Arcilla et al., 1989;
Encarnación et al., 1993). The ophiolite comprises gabbros, a sheeted dike complex and pillow
basalts. The Eocene Angat Ophiolite is in fault contact with pillow basalts associated with Upper
Cretaceous radiolarian chert overlain by Eocene volcaniclastics, referred to as the Montalban
Ophiolite (Encarnación et al., 1993). Similar to the Eocene ophiolites to the north, the Angat
Ophiolite is overlain by arc-derived sediments (Karig, 1983) and has a geochemistry displaying
MORB and IAT characteristics (Yumul, 1993).
To the southeast of Mindoro is the Sibuyan Ophiolite, on the Romblon Island Group. The
Sibuyan Ophiolite is exposed as tectonic slices separated by west-dipping thrust faults
(Dimalanta et al., 2009). A 43.2±2.5 K-Ar whole rock age was obtained from a diorite sample
(Dimalanta et al., 2009). Jurassic to Cretaceous radiolarians were reported from cherts
intercalated with pillow lavas (Maac and Ylade, 1988). However, these may be derived from a
mélange, analogous to the West Luzon Shear Zone exposed to the west of the Zambales
Ophiolite (Karig, 1983). The ophiolite is intruded by 18-20 Ma andesites and rhyolites (Bellon
and Rangin, 1991). The ophiolite is structurally above a tectonic mélange to the northeast that
comprises clasts of ophiolitic material and metasediment in a sheared serpentinite and red
mudstone matrix (Dimalanta et al., 2009). In addition, the ophiolite is structurally above a
metamorphic unit comprising mostly plagioclase-quartz-mica schist. K-Ar mica age
determinations from a quartz-mica schist and a mica schist yielded ages of 12.3±0.2 Ma and
12.2±0.2 Ma (Dimalanta et al., 2009). U-Pb zircon dating of the protolith of the metamorphic
unit yielded a c. 110 Ma maximum depositional age (Knittel et al., 2017).
South of the Romblon Island Group is the island of Panay, which exposes the Antique
Ophiolite. The ophiolite comprises a mantle and crustal section and is exposed in thrust slices
along SE dipping thrust faults (Tamayo et al., 2001). U-Pb ages constrain the crystallization age
of the ophiolite to 44-42 Ma (Mesalles et al., 2018). Calcarenites with Early Eocene foraminifera
comformably overly the pillow basalts (Rangin et al. 1991; Tamayo et al., 2001), but Late
Jurassic to Early Cretaceous and Late Cretaceous radiolarians have also been found in the area,
in unknown tectonic context (McCabe et al., 1982; Rangin et al., 1991). The geochemistry of the
ophiolite is intermediate between MORB and IAT (Tamayo et al., 2001; Yumul et al., 2013). The
ophiolite body is thrust towards the west over Middle Miocene clastic sediments, and these
volcaniclastics, conglomerates, sandstones, and minor carbonates also unconformably cover the
ophiolite. To the east of the ophiolite is a mélange with blocks of ophiolite material in a Middle
Miocene matrix, but the contact between the mélange and the ophiolite is not exposed (Tamayo
et al., 2001). The mélange underthrusts to the east an Oligocene-Early Miocene volcanic arc
sequence (K-Ar ages of 30-21 Ma; Bellon and Rangin, 1991) with overlying marine sediments
(Tamayo et al., 2001).
There is no evidence that the belt of Eocene ophiolites of the northwest Philippines is
separated from the Cretaceous ophiolites by a major thrust fault that may represent a former
subduction zone. Instead, they more likely represent Eocene oceanic crust that formed within
the Cretaceous oceanic crust exposed on the eastern Philippines. Their geochemistry and
overlying Eocene arc-derived sediments suggest that they formed above a subduction zone, that
is related to the Eocene and younger arc rocks that unconformably overlie the Cretaceous
ophiolites to the east. The associated trench was likely to the west, since no evidence exists that
there was an Eocene trench between the Philippines and the Philippine Sea Plate.
Finally, to the west of the Eocene ophiolites, on Mindoro Island, lies the Amnay
Ophiolite, on a ridge that forms the eastern continuation of Palawan orogenic belt (Figure 5).
The Amnay Ophiolite has been interpreted to have formed in the forearc of the subduction zone
that consumed the proto-South China Sea oceanic lithosphere (Yu et al., 2020), and has been
correlated to the Palawan Ophiolite (Advokaat and Van Hinsbergen, 2023). A Middle Oligocene
age was assigned to the ophiolite based on foraminifera in siltstone intercalated with pillow
basalts (Rangin et al., 1985; Sarewitz and Karig, 1986). Zircon U-Pb dating of two metagabbro
samples yielded zircon ages of 23.3±0.2 Ma and 23.6±0.4 Ma (Yu et al., 2020). These ages are
considered magmatic crystallization ages based on the structure and chemical composition of
the zircon grains. A fresh gabbro sample, on the other hand, yielded a 33.0±0.8 Ma zircon U-Pb
age (Yu et al., 2020). The geochemistry of the ophiolite is dominantly MORB with a minor
subduction-related component (Yumul et al, 2009). Structurally below the Amnay Ophiolite are
Jurassic siltstones and Upper Cretaceous (Campanian-Maastrichtian) black shales that are
unconformably overlain by Middle Eocene-Lower Oligocene syn-rift clastic sediments and
Upper Oligocene-Lower Miocene post-rift clastic sediments and carbonates, correlated to the
Palawan Continental Terrane (Marchadier and Rangin, 1990; Advokaat and Van Hinsbergen,
2023).
Southern Philippines Ophiolites; Cebu arc, Southeast Bohol accretionary prism and
trench; Mindanao; Halmahera
The central and southern Philippines are divided by the Philippine Fault Zone into an eastern
belt with a NNW-SSE structural grain, and a western belt with a nearly perpendicular, ENE-SWS
structural grain. The eastern zone contains several prominent ophiolite complexes, from north
to south including the Samar, Tacloban, Malitbog, Dinagat, Surigao, and Pujada ophiolites. The
western zone from north to south includes the Cebu arc, the SE Bohol Ophiolite and
accretionary prism, the proto-SE Bohol Trench, and on Mindanao, the Polanco and Titay
ophiolites (Figure 5).
The Samar Ophiolite is exposed on southernmost Samar Island (Figure 5). Based on
radiolarian biostratigraphy of chert intercalating with pillow basalts, the Samar Ophiolite was
assigned a Late Cretaceous or possibly Early Cretaceous age. This age is supported by K-Ar
whole-rock dating of two basalt samples that yielded ages of 100.2±2.7 Ma and 97.9±2.8 Ma
(Balmater et al., 2015). Based on its geochemistry, the Samar Ophiolite has a supra-subduction
zone signature and is thought to have formed in an intra-arc or forearc setting (Guotana et al.,
2017a, 2018). Unconformably overlying the ophiolite are Upper Oligocene to Lower Pliocene
sediments that incorporate fragments of successively deeper parts of the ophiolitic sequence
and that were likely deposited adjacent to an active arc (Pacle et al., 2017).
Just west of the Samar Ophiolite is the Tacloban Ophiolite, exposed on Leyte Island
(Figure 5). Zircon U-Pb dating of a gabbro from the ophiolite complex yielded two very different
ages: 145.1±3.2 Ma and 124.7±3.3 Ma (Suerte et al., 2005). The Tacloban Ophiolite is
unconformably overlain by Upper Miocene to Lower Pliocene sediments with clasts derived
from the ophiolite (Suerte et al., 2005). On southern Leyte Island lies the Malitbog Ophiolite
(Figure 5). The different units of the ophiolite are separated by northeast trending thrust faults
(Dimalanta et al., 2006). An early Late Cretaceous age was assigned based on foraminifera in
limestones that overlie the ophiolite (Florendo, 1987). Based on geochemistry of the mantle
peridotites, the Tacloban and Malitbog ophiolites were inferred to have formed in a backarc
basin setting (Guotana et al, 2018).
To the east of the Malitbog Ophiolite, on Dinagat Island lies the Dinagat Ophiolite, which
is also exposed in northeastern Mindanao (where it is also known as the Surigao Ophiolite; e.g.,
Yumul, 2003, 2007). The Dinagat Ophiolite comprises a complete mantle-crustal section, and
sporadically, the basalt flows are intercalated with tuffs, tuffaceous sandstones, siltstones, and
shales (Dimalanta et al., 2020). The age of the ophiolite was inferred from an 84.4±4.2 Ma K-Ar
whole rock age (MMAJ-JICA, 1986), although the dated rock type is not clear. Geochemical
analyses revealed a transitional MORB-IAT and supra-subduction zone affinity of the mantle
and volcanic sections (Tamayo et al., 2004; Yumul et al., 1997). In the west, structurally below
the ophiolite are metamorphic rocks consisting of amphibolite schist, quartzo-feldspathic schist,
biotite schist, and metacherts, presumed by Santos (2014) to be Late Cretaceous in age,
although no radiometric ages were reported. The ophiolite is unconformably overlain by Upper
Eocene conglomerates interbedded with calcareous sandstone and mudstone (Santos, 2014).
First clasts of the ophiolite appear in clastic sedimentary rocks of Upper Miocene age (Santos,
2014).
The Pujada Peninsula of southeastern Mindanao exposes the Pujada Ophiolite (Figure 5)
of which zircon U-Pb geochronology from three gabbro samples yielded ages of 90.9±2.7,
90.2±2.0, and 88.4±7.6 Ma (Olfindo et al., 2019). A back-arc basin setting was interpreted for the
generation of the ophiolite, based on its geochemistry (Olfindo et al., 2019). The ophiolite is
thrusted eastwards over amphibolites, which are in turn thrust over greenschist-facies
metamorphosed mafic rocks with an eastward decreasing metamorphic grade (Hawkins et al.,
1985). This partly metamorphosed volcano-sedimentary section has a distinct island arc affinity
(Olfindo et al., 2019). Based on radiolarian and foraminiferal content, a Late Cretaceous age was
assigned to the island-arc volcano-sedimentary sequence (Yumul et al., 2003; Olfindo et al.,
2019). Basalts of the ophiolite are unconformably overlain by Eocene limestones (Mitchell et al.,
1986). Ophiolite clasts of the Pujada ophiolite first appear in an overlying Upper Miocene–
Pliocene turbidite succession (Queaño, 2005).
In summary, the southeastern Philippines ophiolites to the east of the Philippine Fault
systematically reveal Cretaceous ages, perhaps younging southward from ~100 to ~90 Ma
although some outlying ages of 125 or 145 are reported (Tacloban Ophiolite; Suerte et al.,
2005), with geochemical signatures suggesting arc, back-arc, or forearc chemistries. In places,
these are thrusted over metamorphosed volcano-sedimentary rocks with Cretaceous protoliths
and presumably Late Cretaceous ages metamorphic ages that suggest a period of upper plate
shortening above a subduction zone. There is no record of Eocene arc volcanism reported from
the southeastern Philippines, in contrast to Luzon in the northern Philippines, but instead, this
period is characterized by limestone sedimentation. The southeastern Philippine ophiolites
became uplifted and emergent in the late Miocene.
To the west of the Philippine Fault Zone, the central Philippine island of Cebu exposes
To the south of the Cebu arc, the island of Bohol exposes the SE Bohol Ophiolite (Figure
5). The pillows of the ophiolite sequence are overlain with Upper Cretaceous radiolarian chert
(Faustino et al., 2003; Dimalanta et al., 2020) which gives a minimum age for the ophiolite’s
crust. The ophiolite thrusts towards the southeast over non-metamorphosed mélange,
comprising ophiolite-derived chaotically disrupted rock units in a serpentinite matrix (De Jesus
et al., 2000). Structurally below the mélange are metamorphic rocks, including chlorite schists,
quartz-sericite schists, and amphibolites (De Jesus et al., 2000). Collectively, these units were
interpreted as a Late Cretaceous forearc and accretionary prism that formed at the proto-SE
Bohol Trench, which forms a prominent, but tectonically inactive depression to the southeast of
the island (Yumul et al., 2000b; Faustino et al., 2003). The Upper Cretaceous units are
unconformably overlain by lower Miocene to Pleistocene clastics, carbonates, and igneous units
(Faustino et al., 2003). Clasts of the ophiolite are present in middle Miocene clastics of the
overlying sedimentary sequences (Faustino et al., 2003). Zircon U-Pb dating from pyroclastics
and andesites yielded ages between 42.5±1.3 Ma and 30.7±0.2 Ma (Gong et al., 2021), showing
that Eocene-Oligocene magmatism also affected the Bohol region.
To the south of the SE Bohol Trench, on the Zamboanga Peninsula of western Mindanao
are the Titay and Polanco ophiolites, both of unknown age. The Titay Ophiolite complex is
thought to be emplaced onto continental basement that also underlies the Sulu arc (e.g.,
Pubellier et al., 1991; Tamayo et al., 2000; Yumul et al., 2004), correlated to the SW Borneo
Mega-Unit of the SE Asian Tethysides (Advokaat and Van Hinsbergen, 2023). The Titay
Ophiolite is in the north underlain by metamorphic rocks, including metagreywackes,
amphibolites, and quartz-mica schists (Tamayo et al., 2000). A latest Oligocene to early Miocene
age of metamorphism was inferred from 24.6±1.4 and 21.2±1.2 Ma K-Ar ages of amphibole
separates from amphibolites (Tamayo et al., 2000; Yumul et al., 2004). This episode of
metamorphism and associated Miocene volcanism may be related to the Miocene subduction of
the Sulu Sea below the Zamboanga Peninsula along the Sulu Trench (Yumul et al., 2004).
The Molucca Sea Plate was also subducted eastwards, below Halmahera. The forearc of
the Halmahera subduction zone is currently being underthrusted westwards below the forearc
of the Sangihe subduction zone, which means that all oceanic lithosphere of the Molucca Sea has
been consumed (Silver and Moore, 1978; McCaffrey et al., 1980). Arc magmatism related to the
eastward subduction of the Molucca Sea Plate found on Halmahera is young, 8 Ma and younger,
and slightly older on Obi to the south of Halmahera, dated at 11.8±0.7 Ma using K-Ar whole rock
dating (Baker and Malaihollo, 1996). On eastern Halmahera, and the islands of Morotai to the
north, Obi to the south, and Waigeo to the east, ophiolitic rocks are exposed, including
peridotites, gabbros, and few pillow basalts (Hall et al., 1988). From these ophiolitic rocks,
40Ar/39Ar ages of 87.3±7.0 and 73.8±1.5 Ma were obtained from hornblende minerals from
diorite samples (Ballantyne, 1990, 1991). The ophiolitic rocks of the basement complex of
Halmahera are tectonically intercalated with Upper Jurassic/Lower Cretaceous and
carbonaceous turbidites, Upper Cretaceous (Campanian-Maastrichtian) volcaniclastics, and
Eocene pelagic and shallow marine limestones (Hall et al., 1988). This sequence was interpreted
as an east-facing forearc, with ophiolites of similar age and origin as on the southern Philippines
(Hall et al., 1988), such as the Pujada Ophiolite (Olfindo et al., 2019). Oligocene-Miocene
sedimentation on Halmahera is dominated by marine marls and limestones without evidence
for magmatism, until the occurrence of volcaniclastics and lavas in the Miocene (Hall et al.,
1988). On Waigeo, to the east of Halmahera, the ophiolites are deformed and intercalated with
Lower Eocene radiolarian chert (Charlton et al., 1991; Ling et al., 1991) and unconformably
overlain by sandstone with ophiolitic detritus, Lower Oligocene volcaniclastic sandstone, Upper
Oligocene calc-alkaline island arc basalt, and Lower Miocene–Pliocene deep marine limestone
(Charlton et al., 1991), showing that here arc magmatism was active since at least early
Oligocene time.
The connection of the bivergent subduction complexes of the Sangihe and Halmahera
trenches to the trenches southwest and southeast of Mindanao is diffuse and complex. Pubellier
et al. (1999) showed that the connections are essentially two relay ramps. The thrust
displacement on the Sangihe trench decreases northward, and between the islet of Miangas and
southeast Mindanao connects to the Philippine Fault Zone. The convergent motion is northward
increasingly accommodated by the east-dipping Cotabato Trench below southwest Mindanao
(Figure 3). The east-dipping subduction below Halmahera also decreases northward and is
traced towards the Snellius Plateau, which may represent a submerged volcanic arc of
Oligocene age (Pubellier et al., 1999) perhaps correlated to the Oligocene arc sequence of
Waigeo (Pubellier et al., 2004). To the east of the Snellius Plateau, the Philippine Trench
accommodates westward underthrusting of the Philippine Sea Plate, which gradually
disappears southward towards northern Halmahera (Figure 3) (Pubellier et al., 1999).
Southern New Guinea forms the northernmost margin of Australian continental basement,
overthrusted from the north by plates of the Junction Region (Figures 1, 4, and 7). In the
southern part of central New Guinea, Precambrian and Paleozoic basement of the Australian
foreland is exposed (Davies, 2012). This basement is overlain by a series of clastic sediments
interpreted as a foreland basin to advancing nappes from the north. This foreland basin has
been referred to as the Arafura and Fly Platforms (e.g., Pigram and Symonds, 1991). In the
western part, on the Arafura Platform (Figure 7), the Precambrian basement is overlain by
mostly conformable sedimentary sequences that span the upper Proterozoic or Cambrian to the
mid to upper Cenozoic (Davies, 2012). These sediments show a history of shelf sedimentation
up to the late Paleozoic or early Triassic, intruded by Permian-Triassic continental arc
magmatism (Amiruddin, 2009; Crowhurst et al., 2004; Webb and White, 2016; Jost et al., 2018).
This was followed by rifting of the continental margin, which led to the deposition of
mudstones, sandstones, and conglomerate, associated with rift-related volcanism in the Middle
Triassic to Early Jurassic (Home et al., 1990; Pigram and Symonds, 1991). Post-rift passive
margin subsidence in the Late Jurassic to Early Cretaceous led to the deposition of marine
shales interbedded with quartz-rich sandstones. The latest Cretaceous and most of the Cenozoic
was dominated by the deposition of carbonates, with short intervals of clastic sedimentation
interpreted to reflect periods of eustatic sea-level fall (Cloos et al., 2005). Around 30 Ma, an
influx of clastic sediments occurred, which was interpreted as the onset of foreland basin
formation in response to an approaching active margin in the north (Pygram and Symons, 1991;
Quarles van Ufford and Cloos, 2005). Since the late Miocene, thick packages of terrestrial
continental clastic sediments were deposited related to uplift of an advancing Papuan and Irian
Jaya fold-and-thrust belt from the north (Davies, 2012).
The Fly Platform in the east (Figure 7), where the basement is of Permian age, has a
similar Mesozoic and Cenozoic stratigraphy as the Arafura Platform. However, a phase of uplift
and erosion led to a widespread Upper Cretaceous to Eocene hiatus that is not seen in other
parts of New Guinea (Davies, 2012). This uplift is thought to have occurred as a precursor to the
opening of the Coral Sea Basin in the Paleocene (Pygram and Symonds, 1991; Davies, 2012).
Bird’s Head Peninsula
The Sorong Fault that forms the plate boundary between the Philippine Sea Plate and
northwestern New Guinea runs across the northwestern tip of the Bird’s Head Peninsula
(Figure 4 and 7). To the north of the Sorong Fault, the Bird’s Head Peninsula exposes a similar
Oligocene-Miocene oceanic island arc sequence as on Waigeo: basaltic-andesitic lava,
agglomerate and volcaniclastics that yielded K-Ar ages of 31.5-10.5 Ma, sometimes intercalated
with limestone and gabbro intrusions (Pieters et al., 1983; 1989). These island arc volcanics
(known as the Tosem Block) are thrusted along the Koor Fault over an allochthonous unit of
Upper Jurassic to Lower Cretaceous age (known as the Tamrau Block; Figure 7), which consist
of partly metamorphosed shales, siltstones, and sandstones, interpreted as continental slope
deposits of the Australian passive margin. The Mesozoic passive margin sediments are
unconformably overlain by Eocene to middle Miocene limestones (Webb et al., 2019). Both the
Mesozoic and Eocene to Miocene sediments are intruded and overlain by Miocene granitoids
and lavas, including andesitic, dacitic and basaltic tuffs and lavas, volcaniclastics and
intercalated limestones, as well as middle Miocene calcareous mudstones and sandstones
(Webb et al., 2019). K-Ar ages of 20-9 Ma and U-Pb zircon ages of 18-10 Ma were obtained from
the Miocene volcanics (Bladon, 1988; Webb et al., 2020). Based on their dominantly calc-
alkaline nature, it was interpreted that these volcanics formed as part of a continental arc
(Webb et al., 2020). The Koor Fault is sealed by Plio-Pleistocene sediments, which constrains
the timing of thrusting of the Philippine Sea Plate and overlying arc onto the northern Bird’s
Head to late Miocene to Pliocene (Webb et al., 2019). The Tosem and Tamrau blocks are
separated from autochthonous Australian basement (Kemum Block) by the Sorong Fault
(Figure 7). Correlations across the fault based on detrital zircon ages, stratigraphy, and
structural data suggested that the northern tip of the Bird’s Head Peninsula was displaced
westwards by about 300 km along the Sorong Fault since the late Miocene-Pliocene (Webb et
al., 2019).
South of the Sorong Fault, Bird’s Head Peninsula exposes Paleozoic basement of the
Australian continent (the ‘Kemum Block’) intruded by plutons that yielded K-Ar biotite,
muscovite, and hornblende ages between 225 and 295 Ma (Pieters et al., 1983), and overlain by
Paleozoic turbidites, Permian and Mesozoic platform carbonates, Paleocene to Miocene
limestone and Miocene to recent siliciclastics (Pieters et al., 1983; Davies, 2012). In the Bird’s
Neck isthmus to the southeast of the Bird’s Head, this sequence is deformed and locally
metamorphosed in the Lengguru fold-and-thrust belt (Figure 7), a west-verging, N-S striking
fold-and-thrust belt that formed in a short time span of a few million years after 11 Ma, of which
the total amount of shortening is uncertain (Bailly et al., 2009). During formation of the fold-
and-thrust belt some of the transform motion on the Sorong Fault stepped southward towards a
transform fault on Central New Guinea (see below). This was followed by Pliocene E-W
extension leading to core complex exhumation (Bailly et al., 2009).
The Australian foreland of the Arafura and Fly Platforms is overthrusted from the north by the
roughly WNW-ESE trending fold-and-thrust belt, comprising north-dipping, south-verging
thrust faults, referred to as the Papuan fold-and-thrust belt in Papua New Guinea and the Irian
Jaya/Western fold-and-thrust belt in Indonesia (Figure 7). It comprises thrust sheets that
incorporate Australian basement that is overlain by sediments of up to Late Miocene age (Hill,
1999; Cloos et al., 2005). The onset of deformation that formed the southern, continent-derived
fold-and-thrust belt was in the Late Miocene, based on the incorporation of Miocene sediments
in the entire fold-and-thrust belt and on fission-track thermochronology (Hill and Gleadow,
1989; Hill, 1991; Hill and Raza, 1999).
The Irian Ophiolite is exposed in the western part of Central New Guinea (Figure 7). It is
mainly composed of serpentinized peridotites: crustal sections such as gabbros, sheeted dikes
and pillow basalts have not been reported (Weiland, 1999). The ophiolite is thrusted
southwards over mélange that includes metamorphosed pillow basalt, amphibolites with
greenschist and blueschist overprints, and rare eclogites (Weiland, 1999). The ophiolite and
underlying oceanic crustal rocks that were likely accreted from now-subducted oceanic
lithosphere that existed north of the Australian margin lie overthrusted on continental
metasedimentary rocks of the Ruffaer (or Derewo) unit that comprises slate and phyllites
(Cloos et al., 2005), whose protoliths are likely the Jurassic and Cretaceous shales exposed in of
the Irian fold-and-thrust belt and Arafura Platform to the south (Warren and Cloos, 2007).
Radiometric ages from the structurally coherent ophiolite are lacking. However, K-Ar amphibole
age of 57.9 3.9 Ma from a gabbro float sample, and a 68.6 1.1 Ma K-Ar white mica age from
amygdules in a pillow basalt likely date the age of oceanic crust as they were probably derived
from the ophiolite (Weiland, 1999).
The amphibolites underlying the ophiolites yielded different age clusters. The only
available U/Pb zircon age yielded 66.4±0.3 Ma (Weiland, 1999), which is similar to K-Ar ages of
68.4 ± 2.4 Ma (Weiland, 1999) and K-Ar hornblende ages between 61.4 ± 1.5 and 68.3 ± 1.5 Ma
and a K-Ar whole-rock age of 62.8±1.5 Ma obtained from amphibolite float samples (Bladon,
1988; Permana, 1995 in Weiland, 1999). These ages were interpreted by Weiland (1999) as a
metamorphic sole age suggesting incipient subduction around 68 Ma. On the other hand, K-Ar
ages of hornblende obtained from the amphibolites of 182.8 ± 8.0 Ma, 40Ar/39Ar ages between
94.2 ± 0.9 and 171.3 ± 8.1 Ma (Weiland, 1999) and a 147 ± 3.7 Ma K-Ar whole-rock age obtained
by Permana (1995) and reported by Weiland (1999), reveal a wider and much older age range.
These Jurassic amphibolite ages were interpreted as recording dynamic metamorphism that
occurred at mid-ocean ridge during sea floor hydration, thus dating the protolith age, whereby
Cretaceous ages are interpreted as partly reset ages (Weiland, 1999). Such a Jurassic seafloor
spreading age is consistent with the interpreted break-up age of northern Australia, based on
the stratigraphy of the Arafura Platform (Davies, 2012).
To the north of, and likely overlying, the Irian Ophiolite are Oligocene-Miocene volcanic
arc rocks that yielded K-Ar and U-Pb zircon ages of c. 35-24 and 12-10 Ma (Weiland, 1999). The
older suite has an intra-oceanic arc geochemical signature, whereas the younger suite was
interpreted as a volcanic arc that incorporated material derived from a continent (Weiland,
1999), consistent with the ages of the metamorphic rocks reported above.
The April Ophiolite (Figure 7), commonly referred to as the April Ultramafics, occurs as
thrusted slices of mainly peridotite with minor pyroxenite and gabbro and underlying partly
metamorphosed sedimentary and volcanic rocks of up to middle to late Eocene age (Davies,
1982; Davies and Jaques, 1984). The ophiolitic crust is undated, but a Late Cretaceous age was
assigned based on the presence of Upper Cretaceous (possibly Maastrichtian) foraminifera in
limestone intercalated with pillow basalts (Ryburn, 1980). Similar to the Irian Ophiolite to the
west, high pressure – low temperature metamorphic rocks (Om Formation and Tau Blueschists)
are exposed south of and structurally below the ophiolite (Ryburn, 1980; Baldwin et al., 2012),
with 40 – 45 Ma K-Ar ages obtained from sodic amphiboles from the Tau blueschists (Davies,
1982; Rogerson et al., 1987; Weiland, 1999). The Tau blueschists are predominantly mafic
schists with a northward increasing metamorphic grade (Ryburn, 1980). Two amphibolites that
occur close to the April Ophiolite have been dated, yielding K-Ar hornblende ages of 27.2 0.6
and 23.8 2.8 Ma (Page, 1976). Blueschist samples that were collected in the vicinity of the
April Ophiolite yielded K-Ar mica ages of 28.0 0.6 and 24.9 0.4 Ma (Davies, 1982; Rogerson
et al., 1987). To the north, the ultramafic complex is intruded by Early to Late Miocene arc
plutons comprising mostly granodiorite and diorite (Davies, 1980a).
The Marum Ophiolite forms the easternmost ophiolite complex of Central New Guinea
(Figure 7). It consists mainly of peridotites and gabbros that to the south lie thrusted on a thrust
sheet that contains pillow lavas, lava breccia, volcaniclastics, and argillite (Davies and Jaques,
1984). Based on geochemical differences between the basalt sheet (enriched in LREE, Ti, Zr;
transitional MORB; Jaques et al., 1978, 1983) and the ophiolite (depleted in LREE, Ti, Zr, Y) the
basalt sheet is thought to derive from oceanic lithosphere that subducted below the ophiolite
(Davies and Jaques, 1984). The only radiometric ages were reported by Jaques (1981), which
are a K-Ar age of 173 Ma of plagioclase separates from cumulus gabbros and a 59 ± 2.5 Ma age
from hornblende in a granophyric diorite in the upper gabbro sequence. No analytical details
were provided, but these ages fall in the cluster of ages that may represent the protolith of the
subducted lithosphere, and the age of the oldest metamorphism interpreted as metamorphic
sole-related in the Irian Ophiolite. The pillow basalts are intercalated with argilites that contain
poorly preserved radiolaria of probable Eocene age (Jaques, 1981). There is no extensive
metamorphic sequence found below the ophiolite, although some low-grade metasediments are
present (Jaques, 1981). These shales, siltstones and limestones are of Late Cretaceous to Eocene
age and are thought to be derived from the Australian continental margin (Davies and Jaques,
1984).
The total amount of shortening in Papua New Guinea was recently estimated to be c. 500
km based on a cross section balancing analysis (Martin et al., 2023). This includes 220 km of
shortening related to ophiolite emplacement over the Australian continental margin between
roughly 35 and 21 Ma, 190 km of shortening within the Australian passive margin sequence
underneath the ophiolite between 21 and 9 Ma (Martin et al., 2023), and c. 100 km of shortening
within the Papuan fold-and-thrust belt to the south of the ophiolite since 9 Ma, consistent with
earlier estimates (Hill, 1991; Hobson 1986).
The Central Highlands of New Guinea is separated from the coastal mountain ranges by
prominent strike-slip fault zones that form part of the northern plate boundary zone of
Australia (Figures 4 and 7). In the north, these strike-slip faults form the eastward continuation
of the Sorong Fault Zone of the Bird’s Head and include from west to east the Yapen Fault Zone
and the Bewani-Torricelli Fault Zone. The Bewani-Torricelli Fault Zone connects eastwards to
the ridge-transform plate boundary in the Bismarck Sea, which separates the North and South
Bismarck microplates. The North Bismarck microplate is separated in the north and east from
the Caroline and Pacific plates by the Manus and Kilinailai trenches (Figure 4). The Manus
Trench continues westwards as the New Guinea Trench. The South Bismarck microplate is
separated from the Australian Plate by the onshore Ramu-Markham Fault Zone that connects
eastwards with the New Britain Trench (Figure 4; e.g., Holm et al., 2015). The Ramu-Markham
Fault Zone accommodates eastwards increasing convergent motion between the New Guinea
orogen and the South Bismarck Microplate (Koulali et al., 2015).
The coastal ranges also expose ophiolites, intruding and overlying magmatic rocks, and
sedimentary rocks. The geology of the westernmost range, the Foja Range, is poorly known, but
includes ultramafics as well as andesites, basalts and volcaniclastic sediments, presumed to be
of Paleogene age, unconformably overlain by Neogene sediments (Davies, 2012). Biak Island,
north of the Sorong Fault between the Bird’s Head Peninsula and the northern ranges of Central
New Guinea (Figure 7), exposes an ophiolite overlain by Eocene to Lower Oligocene volcanics
and Oligocene to recent shallow-marine carbonates and clastic erosion products thereof
(Saragih et al., 2020). This stratigraphy is comparable to that of the northern Bird’s Head
Peninsula and Waigeo Island to the west.
The Cyclops Ophiolite, at the north coast of New Guinea (Figure 7), comprises
peridotites, cumulate gabbros, dolerites, and lavas including pillow basalts and minor boninites
(Monnier et al., 1999). The ophiolite is deformed by S-dipping thrust faults and is underthrusted
by undated greenschist to amphibolite facies metabasites (Monnier et al., 1999). The ophiolite is
unconformably overlain by Miocene volcaniclastics and limestones. The geochemistry of the
peridotites has been interpreted as supra-subduction zone affinity (Monnier et al., 1999;
Zglinicki et al., 2020), and the crustal series have a geochemistry consistent with a back-arc
basin origin (Monnier et al., 1999). K-Ar whole rock dating yielded a 43 ± 1 Ma age for the
boninite sample, and 29.3 ± 0.7 and 29.5 ± 0.7 Ma ages of basalts with interpreted back-arc
basin affinities (Monnier et al., 1999).
In summary, the geology of the northern mountain ranges of New Guinea is generally
similar to the geology of the Central Highlands. These mountain ranges comprise Mesozoic
(Jurassic to Cretaceous) oceanic crust overlain and intruded by mostly Eocene to early Miocene
volcanics and intrusives, thrusted towards the south over metamorphosed sediments. The
exception is the Cyclops Ophiolite, which is exposed in south-dipping thrust sheets, and of
which Eocene to Oligocene crystallization ages were obtained.
Papuan Peninsula
The geology of the Papuan Peninsula overall follows a similar logic as in central New Guinea. It
exposes an ophiolite belt, underlain in the south by metamorphosed ocean- and continent-
derived, accreted metasedimentary and meta-igneous thrust slices, and overlain by Paleogene
and Neogene volcanics (Figure 7). However, the architecture of the accreted thrust slices differs
from Central New Guinea.
The ophiolite exposed on the Papuan Peninsula is referred to as the Papuan Ultramafic
Belt (e.g., Davies and Smith, 1971; Davies, 1980b; Davies and Jaques, 1984; Lus et al., 2004). The
most complete sequence of this ophiolite is exposed as a NE-ward dipping unit containing
tectonite and cumulate ultramafic rock, gabbroic rock, sheeted dikes, and pillow basalts that are
interbedded with calcareous pelagic sediments. Less complete parts of this ophiolite are found
farther east on the D’Entrecasteaux and Louisiade islands, on the Moresby Seamount, and on
Muyua/Woodlark Island (Davies and Smith, 1971; Davies and Warren, 1988; Monteleone et al.,
2001; Little et al., 2011; Webb et al., 2014; Lindley, 2021). Geochemical signatures of basalts in
the Papuan Ultramafic Belt are interpreted as varying between MORB and supra-subduction
zone signature (Jaques and Chappell, 1980; Whattam et al., 2008; Whattam, 2009). A minimum
age of the ophiolite is provided by 50-55 Ma K-Ar hornblende and plagioclase ages from
plagiogranites and diorites that intrude the ophiolite (Davies and Smith, 1971). Late Paleocene
(58.9±1.1 Ma and 58.8±0.8 Ma) 40Ar/39Ar whole rock ages were obtained from a tholeiitic lava
and boninite from the Dabi Volcanics, which are exposed east of the main ophiolite body
(Walker and McDougall, 1982). Others assigned a Late Cretaceous (Maastrichtian) age to the
ophiolite based on the ages of foraminifera in sediments intercalated with basalts (Davies and
Smith, 1971; Lus et al., 2004). The structural relationship between these stratigraphic units and
the ophiolite is unclear, and as the age of the sediments is older than the crystallization ages of
the crust of the ophiolite, we speculate that the basalts overlain by Maastrichtian sediments are
from the unit that is structurally below the ophiolite. The ophiolite is overlain by Paleocene to
Eocene carbonates, turbidites, and volcanics including pillow lavas, Upper Oligocene to middle
Miocene platform carbonates, turbidites, submarine basalt and volcaniclastics and Miocene to
recent continental clastics derived from contemporary volcanism (Davies and Smith, 1971).
The ophiolite nappe is thrusted over a metamorphic sole along the Owen Stanley Fault
Zone. The metamorphic sole grades downwards from granulite to amphibolite facies rocks (Lus
et al., 2004). The cooling of the metamorphic sole is dated to 58.3 ± 0.4 Ma, obtained from the
mean 40Ar/39Ar age of five samples (Lus et al., 2004), i.e., the same age as the age of the SSZ
ophiolitic crust, as is commonly observed worldwide (e.g., Van Hinsbergen et al., 2015). The
Owen Stanley Fault thus originally formed as a subduction interface, but currently forms the
plate boundary between the Australian and Solomon Sea plates (Davies and Jaques, 1984). The
current relative motion along the Owen Stanley Fault transitions from transtentional in the
southeast to strike-slip and transpressional in the northwest (Benyshek and Taylor, 2021).
A unit of mafic schists is thrusted below the metamorphic sole. This unit, referred to as
the Emo Metamorphics, consists of metamorphic rocks ranging from prehnite-pumpellyite to
greenschist facies, that locally overprint an older blueschist facies metamorphism (Pieters,
1978; Worthing and Crawford, 1996). Cooling of amphibolite minerals has been dated to ~34
Ma and ~14 Ma, based on 40Ar/39Ar ages (Worthing and Crawford, 1996). Protolith rock
constituted gabbro, micro-gabbro and basalt, interbedded with minor volcanoclastic sediments
(Davies and Jaques, 1984; Pieters, 1978; Worthing and Crawford, 1996; Smith, 2013). Basalts
show a geochemical signature intermediate between E-MORB and N-MORB and were
interpreted as back-arc basin-derived (Worthing and Crawford 1996; Smith, 2013; Österle et al.,
2020). An age of the protolith has not been established.
The southernmost, and structurally lowest, tectonic unit of the Papuan Peninsula is the
Aure-Moresby fold-and-thrust belt. The Aure-Moresby fold-and-thrust belt and associated
foreland basin formed since the late Miocene (Ott and Mann, 2015). Here, the orogen of the
Papuan Peninsula is thrust over thinned continental crust of the Papuan and Eastern plateaus
(Ott and Mann, 2015). The Papuan and Eastern Plateaus comprise pre-Mesozoic basement that
was separated from Australia during opening of the Coral Sea Basin (Gaina et al., 1998). The
thinned continental crust of these plateaus is covered with Upper Cretaceous to Paleocene syn-
rift clastics, Paleocene to Oligocene post-rift deep water carbonates and Oligocene to recent
pelagic sediments (Rogerson and Hilyard, 1990; Ott and Mann, 2015).
3.4.5. Louisiade Archipelago and Louisiade Ophiolite
The elevated ridge of the Papuan Peninsula (Pocklington Rise) continues to the east, mostly
submerged with small islands known as the Louisiade Archipelago, with a southward convex
shape (Figure 4). The Pocklington Rise forms the southern margin of the Woodlark Basin and is
in the east overthrust by the Solomon Islands at the San Cristobal Trench. Metamorphosed
mafic volcanics and associated intrusives are exposed on the Deboyne Islands (Davies and
Smith, 1971), and a mélange of ultramafic rocks is thrusted towards the SW over a
metasedimentary schist unit (Webb et al., 2014). The majority of the Louisiade Archipelago
comprises metasediments and intruding metagabbros, both of which underwent prehnite-
pumpellyte to greenschist facies metamorphism (Webb et al., 2014). Based on detrital zircon U-
Pb ages, it was inferred that the protoliths of the metasedimentary rocks are volcaniclastics
derived from a Cretaceous silicic Large Igneous Province of the eastern Australian margin
(Zirakparvar et al., 2013). This led to the interpretation that the metasediments of the Louisiade
Archipelago represent a rifted fragment of the Australian passive margin that became separated
from Australia during opening of the Coral Sea (Webb et al., 2014).
To the east of the Louisiade Archipelago, the Louisiade Ophiolite was recently identified
based on dredging of the northern Louisiade Plateau (Figures 4 and 7). This returned
serpentinized peridotites, basalt breccia, and lavas, together with sedimentary rocks, and
geochemical data suggest it is a supra-subduction zone ophiolite (McCarthy et al., 2022). Based
on a seismic reflection transect, the ophiolite was interpreted as obducted towards the south
onto the Louisiade Plateau, analogous to the Papuan Ultramafic Belt (Davies and Jaques, 1982;
McCarthy et al., 2022).
The Woodlark Basin is bounded by the Pocklington Rise in the south and by the Woodlark Rise
in the north and contains the N-S spreading ridge that separates the Woodlark Microplate from
Australia. The oldest unit exposed on Muyua/Woodlark Island, on the Woodlark Rise, comprises
metamorphosed sandstones and chert beds, including turbidite and shallow marine shelf
deposits of unknown age, in the southeast sometimes intercalated with basalt (Lindley, 2021).
Based on the age of the unconformably overlying Upper Oligocene to Middle Miocene platform
carbonates, the age of the oldest unit is inferred to be Paleocene to Eocene (Joseph and
Finlayson, 1991; Lindley, 2021). Middle Miocene (14-12 Ma) volcanism is not straightforwardly
linked to subduction, and may be extension-related (Lindley, 2021). This episode of post-middle
Miocene extension may be related to the early opening of the Woodlark Basin.
The Woodlark Ridge produced ocean floor with magnetic anomalies initially interpreted
as dating back C2A (3.5 Ma; Weissel et al., 1982). Taylor et al. (1999) subsequently identified
additional anomalies back to C3A (6.2 Ma) in the southeast of the basin. Reconstructed
spreading rates of the Woodlark plate decrease westward and define an Australia-Woodlark
Plate Euler pole on the eastern Papuan Peninsula. To the west of this pole, the Woodlark-
Australia plate boundary is likely formed by the Owen Stanley Fault Zone that ends in a triple
junction with the New Britain Trench and the Ramu-Markham Fault Zone. The New Britain
Trench connects to the San Cristobal Trench, which consumes the Woodlark Ridge (Figure 4).
Towards the west, the amount of extension in the Woodlark Basin decreases and just
east of the Woodlark-Australia Euler pole, on the D’Entrecasteaux Islands, the extension did not
lead to oceanic spreading, but to extension of the orogen and exhumation of previously buried
orogenic units. The D’Entrecasteaux Islands expose one of the youngest metamorphic core
complexes and the youngest exhumed ultrahigh-pressure rock on Earth, which were at a depth
of more than 90 km at about 8 Ma based on U-Pb zircon dating from eclogites and host gneisses,
and garnet Lu-Hf ages on coesite eclogite (Baldwin et al., 1993, 2004, 2008; Monteleone et al.,
2007; Zirakparvar et al., 2011). The metamorphic core complexes, exposing mafic eclogites, are
separated from non-metamorphosed mafic and ultramafic rocks by Pliocene ductile shear zones
and brittle fault zones (Davies and Warren, 1988; Baldwin et al., 1993; Hill, 1994). Final
exhumation of the rocks to the surface occurred since the Plio-Pleistocene and extension is still
active (Baldwin et al., 1993, 2004; Hill and Baldwin, 1993).
To the north of the D’Entrecasteaux Islands is the Trobriand Scarp that forms a marked
northward deepening of bathymetry towards the Solomon Sea basin (Figure 4). Marine
magnetic anomalies were first identified in the Solomon Sea basin by Joshima et al. (1987).
These authors developed different models that could fit the anomaly pattern, and then
preferred a model in which the anomalies formed during chrons C12 to C9 (33-28 Ma),
younging southwards. To explain the absence of a conjugate set of anomalies to the south (the
ophiolites of the Papuan Peninsula and d'Entrecasteaux Islands are considerably older (~59 Ma;
Davies and Smith, 1971; Walker and McDougall, 1982; Rogerson and Hilyard, 1990), they
proposed that the Trobriand Scarp may have been a subduction zone that consumed the
southern conjugate lithosphere. However, there is no clear fault or trench along the Trobriand
Scarp, nor evidence for an accretionary prism. Later, Gaina and Müller (2007) reinterpreted the
marine magnetic anomalies as chrons C19-C15 (41-35 Ma), younging northward. This would
require that the northern conjugate lithosphere has subducted northward along the still-active
San Cristobal and New Britain trenches. Extrapolating the spreading rates reconstructed from
the Solomon Sea floor back in time could well explain the ages of the ophiolites on the Papuan
Peninsula if these are part of the same oceanic lithosphere. Hence, in this scenario there is no
need to invoke that subduction occurred along the Trobriand Scarp. Instead, this scarp likely
represents the northern margin of the continental lithosphere of the Papuan and Eastern
plateaus below the obducted Papuan ophiolites.
3.4.7. North and South Bismarck plates; New Britain; New Ireland
The Melanesian Borderlands occupy the zone of microplates that intervene the Australian and
Pacific plates. This region includes the Bismarck Sea, the islands of Manus, New Britain, New
Ireland, the Solomon Islands and Vanuatu (Figure 4).
The Bismarck Sea hosts the Manus spreading ridge that is estimated to have started
around 3.5 Ma based on the identification of active spreading and marine magnetic anomalies
back to C2A (3 Ma; Taylor, 1979). This spreading ridge links up towards the west with the
Bewani-Torricelli-Yapen-Sorong fault system. The opening of this basin thus constrains the
amount of left-lateral strike-slip on these faults, which is within the range of the estimated 300-
370 km based on lithological constraints (Dow and Sukamto, 1984; Webb et al., 2019
New Britain is part of the South Bismarck Plate and is likely contiguous with the
Adelbert and Finisterre ranges of the eastern Coastal Ranges of central New Guinea (e.g., Abott,
1995). On the North Bismarck Plate are the islands of Manus and New Ireland together with
some smaller islands (Figure 4). Prior to the late Neogene extension in the Manus Basin, all
these islands were part of the same plate and the islands of the Bismarck plates have a similar
tectono-sedimentary architecture. The oldest formations of these islands comprise Upper
Eocene pillow basalts, andesite lavas, andesitic breccia and volcaniclastics interbedded with
coralline and tuffaceous limestone (Francis, 1988; Lindley, 1988, 2006; Stewart and Sandy,
1988; Davies, 2005) that overlie an unknown, but presumably oceanic crust. The age of these
formations is based on paleontological evidence and radiometric ages, including 49.0 ± 5.0 and
45.8 ± 5.0 K-Ar ages from Manus Island (Francis, 1988), and a 37.4±1.7 K-Ar age from a dike
intruding the oldest volcanics on New Ireland (Stewart and Sandy, 1988). On all three islands,
these oldest volcanic units are overlain by Oligocene to middle Miocene volcaniclastics and
limestone and intruded by upper Oligocene to lowermost Miocene arc plutons (Francis, 1988;
Lindley, 1988; Stewart and Sandy, 1988; Davies, 2005; Lindley, 2006). Subsequently, extensive
platform carbonates were deposited during a lull in magmatism from the middle to late Miocene
(Francis, 1988; Lindley, 1988; Stewart and Sandy, 1988). Regional-scale volcanism became
active again in the early Pliocene (Francis, 1988; Lindley, 1988; Stewart and Sandy, 1988).
Recent volcanism on New Britain and on islands to the north of the main island, as well as the
West Bismarck arc, is suggested to be related to subduction at the New Britain Trench (e.g.
Weissel et al., 1982; Woodhead et al., 2010; Baldwin et al., 2012; Holm et al., 2016), while the
extensional opening of the Manus Basin may be responsible for recent volcanism on the islands
northeast of New Ireland (Lindley, 2016).
The modern North Bismarck Plate continues eastward towards the island of Bougainville
(politically part of Papua New Guinea) and the Solomon Islands (Figure 4 and 7). The Solomon
Islands are in an upper plate position relative to the San Cristobal Trench that consumes the
Woodlark microplate, and Australian Plate basins and ridges, such as the Santa Cruz Basin,
along the southwestern margin of the archipelago. The northern boundary of the Solomon
Islands is the North Solomon Trench which forms the continuation of the Manus and Kilinailai
trenches (Tregoning et al., 1998). Oceanic crust of the Ontong Java Plateau on the Pacific Plate is
being subducted below the Solomon Islands along this trench, although relative convergent
motion is slow (~14-23 mm/yr; Tregoning et al., 1998; Phinney et al., 2004; Taira et al., 2004).
The Solomon Islands expose two basement terranes and two generations of arc
volcanism. The uppermost basement terrane is an ophiolite, exposed on the islands of Choiseul,
southern Santa Isabel, Guadalcanal, and Makira (Figure 7). It comprises peridotites, some
gabbro and dolerite, and abundant pillow lavas that are in places intercalated with cherts and
pelagic limestones (Coulson and Vedder, 1986; Ridgway and Coulson, 1987; Berly, 2005; Tejada
et al., 1996; Petterson et al., 1999, 2009). 40Ar/39Ar dating of basalt samples from Santa Isabel
yielded five ages between 60.9±1.2 Ma and 64.0±1.1 Ma, as well as two ages of ~46.5 Ma
(Tejada et al., 1996). A 46.5±1.2 Ma zircon U-Pb age was obtained from a gabbro sample from
Choiseul Island (Battan et al., 2022). The basalts display a wide range in geochemical signatures,
but are generally MORB-like, with some lavas having characteristics of back-arc basin basalts
(Tejada et al., 1996; Berly, 2005) A supra-subduction zone affinity was found in gabbros and
periodites on San Jorge and Santa Isabel islands (Berly, 2005; Berly et al., 2006).
The ultramafic section of the ophiolite overthrusts to the northeast a unit of volcanic
rocks and schists with a metamorphic grade up to amphibolite facies, exposed on Choiseul and
Santa Isabel (Ridgway and Coulson, 1987; Berly, 2005). This unit may represent the
metamorphic sole of the ophiolite that formed during subduction initiation. Radiometric ages
that constrain the metamorphism are sparse, but K-Ar ages of ~36.5 Ma and 44.7±2.1 Ma of
tremolite-actinolite-amphibolite from Nggela Island were obtained by Neef and McDougall
(1976), and Richards et al. (1966) presented K-Ar ages between 32.4±6.8 Ma and 51.5±6.8 Ma,
with a mean of 44±18 Ma of amphibolite and schists from Choiseul. On Santa Isabel are also
some blueschist occurrences reported, although with unknown age (Ota and Kaneko, 2010).
Exposed on the islands of Santa Isabel, Malaita, and Maramasike, to the northeast of the
ophiolite, and structurally underlying the metamorphic rocks, is an OPS sequence comprising
mainly of tholeiitic pillow and flow basalts and some minor gabbro, overlain by Cretaceous-
Pliocene pelagic cherts and limestones (Petterson et al., 1999). 40Ar/39Ar ages of 120.8±2.4 and
121.8±2.9 were obtained from basalts exposed on Malaita, and ages of 122.9±1.5 and between
90 and 95 Ma were obtained from basalts on Santa Isabel (Tejada et al., 1996). These ages are
very similar to the 120.6±0.9 Ma and 88-93 Ma 40Ar/39Ar ages obtained from drill holes on the
Ontong Java Plateau (Tejada et al., 1996; Mahoney et al., 1993). Based on these corresponding
ages and identical isotopic ratios and geochemistry (transitional between tholeiitic N-MORB
and E-MORB) of the Ontong Java Plateau (e.g., Kroenke et al., 1986; Mahoney et al., 1993; Tejada
et al., 1996, 2002; Petterson et al., 1997; Phinney et al., 1999) this sequence is interpreted as
accreted Ontong Java Plateau lavas (Petterson et al., 1999; Phinney et al., 2004). A 44.2±0.2 Ma
40Ar/39Ar age obtained from basalts on Malaita was interpreted lavas that erupted onto the
Ontong Java Plateau when it passed over the Samoan hotspot (Tejada et al., 1996). These
accreted sequences are unconformably overlain by Upper Pliocene-Pleistocene clastic
sedimentary formations (Petterson et al., 1997; Cowley et al., 2004), and accretion of Ontong
Java Plateau rocks to the Solomon Islands thus occurred since c. 4 Ma (e.g., Mann and Taira,
2004; Phinney et al., 2004).
Overlying and intruding the Solomon ophiolite are arc complexes that formed during
two periods separated by a lull. The older episode of arc volcanism occurred in the Oligocene-
Early Miocene, the younger, currently active, episode started in the latest Miocene (Kroenke et
al., 1986; Petterson et al., 1999). Tapster et al. (2014) reported 23.3-25.7 Ma U-Pb zircon ages
from tonalite and diorites, in correspondence with the 24.4 ± 0.3 K-Ar age of a tonalite (Chivas
and McDougall, 1978), both from Guadalcanal. The tonalite and dorite samples of Tapster et al.
(2014) also include zircon xenocryst samples with c. 39-33 Ma and 71-63 Ma U-Pb ages, which
may be recycled zircons from earlier arc volcanism or from the ophiolite. The oldest
volcaniclastics in the Solomon Islands are of Late Oligocene age (Cowley et al., 2004). The oldest
radiometric age reported for the younger volcanic arc sequence is a 6.4 ± 1.9 K-Ar age from
Guadalcanal (Petterson et al., 1999).
3.4.9. Vanuatu; Vitiaz Trench; North Fiji Basin.
Towards the east, the WNW-ESE trending San Cristobal trench changes its orientation and
connects with the NNW-SSE trending New Hebrides trench. The Vanuatu archipelago to the east
of this trench forms an active intra-oceanic arc above the NE-dipping New Hebrides subduction
zone (Figure 4). The active arc overlies an older arc basement, of which the oldest dated rocks
returned 39 ± 5 and 36.7 ± 1 K-Ar ages of hornblende andesite (Taylor et al., 1985), a 35.4 ± 3.8
40Ar/39Ar age of basaltic andesite and a 32.4 ± 0.57 U-Pb zircon age of dolerite (Buys et al.,
2014). These ages are similar to the oldest zircon U-Pb ages reported from Fiji (38.6 ± 0.5 Ma;
Rickard and Williams, 2013), which formed part of the same arc before opening of the North Fiji
Basin since c. 10 Ma (Yan and Kroenke, 1993; Malahoff et al., 1994; Van de Lagemaat et al.,
2018a). In addition to these late Eocene ages, several late Oligocene and early Miocene ages
have been reported from the western belt of Vanuatu (e.g. Mitchell and Warden, 1971; Crawford
et al., 2003; Buys et al., 2014), whereas magmatism in Fiji was most prominent between c. 15
and 5 Ma (Cluzel and Meffre, 2019). Furthermore, a dismembered ophiolite suite is exposed on
Pentecost Island in eastern Vanuatu, which comprises peridotites and gabbros with lenses of
amphibolites and green schist in serpentinites (Mitchell and Warden, 1971; Parrot and Dugas,
1980). An 40Ar/39Ar age of 44.9 ± 3.0 was obtained from a dolerite dyke within the ophiolite,
while an amphibolitic mylonite and a metadolerite of greenschist facies, yielded 40Ar/39Ar ages
of 33.4 ± 0.8 and 35.5 ± 0.6, respectively (Crawford et al., 2003).
The North Solomon Trench connects to the east with the Vitiaz Trench Lineament
(Fairbridge and Stewart, 1960) that currently forms a transform plate boundary between
Cretaceous crust of the Pacific Plate and Neogene crust of the North Fiji Basin (e.g., Pelletier and
Auzende, 1996; Gill et al., 2022). Not much is known about this feature but it is generally
thought that it formed originally as a subduction zone, where the Pacific Plate was being
subducted (south)westwards since the Eocene, until a polarity switch in the Miocene that was
likely associated with the arrival of the Ontong-Java Plateau in the trench (Brocher, 1985;
Pelletier and Auzende, 1996; Crawford et al., 2003).
Finally, the obduction front of ophiolites that is traced from the Papuan Peninsula to the
Louisiade Ophiolites continues from the D’Entrecasteaux Ridge to the south, over New
Caledonia towards the North Island of New Zealand (Figure 4). The architecture and evolution
of this belt was extensively described in Van de Lagemaat et al. (2018a), which we here
summarize and update with recently published results. The New Caledonia Ophiolite (Figure
S2), also referred to as the Peridotite Nappe, is interpreted to have formed in a supra-
subduction zone setting (e.g., Maurizot et al., 2020b). The age of formation of the ophiolite’s
crust is not well constrained, as existing K-Ar ages are considered unreliable (Maurizot et a.,
2020b). The age of the metamorphic sole is c. 56 Ma, based on 40Ar/39Ar hornblende and U-Pb
zircon dating of amphibolites (Cluzel et al., 2012b). Timing of emplacement of the ophiolite is
bracketed by 34 Ma age of the youngest sediments that are thrusted below the ophiolite (Cluzel
et al., 2001), and the c. 27 Ma age of post-obduction plutons that intrude into the Peridotite
Nappe (Paquette and Cluzel, 2007; Sevin et al., 2020). Based on geophysical modelling from
seismic and gravity data, the Peridotite Nappe may be contiguous with the oceanic lithosphere
that underlies the North Loyalty Basin to the north of the main island of New Caledonia (Collot
et al., 1987; Patriat et al., 2018).
The ophiolite is the highest structural unit that is exposed in New Caledonia. Below the
metamorphic sole of the ophiolite occur several units that were incorporated in the NE-dipping
New Caledonia subduction complex (Maurizot et al., 2020b). The uppermost section comprises
oceanic crust as well as a passive margin succession (Poya Terrane), which formed part of the
South Loyalty Basin, the conceptual oceanic basin that was subducted at the New Caledonia
subduction zone. Based on macro-fossils and detrital zircon provenance, the South Loyalty
Basin likely opened in the Late Cretaceous (Campanian) - Paleocene, based on geochemistry as a
back-arc basin (Cluzel et al., 2018). Structurally below the Poya Terrane is the Montagnes
Blanches Nappe, which was derived from the sedimentary cover of passive margin of the
Norfolk Ridge (Maurizot et al., 2020a, b) that forms the eastern margin of the extended
continent Zealandia, the easternmost continental crust of the Australian Plate (Mortimer et al.,
2017). A belt of HP-LT metamorphic rocks (eclogites and blueschists) of oceanic and continental
passive margin origin is exposed in the northeast of the main island of New Caledonia (Maurizot
et al., 2020b), which returned Lu/Hf ages of 38.4-36.5 Ma (Taetz et al., 2021), showing that
subduction continued at least until the latest Eocene. Syn-tectonic trench-fill sediments were
deposited during the Eocene, and unconformably overly a Cretaceous and older accretionary
prism and overlying forearc basin, related to subduction below the entire East Gondwana
margin (Maurizot, 2011; Maurizot and Cluzel, 2014).
The obduction front is inferred to continue submarine along the eastern margin of the
Norfolk Ridge and western margin of the Three Kings Ridge to the southernmost exposures: the
Northland Ophiolite in northernmost North Island, New Zealand (Figure S2). This ophiolite
formed in a supra-subduction zone setting (Whattam et al., 2004, 2005) and its oceanic crust is
thought to be of Paleocene age, based on radiolarians in limestones that are intercalated with
pillow basalts (Hollis and Hanson, 1991). The ophiolite also returned Oligocene ages: between
25.1±1.2 and 26.1±1.0 based on 40Ar/39Ar dating of tholeiitic basalts and 31.6±0.2 Ma 28.3±0.2
based on U/Pb zircon dating of a gabbro and a plagiogranite (Whattam et al., 2005, 2006). The
age of emplacement is constrained by the youngest upper Oligocene sediments that are part of
the ophiolite and the lower Miocene age of the oldest sediments sealing the basal thrust
(Ballance and Spörli, 1979).
4. Paleomagnetic data
We initially made the reconstruction based on marine geophysical and structural geological
constraints without paleomagnetic data as input, after which we tested our reconstruction
against paleomagnetic data. Where necessary, we then iterated the reconstruction within the
previous constraints (see Van Hinsbergen et al. (2020a) for details on the approach).
There is only a limited amount of paleomagnetic data available in the Junction Region, and most
of it was collected from the Philippine Sea Plate. We used the paleomagnetic data compilation of
the Philippine Sea Plate of Van de Lagemaat et al. (2023b). In addition, paleomagnetic data
relevant for the reconstruction are available from the Baliojong accreted OPS units of Sabah,
North Borneo (Van de Lagemaat et al., 2023c) from accreted rocks of the Southern
Palawan/Calatuigas Ophiolite (Almasco et al., 2000), and from the Eocene volcanic rocks of the
Solomon Islands (Musgrave, 1990). (Supplementary Table S1).
As explained in Van de Lagemaat et al. (2023b), paleomagnetic data from the Junction
Region either come from strongly deformed and uplifted plate margins, or from drill cores, and
the declinations of those data that inform vertical axis rotations display strong regional
differences owing to local deformation, or drill core rotation. We therefore only test our
reconstruction against paleolatitude data computed from inclinations. In addition, we tested our
reconstruction against paleolatitudes obtained from igneous data only, as the number of
samples collected at sedimentary localities is insufficient to correct for inclination shallowing,
which makes testing our reconstruction against sedimentary data problematic.
We use the global apparent polar wander path of Vaes et al. (2023) for the major plates
surrounding the Junction Region (Eurasia, Australia, Pacific). For times prior to the 85 Ma
connection of the Pacific Plate to Antarctica, we follow Boschman et al. (2019) in estimating the
relative rotations of the Pacific Plate to South Africa using two mantle reference frames (Torsvik
et al., 2019 for the Pacific, Van der Meer et al., 2010 for the Indo-Atlantic), and used this
connection to predict the position of the Pacific and surrounding plates (e.g., Izanagi) in the
paleomagnetic reference frame. The predicted paleolatitudes of the reconstruction that is
described in section 5 compared to the paleomagnetic data in our compilation is given in Figure
8.
5. Reconstruction
This section presents our reconstruction, which in principle straightforwardly follows from the
reconstruction protocol explained in section 2, and the kinematic and geological constraints
reviewed in section 3 and summarized in Table 1 and Figure 9. GPlates reconstruction files are
available in the Supporting Information. Here, we provide a description of the reconstruction
(see Figures 10 - 33) from the present back to the past, i.e., in order of increasing levels of
interpretation. Each section describes a period back to a moment where we made a key choice
or interpretation.
Our approach to the reconstruction of the Cenozoic Philippine Sea Plate differs
fundamentally from previous reconstructions (Hall, 2002; Gaina and Müller, 2007; Seton and
Müller, 2008; Zahirovic et al., 2014; Wu et al., 2016; Liu et al., 2023). The most important
difference with most previous renditions is that we do not use paleomagnetic data as basis to
infer vertical axis rotations of the Philippine Sea Plate. This is based on a critical re-appraisal of
paleomagnetic data from the Philippine Sea Plate, which showed that the available datasets are
small and do not unequivocally demonstrate whole-plate rotation, but may instead reflect
regional deformation-associated rotations (Van de Lagemaat et al., 2023b).
Previous plate reconstructions that focus on the Junction Region generally only go back
to the early Eocene (Hall, 2002; Gaina and Müller, 2007; Wu et al., 2016; Liu et al., 2023), as this
is the classically inferred age of subduction initiation at the Izu-Bonin-Mariana trench (Stern
and Bloomer, 1992; Reagan et al., 2010). Of the regional reconstructions, only those of Seton
and Müller (2008) and Zahirovic et al. (2014) extend back to the Early Cretaceous. Global plate
and paleogeographic reconstructions (e.g., Seton et al., 2012; Müller et al., 2019; Scotese, 2021)
do go back to the early Mesozoic, but these do not describe the data or rationale behind their
reconstruction of the Junction Region. Hence, our reconstruction is the first to consider the
entire western Pacific/Panthalassa Plate system, starting from modern geological architecture
and developing the reconstruction back in time using systematic plate boundary continuation,
assuming the simplest plate tectonic scenario that satisfies geological observations.
At the end of each reconstruction section, we compare our reconstruction with the
reconstructions of Hall (2002), Gaina and Müller (2007), Zahirovic et al. (2014), model 1 of Wu
et al. (2016), and Liu et al. (2023) back to the Early Cretaceous. We do not compare our
reconstruction to global plate reconstructions or the reconstruction of Seton and Müller (2008),
as these lack detail in the description of the Junction Region, and the global plate
reconstructions generally incorporate one of the regional reconstructions.
5.1. 0 – 15 Ma
The first key interpretation step in our reconstruction coincides with the onset of opening of the
Ayu Trough, estimated to have started sometime between 25 and 15 Ma (Weissel and Anderson,
1978; Fujiwara et al., 1995). Since that opening, the Philippine Sea Plate was connected to the
Caroline Plate. Before that time, arc rocks dredged from the margins of the Ayu Trough suggest
that it was a subduction zone (Kumagai et al., 1996). The youngest of these (c. 20 Ma) shows
that subduction must have continued to at least 20 Ma, and our reconstruction assumed that the
Ayu Trough’s opening started on the young end of the estimated range, around 15 Ma (Figures
10 – 12). This age corresponds to the oldest ages of arc magmatism of the Luzon arc that are
related to westward subduction at the Manila Trench (c. 14 Ma; Maleterre et al., 1988; Defant et
al., 1990). We initially assumed that the Sorol Trough and Mussau Trench/Lyra Trough that
bound the Caroline Plate from the Pacific Plate did not accommodate motion in the last 15 Ma
and use the resulting reconstruction to re-evaluate that assumption below.
The reconstruction of the Ayu Trough restores a ~25° clockwise rotation of the Palau
Basin (the southernmost basin of the Philippine Sea Plate) relative to the Caroline Plate (Figures
10 – 12). Just north of the Ayu Trough Euler pole, this rotation reconstructs sufficient
convergence at the Yap trench to restore an amount of lithosphere that is large enough to
contain the eastern conjugate lithosphere that formed during formation of the Parece Vela
Basin, of which only the western part remains west of Yap Island (Figures 10 - 12). Our
reconstruction thus infers that the onset of extensional opening of the Ayu Trough, in
combination with oblique opening of the Parece Vela basin after c. 20 Ma, initiated subduction
at the Yap trench at the southernmost end of the Parece Vela ridge, locally consuming the
forearc lithosphere.
In the northwest, subduction of the Philippine Sea Plate oceanic basins is restored at the
Nankai and Ryukyu trenches and the extensional opening in the Mariana Trough, active in the
last 6 Ma (Yamazaki et al., 2003), as well as subduction of the West Philippine Basin below the
Philippine Mobile Belt, since 4 Ma (Figures 10 and 11). The amount of convergence
accommodated by this subduction zone decreases both northwards and southwards, as
subduction did not occur to the north of Luzon or to the east of Halmahera. Reconstructing this
lateral gradient in convergence led us to reconstruct a small (~3) clockwise rotation of eastern
Mindanao and the Palau Basin relative to the West Philippine Basin, accommodated along the
Philippine Fault and Mindanao Fracture Zone. Within the Philippines, we restore left-lateral
strike slip faulting of 100-200 km since 4 Ma along the Philippine Fault, East Luzon Transform
Fault, and Sibuyan Sea Fault in line with geological estimates (Aurelio et al. 1991; Barrier et al.,
1991; Mitchell et al., 1986; Cole et al., 1989). The reconstructed strike-slip faulting ensures that
subduction at the Philippine Trench decreased both north and southward.
In the south, the restoration of the Ayu Trough combined with the assumption that the
Caroline Plate was rigidly connected to the Pacific Plate leads to eastward restoration of the
west Philippine plate boundary, reconstructing the now-subducted Molucca Sea Plate between
the Sangihe and Halmahera trenches. This eastward restoration of the Philippine Sea Plate is
accommodated along the Sorong-Yapen-Bewani Torricelli fault system back to 3.5, and along
the New Guinea and Manus trenches north of New Guinea before 3.5 Ma. At 6 Ma, the
Halmahera trench lines up with the western boundary of the obduction front of the Papuan
ophiolites, interpreted as a transform fault. To test whether the west Philippine/Halmahera
trenches continued into New Guinea before 6 Ma, we initially connected the Philippine Sea Plate
to the Papuan ophiolites at this time and restored the New Guinea obduction since ~30 Ma
according to kinematic constraints provided by Martin et al. (2023). However, this leads to large
overlap between the western Philippine Sea Plate domain and the SE Asian Tethysides before 6
Ma, which is clearly inconsistent with the geological record. This overlap is avoided when the
Philippine Sea Plate and the western trenches are reconstructed with the westward moving
Pacific Plate, generating 1200 km of left-lateral transform motion along the northern margin of
New Guinea, presently occupied by the New Guinea and Manus trenches, between 6 and 15 Ma.
This transform motion reconstructs a 1900 km wide Molucca Sea basin at 15 Ma, which
reconstructs as a triangular basin between the southward obducting Papuan ophiolites and the
westward moving southern Philippines/Halmahera ophiolites (Figure 12).
Within New Guinea, we reconstruct 100 km of shortening in the New Guinea fold-and-
thrust belt since 9 Ma following the balanced cross-sections of Hill (1991) and Hobson (1986).
The Sorong-Fault-Bewani-Torricelli fault zone is reconstructed as part of the North Bismarck
Microplate and the relative motion at the transform fault system thus follows from the marine
magnetic anomaly constraints in the Manus Basin (Figures 10 and 11). This leads to c. 300 km of
left-lateral transform motion within northern New Guinea since 3.5 Ma, in line with geological
data (Dow and Sukamto, 1984; Webb et al., 2019). The Woodlark Basin opening since 6.5 Ma is
reconstructed, which is balanced by resurfacing the northern conjugate oceanic lithosphere of
the Solomon Sea that was consumed by subduction at the San Cristobal Trench. Restoring the
western propagation of the Woodlark Basin buries the upper Neogene core complexes of the
D’Entrecasteaux Islands back to their UHP conditions as part of the down-going Australian
lithosphere below the obducting Papuan Ultramafic Belt. In addition, we restore southeastward
roll-back of the New Britain trench, which infers ~225 km of right-lateral strike-slip faulting at
the Ramu-Markham fault zone (Figures 10 and 11). The amount of roll-back follows from the
constraint that motion of the North Bismarck Plate relative to the Papuan ophiolites is pure
strike-slip.
Whereas restoring the westward motion of the Philippine Sea Plate in the last 15 Ma
together with the Pacific Plate avoids overlaps between the western Philippine Sea Plate and the
SE Asian Tethysides, assuming that there was no relative motion between the Pacific and
Caroline Plates back to 15 Ma introduces two other kinematic problems. Firstly, an underlap
(gap) of about 350 km forms between the Caroline Plate and the New Guinea and Manus
trenches along northern New Guinea. This requires that a south-dipping slab should be present
below the northern New Guinea margin, and while this boundary is accommodating some
thrusting alongside dominantly transform motion, there is no evidence for a tomographically
resolvable slab. Secondly, this scenario predicts paleolatitudes for the Philippine Sea Plate that
systematically offset to the north compared to paleolatitude data obtained from igneous rocks
of the Philippine Sea Plate. This problem may be overcome by restoring the Philippine Sea Plate
~3-4° southwards at 15 Ma relative to the Pacific Plate, which is permitted by, and actually
improves, the paleomagnetically predicted paleolatitudes for the Philippine Sea Plate. Candidate
structures to accommodate this motion are the margin between the Philippine Sea Plate and the
Caroline Plate east of the Ayu Trough, and the Izu-Bonin-Marianas trenches farther to the north,
or the Sorol Trough-Lyra Trough boundaries between the Caroline Plate and the Pacific Plate.
The first option would require N-S extension between the Pacific Plate and the Philippine Sea
Plate on the E-W trending segments of the Philippine Sea Plate-Pacific plate boundary, e.g.,
connecting the Mariana Trough and the Yap Trench, that should be still present today, whereas
it is not. In addition, while this scenario does resolve the underlap at the New Guinea trench, the
gap north of the Manus Trench remains and this scenario leads to overlap between the
Philippine Sea Plate and SE Asian Tethysides after 15 Ma. Hence, we favor the second option,
which infers a short-lived south-dipping subduction zone at the Sorol Trough, since 15 Ma, with
the Lyra Trough acting as a transform plate boundary between the Caroline and Pacific plates
during this time (Figures 12 and 13). The asymmetric, southward deepening Sorol Trough was
previously interpreted as a rift (Weissel and Anderson, 1978; Altis, 1999), even though its
morphology is highly asymmetric and normal faults are only found on the southern part of the
rift (Weissel and Anderson, 1978; Dong et al., 2018). We tentatively infer that these normal
faults may instead result from plate bending. There is currently no subducting slab connected to
the Pacific Plate at the Sorol Trough, but a subducted slab has long been interpreted from
tomography below the Caroline Plate (the Caroline Slab; Hall and Spakman, 2002; Van der Meer
et al. 2018), which may represent (part of) this subducted lithosphere. If subduction occurred, it
must thus have ceased sufficiently long ago for the slab to sink to the mantle transition zone,
and we model the cessation of subduction at 7 Ma, although this could be a few million years
earlier or later.
The largest difference between our reconstructions and previous ones in this time
interval is that we reconstruct subduction at the Sorol Trough to avoid major convergence along
the northern margin of New Guinea. Gaina and Müller (2007), Zahirovic et al. (2014), and Wu et
al. (2016) reconstruct up to c. 1000 km of convergence within New Guinea between the
ophiolites and arcs, since 15 Ma. Hall (2002) accommodates the convergence between the
southern margin of the Caroline Plate and the New Guinea arcs (without drawing a subduction
zone plate boundary), while Liu et al. (2023) divide the convergence between a subduction zone
in between the ophiolites and the arc and between the arc and the southern margin of the
Caroline Plate. In addition, Hall (2002) reconstructs major strike slip motion within New
Guinea.
5.2. 15 – 30 Ma
From 15 Ma back in time, we reconstructed the (northern) Caroline Plate as part of the Pacific
Plate. The Philippine Sea plate motion chain (i.e., the collection of the Philippine Sea microplates
that are reconstructed relative to each other with preserved magnetic anomalies) is prior to 15
Ma disconnected from the Caroline (Pacific) Plate, accommodated by subduction below the
eastern margin of the Ayu Trough (Figures 13 – 16), where arc rocks of early Miocene have
been dredged (Kumagai et al., 1996). Instead, we reconstructed the Philippine Sea Plate as
connected to the ophiolites that were overthrusting the northern Australian margin of New
Guinea since 30 Ma (Martin et al., 2023) and evaluated whether that assumption led to
configurations and plate motions elsewhere that conflict with data. With this logic, the relative
motion between the Philippine Sea Plate and the surrounding major plates is constrained back
to the start of ophiolite obduction on New Guinea, which is estimated to have occurred around
30 Ma based on metamorphic ages of continent-derived rocks (Weiland, 1999; N. McQuarrie,
pers. comm.).
The limited convergence between the Philippine Sea Plate and the Australian Plate,
which was restricted to the shortening documented on New Guinea, is in line with the
paleolatitudes predicted from paleomagnetic data of the Philippine Sea Plate (Figure 8; Van de
Lagemaat et al., 2023b). The bulk oceanic crust consumed by post-30 Ma subduction is thus
restored to the north of the Philippine Sea Plate, where it was subducted at the Nankai and
Ryukyu trenches. Within the Philippine Sea Plate, opening of the Shikoku and Parece Vela
Basins is restored (Sdrolias et al., 2004b), which results in a much smaller Philippine Sea Plate
mosaic by 30 Ma (Figure 16).
A key difference between our reconstruction and others in the 15-30 Ma interval is that
we treat the Philippine Mobile belt as a rigid terrane, while other reconstructions incorporate
relative motions between different parts of the Philippines, except for Zahirovic et al. (2014).
Wu et al. (2016) and Liu et al. (2023) reconstruct relative convergence between the east and
west Philippines, accommodated at the location of the present-day Philippine Fault. Hall (2002)
and Gaina and Müller (2007), on the other hand, reconstruct relative convergence between the
north and south Philippines already before c. 10 Ma, accommodated to the south of Luzon.
Further differences between the existing reconstructions in the 15-30 Ma interval are
the configuration and polarity of subduction zone plate boundaries in the New Guinea-
Melanesian region. Hall (2002) and Gaina and Müller (2007) reconstruct a similar northward-
dipping subduction zone below the Papuan ophiolites, but in their models, the Australian
margin enters the trench around 25 Ma, and no convergence is reconstructed in the 15-25 Ma
interval. Moreover, in their models, this subduction zone does not connect with the New
Caledonia subduction zone, as this trench does not exist, or only until 45 Ma. In the model of
Zahirovic et al. (2014), the New Caledonia trench does exist, but is connected to the New Guinea
trench through extensional plate boundaries. In contrast to the other reconstructions, including
ours, the reconstructions of Wu et al. (2016) and Liu et al. (2023) include southward
subduction. Liu et al. (2023) incorporate southward subduction of oceanic crust below the
entire northern margin of New Guinea, while Wu et al. (2016) reconstruct southward intra-
oceanic subduction between New Guinea and the Philippine Sea Plate as well as southward
subduction to the east of New Guinea, at the Trobriand Trough and farther to the east. While the
Trobriand Trough has been interpreted as the site of southward subduction by others (Joshima
et al., 1987), we interpreted the bathymetric depression as a scarp that represents the northern
margin of underthrusted continental lithosphere below the Papuan ophiolites.
5.3. 30 – 45 Ma
Despite the lack of kinematic data that demonstrate relative plate motion amount and
direction at the Junction Region trenches in the 30-45 Ma interval, there is enough geological
data for a well-constrained reconstruction (Figures 16 – 19). At first, we assumed a simplest-
case scenario in which a single plate intervened the Pacific and Australian plate in the Junction
Region, bounded on either side by trenches. This plate includes the Philippine Sea Plate, the
Papuan ophiolites and the Molucca Sea lithosphere that was reconstructed between the Papuan
ophiolites and the southern Philippines and Halmahera at 15 Ma, the eastern continuations of
the Papuan ophiolites in the form of the Louisiade and New Caledonia ophiolites, and the
eastern Philippine Sea Plate arcs (Izu-Bonin-Mariana) and Melanesian arcs (Solomon,
Vitiaz/New Hebrides, Tonga), as well as intervening oceanic basins (Solomon Sea and North
Loyalty Basin) (Figure 13). We restored the available marine magnetic anomaly evidence for
opening of oceanic basins, but there is no direct field evidence for a plate boundary that cut
through this composite series of arcs and basins during this time interval.
First, we reconstructed where the system of two opposing subduction zones must have
terminated. There is positive evidence for subduction below the Solomon Sea ophiolites since
45 Ma (Neef and McDougall, 1976, Richards et al., 1966) and the oldest dated arc rocks on Fiji
are 39 Ma (Rickard and Williams, 2013), showing that the Pacific trench must have continued
beyond Fiji southwards, to the Tonga Trench. However, farther south, subduction at the
Hikurangi trench did not start before c. 28-30 Ma (Van de Lagemaat et al., 2022), and also at the
Alpine Fault there is no evidence for significant pre-30 Ma dextral motion (Kamp, 1986; Furlong
and Kamp, 2013). Instead, at the longitude of the Hikurangi trench, relative plate motion
between the Australian and Pacific plates was only accommodated through eastward
subduction of the Australian Plate before c. 30 Ma, at the subduction zone below the New
Caledonia-Northland ophiolites (Van de Lagemaat et al., 2018a). We thus infer that the
Solomon-Fiji-Tonga trench ended against a transform fault to the south of the North Loyalty
Basin, which we tentatively connect to the Cook transform fault that bounded the Norfolk Basin
during its opening (Figures 17 - 19); Sdrolias et al., 2004a; Van de Lagemaat et al., 2018a),
which is still present in the Tasman Sea, to the north of New Zealand (Figure 4)
Our reconstruction thus infers that to the south of the Cook Fault, there was only one
subduction zone (the Three Kings Ridge trench segment of the New Caledonia subduction zone)
before 30 Ma, and that the oceanic lithosphere that obducted onto Zealandia, including the
Northland Ophiolite, moved as part of the Pacific Plate. The position of this plate boundary
therefore follows directly from Pacific-Australia motion before the moment of obduction (which
we model at 30 Ma following Van de Lagemaat et al., 2018a). We reconstruct the Solomon-Vitiaz
arc at 45 Ma to a position that lines up the Papuan Peninsula-Louisiade trench with the New
Caledonia trench (Figure 19). This position of the Melanesian arc requires that we restore some
clockwise rotation of the Tonga trench segment to avoid overlap with the New Caledonia trench
at 45 Ma.
In the northwest of the Junction Region, our reconstruction ensures that there is
convergence and thus subduction of both the Pacific and Australian plates with the Philippine
Sea Plate and Melanesian arcs. The reconstruction positions the entire Philippine-Melanesian
plate system from the Tonga-New Caledonia region in the southeast to the northwest
Philippines in the northwest such that there is no major convergence between the Philippine
Sea Plate and the SE Asian Tethysides after ~40 Ma because there is little geological evidence
for significant subduction on the northern Philippines after this time in this region. This
reconstruction is in correspondence with paleomagnetic data from the Philippine Sea Plate
(Figure 8), and allows for the opening of the Solomon Sea basin, as dictated by the preserved
magnetic anomalies (Gaina and Müller, 2007) between the two major bounding subduction
zones (Figures 16 - 19).
In our reconstruction, the closure of the Molucca Sea reconstructs the Papuan ophiolites
and the arc rocks of northern New Guinea adjacent to the Cretaceous crust and arc remnants of
the southern Philippine Mobile Belt (Figure 19). The reconstruction of the Molucca Sea in this
way solves and explains several issues: First, it reconstructs sufficient oceanic crust north of the
Australian continental margin for a subduction zone to have existed from 45-30 Ma. Second, this
scenario explains why there are no late Eocene arc volcanics preserved in the southern
Philippines, while they are present in the north: this southern continuation of the arc rifted off
during opening of the Molucca Sea and remained active but is now found on New Guinea. Third,
the reconstruction of the Molucca Sea in this way provides an explanation for the origin of the
Cyclops Ophiolite of New Guinea, from which ages between 43 and 30 Ma were obtained
(Monnier et al., 1999). The reconstruction also explains the abrupt termination of the Papuan
ophiolites just east of the Bird’s Head as a transform (STEP) fault (Figures 16 - 19) and infers
that Miocene subduction initiation of the Halmahera subduction zone occurred on the
northeastern margin of the Molucca Sea back-arc basin (Figure 13).
By 45 Ma, opening of the West Philippine Basin is largely restored (Hilde and Lee, 1984),
which leads to a much smaller Philippine Sea Plate and the N-S length of the Kyushu-Palau ridge
west of the Izu-Bonin Mariana-Palau subduction zone is also significantly reduced (Figure 19).
The extension of the Kyushu-Palau intra-oceanic arc towards the north at 45 Ma is unknown,
but we reconstruct a limited northwestward continuation of the Izu-Bonin Mariana-Palau
trench. We infer that the Pacific Plate was subducting below Japan at the Nankai and Ryukyu
trenches and that the Izu-Bonin Mariana trench ended in a triple junction with transform faults
with perhaps minor convergent components between the Philippine Sea Plate and the SE Asian
Tethysides to the south, and the Pacific Plate and the Proto-South China Sea embayment to the
north.
Another key difference between our reconstruction and the others is the location of
(south)westward subduction initiation at the Solomon-Vitiaz-Tonga trench. All previous models
reconstruct this subduction zone close to the continental margin at subduction initiation, and
subsequently reconstruct trench-roll back and the formation of a back-arc basin, i.e., the
Solomon Sea. Our reconstruction is the only one that reconstructs this subduction zone to start
intra-oceanic (Figure 19) and model the opening of the Solomon Sea in between two opposing
subduction zones. Wu et al. (2016), in addition to their subduction zone that forms along the
New Guinea continental margin, also reconstruct a west-dipping intra-oceanic subduction zone,
which is subsequently consumed when the subduction zone that hosts the Solomon Arc rolls
back.
5.4. 45 – 62 Ma
The logical next step in our reconstruction would be 52 Ma, which corresponds to the age of the
oldest supra-subduction oceanic lithosphere of the Izu-Bonin-Mariana forearc and is the widely
inferred age for the onset of subduction at the Izu-Bonin Mariana trench (Ishizuka et al., 2011a;
Stern et al., 2012; Reagan et al., 2013, 2019). However, as geological data from the Solomon
Islands, Vanuatu, Fiji, and Tonga suggest that southwestward subduction of the Pacific Plate
beneath these arcs did not start before c. 45 Ma (Neef and McDougall, 1976; Richards et al.,
1966; Bloomer et al., 1995; Crawford et al., 2003), it is not straightforward where the Izu-Bonin
Mariana plate boundary continued, or where and why it would have ended. Therefore, we chose
as next step in our reconstruction 62 Ma. North- and eastward subduction of the Australian
Plate at the New Caledonia-New Guinea trench started around 60 Ma as revealed by
metamorphic sole ages of the Papuan and New Caledonia ophiolites (Lus et al., 2004; Cluzel et
al., 2012b). We incorporate a 62 Ma subduction initiation age at the southwestern Philippine
Sea Plate/Melanesian (New Caledonia-New Guinea) subduction zone because at this time a
small change in relative plate motion between the Pacific and Australian plates occurred that
follows from the global plate circuit. We use the constraints from the Papuan and New
Caledonia ophiolites to restore relative motions at the Izu-Bonin Mariana trench from this
perspective (Figures 19 – 23).
As (south)westward subduction of the Pacific Plate south of the Philippine Sea Plate
(Halmahera) did not start before 45 Ma, we reconstruct the Solomon Islands, Vanuatu, and
Tonga, as well as intra-oceanic highs and basins east of the Louisiade-New Caledonia trench,
such as the Three Kings Ridge (Figures 19 - 23), as part of the Pacific Plate. The New Guinea-
New Caledonia subduction zone ended in a zone of diffuse and distributed deformation around
New Zealand, close to the Pacific-Australia Euler pole, and around 40 Ma connected with the
Emerald Basin spreading center (Van de Lagemaat et al., 2018a),
To the north of the Solomon Islands, the Philippine Mobile Belt formed the overriding
plate to the New Guinea-New Caledonia subduction, which was not part of the Pacific Plate
before 45 Ma, as the Izu-Bonin Mariana subduction zone accommodated subduction of the
Pacific Plate below the Philippine Sea Plate. There is no evidence for a discrete plate boundary
between the Philippine Mobile Belt (south of Halmahera) and the Solomon arc that may have
accommodated the inevitable relative motion between the two. We therefore reconstruct that
between 62 and 45 Ma, the Philippines acted as a major relay ramp between two subduction
systems that laterally decreased convergence in opposite direction. The Izu-Bonin Mariana
subduction zone accommodated all Pacific-Australia convergence at the position of the northern
Philippines. The amount of convergence accommodated by this subduction zone decreased
southwards towards Halmahera, and to the southeast, all Australian-Pacific convergence was
accommodated by the New Caledonia-New Guinea subduction zone. We accommodate this
evolution by reconstructing a clockwise rotation of the Philippine Mobile Belt relative to the
Solomon Islands between 62 and 52 Ma that avoids relative convergence between southern
Halmahera and the Pacific Plate. Our restoration of clockwise rotation of the Philippines also
ensures that the amount of convergence accommodated by the New Caledonia-New Guinea
subduction zone decreased northwards and transitioned into a transform fault with both the
Australian Plate and farther north with the SE Asian Tethysides (Figures 20 - 22). This explains
the absence of evidence of subduction beyond the Papuan ophiolites. After 52 Ma, subduction of
the Pacific Plate at the Izu-Bonin-Mariana subduction zone is sustained through opening of the
West Philippine and Palau Basins (Hilde and Lee, 1984; Sasaki et al., 2014). Our reconstruction
thus implies that the Izu-Bonin Mariana subduction zone was already forming by 62 Ma, at a
time that the Philippine Sea Plate only consisted of Cretaceous and latest Jurassic oceanic crust
and arc rocks of the small Huatung Basin (Dechamps et al., 2000) and the Philippine Mobile Belt
(e.g., Geary et al., 1988; Suerte et al., 2005; Dimalanta et al., 2020). We will return to this
element of our reconstruction in the discussion section.
5.5. 62 – 85 Ma
Whereas there are records of subduction that accommodated the convergence of the Pacific
plate system and the Tethyan plate system after 62 Ma and before 85 Ma (e.g., East Gondwana
subduction in New Zealand and New Caledonia (Van de Lagemaat et al., 2023a), and the arcs of
the Philippines and the circum-Proto-South China Sea embayment (section 3; see also Van de
Lagemaat et al., 2023c)), evidence for subduction in the Junction Region in the intervening
period is sparse. The next step in our reconstruction is therefore 85 Ma.
The global plate circuit straightforwardly explains the tectonic quiescence in the 62-85
Ma interval (Figures 24 and 25). This time interval is bracketed between two major plate
motion changes, around 85 Ma and ~50 Ma (as explained, the 62-50 Ma interval accommodated
only minor convergence), and relative motions between the major plates were limited. Australia
moved only ~600 km to the NE relative to Eurasia. Relative motion between the Pacific and
Australian plates was slow, and divergent in the south, accommodated at the Tasman Ridge
(Figures 24 - 26). The northward decrease in extension at the Tasman Ridge indicates that the
Euler Pole of Australia-Pacific motion was located around the northeast corner of the Australian
continental margin, around the Coral Sea. To the north, the Pacific Plate moved in the 85-62 Ma
time interval some ~700 km to the northwest. Prior to the initiation of subduction of the
Australian Plate below the Philippine Sea Plate at the New Caledonia-New Guinea trench at 62
Ma, we reconstruct the Philippine Mobile Belt as part of the Australian Plate (Figures 24 – 26).
Because this reconstruction predicts a NW-SE trend of the Philippine Mobile Belt at 62 Ma, the
boundary between the Philippine Mobile Belt and the Pacific Plate in the 62-85 Ma interval was
predominantly a transform fault with perhaps some obliquely subduction of the Pacific Plate
below the Australian Plate. This oblique subduction may explain the sparse Campanian
accretionary prism rocks identified on Halmahera (Hall et al., 1988) and the 77 Ma diorite
intrusion into the Rapu-Rapu ophiolite in the Philippines (David et al., 1997).
The model of Zahirovic et al. (2014) is the only previous regional reconstruction that
extends this far back in time. In our model, the Philippine Mobile Belt still acts as a rigid plate,
now part of the Australian Plate, whereas the model of Zahirovic et al. (2014) connects the
northernmost Cretaceous arc of the Philippine Sea Plate (the Amami Plateau) to the Pacific Plate
at 65 Ma and reconstruct extension between the Cretaceous arcs between 65 and 85 Ma, which
we restored during the Eocene based on the age constraints in the intervening basins (Hickey-
Vargas, 1998; Ishizuka et al., 2013, 2018). The reconstruction of Zahirovic et al. (2014) suggests
that subduction of the Australian Plate below the Papuan ophiolites already started at 85 Ma,
while the ophiolites of New Guinea and New Caledonia suggest a c. 60 Ma onset of subduction.
Prior to 85 Ma, the western Panthalassa Ocean was surrounded by subduction zones, from
Antarctica to Japan (Figures 26 – 33). Along the east Gondwana margins of New Zealand and
New Caledonia, accretionary prisms formed during subduction of the Phoenix Plate and its
daughter plates that broke up during emplacement of the Ontong-Java-Nui large igneous
province (Van de Lagemaat et al., 2023a, and references therein). Some marginal basin
development may have affected the New Zealand margin in the Triassic-Jurassic (Howell, 1980;
Roser et al., 2002; Van de Lagemaat et al., 2018b), but this is not of consequence for the
reconstruction of the Junction Region. For the Japan margin, major back-arc basin opening, and
closure has recently been postulated for Jurassic and Cretaceous time based on combining
marine magnetic anomaly reconstructions of the northwestern Pacific Plate with analyses of
OPS sequences (Boschman et al., 2021a). However, subduction and arc magmatism along the
eastern Eurasian margins of Indochina and South China west of the Qingdao Line was
continuous from the Jurassic until ~85 Ma and upper plate deformation remained restricted to
Basin & Range-style extension, and prior to that Andean-style orogenesis in the South China
Block (Jahn et al., 1990; Lapierre et al., 1997; Li et al., 2012; J. Li et al., 2014). Finally, to the west,
the SE Asian Tethysides experienced a northward journey from the northwest Australian
margin to the Indochina margin from middle Jurassic to ~85 Ma, accommodated by the closure
and opening of Neotethyan oceanic basins (Hall, 2012; Advokaat and Van Hinsbergen, 2023).
The plate reconstruction of the Junction Region thus hinges on interpretations of the geology
from the Philippines and New Guinea, and the remains of the Proto-South China Sea lithosphere.
We consider the Philippine Mobile Belt as the arc and forearc crust that formed above a
southwest-dipping subduction zone that consumed oceanic crust of the Panthalassa Ocean
(Figures 26 – 33). We refer to this ‘proto-Izu-Bonin Mariana’ subduction zone as the Lagonoy
subduction zone, after the oldest ophiolite exposed in the Philippines. In Permian to Triassic
time this subduction zone was located along the north Australian margin (Figure 33), but in
Late Jurassic time it started to roll back, opening a forearc basin that developed into a back-arc
basin, with some trench-parallel convergence accommodated at the SE Bohol trench (Figures 28
– 33). After 110 Ma, extension jumped back to the forearc. This may either have occurred
because of trench-parallel extension above an oblique subduction zone, such as documented for
the Andaman Sea (Curray, 2005) or the Jurassic Californian ophiolites (Arkula et al., 2023), or
trench-normal, analogous to the ridge jump from the Parece Vela Basin to the Mariana Trough,
or from the South Fiji Basin to the Lau Basin. The Upper Cretaceous crust of the eastern
Philippine Mobile Belt formed shortly after the jump in extension and was subsequently
overlain by arcs. Trench-parallel extension may have played a role at 135 Ma as suggested by
the E-W trending magnetic anomalies of the Huatung Basin (Deschamps et al., 2000), although
this basin may also have opened as a back-arc basin above the proto-SE Bohol Trench. In our
reconstruction, we model a ‘Mariana Trough style’ ridge jump around 110 Ma, inferring trench-
normal extension between 110 and 85 Ma for the Philippine Mobile Belt (Figures 26 – 28),
which may be tested in the future with paleomagnetic data from sheeted dyke sequences (see
e.g., Maffione et al., 2017). This extension and roll-back ceased when the Pacific changed plate
motion at 85 Ma, and the margin became transform dominated (Figure 25).
For times back to 156 Ma, we reconstruct the motion of the Pontus Plate as such that we
avoid extension with the Izanagi Plate. The Telkhinia trench that accommodated subduction of
the Izanagi Plate restores farther into the Panthalassa Ocean back in time, and as a result, its
north-south extent becomes longer in older reconstruction slices (Figures 26 - 33). The triple
junction at e.g., 135 Ma was located farther southeastwards along the East Gondwana margin,
and the Telkhinia subduction zone likely also accommodated the subduction of the Izanagi-
Phoenix ridge, or even part of the Pacific Plate (Figures 28 - 33).
The age of the oldest lithosphere in the Pontus Plate, and hence the maximum age of the
Telkhinia subduction zone, is unknown. Accreted units below the eastern margin of the Pontus
Plate have likely all been lost to subduction. The accreted seamount in the South China margin
with an age of ~154 Ma (Xu et al., 2022) show that the plate contained lithosphere of at least
Late Jurassic age. For ages older than 135 Ma, we have no direct kinematic constraints left from
rocks of the Pontus Plate, and it is possible that the Pontus ‘Ocean’ sensu Van der Meer et al.
(2012) contained multiple plates, separated by ridges. In our reconstruction prior to 135 Ma,
we keep the Telkhinia subduction zone more or less mantle stationary, following the
tomography-based interpretations of the lower mantle below the Pacific Ocean of Van der Meer
et al. (2012).
The Pontus Plate (or further back in time, probably plate system) thus formed a Junction
Region in between the Panthalassa and Tethys domains. The westernmost plate boundary of the
Panthalassa Ocean (or the ‘Thalassa Ocean’ sensu Van der Meer et al., 2012) was formed by the
Telkhinia subduction zone, that connected the subduction zone below Japan with the
subduction zone below East Gondwana. To the west, the subduction zones along the Philippine
Mobile Belt (Lagonoy subduction zone) and the Kuching Zone adjacent to the SW Borneo Block
must have connected through an intra-oceanic trench system to the trenches along Sundaland
and South China. Such trenches were also inferred for Jurassic and older times based on seismic
tomographic images of the lowermost mantle by Van der Meer et al. (2012), but geological relics
of these trenches have not been identified yet. Such relics would have been consumed by
subduction below the SW Borneo Block in the south, the West Sulawesi and Sibumasu terranes
of Sundaland, and depending on how far this subduction zone reached westwards, perhaps even
in de Bangong-Nujiang suture zone between the Lhasa and Qiangtang terranes of western
Burma. Future analysis of accretionary prisms in SE Asia may identify further constraints on the
intra-oceanic subduction systems that must have existed between the Pontus and Neotethys
oceans. The reconstruction at 160 Ma (Figure 33) satisfies all presently available geological
constraints on intra-oceanic subduction that we are aware of from the Junction Region.
The key difference with the reconstruction of Zahirovic et al. (2014) in the oldest part of
our reconstruction is the location of the plate boundary that separates the Panthalassa plates
from the Junction. Where we follow the constraints of Wu et al. (2022) and Van de Lagemaat et
al. (2023c) to infer a subduction zone that extends from the East Gondwana northwards to the
Qingdao Line, and the subduction of the Pontus Plate below the SE China and Indochina
margins, Zahirovic et al. (2014) incorporate the widely held view that the Izanagi Plate
subducted below the entire Eurasian margin. Moreover, their model opens the Proto-South
China Sea as a back-arc basin between c. 65 and 35 Ma, while geological data from surrounding
regions indicate that the oceanic lithosphere of this basin was of Cretaceous age (Dycoco et al.,
2021; Xu et al., 2022; Wang et al., 2023; Van de Lagemaat et al., 2023c).
Our kinematic reconstruction is based on the modern geological architecture of the Junction
Region, in which we assumed a simplest plate model with the least amount of plate boundaries
necessary to explain the present-day geology. In this section we briefly discuss the geodynamic
implications that follow from our reconstruction, and how they may contribute to a better
understanding of geodynamics.
6.1. Small oceanic basins opening in the downgoing plate close to trenches
An interesting finding of our plate reconstruction is that the Junction Region hosts several
examples of relatively small oceanic basins that opened within a down-going plate, in the
proximity of and at ridges parallel to a subduction zone. Opening of oceanic basins above
subduction zones is a common phenomenon, as illustrated by the numerous forearc and back-
arc basins of the West and Southwest Pacific realms, such as the Mariana Trough, Ogasawara
Trough, Parece-Vela Basin, South Fiji Basin, or Lau Basin (Figure 1). Examples of oceanic basins
that opened in a down-going plate are the South China Sea basin, the Santa Cruz Basin, and the
Woodlark Basin (Figure 34). We speculate that there may be causal relationships between
opening of these basins and the nearby subduction zones. The South China Sea basin opened
from the late Eocene to mid-Miocene, during southward subduction of the Proto-South China
Sea, partly contemporaneous with the Sulu Sea that started opening as back-arc basin in the
upper plate (e.g., Advokaat and Van Hinsbergen, 2023). It was recently suggested that the Proto-
South China Sea may have been underlain by an oceanic plateau that initially stopped
subduction in the Mesozoic (Van de Lagemaat et al., 2023c). Those authors tentatively
suggested that when this oceanic plateau was eventually forced to subduct during
counterclockwise vertical-axis rotation of Borneo, the eclogitization of thickened oceanic crust
may have caused an especially strong slab pull, that resulted in the extension and eventual
break-up of the down-going plate and the formation of the South China Sea.
The Santa Cruz Basin formed in the Eocene during eastward subduction at the New
Caledonia subduction zone, while the North Loyalty back-arc basin opened contemporaneously
and in a parallel orientation in the overriding plate (Seton et al., 2016; Van de Lagemaat et al.,
2018a). The eastern margin of the Santa Cruz Basin is flanked by the West Torres Plateau. In the
case of the Santa Cruz Basin, however, extension in the down-going plate ceased in the early
Oligocene, while subduction of the plateau is presently occurring at the New Hebrides trench.
Moreover, it is unknown whether the West Torres Plateau was being subducted during opening
of the Santa Cruz Basin, because its original eastward extension is unknown.
The Woodlark Basin also recently started forming in the down-going Australian Plate,
but at a higher angle to the subduction zone owing to rotational opening around an Euler pole
not far west of the d’Entrecasteaux Islands. In this basin, there is no evidence of the presence of
an oceanic plateau, but it is attached to older, Eocene, oceanic crust of the Solomon Sea. Similar
to the tectonic settings of the South China Sea and Santa Cruz basins, there is also a back-arc
basin in the northwest (the Manus Basin) that forms in an overriding plate position to the
Woodlark Basin. However, in the case of the South China Sea and Santa Cruz Basin, their
opening was preceded a few Ma by the opening of the back-arc basin, while in the case of the
Woodlark Basin, the back-arc basin is younger than the subducting plate basin. Moreover,
where the back-arc and subducting plate basins opened roughly parallel in the case of the South
China Sea and Santa Cruz basins, the orientation of the Manus Basin is at a higher angle with the
Woodlark Basin.
Other than the fact that all these basins formed in a down-going plate in the proximity
and roughly parallel to a subduction zone, the most striking resemblance between the South
China Sea and Santa Cruz basins is that both formed parallel to a back-arc basin in the
overriding plate that started forming a few Ma earlier, but this similarity is not shared with the
Woodlark Basin. The basins that opened in subducting plates all formed in pre-existing weak
zones of that plate; the South China Sea formed within the South China margin accretionary
prism, and the Santa Cruz and Woodlark basis formed at an obducted continent-ocean
transition zone of the Australian Plate.
The consequence of back-arc basins that form during slab roll-back is that subduction
rates exceed the convergence rates between the plates that were originally interacting at the
plate boundary, an effect that is in these cases caused or enhanced by the formation of
subducting plate basins. The reason behind this effect related to the formation of subducting-
plate basins and its connection to subduction zone dynamics are topics that may be explored
further through geodynamic modeling.
Finally, the Coral Sea and Caroline Plate basins also opened in a down-going plate
position, and roughly parallel to the San Cristobal North Solomon subduction zone, respectively,
but at a far larger distance and at a larger scale than the aforementioned basins. A causal
relationship between the opening of the basin and the subduction zone may therefore be less
straightforward. We consider the Coral Sea Basin as a northward extension of the Tasman Sea,
which formed as a marginal basin during continental break-up. Previously, the Caroline Plate
was considered a back-arc basin (e.g., Weissel and Anderson, 1978; Gaina and Müller, 2007; Wu
et al., 2016), but in our reconstruction the plate is separated from the north-dipping New
Guinea subduction zone by the south-dipping North Solomon-Vitiaz subduction zone and the
basin is therefore not in a back-arc position. As the large igneous provinces on the Caroline
Plate clearly indicate its interaction with a mantle plume, we suspect that the formation of
Caroline Plate basins may be related to weakening by a mantle plume, with spreading driven by
slab pull of the already subducting slab.
Placing our reconstruction in a mantle reference frame provides insight into the absolute
motions of plates and subducted slabs (known as ‘slab dragging’; Spakman et al., 2018), during
subduction and after slab break-off. Van de Lagemaat et al. (2018a) showed that the Tonga-
Kermadec slab was dragged northward, laterally through the mantle over about 1200 km since
30 Ma, resulting from the northward, trench-parallel component of the absolute motion of
Pacific Plate. Our reconstruction now assigns an earlier age of subduction initiation of the Tonga
segment of the subduction zone of 45 Ma, which increases the amount of Tonga slab dragging
since subduction initiation to 1600 km (~3.5 cm/yr). The accompanying northward motion of
the Australian Plate resulted in also northward motion of the Tonga-Kermadec trench and
upper plate (Figure 35). Slab dragging is not restricted to trenches that accommodate
subduction between two plates that also share a trench-parallel absolute plate motion
component. Slab dragging also occurs during highly oblique subduction below near-mantle-
stationary upper plates, e.g., the northward dragging of the Burma slab as part of the Indian
Plate (Le Dain et al., 1984; Parsons et al., 2021). In addition to trench-parallel slab dragging, slab
dragging also occurs during trench-perpendicular absolute motion; the motion of upper plates
(and hence trenches) relative to the mantle may result in slab retreat or slab advance, which
causes flat-lying and steep (or even overturned) slabs, respectively (e.g., Qayyum et al., 2022).
We here assess the motion of trenches and slabs relative to the mantle as the result of absolute
plate motion, and how this may have influenced slab geometry, by placing our reconstruction of
the Junction Region in the global moving hotspot reference frame of Doubrovine et al. (2012).
The rapid Australian absolute plate motion must mean that the New Hebrides slab,
which has been subducting and rolling back below Vanuatu, has experienced northward slab
dragging since its formation (c. 10 Ma ago, Yan and Kroenke, 1993). The clockwise rotation of
the trench has resulted in a larger amount of absolute trench motion in the south than in the
north, and a larger northward component of the northern part of the trench. This slab dragging
component may provide a straightforward explanation for the incipient subduction at the
Hunter fracture zone that started in the Plio-Pleistocene (Patriat et al., 2015; 2019; Lallemand
and Arcay, 2021). This transform formed as a STEP fault accommodating New Hebrides trench
roll-back. We foresee that the resistance of the New Hebrides slab against northward slab
dragging, which is known to cause trench-parallel shortening in the downgoing plate (Spakman
et al., 2018), may contribute to the incipient subduction at the Hunter fracture zone.
The Tonga-Kermadec trench was connected northwards with the Solomon-Vitiaz trench,
which also accommodated Pacific-Australia relative motion. The motion of the Pacific Plate
relative to the mantle since 45 Ma is roughly towards the WNW, similar to the orientation of the
North Solomon trench. The subduction of the Pacific Plate at the North Solomon trench is thus
purely the result of the 2200 km (~4.9 cm/yr) northward motion of the North Solomon trench
since 45 Ma, as part of the Australian Plate (Figure 35). The northward motion of the North
Solomon trench leads to slab retreat.
Farther north, the Izu-Bonin Mariana subduction zone also has a N-S orientation. The
Philippine Sea Plate underwent a strong northward motion component (see section 5; also Van
de Lagemaat et al., 2023b). Like the Tonga-Kermadec slab, the Izu-Bonin Mariana slab is located
due west of the present-day trench (Miller et al., 2004; Wu et al., 2016; Van der Meer et al.,
2018), but the subduction zone originated much farther southward (Figure 35). Slab dragging at
the Mariana trench was c. 3000 km (~6.6 cm/yr), i.e., almost twice as fast as the Tonga-
Kermadec subduction zone since 45 Ma. We suggest that the N-S opening of the West Philippine
Basin and the NE-SW opening of the Parece Vela Basin accommodated additional slab dragging
of the Izu-Bonin section driven by absolute motion of the Pacific Plate. Moreover, as subduction
here initiated earlier than in the Tonga-Kermadec case, the total amount of slab dragging is
much larger, in the order of ~4000 km for the northernmost Izu-Bonin trench since 52 Ma
(~7.7 cm/yr) and ~3000 km for the central Mariana trench (~5.8 cm/yr). The recent clockwise
rotation of the Philippine Sea Plate accommodated by extensional motion in the Ayu Trough
since 15 M, has resulted in slab advance of the southern portion of the trench, which explains
the steep nature of the Mariana slab (e.g., Spakman et al., 1989; Van der Hilst, 1991; Miller et al.,
2004). This also suggests that the Mariana Trough back-arc did not open because of slab roll-
back, but rather due to slab resistance against overriding plate retreat.
In the north of the Philippine Sea Plate, the position of the trenches along the Eurasian
margin have been comparatively stable relative to the mantle since the Late Cretaceous. The
total amount of trench retreat of the Ryukyu arc since 70 Ma is about 750 km (~1.1 cm/yr).
However, the opening of the Shikoku Basin in the Philippine Sea Plate, which was accompanied
by trench retreat of the Izu-Bonin trench, has resulted in large-scale (~1600 km since 30 Ma)
trench-parallel slab dragging at the Nankai trench, analogous to the Burma slab.
The Manila and the Halmahera trenches that form the western plate boundary of the
Philippine Sea Plate have experienced large absolute trench motions since subduction initiation
around 15 Ma. In this case, however, the position of the down-going plate relative to the mantle
has been almost stationary throughout the subduction history. So even though the Manila
trench’s upper plate underwent large scale northward, trench-parallel, motion, lateral slab
dragging is minimal here.
In addition to the presently active subduction zones, we may also analyze former
trenches in our reconstruction and make inferences about where we may find slab remnants in
the mantle. The north- and eastward dipping New Guinea-New Caledonia subduction zones
have been subject to slab dragging (Figure 35). Like most trenches in the Junction Region, these
trenches and slabs underwent northward absolute motion, which resulted in trench-parallel
dragging at the New Caledonia trench. Van de Lagemaat et al. (2018a) showed that the South
Loyalty Slab, identified by Schellart et al. (2009), is in a position that is just north of the
reconstructed 30 Ma location of the New Caledonia trench, at which time northward subduction
ceased and the slab likely detached. After the end of subduction, the slab sank vertically into the
mantle, and the inactive trench subsequently overrode the slab during northward motion of the
Australian Plate. The change in orientation from roughly N-S at the New Caledonia trench to
roughly E-W at the New Guinea trench resulted in slab advance rather than lateral dragging of
the slab at the New Guinea trench (Figure 35). During the Eocene, however, slab advance in the
New Guinea sector was largely cancelled by roll-back such that the trench remained in a more
or less mantle-stationary position. The opening of the Molucca Sea and Solomon Sea back-arc
basins occurred mostly due to northward absolute motion components of the Philippine Sea
Plate relative to this mantle-stationary trench. Subsequent slab advance during northward
motion of the New Guinea trench may have resulted in a steeply dipping or overturned slab.
Depending on how the geometry of the slab was affected by slab advance, both the Arafura and
Carpentaria slabs of Van der Meer et al. (2018) may be correlated to the New Guinea subduction
zone. Following slab break-off, Australia overrode these slabs, and their mantle wedges.
Previously, Van Hinsbergen et al. (2020b) inferred that Australia overriding a former mantle
wedge above the Arafura slab led to the enigmatic arc-signature in the Pliocene Ertsberg and
Grasberg magmatic centers in Central New Guinea, close to the obduction front. Those authors
used the reconstructions of Hall (2002) and Zahirovic et al. (2014) who suggested that the
Arafura slab formed entirely at an intra-oceanic subduction zone far north of Australia, prior to
~30 Ma. Our reconstruction instead suggests that the only candidate to have formed the Arafura
slab is the subduction zone that led to the obduction of the Papuan ophiolite. The timing of its
break-off may have occurred anytime after 30 Ma, but the position of the slab suggests that this
occurred around 10 Ma, and that the time delay between mantle wedge formation and plowing
of the Australian margin through the mantle is shorter than interpreted by Van Hinsbergen et al.
(2020b).
In the Cretaceous, the Lagonoy subduction zone underwent trench retreat, as well as
northward trench-parallel slab dragging. Slab dragging at the Telkhinia trench was dominated
by slab advance. During roll-back of the Lagonoy subduction zone, the Telkhinia subduction
zone was progressively consumed by the Lagonoy subduction zone, and we therefore expect
that the Telkhinia slab may have broken in several instances. This may explain some of the slab
remnants that are present in the mantle below the eastern margin of eastern Australia. One of
the slab remnants below Australia is the Lake Eyre slab, which is in the lower mantle below
Lake Eyre, Australia. This slab was previously interpreted using the reconstruction of Hall
(2002) as the detached slab that formed during northward subduction of the Australian Plate
below the Papuan ophiolites (Schellart and Spakman, 2015). In their interpretation subduction
there ceased around 50 Ma, based on the c. 54 Ma age of the metamorphic sole below the
Papuan Ultramafic Belt (Lus et al., 2004). However, as the metamorphic sole forms during
subduction initiation rather than subduction termination, this age does not represent the end of
subduction but the start, and the geology of New Guinea indicates that oceanic subduction
continued into the Oligocene and obduction until the late Miocene. In light of our
reconstruction, it is unlikely that the Lake Eyre slab is correlated to a 60-<30 Ma New Guinea
subduction zone. Instead, when putting our reconstruction in de slab references frame of Van
der Meer et al. (2010), the Lake Eyre slab is in a location that could correspond to the location of
the Lagonoy subduction zone around 85 Ma, just prior to when this plate boundary became
dominated by transform motion owing to the change in absolute motion of the Pacific plate
from west to north (Figure 36). The change in relative plate motion at the plate boundary may
have resulted in slab detachment which subsequently sank vertically into the mantle. This
interpretation would suggest that the Lake Eyre slab is much older than previously thought.
Our analysis of slab dragging in the west Pacific region shows that lateral slab dragging
such as previously shown for the Tonga-Kermadec (Van de Lagemaat et al., 2018a) is a rule
rather than an exception. Lateral slab dragging is a common phenomenon that occurs in all
subduction zones where the down-going plate has a trench-parallel component. In addition, the
absolute motion of upper plates relative to slabs is a key driver of upper plate deformation (Van
Hinsbergen and Schouten, 2021). Slab dragging can only be correctly analyzed in an absolute
plate motion frame, because the absolute plate motion of both the upper plate and down-going
plate as well as their interaction at the trench determine the amount and type of slab dragging.
In addition, as illustrated by the example of the Lake Eyre slab, our reconstruction sheds a new
light on the plate boundary evolution in the Junction Region that requires a reinterpretation of
the geological history of upper and especially lower mantle slabs, such as those listed in the
Atlas of the Underworld (Van der Meer et al., 2018).
The Izu-Bonin Mariana subduction zone is one of the best-studied subduction zones on Earth,
particularly to understand subduction initiation (e.g., Stern and Bloomer, 1992; Stern et al.,
2003, 2012; Reagan et al., 2010; Ishizuka et al., 2011a; Arculus et al., 2015). The age of
subduction initiation at the Izu-Bonin Mariana trench is generally inferred from the oldest
supra-subduction zone gabbros and basalts recovered from the Izu-Bonin-Mariana forearc,
which are ~52 Ma old (Ishizuka et al., 2011a; Reagan et al., 2010, 2013, 2019). This inference is
based on the assumption that the catastrophic extension that led to the overriding plate
spreading centers at which these basalts and gabbros formed along the strike of the Izu-Bonin-
Mariana forearc signals spontaneous subduction initiation (Stern and Bloomer, 1992; Stern et
al., 2012). In such a setting, the area consumed by initial subduction resulting from the
gravitationally driven lithospheric collapse of the down-going plate must be instantaneously
compensated by upper plate extension, and the magmatic rocks formed during that extension
must thus be synchronous with subduction initiation (e.g., Stern and Bloomer, 1992; Stern,
2004; Arculus et al., 2015). However, the Semail Ophiolite of Oman comprises similar forearc
crust that formed through catastrophic extension because of lithospheric collapse, but there,
rocks that formed at the incipient subduction plate boundary are exhumed as metamorphic sole
rocks, which are about 8 Ma older than the age of forearc basalts (Guilmette et al., 2018). Similar
and even larger time delays between the formation of the metamorphic sole and the formation
of forearc crust have been shown since for the Coast Range Ophiolite in California (Mulcahy et
al., 2018), the Halilbağı Complex in Turkey (Pourteau et al., 2019) and the Xigaze Ophiolite in
Tibet (Guilmette et al., 2023). These time lags between metamorphic sole and subsequent
forearc crust formation can only exist in subduction zones that formed due to far-field-forced
convergence, as the subduction interface formed before lithospheric collapse occurred
(Guilmette et al., 2018). The metamorphic sole of the Izu-Bonin Mariana subduction zone
remains buried below the Izu-Bonin Mariana forearc and it is therefore impossible to determine
its age that would directly resolve the debate of whether subduction initiation occurred
spontaneously or was forced. In this case, a kinematic reconstruction provides the next best
insight into the question of whether subduction initiation at the Izu-Bonin Mariana trench must
have been spontaneous or may have been induced.
Based on our reconstruction, we find that there is no need to assume that subduction
initiation at the Izu-Bonin Mariana trench must have been spontaneous. Instead, our
reconstruction shows that it is more likely that subduction initiation at the Izu-Bonin Mariana
trench was forced and occurred around 62 Ma linked to the slow Pacific-Australian plate
convergence that also led to the formation of the New Guinea-New Caledonia subduction zone.
If correct, subduction initiation occurred about 10 Ma before lithospheric collapse and
extension in the Izu-Bonin Mariana forearc, a similar delay as in the Tethyan and Californian
examples. This is based on the reconstructed 62-52 Ma clockwise rotation of the Philippine
Mobile Belt that is required for the northward decrease of convergence to avoid subduction of
the Australian Plate beyond the Papuan ophiolite belt.
In addition to the cause of subduction initiation, also the nature of the Izu-Bonin
Mariana plate boundary before it became a subduction zone has been subject of debate. The
most common suggestion is that subduction initiated along a transform fault between two mid-
ocean ridge segments (e.g., Casey and Dewey, 1984; Dewey and Casey, 2011), whether it be
spontaneous or induced (e.g., Stern and Bloomer, 1992; Hall et al., 2003). This suggestion was
mostly based on the interpretation that extension and spreading in the West Philippine Basin
predated subduction initiation, but it did not explain the presence of oceanic crust in the
overriding Philippine Sea Plate that is up to 100 Ma older (the Lagonoy Ophiolite) at time of
subduction initiation. Our reconstruction shows that the Izu-Bonin Mariana trench initiated
along a pre-existing weakness zone indeed, which had accommodated mostly transform motion
for about 20 Ma resulting from relative plate motion changes around 85 Ma, but that was a
subduction zone before that time: our ‘Lagonoy’ subduction zone that was active in the
Mesozoic.
If the Izu-Bonin Mariana trench formed through forced subduction initiation that
resulted from a small change in Australia-Pacific relative plate motion at 62 Ma together with
the New Guinea-New Caledonia subduction zone, the 52 Ma lithospheric collapse may signal the
effective onset of slab pull, and thus may have been a trigger for the Eocene change in Pacific
Plate motion that resulted in the formation of the Hawaii-Emperor Bend (Torsvik et al., 2017).
Izu-Bonin Mariana subduction initiation as driver for the change in absolute plate motion of the
Pacific Plate was previously suggested based on geodynamic modelling (Faccenna et al., 2012).
Moreover, the possibility that the initiation of effective slab pull that is reflected in lithospheric
collapse following subduction initiation may cause absolute plate motion changes was tested
and shown for the forced subduction initiation event below the Oman and Anatolia ophiolites
that caused a change in African plate motion (Gürer et al., 2022). Our reconstruction shows that
it is feasible that the sequence of events from forced subduction initiation to lithospheric
collapse and a subsequent change in absolute plate motion applies to the Izu-Bonin Mariana
trench and Pacific Plate. To what extent the subduction initiation at the Izu-Bonin Mariana
trench was responsible for the change in absolute plate motion of the Pacific Plate is uncertain,
as the NW Pacific was the location of other drastic plate tectonic changes, including subduction
of the Izanagi-Pacific Ridge that led to the first demonstrable subduction of Pacific Plate
lithosphere (Seton et al., 2015; Wu and Wu, 2019) and a subduction polarity reversal after the
collision of the Olyutorsky arc with Kamchatka (Domeier et al., 2017; Vaes et al., 2019). Based
on the reconstructed 45 Ma age of the Solomon-Vitiaz-Tonga subduction zone, we find it most
likely that subduction initiation there was a result of the change in absolute plate motion change
of the Pacific Plate rather than a cause, but after it happened, it will have contributed to a more
westerly course of absolute Pacific Plate motion. However, if future age constraints provide
evidence for an older age of subduction initiation there, it may have played an active role in
changing the absolute motion of the Pacific Plate.
7. Conclusions
We developed a kinematic restoration of the Junction Region back to the Jurassic, based on the
present-day orogenic record of the circum-Philippine Sea Plate and Australasian region. We
presented a comprehensive review of the orogens of southwestern Japan, Taiwan, the
Philippines and northern Molucca Islands, New Guinea, and the Solomon Islands and adjacent
archipelagoes and ocean basins. Our reconstruction is based on a systematic restoration of the
tectonic motions that are inferred from these orogenic and marine geophysical records, cast in
context of relative motions of the Australian, Eurasian, and Pacific plates, and in context of a
recent restoration of orogenesis in the Tethyan belts of SE Asia. We present our reconstruction
in a series of maps from the present back to the Jurassic and provide GPlates reconstruction
files that can be placed in mantle or paleomagnetic reference frames as basis for geodynamic or
paleoclimatic analysis, respectively. Our conclusions are plentifold, but include the following:
1) The Molucca Sea as well as the Eocene ophiolites of the Philippines and the Cyclops
Ophiolite in New Guinea formed as part of an Eocene back-arc basin above a northward
dipping subduction zone that consumed the Australian Plate. Subduction at this trench
initiated around 62 Ma, was contiguous with the New Caledonia subduction zone, and
ended in the Oligocene.
2) The latest Jurassic oceanic crust that is preserved in the Philippines originated from the
northern margin of the Australian Plate where continental margin subduction was
active during the Permian and Triassic. The formation of the ophiolites records the
onset of oceanic spreading and the formation of an Early Cretaceous back-arc basin
behind a proto-Izu-Bonin Mariana subduction zone which we refer to as the Lagonoy
subduction zone. The Late Cretaceous ophiolites formed when extension relocated to
the forearc.
3) Trench-parallel slab dragging as well as slab retreat and advance are common features
of subduction zones. The previously identified Lake Eyre slab is not related to the New
Guinea subduction zone, but instead may be related to slab break-off at the Lagonoy
subduction zone around 85 Ma, when this plate boundary became a transform fault.
Instead, the Arafura or possibly the Carpentaria slabs may have formed at the New
Guinea subduction zone. A careful re-evaluation of modern mantle structure of the
western Pacific in light of our reconstruction is timely.
4) There is no necessity for spontaneous subduction initiation at the Izu-Bonin Mariana
trench. Our reconstruction predicts forced subduction initiation around 62 Ma, which
subsequently resulted in lithospheric collapse and the formation of the Izu-Bonin
Mariana forearc around 52 Ma. The lithospheric collapse may have been a trigger of the
change in absolute plate motion of the Pacific Plate that caused the formation of the
Hawaii-Emperor Bend.
Acknowledgements
SHAvdL and DJJvH were funded by NWO Vici grant 865.17.001 to DJJvH. We thank an
anonymous reviewer for valuable comments.
References
Advokaat, E. L., Marshall, N. T., Li, S., Spakman, W., Krijgsman, W., & van Hinsbergen, D. J. (2018).
Cenozoic rotation history of Borneo and Sundaland, SE Asia revealed by
paleomagnetism, seismic tomography, and kinematic reconstruction. Tectonics, 37(8),
2486-2512.
Advokaat. E. L., & Van Hinsbergen, D. J. J., (2023). Finding Argoland: reconstructing a lost
continent in SE Asia. Submitted to Gondwana Research.
Almasco, J. N., Rodolfo, K., Fuller, M., & Frost, G. (2000). Paleomagnetism of Palawan,
Philippines. Journal of Asian Earth Sciences, 18(3), 369-389.
Altis, S. (1999). Origin and tectonic evolution of the Caroline Ridge and the Sorol Trough,
western tropical Pacific, from admittance and a tectonic modeling
analysis. Tectonophysics, 313(3), 271-292.
Alvarez-Marron, J., Brown, D., Camanni, G., Wu, Y. M., & Kuo-Chen, H. (2014). Structural
complexities in a foreland thrust belt inherited from the shelf-slope transition: Insights
from the Alishan area of Taiwan. Tectonics, 33(7), 1322-1339.
Amaru, M. L. (2007). Global travel time tomography with 3-D reference models. Ph. D thesis,
Utrecht University, the Netherlands.
Andal, E. S., Arai, S., & Yumul Jr, G. P. (2005). Complete mantle section of a slow-spreading
ridge-derived ophiolite: An example from the Isabela ophiolite in the Philippines. Island
Arc, 14(3), 272-294.
Arai, R., Kodaira, S., Yuka, K., Takahashi, T., Miura, S., & Kaneda, Y. (2017). Crustal structure of
the southern Okinawa Trough: Symmetrical rifting, submarine volcano, and potential
mantle accretion in the continental back-arc basin. Journal of Geophysical Research: Solid
Earth, 122(1), 622-641.
Arcilla, C. A., Ruelo, H. B., & Umbal, J. (1989). The Angat ophiolite, Luzon, Philippines: lithology,
structure, and problems in age interpretation. Tectonophysics, 168(1-3), 127-135.
Arculus, R. J., Ishizuka, O., Bogus, K. A., Gurnis, M., Hickey-Vargas, R., Aljahdali, M. H., ... & Zhang,
Z. (2015). A record of spontaneous subduction initiation in the Izu–Bonin–Mariana
arc. Nature Geoscience, 8(9), 728-733.
Arkula, C., Lom, N., Wakabayashi, J., Rea-Downing, G., Qayyum, A., Dekkers, M. J., ... & van
Hinsbergen, D. J. (2023). The forearc ophiolites of California formed during trench-
parallel spreading: Kinematic reconstruction of the western USA Cordillera since the
Jurassic. Earth-Science Reviews, 237, 104275.
Aurelio, M. A. (2000). Shear partitioning in the Philippines: constraints from Philippine Fault
and global positioning system data. Island Arc, 9(4), 584-597.
Aurelio, M. A., Barrier, E., Rangin, C., & Müller, C. (1991). The Philippine Fault in the late
Cenozoic tectonic evolution of the Bondoc-Masbate-N. Leyte area, central
Philippines. Journal of Southeast Asian Earth Sciences, 6(3-4), 221-238.
Aurelio, M. A., Forbes, M. T., Taguibao, K. J. L., Savella, R. B., Bacud, J. A., Franke, D., ... & Carranza,
C. D. (2014). Middle to Late Cenozoic tectonic events in south and central Palawan
(Philippines) and their implications to the evolution of the south-eastern margin of
South China Sea: Evidence from onshore structural and offshore seismic data. Marine
and Petroleum Geology, 58, 658-673.
Bailly, V., Pubellier, M., Ringenbach, J. C., De Sigoyer, J., & Sapin, F. (2009). Deformation zone
‘jumps’ in a young convergent setting; the Lengguru fold-and-thrust belt, New Guinea
Island. Lithos, 113(1-2), 306-317.
Baker, S., & Malaihollo, J. (1996). Dating of Neogene igneous rocks in the Halmahera region: arc
initiation and development. Geological Society, London, Special Publications, 106(1), 499-
509.
Baldwin, S. L., Fitzgerald, P. G., & Webb, L. E. (2012). Tectonics of the New Guinea region. Annual
Review of Earth and Planetary Sciences, 40, 495-520.
Baldwin, S. L., Lister, G. S., Hill, E. J., Foster, D. A., & McDougall, I. (1993). Thermochronologic
constraints on the tectonic evolution of active metamorphic core complexes,
D'Entrecasteaux Islands, Papua New Guinea. Tectonics, 12(3), 611-628.
Baldwin, S. L., Monteleone, B. D., Webb, L. E., Fitzgerald, P. G., Grove, M., & June Hill, E. (2004).
Pliocene eclogite exhumation at plate tectonic rates in eastern Papua New
Guinea. Nature, 431(7006), 263-267.
Baldwin, S. L., Webb, L. E., & Monteleone, B. D. (2008). Late Miocene coesite-eclogite exhumed in
the Woodlark Rift. Geology, 36(9), 735-738.
Ballance, P. F., & Spörli, K. B. (1979). Northland allochthon. Journal of the Royal Society of New
Zealand, 9(2), 259-275.
Ballantyne, P. (1991). Petrological constraints upon the provenance and genesis of the East
Halmahera ophiolite. Journal of Southeast Asian Earth Sciences, 6(3-4), 259-269.
Ballantyne, P. D.;(1990). The petrology of the ophiolitic rocks of the Halmahera region,
Indonesia. Doctoral thesis (Ph.D), UCL (University College London).
Balmater, H. G., Manalo, P. C., Faustino-Eslava, D. V., Queaño, K. L., Dimalanta, C. B., Guotana, J. M.
R., ... & Yumul Jr, G. P. (2015). Paleomagnetism of the Samar Ophiolite: Implications for
the Cretaceous sub-equatorial position of the Philippine island arc. Tectonophysics, 664,
214-224.
Barrier, E., Huchon, P., & Aurelio, M. (1991). Philippine fault: A key for Philippine
kinematics. Geology, 19(1), 32-35.
Batara, B., & Xu, C. (2022). Evolved magmatic arcs of South Borneo: Insights into Cretaceous slab
subduction. Gondwana Research, 111, 142-164.
Battan, R., Chung, S. L., Komiya, T., Maruyama, S., Lin, A. T., Lee, H. Y., & Iizuka, Y. (2022,
December). Zircon U-Pb Ages and Geochemical Characteristics of Magmatic Rocks from
Choiseul and Santa Isabel, Solomon Islands: Implications for the Magmatic and Tectonic
Evolution in SW Pacific. In AGU Fall Meeting Abstracts (Vol. 2022, pp. V12C-0058).
Beiersdorf, H., Bach, W., Delisle, G., Faber, E., Gerling, P., Hinz, K., & Dheeradilok, P. (1997). Age
and possible modes of formation of the Celebes Sea basement, and thermal regimes
within the accretionary complexes off SW Mindanao and N Sulawesi. In Proceedings of
the International Conference on Stratigraphy and Tectonic Evolution of Southeast Asia
and the South Pacific (pp. 369-387).
Bellon, H. & Rangin, C. (1991). Geochemistry and isotopic dating of Cenozoic volcanic arc
sequences around the Celebes and Sulu Seas. Proceedings of the Ocean Drilling Program
Scientific Results, 124, 321-338
Bellon, H., & Yumul Jr, G. P. (2000). Mio-Pliocene magmatism in the Baguio Mining District
(Luzon, Philippines): age clues to its geodynamic setting. Comptes Rendus de l'Académie
des Sciences-Series IIA-Earth and Planetary Science, 331(4), 295-302.
Benyshek, E. K., & Taylor, B. (2021). Tectonics of the Papua-Woodlark Region. Geochemistry,
Geophysics, Geosystems, 22(1), e2020GC009209.
Bergman, S. C., Hutchison, C. S., Swauger, D. A., & Graves, J. E. (2000). K: Ar ages and
geochemistry of the Sabah Cenozoic volcanic rocks. Geological society of Malaysia
Bulletin 44, 165-171.
Berly, T. (2005). Ultramafic and mafic rock types from Choiseul, Santa Isabel and santa Jorge
(Northeastern Solomon Islands): origins and significance (Doctoral dissertation,
Université Joseph-Fourier-Grenoble I).
Berly, T. J., Hermann, J., Arculus, R. J., & Lapierre, H. (2006). Supra-subduction zone pyroxenites
from San Jorge and Santa Isabel (Solomon Islands). Journal of Petrology, 47(8), 1531-
1555.
Bernard, A., Munschy, M., Rotstein, Y., & Sauter, D. (2005). Refined spreading history at the
Southwest Indian Ridge for the last 96 Ma, with the aid of satellite gravity
data. Geophysical Journal International, 162(3), 765-778.
Billedo, E. B. (1996). The pre-Tertiary ophiolitic complex of northeastern Luzon and the Polillo
group of islands, Philippines. Journal of the Geological Society of the Philippines, 51, 95-
114.
Billedo, E. B., Stephan, J. F., Delteilt, J., Bellon, H., Sajona, F., Feraud, G. (1996). The pre-Tertiary
ophiolitic complex of northeastern Luzon and the Polillo group of islands,
Philippines. Journal of the Geological Society of the Philippines, 51, 95-114.
Bladon, G. M. (1988). Catalogue, appraisal and significance of K-Ar isotopic ages determined for
igneous and metamorphic rocks in Irian Jaya. Geological Research and Development
Centre, Bandung, Indonesia. Preliminary Geological Report, 86.
Boschman, L. M., & Van Hinsbergen, D. J. (2016). On the enigmatic birth of the Pacific Plate
within the Panthalassa Ocean. Science Advances, 2(7), e1600022.
Boschman, L. M., Van der Wiel, E., Flores, K. E., Langereis, C. G., & van Hinsbergen, D. J. (2019).
The Caribbean and Farallon plates connected: Constraints from stratigraphy and
paleomagnetism of the Nicoya Peninsula, Costa Rica. Journal of Geophysical Research:
Solid Earth, 124(7), 6243-6266.
Boschman, L. M., Van Hinsbergen, D. J., Langereis, C. G., Flores, K. E., Kamp, P. J., Kimbrough, D. L.,
... & Spakman, W. (2021b). Reconstructing lost plates of the Panthalassa Ocean through
paleomagnetic data from circum-Pacific accretionary orogens. American Journal of
Science, 321(6), 907-954.
Boschman, L. M., van Hinsbergen, D. J., Torsvik, T. H., Spakman, W., & Pindell, J. L. (2014).
Kinematic reconstruction of the Caribbean region since the Early Jurassic. Earth-Science
Reviews, 138, 102-136.
Boyden, J. A., Müller, R. D., Gurnis, M., Torsvik, T. H., Clark, J. A., Turner, M., ... & Cannon, J. S.
(2011). Next-generation plate-tectonic reconstructions using GPlates. Geoinformatics:
Cyberinfrastructure for the Solid Earth Sciences, Cambridge University
Press, Cambridge (2011), pp. 95-113
Breitfeld, H. T., Hall, R., Galin, T., & BouDagher-Fadel, M. K. (2018). Unravelling the stratigraphy
and sedimentation history of the uppermost Cretaceous to Eocene sediments of the
Kuching Zone in West Sarawak (Malaysia), Borneo. Journal of Asian Earth Sciences, 160,
200-223.
Breitfeld, H. T., Hall, R., Galin, T., Forster, M. A., & BouDagher-Fadel, M. K. (2017). A Triassic to
Cretaceous Sundaland–Pacific subduction margin in West Sarawak,
Borneo. Tectonophysics, 694, 35-56.
Briais, A., Patriat, P., & Tapponnier, P. (1993). Updated interpretation of magnetic anomalies and
seafloor spreading stages in the South China Sea: Implications for the Tertiary tectonics
of Southeast Asia. Journal of Geophysical Research: Solid Earth, 98(B4), 6299-6328.
Brocher, T. M. (1985). On the formation of the Vitiaz Trench lineament and North Fiji Basin.
Geological investigations of the Northern Melanesian Borerland, Circum-Pacific Council for
Energy and Mineral Resources Earth Science Series 3, 13-33.
Brown, D., Alvarez-Marron, J., Camanni, G., Biete, C., Kuo-Chen, H., & Wu, Y. M. (2022). Structure
of the south-central Taiwan fold-and-thrust belt: Testing the viability of the
model. Earth-Science Reviews, 231, 104094.
Brown, D., Alvarez-Marron, J., Biete, C., Kuo-Chen, H., Camanni, G., & Ho, C. W. (2017). How the
structural architecture of the Eurasian continental margin affects the structure,
seismicity, and topography of the south central Taiwan fold-and-thrust
belt. Tectonics, 36(7), 1275-1294.
Brown, D., Ryan, P. D., Byrne, T., Chan, Y. C., Rau, R. J., Lu, C. Y., ... & Wang, Y. J. (2011). The arc–
continent collision in Taiwan. Arc-continent collision, 213-245.
Burton-Johnson, A., Macpherson, C. G., Millar, I. L., Whitehouse, M. J., Ottley, C. J., & Nowell, G. M.
(2020). A Triassic to Jurassic arc in north Borneo: Geochronology, geochemistry, and
genesis of the Segama Valley Felsic Intrusions and the Sabah ophiolite. Gondwana
Research, 84, 229-244.
Buys, J., Spandler, C., Holm, R. J., & Richards, S. W. (2014). Remnants of ancient Australia in
Vanuatu: Implications for crustal evolution in island arcs and tectonic development of
the southwest Pacific. Geology, 42(11), 939-942.
Camanni, G., & Ye, Q. (2022). The significance of fault reactivation on the Wilson cycle
undergone by the northern South China Sea area in the last 60 Myr. Earth-Science
Reviews, 225, 103893.
Cande, S. C., & Stock, J. M. (2004). Pacific—Antarctic—Australia motion and the formation of the
Macquarie Plate. Geophysical Journal International, 157(1), 399-414.
Cande, S. C., Patriat, P., & Dyment, J. (2010). Motion between the Indian, Antarctic and African
plates in the early Cenozoic. Geophysical Journal International, 183(1), 127-149.
Cao, L., Shao, L., Qiao, P., Cui, Y., Zhang, G., & Zhang, X. (2021). Formation and paleogeographic
evolution of the Palawan continental terrane along the Southeast Asian margin revealed
by detrital fingerprints. Bulletin, 133(5-6), 1167-1193.
Cao, L., Shao, L., van Hinsbergen, D. J., Jiang, T., Xu, D., & Cui, Y. (2023). Provenance and evolution
of East Asian large rivers recorded in the East and South China Seas: A review. Geological
Society of America Bulletin.
Casey, J. F., & Dewey, J. F. (1984). Initiation of subduction zones along transform and accreting
plate boundaries, triple-junction evolution, and forearc spreading centres—implications
for ophiolitic geology and obduction. Geological Society, London, Special
Publications, 13(1), 269-290.
Chambers, L. M., Pringle, M. S., & Fitton, J. G. (2004). Phreatomagmatic eruptions on the Ontong
Java Plateau: an Aptian 40Ar/39Ar age for volcaniclastic rocks at ODP Site
1184. Geological Society, London, Special Publications, 229(1), 325-331.
Chandler, M. T., Wessel, P., Taylor, B., Seton, M., Kim, S. S., & Hyeong, K. (2012). Reconstructing
Ontong Java Nui: Implications for Pacific absolute plate motion, hotspot drift and true
polar wander. Earth and Planetary Science Letters, 331, 140-151.
Charlton, T. R., Hall, R., & Partoyo, E. (1991). The geology and tectonic evolution of Waigeo
Island, NE Indonesia. Journal of Southeast Asian Earth Sciences, 6(3-4), 289-297.
Chen, W. S., Chung, S. L., Chou, H. Y., Zugeerbai, Z., Shao, W. Y., & Lee, Y. H. (2017). A
reinterpretation of the metamorphic Yuli belt: Evidence for a middle-late Miocene
accretionary prism in eastern Taiwan. Tectonics, 36(2), 188-206.
Chen, W. S., Huang, Y. C., Liu, C. H., Feng, H. T., Chung, S. L., & Lee, Y. H. (2016). UPb zircon
geochronology constraints on the ages of the Tananao Schist Belt and timing of orogenic
events in Taiwan: Implications for a new tectonic evolution of the South China Block
during the Mesozoic. Tectonophysics, 686, 68-81.
Chivas, A. R., & McDougall, I. (1978). Geochronology of the Koloula porphyry copper prospect,
Guadalcanal, Solomon Islands. Economic Geology, 73(5), 678-689.
Chung, S. L., & Sun, S. S. (1992). A new genetic model for the East Taiwan Ophiolite and its
implications for Dupal domains in the Northern Hemisphere. Earth and Planetary
Science Letters, 109(1-2), 133-145.
Cloos, M., Sapiie, B., van Ufford, A. Q., Weiland, R. J., Warren, P. Q., & McMahon, T. P. (2005).
Collisional delamination in New Guinea: The geotectonics of subducting slab breakoff.
GSA Special Papers.
Cluzel, D., & Meffre, S. (2019). In search of Gondwana heritage in the Outer Melanesian Arc: no
pre-upper Eocene detrital zircons in Viti Levu river sands (Fiji Islands). Australian
Journal of Earth Sciences, 66(2), 265-277.
Cluzel, D., Aitchison, J. C., & Picard, C. (2001). Tectonic accretion and underplating of mafic
terranes in the Late Eocene intraoceanic fore-arc of New Caledonia (Southwest Pacific):
geodynamic implications. Tectonophysics, 340(1-2), 23-59.
Cluzel, D., Jourdan, F., Meffre, S., Maurizot, P., & Lesimple, S. (2012b). The metamorphic sole of
New Caledonia ophiolite: 40Ar/39Ar, U-Pb, and geochemical evidence for subduction
inception at a spreading ridge. Tectonics, 31(3).
Cluzel, D., Maurizot, P., Collot, J., & Sevin, B. (2012a). An outline of the Geology of New Caledonia;
from Permian–Mesozoic Southeast Gondwanaland active margin to Cenozoic obduction
and supergene evolution. Episodes Journal of International Geoscience, 35(1), 72-86.
Cluzel, D., Whitten, M., Meffre, S., Aitchison, J. C., & Maurizot, P. (2018). A Reappraisal of the Poya
Terrane (New Caledonia): Accreted Late Cretaceous-Paleocene Marginal Basin Upper
Crust, Passive Margin Sediments, and Early Eocene E-MORB Sill
Complex. Tectonics, 37(1), 48-70.
Cogné, J. P., Kravchinsky, V. A., Halim, N., & Hankard, F. (2005). Late Jurassic-Early Cretaceous
closure of the Mongol-Okhotsk Ocean demonstrated by new Mesozoic palaeomagnetic
results from the Trans-Baikal area (SE Siberia). Geophysical Journal
International, 163(2), 813-832.
Cole, J., McCabe, R., Moriarty, T., Malicse, J. A., Delfin, F. G., Tebar, H., & Ferrer, H. P. (1989). A
preliminary Neogene paleomagnetic data set from Leyte and its relation to motion on
the Philippine fault. Tectonophysics, 168(1-3), 205-220.
Cole, W. S. (1950). Larger foraminifers from the Palau Islands. US Geological Survey Professional
Paper, B, 221, 21-26.
Collot, J. Y., Malahoff, A., Récy, J., Latham, G., & Missègue, F. (1987). Overthrust emplacement of
New Caledonia ophiolite: geophysical evidence. Tectonics, 6(3), 215-232.
Collot, J., Herzer, R., Lafoy, Y., & Geli, L. (2009). Mesozoic history of the Fairway-Aotea Basin:
Implications for the early stages of Gondwana fragmentation. Geochemistry, Geophysics,
Geosystems, 10(12).
Collot, J., Vendé-Leclerc, M., Rouillard, P., Lafoy, Y., & Géli, L. (2012). Map helps unravel
complexities of the southwestern Pacific Ocean. Eos, Transactions American Geophysical
Union, 93(1), 1-2.
Conand, C., Mouthereau, F., Ganne, J., Lin, A. T. S., Lahfid, A., Daudet, M., Mesalles, L., Giletycz, S., &
Bonzani, M. (2020). Strain partitioning and exhumation in oblique Taiwan collision: Role
of rift architecture and plate kinematics. Tectonics, 39(4), e2019TC005798.
Cosca, M., Arculus, R., Pearce, J., & Mitchell, J. (1998). 40Ar/39Ar and K–Ar geochronological age
constraints for the inception and early evolution of the Izu–Bonin–Mariana arc
system. Island Arc, 7(3), 579-595.
Coulson, F. I., & Vedder, J. G. (1986). Geology of the central and western Solomon Islands.
Geology and Offshore Resources of Pacific Island Arcs–central and western Solomon
Islands, 4, 59-87.
Cowley, S., Mann, P., Coffin, M. F., & Shipley, T. H. (2004). Oligocene to Recent tectonic history of
the Central Solomon intra-arc basin as determined from marine seismic reflection data
and compilation of onland geology. Tectonophysics, 389(3-4), 267-307.
Cox, A., & Hart, R. B. (1986). Plate Tectonics: How It Works. Blackwell Scientific Publications.
Inc, Boston. 392 pp.
Crameri, F., Magni, V., Domeier, M., Shephard, G. E., Chotalia, K., Cooper, G., ... & Thielmann, M.
(2020). A transdisciplinary and community-driven database to unravel subduction zone
initiation. Nature communications, 11(1), 3750.
Crawford, A. J., Meffre, S., & Symonds, P. A. (2003). 120 to 0 Ma tectonic evolution of the
southwest Pacific and analogous geological evolution of the 600 to 220 Ma Tasman Fold
Belt System. Geological Society of Australia Special Publication, 22, 377-397.
Croon, M. B., Cande, S. C., & Stock, J. M. (2008). Revised Pacific-Antarctic plate motions and
geophysics of the Menard Fracture Zone. Geochemistry, Geophysics, Geosystems, 9(7).
Crowhurst, P. V., Maas, R., Hill, K. C., Foster, D. A., & Fanning, C. M. (2004). Isotopic constraints on
crustal architecture and Permo-Triassic tectonics in New Guinea: Possible links with
eastern Australia. Australian Journal of Earth Sciences, 51(1), 107-122.
Curray, J. R. (2005). Tectonics and history of the Andaman Sea region. Journal of Asian Earth
Sciences, 25(1), 187-232.
David Jr, S., Stephan, J. F., Delteil, J., Müller, C., Butterlin, J., Bellon, H., & Billedo, E. (1997).
Geology and tectonic history of Southeastern Luzon, Philippines. Journal of Asian Earth
Sciences, 15(4-5), 435-452.
Davies, H. L. (1980b). Folded thrust fault and associated metamorphism in the Suckling-Dayman
massif, Papua New Guinea. American Journal of Science, 280A, 171–191.
Davies, H. L. (2005). The geology of Bougainville. Bougainville before the conflict (eds AJ Regan &
HM Griffin), 20-30.
Davies, H. L. (2012). The geology of New Guinea-the cordilleran margin of the Australian
continent. Episodes Journal of International Geoscience, 35(1), 87-102.
Davies, H. L., & Jaques, A. L. (1984). Emplacement of ophiolite in Papua New Guinea. Geological
Society, London, Special Publications, 13(1), 341-349.
Davies, H. L., & Smith, I. E. (1971). Geology of eastern Papua. Geological Society of America
Bulletin, 82(12), 3299-3312.
Davies, H. L., & Warren, R. G. (1988). Origin of eclogite-bearing, domed, layered metamorphic
complexes (“core complexes”) in the D'Entrecasteaux islands, Papua New
Guinea. Tectonics, 7(1), 1-21.
Davies, H.L. (1982). Mianmin – 1 : 250 000 geological series. Geological Survey of Papua New
Guinea, Dept. of Minerals and Energy, Port Moresby.
De Jesus, J. V., Yumul, G. P., & Faustino, D. V. (2000). The Cansiwang Melange of southeast Bohol
(central Philippines): origin and tectonic implications. Island Arc, 9(4), 565-574.
Decker, J., Ferdian, F., Morton, A., Fanning, M., & White, L. T. (2017). New geochronology data
from eastern Indonesia–An aid to understanding sedimentary provenance in a frontier
region.
Defant, M. J., Jacques, D., Maury, R. C., de Boer, J., & Joron, J. L. (1989). Geochemistry and tectonic
setting of the Luzon arc, Philippines. Geological Society of America Bulletin, 101(5), 663-
672.
Defant, M. J., Maury, R., Joron, J. L., Feigenson, M. D., Leterrier, J., Bellon, H., ... & Richard, M.
(1990). The geochemistry and tectonic setting of the northern section of the Luzon arc
(the Philippines and Taiwan). Tectonophysics, 183(1-4), 187-205.
DeMets, C., Gordon, R. G., & Argus, D. F. (2010). Geologically current plate motions. Geophysical
journal international, 181(1), 1-80.
DeMets, C., Iaffaldano, G., & Merkouriev, S. (2015). High-resolution Neogene and quaternary
estimates of Nubia-Eurasia-North America plate motion. Geophysical Journal
International, 203(1), 416-427.
DeMets, C., Merkouriev, S., & Sauter, D. (2021). High resolution reconstructions of the Southwest
Indian Ridge, 52 Ma to present: implications for the breakup and absolute motion of the
Africa plate. Geophysical Journal International, 226(3), 1461-1497.
Deng, J., Yang, X., Qi, H., Zhang, Z. F., Mastoi, A. S., & Sun, W. (2017). Early Cretaceous high-Mg
adakites associated with Cu-Au mineralization in the Cebu Island, Central Philippines:
Implication for partial melting of the paleo-Pacific Plate. Ore Geology Reviews, 88, 251-
269.
Deng, J., Yang, X., Qi, H., Zhang, Z. F., Mastoi, A. S., Al Emil, G. B., & Sun, W. (2019). Early
Cretaceous adakite from the Atlas porphyry Cu-Au deposit in Cebu Island, Central
Philippines: Partial melting of subducted oceanic crust. Ore Geology Reviews, 110,
102937.
Deng, J., Yang, X., Zhang, L. P., Duan, L., Mastoi, A. S., & Liu, H. (2020). An overview on the origin
of adakites/adakitic rocks and related porphyry Cu-Au mineralization, Northern Luzon,
Philippines. Ore Geology Reviews, 124, 103610.
Deng, J., Yang, X., Zhang, Z. F., & Santosh, M. (2015). Early Cretaceous arc volcanic suite in Cebu
Island, Central Philippines and its implications on paleo-Pacific plate subduction:
Constraints from geochemistry, zircon U–Pb geochronology and Lu–Hf
isotopes. Lithos, 230, 166-179.
Deschamps, A., & Lallemand, S. (2002). The West Philippine Basin: An Eocene to early Oligocene
back arc basin opened between two opposed subduction zones. Journal of Geophysical
Research: Solid Earth, 107(B12), EPM-1.
Deschamps, A., Monié, P., Lallemand, S., Hsu, S. K., & Yeh, K. Y. (2000). Evidence for Early
Cretaceous oceanic crust trapped in the Philippine Sea Plate. Earth and Planetary Science
Letters, 179(3-4), 503-516.
Dewey, J. F., & Casey, J. F. (2011). The origin of obducted large-slab ophiolite complexes. Arc-
continent collision, 431-444.
Diegor, W. G. (1996). The ophiolitic basement complex of Cebu. Journal of the Geological Society
of the Philippines, 51, 48-60.
Dimalanta, C. B., Faustino-Eslava, D. V., Gabo-Ratio, J. A. S., Marquez, E. J., Padrones, J. T., Payot, B.
D., ... & Yumul Jr, G. P. (2020). Characterization of the proto-Philippine Sea Plate: Evidence
from the emplaced oceanic lithospheric fragments along eastern Philippines. Geoscience
Frontiers, 11(1), 3-21.
Dimalanta, C. B., Ramos, E. G. L., Yumul Jr, G. P., & Bellon, H. (2009). New features from the
Romblon Island Group: Key to understanding the arc–continent collision in Central
Philippines. Tectonophysics, 479(1-2), 120-129.
Dimalanta, C. B., Suerte, L. O., Yumul, G. P., Tamayo, R. A., & Ramos, E. G. L. (2006). A Cretaceous
supra-subduction oceanic basin source for Central Philippine ophiolitic basement
complexes: Geological and geophysical constraints. Geosciences Journal, 10, 305-320.
Domeier, M., Shephard, G. E., Jakob, J., Gaina, C., Doubrovine, P. V., & Torsvik, T. H. (2017).
Intraoceanic subduction spanned the Pacific in the Late Cretaceous–Paleocene. Science
Advances, 3(11), eaao2303.
Dong, D., Zhang, Z., Bai, Y., Fan, J., & Zhang, G. (2018). Topographic and sedimentary features in
the Yap subduction zone and their implications for the Caroline Ridge
subduction. Tectonophysics, 722, 410-421.
Doo, W. B., Hsu, S. K., Yeh, Y. C., Tsai, C. H., & Chang, C. M. (2015). Age and tectonic evolution of
the northwest corner of the West Philippine Basin. Marine Geophysical Research, 36,
113-125.
Doubrovine, P. V., Steinberger, B., & Torsvik, T. H. (2012). Absolute plate motions in a reference
frame defined by moving hot spots in the Pacific, Atlantic, and Indian oceans. Journal of
Geophysical Research: Solid Earth, 117(B9).
Doust, H. (1990). Geology of the Sepik basin, Papua New Guinea. Papua New Guinea (PNG)
Proceedings of the First PNG Petroleum Convention, 461-478.
Dow, D. B., & Sukamto, R. (1984). Western Irian Jaya: the end-product of oblique plate
convergence in the late Tertiary. Tectonophysics, 106(1-2), 109-139.
Dycoco, J. M. A., Payot, B. D., Valera, G. T. V., Labis, F. A. C., Pasco, J. A., Perez, A. D., & Tani, K.
(2021). Juxtaposition of Cenozoic and Mesozoic ophiolites in Palawan island,
Philippines: New insights on the evolution of the Proto-South China
Sea. Tectonophysics, 819, 229085.
Encarnación, J. (2004). Multiple ophiolite generation preserved in the northern Philippines and
the growth of an island arc complex. Tectonophysics, 392(1-4), 103-130.
Encarnación, J. P., Mukasa, S. B., & Obille Jr, E. C. (1993). Zircon U-Pb geochronology of the
Zambales and Angat Ophiolites, Luzon, Philippines: Evidence for an Eocene arc-back arc
pair. Journal of Geophysical Research: Solid Earth, 98(B11), 19991-20004.
Engebretson, D. C., Cox, A., & Gordon, R. G. (1985). Relative motions between oceanic and
continental plates in the Pacific basin. GSA Special Papers, 206.
Evans, C. A., Hawkins, J. W., & Moore, G. F. (1983). Petrology and geochemistry of ophiolitic and
associated volcanic rocks on the Talaud Islands, Molucca Sea Collision Zone, northeast
Indonesia. Geodynamics of the Western Pacific-Indonesian Region, 11, 159-172.
Fabbri, O., Monié, P., & Fournier, M. (2004). Transtensional deformation at the junction between
the Okinawa trough back-arc basin and the SW Japan island arc. Geological Society,
London, Special Publications, 227(1), 297-312.
Faccenna, C., Becker, T. W., Lallemand, S., & Steinberger, B. (2012). On the role of slab pull in the
Cenozoic motion of the Pacific plate. Geophysical Research Letters, 39(3).
Fairbridge, R. W., & Stewart Jr, H. B. (1960). Alexa Bank, a drowned atoll on the Melanesian
Border Plateau. Deep Sea Research (1953), 7(2), 100-116.
Falvey, D. A., & Taylor, L. W. (1974). Queensland Plateau and Coral Sea Basin: structural and
time-stratigraphic patterns. Exploration Geophysics, 5(4), 123-126.
Faure, M., & Ishida, K. (1990). The Mid-Upper Jurassic olistostrome of the west Philippines: a
distinctive key-marker for the North Palawan block. Journal of Southeast Asian Earth
Sciences, 4(1), 61-67.
Faustino, D. V., Yumul, G. P., De Jesus, J. V., Dimalanta, C. B., Aitchison, J. C., Zhou, M. F., ... & De
Leon, M. M. (2003). Geology of southeast Bohol, central Philippines: accretion and
sedimentation in a marginal basin. Australian Journal of Earth Sciences, 50(4), 571-583.
Fisher, D. M., Lu, C.-Y., & Chu, H.-T. (2002). Taiwan slate belt: Insights into the ductile interior of
an arc-continent collision. Geological Society of America: Special Paper, 358, 93–106.
Francis, G. (1988). Stratigraphy of Manus Island, Western New Ireland Basin, Papua New
Guinea. Geology and Offshore Resources of Pacific Island Arcs–New Ireland and Manus
Region, Papua New Guinea, 9, 31-40.
Fujiwara, T., Tamaki, K., Fujimoto, H., Ishii, T., Seama, N., Toh, H., ... & Kinoshita, H. (1995).
Morphological studies of the Ayu trough, Philippine sea–Caroline plate
boundary. Geophysical Research Letters, 22(2), 109-112.
Fuller, M., Haston, R., & Almasco, J. (1989). Paleomagnetism of the Zambales ophiolite, Luzon,
northern Philippines. Tectonophysics, 168(1-3), 171-203.
Fuller, M., Haston, R., Lin, J. L., Richter, B., Schmidtke, E., & Almasco, J. (1991). Tertiary
paleomagnetism of regions around the South China Sea. Journal of Southeast Asian Earth
Sciences, 6(3-4), 161-184.
Gans, P. B., Mortimer, N., Patriat, M., Turnbull, R. E., Crundwell, M. P., Agranier, A., ... & Collot, J.
(2023). Detailed 40Ar/39Ar geochronology of the Loyalty and Three Kings Ridges
clarifies the extent and sequential development of Eocene to Miocene southwest Pacific
remnant volcanic arcs. Geochemistry, Geophysics, Geosystems, 24(2), e2022GC010670.
Gaina, C., & Müller, D. (2007). Cenozoic tectonic and depth/age evolution of the Indonesian
gateway and associated back-arc basins. Earth-Science Reviews, 83(3-4), 177-203.
Gaina, C., Müller, D. R., Royer, J. Y., Stock, J., Hardebeck, J., & Symonds, P. (1998). The tectonic
history of the Tasman Sea: a puzzle with 13 pieces. Journal of Geophysical Research: Solid
Earth, 103(B6), 12413-12433.
Gaina, C., Müller, R. D., Royer, J. Y., & Symonds, P. (1999). Evolution of the Louisiade triple
junction. Journal of Geophysical Research: Solid Earth, 104(B6), 12927-12939.
Garrison, R. E., Espiritu, E., Horan, L. J., & Mack, L. E. (1979). Petrology, sedimentology, and
diagenesis of hemipelagic limestone and tuffaceous turbidities in the Aksitero
Formation, central Luzon, Philippines. U.S. Geological Survey Professional Papers 1112,
1–16
Geary, E. E., & Kay, R. W. (1989). Identification of an early cretaceous ophiolite in the Camarines
Norte-Calaguas Islands basement complex, eastern Luzon,
Philippines. Tectonophysics, 168(1-3), 109-126.
Geary, E. E., Harrison, T. M., & Heizler, M. (1988). Diverse ages and origins of basement
complexes, Luzon, Philippines. Geology, 16(4), 341-344.
Geary, E.E. (1986) Tectonic significance of basement complexes and ophiolites in the northern
Philippines: Results of geological, geochronological and geochemical investigations
(Ph.D. thesis). Ithaca, New York, Cornell University, 221 p.
Gill, J., Todd, E., Hoernle, K., Hauff, F., Price, A. A., & Jackson, M. G. (2022). Breaking up is hard to
do: Magmatism during oceanic arc breakup, subduction reversal, and
cessation. Geochemistry, Geophysics, Geosystems, 23(12), e2022GC010663.
Gong, L., Hollings, P., Zhang, Y., Tian, J., Li, D., & Chen, H. (2021). Contribution of an Eastern
Indochina-derived fragment to the formation of island arc systems in the Philippine
Mobile Belt. GSA Bulletin, 133(9-10), 1979-1995.
Granot, R., & Dyment, J. (2018). Late Cenozoic unification of East and West Antarctica. Nature
Communications, 9(1), 3189.
Granot, R., Cande, S. C., Stock, J. M., & Damaske, D. (2013). Revised Eocene-Oligocene kinematics
for the West Antarctic rift system. Geophysical Research Letters, 40(2), 279-284.
Griffin, T. J., & Moresby, P. (1983). Granitoids of the Tertiary continent-island are collision zone,
Papua New Guinea. Geological Society of America, Memoir i59, 61-76.
Grobys, J. W., Gohl, K., & Eagles, G. (2008). Quantitative tectonic reconstructions of Zealandia
based on crustal thickness estimates. Geochemistry, Geophysics, Geosystems, 9(1).
Guilmette, C., Smit, M. A., van Hinsbergen, D. J., Gürer, D., Corfu, F., Charette, B., ... & Savard, D.
(2018). Forced subduction initiation recorded in the sole and crust of the Semail
Ophiolite of Oman. Nature Geoscience, 11(9), 688-695.
Guilmette, C., van Hinsbergen, D., Smit, M., Godet, A., Fournier-Roy, F., Butler, J., ... & Hodges, K.
(2023). Formation of the Xigaze Metamorphic Sole under the Tibetan continental
margin reveals generic characteristics of the subduction initiation process. Preprint
available via: https://doi.org/10.21203/rs.3.rs-2621281/v1
Gungor, A., Lee, G. H., Kim, H. J., Han, H. C., Kang, M. H., Kim, J., & Sunwoo, D. (2012). Structural
characteristics of the northern Okinawa Trough and adjacent areas from regional
seismic reflection data: geologic and tectonic implications. Tectonophysics, 522, 198-207.
Guotana, J. M. R., Payot, B. D., Dimalanta, C. B., Ramos, N. T., Faustino-Eslava, D. V., Queaño, K. L.,
& Yumul Jr, G. P. (2017). Arc and backarc geochemical signatures of the proto-Philippine
Sea Plate: Insights from the petrography and geochemistry of the Samar Ophiolite
volcanic section. Journal of Asian Earth Sciences, 142, 77-92.
Guotana, J. M. R., Payot, B. D., Dimalanta, C. B., Ramos, N. T., Faustino-Eslava, D. V., Queaño, K. L.,
& Yumul Jr, G. P. (2018). Petrological and geochemical characteristics of the Samar
Ophiolite ultramafic section: implications on the origins of the ophiolites in Samar and
Leyte islands, Philippines. International Geology Review, 60(4), 401-417.
Gürer, D., Granot, R., & van Hinsbergen, D. J. (2022). Plate tectonic chain reaction revealed by
noise in the Cretaceous quiet zone. Nature Geoscience, 15(3), 233-239.
Gurnis, M., Hall, C., & Lavier, L. (2004). Evolving force balance during incipient
subduction. Geochemistry, Geophysics, Geosystems, 5(7).
Hall, C. E., Gurnis, M., Sdrolias, M., Lavier, L. L., & Müller, R. D. (2003). Catastrophic initiation of
subduction following forced convergence across fracture zones. Earth and Planetary
Science Letters, 212(1-2), 15-30.
Hall, R. (2002). Cenozoic geological and plate tectonic evolution of SE Asia and the SW Pacific:
computer-based reconstructions, model and animations. Journal of Asian earth
sciences, 20(4), 353-431.
Hall, R. (2012). Late Jurassic–Cenozoic reconstructions of the Indonesian region and the Indian
Ocean. Tectonophysics, 570, 1-41.
Hall, R., & Breitfeld, H. T. (2017). Nature and demise of the proto-South China Sea. Bulleting of
the Geological Society of Malaysia, 63, 61-76.
Hall, R., & Spakman, W. (2002). Subducted slabs beneath the eastern Indonesia–Tonga region:
insights from tomography. Earth and Planetary Science Letters, 201(2), 321-336.
Hall, R., Audley-Charles, M. G., Banner, F. T., Hidayat, S., & Tobing, S. L. (1988). Basement rocks of
the Halmahera region, eastern Indonesia: a Late Cretaceous–Early Tertiary arc and fore-
arc. Journal of the Geological Society, 145(1), 65-84.
Haston, R., Fuller, M., & Schmidtke, E. (1988). Paleomagnetic results from Palau, West Caroline
Islands: a constraint on Philippine Sea plate motion. Geology, 16(7), 654-657.
Hawkins, J. W. (1985). Geology of the composite terranes of east and central Mindanao. Circum
Pacific Council Publications, Earth Science Series 1, 437-463.
Hawkins, J. W., & Evans, C. A. (1983). Geology of the Zambales Range, Luzon, Philippine Islands:
Ophiolite derived from an island arc-back arc basin pair. Washington DC American
Geophysical Union Geophysical Monograph Series, 27, 95-123.
Hawkins, J. W., & Ishizuka, O. (2009). Petrologic evolution of Palau, a nascent island arc. Island
Arc, 18(4), 599-641.
Hawkins, J., & Batiza, R. (1977). Metamorphic rocks of the Yap arc-trench system. Earth and
Planetary Science Letters, 37(2), 216-229.
Hegarty, K. A., Weissel, J. K., & Hayes, D. E. (1983). Convergence at the Caroline-Pacific plate
boundary: collision and subduction. Washington DC American Geophysical Union
Geophysical Monograph Series, 27, 326-348.
Herzer, R. H., Davy, B. W., Mortimer, N., Quilty, P. G., Chaprioniere, G. C. H., Jones, C. M., et al.
(2009). Seismic stratigraphy and structure of the Northland Plateau and the
development of the Vening Meinesz transform margin, SW Pacific Ocean. Marine
Geophysical Researches, 30(1), 21–60.
Hickey-Vargas, R., Ishizuka, O., & Bizimis, M. (2013). Age and geochemistry of volcanic clasts
from DSDP Site 445, Daito Ridge and relationship to Minami-Daito Basin and early Izu-
Bonin arc magmatism. Journal of Asian Earth Sciences, 70, 193-208.
Hickey-Vargas, R. (1998). Origin of the Indian Ocean-type isotopic signature in basalts from
Philippine Sea plate spreading centers: An assessment of local versus large-scale
processes. Journal of Geophysical Research: Solid Earth, 103(B9), 20963-20979.
Hickey-Vargas, R. (2005). Basalt and tonalite from the Amami Plateau, northern West Philippine
Basin: New Early Cretaceous ages and geochemical results, and their petrologic and
tectonic implications. Island Arc, 14(4), 653-665.
Hilde, T. W., & Lee, C. S. (1984). Origin and evolution of the West Philippine Basin: a new
interpretation. Tectonophysics, 102(1-4), 85-104.
Hill, E. J. (1994). Geometry and kinematics of shear zones formed during continental extension
in eastern Papua New Guinea. Journal of Structural Geology, 16(8), 1093-1105.
Hill, E. J., & Baldwin, S. L. (1993). Exhumation of high-pressure metamorphic rocks during
crustal extension in the D'Entrecasteaux region, Papua New Guinea. Journal of
Metamorphic Geology, 11(2), 261-277.
Hill, K. C. (1991). Structure of the Papuan fold belt, Papua New Guinea. AAPG bulletin, 75(5),
857-872.
Hill, K. C., & Gleadow, A. J. W. (1989). Uplift and thermal history of the Papuan Fold Belt, Papua
New Guinea: Apatite fission track analysis. Australian Journal of Earth Sciences, 36(4),
515-539.
Hill, K. C., & Hall, R. (2002). Mesozoic–Cainozoic evolution of Australia's New Guinea margin in a
West Pacific context. Defining Australia: the Australian Plate as Part of Planet Earth:
Geological Society of America and Geological Society of Australia, joint publication, special
paper, 1-43.
Hill, K. C., & Raza, A. (1999). Arc-continent collision in Papua Guinea: Constraints from fission
track thermochronology. Tectonics, 18(6), 950-966.
Hinz, K., Block, M., Kudrass, H. R., & Meyer, H. (1991). Structural elements of the Sulu Sea,
Philippines. AAPG Bulletin 78.
Ho, C. S. (1969). Geological significance of potassium-argon ages of the Chimei igneous complex
in eastern Taiwan. Bull. Geol. Surv. Taiwan, 20, 63-74.
Ho, C. S. (1986). A synthesis of the geologic evolution of Taiwan. Tectonophysics, 125(1-3), 1-16.
Hobson, D. M. (1986). A thin skinned model for the Papuan thrust belt and some implications
for hydrocarbon exploration. The APPEA Journal, 26(1), 214-225.
Hollings, P., Wolfe, R., Cooke, D. R., & Waters, P. J. (2011). Geochemistry of Tertiary igneous
rocks of northern Luzon, Philippines: Evidence for a back-arc setting for alkalic
porphyry copper-gold deposits and a case for slab roll-back?. Economic Geology, 106(8),
1257-1277.
Hollis, C. J., & Hanson, J. A. (1991). Well-preserved late Paleocene Radiolaria from Tangihua
Complex, Camp Bay, eastern Northland. Tane, 33, 65-76.
Holm, R. J., Rosenbaum, G., & Richards, S. W. (2016). Post 8 Ma reconstruction of Papua New
Guinea and Solomon Islands: Microplate tectonics in a convergent plate boundary
setting. Earth-Science Reviews, 156, 66-81.
Holm, R. J., Spandler, C., & Richards, S. W. (2015). Continental collision, orogenesis and arc
magmatism of the Miocene Maramuni arc, Papua New Guinea. Gondwana
Research, 28(3), 1117-1136.
Home, P. C., Dalton, D. G., & Brannan, J. (1990). Geological evolution of the western Papuan
basin. Proceedings of the First PNG Petroleum Convention, 107-117.
Honza, E., & Fujioka, K. (2004). Formation of arcs and backarc basins inferred from the tectonic
evolution of Southeast Asia since the Late Cretaceous. Tectonophysics, 384(1-4), 23-53.
Hou, Y., Shao, L., Cui, Y., Allen, M. B., Zhu, W., Qiao, P., ... & Goh, T. L. (2022). Sediment features
and provenance analysis of the late Mesozoic–early Cenozoic strata of the Ryukyu
Islands: Implications for palaeogeography of East China Sea. Marine and Petroleum
Geology, 145, 105840.
Howell, D. G. (1980). Mesozoic accretion of exotic terranes along the New Zealand segment of
Gondwanaland. Geology, 8(10), 487-491.
Huang, C. Y., Chen, W. H., Wang, M. H., Lin, C. T., Yang, S., Li, X., ... & Harris, R. (2018). Juxtaposed
sequence stratigraphy, temporal-spatial variations of sedimentation and development of
modern-forming forearc Lichi Mélange in North Luzon Trough forearc basin onshore
and offshore eastern Taiwan: An overview. Earth-science reviews, 182, 102-140.
Huang, C. Y., Wang, P., Yu, M., You, C. F., Liu, C. S., Zhao, X., ... & Yumul Jr, G. P. (2019). Potential
role of strike-slip faults in opening up the South China Sea. National Science Review, 6(5),
891-901.
Huang, C. Y., Yuan, P. B., & Tsao, S. J. (2006). Temporal and spatial records of active arc-
continent collision in Taiwan: A synthesis. Geological Society of America Bulletin, 118(3-
4), 274-288.
Hutchison, C. S., Bergman, S. C., Swauger, D. A., & Graves, J. E. (2000). A Miocene collisional belt
in north Borneo: uplift mechanism and isostatic adjustment quantified by
thermochronology. Journal of the Geological Society, 157(4), 783-793.
Hutchison, D. S. (1975). Basement geology of the North Sepik region, Papua New Guinea. Bureau
of Mineral Resources, Australia, Record, 162.
Hutchison, D. S., & Norvick, M. (1978). Wewak, Papua New Guinea 1: 250,000 geological series.
Geological Survey of Papua New Guinea, Dept. of Minerals and Energy, Port Moresby.
Hutchison, D. S., & Norvick, M. (1980). Geology of the North Sepik region, Papua New Guinea.
Bureau of Mineral Resources.
Imai, A. (2001). Generation and evolution of ore fluids for porphyry Cu-Au mineralization of the
Santo Tomas II (Philex) deposit, Philippines. Resource Geology, 51(2), 71-96.
Ishizuka, O., Hickey-Vargas, R., Arculus, R. J., Yogodzinski, G. M., Savov, I. P., Kusano, Y., ... & Sudo,
M. (2018). Age of Izu–Bonin–Mariana arc basement. Earth and Planetary Science
Letters, 481, 80-90.
Ishizuka, O., Kimura, J. I., Li, Y. B., Stern, R. J., Reagan, M. K., Taylor, R. N., ... & Haraguchi, S.
(2006). Early stages in the evolution of Izu–Bonin arc volcanism: New age, chemical, and
isotopic constraints. Earth and Planetary Science Letters, 250(1-2), 385-401.
Ishizuka, O., Tani, K., Harigane, Y., Yamazaki, T., & Ohara, Y. (2015). Geologic and
geochronological constraints on the Philippines Sea tectonics around 50 Ma. JpGU 2015
Abstract.
Ishizuka, O., Tani, K., Reagan, M. K., Kanayama, K., Umino, S., Harigane, Y., ... & Dunkley, D. J.
(2011a). The timescales of subduction initiation and subsequent evolution of an oceanic
island arc. Earth and Planetary Science Letters, 306(3-4), 229-240.
Ishizuka, O., Tani, K., Taylor, R. N., Umino, S., Sakamoto, I., Yokoyama, Y., ... & Sekimoto, S. (2022).
Origin and age of magmatism in the northern Philippine Sea basins. Geochemistry,
Geophysics, Geosystems, 23(4).
Ishizuka, O., Taylor, R. N., Ohara, Y., & Yuasa, M. (2013). Upwelling, rifting, and age-progressive
magmatism from the Oki-Daito mantle plume. Geology, 41(9), 1011-1014.
Ishizuka, O., Taylor, R. N., Yuasa, M., & Ohara, Y. (2011b). Making and breaking an island arc: A
new perspective from the Oligocene Kyushu-Palau arc, Philippine Sea. Geochemistry,
Geophysics, Geosystems, 12(5).
Ishizuka, O., Yuasa, M., Tamura, Y., Shukuno, H., Stern, R. J., Naka, J., ... & Taylor, R. N. (2010).
Migrating shoshonitic magmatism tracks Izu–Bonin–Mariana intra-oceanic arc rift
propagation. Earth and Planetary Science Letters, 294(1-2), 111-122.
Isozaki, Y., Aoki, K., Nakama, T., & Yanai, S. (2010). New insight into a subduction-related
orogen: A reappraisal of the geotectonic framework and evolution of the Japanese
Islands. Gondwana Research, 18(1), 82-105.
Isozaki, Y., Maruyama, S., & Furuoka, F. (1990). Accreted oceanic materials in Japan.
Tectonophysics, 181(1-4), 179-205.
Ito, T., Kojima, Y., Kodaira, S., Sato, H., Kaneda, Y., Iwasaki, T., ... & Ikawa, T. (2009). Crustal
structure of southwest Japan, revealed by the integrated seismic experiment Southwest
Japan 2002. Tectonophysics, 472(1-4), 124-134.
Jahn, B. M. (1986). Mid-ocean ridge or marginal basin origin of the East Taiwan Ophiolite:
chemical and isotopic evidence. Contributions to Mineralogy and Petrology, 92(2), 194-
206.
Jahn, B. M., Chi, W. R., & Yui, T. F. (1992). A Late Permian formation of Taiwan (marbles from
Chia-Li well No. 1): Pb-Pb isochron and Sr isotopic evidence, and its regional geological
significance. Journal of the Geological Society China, 35, 193-218.
Jahn, B. M., Zhou, X. H., & Li, J. L. (1990). Formation and tectonic evolution of southeastern China
and Taiwan: isotopic and geochemical constraints. Tectonophysics, 183(1-4), 145-160.
Jaques, A. L. (1976). High-K2O island-arc volcanic rocks from the Finisterre and Adelbert
Ranges, northern Papua New Guinea. Geological Society of America Bulletin, 87(6), 861-
867.
Jaques, A. L. (1981). Petrology and petrogenesis of cumulate peridotites and gabbros from the
Marum ophiolite complex, northern Papua New Guinea. Journal of Petrology, 22(1), 1-40.
Jaques, A. L., & BW, C. (1978). Geochemistry of LIL-element enriched tholeiites from the Marum
ophiolite complex, northern Papua New Guinea. Australian Journal of Geology and
Geophysics 3(4), 297-310.
Jaques, A. L., & Chappell, B. W. (1980). Petrology and trace element geochemistry of the Papuan
ultramafic belt. Contributions to Mineralogy and Petrology, 75(1), 55-70.
Jaques, A. L., & Robinson, G.P. (1976). Madang - Papua New Guinea 1: 250,000 geological series.
Geological Survey of Papua New Guinea, Dept. of Minerals and Energy, Port Moresby.
Jaques, A. L., & Robinson, G.P. (1980). Bogia - Papua New Guinea 1: 250,000 geological series.
Geological Survey of Papua New Guinea, Dept. of Minerals and Energy, Port Moresby.
Jaques, A. L., Chappell, B. W., & Taylor, S. R. (1983). Geochemistry of cumulus peridotites and
gabbros from the Marum Ophiolite Complex, northern Papua New Guinea. Contributions
to Mineralogy and Petrology, 82, 154-164.
Johnson, T. L. (1979). Alternative model for emplacement of the Papuan ophiolite, Papua New
Guinea. Geology, 7(10), 495-498.
Jolivet, L., Tamaki, K., & Fournier, M. (1994). Japan Sea, opening history and mechanism: A
synthesis. Journal of Geophysical Research: Solid Earth, 99(B11), 22237-22259.
Joseph, L. E., & Finlayson, E. J. (1991). A revised stratigraphy of Muyua (Woodlark Island).
Geological Survey of Papua New Guinea.
Joshima, M., Okuda, Y., Murakami, F., Kishimoto, K., & Honza, E. (1986). Age of the Solomon Sea
Basin from magnetic lineations. Geo-Marine Letters, 6, 229-234.
Jost, B. M., Webb, M., & White, L. T. (2018). The Mesozoic and Palaeozoic granitoids of north-
western New Guinea. Lithos, 312, 223-243.
Karig, D. E. (1971). Structural history of the Mariana island arc system. Geological Society of
America Bulletin, 82(2), 323-344.
Karig, D. E. (1983). Accreted terranes in the northern part of the Philippine archipelago.
Tectonics, 2(2), 211-236.
Keenan, T. E., Encarnación, J., Buchwaldt, R., Fernandez, D., Mattinson, J., Rasoazanamparany, C.,
& Luetkemeyer, P. B. (2016). Rapid conversion of an oceanic spreading center to a
subduction zone inferred from high-precision geochronology. Proceedings of the
National Academy of Sciences, 113(47), E7359-E7366.
Kiminami, K., Miyashita, S., & Kawabata, K. (1994). Ridge collision and in situ greenstones in
accretionary complexes: An example from the Late Cretaceous Ryukyu Islands and
southwest Japan margin. Island Arc, 3(2), 103-111.
Kimura, G., Hashimoto, Y., Kitamura, Y., Yamaguchi, A., & Koge, H. (2014). Middle Miocene swift
migration of the TTT triple junction and rapid crustal growth in southwest Japan: A
review. Tectonics, 33(7), 1219-1238.
Kizaki, K. (1986). Geology and tectonics of the Ryukyu Islands. Tectonophysics, 125(1-3), 193-
207.
Knittel, U. (1983). Age of the Cordon Syenite Complex and its implication on the Mid-Tertiary
history of North Luzon. Philippine Geol, 37(2), 22-31.
Knittel, U., Walia, M., Suzuki, S., Dimalanta, C. B., Tamayo, R., Yang, T. F., & Yumul Jr, G. P. (2017).
Diverse protolith ages for the Mindoro and Romblon Metamorphics (Philippines):
Evidence from single zircon U–Pb dating. Island Arc, 26(1), e12160.
Kobayashi, K., Fujioka, K., Fujiwara, T., Iwabuchi, Y., & Kitazato, H. (1997). Why is the Palau
Trench so deep? Deep-sea trench without plate convergence. Proceedings of the Japan
Academy, Series B, 73(6), 89-94.
Koulali, A., Tregoning, P., McClusky, S., Stanaway, R., Wallace, L., & Lister, G. (2015). New Insights
into the present-day kinematics of the central and western Papua New Guinea from
GPS. Geophysical Journal International, 202(2), 993-1004.
Kravchinsky, V. A., Cogné, J. P., Harbert, W. P., & Kuzmin, M. I. (2002). Evolution of the Mongol-
Okhotsk Ocean as constrained by new palaeomagnetic data from the Mongol-Okhotsk
suture zone, Siberia. Geophysical Journal International, 148(1), 34-57.
Kroenke, L. (1981). Site 451: East edge of the West Mariana Ridge. Initial Report of the Deep Sea
Drilling Project, 59, 405-483.
Kroenke, L. W., & Eade, J. V. (1982). Three Kings Ridge: A west-facing arc. Geo-Marine
Letters, 2(1-2), 5–10.
Kroenke, L. W., Resig, J. M., & Cooper, P. A. (1986). Tectonics of the Southeastern Solomon
Islands: Formation of the Malaita Anticlinorium. Geology and Offshore Resources of
Pacific Island Arcs–central and western Solomon Islands, 4, 109-116.
Kumagai, H., Kaneoka, I., & Ishii, T. (1996). The active period of the Ayu Trough estimated from
K-Ar ages: The southeastern spreading center of Philippine Sea Plate. Geochemical
Journal, 30(2), 81-87.
Lafoy, Y., Géli, L., Klingelhoefer, F., Vially, R., Sichler, B., & Nouzé, H. (2005). Discovery of
continental stretching and oceanic spreading in the Tasman Sea. Eos, Transactions
American Geophysical Union, 86(10), 101-105.
Lai, Y. M., Song, S. R., Lo, C. H., Lin, T. H., Chu, M. F., & Chung, S. L. (2017). Age, geochemical and
isotopic variations in volcanic rocks from the Coastal Range of Taiwan: Implications for
magma generation in the Northern Luzon Arc. Lithos, 272, 92-115.
Lallemand, S., & Arcay, D. (2021). Subduction initiation from the earliest stages to self-sustained
subduction: Insights from the analysis of 70 Cenozoic sites. Earth-Science Reviews, 221,
103779.
Landmesser, C., Andrews, J. & Packham, G. (1973). Aspects of the geology of the eastern Coral
Sea and the western New Hebrides Basin. Initial Reports of the Deep Sea Drilling Project,
30, 647-662
Lapierre, H., Jahn, B. M., Charvet, J., & Yu, Y. W. (1997). Mesozoic felsic arc magmatism and
continental olivine tholeiites in Zhejiang Province and their relationship with the tectonic
activity in southeastern China. Tectonophysics, 274(4), 321-338.
Lapouille, A. (1982). Etude des bassins marginaux fossiles du Sud-Ouest Pacifique: bassin Nord
d'Entrecasteaux, bassin Nord-Loyaute, bassin Sud-Fidjien, Contribution a l'etude
geodynamique du Sud-Ouest Pacifique, Travaux et Documents de l'ORSTOM,147, 409-438.
Larson, R. L., & Chase, C. G. (1972). Late Mesozoic evolution of the western Pacific Ocean.
Geological Society of America Bulletin, 83(12), 3627-3644.
Launay, J., Dupont, J., & Lapouille, A. (1982). The Three Kings Ridge and the Norfolk Basin
(southwest Pacific): An attempt at structural interpretation. South Pacific Marine
Geological Notes, 2, 121–130.
Le Dain, A. Y., Tapponnier, P., & Molnar, P. (1984). Active faulting and tectonics of Burma and
surrounding regions. Journal of Geophysical Research: Solid Earth, 89(B1), 453-472.
Lebrun, J. F., Lamarche, G., Collot, J. Y., & Delteil, J. (2000). Abrupt strike-slip fault to subduction
transition: The Alpine fault-Puysegur trench connection, New Zealand. Tectonics, 19(4),
688-706.
Lee, C. S., Shor Jr, G. G., Bibee, L. D., Lu, R. S., & Hilde, T. W. (1980). Okinawa Trough: Origin of a
back-arc basin. Marine Geology, 35(1-3), 219-241.
Lee, T. Y., & Lawver, L. A. (1994). Cenozoic plate reconstruction of the South China Sea region.
Tectonophysics, 235(1-2), 149-180.
Li, C. F., Xu, X., Lin, J., Sun, Z., Zhu, J., Yao, Y., ... & Zhang, G. L. (2014). Ages and magnetic
structures of the South China Sea constrained by deep tow magnetic surveys and IODP
Expedition 349. Geochemistry, Geophysics, Geosystems, 15(12), 4958-4983.
Li, J., Zhang, Y., Dong, S., & Johnston, S. T. (2014). Cretaceous tectonic evolution of South China: A
preliminary synthesis. Earth-Science Reviews, 134, 98-136.
Li, Z. X., Li, X. H., Chung, S. L., Lo, C. H., Xu, X., & Li, W. X. (2012). Magmatic switch-on and switch-
off along the South China continental margin since the Permian: Transition from an
Andean-type to a Western Pacific-type plate boundary. Tectonophysics, 532, 271-290.
Lin, C. T., Harris, R., Sun, W. D., & Zhang, G. L. (2019). Geochemical and geochronological
constraints on the origin and emplacement of the East Taiwan Ophiolite. Geochemistry,
Geophysics, Geosystems, 20(4), 2110-2133.
Lindley, D. (1988). Early Cainozoic stratigraphy and structure of the Gazelle Peninsula, east New
Britain: An example of extensional tectonics in the New Britain arc-trench
complex. Journal of the Geological Society of Australia, 35(2), 231-244.
Lindley, I. D. (2006). New Britain Trench, Papua New Guinea: An extensional element in a
regional sinistral strike-slip system. New Concepts in Global Tectonics Newsletter, 41, 15-
27.
Lindley, I. D. (2016). Plate flexure and volcanism: late cenozoic tectonics of the tabar–lihir–
tanga–feni alkalic province, new Ireland basin, Papua New Guinea. Tectonophysics, 677,
312-323.
Ling, H. Y., Hall, R., & Nichols, G. J. (1991). Early Eocene Radiolaria from Waigeo Island, Eastern
Indonesia. Journal of Southeast Asian Earth Sciences, 6(3-4), 299-305.
Little, T. A., Hacker, B. R., Gordon, S. M., Baldwin, S. L., Fitzgerald, P. G., Ellis, S., & Korchinski, M.
(2011). Diapiric exhumation of Earth's youngest (UHP) eclogites in the gneiss domes of
the D'Entrecasteaux Islands, Papua New Guinea. Tectonophysics, 510(1-2), 39-68.
Liu, B., Li, S. Z., Suo, Y. H., Li, G. X., Dai, L. M., Somerville, I. D., ... & Yu, S. (2016). The geological
nature and geodynamics of the Okinawa Trough, Western Pacific. Geological Journal, 51,
416-428.
Liu, B., Wu, J. H., Li, H., Wu, Q. H., Evans, N. J., Kong, H., & Xi, X. S. (2020). Geochronology,
geochemistry and petrogenesis of the Dengfuxian lamprophyres: Implications for the
early Cretaceous tectonic evolution of the South China Block. Geochemistry, 80(2),
125598.
Liu, J., Li, S., Cao, X., Dong, H., Suo, Y., Jiang, Z., Zhou, J., Li., X., Zhang, R., Liu, L., & Foulger, G. R.
(2023). Back-Arc Tectonics and Plate Reconstruction of the Philippine Sea-South China
Sea Region Since the Eocene. Geophysical Research Letters, 50(5), e2022GL102154.
Lo, C. H., Onstott, T. C., Chen, C. H., & Lee, T. (1994). An assessment of 40Ar/39Ar dating for the
whole-rock volcanic samples from the Luzon Arc near Taiwan. Chemical Geology, 114(1-
2), 157-178.
Lo, Y. C., Chih-Tung, C., Lo, C. H., & Sun-Lin, C. (2020). Ages of ophiolitic rocks along plate suture
in Taiwan orogen: Fate of the South China Sea from subduction to collision. TAO:
Terrestrial, Atmospheric and Oceanic Sciences, 31(4), 3.
Lus, W. Y., McDougall, I., & Davies, H. L. (2004). Age of the metamorphic sole of the Papuan
Ultramafic Belt ophiolite, Papua New Guinea. Tectonophysics, 392(1-4), 85-101.
Maac, Y. O., & Ylade, E. D. (1988). Stratigraphic and paleontologic studies of Tablas,
Romblon. Report of Research and Development Cooperation ITIT Project, (8319), 44-67.
Maffione, M., van Hinsbergen, D. J., de Gelder, G. I., van der Goes, F. C., & Morris, A. (2017).
Kinematics of Late Cretaceous subduction initiation in the Neo-Tethys Ocean
reconstructed from ophiolites of Turkey, Cyprus, and Syria. Journal of Geophysical
Research: Solid Earth, 122(5), 3953-3976.
Mahoney, J. J., Storey, M., Duncan, R. A., Spencer, K. J., & Pringle, M. (1993). Geochemistry and age
of the Ontong Java Plateau. The Mesozoic Pacific: Geology, Tectonics, and Volcanism.
Geophysical Monograph Series, 77, 233-261.
Malahoff, A., Feden, R. H., & Fleming, H. S. (1982). Magnetic anomalies and tectonic fabric of
marginal basins north of New Zealand. Journal of Geophysical Research, 87(B5), 4109–
4125.
Malahoff, A., Kroenke, L. W., Cherkis, N., & Brozena, J. (1994). Magnetic and tectonic fabric of the
North Fiji Basin and Lau Basin. Basin Formation, Ridge Crest Processes, and
Metallogenesis in the North Fiji Basin, 49-61.
Malaihollo, J. F., & Hall, R. (1996). The geology and tectonic evolution of the Bacan region, east
Indonesia. Geological Society, London, Special Publications, 106(1), 483-497.
Maleterre, P. (1988). The Southern Central Cordillera of Luzon; a multistage Upper Eocene to
Pleistocene arc-deformed on the northern end of the Philippine strike-slip fault.
In International Symposium on the Geodynamic Evolution of Eastern Eurasian Margin
(Abstracts), Paris 13-20 September 1988.
Manalo, P. C., Dimalanta, C. B., Faustino-Eslava, D. V., Payot, B. D., Ramos, N. T., Queaño, K. L., ... &
Yumul Jr, G. P. (2015). Geochemical and Geophysical Characteristics of the Balud
Ophiolitic Complex (BOC), Masbate Island, Philippines: Implications for its Generation,
Evolution and Emplacement. Terrestrial, Atmospheric & Oceanic Sciences, 26(6).
Marchadier, Y., & Rangin, C. (1990). Polyphase tectonics at the southern tip of the Manila trench,
Mindoro-Tablas Islands, Philippines. Tectonophysics, 183(1-4), 273-287.
Martin, A. K. (2011). Double saloon door tectonics in the Japan Sea, Fossa magna, and the
Japanese Island arc. Tectonophysics, 498(1-4), 45-65.
Martin, P. E., Macdonald, F. A., McQuarrie, N., Flowers, R. M., Weiland, R. J., Maffre, P. J. Y. (2023).
The rise of New Guinea and the fall of Neogene global temperatures. Proceedings of the
National Academy of Sciences, in press.
Matthews, K. J., Müller, R. D., Wessel, P., & Whittaker, J. M. (2011). The tectonic fabric of the
ocean basins. Journal of Geophysical Research: Solid Earth, 116(B12).
Maurizot, P., Cluzel, D., Meffre, S., Campbell, H. J., Collot, J., & Sevin, B. (2020a). Pre-Late
Cretaceous basement terranes of the Gondwana active margin of New
Caledonia. Geological Society, London, Memoirs, 51(1), 27-52.
Maurizot, P., Cluzel, D., Patriat, M., Collot, J., Iseppi, M., Lesimple, S., ... & Davies, H. L. (2020b).
The Eocene subduction–obduction complex of New Caledonia. Geological Society,
London, Memoirs, 51(1), 93-130.
McCabe, R., Almasco, J., & Diegor, W. (1982). Geologic and paleomagnetic evidence for a possible
Miocene collision in western Panay, central Philippines. Geology, 10(6), 325-329.
Mccaffrey, R., Silver, E. A., & Raitt, R. W. (1980). Crustal structure of the Molucca Sea collision
zone, Indonesia. In The tectonic and geologic evolution of Southeast Asian seas and
islands, Geophysical Monograph 23, 161-177.
McCarthy, A., Magri, L., Sauermilch, I., Fox, J., Seton, M., Mohn, G., ... & Whittaker, J. M. (2022). The
Louisiade ophiolite: a missing link in the western Pacific. Terra Nova, 34(2), 146-154.
Meijer, A., Reagan, M., Ellis, H., Shafiqullah, M., Sutter, J., Damon, P., & Kling, S. (1983).
Chronology of volcanic events in the eastern Philippine Sea. Washington DC American
Geophysical Union Geophysical Monograph Series, 27, 349-359.
Mesalles, L., Walia, M., Lee, H., & Lee, Y. H. (2018, December). Cenozoic evolution of the Panay
Island, Philippines: magmatic chronology and arc-continent collision. In AGU Fall
Meeting Abstracts (Vol. 2018, pp. T23A-0343).
Metcalfe, I. (2013). Gondwana dispersion and Asian accretion: Tectonic and palaeogeographic
evolution of eastern Tethys. Journal of Asian Earth Sciences, 66, 1-33.
Miki, M. (1995). Two-phase opening model for the Okinawa Trough inferred from
paleomagnetic study of the Ryukyu arc. Journal of Geophysical Research: Solid
Earth, 100(B5), 8169-8184.
Miller, M. S., Kennett, B. L. N., & Lister, G. S. (2004). Imaging changes in morphology, geometry,
and physical properties of the subducting Pacific plate along the Izu–Bonin–Mariana
arc. Earth and Planetary Science Letters, 224(3-4), 363-370.
Mitchell, A. H. G., & Warden, A. J. (1971). Geological evolution of the New Hebrides island
arc. Journal of the Geological Society, 127(5), 501-529.
Mitchell, A. H. G., Hernandez, F. T., & Dela Cruz, A. P. (1986). Cenozoic evolution of the Philippine
Archipelago. Journal of Southeast Asian Earth Sciences, 1(1), 3-22.
MMAJ-JICA (1977). Report on geological survey of northeastern Luzon. Phase III: Metal Mining
Agency of Japan and Japan International Cooperation Agency, Tokyo.
MMAJ-JICA. (1986). Report on Mineral Exploration, Mineral Deposits and Tectonics of Two
Contrasting Geologic Environments in the Republic of the Philippines, Phase II (Masbate
and Leyte Areas).
MMAJ-JICA. (1987). Report on the mineral exploration, mineral deposits and tectonics of two
contrasting geological environments in the Philippines, Phase 3 (part1), northern Sierra
Madre area. Government of Japan, 403.
Monnier, C., Girardeau, J., Maury, R. C., & Cotten, J. (1995). Back-arc basin origin for the East
Sulawesi ophiolite (eastern Indonesia). Geology, 23(9), 851-854.
Monnier, C., Girardeau, J., Pubellier, M., Polvé, M., Permana, H., & Bellon, H. (1999). Petrology
and geochemistry of the Cyclops ophiolites (Irian Jaya, East Indonesia): consequences
for the Cenozoic evolution of the north Australian margin. Mineralogy and
Petrology, 65(1-2), 1-28.
Monteleone, B. D., Baldwin, S. L., Ireland, T. R., & Fitzgerald, P. G. (2001). Thermochronologic
constraints for the tectonic evolution of the Moresby seamount, Woodlark Basin, Papua
New Guinea. In Proc. ODP, Sci. Results (Vol. 180, pp. 1-34). Ocean Drilling Program.
Monteleone, B. D., Baldwin, S. L., Webb, L. E., Fitzgerald, P. G., Grove, M., & Schmitt, A. K. (2007).
Late Miocene–Pliocene eclogite facies metamorphism, D'Entrecasteaux Islands, SE
Papua New Guinea. Journal of Metamorphic Geology, 25(2), 245-265.
Moore, G. F., Kadarisman, D., Evans, C. A., & Hawkins, J. W. (1981). Geology of the Talaud islands,
Molucca sea collision zone, northeast Indonesia. Journal of Structural Geology, 3(4), 467-
475.
Morishita, T., Tani, K. I., Soda, Y., Tamura, A., Mizukami, T., & Ghosh, B. (2018). The uppermost
mantle section below a remnant proto-Philippine Sea island arc: Insights from the
peridotite fragments from the Daito Ridge. American Mineralogist: Journal of Earth and
Planetary Materials, 103(7), 1151-1160.
Morrice, M. G., Jezek, P. A., Gill, J. B., Whitford, D. J., & Monoarfa, M. (1983). An introduction to the
Sangihe arc: Volcanism accompanying arc—arc collision in the Molucca Sea,
Indonesia. Journal of Volcanology and Geothermal Research, 19(1-2), 135-165.
Mortimer, N. (2004). New Zealand's geological foundations. Gondwana research, 7(1), 261-272.
Mortimer, N., Campbell, H. J., Tulloch, A. J., King, P. R., Stagpoole, V. M., Wood, R. A., ... & Seton, M.
(2017). Zealandia: Earth’s hidden continent. GSA today, 27(3), 27-35.
Mortimer, N., Gans, P. B., Palin, J. M., Herzer, R. H., Pelletier, B., & Monzier, M. (2014a). Eocene
and Oligocene basins and ridges of the Coral Sea-New Caledonia region: Tectonic link
between Melanesia, Fiji, and Zealandia. Tectonics, 33(7), 1386-1407.
Mortimer, N., Herzer, R. H., Gans, P. B., Laporte-Magoni, C., Calvert, A. T., & 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 Geology,
237(1-2), 1–24.
Mortimer, N., Herzer, R. H., Gans, P. B., Parkinson, D. L., & Seward, D. (1998). Basement geology
from Three Kings Ridge to West Norfolk Ridge, southwest Pacific Ocean: evidence from
petrology, geochemistry and isotopic dating of dredge samples. Marine Geology, 148(3-
4), 135-162.
Mortimer, N., Rattenbury, M. S., King, P. R., Bland, K. J., Barrell, D. J. A., Bache, F., ... & Turnbull, R.
E. (2014b). High-level stratigraphic scheme for New Zealand rocks. New Zealand Journal
of Geology and Geophysics, 57(4), 402-419.
Mouthereau, F., & Lacombe, O. (2006). Inversion of the Paleogene Chinese continental margin
and thick-skinned deformation in the Western Foreland of Taiwan. Journal of Structural
Geology, 28(11), 1977-1993.
Mouthereau, F., Lacombe, O., Deffontaines, B., Angelier, J., & Brusset, S. (2001). Deformation
history of the southwestern Taiwan foreland thrust belt: insights from tectono-
sedimentary analyses and balanced cross-sections. Tectonophysics, 333(1-2), 293-322.
Mueller, C. O., & Jokat, W. (2019). The initial Gondwana break-up: a synthesis based on new
potential field data of the Africa-Antarctica Corridor. Tectonophysics, 750, 301-328.
Mulcahy, S. R., Starnes, J. K., Day, H. W., Coble, M. A., & Vervoort, J. D. (2018). Early onset of
Franciscan subduction. Tectonics, 37(5), 1194-1209.
Müller, R. D., Cannon, J., Qin, X., Watson, R. J., Gurnis, M., Williams, S., ... & Zahirovic, S. (2018).
GPlates: building a virtual Earth through deep time. Geochemistry, Geophysics,
Geosystems, 19(7), 2243-2261.
Müller, R. D., Royer, J. Y., Cande, S. C., Roest, W. R., & Maschenkov, S. (1999). New constraints on
the Late Cretaceous/Tertiary plate tectonic evolution of the Caribbean. Sedimentary
basins of the world, 4, 33-59.
Müller, R. D., Zahirovic, S., Williams, S. E., Cannon, J., Seton, M., Bower, D. J., Tetley, M. G., Heine,
C., Le Breton, E., Liu, S., Russel, S. H. J., Yang, T., Leonard, J. & Gurnis, M. (2019). A global
plate model including lithospheric deformation along major rifts and orogens since the
Triassic. Tectonics, 38(6), 1884-1907.
Nakanishi, M., & Winterer, E. L. (1998). Tectonic history of the Pacific-Farallon-Phoenix triple
junction from Late Jurassic to Early Cretaceous: an abandoned Mesozoic spreading
system in the central Pacific basin. Journal of Geophysical Research: Solid Earth, 103(B6),
12453-12468.
Nakanishi, M., Tamaki, K., & Kobayashi, K. (1992). Magnetic anomaly lineations from Late
Jurassic to Early Cretaceous in the west-central Pacific Ocean. Geophysical Journal
International, 109(3), 701-719.
Neef, G. & McDougall I. (1976) Potassium-argon ages on rocks from Small Nggela Island, British
Solomon Islands, SW Pacific, Pacific Geology (11), 81-86.
Nichols, G., & Hall, R. (1999). History of the Celebes Sea Basin based on its stratigraphic and
sedimentological record. Journal of Asian Earth Sciences, 17(1-2), 47-59.
Nishimura, T. (2011). Back-arc spreading of the northern Izu–Ogasawara (Bonin) Islands arc
clarified by GPS data. Tectonophysics, 512(1-4), 60-67.
Nong, A. T., Hauzenberger, C. A., Gallhofer, D., & Dinh, S. Q. (2021). Geochemistry and zircon UPb
geochronology of Late Mesozoic igneous rocks from SW Vietnam–SE Cambodia:
Implications for episodic magmatism in the context of the Paleo-Pacific
subduction. Lithos, 390, 106101.
Nong, A. T., Hauzenberger, C. A., Gallhofer, D., Skrzypek, E., & Dinh, S. Q. (2022). Geochemical and
zircon U-Pb geochronological constraints on late mesozoic Paleo-Pacific subduction-
related volcanism in southern Vietnam. Mineralogy and Petrology, 116(5), 349-368.
O'Neill, C., Müller, D., & Steinberger, B. (2005). On the uncertainties in hot spot reconstructions
and the significance of moving hot spot reference frames. Geochemistry, Geophysics,
Geosystems, 6(4).
Ogg, J. G. (2020). Geomagnetic polarity time scale. In Geologic time scale 2020 (pp. 159-192).
Elsevier.
Okada, A. (1973). On the quaternary faulting along the median tectonic line. The Median
Tectonic Line, 46, 49–86.
Olfindo, V. S. V., Payot, B. D., Valera, G. T. V., Gadot Jr, E. G., Villaplaza, B. R. B., Tani, K., ... & Yumul
Jr, G. P. (2019). Petrographic and geochemical characterization of the crustal section of
the Pujada Ophiolite, southeastern Mindanao, Philippines: Insights to the tectonic
evolution of the northern Molucca Sea Collision Complex. Journal of Asian Earth
Sciences, 184, 103994.
Österle, J. E., Little, T. A., Seward, D., Stockli, D. F., & Gamble, J. (2020). The petrology,
geochronology and tectono-magmatic setting of igneous rocks in the Suckling-Dayman
metamorphic core complex, Papua New Guinea. Gondwana Research, 83, 390-414.
Ota, T., & Kaneko, Y. (2010). Blueschists, eclogites, and subduction zone tectonics: Insights from
a review of Late Miocene blueschists and eclogites, and related young high-pressure
metamorphic rocks. Gondwana Research, 18, 167-188.
Ott, B., & Mann, P. (2015). Late Miocene to Recent formation of the Aure-Moresby fold-thrust
belt and foreland basin as a consequence of Woodlark microplate rotation, Papua New
Guinea. Geochemistry, Geophysics, Geosystems, 16(6), 1988-2004.
Pacle, N. A. D., Dimalanta, C. B., Ramos, N. T., Payot, B. D., Faustino-Eslava, D. V., Queaño, K. L., &
Yumul Jr, G. P. (2017). Petrography and geochemistry of Cenozoic sedimentary
sequences of the southern Samar Island, Philippines: clues to the unroofing history of an
ancient subduction zone. Journal of Asian Earth Sciences, 142, 3-19.
Page, R. W. (1976). Geochronology of Igneous and Metamorphic Rocks in the New Guinea
Highlands. Bureau of Mineral Resources, Canberra, Australia.
Pain, C. F. (1983). Volcanic rocks and surfaces as indicators of landform age: The Astrolabe
Agglomerate, Papua New Guinea. Australian Geographer, 15(6), 376-381.
Paquette, J. L., & Cluzel, D. (2007). U–Pb zircon dating of post-obduction volcanic-arc granitoids
and a granulite-facies xenolith from New Caledonia. Inference on Southwest Pacific
geodynamic models. International Journal of Earth Sciences, 96, 613-622.
Parrot, J. F., & Dugas, F. (1980). The disrupted ophiolitic belt of the Southwest Pacific: evidence
of an Eocene subduction zone. Tectonophysics, 66(4), 349-372.
Parsons, A. J., Sigloch, K., & Hosseini, K. (2021). Australian Plate Subduction is Responsible for
Northward Motion of the India-Asia Collision Zone and∼ 1,000 km Lateral Migration of
the Indian Slab. Geophysical Research Letters, 48(18), e2021GL094904.
Pasco, J. A., Dycoco, J. M. A., Valera, G. T. V., Payot, B. D., Pillejera, J. D. B., Uy, F. A. A. E., ... &
Dimalanta, C. B. (2019). Petrogenesis of ultramafic-mafic clasts in the Dos Hermanos
Mélange, Ilocos Norte: insights to the evolution of western Luzon, Philippines. Journal of
Asian Earth Sciences, 184, 104004.
Patriat, M., Collot, J., Danyushevsky, L., Fabre, M., Meffre, S., Falloon, T., ... & Fournier, M. (2015).
Propagation of back-arc extension into the arc lithosphere in the southern New
Hebrides volcanic arc. Geochemistry, Geophysics, Geosystems, 16(9), 3142-3159.
Patriat, M., Collot, J., Etienne, S., Poli, S., Clerc, C., Mortimer, N., ... & VESPA scientific voyage team.
(2018). New Caledonia obducted Peridotite Nappe: offshore extent and implications for
obduction and postobduction processes. Tectonics, 37(4), 1077-1096.
Patriat, M., Falloon, T., Danyushevsky, L., Collot, J., Jean, M. M., Hoernle, K., ... & Feig, S. T. (2019).
Subduction initiation terranes exposed at the front of a 2 Ma volcanically-active
subduction zone. Earth and Planetary Science Letters, 508, 30-40.
Pelletier, B., & Auzende, J. M. (1996). Geometry and structure of the Vitiaz trench lineament (SW
Pacific). Marine Geophysical Researches, 18, 305-335.
Permana, H. (1995). Etude des ophiolites des Weylands (Irian Jaya): origine et âge de mise en
place. Comparaison avec celles de la Haute Chaîne centrale. Mémoire DEA, Université de
Brest, France.
Petterson, M. G., Babbs, T., Neal, C. R., Mahoney, J. J., Saunders, A. D., Duncan, R. A., ... & Natogga,
D. (1999). Geological–tectonic framework of Solomon Islands, SW Pacific: crustal
accretion and growth within an intra-oceanic setting. Tectonophysics, 301(1-2), 35-60.
Petterson, M. G., Coleman, P. J., Tolia, D., Mahoa, H., & Magu, R. (2009). Application of terrain
modelling of the Solomon Islands, Southwest Pacific, to the metallogenesis and mineral
exploration in composite arc-ocean floor terrain collages. Pacific Minerals in the New
Millennium: Science, Exploration, Mining, and Community, 99-120.
Petterson, M. G., Neal, C. R., Mahoney, J. J., Kroenke, L. W., Saunders, A. D., Babbs, T. L., ... &
McGrail, B. (1997). Structure and deformation of north and central Malaita, Solomon
Islands: tectonic implications for the Ontong Java Plateau-Solomon arc collision, and for
the fate of oceanic plateaus. Tectonophysics, 283(1-4), 1-33.
Phinney, E. J., Mann, P., Coffin, M. F., & Shipley, T. H. (1999). Sequence stratigraphy, structure,
and tectonic history of the southwestern Ontong Java Plateau adjacent to the North
Solomon Trench and Solomon Islands arc. Journal of Geophysical Research: Solid
Earth, 104(B9), 20449-20466.
Phinney, E. J., Mann, P., Coffin, M. F., & Shipley, T. H. (2004). Sequence stratigraphy, structural
style, and age of deformation of the Malaita accretionary prism (Solomon arc–Ontong
Java Plateau convergent zone). Tectonophysics, 389(3-4), 221-246.
Pieters, P. E. (1976). Port Moresby-Kalo-Aroa - Papua New Guinea 1: 250,000 geological series.
Geological Survey of Papua New Guinea, Dept. of Minerals and Energy, Port Moresby.
Pieters, P. E., Hartono, U., & Amri, C. (1989). Geology of the Mar sheet area, Irian Jaya. Geological
Research and Development Centre, Bandung (62pp.).
Pieters, P. E., Pigram, C. J., Trail, D. S., Dow, D. B., Ratman, N., & Sukamto, R. (1983). The
stratigraphy of western Irian Jaya. Proceedings Indonesian Petroleum Association,
Twelfth Annual Convention. 229-261.
Pigram, C. J., & Symonds, P. A. (1991). A review of the timing of the major tectonic events in the
New Guinea Orogen. Journal of Southeast Asian Earth Sciences, 6(3-4), 307-318.
Porth H., Muller C. & Daniels C.H. von (1989). The sedimentary formations of the Visayan Basin,
Philippines. In: PORTH H. & DANIELS C.H. (eds.), On the Geology and Hydrocarbon
Prospects of the Visayan Basin, Philippines.- Geologisches Jahrbuch, Stuttgart, Heft 70, p.
29-87.
Pourteau, A., Scherer, E. E., Schorn, S., Bast, R., Schmidt, A., & Ebert, L. (2019). Thermal evolution
of an ancient subduction interface revealed by Lu–Hf garnet geochronology, Halilbağı
Complex (Anatolia). Geoscience Frontiers, 10(1), 127-148.
Pubellier, M., Bader, A. G., Rangin, C., Deffontaines, B., & Quebral, R. (1999). Upper plate
deformation induced by subduction of a volcanic arc: the Snellius Plateau (Molucca Sea,
Indonesia and Mindanao, Philippines). Tectonophysics, 304(4), 345-368.
Pubellier, M., Monnier, C., Maury, R., & Tamayo, R. (2004). Plate kinematics, origin and tectonic
emplacement of supra-subduction ophiolites in SE Asia. Tectonophysics, 392(1-4), 9-36.
Pubellier, M., Quebral, R., Rangin, C., Deffontaines, B., Muller, C., Butterlin, J., & Manzano, J.
(1991). The Mindanao collision zone: a soft collision event within a continuous Neogene
strike-slip setting. Journal of Southeast Asian Earth Sciences, 6(3-4), 239-248.
Qayyum, A., Lom, N., Advokaat, E. L., Spakman, W., Van Der Meer, D. G., & van Hinsbergen, D. J.
(2022). Subduction and Slab Detachment Under Moving Trenches During Ongoing
India-Asia Convergence. Geochemistry, Geophysics, Geosystems, 23(11), e2022GC010336.
Qian, S., Zhang, X., Wu, J., Lallemand, S., Nichols, A. R., Huang, C., ... & Zhou, H. (2021). First
identification of a Cathaysian continental fragment beneath the Gagua Ridge, Philippine
Sea, and its tectonic implications. Geology, 49(11), 1332-1336.
Qian, X., Yu, Y., Wang, Y., Gan, C., Zhang, Y., & Asis, J. B. (2022). Late Cretaceous Nature of SW
Borneo and Paleo-Pacific Subduction: New Insights from the Granitoids in the Schwaner
Mountains. Lithosphere, 2022(1), 8483732.
Quarles van Ufford, A. & Cloos, M. (2005). Cenozoic tectonics of New Guinea. AAPG bulletin,
89(1), 119-140.
Queaño, K. L. (2005). Upper Miocene to Lower Pliocene Sigaboy formation turbidites, on the
Pujada Peninsula, Mindanao, Philippines: internal structures, composition, depositional
elements and reservoir characteristics. Journal of Asian Earth Sciences, 25(3), 387-402.
Queaño, K. L., Ali, J. R., Aitchison, J. C., Yumul, G. P., Pubellier, M., & Dimalanta, C. B. (2008).
Geochemistry of Cretaceous to Eocene ophiolitic rocks of the Central Cordillera:
Implications for Mesozoic-early Cenozoic evolution of the northern
Philippines. International Geology Review, 50(4), 407-421.
Queaño, K. L., Dimalanta, C. B., Yumul Jr, G. P., Marquez, E. J., Faustino-Eslava, D. V., Suzuki, S., &
Ishida, K. (2017b). Stratigraphic units overlying the Zambales Ophiolite Complex (ZOC)
in Luzon,(Philippines): Tectonostratigraphic significance and regional
implications. Journal of Asian Earth Sciences, 142, 20-31.
Queaño, K. L., Marquez, E. J., Aitchison, J. C., & Ali, J. R. (2013). Radiolarian biostratigraphic data
from the Casiguran Ophiolite, northern Sierra Madre, Luzon, Philippines: stratigraphic
and tectonic implications. Journal of Asian Earth Sciences, 65, 131-142.
Queaño, K. L., Marquez, E. J., Dimalanta, C. B., Aitchison, J. C., Ali, J. R., & Yumul Jr, G. P. (2017a).
Mesozoic radiolarian faunas from the northwest Ilocos Region, Luzon, Philippines and
their tectonic significance. Island Arc, 26(4), e12195.
Queaño, K. L., Yumul Jr, G. P., Marquez, E. J., Gabo-Ratio, J. A., Payot, B. D., & Dimalanta, C. B.
(2020). Consumed tectonic plates in Southeast Asia: Markers from the Mesozoic to early
Cenozoic stratigraphic units in the northern and central Philippines. Journal of Asian
Earth Sciences: X, 4, 100033.
Quebral, R. D., Pubellier, M., & Rangin, C. (1996). The onset of movement on the Philippine Fault
in eastern Mindanao: A transition from a collision to a strike-slip
environment. Tectonics, 15(4), 713-726.
Raimbourg, H., Augier, R., Famin, V., Gadenne, L., Palazzin, G., Yamaguchi, A., & Kimura, G.
(2014). Long-term evolution of an accretionary prism: The case study of the Shimanto
Belt, Kyushu, Japan. Tectonics, 33(6), 936-959.
Rangin, C. (1991). The Philippine Mobile Belt: a complex plate boundary. Journal of Southeast
Asian Earth Sciences, 6(3-4), 209-220.
Rangin, C., Bellon, H., Benard, F., Letouzey, J., Muller, C., & Sanudin, T. A. H. I. R. (1990). Neogene
arc-continent collision in Sabah, northern Borneo (Malaysia). Tectonophysics, 183(1-4),
305-319.
Rangin, C., Dahrin, D., Quebral, R., & Modec Scientific Party. (1996). Collision and strike-slip
faulting in the northern Molucca Sea (Philippines and Indonesia): preliminary results of
a morphotectonic study. Geological Society, London, Special Publications, 106(1), 29-46.
Rangin, C., Spakman, W., Pubellier, M., & Bijwaard, H. (1999). Tomographic and geological
constraints on subduction along the eastern Sundaland continental margin (South-East
Asia). Bulletin de la Société géologique de France, 170(6), 775-788.
Rangin, C., Stephan, J. F., & Muller, C. (1985). Middle Oligocene oceanic crust of South China Sea
jammed into Mindoro collision zone (Philippines). Geology, 13(6), 425-428.
Rangin, C., Stephan, J. F., Butterlin, J., Bellon, H., Muller, C., Chorowicz, J., & Baladad, D. (1991).
Collision néogène d’arcs volcaniques dans le centre des Philippines: stratigraphie et
structure de la chaîne d’Antique (île de Panay). Bulletin Societe Géologique du
France, 162(3), 465-477.
Reagan, M. K., Hanan, B. B., Heizler, M. T., Hartman, B. S., & Hickey-Vargas, R. (2008).
Petrogenesis of volcanic rocks from Saipan and Rota, Mariana Islands, and implications
for the evolution of nascent island arcs. Journal of Petrology, 49(3), 441-464.
Reagan, M. K., Heaton, D. E., Schmitz, M. D., Pearce, J. A., Shervais, J. W., & Koppers, A. A. (2019).
Forearc ages reveal extensive short-lived and rapid seafloor spreading following
subduction initiation. Earth and Planetary Science Letters, 506, 520-529.
Reagan, M. K., Ishizuka, O., Stern, R. J., Kelley, K. A., Ohara, Y., Blichert-Toft, J., ... & Woods, M.
(2010). Fore-arc basalts and subduction initiation in the Izu-Bonin-Mariana
system. Geochemistry, Geophysics, Geosystems, 11(3).
Reagan, M. K., McClelland, W. C., Girard, G., Goff, K. R., Peate, D. W., Ohara, Y., & Stern, R. J. (2013).
The geology of the southern Mariana fore-arc crust: Implications for the scale of Eocene
volcanism in the western Pacific. Earth and Planetary Science Letters, 380, 41-51.
Récy, J. (1977). Fossil subduction zones: examples in the South West Pacific, International
Symposium Geodynamics in S.W. Pacific, 345-356.
Ren, J., Niu, B., Wang, J., Jin, X., Zhao, L., & Liu, R. (2013). Advances in research of Asian geology—
A summary of 1: 5M International Geological Map of Asia project. Journal of Asian Earth
Sciences, 72, 3-11.
Richard, M., Bellon, H., Maury, R., Barrier, E., & Wen-Shing, J. (1986). Miocene to recent calc-
alkalic volcanism in eastern Taiwan: K-Ar ages and petrography. Tectonophysics, 125(1-
3), 87-102.
Richards, J. R., Cooper, J. A., Webb, A. W., & Coleman, P. J. (1966). Potassium–Argon
Measurements of the Age of Basal Schists in the British Solomon Islands. Nature, 211,
1251-1252.
Rickard, M. J., & Williams, I. S. (2013). No zircon U–Pb evidence for a Precambrian component in
the Late Eocene Yavuna trondhjemite, Fiji. Australian Journal of Earth Sciences, 60(4),
521-525.
Ridgway, J., & Coulson, F. I. E. (1987). The Geology of Choiseul and the Shortland Islands, Solomon
Islands (Vol. 8). HM Stationery Office.
Ridley, W. I., Rhodes, J. M., REID, A. M., Jakes, P., Shih, C. Y., & Bass, M. N. (1974). Basalts from leg
6 of the deep-sea drilling project. Journal of petrology, 15(1), 140-159.
Ringenbach, J. C., Stephan, J. F., Maleterre, P., & Bellon, H. (1990). Structure and geological
history of the Lepanto-Cervantes releasing bend on the Abra River fault, Luzon Central
Cordillera, Philippines. Tectonophysics, 183(1-4), 225-241.
Rioux, M., Bowring, S., Kelemen, P., Gordon, S., Miller, R. and Dudás, F. (2013). Tectonic
development of the Samail ophiolite: High-precision U-Pb zircon geochronology and Sm-
Nd isotopic constraints on crustal growth and emplacement. Journal of Geophysical
Research: Solid Earth, 118, 2085-2101.
Rodrigo, J. D., Gabo-Ratio, J. A. S., Queaño, K. L., Fernando, A. G. S., de Silva Jr, L. P., Yonezu, K., &
Zhang, Y. (2020). Geochemistry of the Late Cretaceous Pandan Formation in Cebu Island,
Central Philippines: sediment contributions from the Australian plate margin during the
Mesozoic. The Depositional Record, 6(2), 309-330.
Rodrigo, J., & Schlagintweit, F. (2022). The Lower Cretaceous Tuburan Limestone of Cebu Island,
Philippines: Microfacies, micropalaeontology, biostratigraphy, and palaeogeographic
perspectives. Carnets Geol., 22(14), 661-679.
Rodrigo, J., Tsutsumi, Y., Tani, K., Haga, T., & Aira, J. (2021). Tectonostratigraphy of the basement
complex of Cebu Island, central Philippines: New constraints from petrography,
micropaleontology and geochronology. In The Second International Symposium of the
International Geoscience Programme Project 679, 19.
Rodriguez-Roa, F. A., & Wiltschko, D. V. (2010). Thrust belt architecture of the central and
southern Western Foothills of Taiwan. Geological Society, London, Special
Publications, 348(1), 137-168.
Roeser, H. A. (1991). Age of the crust of the southeast Sulu Sea basin based on magnetic
anomalies and age determined at site 768. In Proceedings of the Ocean Drilling Program,
Scientific Results 124, 339-343.
Rogerson, R. (1987). The geology and mineral resources of the Sepik headwaters region, Papua
New Guinea (Vol. 12). Geological Survey of Papua New Guinea.
Rogerson, R. J., & Hilyard, D. B. (1990). Scrapland: a suspect composite terrane in Papua New
Guinea. Papua New Guinea (PNG) Petroleum Convention Proceedings.
Roser, B. P., Coombs, D. S., Korsch, R. J., & Campbell, J. D. (2002). Whole-rock geochemical
variations and evolution of the arc-derived Murihiku Terrane, New Zealand. Geological
Magazine, 139(6), 665-685.
Ruellan, E., Delteil, J., Wright, I., & Matsumoto, T. (2003). From rifting to active spreading in the
Lau Basin-Havre Trough backarc system (SW Pacific): Locking/unlocking induced by
seamount chain subduction. Geochemistry, Geophysics, Geosystems, 4(5), 8909.
Ryburn, R. J. (1980). Blueschists and associated rocks in the south Sepik region, Papua New
Guinea; field relations, petrology, mineralogy, metamorphism and tectonic
setting. Doctoral dissertation, University of Auckland.
Rytuba, J. J., & Miller, W. R. (1990). Geology and geochemistry of epithermal precious metal vein
systems in the intra-oceanic arcs of Palau and Yap, western Pacific. Journal of
Geochemical Exploration, 35(1-3), 413-447.
Saito, M. (2008). Rapid evolution of the Eocene accretionary complex (Hyuga Group) of the
Shimanto terrane in southeastern Kyushu, southwestern Japan. Island Arc, 17(2), 242-
260.
Sajona, F. G., Bellon, H., Maury, R. C., Pubellier, M., Quebral, R. D., Cotten, J., ... & Pamatian, P.
(1997). Tertiary and quaternary magmatism in Mindanao and Leyte (Philippines):
geochronology, geochemistry and tectonic setting. Journal of Asian Earth Sciences, 15(2-
3), 121-153.
Sandmann, S., Nagel, T. J., Froitzheim, N., Ustaszewski, K., & Münker, C. (2015). Late Miocene to
Early Pliocene blueschist from Taiwan and its exhumation via forearc extraction. Terra
Nova, 27(4), 285-291.
Santos, R. A. (2014). Mineral Resource Estimate Report of the Islands of Nonoc, Awasan, and
Hanigad and Part of South Dinagat for Pacific Nickel Philippines Inc. (PNPI). MPSA No.
072-1997-XIII.
Saputra, A., Hall, R., & White, L. T. (2014). Development of the Sorong Fault Zone North of
Misool, Eastern Indonesia. Indonesian Petroleum Association, 38th Annual Convention
Proceedings, IPA 14-G-086.
Saragih, R. Y., Sunan, H. L., Slameto, E., & Nurdiana, I. (2020, December). Rifting and Lifting
Neogene Age in Biak-Yapen Basin Based on Structural Trajectory Analogy in Biak and
Supiori Island. In IOP Conference Series: Materials Science and Engineering (Vol. 982, No.
1, p. 012042). IOP Publishing.
Sarewitz, D. R., & Karig, D. E. (1986). Processes of allochthonous terrane evolution, Mindoro
Island, Philippines. Tectonics, 5(4), 525-552.
Sasaki, T., Yamazaki, T., & Ishizuka, O. (2014). A revised spreading model of the West Philippine
Basin. Earth, Planets and Space, 66, 1-9.
Sato, T., Kasahara, J., Katao, H., Tomiyama, N., Mochizuki, K., & Koresawa, S. (1997). Seismic
observations at the Yap Islands and the northern Yap Trench. Tectonophysics, 271(3-4),
285-294.
Schellart, W. P., & Spakman, W. (2015). Australian plate motion and topography linked to fossil
New Guinea slab below Lake Eyre. Earth and Planetary Science Letters, 421, 107-116.
Schellart, W. P., Kennett, B. L. N., Spakman, W., & Amaru, M. (2009). Plate reconstructions and
tomography reveal a fossil lower mantle slab below the Tasman Sea. Earth and
Planetary Science Letters, 278(3-4), 143-151.
Schellart, W. P., Lister, G. S., & Toy, V. G. (2006). A Late Cretaceous and Cenozoic reconstruction
of the Southwest Pacific region: tectonics controlled by subduction and slab rollback
processes. Earth-Science Reviews, 76(3-4), 191-233.
Schlüter, H. U., HINz, K., & Block, M. (1996). Tectono-stratigraphic terranes and detachment
faulting of the South China Sea and Sulu Sea. Marine Geology, 130(1-2), 39-78.
Schweller, W. J., Karig, D. E., & Bachman, S. D. (1983). The Tectonic and Geologic Evolution of
Southeast Asian Seas and Islands: Part 2. American Geophysical Union Geophysical
Monograph Series 27, 95-123.
Scotese, C. R. (2021). An atlas of Phanerozoic paleogeographic maps: the seas come in and the
seas go out. Annual Review of Earth and Planetary Sciences, 49, 679-728.
Sdrolias, M., Müller, R. D., & Gaina, C. (2001). Plate tectonic evolution of eastern Australian
marginal ocean basins. In K. C. Hill & T. Bernecker (Eds.), Eastern Australian Basins
Symposium (pp. 227–237). Melbourne: Petroleum Exploration Society of Australia
Special Publication.
Sdrolias, M., Müller, R. D., Mauffret, A., & Bernardel, G. (2004a). Enigmatic formation of the
Norfolk Basin, SW Pacific: A plume influence on back-arc extension. Geochemistry,
Geophysics, Geosystems, 5, Q06005.
Sdrolias, M., Roest, W. R., & Müller, R. D. (2004b). An expression of Philippine Sea plate rotation:
the Parece Vela and Shikoku basins. Tectonophysics, 394(1-2), 69-86.
Seton, M., & Müller, R. D. (2008). Reconstructing the junction between Panthalassa and Tethys
since the Early Cretaceous. PESA Eastern Australian Basin Symposium III abstract, 263-
266.
Seton, M., Flament, N., Whittaker, J., Müller, R. D., Gurnis, M., & Bower, D. J. (2015). Ridge
subduction sparked reorganization of the Pacific plate-mantle system 60–50 million
years ago. Geophysical Research Letters, 42(6), 1732-1740.
Seton, M., Mortimer, N., Williams, S., Quilty, P., Gans, P., Meffre, S., ... & Matthews, K. J. (2016).
Melanesian back-arc basin and arc development: Constraints from the eastern Coral
Sea. Gondwana Research, 39, 77-95.
Seton, M., Müller, R. D., Zahirovic, S., Gaina, C., Torsvik, T., Shephard, G., Talsma, A., Gurnis, M.,
Turner, M., Maus, S. & Chandler, M. (2012). Global continental and ocean basin
reconstructions since 200 Ma. Earth-Science Reviews, 113(3-4), 212-270.
Seton, M., Whittaker, J. M., Wessel, P., Müller, R. D., DeMets, C., Merkouriev, S., ... & Williams, S. E.
(2014). Community infrastructure and repository for marine magnetic identifications.
Geochemistry, Geophysics, Geosystems, 15(4), 1629-1641.
Sevin, B., Maurizot, P., Cluzel, D., Tournadour, E., Etienne, S., Folcher, N., ... & Patriat, M. (2020).
Post-obduction evolution of New Caledonia. Geological Society, London, Memoirs, 51(1),
147-188.
Shao, L., Cao, L., Qiao, P., Zhang, X., Li, Q., & van Hinsbergen, D. J. (2017). Cretaceous–Eocene
provenance connections between the Palawan Continental Terrane and the northern
South China Sea margin. Earth and Planetary Science Letters, 477, 97-107.
Shiraki, K. (1971). Metamorphic basement rocks of Yap Islands, western Pacific: Possible
oceanic crust beneath an island arc. Earth and Planetary Science Letters, 13(1), 167-174.
Sibuet, J. C., Hsu, S. K., Shyu, C. T., & Liu, C. S. (1995). Structural and kinematic evolutions of the
Okinawa Trough backarc basin. Backarc Basins: Tectonics and Magmatism, 343-379.
Sibuet, J. C., Letouzey, J., Barbier, F., Charvet, J., Foucher, J. P., Hilde, T. W., ... & Stéphan, J. F.
(1987). Back arc extension in the Okinawa Trough. Journal of Geophysical Research: Solid
Earth, 92(B13), 14041-14063.
Silver, E. A., & Moore, J. C. (1978). The Molucca sea collision zone, Indonesia. Journal of
Geophysical Research: Solid Earth, 83(B4), 1681-1691.
Silver, E. A., & Rangin, C. (1991). Leg 124 Tectonic synthesis. In Proceedings of the Ocean Drilling
Program, 124, 3-9.
Smith, I. E. (2013). The chemical characterization and tectonic significance of ophiolite terrains
in southeastern Papua New Guinea. Tectonics, 32(2), 159-170.
Smith, R. B., Betzler, C., Brass, G. W., Huang, Z., Linsley, B. K., Menill, D., ... & Spadea, P. (1990).
Depositional history of the Celebes Sea from ODP Sites 767 and 770. Geophysical
Research Letters, 17(11), 2061-2064.
Spakman, W., Chertova, M. V., van den Berg, A., & van Hinsbergen, D. J. (2018). Puzzling features
of western Mediterranean tectonics explained by slab dragging. Nature
Geoscience, 11(3), 211-216.
Spakman, W., Stein, S., van der Hilst, R., & Wortel, R. (1989). Resolution experiments for NW
Pacific subduction zone tomography. Geophysical Research Letters, 16(10), 1097-1100.
Srivastava, S. P., & Roest, W. R. (1996). Comment on “Porcupine plate hypothesis” by MF Gerstell
and JM Stock (Marine Geophysical Researches 16, pp. 315–323, 1994). Marine
Geophysical Researches, 18(5), 589-593.
Stampfli, G. M., & Borel, G. D. (2002). A plate tectonic model for the Paleozoic and Mesozoic
constrained by dynamic plate boundaries and restored synthetic oceanic
isochrons. Earth and Planetary science letters, 196(1-2), 17-33.
Stern, R. J. (2004). Subduction initiation: spontaneous and induced. Earth and Planetary Science
Letters, 226(3-4), 275-292.
Stern, R. J., & Bloomer, S. H. (1992). Subduction zone infancy: examples from the Eocene Izu-
Bonin-Mariana and Jurassic California arcs. Geological Society of America
Bulletin, 104(12), 1621-1636.
Stern, R. J., & Gerya, T. (2018). Subduction initiation in nature and models: A review.
Tectonophysics, 746, 173-198.
Stern, R. J., Fouch, M. J., & Klemperer, S. L. (2003). An overview of the Izu-Bonin-Mariana
subduction factory. Geophysical Monograph, American Geophysical Union, 138, 175-222.
Stern, R. J., Reagan, M., Ishizuka, O., Ohara, Y., & Whattam, S. (2012). To understand subduction
initiation, study forearc crust: To understand forearc crust, study
ophiolites. Lithosphere, 4(6), 469-483.
Stewart, W. D., & Sandy, M. J. (1988). Geology of New Ireland and Djaul Islands, northeastern
Papua New Guinea. Geology and Offshore Resources of Pacific Island Arcs–New Ireland
and Manus Region, Papua New Guinea, 9, 13-30.
Suerte, L. O., Yumul, G. P., Tamayo, R. A., Dimalanta, C. B., Zhou, M. F., Maury, R. C., ... & Balce, C. L.
(2005). Geology, geochemistry and U-Pb SHRIMP age of the Tacloban Ophiolite Complex,
Leyte Island (Central Philippines): Implications for the existence and extent of the
proto-Philippine Sea Plate. Resource Geology, 55(3), 207-216.
Sugiyama, Y. (1994). Neotectonics of Southwest Japan due to the right-oblique subduction of the
Philippine Sea plate. Geofisica International, 33(1), 53-76.
Suppe, J. & Chi, W. R. (1985). Tectonic implications of Miocene sediments of Lan-Hsu island,
northern Luzon arc. Petroleum Geology of Taiwan, 21, 93-106.
Suppe, J. O. H. N. (1980). A retrodeformable cross section of northern Taiwan. In Proc. Geol. Soc.
China, 23, 46-55.
Sutherland, R., Collot, J., Bache, F., Henrys, S., Barker, D., Browne, G. H., et al. (2017). Widespread
compression associated with Eocene Tonga-Kermadec subduction
initiation. Geology, 45, 355–358.
Taetz, S., Scherer, E.E., Bröcker, M., Spandler, C., and John, Tim (2021). Petrological and Lu-Hf
age constraints for eclogitic rocks from the Pam Peninsula, New Caledonia. Lithos, 388-
389, 106073
Taira, A. (1988). The Shimanto belt in Shikoku, Japan-evolution of Cretaceous to Miocene
accretionary prism. The Shimanto belt, Southwest Japan-Studies on the evolution of an
accretionary prism.
Taira, A., Mann, P., & Rahardiawan, R. (2004). Incipient subduction of the Ontong Java Plateau
along the North Solomon trench. Tectonophysics, 389(3-4), 247-266.
Taira, A., Okada, H., Whitaker, J. H., & Smith, A. J. (1982). The Shimanto Belt of Japan: Cretaceous-
lower Miocene active-margin sedimentation. Geological Society, London, Special
Publications, 10(1), 5-26.
Tamayo Jr, R. A., Maury, R. C., Yumul Jr, G. P., Polvé, M., Cotten, J., Dimantala, C. B., & Olaguera, F.
O. (2004). Subduction-related magmatic imprint of most Philippine ophiolites:
implications on the early geodynamic evolution of the Philippine archipelago. Bulletin de
la Société Géologique de France, 175(5), 443-460.
Tamayo Jr, R. A., Yumul Jr, G. P., Maury, R. C., Polvé, M., Cotten, J., & Bohn, M. (2001).
Petrochemical investigation of the Antique Ophiolite (Philippines): Implications on
volcanogenic massive sulfide and podiform chromitite deposits. Resource Geology, 51(2),
145-164.
Tamayo, R. A., Yumul, G. P., Maury, R. C., Bellon, H., Cotten, J., Polvé, M., ... & Querubin, C. (2000).
Complex origin for the south-western Zamboanga metamorphic basement complex,
Western Mindanao, Philippines. Island Arc, 9(4), 638-652.
Tamayo, R. J., Yumul Jr., G. P., Santos, R. A., Jumawan, F., Rodolfo, K. S. (1998). Petrology and
mineral chemistry of a back-arc upper mantle suite: Example from the Camarines Norte
Ophiolite complex, South Luzon. Journal of the Geological Society of the Philippines, 51, 1-
23.
Tanaka, G., & Nomura, S. I. (2009). Late Miocene and Pliocene Ostracoda from the Shimajiri
Group, Kume-jima Island, Japan: Biogeographical significance of the timing of the
formation of back-arc basin (Okinawa Trough). Palaeogeography, Palaeoclimatology,
Palaeoecology, 276(1-4), 56-68.
Tapster, S., Roberts, N. M. W., Petterson, M. G., Saunders, A. D., & Naden, J. (2014). From
continent to intra-oceanic arc: Zircon xenocrysts record the crustal evolution of the
Solomon island arc. Geology, 42(12), 1087-1090.
Taylor, B. (1979). Bismarck Sea: evolution of a back-arc basin. Geology, 7(4), 171-174.
Taylor, B. (2006). The single largest oceanic plateau: Ontong Java–Manihiki–Hikurangi. Earth
and Planetary Science Letters, 241(3-4), 372-380.
Taylor, B., Goodliffe, A. M., & Martinez, F. (1999). How continents break up: insights from Papua
New Guinea. Journal of Geophysical Research: Solid Earth, 104(B4), 7497-751
Taylor, F. W., Jouannic, C., & Bloom, A. L. (1985). Quaternary uplift of the Torres Islands,
northern New Hebrides frontal arc: Comparison with Santo and Malekula Islands,
central New Hebrides frontal arc. The Journal of Geology, 93(4), 419-438.
Tejada, M. L. G., & Castillo, P. R. (2002). In search of a common ground: Geochemical study of
ancient oceanic crust in eastern Luzon, Philippines. Geochimica et Cosmochimica Acta
56(66), 767.
Tejada, M. L. G., Mahoney, J. J., Duncan, R. A., & Hawkins, M. P. (1996). Age and geochemistry of
basement and alkalic rocks of Malaita and Santa Isabel, Solomon Islands, southern
margin of Ontong Java Plateau. Journal of Petrology, 37(2), 361-394.
Tejada, M. L. G., Mahoney, J. J., Neal, C. R., Duncan, R. A., & Petterson, M. G. (2002). Basement
geochemistry and geochronology of Central Malaita, Solomon Islands, with implications
for the origin and evolution of the Ontong Java Plateau. Journal of Petrology, 43(3), 449-
484.
Torsvik, T. H., Doubrovine, P. V., Steinberger, B., Gaina, C., Spakman, W., & Domeier, M. (2017).
Pacific plate motion change caused the Hawaiian-Emperor Bend. Nature
Communications, 8(1), 15660.
Torsvik, T. H., Müller, R. D., Van der Voo, R., Steinberger, B., & Gaina, C. (2008). Global plate
motion frames: toward a unified model. Reviews of geophysics, 46(3).
Torsvik, T. H., Steinberger, B., Shephard, G. E., Doubrovine, P. V., Gaina, C., Domeier, M., ... &
Sager, W. W. (2019). Pacific-Panthalassic reconstructions: Overview, errata and the way
forward. Geochemistry, Geophysics, Geosystems, 20(7), 3659-3689.
Torsvik, T. H., Van der Voo, R., Preeden, U., Mac Niocaill, C., Steinberger, B., Doubrovine, P. V., ... &
Cocks, L. R. M. (2012). Phanerozoic polar wander, palaeogeography and
dynamics. Earth-Science Reviews, 114(3-4), 325-368.
Torsvik, T., & Cocks, L. (2017). Earth History and Palaeogeography. Cambridge: Cambridge
University Press. doi:10.1017/9781316225523.
Tregoning, P., Tan, F., Gilliland, J., McQueen, H., & Lambeck, K. (1998). Present-day crustal
motion in the Solomon Islands from GPS observations. Geophysical research
letters, 25(19), 3627-3630.
Ujiie, K. (1997). Off-scraping accretionary process under the subduction of young oceanic crust:
The Shimanto Belt of Okinawa Island, Ryukyu Arc. Tectonics, 16(2), 305-322.
Ujiie, K. (2002). Evolution and kinematics of an ancient décollement zone, mélange in the
Shimanto accretionary complex of Okinawa Island, Ryukyu Arc. Journal of Structural
Geology, 24(5), 937-952.
Vaes, B., Van Hinsbergen, D. J. J., Van de Lagemaat, S. H. A., Van der Wiel, E., Lom, N., Advokaat, E.,
... & Langereis, C., (2023). A global apparent polar wander path for the last 320 Ma
calculated from site-level paleomagnetic data. Earth-Science Reviews, in press.
Vaes, B., Van Hinsbergen, D. J., & Boschman, L. M. (2019). Reconstruction of subduction and
back-arc spreading in the NW Pacific and Aleutian Basin: Clues to causes of Cretaceous
and Eocene plate reorganizations. Tectonics, 38(4), 1367-1413.
Vaes, B., Van Hinsbergen, D. J., & Boschman, L. M. (2019). Reconstruction of subduction and
back-arc spreading in the NW Pacific and Aleutian Basin: Clues to causes of Cretaceous
and Eocene plate reorganizations. Tectonics, 38(4), 1367-1413.
Van de Lagemaat, S. H. A., Boschman, L. M., Kamp, P. J. J., Langereis, C. G., & van Hinsbergen, D. J.
J. (2018b). Post-remagnetisation vertical axis rotation and tilting of the Murihiku
Terrane (North Island, New Zealand). New Zealand Journal of Geology and
Geophysics, 61(1), 9-25.
Van de Lagemaat, S. H. A., Cao, L., Asis, J., Advokaat, E.L., Mason, P.R.D., Xu, D., Dekkers, M.J., and
van Hinsbergen, D.J.J., (2023b). Causes of Late Cretaceous subduction termination below
South China and Borneo: Was the Proto-South China Sea underlain by an oceanic
plateau? Submitted to Geoscience Frontiers. Preprint:
https://doi.org/10.31223/X5ZW9G
Van de Lagemaat, S. H. A., Kamp, P. J. J., Boschman, L. M., & Van Hinsbergen, D. J. J. (2023a).
Reconciling the Cretaceous breakup and demise of the Phoenix Plate with East
Gondwana orogenesis in New Zealand. Earth-Science Reviews, 236, 104276.
Van de Lagemaat, S. H. A., Mering, J. A., & Kamp, P. J. J. (2022). Geochemistry of syntectonic
carbonate veins within Late Cretaceous turbidites, Hikurangi Margin (New Zealand):
Implications for a mid-Oligocene age of subduction initiation. Geochemistry, Geophysics,
Geosystems, 23(5), e2021GC010125.
Van de Lagemaat, S. H. A., Pastor-Galán, D., Zanderink, B.B.G., Villareal, M.J., Jenson, J.W., Dekkers,
M.J., van Hinsbergen, D.J.J., (2023b), A critical reappraisal of paleomagnetic evidence for
Philippine Sea Plate rotation. Tectonophysics, 863, 230010.
Van de Lagemaat, S. H. A., Swart, M. L., Vaes, B., Kosters, M. E., Boschman, L. M., Burton-Johnson,
A., Bijl, P. K., Spakman, W. & Van Hinsbergen, D. J. (2021). Subduction initiation in the
Scotia Sea region and opening of the Drake Passage: When and why?. Earth-Science
Reviews, 215, 103551.
Van de Lagemaat, S. H. A., Van Hinsbergen, D. J. J., Boschman, L. M., Kamp, P. J. J., & Spakman, W.
(2018a). Southwest Pacific absolute plate kinematic reconstruction reveals major
Cenozoic Tonga-Kermadec slab dragging. Tectonics, 37(8), 2647-2674.
Van den Broek, J. M., & Gaina, C. (2020). Microcontinents and continental fragments associated
with subduction systems. Tectonics, 39(8), e2020TC006063.
Van der Hist, R., Engdahl, R., Spakman, W., & Nolet, G. (1991). Tomographic imaging of
subducted lithosphere below northwest Pacific island arcs. Nature, 353(6339), 37-43.
Van der Meer, D. G., Spakman, W., Van Hinsbergen, D. J., Amaru, M. L., & Torsvik, T. H. (2010).
Towards absolute plate motions constrained by lower-mantle slab remnants. Nature
Geoscience, 3(1), 36-40.
Van der Meer, D. G., Torsvik, T. H., Spakman, W., Van Hinsbergen, D. J. J., & Amaru, M. L. (2012).
Intra-Panthalassa Ocean subduction zones revealed by fossil arcs and mantle
structure. Nature Geoscience, 5(3), 215-219.
Van der Meer, D. G., Van Hinsbergen, D. J., & Spakman, W. (2018). Atlas of the underworld: Slab
remnants in the mantle, their sinking history, and a new outlook on lower mantle
viscosity. Tectonophysics, 723, 309-448.
Van der Voo, R., van Hinsbergen, D. J., Domeier, M., Spakman, W., & Torsvik, T. H. (2015). Latest
Jurassic–earliest Cretaceous closure of the Mongol-Okhotsk Ocean: A paleomagnetic and
seismological-tomographic analysis. Geological Society of America Special Papers, 513,
589-606.
Van Hinsbergen, D. J. J., & Schouten, T. L. (2021). Deciphering paleogeography from orogenic
architecture: constructing orogens in a future supercontinent as thought
experiment. American Journal of Science, 321(6), 955-1031.
Van Hinsbergen, D. J. J., Steinberger, B., Guilmette, C., Maffione, M., Gürer, D., Peters, K., ... &
Spakman, W. (2021). A record of plume-induced plate rotation triggering subduction
initiation. Nature Geoscience, 14(8), 626-630.
Van Hinsbergen, D. J. J., Torsvik, T. H., Schmid, S. M., Maţenco, L. C., Maffione, M., Vissers, R. L.,
Gürer, D. & Spakman, W. (2020). Orogenic architecture of the Mediterranean region and
kinematic reconstruction of its tectonic evolution since the Triassic. Gondwana
Research, 81, 79-229.
Van Hinsbergen, D. J., Peters, K., Maffione, M., Spakman, W., Guilmette, C., Thieulot, C., ... &
Kaymakcı, N. (2015). Dynamics of intraoceanic subduction initiation: 2.
Suprasubduction zone ophiolite formation and metamorphic sole exhumation in context
of absolute plate motions. Geochemistry, Geophysics, Geosystems, 16(6), 1771-1785.
Van Hinsbergen, D. J., Spakman, W., de Boorder, H., Van Dongen, M., Jowitt, S. M., & Mason, P. R.
(2020b). Arc-type magmatism due to continental-edge plowing through ancient
subduction-enriched mantle. Geophysical Research Letters, 47(9), e2020GL087484.
Van Hinsbergen, D. J., Torsvik, T. H., Schmid, S. M., Maţenco, L. C., Maffione, M., Vissers, R. L., ... &
Spakman, W. (2020a). Orogenic architecture of the Mediterranean region and kinematic
reconstruction of its tectonic evolution since the Triassic. Gondwana Research, 81, 79-
229.
Van Horne, A., Sato, H., & Ishiyama, T. (2017). Evolution of the Sea of Japan back-arc and some
unsolved issues. Tectonophysics, 710, 6-20.
Veevers, J. J., Powell, C. M., & Roots, S. R. (1991). Review of seafloor spreading around Australia.
I. Synthesis of the patterns of spreading. Australian journal of earth sciences, 38(4), 373-
389.
Vissers, R. L. M., & Meijer, P. T. (2012a). Mesozoic rotation of Iberia: Subduction in the
Pyrenees?. Earth-Science Reviews, 110(1-4), 93-110.
Vissers, R. L. M., & Meijer, P. T. (2012b). Iberian plate kinematics and Alpine collision in the
Pyrenees. Earth-Science Reviews, 114(1-2), 61-83.
Wai, K. M., Abbott, M. J., & Grady, A. E. (1994). The Sadowa Igneous Complex, Eastern Papua New
Guinea: Ophiolite or not?. Goldschmidt Conference Edinburgh, 949-950.
Wakita, K. (2015). OPS mélange: A new term for mélanges of convergent margins of the
world. International Geology Review, 57(5-8), 529-539.
Wakita, K., & Metcalfe, I. (2005). Ocean plate stratigraphy in East and Southeast Asia. Journal of
Asian Earth Sciences, 24(6), 679-702.
Walker, D. A., & McDougall, I. (1982). 40Ar/39Ar and K-Ar dating of altered glassy volcanic rocks:
the Dabi Volcanics, PNG. Geochimica et Cosmochimica Acta, 46(11), 2181-2190.
Wallis, S. R., Yamaoka, K., Mori, H., Ishiwatari, A., Miyazaki, K., & Ueda, H. (2020). The basement
geology of Japan from A to Z. Island Arc, 29(1), e12339.
Wang, Y., Qian, X., Asis, J. B., Cawood, P. A., Wu, S., Zhang, Y., ... & Lu, X. (2023). “Where, when and
why” for the arc-trench gap from Mesozoic Paleo-Pacific subduction zone: Sabah
Triassic-Cretaceous igneous records in East Borneo. Gondwana Research, 117, 117-138.
Warren, P. Q., & Cloos, M. (2007). Petrology and tectonics of the Derewo metamorphic belt, west
New Guinea. International Geology Review, 49(6), 520-553.
Waters, P. J., Cooke, D. R., Gonzales, R. I., & Phillips, D. (2011). Porphyry and epithermal deposits
and 40Ar/39Ar geochronology of the Baguio district, Philippines. Economic
Geology, 106(8), 1335-1363.
Watts, A. B., Weissel, J. K., & Larson, L. R. (1977). Sea-floor spreading in marginal basins of the
western Pacific. Tectonophysics, 37(1-3), 167–181.
Webb, L. E., Baldwin, S. L., & Fitzgerald, P. G. (2014). The Early-Middle Miocene subduction
complex of the Louisiade Archipelago, southern margin of the Woodlark
Rift. Geochemistry, Geophysics, Geosystems, 15(10), 4024-4046.
Webb, M., & White, L. T. (2016). Age and nature of Triassic magmatism in the Netoni Intrusive
Complex, West Papua, Indonesia. Journal of Asian Earth Sciences, 132, 58-74.
Webb, M., White, L. T., Jost, B. M., & Tiranda, H. (2019). The Tamrau Block of NW New Guinea
records late Miocene–Pliocene collision at the northern tip of the Australian
Plate. Journal of Asian Earth Sciences, 179, 238-260.
Webb, M., White, L. T., Jost, B. M., Tiranda, H., & BouDagher-Fadel, M. (2020). The history of
Cenozoic magmatism and collision in NW New Guinea–New insights into the tectonic
evolution of the northernmost margin of the Australian Plate. Gondwana Research, 82,
12-38.
Weiland Jr, R. J. (1999). Emplacement of the Irian ophiolite and unroofing of the Ruffaer
metamorphic belt of Irian Jaya, Indonesia. PhD Thesis. The University of Texas at Austin.
Weissel, J. K. (1980). Evidence for Eocene oceanic crust in the Celebes Basin. Washington DC
American Geophysical Union Geophysical Monograph Series, 23, 37-47.
Weissel, J. K., & Anderson, R. N. (1978). Is there a Caroline plate?. Earth and Planetary Science
Letters, 41(2), 143-158.
Weissel, J. K., & Watts, A. B. (1979). Tectonic evolution of the Coral Sea basin. Journal of
Geophysical Research: Solid Earth, 84(B9), 4572-4582.
Weissel, J. K., Taylor, B., & Karner, G. D. (1982). The opening of the Woodlark Basin, subduction
of the Woodlark spreading system, and the evolution of northern Melanesia since mid-
Pliocene time. Tectonophysics, 87(1-4), 253-277.
Wessel, P., Matthews, K. J., Müller, R. D., Mazzoni, A., Whittaker, J. M., Myhill, R., & Chandler, M. T.
(2015). Semiautomatic fracture zone tracking (Vol. 16, No. 7, pp. 2462-2472).
Whattam, S. A. (2009). Arc-continent collisional orogenesis in the SW Pacific and the nature,
source and correlation of emplaced ophiolitic nappe components. Lithos, 113(1-2), 88-
114.
Whattam, S. A., Malpas, J. G., Ali, J. R., Smith, I. E., & Lo, C. H. (2004). Origin of the Northland
Ophiolite, northern New Zealand: discussion of new data and reassessment of the
model. New Zealand Journal of Geology and Geophysics, 47(3), 383-389.
Whattam, S. A., Malpas, J., Ali, J. R., & Smith, I. E. M. (2008). New SW Pacific tectonic model:
Cyclical intraoceanic magmatic arc construction and near-coeval emplacement along the
Australia-Pacific margin in the Cenozoic. Geochemistry, Geophysics, Geosystems, 9,
Q03021.
Whattam, S. A., Malpas, J., Ali, J. R., Lo, C. H., & Smith, I. E. (2005). Formation and emplacement of
the Northland ophiolite, northern New Zealand: SW Pacific tectonic implications. Journal
of the Geological Society, 162(2), 225-241.
Whattam, S. A., Malpas, J., Smith, I. E. M., & Ali, J. R. (2006). Link between SSZ ophiolite
formation, emplacement and arc inception, Northland, New Zealand: U-Pb SHRIMP
constraints; Cenozoic SW Pacific tectonic implications. Earth and Planetary Science
Letters, 250(3-4), 606–632.
Whittaker, J. M., Goncharov, A., Williams, S. E., Müller, R. D., & Leitchenkov, G. (2013). Global
sediment thickness data set updated for the Australian-Antarctic Southern
Ocean. Geochemistry, Geophysics, Geosystems, 14(8), 3297-3305.
Whittaker, J. M., Muller, R. D., Leitchenkov, G., Stagg, H., Sdrolias, M., Gaina, C., & Goncharov, A.
(2007). Major Australian-Antarctic plate reorganization at Hawaiian-Emperor bend
time. Science, 318(5847), 83-86.
Williams, S. E., Whittaker, J. M., & Müller, R. D. (2011). Full-fit, palinspastic reconstruction of the
conjugate Australian-Antarctic margins. Tectonics, 30(6).
Wintsch, R. P., & Li, X. H. (2014). Hf and O isotopic evidence for metamorphic crystallization of
zircon during contact metamorphism of Fenniaolin metabasalts, Tananao complex,
Taiwan. Lithos, 205, 142-147.
Wintsch, R. P., Yang, H. J., Li, X. H., & Tung, K. A. (2011). Geochronologic evidence for a cold arc–
continent collision: The Taiwan orogeny. Lithos, 125(1-2), 236-248.
Woodhead, J., Hergt, J., Sandiford, M., & Johnson, W. (2010). The big crunch: Physical and
chemical expressions of arc/continent collision in the Western Bismarck arc. Journal of
Volcanology and Geothermal Research, 190(1-2), 11-24.
Worthing, M. A., & Crawford, A. J. (1996). The igneous geochemistry and tectonic setting of
metabasites from the Emo Metamorphics, Papua New Guinea; a record of the evolution
and destruction of a backarc basin. Mineralogy and Petrology, 58(1-2), 79-100.
Wright, N. M., Müller, R. D., Seton, M., & Williams, S. E. (2015). Revision of Paleogene plate
motions in the Pacific and implications for the Hawaiian-Emperor bend. Geology, 43(5),
455-458.
Wright, N. M., Seton, M., Williams, S. E., & Mueller, R. D. (2016). The Late Cretaceous to recent
tectonic history of the Pacific Ocean basin. Earth-Science Reviews, 154, 138-173.
Wu, J. T. J., & Wu, J. (2019). Izanagi-Pacific ridge subduction revealed by a 56 to 46 Ma magmatic
gap along the northeast Asian margin. Geology, 47(10), 953-957.
Wu, J., Lin, Y. A., Flament, N., Wu, J. T. J., & Liu, Y. (2022). Northwest Pacific-Izanagi plate
tectonics since Cretaceous times from western Pacific mantle structure. Earth and
Planetary Science Letters, 583, 117445.
Wu, J., Suppe, J., Lu, R., & Kanda, R. (2016). Philippine Sea and East Asian plate tectonics since 52
Ma constrained by new subducted slab reconstruction methods. Journal of Geophysical
Research: Solid Earth, 121(6), 4670-4741.
Xu, Y., Yan, Q., Shi, X., Jichao, Y., Deng, X., Xu, W., & Jing, C. (2022). Discovery of Late Mesozoic
volcanic seamounts at the ocean-continent transition zone in the Northeastern margin
of South China Sea and its tectonic implication. Gondwana Research.
Yamazaki, T., Seama, N., Okino, K., Kitada, K., Joshima, M., Oda, H., & Naka, J. (2003). Spreading
process of the northern Mariana Trough: Rifting-spreading transition at 22
N. Geochemistry, Geophysics, Geosystems, 4(9).
Yan, C. Y. & Kroenke, L. W. (1993). A plate reconstruction of the southwest Pacific 0-100 Ma.
In Proceedings of Ocean Drilling Program, Scientific Results, 130, 697-709.
Yan, S., Yan, Q., Shi, X., Yuan, L., Liu, Y., Yang, G., & Ye, X. (2022). The dynamics of the Sorol
Trough magmatic system: Insights from bulk-rock chemistry and mineral geochemistry
of basaltic rocks. Geological Journal, 57(10), 4074-4089.
Yang, K. M., Huang, S. T., Jong-Chang, W., Ting, H. H., Wen-Wei, M., Lee, M., ... & Chang-Jie, L.
(2007). 3D geometry of the Chelungpu thrust system in central Taiwan: Its implications
for active tectonics. TAO: Terrestrial, Atmospheric and Oceanic Sciences, 18(2), 143.
Yang, K. M., Rau, R. J., Chang, H. Y., Hsieh, C. Y., Ting, H. H., Huang, S. T., ... & Tang, Y. J. (2016). The
role of basement-involved normal faults in the recent tectonics of western
Taiwan. Geological Magazine, 153(5-6), 1166-1191.
Yang, T. F., Tien, J. L., Chen, C. H., Lee, T., & Punongbayan, R. S. (1995). Fission-track dating of
volcanics in the northern part of the Taiwan-Luzon Arc: eruption ages and evidence for
crustal contamination. Journal of Southeast Asian Earth Sciences, 11(2), 81-93.
Yeh, K. Y., & Cheng, Y. N. (2001). The first finding of early Cretaceous radiolarians from Lanyu,
the Philippine Sea Plate. Collection and Research, (13), 111-145.
Yu, M., Dilek, Y., Yumul Jr, G. P., Yan, Y., Dimalanta, C. B., & Huang, C. Y. (2020). Slab-controlled
elemental–isotopic enrichments during subduction initiation magmatism and variations
in forearc chemostratigraphy. Earth and Planetary Science Letters, 538, 116217.
Yui, T. F., Maki, K., Lan, C. Y., Hirata, T., Chu, H. T., Kon, Y., Yokoyama, T. D., Jahn, B. M., & Ernst, W.
G. (2012). Detrital zircons from the Tananao metamorphic complex of Taiwan:
Implications for sediment provenance and Mesozoic tectonics. Tectonophysics, 541, 31-42.
Yumul Jr, G. P. (1993). Angat Ophiolitic Complex, Luzon, Philippines: a Cretaceous dismembered
marginal basin ophiolitic complex. Journal of Southeast Asian Earth Sciences, 8(1-4), 529-
537.
Yumul Jr, G. P., Dimalanta, C. B., & Jumawan, F. T. (2000a). Geology of the southern Zambales
ophiolite complex, Luzon, Philippines. Island Arc, 9(4), 542-555.
Yumul Jr, G. P., Dimalanta, C. B., Maglambayan, V. B., & Tamayo Jr, R. A. (2003). Mineralization
controls in island arc settings: Insights from Philippine metallic deposits. Gondwana
Research, 6(4), 767-776.
Yumul Jr, G. P., Dimalanta, C. B., Tamayo Jr, R. A., & Faustino-Eslava, D. V. (2013). Geological
features of a collision zone marker: the Antique Ophiolite Complex (Western Panay,
Philippines). Journal of Asian Earth Sciences, 65, 53-63.
Yumul Jr, G. P., Dimalanta, C. B., Tamayo Jr, R. A., & Zhou, M. F. (2006). Geology and geochemistry
of the Rapu-Rapu ophiolite complex, eastern Philippines: possible fragment of the proto-
Philippine Sea Plate. International Geology Review, 48(4), 329-348.
Yumul Jr, G. P., Dimalanta, C. B., Tamayo Jr, R. A., Maury, R. C., Bellon, H., Polvé, M., ... & Cotten, J.
(2004). Geology of the Zamboanga Peninsula, Mindanao, Philippines: An enigmatic
South China continental fragment? Geological Society, London, Special
Publications, 226(1), 289-312.
Yumul Jr, G. P., Dimalanta, C. B., Tamayo, R. A., & Barretto, J. A. L. (2000b). Contrasting
morphological trends of islands in Central Philippines: speculation on their origin. Island
Arc, 9(4), 627-637.
Yumul Jr, G. P., Jumawan, F. T., & Dimalanta, C. B. (2009). Geology, geochemistry and chromite
mineralization potential of the Amnay Ophiolitic Complex, Mindoro,
Philippines. Resource geology, 59(3), 263-281.
Zahirovic, S., Seton, M., & Müller, R. D. (2014). The Cretaceous and Cenozoic tectonic evolution of
Southeast Asia. Solid Earth, 5(1), 227-273.
Zamoras, L. R., & Matsuoka, A. (2001). The Malampaya Sound Group in the Calamian Islands,
North Palawan Block (Philippines). The Journal of the Geological Society of Japan, 107(5),
XI-XII.
Zamoras, L. R., & Matsuoka, A. (2004). Accretion and postaccretion tectonics of the Calamian
Islands, North Palawan block, Philippines. Island Arc, 13(4), 506-519.
Zglinicki, K., Szamałek, K., & Górska, I. (2020). The Cyclops Ophiolite as a Source of High-Cr
Spinels from Marine Sediments on the Jayapura Regency Coast (New Guinea,
Indonesia). Minerals, 10(9), 735.
Zhang, G., Zhang, J., Wang, S., & Zhao, J. (2020). Geochemical and chronological constraints on
the mantle plume origin of the Caroline Plateau. Chemical Geology, 540, 119566.
Zhang, J., & Zhang, G. (2020). Geochemical and chronological evidence for collision of proto-Yap
arc/Caroline plateau and rejuvenated plate subduction at Yap trench. Lithos, 370,
105616.
Zirakparvar, N. A., Baldwin, S. L., & Vervoort, J. D. (2011). Lu–Hf garnet geochronology applied to
plate boundary zones: Insights from the (U) HP terrane exhumed within the Woodlark
Rift. Earth and Planetary Science Letters, 309(1-2), 56-66.
Zirakparvar, N. A., Baldwin, S. L., & Vervoort, J. D. (2013). The origin and geochemical evolution
of the Woodlark Rift of Papua New Guinea. Gondwana Research, 23(3), 931-943.
Figure 1: Map of the Pangea-Tethys and Panthalassa plate tectonic realms separated by the
Juction Region. Present-day plate boundaries are red (modified from Bird, 2003), relevant
former plate boundaries are gray. Dark shaded areas behind colored realms are present-day
subaerially exposed crust, the light shaded areas are submerged continental crust and
thickened oceanic crust, i.e.,oceanic plateaus and island arcs.
Figure 2: Map of the reconstructed region. A0 geographic and tectonic maps of the region are
provided in the supporting information as Figure S1 and Figure S2. Present-day plate
boundaries are red (modified from Bird, 2003), relevant former plate boundaries are gray.
Marine magnetic anomalies are indicated by black lines, fracture zones are indicated by white
lines (both based on the GSFML database, Matthews et al., 2011; Seton et al., 2014; Wessel et al.,
2015, and references therein). Background image is ETOPO 2022 15 Arc-Second Global Relief
Model (NOAA National Centers for Environmental Information, 2022).
Figure 3: Geographic (A) and tectonic (B) maps of the Philippine Sea Plate region. Present-day
plate boundaries are red (modified from Bird, 2003), relevant former plate boundaries are gray.
Marine magnetic anomalies are indicated by white lines, fracture zones are indicated by black
lines (both based on the GSFML database, Matthews et al., 2011; Seton et al., 2014; Wessel et al.,
2015, and references therein).
AP = Amami Plateau; ASB = Amami-Sankaku Basin; BbIG = Babuyan Island Group; BHP = Bird's
Head Peninsula; BTF = Bewani-Torricelli Fault; BtIG = Batan Island Group; CC = Central
Cordillera; CIR = Caroline Islands Ridge; ELTF = East Luzon Transform Fault; ELTr = East Luzon
Trench; GR = Gagua Ridge; HTr = Halmahera Trench; HB = Huatung Basin; KB = Kikai Basin;
KDB = Kita-Daito Basin; KG = Kerama Gap; LI = Lanyu Island; MAE = Minami-Amami
Escarpment; MDB = Minami-Daito Basin; MS = Molucca Sea; Ne Tr = Negros Trench; OI =
Okinawa Island; OT = Ogasawara Trough; RIG = Romblon Island Group; RMF = Ramu-Markham
Fault; Sa Tr = Sangihe Trench; SCDL = Sindangan-Cotabato-Daguma Linemaent; SMR = Sierra
Madre Range; SP = Snellius Plateau; SSF = Sibuyan Sea Fault; Su Tr = Sulu Trench; WCR = West
Caroline Rise; ZP = Zamboanga Peninsula.
Figure 4: Geographic (A) and tectonic (B) maps of the Melanesian – SW Pacific region. Present-
day plate boundaries in red (modified from Bird, 2003), relevant former plate boundaries in
gray. Marine magnetic anomalies are indicated by white lines, fracture zones are indicated by
black lines (both based on the GSFML database, Matthews et al., 2011; Seton et al., 2014; Wessel
et al., 2015, and references therein).
Figure 5: Geological map of Taiwan, the Philippines, and the northern Molucca Islands. Based
on Ren et al. (2003). FTB = Fold-and-Thrust Belt
Figure 6: Tectonic map of the Philippine Sea Plate, highlighting the different microplates that
formed as a result of oceanic spreading at different spreading centers. Present-day plate
boundaries are red (modified from Bird, 2003), former plate boundaries are white.
KB = Kikai Basin; KDB = Kita-Daito Basin; MAE = Minami-Amami Escarpment; ASB = Amami-
Sankaku Basin; MDB = Minami-Daito Basin; HB = Huatung Basin.
Figure 7: Geological map of New Guinea and the Solomon Islands, based on Davies (2012);
Tejada et
al. (1996); Petterson et al. (1999); Tapster et al. (2014). FTB = Fold-and-Thrust Belt
Figure 10: Tectonic map at 0 Ma. The colors mark the plate tectonic affinity of lithosphere. The
shaded areas behind the colors are continental crust or thickened oceanic crust, i.e., island arcs and
oceanic plateaus. Present-day coastlines are shown throughout the reconstruction (Figures 10 - 33)
for reference.
Figure 27: Snapshot of the kinematic reconstruction at 100 Ma. Note the more southerly
projection compared to the previous figures.
Figure 34: Map of the oceanic basins in the Junction Region colored by their tectonic setting.
Back-arc basins formed above active subduction zones. Marginal basins formed by breaking of a
continental margin in absence of a subduction zone. The marginal basins east of Australia
formed after the end of Mesozoic subduction and before the onset of Tonga-Kermadec
subduction. The subducting plate basins formed at ridges that were created within a downgoing
plate close to a subduction zone.
Figure 35: Maps showing the locations of subduction zones through time in the mantle
reference frame of Doubrovine et al. (2012). A) Izu-Bonin Mariana subduction zone; B) Tonga-
Kermadec(-Vitiaz-Solomon) subduction zone; C) New Guinea-New Caledonia subduction zone.
Figure 36: Reconstructed position of the Lagonoy and Telkhinia subduction zones during the
mid to late Cretaceous in the slab reference frame of Van der Meer et al. (2010) projected on a
seismic tomographic image of the present-day mantle structure at a depth of 1050 km, based on
the UU-P07 tomographic model (Amaru, 2007).
Author biography
Suzanna van de Lagemaat recently received her PhD from Utrecht University. Her PhD
research focused on the plate tectonic evolution of the western and southern Panthalassa
realm, spanning from Japan to Patagonia. Van de Lagemaat uses a combination of field-
based research, including paleomagnetism and geochronology, and extensive literature
research to make kinematic reconstructions of plate tectonics and orogenesis in the
Panthalassa realm during the Mesozoic and Cenozoic. She is also interested in using these
kinematic reconstructions to make inferences about geodynamic processes.
Douwe van Hinsbergen (PhD, Utrecht University, 2004) is full professor of global tectonics
and paleogeography at Utrecht University, the Netherlands, where he has taught since
2012. Van Hinsbergen studies and kinematically reconstructs plate tectonics, orogenesis,
and paleogeography across the globe. He closely collaborates with seismologists to
reconstruct mantle convection and with modelers of geodynamics and paleoclimate to
advance understanding in the physics driving geological processes particularly related to
plate tectonics. Van Hinsbergen is author or co-author of over 200 articles in peer-reviewed
international journals.
GR Focus Review
Timeframe
Location Interpretation Type of data used Anomalies (Ma) Reference
Mariana trough Back-arc extension Marine magnetic anomalies 3n.3 - 0 4.8 - 0 Yamazaki et al. (2003)
Kita Daito Basin Back-arc extension Crustal structures (refraction data) 46 - 41 Ishizuka et al. (2022)
Amami Sankaku
Basin Back-arc extension 49.3 - 46.8 Ma drilled basalts 50 - 46 Ishizuka et al. (2018)
Ayu Trough Subduction Dredged arc lavas > 15 Kamasaki et al. (1996); This study
Philippines
Telkhinia subduction
zone Subduction Paleomagnetic data > 160 - 30 (?) Van de Lagemaat et al. (2023c)
Philippine Mobile
Belt Clockwise rotation Double relay ramp 62 - 45 This study.
Caroline Plate
C16n.1n y -
West Caroline Basin Oceanic spreading Marine magnetic anomalies C11n.2n o 35.58 - 28.0 Gaina and Muller (2007)
C16n.1n y -
East Caroline Basin Oceanic spreading Marine magnetic anomalies C8n.2n o 35.58 - 25.0 Gaina and Muller (2007)
Molucca Sea
New Guinea
Sorong-Yapen-
Bewani Torricelli 300 km left-lateral Taylor (1979); Webb et al. (2019);
Fault strike slip Opening of Manus Basin 3.5 - 0 This study
210 km shortening
related to ophiolite
Central New Guinea emplacement Balanced cross-sections 35 - 21 Martin et al. (2023)
Central New Guinea 190 km shortening of Balanced cross-sections 21 - 9 Martin et al. (2023)
the Australian margin
underneath the
ophiolite
100 km shortening in
the Papuan & Irian
Jaya Fold-and-thrust Hobson (1986); Hill (1991); Martin et
Central New Guinea belt Balanced cross-sections 9-0 al. (2023)
Counterclockwise
rotation New Guinea
Central New Guinea obduction front Avoid overlap with SE Asian Tethysides 30 - 15 This study
Melanesian
Borderlands
C2A -
Bismarck Sea Back-arc extension Marine magnetic anomalies present-day 3.5 - 0 Taylor (1979)
C19n o -
Solomon Sea Back-arc extension Marine magnetic anomalies C16n.1n y 41.18 - 35.58 Gaina and Müller (2007)
Solomon Sea End of spreading New Guinea ophiolite obduction 30 This study
Subducting plate C3A - Weissel et al. (1982); Taylor et al.
Woodlark Basin spreading Marine magnetic anomalies present-day 6.3 - 0 (1990)
Santa Cruz
Basin/South Rennell Subducting plate C19n o -
Trough spreading Marine magnetic anomalies C13n o 40.07 - 30 Seton et al. (2016)
Solomon - Vitiaz - Southward subduction Ages of Solomon and Pentecost Islands Neef and McDougall (1976);
Tonga arc initiation ophiolites 45 Crawford et al. (2003)
SW Pacific
C3A -
Lau Basin Back-arc extension Marine magnetic anomalies present-day 7-0 Yan and Kroenke (1993)
Yan and Kroenke (1993); Van de
Havre Trough Back-arc extension In tandem with Lau Basin 7-0 Lagemaat et al. (2018a)
C4 - present-
North Fiji Basin Back-arc extension Marine magnetic anomalies day 10 - 0 Yan and Kroenke (1993)
C20n o -
North Loyalty Basin Back-arc extension Marine magnetic anomalies C16n o 43.45 - 35 Sdrolias et al. (2003)
Marginal basin
South Loyalty Basin formation Misfit North and South Zealandia 79 - 62 Van de Lagemaat et al. (2023a)
Fairway-Aotea and
New Caledonia Marginal basin Grobys et al. (2008); Van de
basins formation Crustal thickness estimates 92 - 84 Lagemaat et al. (2023a)
Gaina et al. (1998); Grobys et al.
Marginal basin C33n o - (2008); Van de Lagemaat et al.
Tasman Sea formation Marine magnetic anomalies C24n o 95 - 52 (2023a)
Highlights
Declaration of interests
☒ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work
reported in this paper.
☐ The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: