You are on page 1of 27

RESEARCH ARTICLE Sedimentary Record of Arc‐Continent Collision Along

10.1029/2019TC005741
Mesozoic SW North America (Siuna Belt, Nicaragua)
Special Section: Goran Andjić1 , Javier Escuder‐Viruete2 , Claudia Baumgartner‐Mora3,
Collisional orogenic systems as
recorders of collisions between Peter O. Baumgartner3 , Simon F. Mitchell4 , Michèle Caron5, and Esmeralda Caus6
arc and continents 1
School of Earth and Environmental Sciences, The University of Queensland, St Lucia, Queensland, Australia, 2Instituto
Geológico y Minero de España, Madrid, Spain, 3Institut des Sciences de la Terre, Université de Lausanne, Lausanne,
Key Points: Switzerland, 4Department of Geography and Geology, The University of the West Indies, Kingston, Jamaica, 5Department
• The Siuna Belt of northern of Geosciences, Université de Fribourg, Fribourg, Switzerland, 6Department of Geology, Autonomous University of
Nicaragua exposes the suture zone
between the Siuna Intraoceanic Arc
Barcelona, Bellaterra, Spain
and the continental Chortis Block
• Rocks within the suture zone
suggest that the Siuna Island Arc Abstract The western margin of the Caribbean Plate is a typical example where oceanic and continental
represents an exotic arc that terranes have amalgamated by subduction, collision, and strike‐slip processes. The boundaries between
accreted to the Chortis Block
approximately 134‐131 Ma
these blocks, as well as their tectonostratigraphic records, are generally covered by younger deposits and
• Postcollisional extension led to the dense tropical vegetation, which may hamper reconstructing the accretionary evolution of the convergent
rifting of the Chortis Block margin. In that context, the study of overlap sedimentary assemblages represents an important tool to
constrain the accretion timing of terranes. In northern Central America, the geology of the suture zone
between the Chortis Block and the exotic Siuna Intraoceanic Arc indicates that the two terranes were
Correspondence to: assembled together during a Hauterivian arc‐continent collision (approximately 134–131 Ma). The exotic
G. Andjić, origin of the intraoceanic arc is based on the nature of metamorphic blocks within and the kinematics of the
g.andjic@uq.edu.au
Siuna Serpentinite Mélange. The short duration of the collision event is suggested by coeval exhumation
of the Siuna Serpentinite Mélange and Chortis‐derived coarse sedimentation (El Amparo Formation) along
Citation: the suture zone, rapidly followed by onset of pelagic sedimentation (Rio Matis Formation). Although the
Andjić, G., Escuder‐Viruete, J.,
Baumgartner‐Mora, C., Baumgartner, collision appears to have been short lived and preserved only in the suture zone, postcollisional extension
P. O., Mitchell, S. F., Caron, M., & Caus, affected intra‐arc to back‐arc settings of the Chortis Block and led to the formation of kilometer‐thick
E. (2019). Sedimentary record of extensional basins. We envisage that the convergent margin inboard of SW Mexico—represented by the
arc‐continent collision along Mesozoic
SW North America (Siuna Belt, fringing Guerrero Intraoceanic Arc and the Mixteca continental terrane—underwent similar postcollisional
Nicaragua). Tectonics, 38, 4399–4425. extension, whereas the western margin of the proto‐Caribbean oceanic realm experienced onset of WSW
https://doi.org/10.1029/2019TC005741 dipping subduction beneath the accreted Siuna Intraoceanic Arc.

Received 2 JUL 2019


Accepted 5 OCT 2019
Accepted article online 24 OCT 2019
Published online 21 DEC 2019
1. Introduction
Suture zones between intraoceanic arcs and continents are the inevitable result of plate tectonics and the
witness of fundamental processes such as the growth/destruction of Earth's crust, exhumation of deeply sub-
ducted oceanic and continental material, and Earth's surface vertical motions. Recognition and dating of
ancient arc‐arc or arc‐continent collision zones is a fundamental prerequisite in order to reconstruct
Earth's crust history (e.g., Cawood et al., 2009). Disparate, scattered remnants of accreted intraoceanic arcs
seem to be the rule as processes that can remove them from Earth's surface (subduction and burial) appear to
be more efficient that those favoring their preservation (accretion; e.g., Stern & Scholl, 2010). Consequently,
this complicates the recognition of once continuous, plate boundary‐wide intraoceanic arc systems (Hébert
et al., 2012). Moreover, arc collision events are accompanied by processes such as volcanic arc cessation,
mélange formation/exhumation, shortening, and rock/surface uplift (Draut et al., 2008). The identification
of past arc collision events may be ambiguous or overlooked in the case when the above processes cannot be
tied to any exposed terrane, that is, the response of the passive or convergent margin to arc collision is pre-
served but not the colliding arc itself (Hall & Wilson, 2000).
Several processes decrease the preservation potential of intraoceanic arcs at different stages of their evolu-
tion. (1) Nonsteady subduction erosion can affect intraoceanic arcs in two ways: (a) tectonic erosion during
the lifespan of an intraoceanic arc, with the removal of tens of kilometers of forearc crust (e.g., Azuero/Golfito
©2019. American Geophysical Union.
Arc in Costa Rica, Buchs et al., 2011; Talkeetna Arc in Alaska, Clift et al., 2005; European Variscides, Oncken,
All Rights Reserved. 1998) and (b) tectonic erosion of an intraoceanic arc once it is accreted to a convergent margin (Clift et al.,

ANDJIĆ ET AL. 4399


Tectonics 10.1029/2019TC005741

2009). (2) Intraoceanic arcs may be underplated or subducted during arc‐arc collision events (Hall & Wilson,
2000; Tetreault & Buiter, 2012, 2014; Yamamoto et al., 2009). (3) Chemical and mechanical weathering of
igneous rocks, especially in tropical climates, may impede the study of intraoceanic arc records (Hall &
Smyth, 2008; Hastie et al., 2007). (4) In long‐lived convergent margins, thick volcanic piles may prevent
the direct observation of older autochthonous or allochthonous volcanic formations and ophiolitic remnants
(Andjić et al., 2018, 2019).
Among the key markers of arc‐continent collision zones, exhumed high‐pressure serpentinite‐matrix
mélanges provide precious information on subduction initiation/duration/termination as well as the nature
and age of plates involved during convergence episodes (Agard et al., 2009; Escuder‐Viruete et al., 2013;
Harlow et al., 2004; Krebs et al., 2008; Shervais et al., 2011). Overlap or molasse sedimentary rocks overlying
exhumed ophiolitic mélanges provide equally important information on the timing of tectonic events and
the nature of the terranes involved in arc collision zones (Nokleberg et al., 2000). The combined study of
exhumed ophiolitic mélanges and overlap assemblages is especially relevant in the cases where the colliding
objects have subsequently been removed from Earth's surface by tectonic/physical processes or laterally dis-
placed many tens or hundreds of kilometers (Aitchison et al., 2011; Hildebrand, 2013; Johnston, 2001, 2008).
The Siuna Belt of inland Nicaragua is one such place where the lack of exposure and detailed knowledge of
Cretaceous arc‐related units represents a major gap in the reconstruction of the Circum‐Caribbean terranes
in general, and the enigmatic history of arc‐continent collision in Central America in particular. In this
paper we present new sedimentary constraints on the early evolution of the only known exposure of the
suture zone that separates two of the largest terranes in Central America, that is, the Siuna Intraoceanic
Arc and the continental Chortis Block. This paper complements our recent study of the evolution of the
Siuna Serpentinite Mélange (SSM; Escuder‐Viruete et al., 2019) and provides new constraints on the exhu-
mation timing of the mélange along the suture zone between the two blocks, as well as the postcollisional
evolution of the convergent margin. We also propose a model for the Early Cretaceous evolution of the mar-
gins of SW North America and the western proto‐Caribbean, some key steps of which may be explained by
the collision of the Siuna Intraoceanic Arc.

2. Northern Central America: A Collage of Continental and Oceanic Terranes


Representing the western margin of the Caribbean Plate, the isthmus of Central America consists of a com-
plex geological collage of predominantly Mesozoic‐Cenozoic oceanic assemblages of suprasubduction and
intraplate affinities with subordinate continental terranes restricted to its northern part (Rogers et al.,
2007; Baumgartner et al., 2008; Buchs et al., 2011; Figure 1). The geology of northern Central America is char-
acterized by a core of continental crust, the Chortis Block, with which oceanic terranes were tectonically jux-
taposed by collision and strike‐slip processes during the Cretaceous‐Cenozoic (Mann et al., 2007; Pindell
et al., 2012). Based on a multidisciplinary data set, Rogers, Mann, and Emmet (2007) established that the
Chortis Block comprised two distinct terranes; the Central Chortis Terrane consists of greenschist‐ to
amphibolite‐grade continental rocks of Precambrian‐Paleozoic age, whereas the Eastern Chortis Terrane dis-
plays Jurassic nonmetamorphic to greenschist‐grade (meta)sedimentary rocks of continental provenance
(Agua Fria Formation; Figure 2). The Chortis terranes are covered by a Cretaceous sedimentary‐volcanic
overlap succession formed in extensional basins (Gordon, 1990, 1993; Rogers et al., 2007; Rogers et al.,
2007). Prior to the Cenozoic, the Chortis Block was possibly located in the prolongation of Mixteca/Xolapa
terranes of SW Mexico, as suggested by a wide range of geological similarities and the paleomagnetic history
of the Chortis Block (Rogers, Mann, & Emmet, 2007; Sierra‐Rojas et al., 2016; Molina Garza et al., 2017;
Figure 2b).
To the NW, the Central Chortis Terrane is bordered by the Guatemala Suture Zone, which marks the trans-
form boundary between the Caribbean Plate and the North American Plate (Donnelly et al., 1990; Flores
et al., 2015; Figure 1). The Guatemala Suture Zone displays several belts of oceanic and continental high‐
pressure rocks that were emplaced during at least two Cretaceous collision events and later juxtaposed by
Cenozoic strike‐slip motion along the Motagua‐Cayman fault system (Flores et al., 2013; Harlow et al.,
2004; Pindell et al., 2012; Ratschbacher et al., 2009; Torres‐de León et al., 2012). To the southwest, the
Chortis continental terranes are possibly bordered by accreted oceanic terranes, as suggested by the

ANDJIĆ ET AL. 4400


Tectonics 10.1029/2019TC005741

Figure 1. Plate tectonic map displaying the main terranes of Middle America (modified from Sedlock et al., 1993; Mann et al., 2007; Rogers et al., 2007a;
Baumgartner et al., 2008; Centeno‐García et al., 2008; Ratschbacher et al., 2009; Flores et al., 2013, 2015; Padilla y Sánchez et al., 2013; Carvajal‐Arenas &
Mann, 2018; Andjić et al., 2018, 2019; Sanchez et al., 2019). White arrows show the azimuth and rate (mm/year) of oceanic crust convergence relative to North
America (after Ferrari et al., 2012) and the Caribbean Plate (after Kobayashi et al., 2014). Triangles represent Quaternary volcanoes (after Rogers et al., 2007a). The
black rectangle represents the location of the geological map in Figure 2a. MAT = Middle America Trench.

geochemistry of Quaternary volcanoes and gravity anomalies (Carr et al., 2003; Rogers, Mann, & Emmet,
2007; Figures 1 and 2).
To the southeast, the Eastern Chortis Terrane is bordered by an accreted intraoceanic arc assemblage of
Jurassic‐Early Cretaceous age, called the Siuna Terrane (Rogers, Mann, & Emmet, 2007; Rogers, Mann,
Emmet, & Venable, 2007; Venable, 1994) or the Mesquito Composite Oceanic Terrane (Baumgartner et al.,
2008; Flores et al., 2015; Figures 1 and 2). Besides evidence built on field, geological, petrological, and geo-
chemical data (Baumgartner et al., 2008; Escuder‐Viruete et al., 2019; Flores et al., 2015), the distinction
between the Chortis continental terrane and the Siuna Intraoceanic Arc terrane is further based on their con-
spicuously different Pb isotope and gravity signatures (Rogers, Mann, & Emmet, 2007; Sundblad et al., 1991;
Venable, 1994). The recognition of the exotic Siuna Intraoceanic Arc relies on the geology of the Siuna Belt,
which was exposed due to Cenozoic strike‐slip fault activity (Escuder‐Viruete et al., 2019; Flores et al., 2015).
The bulk of the Siuna Intraoceanic Arc is covered by younger sedimentary and volcanic deposits elsewhere
(Baumgartner et al., 2008), including offshore eastern Nicaragua (Carvajal‐Arenas & Mann, 2018; Sanchez
et al., 2019). Based on geological and age similarities, Baumgartner et al. (2008) included the Siuna
Intraoceanic Arc in the Mesquito Composite Oceanic Terrane together with other oceanic assemblages, that
is, the El Castillo Mélange (S Nicaragua), the Santa Elena Intraoceanic Arc (N Costa Rica), and rocks of DSDP
Leg 67/84 (Guatemala forearc; Figure 1). The latter three exposures may be related to a distinct intraoceanic
arc that accreted during the mid‐Cretaceous, as suggested by the evolution of the Santa Elena Intraoceanic
Arc and its similarities with that of the rocks of the DSDP Leg 67/84 (Escuder‐Viruete et al., 2015;
Escuder‐Viruete & Baumgartner, 2014; Geldmacher et al., 2008). In contrast, any bathymetric feature that

ANDJIĆ ET AL. 4401


Tectonics 10.1029/2019TC005741

Figure 2. Geology of the eastern Chortis Block, northern Central America. (a) Geological map of eastern Honduras and northern Nicaragua (modified from Rogers
et al., 2007a). The Chortis Block comprises two terranes, the Central Chortis Terrane (CCT) and the Eastern Chortis Terrane (ECT). The suture between
the ECT and the Siuna Intraoceanic Arc (SIA) is exposed in the Siuna Belt (northern Nicaragua). The red rectangle represents the area covered by the geological
map in Figure 3b. (b) Chronostratigraphic chart of the Siuna Belt (column 1, modified from: Venable, 1994; Baumgartner et al., 2008; Flores, 2009; Flores et al.,
2015; Escuder‐Viruete et al., 2019) and the Chortis Block (columns 2 and 3, modified from: Gordon, 1990; Scott & Finch, 1999; Rogers, Mann, & Emmet, 2007,
Rogers, Mann, Emmet, & Venable, 2007b, Rogers, Mann, Scott, & Patino, 2007c). Note the chronostratigraphy of the Mixteca Terrane (SW Mexico) for comparison
(modified from Sierra‐Rojas et al., 2016; Molina Garza et al., 2017). The timescale is after Ogg et al. (2016). EAF = El Amparo Formation.

ANDJIĆ ET AL. 4402


Tectonics 10.1029/2019TC005741

may have been accreted to the SW margin of Mexico—represented by Guerrero and Xolapa terranes—is
either located offshore or was removed by tectonic erosion. The latter process may explain the close proximity
of these Mesozoic autochthonous arc terranes to the trench (approximately 60 km; Figure 1). Clift and
Vannucchi (2004) indicated an average rate of tectonic erosion of 1 km/m.y. since 23 Ma for SW Mexico,
which is high enough to have caused removal of the remnants of an accreted intraoceanic arc of average
thickness (approximately 26 km; Tetreault & Buiter, 2014) during the Neogene.
Late Cretaceous shortening affected both the Central and Eastern Chortis terranes. NE‐SW oriented short-
ening (present‐day coordinates) in the Central Chortis Terrane occurred during a regional Laramide age
deformation, the origin of which remains poorly constrained (Burkart et al., 1987; Donnelly et al., 1990;
Rogers, Mann, & Emmet, 2007, Rogers, Mann, Emmet, & Venable, 2007, Rogers, Mann, Scott, & Patino,
2007; Fitz‐Díaz et al., 2018; Figure 2a). In contrast, NW‐SE oriented shortening (present‐day coordinates)
in the Eastern Chortis Terrane (Colon Fold Belt; Rogers, Mann, Emmet, & Venable, 2007; Figure 2a) is con-
sidered as the result of the diachronous insertion of the Caribbean Large Igneous Province between the
Americas, which started during the Coniacian (approximately 89 Ma; Andjić et al., 2019).

3. Geology of the Siuna Belt: Insight Into an Arc‐Continent Suture Zone


The Siuna Belt consists of a serpentinite‐matrix mélange, the SSM, that is unconformably or tectonically
overlain by a Hauterivian to Albian sedimentary‐volcanic succession (Venable, 1994; Baumgartner et al.,
2008; Flores et al., 2015; Escuder‐Viruete et al., 2019; Figures 3 and 4). The belt is bounded by thrust faults
on its western and eastern sides and flanked by Late Cretaceous to Eocene volcanic and volcaniclastic rocks.
Dikes dated at 75 Ma and plutons dated at 60 Ma intrude the Siuna Belt and the adjacent volcano‐
sedimentary assemblage (Venable, 1994; Figure 3b).

3.1. The SSM


The SSM represents an approximately 2.5‐km‐thick tectonic unit characterized by a block‐in‐matrix fabric
consisting of decimeter‐ to hectometer‐sized metamorphic blocks embedded in a foliated serpentinite
(Baumgartner et al., 2008; Flores, 2009; Flores et al., 2015; Escuder‐Viruete et al., 2019; Figures 3 and 4).
The serpentinite‐matrix is interpreted as the result of hydration, ductile deformation, and metamorphism
of mantle peridotites in a subduction channel (Escuder‐Viruete et al., 2019; Flores et al., 2015). Blocks of
red ribbon radiolarite of Middle Jurassic age suggest material incorporation from a subducting oceanic plate
(Baumgartner et al., 2008). Moreover, ultramafic blocks within the mélange, including peridotites, present
abyssal, and suprasubduction zone affinities. A group of ultramafic blocks (i.e., harzburgites) with
suprasubduction signatures represents the complementary residue of melts that produced the protoliths of
metaigneous mafic blocks displaying Island Arc Tholeiitic (IAT) and boninitic affinities (Escuder‐Viruete
et al., 2019). Ultramafic and mafic blocks reached metamorphic conditions ranging from greenschist and
blueschist facies to amphibolite and eclogites facies (Escuder‐Viruete et al., 2019; Flores et al., 2015). An
arc‐related amphibolite block preserved a mineral assemblage that recorded near‐metamorphic peak
conditions at 148.9 ± 2.9 Ma (40Ar‐39Ar on amphibole, Tithonian; Escuder‐Viruete et al., 2019; Figure 3a).
Overall, the exotic blocks within the SSM suggest that both the upper plate (intraoceanic arc) and the lower
plate (oceanic crust) contributed to the development of the serpentinite‐matrix mélange at high‐pressure
metamorphic conditions. Exhumation of blocks appears to have started in the early Valanginian, as
suggested by 40Ar‐39Ar dating on phengite crystals recovered from a mica‐rich rock (139.2 ± 0.4 Ma;
Flores et al., 2015). Final exhumation of the SSM occurred during the Hauterivian, as deduced from the
overlapping ages between the youngest cooling event recorded in blocks within the mélange (40Ar‐39Ar,
130.6 ± 1.1 Ma on amphibole, 132.8 ± 0.5 Ma on phengite; Escuder‐Viruete et al., 2019) and the base of
the overlying sedimentary cover (see below and Figure 3a).

3.2. The Sedimentary‐Volcanic Overlap Assemblage


The SSM is unconformably or tectonically overlain by a sedimentary‐volcanic overlap assemblage domi-
nated by thinly bedded pelagic‐hemipelagic limestone and limey siltstone interbedded with volcaniclastic
turbidites (Baumgartner et al., 2008; Flores et al., 2015; Venable, 1994); the composite thickness of the over-
lap assemblage is estimated to be at least 1,500 m. This carbonate formation here is called the Rio Matis
Formation, to which we attribute a late Hauterivian to Albian age (Figure 3a). Venable (1994) indicated

ANDJIĆ ET AL. 4403


Tectonics 10.1029/2019TC005741

Figure 3. Geology of the Siuna area, northern Nicaragua. (a) Chronostratigraphic chart of the Siuna Belt (modified from Venable, 1994; Baumgartner et al., 2008;
Flores, 2009; Flores et al., 2015; Escuder‐Viruete et al., 2019). An overlap assemblage composed of two distinct formations—the El Amparo and Rio Matis
Formations—overlies the Siuna Serpentinite Mélange. The timescale is from Ogg et al. (2016). (b) Geological map of the Siuna Belt displaying major rock units,
thrusts, and faults. Note the position of cross sections depicted in Figures 4a, 4b, and 4c.

ANDJIĆ ET AL. 4404


Tectonics 10.1029/2019TC005741

Figure 4. Structure of the Siuna Belt along key transects (modified after Escuder‐Viruete et al., 2019). (a)–(c) Cross sections through the Siuna Belt showing its
general architecture (no vertical exaggeration). See Figure 3b for the position of the three transects (I‐I′, II‐II′, and III‐III′). The color code is the same as in
Figure 3b, except for the blocks in the Siuna Serpentinite Mélange. Unit abbreviations: SSM = Siuna Serpentinite Mélange; LKea = El Amparo Formation;
LKrm = Rio Matis Formation; LKv = Lower Cretaceous volcanics; UKv = Upper Cretaceous volcanics; Pv = Paleogene volcanics; Pp = Paleogene plutons.
(d, e) Stereographic projections of the poles to mean foliation (Sp). (f, g) Stereographic projections of the mean lineation (Lp). Lineation was measured on oriented
antigorite fibers, chlorite flakes, and aligned magnetite grains. (d, f) Equal angle, lower hemisphere projections. (e, g) Equal area, lower hemisphere projections.
See Escuder‐Viruete et al. (2019) for the detailed kinematic interpretation of Sp and Lp.

that andesitic volcanic flows, pyroclastic rocks, and tuffs are intercalated with the Rio Matis Formation. In
places, carbonate material of shallow water origin is interstratified with the thinly bedded deep water
lithologies and generally appear as aligned metric boulders and large elongated bodies up to 0.1 km3 in
volume (Venable, 1994; Flores et al., 2015; Figure 3b). Two types of conglomerate were recognized in the
overlap succession by Venable (1994), a volcanic‐rich type and a volcanic‐poor type. Here we consider the
volcanic‐poor conglomerate as a separate formation—the El Amparo Formation—underlying the Rio
Matis Formation (see below).

3.3. Structure of the Siuna Belt


The Siuna Belt was deformed by brittle thrust, strike‐slip, and normal faults (Venable, 1994; Escuder‐Viruete
et al., 2019; Figures 3b and 4). N‐S to NE‐SW trending, middle‐angle E and W dipping thrusts bound the
Siuna Belt to the west and the east (Figure 3b), as well as separate the SSM from the overlying overlap
sequence in places (e.g., the Siuna mine). A discrete phase of NE‐SW to ENE‐WSW trending, subvertical nor-
mal, and strike‐slip faulting affected the whole belt (kilometer‐spaced faults in Figures 3a and 4c). Although
deformation clearly postdates the Early Cretaceous deposition of the overlap succession, its precise timing
remains unknown due to poor outcrop conditions. Venable (1994) argued that deformation took place prior
to the emplacement of late Campanian dikes. Similarly, Escuder‐Viruete et al. (2019) indicated that folds and
thrusts did not affect the Cenozoic volcanic pile, whereas strike‐slip faults were reactivated during the
Neogene‐Quaternary. Flores et al. (2015) considered that the Siuna Belt owed its present configuration to

ANDJIĆ ET AL. 4405


Tectonics 10.1029/2019TC005741

Figure 5. Three‐step sketch of the evolution of the convergent margin of SW North America inboard of the Chortis Block and its response to the collision of the
Siuna Intraoceanic Arc. The red arrow indicates mélange exhumation. The drawing style is after Stampfli et al. (2013). See text for discussion.

a post‐Eocene positive flower structure related to a fault system parallel to the one of the Guayape Fault (e.g.,
Gordon & Muehlberger, 1994; Figure 2a).

4. Stratigraphy of the Overlap Succession


4.1. El Amparo Formation (Basal Conglomerate, Hauterivian)
The El Amparo Formation is the oldest lithology of the overlap succession (Figures 2b and 3a). It consists of
granule to pebble, clast‐supported polymict conglomerate beds, with a total thickness not exceeding 20 m.
The best exposure is located 800 m northwest of the Cerro Waylawas, where a 15‐m‐thick sequence of con-
glomerates occurs; a section of similar thickness is exposed at the northern tip of the Cerro Las Delicias
(Figure 3b). The finest‐grained levels of the formation are locally low‐angle cross‐bedded granule conglom-
erates. In the uppermost part of the formation, decimeter‐bedded conglomerate alternates with decimeter‐
bedded siltstone. The average proportions of clasts within the conglomerate are the following

ANDJIĆ ET AL. 4406


Tectonics 10.1029/2019TC005741

Figure 6. Four‐step scenario of the geodynamic evolution of the convergent margins of SW North America during the latest Middle Jurassic to late Early
Cretaceous (modified after Sedlock et al., 1993; Flores, 2009; Bandini et al., 2011; Pindell et al., 2012; Flores et al., 2013, 2015; Madrigal et al., 2016; Andjić et al.,
2018; Martini & Ortega‐Gutiérrez, 2018). Note the position of the Siuna Intraoceanic Arc (SIA) through time. See Figure 7 for transects inboard of Guerrero
(dashed lines number 1), Chortis (dashed lines number 2), and W proto‐Caribbean (dashed lines number 3) margins. Abbreviations: ACG = Arperos‐Cuicateco‐
Guatemala back‐arc basin; CA = Caribbean Arc; FAR = Farallon Plate; IATF = Inter‐American Transform Fault; MEZ = Mezcalera Plate; MT = Manzanillo
Terrane; NC = Nicoya Complex; SEA = Santa Elena Intraoceanic Arc; WRA = Wrangellia. See text for discussion.

(Figure A1h): fine‐ to medium‐grained quartz wacke (30%), quartz siltstone (30%), polycrystalline quartz
(30%), microcrystalline quartz (or chert, 5%), and tuff/radiolarian‐bearing tuffaceous siltstone (5%). Lithic
and quartz clasts are generally subrounded to rounded. Scarce hemipelagic matrix within the
conglomerate contains radiolarians, indicating an offshore setting of deposition.
Quartz siltstone and quartz wacke occasionally show occurrence of incipient to well‐developed cleavage.
In some cases, quartz wacke clasts display a foliation formed by dissolution during pressure processes in
very low to low degree metamorphic conditions, which is suggested by chlorite neoblastesis. Conversely,
tuffaceous clasts do not show deformation features. In most cases, polycrystalline quartz clasts correspond
to fragments of quartzite (Figures A1i and A1j). Clasts of quartzite generally present a low‐grade fabric
defined by elongated quartz grains and rare subparallel white micas, including microtextures of both
grain boundary migration and rotation of subgrains that were formed by intracrystalline plastic deforma-
tion, possibly in low‐T greenschist facies conditions, and recrystallized quartz ribbons, which are micro-
textures typical of mylonitic quartzite that recrystallized in high‐T greenschist facies to low‐T
amphibolite facies conditions.

ANDJIĆ ET AL. 4407


Tectonics 10.1029/2019TC005741

Figure 7. Four‐step sketches of the geodynamic evolution of the convergent margins of SW North America during the latest Middle Jurassic to late Early
Cretaceous (modified after Sedlock et al., 1993; Flores, 2009; Bandini et al., 2011; Pindell et al., 2012; Flores et al., 2013, 2015; Escuder‐Viruete & Baumgartner,
2014; Martini et al., 2014; Escuder‐Viruete & Castillo‐Carrión, 2016; Madrigal et al., 2016; Andjić et al., 2018; Martini & Ortega‐Gutiérrez, 2018). The drawing
style is after Stampfli et al. (2013). Note the position of the Siuna Intraoceanic Arc through time. See Figure 6 for the location of the Guerrero Terrane, the Chortis
Block, and the W proto‐Caribbean realm, as well as corresponding transects (numbers 1 to 3). The ARPEROS basin corresponds to the Arperos‐Cuicateco‐
Guatemala (ACG) back‐arc basin in Figure 6. Note also their present‐day position in Figure 1. See text for discussion.

4.2. Rio Matis Formation (Late Hauterivian to Late Albian)


The bulk of the Rio Matis Formation comprises dark gray, thinly bedded, pelagic to silty hemipelagic lime-
stone (Figure A1a) commonly with planktonic foraminifera and radiolarians. The formation is locally up to
1,500 m thick (Figure 4). Four types of lithologies are commonly interstratified with the thinly bedded
limestone (Figure 3a).
(1) Volcaniclastic deposits range from siltstone to conglomerate. Centimeter‐thick beds of radiolarian‐
bearing tuffaceous siltstone are present at all the levels of the formation; metric packages of such lithology
crop out along the main road (1 km south of Tadasna; Figure A1c). The coarsest volcaniclastic deposits cor-
respond to decimeter‐thick, fine‐ to coarse‐grained, arkosic‐lithic arenite/wacke, and granule‐to‐pebble
polymict conglomerate beds (Figure A1d). Andesite, fine‐ to coarse‐grained tuff, and tuffaceous siltstone
are the most common rock clasts within these deposits; altered glassy volcanic clasts, and matrix give a dis-
tinctive green color to the volcaniclastic deposits. The most common individual grains correspond to plagi-
oclase, clinopyroxene, amphibole, and opaque minerals. Subordinate rock clasts (<5%) are quartz arenite,
shallow water limestone, and red radiolarian chert (Figure A1g), whereas subordinate minerals (<5%) cor-
respond to quartz (polycrystalline, microcrystalline, and monocrystalline) and serpentine. Serpentine occurs
only as fine to medium sand in few arkosic‐lithic arenite beds (Figures A1e and A1f). Serpentine grains
mainly correspond to massive and foliated antigorite and rare chrysotile‐lizardite, which is in accordance

ANDJIĆ ET AL. 4408


Tectonics 10.1029/2019TC005741

with the predominance of antigorite in the underlying serpentinite‐matrix mélange (Escuder‐Viruete et al.,
2019). The hosting arkosic‐lithic arenite beds are dominated by calcified plagioclase grains, a feature that has
not been observed in lithologies free of serpentinite clasts.
(2) Black rudist‐rich limestone olistoliths occur as decimeter to decameter‐sized boulders. Due to the weak-
ness of the (hemi)pelagic matrix to weathering, olistotromes are only preserved in cases where limestone
boulders are decimetric in size. Centimeter‐sized rudists are commonly found in life position as well as
clasts. Other common constituents in the grainstone/packstone matrix include rounded bioclasts of mol-
luscs, green algae, Solenoporacean red algae (Parachaetetes), echinoids, agglutinated foraminifera, and
miliolid foraminifera.
(3) Along the Siuna Belt, decameter‐sized gray limestone channels coincide with prominent NNE‐SSW
oriented reliefs, such as the Cerro Waylawas, Loma La Calera, and Cerro Las Delicias (Figure 3b). This mas-
sive limestone consists of wackstone/packstone dominated by micritic peloids. Common bioclasts include
Orbitolina, miliolids, corals, molluscs (including rudists), green algae, Solenoporacean red algae
(Parachaetetes), and echinoids. Minor detrital grains include plagioclase and quartz crystals, and clasts of
tuffs/aphanitic volcanic rocks. Centimeter‐sized lenses of black chert commonly occur within the limestone.
The depositional contact with the encompassing pelagic sedimentary rocks has not been observed, probably
due to a combination of thrust faulting, weathering, and differential compaction (Venable, 1994).
Nevertheless, the original channel configuration is deduced from (a) the NNE‐SSW orientation and elon-
gated shape of the limestone hills, (b) the age of the limestone hills (see below) encompassed by that of
the enclosing pelagic sedimentary rocks, and (c) the carbonate material that forms the limestone hills was
also reworked in turbidites that are interbedded with the pelagic sedimentary rocks (see Point 4 below).
(4) Decimeter‐bedded calciturbidites range from coarse‐grained arenite to pebble conglomerate/breccia.
These deposits are composed of clasts of the same material as the decameter‐sized limestone channels
(see Point 3 above).

5. Biostratigraphy of the Rio Matis Formation


5.1. Radiolarian Biochronology
Radiolarians were extracted from two ash‐rich beds of the lower Rio Matis Formation. Sample NI14‐28 was
recovered at the base of the unit, 0.5 m above the boundary with the underlying conglomerate of the El
Amparo Formation (Figures 3, A1, and A2). Sample NI14‐52 was recovered higher in the Rio Matis
Formation, along the main road 1 km SSW of Tadasna (Figures 3 and A2).
Sample NI14‐28 contains a diverse radiolarian assemblage (Figure A2), with the co‐occurrence of the follow-
ing age‐diagnostic species: Cana septemporatus, Obesacapsula verbana, and Squinabollum simplex.
According to Baumgartner et al. (1995), the co‐occurrence of these taxa is characteristic of their UAZone
20, which indicates a late Hauterivian‐earliest Barremian age. The lower boundary of their UAZone 20 coin-
cides with the last appearance of nannofossil Cruciellipsis cuvilleri (base of polarity chron M8), whereas the
upper boundary of the zone coincides with the last appearance of nannofossil Calcicalathina oblongata
(approximately 0.5 Ma above the base of polarity chron M3; Ogg et al., 2016). This interval currently corre-
sponds to a numerical age of approximately 133.2–130.2 Ma (Ogg et al., 2016; Figure 3a).
Sample NI14‐52 is characterized by the co‐occurrence of Pseudodictyomitra carpatica, Pseudodictyomitra
hornatissima, Pseudodictyomitra lodogaensis, and Parvimitrella communis (Figure A2). According to
O'Dogherty (1994), this assemblage indicates a late early Aptian age (Zone UA 5), which currently corre-
sponds to a numerical age of approximately 124–123 Ma (Ogg et al., 2016; Figure 3a).

5.2. Planktonic Foraminifera Biochronology


One sample of pelagic limestone from the upper Rio Matis Formation yielded datable planktonic foramini-
fera that were studied in randomly oriented thin sections (Figure A3). Sample NI14‐32 contains Biticinella
breggiensis, Thalmanninella appenninica, Thalmanninella gandolfii, Praeglobotruncana delrioensis,
Praeglobotruncana stephani, Pseudothalmanninella ticinensis, and Ticinella madecassiana. This assemblage
is characteristic of the lower Thalmanninella appenninica zone (latest Albian; Premoli Silva & Sliter, 2003),
which currently corresponds to a numerical age of approximately 102–101 Ma (Ogg et al., 2016; Figure 3a).

ANDJIĆ ET AL. 4409


Tectonics 10.1029/2019TC005741

5.3. Larger Benthic Foraminifera Biochronology


Larger benthic foraminifera were recovered north of Siuna from (a) calciturbidite beds of the Rio Matis
Formation, along the Rio Matis (Figures 3 and A4) and (b) the massive, bioclastic limestone package that
forms the bulk of the Cerro Las Delicias (Figures 3 and A4). In both localities, the abundant larger forami-
nifera are conical orbitolinids in which we distinguish at least two different species of Orbitolina
(Mesorbitolina). A sample from the Cerro Las Delicias (numbers 2 and 3, and 6 in Figure A4) revealed an
axial section of O. (M.) cf. texana (Roemer), first published by Baumgartner‐Mora et al. (2013),
Baumgartner‐Mora et al., 2013) as O. texana. The measurements of the embryonic apparatus and the embryo
of our specimen plots beyond the area attributed to O. (M.) texana by Douglas (1960). Our values plot also
beyond the largest values of Hofker's (1963) group III (equivalent to the subgenus Mesorbitolina) suggesting
a late Albian minimum age. O. (M) texana appears to be restricted to the late Aptian‐early Albian interval, as
reported form the Glen Formation in Texas, as well as in Guatemala, Venezuela, and Honduras (Schroeder,
1975; Scott & Gonzalez‐Leon, 1991; Vaughan, 1932). Görög and Arnaud‐Vanneau (1996) reported O. (M) tex-
ana with other species, such as O. concava and O. ovalis in Venezuela. In conclusion, our specimen could be
a late O. (M) texana or an undescribed upper Albian new species.
5.4. Rudist Biochronology
We photographed rudist sections in outcrops from limestone olistoliths in the Rio Matis Formation exposed
in the Rio Matis and adjacent hills to the north of Siuna (Figures 3 and A5). Several localities yield typical
Lower Albian rudist assemblages including Coalcomana spp., of which some could be determined as
Coalcomana ramosa (Figure A5). Other forms include Caprinuloidea spp., of which one could be deter-
mined as Caprinuloidea cf. C. multitubifera. Additional forms include Toucasia sp. and a polyconitid. At
the Lomas El Dorado locality (Figure A5), the rudist‐rich olistoliths sit on late Albian pelagic limestone
(sample NI14‐32 above).
5.5. Summary and Stratigraphic Implications of the Biochronologic Ages
Deposition of the Rio Matis Formation spanned the late Hauterivian‐late Albian time interval (approxi-
mately 133–100 Ma). Pelagic sedimentation was continuous throughout that time but was regularly inter-
rupted by the settling of ash and deposition of volcaniclastic gravity flows. Deposition of shallow water
carbonate material occurred at least from the late Aptian‐early Albian and continued until the late Albian.
However, most of the shallow water‐derived boulders and channels (e.g., Cerro Waylawas, Cerro La
Calera, and Lomas El Caracol/Sauceria west of Tadasna) remain undated and may be of older than the late
Aptian. Finally, a Hauterivian age is assigned to the El Amparo Formation, given the late Hauterivian‐
earliest Barremian age of the base of the overlying Rio Matis Formation

6. Discussion
6.1. Provenance of the Overlap Succession
6.1.1. Source of the Detrital Material in the El Amparo and Rio Matis Formations
Besides a change in sedimentation style, a major shift in provenance occurred at the boundary between the
El Amparo and Rio Matis Formations. The Hauterivian El Amparo Formation is dominated by quartzose
material whereas the detrital beds of the late Hauterivian to late Albian Rio Matis Formation are dominated
by volcaniclastic material (Figures 3 and A1).
The conglomerate beds of the El Amparo Formation consist mainly of compositionally and texturally
mature sedimentary clasts, such as quartz siltstone/sandstone occasionally displaying very low to low meta-
morphic degrees and greenschist‐grade quartzite, and minor unmetamorphosed tuff/tuffaceous siltstone
(Figure A1h). The lithologies of the El Amparo Formation indicate that it was sourced from the upper crustal
levels of a continental block, as it chiefly contains clasts of poorly metamorphic quartzose cover strata (e.g.,
Garzanti, 2016). Conversely, the provenance of the scarce tuffaceous clasts of the El Amparo Formation
remains ambiguous, given that such deposits are (a) known from the Jurassic Agua Fria Formation of the
East Chortis Terrane (Gordon, 1990); (b) formed in any deep‐marine setting located relatively close to a con-
vergent margin; and (c) interstratified throughout the overlap succession, including as clasts in quartzose
and volcaniclastic conglomerates. Nonetheless, the most likely source for the bulk of the El Amparo
Formation is the Jurassic Agua Fria Formation, which comprises up to greenschist‐grade tuff, shale,

ANDJIĆ ET AL. 4410


Tectonics 10.1029/2019TC005741

sandstone, quartzite, and quartz‐pebble conglomerate (Gordon, 1993; Figure 2); the metasedimentary rocks
embedded in the Siuna serpentinite‐matrix mélange are not a likely source since they display distinct
mineral assemblages indicative of higher metamorphic grades (Flores et al., 2015). Moreover, the Agua
Fria Formation is the only formation known to have been exposed to erosion east of the Guayape Fault dur-
ing the earliest Cretaceous (Viland et al., 1996; Rogers, Mann, Emmet, & Venable, 2007; Figure 2). Although
being possibly penecontemporaneous and of volcanic‐poor facies, the El Amparo and Tepemechin
Formations are not lateral equivalents: (a) The El Amparo Formation contains much less metamorphic frag-
ments than the Tepemechin Formation (Venable, 1994); (b) the Tepemechin Formation does not outcrop
east of the Guayape Fault (Gordon, 1993; Rogers, Mann, Emmet, & Venable, 2007; Figure 2); and (c) the
El Amparo Formation was deposited in a deep‐marine setting (see section 4.1.), whereas the Tepemechin
Formation was formed in a continental setting (Gordon, 1990).
The volcaniclastic material embedded in the overlying Rio Matis Formation is predominantly composed of
andesitic clasts/minerals and clasts of tuffs/tuffaceous siltstones (Figures 3, A1c, and A1d), which are lithol-
ogies typical of convergent margin settings. Since geochemical analysis has not yet been performed, it is not
currently possible to discriminate between intraoceanic arc and continental arc provenance for the volcani-
clastic material. Andesitic volcanic flows and tuff have been reported from Late Jurassic Agua Fria and
Aptian‐Albian Manto Formations of the Chortis Terranes (Gordon, 1990, 1993; Rogers, Mann, Scott, &
Patino, 2007c; Figure 2), as well as from Lower Cretaceous volcanic units outcropping along the Siuna
Belt (Venable, 1994; Flores, 2009; Flores et al., 2015: Figure 3). Besides possibly containing a mix of
volcaniclastic material from different sources, the volcaniclastic beds of the Rio Matis Formation display det-
rital grains that reveal at least two additional sediment sources. Minor proportions of quartzite and ophiolitic
detritus (i.e., serpentinite and red radiolarian chert) document the contribution of sediments sourced from
continental and oceanic terranes, respectively. The quartzite clasts are similar to those of the El Amparo
Formation and were probably sourced from the same lithologies of the Agua Fria Formation. The latter
was locally exposed to subaerial erosion along the Eastern Chortis Terrane at least until the Barremian‐
Aptian, when it was unconformably overlain by the Atima Formation (Rogers, Mann, Emmet, & Venable,
2007; Scott & Finch, 1999). In contrast, the ophiolitic material was sourced from the SSM (Figures 2 and
3), which is the only known ophiolitic source in the area.
6.1.2. Source of the Shallow Water Carbonate Material in the Rio Matis Formation
Shallow water carbonate gravity flow sedimentation initiated no later than the late Aptian (Figure 3).
Bioclasts and carbonate grains were eroded off rudist reefs and mud‐rich lagoons typical of tropical to
subtropical carbonate platforms. The carbonate material was transported from platform to basin settings
within three types of deposits (see section 4.2.): (a) decameter‐sized channelized debris flows made of micrite
and sand‐sized bioclasts/micritic peloids, (b) olistostromes of decimeter‐ to decameter‐sized limestone
boulders, and (c) decimeter‐bedded calciturbidites. In the Colon Fold Belt (approximately 120 km N of
Siuna; Figure 2a), the Aptian‐Albian shallow‐shelf Atima Formation consists of a shallowing upward
sequence of peloid‐rich rudist‐bearing wackstone/packstone/grainstone beds containing bioclasts such as
Orbitolina, Solenoporacean red algae (Parachaetetes texana), and green algae (Scott & Finch, 1999). The
Aptian‐Albian Atima Formation of the Eastern Chortis Block represents a likely source for the shallow
water carbonate material of the Rio Matis Formation because of (a) the coeval shallow water‐derived
carbonate gravity flow sedimentation in the deep water Rio Matis Formation and shallow water deposition
in the Atima Formation; (b) the similar composition of the shallow water carbonate material in the two
formations; and (c) the NNE‐SSW to NNW‐SSE trends of the largest carbonate bodies in the Rio Matis
Formation (Figure 3b), supporting sediment pathways between the two areas.
6.2. Mode and Timing of the Collision of the Siuna Intraoceanic Arc With the Chortis Block and
Postcollisional Evolution
Based on field, geochemical, geochronological, and microstructural data, we have recently established that
the SSM represents a subduction channel to a Jurassic‐Early Cretaceous intraoceanic arc (Escuder‐Viruete
et al., 2019); enclosed sedimentary and igneous blocks of oceanic plate and intraoceanic arc affinities were
subducted up to eclogite facies conditions prior to exhumation during the Hauterivian. The exhumation
event is interpreted as the result of the collision between the Siuna Intraoceanic Arc and the Chortis
Block, which is supported by (a) the position of the SSM along a sharp NE‐SW oriented boundary separating
rocks of continental and oceanic affinities (Figures 1 and 2); (b) the affinity of the overlap succession with the

ANDJIĆ ET AL. 4411


Tectonics 10.1029/2019TC005741

sedimentary cover of the Chortis Block (see above and Flores et al., 2015; Figure 2b); and (c) quartzites with
continental affinity in the SSM (Flores et al., 2015), suggesting the subduction of Chortis‐derived sedimen-
tary material prior to final arc‐continent collision (Escuder‐Viruete et al., 2019).
Our preferred geodynamic scenario is that of a Molucca Sea‐type collision (e.g., Cardwell et al., 1980; Flores,
2009; Zagorevski et al., 2008) between an intraoceanic arc and a continental arc (Figure 5a), given that (a) the
SSM, which presents characteristics of a subduction channel related to an intraoceanic arc (Escuder‐Viruete
et al., 2019; Flores et al., 2015; Venable, 1994), displays microstructures indicating a top‐to‐the‐NNE inverse
shear sense, suggesting that the subduction zone dipped toward the SSW in present‐day coordinates, that is,
opposite to the Chortis Block (Escuder‐Viruete et al., 2019; Figure 4), and (b) arc volcanism was active on the
continental Chortis Block and neighboring Mixteca and Xolapa terranes (SW Mexico) prior to the collision
event (Campa‐Uranga et al., 2012; Ducea et al., 2004; Gordon, 1993; Rogers, Mann, & Emmet, 2007).
Based on the new results obtained from the sedimentary overlap, we are able to further unravel the mode
and timing of arc‐continent collision. The Hauterivian age of (i) the youngest cooling event deduced from
metamorphic blocks enclosed within the SSM and (ii) the oldest rocks of the deep water overlap succession
(Figure 3a) indicate that the SSM was exhumed to the seafloor by that time. Moreover, the provenance of the
overlap El Amparo Formation shows that sediments sourced from the Eastern Chortis Terrane were shed on
the SSM during its exhumation along the suture zone (Figure 5b). Therefore, we infer that the attempted
subduction of the Chortis forearc triggered the exhumation of the SSM (e.g., forced return by subduction
of continental material; Brun & Faccenna, 2008; Agard et al., 2009) and caused the development of a flexural
bulge (e.g., Jacobi, 1981) in the Chortis continental crust (Figure 5b). The upwarping may have been respon-
sible for a short‐lived uplift event that generated the conglomeratic material of the Hauterivian El Amparo
Formation (Figure 5b).
The arc‐continent collision event did not produce obvious contractional structures along the convergent
margin inboard of the Chortis Block and SW Mexico. Rather, the system underwent extension shortly after
the onset of collision, which is attested by the numerous Barremian‐Aptian intra‐arc to back‐arc rifts that
formed along the continental convergent margin inboard of Mixteca (Sierra‐Rojas et al., 2016; Sierra‐Rojas
& Molina‐Garza, 2014) and Chortis (Rogers, Mann, Emmet, & Venable, 2007, Rogers, Mann, Scott, &
Patino, 2007c) terranes. Extension was possibly initiated by the detachment of the double‐dipping oceanic
lithosphere from the overlying colliding arcs (e.g., Zhang et al., 2017) and continued as an east dipping slab
rolled back to establish a new subduction zone (Figures 5b and 5c).
Moreover, a Molucca Sea‐type collision model implies a stage of volcanic arc quiescence (e.g., Hall & Wilson,
2000), which coincides with the time interval between the sinking onset of the subducting slab and renewed
subduction of a succeeding slab. In the case of the Molucca Sea collision zone, such a time interval may be as
short as 2 m.y. (Hall & Smyth, 2008). Therefore, lulls in arc volcanism during ancient collision events such as
the one between Siuna and Chortis/SW Mexico may be easily overlooked, notably due to (a) the fact that
ages of magmatic and volcaniclastic rocks in SW Mexico are generally characterized by uncertainties of at
least ±1 Ma (Sierra‐Rojas et al., 2016), which leads to an apparently continuous record of volcanic arc activ-
ity during the 140‐ to 130‐Ma interval, and (b) the fact that Lower Cretaceous volcanic rocks of the Chortis
Block remain poorly dated. The scarcity of volcanic rocks of Earliest Cretaceous age on the Chortis Block
may also be explained by oblique subduction beneath SW North America during this time, which could have
led to the development of a magmatic‐poor margin (Pindell et al., 2012).
After a period from which few magmatic ages were reported (i.e., Berriasian‐Valanginian, approximately
145–135 Ma; Talavera‐Mendoza et al., 2007; Campa‐Uranga et al., 2012; Ortega‐Flores et al., 2013; 2015;
Molina Garza et al., 2017), increased volcanic arc activity appears to have been associated with extensional
intra‐arc to back‐arc basin formation in the Chortis Block and Mixteca Terrane from at least the Barremian
(approximately 130 Ma; Sierra‐Rojas et al., 2016). The rejuvenation of continental arc volcanism associated
with an extensional regime is a typical phase of early subduction along continental arcs (Busby, 2012), which
in this case follows arc‐continent collision. According to Busby (2004, 2012), this regime leads to the devel-
opment of early‐stage low‐lying grabens filled with kilometers of volcanic and volcaniclastic/carbonate/cra-
ton‐derived sedimentary rocks formed in nonmarine to deep‐marine conditions. These features characterize
the Barremian‐Aptian basins of Mixteca and Chortis terranes, which show remarkable similarities in terms
of extension timing and infill deposits (Molina Garza et al., 2017). In the basins of both terranes, continental

ANDJIĆ ET AL. 4412


Tectonics 10.1029/2019TC005741

sedimentation of basement‐derived conglomerates are succeeded by intercalations of shallow‐marine


carbonate/volcaniclastic rocks and volcanic flows (Rogers, Mann, Scott, & Patino, 2007; Sierra‐Rojas et al.,
2016; Figure 5c). The Siuna area was located in a forearc basin where deep‐marine deposits were interstra-
tified with andesitic flows (Flores, 2009; Flores et al., 2015; Venable, 1994), suggesting that postcollisional arc
volcanic activity established in the vicinity of the suture between the Siuna Intraoceanic Arc and the Chortis
Block (Figure 5c).

7. Margin‐Scale Geodynamic Scenario


Numerous models have been proposed for the Late Jurassic‐Early Cretaceous evolution of Middle America
(e.g., Boschman et al., 2014; Boschman et al., 2018; Centeno‐García et al., 2008; Dickinson & Lawton, 2001;
Fitz‐Díaz et al., 2018; Flores, 2009; Mann et al., 2007; Peña‐Alonso et al., 2018; Pindell & Kennan, 2009;
Sedlock et al., 1993; Umhoefer, 2003). Given that it is beyond the scope of this paper to comprehensively
review existing models, we propose here a four‐step geodynamic scenario that, in our view, represents a sim-
ple framework for the Late Jurassic‐Early Cretaceous evolution of the convergent margins of the Chortis
Block and WSW Mexico (mainly based on Flores, 2009; Bandini et al., 2011; Madrigal et al., 2016).

7.1. Middle Jurassic (Approximately 174–163 Ma): Extensional Continental Convergent Margin
Eastward subduction beneath western Mexico produced the continental Nazas arc (Bartolini et al., 2003;
Dickinson & Lawton, 2001). Lawton and Molina‐Garza (2014) postulated that the Nazas arc was composed
of basins similar to the arc‐graben depression of western United States, where volcanism and volcaniclastic
and siliciclastic sedimentation co‐occurred. Fitz‐Díaz et al. (2018) and Martini and Ortega‐Gutiérrez (2018)
suggested that the extensional regime was a combination of slab rollback and North America‐South America
separation (i.e., Pangea breakup and opening of the Gulf of Mexico). We envisage that the Chortis Block was
in the southward prolongation of the Mexican Nazas extensional arc (Figures 6a and 7a), as suggested by the
geology of the Agua Fria Formation. Gordon (1990) describes the depositional environment of the Agua Fria
Formation as a nonmarine to deep‐marine basin where dominantly siliciclastic sediments were interstrati-
fied with volcanic flows and tuffs (>500 m thick east of the Guayape Fault). The depositional architecture
was controlled by normal faults, the activity of which led to kilometric thickness variations of the Agua
Fria Formation over short distances.
The size and age of the Mezcalera Plate (e.g., Dickinson & Lawton, 2001; Figures 6a and 7a), which was sub-
ducted beneath southwest North America, remains poorly constrained due to the scarcity of its remnants.
Bajocian red ribbon radiolarite in sedimentary contact with arc‐derived greenstone within the SSM
(Figures 2 and 3) suggests that (a) the convergent margin inboard of the Siuna Intraoceanic Arc had been
active since at least 170 Ma (Baumgartner et al., 2008; Flores et al., 2015), which is in accordance with the
148‐Ma metamorphic peak reported from an IAT‐related amphibolite block within the SSM (Escuder‐
Viruete et al., 2019); (b) the basement of the Siuna Intraoceanic Arc was a Middle Jurassic (and possibly
older) oceanic crust of the Mezcalera Plate; and (c) intraoceanic arc volcanism took place in the Pacific
realm, given that red ribbon radiolarite deposition did not occur along the American margins and in the
proto‐Caribbean realm (Bandini et al., 2011; Baumgartner, 2013; Figure 6a). Assuming translation rates
up to 11 cm/year (e.g., Holt et al., 2017; Johnston & Borel, 2007), approximately 40 m.y. of convergence
recorded in the SSM suggest that the Siuna Intraoceanic Arc could have initiated as far as 4,400 km from
North America during the Middle Jurassic. Models of Zhang et al. (2017) suggest that plate widths much lar-
ger than 2,000 km are not sustainable in natural Molucca Sea‐type systems unless a dense slab is involved.
Therefore, the model in Figure 6a shows a reconstruction of the Mesozoic Mezcalera Ocean, which likely
comprised several oceanic plates of limited size. These plates were eventually subducted along the western
Americas in a southward zipper closure during the Early Cretaceous (e.g., Bandini et al., 2011; Figure 6a).

7.2. Late Jurassic (Approximately 163–145 Ma): Forearc to Back‐Arc Rifting and Spreading
Rollback of the subducting plate eventually led to the rifting of the convergent margin and opening of the
Arperos back‐arc basin; oceanic spreading initiated during the Early Cretaceous (Sedlock et al., 1993;
Martini et al., 2011; Figures 6b and 7b). Coeval back‐arc opening occurred along western South America
(Braz et al., 2018; Figure 6b). Rifting‐separation histories of the arc seem to have differed between western
Mexico and southwestern Mexico. Along western Mexico, the locus of rupture was located in the forearc,

ANDJIĆ ET AL. 4413


Tectonics 10.1029/2019TC005741

creating the intraoceanic arc of the Guerrero Terrane, as demonstrated by the scarce continent‐derived
turbidites with zircons of Gondwana affinity (Arteaga Complex) and overall lack of continental crust in
the basement of the intraoceanic arc (Centeno‐García et al., 2008, Centeno‐García et al., 2011; Figures 6b
and 7b). In contrast, along southwestern Mexico, a continental ribbon made up of the Mixteca and
Chortis terranes was separated from the North American Plate following the onset of rifting in proximal
forearc to intra‐arc settings (Helbig et al., 2012; Figures 6b and 7b). The reason for the migration of the rift
location along the convergent margin is uncertain, given that the opening of back‐arc basins is subject to
multiple controls (Martinez & Taylor, 2006). We speculate that the different styles of rifting along the margin
was due to a combination of variable crustal structure as well as different tectonic configurations at the ends
of both systems (incipient formation of oceanic crust in the proto‐Caribbean to the S versus stable cratonic
area to the N, Pindell & Kennan, 2009; see also Taylor, 1992; Figure 6b).
The Chortis Block was located in the structurally complex intersection between the convergent margin of
SW America and the extensional margin of the W proto‐Caribbean (Figure 6b). Cessation of the arc activity
as recorded in the Agua Fria Formation suggests that the locus of magmatism shifted elsewhere as the slab
rolled back; strike‐slip faulting of the convergent margin related to the opening of the proto‐Caribbean could
have also played a role in the migration/cessation of arc volcanic activity along the Chortis Block (Pindell &
Kennan, 2009). Moreover, cessation of marine sedimentation in the interior of the Chortis Block implies that
this area became the WNW shoulder of a rift that gave way to seafloor spreading in the Arperos Basin and/or
in the western proto‐Caribbean (Pindell & Kennan, 2009; Figure 7b).

7.3. Hauterivian‐Aptian (Approximately 134–113 Ma): Diachronous Closure of the


Mezcalera Ocean
The collision of the Siuna Intraoceanic Arc with the SW margin of Mexico was a short‐lived event that
occurred during the late Valanginian‐Hauterivian, as implied by the succession of mélange exhumation,
overlap sedimentation, and convergent margin extension within approximately 5 m.y. (Figures 2, 3, and
5). Postcollisional extension affected the fringing intraoceanic arc to continental arc of WSW Mexico. This
extensional event initiated the extensional episode that formed the bulk of the Guerrero Intraoceanic Arc
(“Early Cretaceous extensional arc assemblage” of Centeno‐García et al., 2011) as well as intra‐arc to back‐
arc rifting of the Mixteca Terrane and Chortis Block (Figures 5c, 6c, and 7c).
Additional evidence points toward an ocean basin closure along SW North America during the Early
Cretaceous. Flores (2009) and Flores et al. (2015) argued for a southward zipper closure of the Mezcalera
Ocean that separated the Siuna Intraoceanic Arc from North America. In their model, the diachronous clo-
sure of the ocean basin is revealed by the southward younging of metamorphic peak and exhumation ages
obtained from blocks within mélanges that crop out in Baja California (Cedros Island, Baldwin &
Harrison, 1992), Siuna (Baumgartner et al., 2008), and South Motagua (Flores et al., 2013). Ocean basin clo-
sure occurred along the western proto‐Caribbean realm later than inboard of the Chortis Block (Bandini
et al., 2011; Figures 6 and 7). During the Hauterivian‐Aptian, the western margin of the proto‐Caribbean
oceanic realm was bounded by an IAT (Pindell et al., 2012; Escuder‐Viruete & Castillo‐Carrión, 2016;
Torró et al., 2017; Figures 6c and 7c) and/or a west facing intraoceanic arc system (Escuder‐Viruete et al.,
2013; Neill et al., 2010, 2012). Several key points suggest an arc collision occurred inboard of the western
proto‐Caribbean realm during the Aptian. (1) Exhumation ages of blocks within the Pacific‐related South
Motagua Mélange yielded dates of 125–113 Ma (Brueckner et al., 2009; Flores et al., 2013, 2015; Harlow
et al., 2004). The South Motagua Mélange was juxtaposed with the Chortis Block by strike‐slip tectonics dur-
ing the Cenozoic (Flores & Harlow, 2013; Pindell et al., 2012; Torres‐de León et al., 2012). (2) Aptian collision
of an intraoceanic arc is documented in Hispaniola from a range of events such as batholith exhumation,
back‐arc deformation, and arc volcanic activity cessation (Escuder‐Viruete & Pérez‐Estaún, 2013). (3)
Inception of SW dipping subduction occurred inboard Hispaniola and Cuba during the 125‐ to 110‐Ma inter-
val (Escuder‐Viruete et al., 2013; Krebs et al., 2008; Lázaro et al., 2016), an event possibly triggered by arc col-
lision (Escuder‐Viruete & Castillo‐Carrión, 2016; Figure 7d). (4) Back‐arc deformation occurred at
approximately 106 Ma in the Late Jurassic arc basement of La Désirade Island (Guadeloupe; Corsini et al.,
2011), which is approximately 5 m.y. later than in Hispaniola (Escuder‐Viruete et al., 2009). Diachronous
back‐arc deformation may be explained by a more southern position of La Désirade Island along the Siuna
Intraoceanic Arc (Figure 6d).

ANDJIĆ ET AL. 4414


Tectonics 10.1029/2019TC005741

7.4. Albian (Approximately 113–100 Ma): Closure of the Arperos Back‐Arc Basin
The Albian was characterized by the closure of back‐arc basins along SW North America (Martini et al.,
2014) and W South America (Maloney et al., 2011; Braz et al., 2018; Figure 6d), possibly in response to an
increased westward velocity of North America and South America (Spikings et al., 2015; Umhoefer, 2003).
Along SW North America, closure of the Arperos Basin coincided with the accretion of the Alisitos and
Guerrero fringing intraoceanic arcs (Busby, 2004; Centeno‐García et al., 2011; Figures 6d and 7d). Onset
of back‐arc closure could have initiated as early as the Barremian (Bandini et al., 2011), as suggested by
the formation of the North Motagua Mélange (Harlow et al., 2004; Brueckner et al., 2009; Flores et al.,
2015; Figures 6c and 7c). At the junction between the Arperos and proto‐Caribbean basins, back‐arc closure
led to the subduction of the Maya Block (Bandini et al., 2011; Flores et al., 2015) thus forming the Chuacús
Complex (approximately 100‐Ma metamorphic peak; Maldonado et al., 2017; Figures 6d and 7d). The restric-
tion of metamorphic events to the southern part of the back‐arc suggests that the basin was significantly
wider in its southern segment, possibly due to the proximity of the spreading proto‐Caribbean oceanic floor,
thus allowing for further shortening to occur during subsequent back‐arc closure (Figures 6 and 7).

8. Conclusions
Presently exposed in the heart of central Nicaragua, the Siuna Belt represents a suture zone that recorded an
arc‐continent collision between the exotic Siuna Intraoceanic Arc and the southwest margin of North
America represented by the Chortis Block (at approximately 134–131 Ma; Figure 5). The Hauterivian timing
of the collision event is evidenced by the youngest exhumation ages of metamorphic blocks within the SSM
and the depositional age of the overlying Chortis Block‐derived conglomerate (El Amparo Formation;
Figure 3). Discrete, fine‐grained ophiolitic detritus in the deep‐marine overlap sedimentary succession indi-
cates that the SSM was exhumed along low‐relief structures below the sea level (Figure 5b).
Previously thought to result from the collision between the Greater Antillean Arc and SW North America dur-
ing the Campanian (Rogers, Mann, & Emmet, 2007; Venable, 1994), the Siuna Belt actually recorded an earlier
collision between an exotic intraoceanic arc system and SW North America that led to a Molucca Sea‐type clo-
sure of the Mesozoic Mezcalera Ocean (Figures 5 and 6). We envisage that this episode triggered the following
Barremian‐Aptian events along the margins of SW North America and the W proto‐Caribbean, from north to
south: (1) main extensional and volcanic phases of the Guerrero fringing intraoceanic arc (e.g., Centeno‐García
et al., 2011); (2) intra‐arc to back‐arc rifting of and renewed arc volcanic activity on Chortis and Mixteca ter-
ranes (Rogers, Mann, Scott, & Patino, 2007; Sierra‐Rojas et al., 2016); (3) onset of WSW dipping subduction
along the western margin of the proto‐Caribbean (Caribbean Arc; Escuder‐Viruete et al., 2013), which incor-
porated exhumed mélanges related to the arc collision (e.g., South Motagua Mélange; Pindell et al., 2012;
Flores et al., 2013). The Colon Fold Belt of the Eastern Chortis Terrane did not record the collision of the
Siuna Intraoceanic Arc; rather, it formed during the Late Cretaceous time‐transgressive collision of the
Caribbean Large Igneous Province with North America (Andjić et al., 2019).
We emphasize the key role of overlap sedimentary assemblages in deciphering arc‐continent collision
events. In particular, biostratigraphic data provide important constraints on the timing of tectonic events
related to the collision of intraoceanic arcs and the postcollisional evolution of convergent margins, such
as the syncollisional record of margin uplift (e.g., El Amparo Formation) followed by postcollisional subsi-
dence and renewed subduction and arc volcanic activity (e.g., Rio Matis Formation). Moreover, when com-
bined together, cooling ages of metamorphic blocks within mélanges and depositional ages of overlap
sedimentary rocks can provide valuable ties on the timing of subduction channel exhumation, the rate of
which could have reached up to 20 mm/year in the case of the SSM (Escuder‐Viruete et al., 2019).

Appendix A

Figure A1 shows the facies of the Hauterivian El Amparo and latest Hauterivian–earliest Barremian to
Albian Rio Matis formations. Figures A2 to A5 illustrate the micro‐ and macrofossils that are used to con-
strain the timing of deposition of the Rio Matis Formation.

ANDJIĆ ET AL. 4415


Tectonics 10.1029/2019TC005741

Figure A1. Field photographs and thin section photomicrographs of the Rio Matis Formation (a–g) and the underlying El Amparo Formation (h–i), Siuna area,
northern Nicaragua. See Figure 3b for names of localities. (a) Volcaniclastic turbidite overlying a dark pelagic limestone (Rio Matis, N13°48′38.9″/W84°45′
37.5″). The GPS device is 10 cm long. (b) Calciturbidite composed of a poorly sorted conglomeratic level overlain by a sandy well‐sorted level (Rio Matis, N13°48′
17.6″/W84°45′53.4″). Plane polarized light (PPL). O = Orbitolinid; Ga = Green alga. (c) Tuffaceous turbidite composed of clasts of radiolarian‐bearing tuffaceous
siltstone enclosed in a sandy crystal ash matrix (locality 1 km south of Tadasna, N13°34′53.3″/W84°48′51.8″). Early Aptian radiolarians were extracted from the
overlying tuffaceous siltstone (see sample NI14‐52 in Figure A2). The GPS device is approximately 5 cm long. (d) Clast‐supported volcaniclastic conglomerate
composed of clasts of lithic wacke (Lw), andesite (And), tuff (T), altered volcanic glass/material (Ag), and scarce polycrystalline quartz (El Caracol, N13°35′40.9″/
W84°49′14.3″). PPL. (e), (f) Serpentine grains (antigorite) in an arkosic‐lithic arenite (Rio Matis, N13°45′37.6″/W84°47′37.0″). Cross‐polarized light (CPL).
(g) Red radiolarian chert grain (El Caracol, N13°35′40.9″/W84°49′14.3″). PPL. (h) Clast‐supported conglomerate composed of clasts of quartz wacke (Qw), quartz
arenite (Qa), polycrystalline quartz (Qp), and scarce radiolarian‐bearing tuffaceous siltstone (El Amparo, N13°39′59.2″/W84°49′44.4″). Note the incipient
cleavage in the largest clast of quartz wacke. The red rectangle indicates the quartz clast illustrated in (i). PPL. (i, j) Clasts of quartzite displaying recrystallized
quartz ribbons (for [i], same sample as [h]; for [j], Las Delicias, N13°48′07.9″/W84°45′23.0″). CPL.

ANDJIĆ ET AL. 4416


Tectonics 10.1029/2019TC005741

Figure A2. Scanning electron microscope images of radiolarians from the Rio Matis Formation, Siuna area, northern Nicaragua. Sample NI14‐52 (early Aptian),
N13°34′53.3″/W84°48′51.8″: (1) Pseudodictyomitra carpatica (Lozyniak); (2), (3), and (4) Pseudodictyomitra lodogaensis Pessagno; (5) and (6) Pseudodictyomitra
hornatissima (Squinabol); (7) Thanarla pulchra (Squinabol); (8) and (9) Parvimitrella communis (Squinabol); (10) Svinitzium puga (Schaaf). Sample NI14‐28 (late
Hauterivian‐earliest Barremian), N13°39′59.2″/W84°49′44.4″: (11) Loopus sp.; (12) Pseudodictyomitra sp.; (13) Pseudodictyomitra leptoconica (Foreman); (14)
Pseudodictyomitra carpatica (Lozyniak); (15) Pseudodictyomitra suyarii Dumitrica, Immenhauser, & Dumitrica‐Jud; (16) Loopus blabla (Schaaf) sensu Dumitrica,
Immenhauser, & Dumitrica‐Jud; (17) Caneta sp.; (18) Thanarla pacifica Nakaseko & Nishimura; (19) Pantanellium sp.; (20) Dibolachras polylophia (Foreman); (21)
Dibolachras limatum (Foreman); (22) Hiscocapsa kaminogoensis (Aita); (23) Dibolachras cf. D. coronata (Steiger); (24) Hemicryptocapsa capita (Tan); (25)
Hiscocapsa uterculus (Parona); (26) Squinabollum asseni (Tan); (27) and (28) Squinabollum simplex (Taketani); (29) and (30) Cana septemporatus (Parona); (31) and
(32) Obesacapsula verbana (Parona).

ANDJIĆ ET AL. 4417


Tectonics 10.1029/2019TC005741

Figure A3. Transmitted light images of planktonic foraminifera from the Rio Matis Formation, Siuna area, northern Nicaragua. Scale for all images is 0.1 mm as in
(1). Sample NI14‐32 (Thalmanninella appenninica zone, latest Albian), N13°47′40.9″/W84°46′23.8″: (1) and (2) Biticinella breggiensis (Gandolfi); (3) and (4)
Ticinella madecassiana Sigal; (5), (6), and (7) Praeglobotruncana delrioensis (Plummer); (8) and (9) Praeglobotruncana stephani (Gandolfi); (10)
Pseudothalmanninella ticinensis (Gandolfi); (11) and (12) Thalmanninella gandolfii Luterbacher & Premoli Silva; (13), (14), (15), and (16) Thalmanninella appen-
ninica (Renz).

ANDJIĆ ET AL. 4418


Tectonics 10.1029/2019TC005741

Figure A4. Samples from limestone blocks and chanel fills in the Rio Matis Formation, Siuna area, northern Nicaragua. Scale bar is 2 mm, except when indicated
otherwise. (1) Sample NI14‐19C(3). Two subaxial sections of Orbitolina (Mesorbitolina) cf. texana (Roemer) showing a narrow exoskeleton in the marginal area of
the cone and a well‐developed endoskeleton in the radial and central area. (2), (3), and (6) Sample 05‐01‐26‐07 (Las Delicias, collected by P. O. B. and Kennet Flores
in 2007). Axial section of Orbitolina (Mesorbitolina) cf. texana (Roemer), after Baumgartner‐Mora et al. (2013). (2) and (3) show the large, complex embryonic
apparatus, with a flat embryo (e), numerous supraembryonic (se), and periembryonic chambers (pe), suggesting a late Albian age (see text). Scale bar in 3 is 0.5 mm.
(4) Sample ORB4(10). Subaxial section of Orbitolina (Mesorbitolina) sp. (4) and (7) are specimens with abundant agglutinated sponge raxes that co‐occur with
Orbitolina (Mesorbitolina) cf. texana without sponge raxes. (5) Sample NI14‐19C(3). Almost centered oblique section near the axial plane of Orbitolina
(Mesorbitolina) cf. texana (Roemer) cutting tangentially the exoskeleton (beams and rafters are visible in the marginal area of the cone) and obliquely, from top to
bottom, the radial and central parts, the interior of the cut permits to observe the alternating septula (black) and chamberlets (white). Scale bar 1 mm. (7) Sample NI
14‐02a. Transverse oblique section of Orbitolina (Mesorbitolina) sp. showing exo‐ and endoskeleton. (8), (9), and (11) Microfacies constituted by rounded
extraclasts of pelletoidal packstone with some small foraminifera and Orbitolina (Mesorbitolina) cf. texana. (8) Sample ORB2(6). (9) Sample ORB3(9). (11) Detail of
(8) showing oblique section of embryonic apparatus measuring at least 450 μm; scale bar 200 μm. (10) Sample NI14‐08a(18). Detail of a centimeter‐size extraclast
of a pelletoidal packstone rich in Orbitolina (Mesorbitolina) cf. texana (two subaxial sections) and other small foraminifera. (12) Sample ORB6(8)b. Transverse
section of a small Paracoskinolina? showing the exoskeleton constituted only by beams (of two orders) and endoskeleton with pillars. Note that the external wall is
thicker than the epidermis of Orbitolina, the exoskeleton is simpler and penetrates deeper to the interior of the chamber, and the central area is occupied by
pillars instead of septula forming chamberlets as in Orbitolina. Scale bar is 200 μm. Sample localities: NI14‐19 (N13°47′25.2″/W84°45′52.3″); 05‐01‐26‐07, ORB, and
NI14‐08 (N13°47′30.0″/W84°46′06.5″); NI14‐02 (N13°48′17.6″/W84°45′53.4″).

ANDJIĆ ET AL. 4419


Tectonics 10.1029/2019TC005741

Figure A5. Outcrop photographs of lower Albian rudists form the Siuna area, northern Nicaragua. Scale for all images is 1 cm as in (1), except for (8) where scale is
3 cm. (1), (2), (3) Coalcomana sp. (4), (5), (6), and (11) Caprinuloidea perfecta Palmer. (7) Toucasia sp. (8) Polyconitid. (9) Coalcomana ramosa (Boehm).
(10) Caprinuloidea cf. C. multitubifera Palmer. GPS locations: (1), (2), (4), (5), and (7) N13°47′39.2″/W84°46′28.1″; (3) N13°48.377′/W84°45.802′; (6), (8), and
(10) N13°48.375′/W84°45.793′; (9) N13°48.293′/W84°45.890′; (11) N13°48.375′/W84°45.743′.

ANDJIĆ ET AL. 4420


Tectonics 10.1029/2019TC005741

Acknowledgments References
We thank the Instituto de Geología y
Agard, P., Yamato, P., Jolivet, L., & Burov, E. (2009). Exhumation of oceanic blueschists and eclogites in subduction zones: Timing and
Geofísica (Universidad Nacional
mechanisms. Earth‐Science Reviews, 92(1–2), 53–79. https://doi.org/10.1016/j.earscirev.2008.11.002
Autónoma de Nicaragua) and espe-
Aitchison, J. C., Xia, X., Baxter, A. T., & Ali, J. R. (2011). Detrital zircon U‐Pb ages along the Yarlung‐Tsangpo suture zone, Tibet:
cially to Dionisio Rodríguez Altamirano
Implications for oblique convergence and collision between India and Asia. Gondwana Research, 20, 691–709. https://doi.org/10.1016/j.
for logistical and administrative sup-
gr.2011.04.002
port; Elliet Pérez Romero, Mélida
Andjić, G., Baumgartner, P. O., & Baumgartner‐Mora, C. (2018). Rapid vertical motions and formation of volcanic arc gaps: Plateau collision
Schliz, and Bernard Mercier de Lépinay
recorded in the forearc geological evolution (Costa Rica margin). Basin Research, 30(5), 863–894. https://doi.org/10.1111/bre.12284
for the help in the field; Kennet Flores
Andjić, G., Baumgartner, P. O., & Baumgartner‐Mora, C. (2019). Collision of the Caribbean Large Igneous Province with the Americas:
for sharing thin sections of the study
Earliest evidence from the forearc of Costa Rica. Geological Society of America Bulletin, 131(9‐10), 1555–1580. https://doi.org/10.1130/
area; Gérard Stampfli for passing on
B35037.1
some figure templates; Associate Editor
Baldwin, S. L., & Harrison, T. M. (1992). The P‐T‐t history of blocks in serpentinite‐matrix mélange, west‐central Baja California. Geological
Jason Ali and reviewer Rob Rogers for
Society of America Bulletin, 104(1), 18–31. https://doi.org/10.1130/0016‐7606(1992)104<0018:TPTTHO>2.3.CO;2
their helpful and insightful reviews that
Bandini, A. N., Baumgartner, P. O., Flores, K., Dumitrica, P., Hochard, C., Stampfli, G. M., & Jackett, S.‐J. (2011). Aalenian to Cenomanian
greatly improved the manuscript; and
Radiolaria of the Bermeja Complex (Puerto Rico) and Pacific origin of radiolarites on the Caribbean Plate. Swiss Journal of Geosciences,
Editor‐in‐Chief John Geissman for the
104(3), 367–408. https://doi.org/10.1007/s00015‐011‐0072‐2
careful editorial handling. This
Bartolini, C., Lang, H., & Spell, T. (2003). Geochronology, geochemistry, and tectonic setting of the Mesozoic Nazas arc in north‐central
research was supported by funds of the
Mexico, and its continuation to northern South America. In C. Bartolini, R. T. Buffler, & J. Blickwede (Eds.), The Circum‐Gulf of Mexico
Swiss National Science Foundation
and the Caribbean: Hydrocarbon habitats, basin formation, and plate tectonics, American Association of Petroleum Geologists Memoir (Vol.
(projects 162670 and 143894) and the
79, pp. 427–461). Tulsa, Oklahoma, U.S.A.: The American Association of Petroleum Geologists.
Herbette Foundation at the University
Baumgartner, P. O., Flores, K., Bandini, A. N., Girault, F., & Cruz, D. (2008). Upper Triassic to Cretaceous radiolaria from Nicaragua and
of Lausanne (to P. O. Baumgartner).
northern Costa Rica—The Mesquito composite oceanic terrane. Ofioliti, 33, 1–19.
The data for this paper are contained in
Baumgartner, P. O., O'Dogherty, L., Gorican, S., Urquhart, E., Pillevuit, A., & De Wever, P. (Eds) (1995). Middle Jurassic to Lower
the text, figures, and appendix and can
Cretaceous radiolaria of Tethys: Occurrences, systematics, biochronology. In Mémoires de Géologie (Vol. 23, 1162 p.). Lausanne,
also be found in the data repository
Switzerland: Université de Lausanne.
PANGAEA Data Archiving &
Baumgartner‐Mora, C., Baumgartner, P.O., Andjic, G., & Barat, F. (2013). Mid Cretaceous to Oligocene rise of the Middle American
Publication at https://doi.pangaea.de/
landbridge—documented by south‐eastwards younging Larger Foraminifera in shallow water carbonates (Nicaragua—Costa Rica—
10.1594/PANGAEA.907429.
Panama). 11th Swiss Geoscience Meeting, Lausanne, Abstract Volume pp. 187‐188.
Baumgartner‐Mora, C., Baumgartner, P. O., & Barat, F. (2013). Middle Cretaceous to Oligocene rise of the Middle American landbridge—
documented by south‐eastwards younging shallow water carbonates. Geophysical Research Abstracts, 15. EGU2013‐12742‐2
Boschman, L. M., Garza, R. S. M., Langereis, C. G., & van Hinsbergen, D. J. (2018). Paleomagnetic constraints on the kinematic relationship
between the Guerrero terrane (Mexico) and North America since Early Cretaceous time. Geological Society of America Bulletin, 130(7–8),
1131–1142. https://doi.org/10.1130/B31916.1
Boschman, L. M., van Hinsbergen, D. J. 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. https://doi.org/10.1016/j.earscirev.2014.08.007
Braz, C., Seton, M., Flament, N., & Müller, D. (2018). Geodynamic reconstruction of an accreted Cretaceous back‐arc basin in the Northern
Andes. Journal of Geodynamics, 121, 115–132. https://doi.org/10.1016/j.jog.2018.09.008
Brueckner, H. K., Avé Lallemant, H. G., Sisson, V. B., Harlow, G. E., Hemming, S. R., Martens, U., et al. (2009). Metamorphic reworking of
a high pressure–low temperature mélange along the Motagua fault, Guatemala: A record of Neocomian and Maastrichtian transpres-
sional tectonics. Earth and Planetary Science Letters, 284(1–2), 228–235. https://doi.org/10.1016/j.epsl.2009.04.032
Brun, J. P., & Faccenna, C. (2008). Exhumation of high‐pressure rocks driven by slab rollback. Earth and Planetary Science Letters, 272(1‐2),
1–7. https://doi.org/10.1016/j.epsl.2008.02.038
Buchs, D. M., Baumgartner, P. O., Baumgartner‐Mora, C., Flores, K., & Bandini, A. N. (2011). Late Cretaceous to Miocene tectonostrati-
graphy of the Azuero area (west Panama) and the discontinuous accretion and subduction erosion along the Middle American margin.
Tectonophysics, 512(1–4), 31–46. https://doi.org/10.1016/j.tecto.2011.09.010
Burkart, B., Denton, B., Dengo, C., & Moreno, G. (1987). Tectonic wedges and offset Laramide structures along the Polochic fault of
Guatemala and Chimpas, Mexico: Reaffirmation of large Neogene displacement. Tectonics, 9, 411–422.
Busby, C. J. (2004). Continental growth at convergent margins facing large ocean basins: A case study from Mesozoic Baja California,
Mexico. Tectonophysics, 392(1‐4), 241–277. https://doi.org/10.1016/j.tecto.2004.04.017
Busby, C. J. (2012). Extensional and transtensional continental arc basins: Case studies from the southwestern United States. In C. Busby, &
A. Azor (Eds.), Tectonics of sedimentary basins: Recent advances (pp. 382–404). Oxford, UK: Blackwell Publishing. https://doi.org/
10.1002/9781444347166.ch19
Campa‐Uranga, M. F., Torres de Leon, R., Iriondo, A., & Premo, W. R. (2012). Caracterizacion geologica de los ensambles metamorficos de
Taxco y Taxco el Viejo, Guerrero, Mexico. Boletín de la Sociedad Geológica Mexicana, 64(3), 369–385. https://doi.org/10.18268/
BSGM2012v64n3a8
Cardwell, R. K., Isacks, B. L., & Karig, D. E. (1980). The spatial distribution of earthquakes, focal mechanism solutions and subducted
lithosphere in the Philippines and northeast Indonesian islands. In D. E. Hayes (Ed.), The Tectonic and Geologic Evolution of Southeast
Asian Seas and Islands, Geophysical Monograph Series (Vol. 23, pp. 1–35). Washington, D.C.: American Geophysical Union. https://doi.
org/10.1029/GM023p0001
Carr, M., Feigenson, M., Patino, L., & Walker, J. (2003). Volcanism and geochemistry in Central America: Progress and problems. In J. Eiler
& G. Abers (Eds.), The subduction factory, Geophysical Monograph Series (Vol. 138, pp. 153–179). Washington, D.C.: American
Geophysical Union.
Carvajal‐Arenas, L. C., & Mann, P. (2018). Western Caribbean intraplate deformation: Defining a continuous and active microplate
boundary along the San Andres rift and Hess Escarpment fault zone, Colombian Caribbean Sea. American Association of Petroleum
Geologists Bulletin, 102(08), 1523–1563. https://doi.org/10.1306/12081717221
Cawood, P. A., Kröner, A., Collins, W. J., Kusky, T. M., Mooney, W. D., & Windley, B. F. (2009). Accretionary orogens through Earth
history. In P. A. Cawood, & A. Kröner (Eds.), Earth Accretionary Systems in Space and Time, Geological Society of London Special
Publication (Vol. 318, pp. 1–36). London, UK: The Geological Society.
Centeno‐García, E., Busby, C., Busby, M., & Gehrels, G. (2011). Evolution of the Guerrero composite terrane along the Mexican margin,
from extensional fringing arc to contractional continental arc. Geological Society of America Bulletin, 123(9‐10), 1776–1797. https://doi.
org/10.1130/B30057.1

ANDJIĆ ET AL. 4421


Tectonics 10.1029/2019TC005741

Centeno‐García, E., Guerrero‐Suastegui, M., & Talavera‐Mendonza, O. (2008). The Guerrero composite terrane of western Mexico:
Collision and subsequent rifting in a suprasubduction zone. In A. E. Draut, P. D. Clift, & D. W. Scholl (Eds.), Formation and applications
of the sedimentary record in arc collision zones, Geological Society of America Special Paper (Vol. 436, pp. 279–308). Boulder, Colorado:
The Geological Society of America.
Clift, P. D., Pavlis, T., DeBari, S. M., Draut, A. E., Rioux, M., & Kelemen, P. B. (2005). Subduction erosion of the Jurassic Talkeetna‐Bonanza
Arc and the Mesozoic accretionary tectonics of western North America. Geology, 33(11), 881–884. https://doi.org/10.1130/G21822.1
Clift, P. D., Schouten, H., & Vannucchi, P. (2009). Arc–continent collisions, subduction mass recycling and the maintenance of the con-
tinental crust. In P. A. Cawood, & A. Kröner (Eds.), Earth Accretionary Systems in Space and Time, Geological Society of London Special
Publication (Vol. 318, pp. 75–103). London, UK: The Geological Society.
Clift, P. D., & Vannucchi, P. (2004). Controls on tectonic accretion versus erosion in subduction zones: Implications for the origin and
recycling of the continental crust. Review of Geophysics, 42(2), RG2001. https://doi.org/10.1029/2003RG000127
Corsini, M., Lardeaux, J. M., Verati, C., Voitus, E., & Balagne, M. (2011). Discovery of Lower Cretaceous synmetamorphic thrust tectonics
in French Lesser Antilles (La Désirade Island, Guadeloupe): Implications for Caribbean geodynamics. Tectonics, 30, TC4005. https://doi.
org/10.1029/2011TC002875
Dickinson, W. R., & Lawton, T. F. (2001). Carboniferous to Cretaceous assembly and fragmentation of Mexico. Geological Society of America
Bulletin, 113(9), 1142–1160. https://doi.org/10.1130/0016‐7606(2001)113<1142:CTCAAF>2.0.CO;2
Donnelly, T. W., Horne, G. S., Finch, R. C., & Lopez‐Ramos, E. (1990). Northern Central America; the Maya and Chortis blocks. In G.
Dengo, & J. E. Case (Eds.), The Caribbean region, The Geology of North America (Vol. H, pp. 37–76). Boulder, Colorado: The Geological
Society of America.
Douglas, R. C. (1960). Revision of the family Orbitolinidae. Micropaleontology, 6(3), 249–270. https://doi.org/10.2307/1484232
Draut, A. E., Clift, P. D., & Scholl, D. W. (Eds) (2008). Formation and Applications of the Sedimentary Record in Arc Collision Zones,
Geological Society of America Special Paper (Vol. 436, 403 p.). Boulder, Colorado: The Geological Society of America. https://doi.org/
10.1130/SPE436
Ducea, M., Gehrels, G. E., Shoemaker, S., Ruiz, J., & Valencia, V. A. (2004). Geologic evolution of the Xolapa Complex, southern Mexico:
Evidence from U‐Pb zircon geochronology. Geological Society of America Bulletin, 116(7), 1016–1025. https://doi.org/10.1130/B25467.1
Escuder‐Viruete, J., Andjić, G., Baumgartner‐Mora, C., Baumgartner, P. O., Castillo‐Carrion, M., & Gabites, J. (2019). Origin and geody-
namic significance of the Siuna Serpentinite Mélange, Northeast Nicaragua: Insights from the large‐scale structure, petrology and
geochemistry of the ultramafic blocks. Lithos, 340‐341, 1–19. https://doi.org/10.1016/j.lithos.2019.05.002
Escuder‐Viruete, J., & Baumgartner, P. O. (2014). Structural evolution and deformation kinematics of a subduction‐related serpentinite‐
matrix mélange, Santa Elena Peninsula, northwest Costa Rica. Journal of Structural Geology, 66, 356–381. https://doi.org/10.1016/j.
jsg.2014.06.003
Escuder‐Viruete, J., Baumgartner, P. O., & Castillo‐Carrión, M. (2015). Compositional diversity in ophiolitic peridotites as result of a multi‐
process history: The Santa Elena ophiolite, northwest Costa Rica. Lithos, 231, 16–34. https://doi.org/10.1016/j.lithos.2015.05.019
Escuder‐Viruete, J., & Castillo‐Carrión, M. (2016). Subduction of fore‐arc crust beneath an intra‐oceanic arc: The high‐P Cuaba mafic
gneisess and amphibolites of the Rio San Juan Complex, Dominican Republic. Lithos, 262, 298–319. https://doi.org/10.1016/j.
lithos.2016.07.024
Escuder‐Viruete, J., & Pérez‐Estaún, A. (2013). Contrasting exhumation P–T paths followed by high‐P rocks in the northern Caribbean
subduction–accretionary complex: Insights from the structural geology, microtextures and equilibrium assemblage diagrams. Lithos,
160‐161, 117–144. https://doi.org/10.1016/j.lithos.2012.11.028
Escuder‐Viruete, J., Pérez‐Estaún, A., Weis, D., & Friedman, R. (2009). Geochemical characteristics of the Río Verde Complex, Central
Hispaniola: Implications for the paleotectonic reconstruction of the Lower Cretaceous Caribbean island‐arc. Lithos, 114, 168–185.
Escuder‐Viruete, J., Valverde‐Vaquero, P., Rojas‐Agramonte, Y., Jabites, J., Carrión‐Castillo, M., & Pérez‐Estaún, A. (2013). Timing of
deformational events in the Río San Juan complex: Implications for the tectonic controls on the exhumation of high‐P rocks in the
northern Caribbean subduction‐accretionary prism. Lithos, 177, 416–435. https://doi.org/10.1016/j.lithos.2013.07.006
Ferrari, L., Orozco‐Esquivel, T., Manea, V., & Manea, M. (2012). The dynamic history of the Trans‐Mexican Volcanic Belt and the Mexico
subduction zone. Tectonophysics, 522, 122–149.
Fitz‐Díaz, E., Lawton, T. F., Juárez‐Arriaga, E., & Chávez‐ Cabello, G. (2018). The Cretaceous–Paleogene Mexican orogen: Structure, basin
development, magmatism and tectonics. Earth‐Science Reviews, 183, 56–84. https://doi.org/10.1016/j.earscirev.2017.03.002
Flores, K. (2009). Mesozoic oceanic terranes of Southern Central America: Geology, geochemistry and geodynamics. Ph.D. thesis. Switzerland:
University of Lausanne. 290 p
Flores, K., Skora, S., Martin, C., Harlow, G. E., Rodríguez, D., & Baumgartner, P. O. (2015). Metamorphic history of riebeckite‐ and
aegirine‐augite‐bearing high‐pressure–low‐temperature blocks within the Siuna Serpentinite Mélange, northeastern Nicaragua.
International Geology Review, 57(5–8), 943–977. https://doi.org/10.1080/00206814.2015.1027747
Flores, K. E., & Harlow, G. E. (2013). Geodynamic evolution and tectonic history of the ophiolites and serpentinite mélanges in the
Guatemala Suture Zone. Paper presented at 109th Annual Meeting of the Cordilleran section of the Geological Society of America.
Abstracts with Programs, 45(6), 55.
Flores, K. E., Martens, U. C., Harlow, G. E., Brueckner, H. K., & Pearson, N. J. (2013). Jadeitite formed during subduction: In situ zircon
geochronology constraints from two different tectonic events within the Guatemala Suture Zone. Earth and Planetary Science Letters,
371‐372, 67–81. https://doi.org/10.1016/j.epsl.2013.04.015
Garzanti, E. (2016). From static to dynamic provenance analysis—Sedimentary petrology upgraded. Sedimentary Geology, 336, 3–13.
https://doi.org/10.1016/j.sedgeo.2015.07.010
Geldmacher, J., Hoernle, K., Van den Bogaard, P., Hauff, F., & Klügel, A. (2008). Age and geochemistry of the Central American forearc
basement (DSDP Leg 67 and 84): Insights into Mesozoic arc volcanism and seamount accretion on the fringe of the Caribbean LIP.
Journal of Petrology, 49(10), 1781–1815. https://doi.org/10.1093/petrology/egn046
Gordon, M., & Muehlberger, W. (1994). Rotation of the Chortís block causes dextral slip on the Guayape fault. Tectonics, 13(4), 858–872.
https://doi.org/10.1029/94TC00923
Gordon, M. B. (1990). Strike‐slip faulting and basin formation at the Guayape Fault‐Valle de Catacamas Intersection, Honduras, Central
America. Ph.D. thesis. Austin, TX: Department of Geological Sciences, University of Texas at Austin. 259 p.
Gordon, M. B. (1993). Revised Jurassic and Early Cretaceous (Pre‐Yojoa Group) stratigraphy of the Chortis block: Paleogeographic and
tectonic implications. In J. L. Pindell, & R. F. Perkins (Eds.), Mesozoic and Early Cenozoic development of the Gulf of Mexico and
Caribbean region: A context for hydrocarbon exploration (pp. 143–154). Austin, TX: Gulf Coast Section Society of Economic
Paleontologists and Mineralogists Foundation.

ANDJIĆ ET AL. 4422


Tectonics 10.1029/2019TC005741

Görög, A., & Arnaud‐Vanneau, A. (1996). Lower Cretaceous Orbitolinas from Venezuela. Micropalentology, 42(1), 65–78. https://doi.org/
10.2307/1485984
Hall, R., & Smyth, H. R. (2008). Cenozoic arc processes in Indonesia: Identification of the key influences on the stratigraphic record in
active volcanic arcs. In A. E. Draut, P. D. Clift, & D. W. Scholl (Eds.), Formation and applications of the sedimentary record in arc collision
zones, Geological Society of America Special Paper (Vol. 436, pp. 27–54). Boulder, Colorado: The Geological Society of America.
Hall, R., & Wilson, M. E. J. (2000). Neogene sutures in eastern Indonesia. Journal of Asian Earth Sciences, 18(6), 781–808. https://doi.org/
10.1016/S1367‐9120(00)00040‐7
Harlow, G. E., Hemming, S. R., Avé Lallemant, H. G., Sisson, V. B., & Sorensen, S. S. (2004). Two high pressure‐low‐temperature
serpentinite‐matrix mélange belts, Motagua fault zone, Guatemala: A record of Aptian and Maastrichtian collisions. Geology, 32(1),
17–20. https://doi.org/10.1130/G19990.1
Hastie, A. R., Kerr, A. C., Pearce, J. A., & Mitchell, S. F. (2007). Classification of altered volcanic island arc rocks using immobile trace
elements: Development of the Co–Th discrimination diagram. Journal of Petrology, 48(12), 2341–2357. https://doi.org/10.1093/petrol-
ogy/egm062
Hébert, R., Bezard, R., Guilmette, C., Dostal, J., Wang, C., & Liu, Z. (2012). The Indus–Yarlung Zangbo ophiolites from Nanga Parbat to
Namche Barwa syntaxes, southern Tibet: First synthesis of petrology, geochemistry, and geochronology with incidences on geodynamic
reconstructions of Neo‐Tethys. Gondwana Resaarch, 22(2), 377–397. https://doi.org/10.1016/j.gr.2011.10.013
Helbig, M., Keppie, J. D., Murphy, J. B., & Solari, L. A. (2012). U‐Pb geochronological constraints on the Triassic–Jurassic Ayu Complex,
southern Mexico: Derivation from the western margin of Pangea‐A. Gondwana Research, 22(3‐4), 910–927. https://doi.org/10.1016/j.
gr.2012.03.004
Hildebrand, R. S. (2013). Mesozoic Assembly of the North American Cordillera. In Geological Society of America Special Paper (Vol. 495, 169 p.).
Boulder, Colorado: The Geological Society of America. https://doi.org/10.1130/2013.2495
Hofker, J. (1963). Studies on the genus Orbitolina (Foraminiferida). Leidse Geologische Medelingen, 29, 181–253.
Holt, A. F., Royden, L. H., & Becker, T. W. (2017). The dynamics of double slab subduction. Geophysical Journal International, 209,
ggw496–ggw265. https://doi.org/10.1093/gji/ggw496
Jacobi, R. D. (1981). Peripheral bulge—A causal mechanism for the Lower/Middle Ordovician unconformity along the western margin of
the Northern Appalachians. Earth and Planetary Science Letters, 56, 245–251. https://doi.org/10.1016/0012‐821X(81)90131‐X
Johnston, S. T. (2001). The great Alaskan terrane wreck: Reconciliation of paleomagnetic and geological data in the northern Cordillera.
Earth and Planetary Science Letters, 193(3‐4), 259–272. https://doi.org/10.1016/S0012‐821X(01)00516‐7
Johnston, S. T. (2008). The Cordilleran ribbon continent of North America. Annual Reviews of Earth and Planetary Sciences, 36(1), 495–530.
https://doi.org/10.1146/annurev.earth.36.031207.124331
Johnston, S. T., & Borel, G. D. (2007). The odyssey of the Cache Creek terrane, Canadian Cordillera: Implications for accretionary orogens,
tectonic setting of Panthalassa, the Pacific superwell, and break‐up of Pangea. Earth and Planetary Science Letters, 253(3‐4), 415–428.
https://doi.org/10.1016/j.epsl.2006.11.002
Kobayashi, D., LaFemina, P., Geirsson, H., Chichaco, E., Abrego, A. A., Mora, H., & Camacho, E. (2014). Kinematics of the western
Caribbean: Collision of the Cocos Ridge and upper plate deformation. Geochemistry Geophysics Geosystems, 15(5), 1671–1683. https://
doi.org/10.1002/2014GC005234
Krebs, M., Maresch, W. V., Schertl, H.‐P., Münker, C., Baumann, A., Draper, G., et al. (2008). The dynamics of intra‐oceanic subduction
zones: A direct comparison between fossil petrological evidence (Rio San Juan Complex, Dominican Republic) and numerical simula-
tion. Lithos, 103(1‐2), 106–137. https://doi.org/10.1016/j.lithos.2007.09.003
Lawton, T. F., & Molina‐Garza, R. S. (2014). U‐Pb geochronology of the type Nazas Formation and superjacent strata, northeastern
Durango, Mexico: Implications of a Jurassic age for continental‐arc magmatism in north‐central Mexico. Geological Society of America
Bulletin, 126(9‐10), 1181–1199. https://doi.org/10.1130/B30827.1
Lázaro, C., Blanco‐Quintero, I. F., Proenza, J. A., Rojas‐Agramonte, Y., Neubauer, F., Núñez‐Cambra, K., & García‐Casco, A. (2016).
Petrogenesis and 40Ar/39Ar dating of proto‐forearc crust in the Early Cretaceous Caribbean arc: The La Tinta mélange (eastern Cuba) and its
easterly correlation in Hispaniola. International Geology Review, 58(8), 1020–1040. https://doi.org/10.1080/00206814.2015.1118647
Madrigal, P., Gazel, E., Flores, K. E., Bizimis, M., & Brian, J. (2016). Record of massive upwellings from the Pacific large low shear velocity
province. Nature Communications, 7(1), 13309. https://doi.org/10.1038/ncomms13309
Maldonado, R., Weber, B., Ortega‐Gutiérrez, F., & Solari, L. A. (2017). High‐pressure metamorphic evolution of eclogite and associated
metapelite from the Chuacús complex (Guatemala Suture Zone): Constraints from phase equilibria modelling coupled with Lu‐Hf and
U‐Pb geochronology. Journal of Metamorphic Geology, 36(1), 95–124. https://doi.org/10.1111/jmg.12285
Maloney, K. T., Clarke, G. L., Klepeis, K. A., Fanning, C. M., & Wang, W. (2011). Crustal growth during back‐arc closure: Cretaceous
exhumation history of Cordillera Darwin, southern Patagonia. Journal of Metamorphic Geology, 29(6), 649–672. https://doi.org/10.1111/
j.1525‐1314.2011.00934.x
Mann, P., Rogers, R. D., & Gahagan, L. (2007). Overview of plate tectonic history and its unresolved tectonic problems. In J. Bundschuh, &
G. E. Alvarado (Eds.), Central America: Geology, resources, hazards (Vol. 1, pp. 201–241). London, UK: Taylor & Francis.
Martinez, F., & Taylor, B. (2006). Modes of crustal accretion in back‐arc basins: Inferences from the Lau Basin. In D. M. Christie, C. R.
Fisher, S. M. Lee, & S. Givens (Eds.), Back‐arc spreading systems: Geological, biological, chemical, and physical interactions, Geophysical
Monographs Series (Vol. 166, pp. 5–30). Washington, D.C.: American Geophysical Union. https://doi.org/10.1029/166gm03
Martini, M., Mori, L., Solari, L., & Centeno‐García, E. (2011). Sandstone provenance of the Arperos Basin (Sierra de Guanajuato, central
Mexico): Late Jurassic–Early Cretaceous back‐arc spreading as the foundation of the Guerrero terrane. Journal of Geology, 119(6),
597–617. https://doi.org/10.1086/661989
Martini, M., & Ortega‐Gutiérrez, F. (2018). Tectono‐stratigraphic evolution of eastern Mexico during the break‐up of Pangea: A review.
Earth‐Science Reviews, 183, 38–55. https://doi.org/10.1016/j.earscirev.2016.06.013
Martini, M., Solari, L., & López‐Martínez, M. (2014). Correlating the Arperos Basin from Guanajuato, central Mexico, to Santo Tomás,
southern Mexico: Implications for the paleogeography and origin of the Guerrero terrane. Geosphere, 10(6), 1385–1401. https://doi.org/
10.1130/GES01055.1
Molina Garza, R. S., van Hinsbergen, D. J. J., Boschman, L. M., Rogers, R. D., & Ganerød, M. (2017). Large‐scale rotations of the Chortis
Block (Honduras) at the southern termination of the Laramide flat slab. Tectonophysics, 760, 36–57. https://doi.org/10.1016/j.
tecto.2017.11.026
Neill, I., Gibbs, J. A., Hastie, A. R., & Kerr, A. C. (2010). Origin of the volcanic complexes of La Désirade, Lesser Antilles: Implications for
tectonic reconstruction of the Late Jurassic to Cretaceous Pacific‐proto‐Caribbean margin. Lithos, 120(3‐4), 407–420. https://doi.org/
10.1016/j.lithos.2010.08.026

ANDJIĆ ET AL. 4423


Tectonics 10.1029/2019TC005741

Neill, I., Kerr, A. C., Hastie, A. R., Pindell, J. L., Millar, I. L., & Atkinson, N. (2012). Age and petrogenesis of the Lower Cretaceous North
Coast Schist of Tobago, a fragment of the proto‐Greater Antilles inter‐American arc system. Journal of Geology, 120(4), 367–384. https://
doi.org/10.1086/665798
Nokleberg, W. J., Parfenov, L. M., Monger, J. W. H., Norton, I. O., Khanchuk, A. I., Stone, D., et al. (2000). Phanerozoic tectonic evolution of
the Circum‐North Pacific, U.S. Geological Survey Professional Paper (Vol. 1626, 122 p). Denver, Colorado: U.S. Geological Survey.
O'Dogherty, L. (1994). Biochronology and paleontology of Mid‐Cretaceous radiolarians from Northern Apennines (Italy) and Betic Cordillera
(Spain), Mémoires de Géologie (Vol. 21, 415 p.). Lausanne, Switzerland: Université de Lausanne.
Ogg, J. G., Ogg, G. M., & Gradstein, F. M. (2016). A concise geologic time scale (240 p.). Amsterdam, Netherlands: Elsevier B.V. https://doi.
org/10.1016/C2009‐0‐64442‐1
Oncken, O. (1998). Evidence for precollisional subduction erosion in ancient collisional belts: The case of the Mid‐European Variscides.
Geology, 26(12), 1075–1078. https://doi.org/10.1130/0091‐7613(1998)026<1075:EFPSEI>2.3.CO;2
Ortega‐Flores, B., Solari, L., Lawton, T. F., & Ortega‐Obregon, C. (2013). Detrital zircon record of major Middle Triassic–Early Cretaceous
provenance shift, central Mexico: Demise of Gondwanan continental fluvial systems and onset of back‐arc volcanism and sedimenta-
tion. International Geology Review, 56, 237–261.
Padilla y Sánchez, R. J., Domínguez Trejo, I., López Azcárraga, A. G., Mota Nieto, J., Fuentes Menes, A. O., Rosique Naranjo, F., et al.
(2013). Tectonic Map of Mexico, American Association of Petroleum Geologists GIS Open Files series. Mexico, D.F.: National Autonomous
University of Mexico.
Peña‐Alonso, T. A., Molina‐Garza, R. S., Villalobos‐Escobar, G., Estrada‐Carmona, J., Levresse, G., & Solari, L. (2018). The opening and
closure of the Jurassic‐Cretaceous Xolapa basin, southern Mexico. Journal of South American Earth Sciences, 88, 599–620. https://doi.
org/10.1016/j.jsames.2018.10.003
Pindell, J. L., & Kennan, L. (2009). Tectonic evolution of the Gulf of Mexico, Caribbean and northern South America in the mantle
reference frame: An update. In K. H. James, M. A. Lorente, & J. L. Pindell (Eds.), The Origin and Evolution of the Caribbean Plate:
Geological Society of London Special Publication (Vol. 328, pp. 1–55). London, UK: The Geological Society.
Pindell, J. L., Maresch, W. V., Martens, U., & Stanek, K. (2012). The Greater Antillean Arc: Early Cretaceous origin and proposed rela-
tionship to Central American subduction mélanges: Implications for models of Caribbean evolution. International Geology Review, 54(2),
131–143. https://doi.org/10.1080/00206814.2010.510008
Premoli Silva, I., & Sliter, W.V. (2003). Practical manual of Cretaceous planktonic foraminifera (462 p.). Eni Italia.
Ratschbacher, L., Franz, L., Min, M., Bachmann, R., Martens, U., Stanek, K., et al. (2009). The North American‐Caribbean Plate boundary
in Mexico‐Guatemala‐Honduras. In K. H. James, M. A. Lorente, & J. L. Pindell (Eds.), The origin and evolution of the Caribbean Plate:
Geological Society of London Special Publication (Vol. 328, pp. 219–293). London, UK: The Geological Society. https://doi.org/10.1144/
SP328.11
Rogers, R. D., Mann, P., & Emmet, P. A. (2007). Tectonic terranes of the Chortis block based on integration of regional aeromagnetic
and geologic data. In P. Mann (Ed.), Geologic and tectonic development of the Caribbean plate in northern Central America, Geological
Society of America Special Paper (Vol. 428, pp. 65–88). Boulder, Colorado: The Geological Society of America. https://doi.org/10.1130/
2007.2428(04)
Rogers, R. D., Mann, P., Emmet, P. A., & Venable, M. E. (2007). Colon fold belt of Honduras: Evidence for Late Cretaceous collision
between the continental Chortis block and intra‐oceanic Caribbean arc. In P. Mann (Ed.), Geologic and tectonic development of the
Caribbean Plate in northern Central America, Geological Society of America Special Paper (Vol. 428, pp. 129–149). Boulder, Colorado: The
Geological Society of America. doi: https://doi.org/10.1130/2007.2428(06)
Rogers, R. D., Mann, P., Scott, R. W., & Patino, L. (2007). Cretaceous intra‐arc rifting, sedimentation, and basin inversion in east‐central
Honduras. In P. Mann (Ed.), Geologic and tectonic development of the Caribbean Plate in northern Central America, Geological Society of
America Special Paper (Vol. 428, pp. 89–128). Boulder, Colorado: The Geological Society of America. doi: https://doi.org/10.1130/
2007.2428(05)
Sanchez, J., Mann, P., Carvajal‐Arenas, L. C., & Bernal‐Olaya, R. (2019). Regional transect across the western Caribbean Sea based on
integration of geologic, seismic reflection, gravity, and magnetic data. American Association of Petroleum Geologists Bulletin, 103(2),
303–343. https://doi.org/10.1306/05111816516
Schroeder, R. (1975). General evolutionary trends in Orbitolinas (pp. 117–128). Special Issue: Revista Espanola de Micropaleontogia.
Scott, R., & Finch, R. (1999). Cretaceous carbonate biostratigraphy and environments in Honduras. In P. Mann (Ed.), Caribbean basins,
Sedimentary basins of the world series (Vol. 4, pp. 151–166). Amsterdam: Elsevier.
Scott, R., & Gonzalez‐Leon, C. (1991). Paleontology and biostratigraphy of Cretaceous rocks, Lampazos area, Sonora, Mexico. In E. Pérez‐
Segura, & C. Jacques‐Ayala (Eds.), Studies of Sonoran Geology, Geological Society of America Special Paper (Vol. 254, pp. 51–67). Boulder,
Colorado: The Geological Society of America.
Sedlock, R. L., Ortega‐Gutiérrez, F., & Speed, R. C. (1993). Tectonostratigraphic terranes and tectonic evolution of Mexico. In Geological
Society of America Special Paper (Vol. 278, 142 p.). Boulder, Colorado: The Geological Society of America. doi: https://doi.org/10.1130/
SPE278
Shervais, J. W., Choi, S. H., Sharp, W. D., Ross, J., Zoglman‐Schuman, M., & Mukasa, S. B. (2011). Serpentinite matrix mélange:
Implications of mixed provenance for mélange formation. In J. Wakabayashi, & Y. Dilek (Eds.), Mélanges: Processes of formation and
societal significance, Geological Society of America Special Paper (Vol. 480, pp. 1–30). Boulder, Colorado: The Geological Society of
America. doi: https://doi.org/10.1130/2011.2480(01)
Sierra‐Rojas, M. I., & Molina‐Garza, R. S. (2014). La Formacion Zicapa del sur de Mexico: Revision estratigrafica, sedimentologıa y
ambientes sedimentarios. Revista Mexicana de Ciencias Geologicas, 3, 1174–1189.
Sierra‐Rojas, M. I., Molina‐Garza, R. S., & Lawton, T. F. (2016). The Aptian Atzompa formation in south central Mexico: Record of evo-
lution from extensional back‐arc basin margin to carbonate platform. Journal of Sedimentary Research, 86(6), 712–733. https://doi.org/
10.2110/jsr.2016.45
Spikings, R., Cochrane, R., Villagómez, D., Van der Lelij, R., Vallejo, C., Winkler, W., et al. (2015). The geological history of northwestern
South America: From Pangaea to the early collision of the Caribbean large Igneous Province (290–75 Ma). Gondwana Research, 27,
95–139. https://doi.org/10.1016/j.gr.2014.06.004
Spikings, R. A., Crowhurst, P. V., Winkler, W., & Villagomez, D. (2010). Syn‐ and post accretionary cooling history of the Ecuadorian Andes
constrained by their in‐situ and detrital thermochronometric record. Journal of South American Earth Sciences, 30(3‐4), 121–133.
https://doi.org/10.1016/j.jsames.2010.04.002
Stampfli, G. M., Hochard, C., Vérard, C., Wilhem, C., & von Raumer, J. (2013). The formation of Pangea. Tectonophysics, 593, 1–19.
https://doi.org/10.1016/j.tecto.2013.02.037

ANDJIĆ ET AL. 4424


Tectonics 10.1029/2019TC005741

Stern, R. J., & Scholl, D. W. (2010). Yin and yang of continental crust creation and destruction by plate tectonic processes. International
Geology Review, 52(1), 1–31. https://doi.org/10.1080/00206810903332322
Sundblad, K., Cumming, G., & Krstic, D. (1991). Lead isotope evidence for the formation of epithermal gold quartz veins in the Chortis
block, Nicaragua. Economic Geology and the Bulletin of the Society of Economic Geologists, 86(5), 944–959. https://doi.org/10.2113/
gsecongeo.86.5.944
Talavera‐Mendoza, O., Ruiz, J., Gehrels, G., Valencia, V., & Centeno‐García, E. (2007). Detrital zircon U/Pb geochronology of southern
Guerrero and western Mixteca arc successions (southern Mexico): New insights for the tectonic evolution of southwestern North
America during the late Mesozoic. Geological Society of America Bulletin, 119(9‐10), 1052–1065. https://doi.org/10.1130/B26016.1
Taylor, B. (1992). Rifting and the volcanic‐tectonic evolution of the Izu‐Bonin‐Mariana arc. In B. Taylor, K. Fujioka, et al. (Eds.),
Proceedings of the Ocean Drilling Program, Scientific Results (Vol. 126, pp. 627–665). College Station, TX: Ocean Drilling Program.
Tetreault, J. L., & Buiter, S. (2012). Geodynamic models of terrane accretion: Testing the fate of island arcs, oceanic plateaus, and conti-
nental fragments in subduction zones. Journal of Geophysical Research, 117(B8), B08403. https://doi.org/10.1029/2012JB009316
Tetreault, J. L., & Buiter, S. J. H. (2014). Future accreted terranes: A compilation of island arcs, oceanic plateaus, submarine ridges, sea-
mounts, and continental fragments. Solid Earth, 5(2), 1243–1275. https://doi.org/10.5194/se‐5‐1243‐2014
Torres‐de León, R., Solari, L. A., Ortega‐Gutiérrez, F., & Martens, U. (2012). The Chortis Block southwestern México connections.
American Journal of Science, 312(3), 288–313. https://doi.org/10.2475/03.2012.02
Torró, L., Proenza, J. A., Marchesi, C., García‐Casco, A., & Lewis, J. F. (2017). Petrogenesis of meta‐volcanic rocks from the Maimón
Formation (Dominican Republic): Geochemical record of the nascent Greater Antilles paleo‐arc. Lithos, 278–281, 255–273.
Umhoefer, P. J. (2003). A model for the North America Cordillera in the Early Cretaceous: Tectonic escape related to arc collision of the
Guerrero terrane and a change in North America plate motion. In S. E. Johnson, S. R. Paterson, J. M. Fletcher, G. H. Girty, D. L.
Kimbrough, & A. Martín‐Barajas (Eds.), Tectonic evolution of northwestern México and the southwestern USA, Geological Society of
America Special Paper (Vol. 374, pp. 117–134). Boulder, Colorado: The Geological Society of America.
Vaughan, T. W. (1932). The foraminiferal genus Orbitolina in Guatemala and Venezuela. Proceedings of the National Academy of Sciences,
18(10), 609–610. https://doi.org/10.1073/pnas.18.10.609
Venable, M. E. (1994). A geologic, tectonic, and metallogenic evaluation of Siuna terrane (154 p.). Ph.D. thesis. Tucson, Arizona: University of
Arizona.
Viland, J., Henry, B., Calix, R., & Diaz, C. (1996). Late Jurassic deformation in Honduras: Proposals for a revised regional stratigraphy.
Journal of South American Earth Sciences, 9(3‐4), 153–160. https://doi.org/10.1016/0895‐9811(96)00002‐8
Yamamoto, S., Senshu, H., Rino, S., Omori, S., & Maruyama, S. (2009). Granite subduction: Arc subduction, tectonic erosion and sediment
subduction. Gondwana Research, 15(3‐4), 443–453. https://doi.org/10.1016/j.gr.2008.12.009
Zagorevski, A., van Staal, C. R., McNicoll, V., Rogers, N., & Valverde‐Vaquero, P. (2008). Tectonic architecture of an arc‐arc collision zone,
Newfoundland Appalachians. In A. E. Draut, P. D. Clift, & D. W. Scholl (Eds.), Formation and applications of the sedimentary record in
arc collision zones, Geological Society of America Special Paper (Vol. 436, pp. 309–333). Boulder, Colorado: The Geological Society of
America. doi: https://doi.org/10.1130/2008.2436(14)
Zhang, Q., Guo, F., Zhao, L., & Wu, Y. (2017). Geodynamics of divergent double subduction: 3‐D numerical modeling of a Cenozoic
example in the Molucca Sea region, Indonesia. Journal of Geophysical Research: Solid Earth, 122, 3977–3998. https://doi.org/10.1002/
2017JB013991

ANDJIĆ ET AL. 4425

You might also like