You are on page 1of 11

Ridge Subduction and Slab Windows

Derek J Thorkelson, Department of Earth Sciences, Simon Fraser University, Burnaby, BC, Canada
© 2020 Elsevier Inc. All rights reserved.

Introduction 1
Terminology 1
Growth of the Slab Window Paradigm 2
Identification of Triple Junctions 2
Recognition of Slab Windows 3
Toward a Comprehensive Model 3
Active Slab Windows 4
The Northern Cordilleran Slab Window 5
Fraternal With the Californian Slab Window 5
Origin of Northern Cordilleran Slab Window 5
Degradation of the Southern Margin 5
Volcanism of the Northern Cordilleran Slab Window 5
The Californian Slab Window 5
Introduction 5
Physical Controls on Californian Slab Window Formation 6
Contributions From Thermal Erosion 6
A Unified Perspective of Window Formation 6
Volcanism and Tectonism 6
Relating Crustal Extension to Ridge Subduction 7
The Central American Slab Window 7
The Patagonian Slab Window 7
The Solomon Slab Window 8
Changes in Mantle Conditions and Magmatism 8
Forearc Magmatism and Metamorphism 9
Metamorphism, Magmatism and Deformation in Older Slab Windows 9
Ophiolite Obduction 9
Paleozoic and Precambrian Slab Windows 9
Changes in Elevation and Landscape 10
Recognizing Slab Windows as Components of the Total Earth System 10
References 10
Further Reading 11

Introduction

The Earth’s lithosphere consists of numerous tectonic plates which are separated by three main types of plate boundaries: spreading
ridges, transform faults and subduction zones. Slab windows form where these features occur at the same location, i.e., at sites of
spreading ridge and transform fault subduction. The coincidence of these tectonic features leads to structural, thermal and
geochemical anomalies in the crust and mantle. Four active slab windows are present along the western margins of North America
and South America, and one is located in the western Pacific region. Many more occurred in earlier times, probably beginning with
the onset of plate tectonics in the Precambrian. As such, slab windows have influenced the tectonic development of the continents
and affected the condition of the mantle throughout much of Earth history.

Terminology

The term “ridge subduction,” in the context of slab window formation, refers to subduction of a mid-ocean spreading ridge rather
than an “aseismic ridge” such as the Ninety East or Carnegie ridges. An aseismic ridge is typically a mantle hot-spot track within a
plate, whereas a spreading ridge is a divergent boundary between two oceanic plates.
The term “slab window” refers to a region in the mantle beneath a convergent plate margin that is largely free of subducted
lithosphere (known as “slab”). The term should be restricted to sites of spreading ridge subduction rather than (a) vertical “tears”
within a slab, such as those beneath South America, Greece and western Turkey, and (b) gaps generated by “slab breakoff” such as
those beneath Tibet, central Turkey and the western Pacific. However, localized tearing, breakoff and thermal erosion of subducting
oceanic lithosphere readily occur along the margins of a slab window, leading to the formation of microplates and modification of

Encyclopedia of Geology, 2nd edition https://doi.org/10.1016/B978-0-08-102908-4.00082-5 1


2 Ridge Subduction and Slab Windows

the slab window shape. The term slab window can therefore be applied liberally to complex slab-free areas involving torn and
disintegrating slabs in the vicinity of ridge-trench intersections. The terms “asthenospheric window” and “slab gap” have been used
for a variety of slab-free regions including slab windows.

Growth of the Slab Window Paradigm


Identification of Triple Junctions
The concept of spreading ridge subduction as an important tectonic process began with the identification of plate-boundary triple
junctions in the late 1960s (McKenzie and Morgan, 1969). The principal type of triple junction related to ridge subduction is termed
“ridge-trench-trench” (RTT). A trench is the uppermost part of a subduction zone. The RTT configuration refers to three tectonic
plates, two of which are diverging from one another along a spreading ridge and are concurrently subducting beneath the third plate
at a trench, which can be thought of as two trenches, one on each side of the triple junction (Fig. 1A). Another relevant triple
junction is termed “ridge-trench-transform fault,” (RTF) (Fig. 1B). It refers to a configuration in which a ridge intersects a continental

Triple Junctions and Slab Window Environments


(A) (B)
A A
SW
SW
F
T T T
R
R

B C B C

(C) (D)
A A
SW

SW
OC
T F T T D
B R
B
R

C C

Legend
Extensional
Surface structures and
projection A Overriding Oceanic rotation of
plate A
R

of subducted spreading ridge overriding plate


oceanic plate with motion in divergent (D)
Oceanic vector (full environment
Motion vector B plate B or C spreading rate)
relative to on sea floor Original coastline
overriding plate OC prior to ridge
subduction
Transform or
F Triple junction
Trench (upper part transcurrent
T
of subduction zone) fault SW Slab window
Fig. 1 Examples of triple junctions related to ridge subduction and slab window formation. (A) RTT (ridge-trench-trench) environment where both oceanic plates
B and C are subducting beneath overriding plate A. Slab window (SW) forms between the downgoing slabs. Degradation of slab window edges shown by irregular
linework. (B) RTF (ridge-trench-transform fault) environment where oceanic plate B is subducting but oceanic plate C is moving parallel to coastline. (C) TTF (trench-
trench-transform fault) environment assuming subduction of fault F was preceded by subduction of a spreading ridge segment, forming slab window. (D) RTD (ridge-
trench-divergent) environment generated by subduction of ridge and unstable interaction of plate C with overriding plate A. As ridge is subducted, plate C comes into
contact with, but moves away from, overriding plate A, causing plate A to extend outward until coastline becomes parallel to plate C motion vector (shown), with
possible addition of a spreading ridge (as in the Gulf of California). Slab window underlies entire region to the right of triple junction. Based on McKenzie DP and
Morgan WJ (1969) Evolution of triple junctions. Nature 224(5215): 125–133 and Thorkelson DJ (1996) Subduction of diverging plates and the principles of slab
window formation. Tectonophysics 255(1–2): 47–63.
Ridge Subduction and Slab Windows 3

margin with a trench on one side of the triple junction, and a transform fault (more acceptably termed a transcurrent fault in this
case) on the other side. A third case, termed TTF, occurs where a transform fault that offsets a spreading ridge intersects a trench
(Fig. 1C). A fourth is a special version of an RTF in which one of the two oceanic plates moves away from the overriding plate
(Fig. 1D). This configuration can be abbreviated RTD, where “D” refers to “divergent” in order to designate the particular kinematics
of this situation. This type of configuration occurred along coastal California in Neogene time, as described below. Other, more
complex configurations are possible, and styles may change over geological time as spreading and subduction continue, and the
location of the triple junction changes.

Recognition of Slab Windows


The idea that spreading ridge subduction could lead to a gap between subducted plates was formulated in the 1970s from studies of
Japan, the Aleutian Islands and Sumatra (Uyeda and Miyashiro, 1974; Marshak and Karig, 1977; DeLong et al., 1979). The gap was
identified as a locus of anomalously high heat flow, especially in the forearc. These early studies formed the basis for two landmark
papers on the consequences of ridge subduction beneath southwestern North America (Dickinson and Snyder, 1979) and southern
South America (Forsythe and Nelson, 1985). The authors demonstrated that subduction of a spreading ridge would generate a slab
window with a predictable size and shape, based on the configuration and motion of the three tectonic plates in the triple junction
area. They showed that (a) plate growth at a spreading ridge will stop as the ridge enters the subduction zone, and (b) the
subducting spreading ridge will split apart and expand into a slab window. The outcome is an “unzipping” of the divergent plate
boundary as it descends into the mantle. The authors also demonstrated that activity in an existing chain of arc volcanoes, such as
those in the Andes, may be interrupted above a slab window and supplanted by a broader volcanic field with different chemical and
physical characteristics, such as mafic alkaline lavas and cinder cones.

Toward a Comprehensive Model


In order to visualize the range of possible slab window styles and shapes, idealized diagrams of slab windows were developed using
principles of plate kinematics and structural geology. Shapes, sizes and motions of slab windows were predicted for a variety of plate
configurations, yielding a series of type-examples that could be applied to real-world situations (Thorkelson, 1996). This approach
was coupled with an explosion of information from case studies of ridge subduction and advancements in understanding the
lithosphere-asthenosphere system. The result was a new paradigm of slab window tectonics that merged plate kinematics with
aspects of metamorphism, magmatism, and mantle dynamics (Sisson et al., 2003). The main features of slab windows, as currently
understood, are provided in Fig. 2.

Fig. 2 Block diagram of slab window paradigm showing selected features associated with ridge subduction as described in text. Two oceanic plates (green and
blue) are subducting beneath an overriding plate (transparent grey). Spreading ridge intersects leading edge of overriding plate at an RTT triple junction.
Metasomatism-related melting of mantle wedge above the subducting plates yields volcanic arcs. Intervening “slab window volcanoes” occur (1) in forearc due to
blow-torch effect (red), (2) above slab edges due to partial melting of subducted oceanic crust (pink; potentially adakitic in composition) and (3) above slab window in
response to decompression-melting of upwelling sub-slab mantle (purple). Edges of subducting plates are drawn in two styles, with blue plate showing
“geometrical” edge based on plate kinematics, the tenet that plate growth ceases at the triple junction, and the (unlikely) assumption that the slab does not undergo
further changes in shape. The green plate shows a more plausible “effective edge” which bears the effects of heating, partial melting and fragmentation. Modified
from Thorkelson DJ (1996) Subduction of diverging plates and the principles of slab window formation. Tectonophysics 255(1–2): 47–63; Sisson V, Pavlis T, Roeske
S and Thorkelson D (2003) Introduction: An overview of ridge-trench interactions in modern and ancient settings, In: Geology of a Transpressional Orogen Developed
During Ridge-Trench Interaction Along the North Pacific Margin, vol. 371, pp. 1–18. Geological Society of America and Thorkelson DJ and Breitsprecher K (2005)
Partial melting of slab window margins: Genesis of adakitic and non-adakitic magmas. Lithos 79(1–2): 25–41.
4 Ridge Subduction and Slab Windows

Active Slab Windows

Slab windows that occur at sites of modern ridge-trench interactions are considered “active.” They include the Northern Cordilleran
slab window, the Californian slab window, the Central American slab window, the Patagonian slab window, and the Solomon slab
window (Fig. 3; Table 1).

Fig. 3 Plate tectonic map showing active slab windows (red) and locations that are likely to have been affected by previous Cenozoic slab windows (pink) beneath
the Americas and Antarctica Peninsula. Thorkelson DJ, Madsen JK and Sluggett (2011) Mantle flow through the Northern Cordilleran slab window revealed by
volcanic geochemistry. Geology 39: 267–270. https://doi.org/10.1130/G31522.1, courtesy Geological Society of America.

Table 1 Active slab windows.

Slab window Age of Plate tectonic context


Name and location activity

Northern Cordilleran Oligocene- Slab window was produced by subduction of the East Pacific rise beneath the North American plate. Window is
slab window Present flanked to the south by the subducting Juan de Fuca plate and the stagnant Explorer microplate, and to the
Western Canada northwest by the Pacific plate. Juan de Fuca and Explorer plates are fragments of the former Farallon plate.
Northern Cordilleran slab window is fraternal with the Californian slab window.
Californian slab window Oligocene- Slab window occurred from subduction of the East Pacific rise beneath the North American plate. Window is flanked
Southwestern United Present to the southwest by the Monterey, Guadalupe, Magdalena and Rivera microplates, and to the northwest by the
States Juan de Fuca plate. Window shape was strongly influenced by the formation and rotation of the microplates, which
are fragments of the former Farallon plate. The Californian slab window is fraternal with the Northern Cordilleran
slab window.
Central American slab Miocene- Slab window was generated by subduction of the Nazca-Cocos spreading ridge beneath the Caribbean plate.
window Present Window is flanked to the south by the Nazca plate and to the north by the Cocos plate. Window formed during
Panama emergence of the Isthmus of Panama and subduction of the aseismic Cocos ridge, a track of the Galapagos
hotspot.
Patagonian slab Oligocene- Slab window was produced by subduction of the Nazca rise beneath southern South America. Window is flanked to
window Present the north by the Nazca plate and to the south by the Antarctic plate. Concurrent ophiolite obduction and formation
Southern Chile and of the Chonos and Cabo Elena microplates modified the window slightly. Window formation was partly
Argentina synchronous with growth of the Antarctic Peninsula slab window, which became inactive in the Pliocene.
Solomon slab window Miocene- Slab window was caused by subduction of diverging oceanic crust of the Woodlark basin beneath the Solomon
Solomon Islands Present islands, east of New Guinea. The subducting plates are the Woodlark plate, a small and probably ephemeral plate,
and the Australian plate. The area of oceanic crust being subducted and the slab window are small in comparison
to the other active slab window localities.
Ridge Subduction and Slab Windows 5

The Northern Cordilleran Slab Window


Fraternal With the Californian Slab Window
The Northern Cordilleran and Californian slab windows (Fig. 3) occurred from subduction of the East Pacific rise beneath western
North America. The slab windows are “fraternal” because they grew in response to concurrent subduction of the same ridge-
transform system at separate locations (Thorkelson and Taylor, 1989; Thorkelson, 1996). This configuration developed because the
spreading ridge is offset by long transform faults, most notably the Mendocino transform. The two slab windows grew synchro-
nously, separated by a large fragment of the former Farallon plate, the Juan de Fuca plate (Fig. 3). The two slab windows share many
similarities, although the Californian slab window is tectonically more complicated.

Origin of Northern Cordilleran Slab Window


Opening of the Northern Cordilleran slab window began when the East Pacific rise (spreading ridge) intersected the continental
margin of British Columbia, western Canada, in the Eocene to early Oligocene. The spreading ridge was flanked by the Pacific plate
to the west and the Farallon plate to the east. This intersection occurred after a previous regime of plate interactions involving the
Kula and Farallon plates. The Pacific plate moved northward and the Farallon moved eastward, and the slab window formed as an
expansive gap, emanating from near the northern tip of Vancouver Island (Madsen et al., 2006). The northward-moving Pacific
plate moved in a strike-slip manner along the coastline until it reached the Aleutian trench where it descended into the subduction
zone below Alaska and the Aleutian Islands. In contrast, the Farallon plate subducted eastward along the coastline from southern
British Columbia to California. The Farallon plate was largely consumed and, along this part of North America, became limited to
the Juan de Fuca plate and a daughter microplate, the Explorer plate. The Explorer plate is the broken-off northern tip of the Juan de
Fuca plate, and its separation occurred in the Pliocene. It is now jammed under the continental margin, nearly stationary and
tectonically stagnant (Polat et al., 2018).

Degradation of the Southern Margin


Tectonic modeling of the Northern Cordilleran slab window by Madsen et al. (2006) showed that the northern edge of the Juan de
Fuca (including Explorer) plate should extend from near the northern tip of Vancouver Island northeastward, toward central British
Columbia. This “geometrical” edge was constructed using Cenozoic plate configurations and kinematics (e.g., Wilson, 1988), plus
the assumption that the plate edge (slab window margin) remained unaffected by physical or thermal degradation. Subsequent
tomographic modeling by Mercier et al. (2009) showed that the actual or “effective” edge of the Juan de Fuca plate lies significantly
south of the modeled geometrical edge (Thorkelson et al., 2011), and appears to have an irregular shape with a maximum depth of
300 km. The difference between the geometrical and effective slab edge is attributed to a combination of thermal erosion and
physical disintegration during the downward passage of the slab through the hot mantle, resulting in tearing, crumbling and partial
melting of the slab edge (cf., Thorkelson and Breitsprecher, 2005).

Volcanism of the Northern Cordilleran Slab Window


Volcanism associated with the Northern Cordilleran slab window varies profoundly from one side to the other. Presently, the Juan
de Fuca plate is vigorously subducting beneath southernmost British Columbia and the northwestern United States, yielding the
Cascade-Garibaldi volcanic arc. Far to the north, subduction of the Pacific plate beneath Alaska is generating the Aleutian volcanic
arc. In the intervening region, above the Northern Cordilleran slab window, the volcanism is different, both chemically and
physically (Thorkelson et al., 2011). Whereas the two volcanic arcs consist mainly of explosive, steep-sided stratovolcanoes with
mainly calc-alkaline compositions, the slab window volcanic field consists of more mafic, less explosive volcanoes with gentler
slopes and more alkalic compositions. Prominent volcanic regions include Mount Edziza in northern British Columbia and the
Chilcotin Group in southern British Columbia. The slab window volcanic field has an “intraplate” character, broadly similar to
some continental rifts and some hotspot-generated ocean islands, and is sandwiched between normal volcanic arcs. In addition, the
volcanic arc to slab window transition is marked by the presence of adakites, which are volcanic rocks with compositions consistent
with melting of the subducted oceanic crust. Adakites are present in the Wrangel volcanic field above the eastern edge of the
subducted Pacific plate, and in the Garibaldi field above the northern edge of the Juan de Fuca plate. This arrangement is remarkably
consistent with the predictive view of adakite melt generation along the margins of slab windows as established in theoretical and
empirical studies (Kay et al., 1993; Thorkelson and Breitsprecher, 2005) and with the geophysically imaged degradation of the Juan
de Fuca plate.

The Californian Slab Window


Introduction
Formation of the Californian slab window (Fig. 3) occurred from subduction of the East Pacific rise (Farallon-Pacific spreading
ridge) in the Cenozoic (Dickinson and Snyder, 1979). The process began where a segment of the rise, between the Pioneer and
6 Ridge Subduction and Slab Windows

Murray fracture zones, approached the continental margin along the Californian coast at approximately 30 Ma. Ridge subduction
progressed in a complex manner to the north and south, opening a huge slab-free region beneath California and northern Mexico.
As with its fraternal slab window to the north, the Californian slab window remains active and continues to affect the North
American plate. The ridge-continent interaction also resulted in capture of a sliver of the continent by the Pacific plate through
formation of new sea floor in the Gulf of California, separation of Baja California from the rest of Mexico, and the formation of the
San Andreas fault.
Despite its fundamental simplicity, the history of East Pacific rise subduction is complex. Entry of the actual ridge into the trench
was impeded by local changes to the plate configuration along with physical and thermal degradation of the downgoing Farallon
slab. As such, the original “geometrical” shape of the slab window has been supplanted by more advanced versions (Wilson et al.,
2005), although some details of window evolution remain uncertain.

Physical Controls on Californian Slab Window Formation


As the ridge approached the continental margin, the age of the Farallon plate entering the trench grew younger (with “zero age” at
the spreading ridge). The youthful parts of the downgoing plate were relatively hot, thin, weak and buoyant. Consequently, the
deeper, denser parts of the subducting Farallon slab began to locally pull away, gradually stranding a shallower portion of the
Farallon, forming the Monterey microplate (Lonsdale, 1991; Wilson et al., 2005). Part of this orphaned fragment remained
unsubducted on the sea-floor, with the rest jammed beneath the continental margin, a process which parallels the current plight
of the Explorer plate, discussed earlier. The process of plate rupture and microplate formation continued locally to the south,
leading to the formation of the Guadalupe, Magdalena and Rivera microplates (Fletcher et al., 2007). As such, ridge subduction per
se occurred only where microplates did not form, and is specifically responsible for only part of the overall slab window shape.

Contributions From Thermal Erosion


In addition to plate tearing and the stranding of microplates, the process of thermal erosion has played an important role.
Severinghaus and Atwater (1990) provided an extreme view in which a large region of the downgoing Farallon slab was heated
to the point where it was effectively resorbed into the mantle, leaving a large “slab gap” between older and younger parts of the
Farallon plate. A supportive but more modest model of thermal erosion, in which only the fringes of subducting slabs are
thoroughly degraded, was provided by Thorkelson and Breitsprecher (2005). Consequently, the degree to which thermal erosion
has contributed to the window beneath California and adjacent Mexico is not precisely known.

A Unified Perspective of Window Formation


The Californian slab window is the product of three main process: (1) ridge subduction, (2) slab tearing and microplate formation,
and (3) thermal erosion. Attempts to subdivide the current, vast slab-free region into specific domains generated by each of these
processes could be pursued; however, the perspective offered here is that this slab window is the product of several styles of slab
removal and no-growth acting synchronously to yield a slab-free breach of a previously intact and long-lived subduction zone. Each
of these processes was unquestionably driven by the approach of the spreading ridge to, and locally into, the subduction zone,
consistent with other active slab windows. The complex mechanical and thermal history of this case broadens the view of how ridges
and trenches can interact and expands the applicability of the slab window paradigm (Fig. 2).

Volcanism and Tectonism


The Neogene volcanic activity in the southwestern United States is as complex as its tectonic history. However, some fundamental
observations regarding distribution and composition illuminate the importance of the slab window on mantle and crustal
evolution. First, the Cascade volcanic arc lies only above the Juan de Fuca plate, with the southernmost volcano, Lassen Peak,
sitting above the subducted Mendocino transform fault, i.e., at the northern limit of the Californian slab window. The arc comprises
a chain of steep-sided stratovolcanoes with strong subduction-related geochemical compositions. As such, the Cascades are
comparable to volcanic arcs elsewhere in the world, including the Aleutian arc in Alaska and the Andes in South America.
South of the Cascade arc, numerous volcanoes form a broad field extending from the Californian and Mexican coastline as far
inboard as New Mexico. Overall, the volcanoes have less volcanic arc or “supra-subduction zone” character, both chemically and
physically. Relative to the Cascade volcanoes, those above the slab window are typically smaller, with gentler profiles and
compositions closer to rift or ocean-island volcanoes than volcanic arcs (Cole and Basu, 1992). As such, this field is broadly
analogous to the slab window volcanic field above the Northern Cordilleran slab window. However, volcano compositions above
the Californian slab window are more varied, with mafic volcanoes, such as those in the Californian coastal volcanic field and the
Pinacate volcanic field of Mexico and Arizona, scattered among intermediate to felsic eruptive centers, such as the Long Valley
caldera in California and the Chiricahua ignimbrite field in Arizona.
The Neogene Jemez volcanic field of New Mexico lies 1000 km from the California coastline. Whether it and other far-inland
volcanoes should be included in the slab window field is debatable because other factors, such as rifting along the Rio Grande
Ridge Subduction and Slab Windows 7

valley, may be viewed as more important controls. However, the issue of how far the effects of a slab window can extend is not
unique to this region.

Relating Crustal Extension to Ridge Subduction


The area above the California slab window has undergone large-magnitude extensional tectonism, forming the Basin and Range
geological province (Spencer et al., 1995), strike-slip fault zones along the San Andreas and Walker Lane corridors, and localized
zones of mantle delamination. All of these processes have complicated any simple signal that may have been generated by the slab
window alone. For that reason, volcanism in the Californian slab window region is more difficult to interpret than above the
Northern Cordilleran slab window where post-Eocene deformation was minimal. Nevertheless, the fact that North America was
undergoing widespread extension and magmatism means that decompression-melting of the underlying mantle may have been a
principle cause of mafic magma flux into the crust, and subsequent crustal melting, and these processes could be viewed as
completely independent from any changes to mantle and crust imparted by the window.
In a general way, however, the extensional tectonism of the southwestern United States and northern Mexico are related to slab
window formation because the extension was largely triggered by a transition from Farallon plate subduction to Pacific plate-driven
transtension. The Pacific plate, at the time of its contact with North America in the Oligocene, was moving slightly away from the
continent, and its divergence led to stretching and thinning of the North American plate (the RTD style of triple junction, as
described previously). The Pacific plate came into contact with North America via intersection between the East Pacific rise and the
North American trench, and hence both slab window formation and extensional tectonism sprang from the same basic tectonic
phenomenon. As such, extensional and transcurrent tectonism in the overriding plate are regarded as potential consequences of
ridge subduction and fall within the slab window paradigm (Fig. 2).

The Central American Slab Window

The Central American slab window (Johnston and Thorkelson, 1997) reflects subduction of the diverging Cocos and Nazca oceanic
plates beneath Central America (Fig. 3). The Nazca-Cocos ridge, which separates the two oceanic plates, formed in the Oligocene
and underwent two significant reorganizations. Despite these adjustments, the ridge and its transform faults subducted nearly
continuously from the Oligocene to the present day, with the Nazca-Cocos-Caribbean triple junction positioned along the coastline
of Costa Rica and Panama (McGirr et al., 2020).
Two slab windows (McGirr et al., 2020), or one slab window with two major dilation zones, formed from the Miocene to the
present. Opening of the slab window(s) triggered two changes in the geochemical composition of the Central American volcanic arc.
The first was a decrease in “arc character,” with volatile-sensitive trace element ratios such as Ba/La progressively dropping from
Guatemala to Costa Rica and Panama. These variations were accompanied by progressively lower 143Nd/144Nd, and together they
indicate a gradual southward enrichment in the composition of the mantle source, from arc-type to ocean-island-type. The second
change was the eruption of adakitic lava in southern Costa Rica and Panama, near the slab window margin. Genesis of these
adakites is consistent with melting of the Cocos and Nazca plate edges.

The Patagonian Slab Window

Subduction of the Chile rise, the spreading ridge-transform system between the Nazca and Antarctic plates, is taking place beneath
southern South America (Fig. 3). The ridge subduction was preceded by millions of years of normal subduction of the Nazca plate.
Subduction of the ridge began in the Miocene and continued unabated to the present day (Forsythe and Nelson, 1985). The first
intersection occurred between a segment of the spreading ridge and the South American trench, followed by alternating transform
and ridge segments. The initial intersection likely formed a single RTT triple junction, but as subduction proceeded it divided into
two triple junctions, with one migrating northward and the other southward (Breitsprecher and Thorkelson, 2009). As the triple
junctions moved apart, the Antarctic plate began to subduct in the intervening region, and the Nazca plate became restricted to the
north. A large slab window formed as successive ridge segments entered the trench, subducted and separated. Tearing of the
subducting lithosphere resulted in microplate formation in at least two locations but was overall modest in its effect. The
Patagonian case stands as a modern example of a slab window generated principally by ridge subduction.
The distribution and timing of volcanism in Patagonia is closely linked to the history of slab window formation. Outpourings of
mainly mafic lava with ocean-island-like compositions formed several large volcanic regions such as the Pali Aike and Meseta del
Lago Buenos Aires fields in Argentina (Gorring et al., 2003). The location of these “plateau lavas” broadly reflects the expansion of
the slab window from the Miocene to the present day. To the north of the slab window lie arc volcanoes of the Andes. Mt. Hudson
sits at the southern end of this volcanic chain, just north of the modern slab window. As such, Mt. Hudson is equivalent to Mt.
Lassen, which lies at the southern end of the Cascade arc in northern California. To the west of the Patagonian slab window lies the
Austral volcanic zone, which is largely adakitic in composition. It was generated mainly through partial melting of the leading edge
of the newly subducted Antarctic plate (Stern and Kilian, 1996). In several million years from now, the Antarctic plate will have
8 Ridge Subduction and Slab Windows

subducted deeply enough to trigger normal arc magmatism, generating a volcanic arc slightly to the east of the current Austral zone
(Breitsprecher and Thorkelson, 2009).

The Solomon Slab Window

The Woodlark basin lies east of New Guinea and southwest of the Solomon Islands. The basin is smaller than California and is
underlain by oceanic lithosphere, which is divided by the Woodlark spreading center, a spreading ridge that is offset by several
transform faults. The ridge-transform system separates the Woodlark plate to the north from the Australian plate to the south. The
oceanic lithosphere and its ridge are subducting northeastward along the San Cristobal trench beneath the Solomon Island arc on
the Pacific plate, forming the Solomon slab window (Chadwick et al., 2009). The current subduction zone formed in response to
jamming of the Melanesia trench to the north of the Solomon Islands by the Ontong Java oceanic plateau. The slab window began
to form in the Miocene.
The distribution of earthquake hypocenters indicates that subduction beneath the Pacific plate is occurring at a moderate angle
involving downgoing slabs that have not yet descended far into the asthenosphere. Compared to many sites of subduction, the San
Cristobal trench is neither bathymetrically deep nor well defined. These attributes suggest that the young lithosphere of the
Woodlark basin is buoyant and resistant to deep subduction (Mann et al., 1998) and that the slab window will become inactive
within a few million years.
The New Georgia group of islands, part of the Solomon arc, overlies the subducting ridge. Volcanoes of New Georgia record the
eruption of two types of lava with distinctive compositions (Schuth et al., 2004). One is basalt with characteristics similar to mafic
lavas at other island arc volcanoes. The other is picrite, a relatively rare form of basalt with unusually high concentrations of
transition metals, particularly Mg, Ni and Cr. Both the basalt and picrite display low Nb/La and Ta/Th, which indicate that the
magmas were generated by partial melting of mantle which had been hydrated by subduction. However, the high transition metal
contents of the picrites reflect anomalously high mantle temperatures and high percentages of melting. The high temperatures were
generated by the uprise of hot asthenosphere through the slab window, displacing the mantle wedge, triggering anomalously high
degrees of melting of the subduction-metasomatized peridotite.

Changes in Mantle Conditions and Magmatism

The effects of slab windows as described above are strongly tied to changes in the mantle environment as ridge subduction replaces
normal subduction. Changes to the mantle typically involve temperature, composition, and patterns of flow.
As ridge subduction proceeds, the previously subducted slab from one or two plates moves away and is replaced by mantle
material (mainly peridotite) in the slab window (Fig. 2). As this occurs, the supra-slab mantle (mantle wedge) comes into direct
contact with sub-slab mantle. The supra-slab mantle is typically hydrated and chemically enriched in incompatible elements,
notably the alkalies and alkaline earth elements, from the expulsion of melt and aqueous fluids from downgoing oceanic
lithosphere. The sub-slab mantle is typically drier, hotter, and relatively depleted in incompatible elements, but may also contain
enriched peridotite from mantle plumes, or streaks of pyroxenite or eclogite.
A slab window is a breach in an otherwise extensive slab barrier that separates the sub- and supra-slab mantle (mantle wedge)
reservoirs. The window thereby permits passage of mantle from one reservoir into the other, i.e., flow of sub-slab mantle into the
former mantle wedge, or vice versa. Examples of both exist, and together they underline the variability of mantle flow patterns
through slab windows.
Upward vertical flow of sub-slab mantle is evident in all four of the slab windows along the Americas. In the Northern
Cordilleran slab window, upwelling of sub-slab mantle in the Neogene displaced the mantle wedge that was previously metaso-
matized during the Paleogene and Mesozoic (Thorkelson et al., 2011). The ascent of drier, sub-slab mantle is indicated by the
Eocene-Miocene transition in volcanic geochemistry. Upwelling is also evident in the Californian slab window, although the degree
to which the up-flow was generated by a broad pattern of flow through the slab window rather than convection generated by
extension and/or delamination of the overriding lithosphere is difficult to quantify. Nevertheless, the Neogene transition from arc-
like to intraplate-like compositions in the southwestern United States implies the involvement of mantle that was little affected by
slab metasomatism and is consistent with the rise of sub-slab mantle.
In the Central American and Patagonian slab windows, horizontal mantle flow accompanied upwelling. In Central America,
up-flow carried sub-slab mantle that was chemically modified by the Galapagos hotspot into the mantle wedge of the Central
American volcanic arc, leading to a “plume” geochemical signature in the volcanoes near the window. The geochemical trends show
that the mantle flowed upward and then laterally for hundreds of km beneath the Caribbean plate and the Central American arc,
“contaminating” the magma source region.
In Patagonia, the up-flowing sub-slab mantle has plume-like character, as indicated by the composition of Neogene plateau
lavas in Chile and Argentina (Gorring et al., 2003). The source of the plume-like compositions is not well understood. Upwelling of
the sub-slab mantle was accompanied by westward flow of the supra-slab mantle. Arc-metasomatized asthenosphere that originated
under the South American continental margin passed beneath the Nazca and Antarctic plates for hundreds of km, contaminating
the melt source-region of the southern Chile rise (Klein and Karsten, 1995).
Ridge Subduction and Slab Windows 9

Lateral mantle flow from the mantle wedge to beneath the subducting plates is also evident in the Solomon slab window region
(Chadwick et al., 2009). Volcanoes in the Woodlark basin, which consists of two plates subducting beneath the Solomon Islands,
have arc-like compositions that were derived from the Solomon arc mantle wedge.

Forearc Magmatism and Metamorphism

Where a spreading ridge descends beneath the overriding plate, the forearc region may undergo anomalous heating, fluid activity
and magmatism. This process, termed the “blow-torch effect” (DeLong et al., 1979), is expected to be most pronounced near the
triple junction, where the subducting ridge is narrow and heat transfer from the underlying mantle is strongly focused (Fig. 2).
Heating, metamorphism, and thermal weakening of the forearc may last for several million years during and after the passage of a
subducting ridge (Groome and Thorkelson, 2009).
Forearc magmatism is evident in some of the active slab windows. Above the Northern Cordilleran slab window, Neogene
volcanoes and shallow intrusions are located on western Vancouver Island and the Haida Gwaii (Queen Charlotte Islands). Above
the Californian window, Neogene volcanoes occur in the Coast Ranges of California and Baja California. In the Solomon arc—
Woodlark basin, volcanoes extend from the main Solomon arc toward the trench, and beyond the forearc into the Woodlark basin,
defining a volcanic field produced by a combination of arc and mid-ocean ridge processes. Near-trench intrusions also occur near
the Taitao Peninsula in Chile.
Igneous activity near triple junctions is likely to begin with injection of mafic dikes into the forearc, followed by crustal melting,
assimilation and mixing. These processes account for the mafic to felsic range of compositions in near-trench ridge subduction
settings.
Metamorphism at active slab windows is difficult to assess because rocks in those locations have not been markedly exhumed,
and the processes and products of metamorphism remain deeply buried.

Metamorphism, Magmatism and Deformation in Older Slab Windows


A complex history of slab windows beneath western North America has been proposed for intervals in the Late Cretaceous to Eocene
(Thorkelson and Taylor, 1989; Haeussler et al., 2003; Sisson et al., 2003; Madsen et al., 2006). The windows were generated by
subduction of rapidly migrating ridges that separated the Farallon, Kula and Resurrection plates. Studies of near-trench regions of
Alaska and British Columbia indicate that ridge subduction generated metamorphism and fluid activity under higher temperatures
than normally associated with forearc environments. Magmas derived from both mantle and crustal sources were emplaced as
plutons, and fluids with locally economic gold abundances were emplaced as veins (Fig. 2). Ductile and brittle deformation were
punctuated by intervals of rapid exhumation. These features are broadly consistent with the expected outcomes of anomalous
heating and stress fields in a forearc region subjected to subduction of a rapidly migrating spreading ridge. However, the precise
location of the ridge-trench intersections, and their spatial and temporal ties to the ascribed features, cannot be independently
demonstrated through unique plate reconstructions because too much of the ocean floor has since been subducted, and the record
of plate shapes and kinematics has been lost. Nevertheless, the style of metamorphism, magmatism and deformation expressed
along this continental margin serves as a plausible record of ridge subduction with probable application to other ancient slab
windows.
Ridges separating the Pacific, Izanagi and Kula plates subducted beneath eastern Asia in Mesozoic to Paleogene time (Kinoshita,
1999). Studies on metamorphism and magmatism in Japan showed the development of magmatic and metamorphic belts that may
have been generated by ridge subduction. Tomographic and plate kinematic modeling confirm an Eocene event of Pacific-Izanagi
ridge subduction which led to widespread geochemical changes to the mantle beneath the Asian continental margin (Wu and Wu,
2019). Subduction of other ridges also occurred in the southeast Pacific region. The Farallon—Phoenix ridge subducted somewhere
along the coast of Chile in Late Cretaceous to Neogene time, and the Antarctic—Phoenix ridge subducted beneath the Antarctic
Peninsula in Paleogene and Neogene time, ending in the Pliocene (Hole and Larter, 1993).

Ophiolite Obduction

The process of ridge subduction may be conducive to the obduction of ophiolites (Fig. 2). A spreading ridge is bathymetrically
higher, hotter and weaker than the flanking older ocean floor. As a ridge encounters a trench, it will have a greater likelihood of
becoming detached from the oceanic plate(s) and thrust onto the forearc of the overriding plate. This process was apparently the
cause of obduction of the Resurrection ophiolite in southeastern Alaska (Kusky and Young, 1999), and possibly the Taitao ophiolite
of southern Chile (Forsythe and Nelson, 1985).

Paleozoic and Precambrian Slab Windows

Numerous events of ridge subduction and slab window formation occurred early in Earth history. For example, in the Central Asian
Orogenic Belt, slab windows of Paleozoic to early Mesozoic age have been proposed to account for patterns of igneous geochemistry
10 Ridge Subduction and Slab Windows

that differ from those related to normal subduction. Plate tectonics started in the Precambrian, possibly in the Archean. Higher heat
flow in the Precambrian may have led to more vigorous plate activity including a greater number of ridge-trench encounters.
As such, ridge subduction would have played an important geodynamic role in the early Earth, possibly triggering “blooms” of
granitoid formation, particularly those with trondhjemite or tonalite compositions that are common in Archean cratons. The effects
of ridge subduction on the long-term evolution of the crust and mantle are not fully understood, but are likely to be profound,
particularly if plate tectonics began in the Archean.

Changes in Elevation and Landscape

The opening of a slab window may induce uplift of the overriding plate. Studies in Patagonia, the Antarctic Peninsula and Alaska
collectively indicate uplift of hundreds of meters to perhaps kilometers. In British Columbia, uplift may also have been driven by
slab window formation, as the region above the slab window is topographically high. Uplift related to the California slab window
may also have occurred, but the region is complicated by a history of pronounced extension and transcurrent faulting, so isolating
the potential slab window effects from other factors is problematic.
Uplift above slab windows may be the result of three main factors. First, the flow field of the mantle will be different from that
above an arc because the absence of slabs will reduce or eliminate corner flow. Instead, the flow field may be dominated by
upwelling and/or lateral (horizontal) mantle flow. Second, the temperature of the mantle will rise, as mantle that has not been
chilled in a subduction environment will gradually displace mantle formerly located in the mantle wedge. Third, the stress field in
the overriding plate will likely be different, as two oceanic plates, not one, will be present along the leading edge of the overriding
plate. This situation may lead to either an increase or decrease in compressional stress and may produce different stress fields on
opposite sides of the triple junction, leading to transcurrent faulting. Each case of ridge subduction is unique and needs to be
examined individually to determine how slab window formation is affecting features such as topography, drainage and crustal
exhumation.

Recognizing Slab Windows as Components of the Total Earth System

Slab windows are unique and important features that can affect a region in a variety of ways, from mantle flow patterns to magma
genesis and geomorphology. However, it is critical to realize that a slab window forms only part of an overall geological
environment. The overriding plate is likely to be affected by structural, thermal and magmatic processes that are independent
from, or blended with, the effects of a slab window. Which magmatic and tectonic features should be ascribed to slab window
formation, and which are better rationalized by other processes, are complex questions requiring nuanced answers.
The southwestern United States is prime example of a region which, during slab window formation, has been affected by a range
of geological processes. These include crustal extension, mantle delamination, strike-slip faulting, microplate formation and capture
of a continental sliver by an invasive spreading ridge. All of these processes are arguably linked to interactions among the North
American, Farallon and Pacific plates, but to what degree they should be ascribed to subduction of the East Pacific rise and growth of
the Californian slab window is a matter of judgment and emphasis. An additional complication is the pre-Oligocene geological
history of the region involving magmatism and contractional, transcurrent and extensional deformation, dating back to the
Precambrian. Accordingly, the early Cenozoic North American lithosphere varied spatially in age, composition, fabric and strength,
and all of these attributes undoubtedly played roles in the late Cenozoic record of magmatism, deformation and landscape
evolution. Nevertheless, the most consistently identified effect of slab windows is a change in mantle conditions, from a hydrous
mantle wedge during subduction to drier, hotter asthenosphere in response to slab window formation. It is difficult to imagine that
the geological evolution of the southwestern United States was not affected by this fundamental change.
An additional challenge in the study of ridge subduction is the integration of new geological and geophysical information. For
example, the shape of a slab window as originally proposed may need to be redrawn to accommodate newly recognized thermal
erosion or microplate formation, or improvements in plate reconstructions. Perceived mantle flow patterns may need to be adjusted
to accommodate new measurements of mantle velocity or anisotropy, or the depth and source of magma generation. The
progression of thought on how ridge subduction in the broadest sense affects a geological region brings us closer to a more
complete understanding of how plate tectonics operates and has affected continental history and planetary evolution.

References
Breitsprecher K and Thorkelson DJ (2009) Neogene kinematic history of Nazca–Antarctic–Phoenix slab windows beneath Patagonia and the Antarctic Peninsula. Tectonophysics
464(1–4): 10–20.
Chadwick J, Perfit M, McInnes B, Kamenov G, Plank T, Jonasson I, and Chadwick C (2009) Arc lavas on both sides of a trench: Slab window effects at the Solomon Islands triple
junction, SW Pacific. Earth and Planetary Science Letters 279(3–4): 293–302.
Cole RB and Basu AR (1992) Middle tertiary volcanism during ridge-trench interactions in Western California. Science 258(5083): 793–796.
DeLong SE, Schwarz WM, and Anderson RN (1979) Thermal effects of ridge subduction. Earth and Planetary Science Letters 44(2): 239–246.
Dickinson WR and Snyder WS (1979) Geometry of subducted slabs related to San Andreas transform. Journal of Geology 87(6): 609–627.
Ridge Subduction and Slab Windows 11

Fletcher JM, Grove M, Kimbrough D, Lovera O, and Gehrels GE (2007) Ridge-trench interactions and the Neogene tectonic evolution of the Magdalena shelf and southern Gulf of
California: Insights from detrital zircon U-Pb ages from the Magdalena fan and adjacent areas. Geological Society of America Bulletin 119(11 − 12): 1313–1336.
Forsythe R and Nelson E (1985) Geological manifestations of ridge collision: Evidence from the Golfo de Penas-Taitao Basin, southern Chile. Tectonics 4(5): 477–495.
Gorring M, Singer B, Gowers J, and Kay SM (2003) Plio–Pleistocene basalts from the Meseta del Lago Buenos Aires, Argentina: Evidence for asthenosphere–lithosphere interactions
during slab window magmatism. Chemical Geology 193(3–4): 215–235.
Groome WG and Thorkelson DJ (2009) The three-dimensional thermo-mechanical signature of ridge subduction and slab window migration. Tectonophysics 464(1–4): 70–83.
Haeussler PJ, Bradley DC, Wells RE, and Miller ML (2003) Life and death of the resurrection plate: Evidence for its existence and subduction in the northeastern Pacific in
Paleocene–Eocene time. Geological Society of America Bulletin 115(7): 867–880.
Hole MJ and Larter RD (1993) Trench-proximal volcanism following ridge crest-trench collision along the Antarctic Peninsula. Tectonics 12: 897–910.
Johnston ST and Thorkelson DJ (1997) Cocos-Nazca slab window beneath Central America. Earth and Planetary Science Letters 146(3): 465–474.
Kay SM, Ramos VA, and Marquez M (1993) Evidence in Cerro Pampa volcanic rocks for slab-melting prior to ridge-trench collision in southern South America. Journal of Geology
101(6): 703–714.
Kinoshita O (1999) A migration model of magmatism explaining a ridge subduction, and its details on a statistical analysis of the granite ages in Cretaceous Southwest Japan. Island
Arc 8(2): 181–189.
Klein EM and Karsten JL (1995) Ocean-ridge basalts with convergent-margin geochemical affinities from the Chile Ridge. Nature 374(6517): 52–57.
Kusky TM and Young CP (1999) Emplacement of the resurrection Peninsula ophiolite in the southern Alaska forearc during a ridge-trench encounter. Journal of Geophysical Research -
Solid Earth 104(B12): 29025–29054.
Lonsdale P (1991) Structural patterns of the pacific floor offshore of Peninsular California. In: The Gulf and Peninsular Province of the Californias, 87–125. American Association of
Petroleum Geologists Memoir 47.
Madsen JK, Thorkelson DJ, Friedman RM, and Marshall DD (2006) Cenozoic to recent plate configurations in the Pacific Basin: Ridge subduction and slab window magmatism in
western North America. Geosphere 2(1): 11–34.
Mann P, Taylor FW, Lagoe MB, Quarles A, and Burr G (1998) Accelerating late quaternary uplift of the New Georgia Island Group (Solomon island arc) in response to subduction of the
recently active woodlark spreading center and Coleman seamount. Tectonophysics 295(3–4): 259–306.
Marshak RS and Karig DE (1977) Triple junctions as a cause for anomalously near-trench igneous activity between the trench and volcanic arc. Geology 5(4): 233–236.
McGirr R, Seton M, and Williams S (2020) Kinematic and geodynamic evolution of the Panama Isthmus region: Implications for Central American Seaway closure. Geological Society of
America Bulletin. https://doi.org/10.1130/B35595.1.
McKenzie DP and Morgan WJ (1969) Evolution of triple junctions. Nature 224(5215): 125–133.
Mercier J-P, Bostock MG, Cassidy JF, Dueker K, Gaherty JB, Garnero EJ, Revenaugh J, and Zandt G (2009) Body-wave tomography of western Canada. Tectonophysics 475(3–4):
480–492.
Polat A, Frei R, Longstaffe FJ, Thorkelson DJ, and Friedman E (2018) Petrology and geochemistry of the Tasse mantle xenoliths of the Canadian Cordillera: A record of Archean to
quaternary mantle growth, metasomatism, removal, and melting. Tectonophysics 737: 1–26.
Schuth S, Rohrbach A, Munker C, Ballhaus C, Garbe-Schonberg D, and Qopoto C (2004) Geochemical constraints on the petrogenesis of arc picrites and basalts, New Georgia Group,
Solomon Islands. Contributions to Mineralogy and Petrology 148(3): 288–304.
Severinghaus J and Atwater T (1990) Cenozoic geometry and thermal state of the subducting slabs beneath western North America. In: Basin and Range Extensional Tectonics Near
the Latitude of Las Vegas, Nevada, 1–22.. Geological Society of America Memoir 176.
Sisson V, Pavlis T, Roeske S, and Thorkelson D (2003) Introduction: An overview of ridge-trench interactions in modern and ancient settings. Geology of a Transpressional Orogen
Developed During Ridge-Trench Interaction Along the North Pacific Margin. vol. 371, pp. 1–18. Geological Society of America.
Spencer JE, Richard SM, Reynolds SJ, Miller RJ, Shafiqullah M, Gilbert WG, and Grubensky MJ (1995) Spatial and temporal relationships between mid-tertiary magmatism and
extension in southwestern Arizona. Journal of Geophysical Research - Solid Earth 100(B6): 10321–10351.
Stern CR and Kilian R (1996) Role of the subducted slab, mantle wedge and continental crust in the generation of adakites from the Andean Austral Volcanic Zone. Contributions to
Mineralogy and Petrology 123(3): 263–281.
Thorkelson DJ (1996) Subduction of diverging plates and the principles of slab window formation. Tectonophysics 255(1–2): 47–63.
Thorkelson DJ and Breitsprecher K (2005) Partial melting of slab window margins: Genesis of adakitic and non-adakitic magmas. Lithos 79(1–2): 25–41.
Thorkelson DJ and Taylor RP (1989) Cordilleran slab windows. Geology 17(9): 833–836.
Thorkelson DJ, Madsen JK, and Sluggett CL (2011) Mantle flow through the Northern Cordilleran slab window revealed by volcanic geochemistry. Geology 39(3): 267–270.
Uyeda S and Miyashiro A (1974) Plate tectonics and the Japanese Islands: A synthesis. Geological Society of America Bulletin 85(7): 1159–1170.
Wilson DS (1988) Tectonic history of the Juan de Fuca Ridge over the last 40 million years. Journal of Geophysical Research - Solid Earth 93(B10): 11863–11876.
Wilson DS, McCrory PA, and Stanley RG (2005) Implications of volcanism in coastal California for the Neogene deformation history of western North America. Tectonics 24(3).
Wu TJ and 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.

Further Reading
Dickinson WR and Snyder WS (1979) Geometry of triple junctions related to San Andreas transform. Journal of Geophysical Research - Solid Earth 84(B2): 561–572.
Edwards BR and Russell JK (2000) Distribution, nature, and origin of Neogene–Quaternary magmatism in the northern Cordilleran volcanic province, Canada. Geological Society of
America Bulletin 112(8): 1280–1295.
Eyuboglu Y, Dudas FO, Thorkelson D, Santosh M, Zhu Di-C, Liu Z, Chattergee N, Yi K, and Santosh M (2017) Eocene granitoids of northern Turkey: Polybaric magmatism in an evolving
arc-slab window system. Gondwana Research 50: 311–345.
Furlong KP and Schwartz SY (2004) Influence of the Mendocino triple junction on the tectonics of coastal California. Annual Review of Earth and Planetary Sciences 32: 403–433.
Guillaume B, Martinod J, Husson L, Roddaz M, and Riquelme R (2009) Neogene uplift of central eastern Patagonia: Dynamic response to active spreading ridge subduction?
Tectonics 28(2).
Hole MJ, Rogers G, Saunders AD, and Storey M (1991) Relation between alkalic volcanism and slab-window formation. Geology 19(6): 657–660.
Russo RM, VanDecar JC, Mocanu VI, Gallego A, and Murdie RE (2010) Subduction of the Chile Ridge: Upper mantle structure and flow. GSA Today 20(9): 4–10.
Sisson VB, Roeske SM, and Pavlis TL (eds.) (2003) Geology of a Transpressional Orogen Developed During Ridge-Trench Interaction Along the North Pacific Margin. vol. 371.
Geological Society of America.

You might also like