Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

RESEARCH ARTICLE Effects of Tectonic Quiescence Between Orogeny and Rifting

10.1029/2022TC007492
Claudio Alejandro Salazar-Mora1 and Victor Sacek2
Key Points:
Instituto de Geociências. Universidade de São Paulo, São Paulo, Brazil, 2Instituto de Astronomia, Geofísica e Ciências
1
• T ectonic quiescence duration controls
Atmosféricas, Universidade de São Paulo, Rua do Matão, São Paulo, Brazil
the pre-rift lithospheric thermal
structure
• Tectonic quiescence duration and
orogen size control rifted margins Abstract Tectonic quiescence is a period of nearly null relative plate motion. During the Atlantic Ocean
architecture opening, rifting cut across continental segments of supercontinent Pangea that had been through different
• Preservation of crust in the fossil
subduction zone after rifting is
periods of quiescence, which varied from tens to hundreds of Myr. Using numerical models, we explored the
controlled by quiescence time effects of quiescence between orogeny and rifting on the pre-rift lithosphere and subsequent conjugate rifted
margin configuration. Our model results showed that quiescence of ∼30–60 Myr caused wide orogenic wedges
Supporting Information:
(>300 km) to be heated beneath their centers, which led to the breakup of the lithospheric mantle before the
Supporting Information may be found in breakup of the continental crust during rifting, resulting in hyperextended conjugate rifted margins. In these
the online version of this article. scenarios, continental lithosphere breakup away from the suture contributed to the preservation of previously
subducted lower crust along the subduction zone. Long periods of tectonic quiescence (∼100 Myr) experienced
Correspondence to: by a wide orogen permitted efficient orogenic collapse and lateral spreading, with pronounced lithospheric
C. A. Salazar-Mora, thermal weakening. Consequently, post-quiescence rifting reactivated the suture and effectively drove plate
claudio.mora@usp.br eduction, exhuming most of the subducted continental lower crust and mantle lithosphere. Ultimately, ultra-
wide domains developed into asymmetric conjugate margins. Contrastingly, in narrow orogenic wedges
Citation: (<200 km), different quiescence durations had minor effects on the margin architecture. In these cases, a
Salazar-Mora, C. A., & Sacek, V. longer quiescence time contributed to a larger preservation of subducted crust in the fossil subduction zone
(2023). Effects of tectonic quiescence
between orogeny and rifting. Tectonics, after rifting. Our models agree with conjugate rifted margins in the North and South Atlantic, which probably
42, e2022TC007492. https://doi. developed after the collapse of wide orogens and subsequent periods of tectonic quiescence.
org/10.1029/2022TC007492

Received 6 JUL 2022 1. Introduction


Accepted 9 JAN 2023
The parallelism between old orogenic belts and young rift systems emphasizes the importance of orogenic inher-
itance within the continental lithosphere during the break-up of continents (Vauchez et al., 1997). The pre-rift
continental lithosphere is largely heterogeneous and comprises at least three types of inheritances, namely, ther-
mal, compositional, and structural (Manatschal et al., 2015). The relative importance of these types of inher-
itances on rifted margin formation was explored through numerical models (e.g., Gouiza & Naliboff, 2021;
Peron-Pinvidic et al., 2022; Petersen & Schiffer, 2016; Salazar-Mora et al., 2018), but the role of the time interval
of tectonic quiescence between the orogenic and the rifting phase is still unexplored. In fact, the opening of the
Atlantic Ocean in its North, Central, and South segments occurred with different periods of time between the
amalgamation of Gondwana-Pangea and the onset of Pangea rifting (Figure 1).

In the North Atlantic, particularly between Greenland and Norway, the closure of the Iapetus Ocean during the
convergence of Baltica and Laurentia culminated in the final continent-continent collision of the Scandinavian
Caledonides. This collisional event is referred to as the Scandian Phase, and the syn-collisional magmatism
associated with it started at ∼438 Ma and lasted until ∼415 Ma (Slagstad & Kirkland, 2018). High-to ultra-
high pressure metamorphism at 420-415 Ma also suggests that collision persisted until ∼415 Ma. The onset of
orogenic collapse is suggested to have started at ∼405 Ma, based on the reversal of tectonic vergence in the basal
Caledonian detachment in southern Norway (Fossen, 2000). The transition from collapse-related extension to
that associated with rifting is still somewhat debatable. Peron-Pinvidic and Osmundsen (2020) and references
therein show that a period of 30–40 Myr must have passed between the end of orogenic collapse to the onset of
stretching during early rifting associated with plate divergence. In the North Atlantic Caledonides, extensional
activity is recorded during Devonian basin filling, between 385 and 398 Ma (Fossen, 2010), which was followed
by ultrahigh-pressure metamorphism in the Greenland Caledonides at 365–350 Ma that, in turn, involves plate
© 2023. American Geophysical Union. convergence, therefore requiring the transition to post-orogenic extension dominated by high-angle faults to be
All Rights Reserved. younger than 345 Ma (Gilotti & McClelland, 2008).

SALAZAR-MORA AND SACEK 1 of 19


Tectonics 10.1029/2022TC007492

Figure 1. (a) Chart showing the tectonic episodes of orogeny, tectonic quiescence, onset of extension, and continental break-up of the South, Central, and North
Atlantic rifted margins and their ancient orogenic and post-orogenic history. (b) Lower Carboniferous paleogeographic reconstruction of the continents, showing
the Pan-African, Appalachians, and Caledonide orogens. (c) Upper Triassic paleogeographic reconstruction of Pangea showing Laurentia, Baltica, and Gondwana.
Dashed lines show the locus and time of continental breakup. Figures b and c were adapted from Scotese PALEOMAP Project (https://www.earthbyte.org/paleodem-
resource-scotese-and-wright-2018/). 1(Caxito et al., 2022), 2(Nürnberg & Müller, 1991), 3(Moulin et al., 2010), 4(Hatcher et al., 2007), 5(Ma et al., 2019), 6(Slagstad &
Kirkland, 2018), 7(Gilotti & McClelland, 2008), 8(Peron-Pinvidic & Osmundsen, 2020):

The rifted margins of the Central Atlantic developed on a continental lithosphere that had been previously amal-
gamated in the final stages of the Appalachian Orogen development (Figure 1), namely the Alleghanian collision.
The magmatism associated with the various stages of this collision span between 300 and 265 Ma, with the main
uplift of the mountain chain ∼325–320 Ma (Hatcher et al., 2007). The same authors understood that the time
between the end of the collision (∼265 Ma) and the initial stages of the Atlantic Ocean opening in the region
(∼215 Ma) is about 50 Myr. Nevertheless, structural and cooling ages data suggest that the extensional collapse of
the Alleghanian orogen occurred with extensional detachment faulting during a period between 250 and 237 Ma
(Ma et al., 2019). If we assume that this extension during the orogenic collapse took place during the conver-
gence, then the tectonic quiescence period would have started at 237 Ma, which means it lasted for ∼22 Myr until
the onset of Pangea breakup related extension at 215 Ma (Figure 1).

Toward the South Atlantic rifted margins, the evolutionary tectonic history is significantly older than the cases
shown above. The South Atlantic Brasiliano Orogenic System (SABOS) of Western Gondwana had its syn-to the
late-collisional course in the late Ediacaran to early Cambrian, which was diachronous throughout the orogenic
system (Caxito et al., 2022). While in the northern part (Araçuaí orogen) the collisional phase is between 580
and 540 Ma, in the southern portion it began rather earlier, at 620 Ma. The orogenic collapse is better constrained

SALAZAR-MORA AND SACEK 2 of 19


Tectonics 10.1029/2022TC007492

in the SABOS northern portion, between 530 and 490 Ma and with main structural lineaments reactivation
∼500-490 Ma (Santiago et al., 2020). The southern segment of the SABOS was dominated by wrench tectonics in
the post-orogenic phase, with strike-slip ranging broadly from 625 to 530 Ma (Caxito et al., 2022, and references
therein). The onset of extension between South America and Africa is between 150 and 130 Ma (Nürnberg &
Müller, 1991), which entails ∼340 Myr of tectonic quiescence between the end of the orogenic history and onset
of extension.
Tectonic quiescence represents the period of time when any rate of relative motion between two continental plates
is practically null. The transition from orogeny to rifting is rather complex, particularly because of the phase of
orogenic extensional collapse, driven by the sharp elevation gradients (Dewey, 1988) between the mountain and
its forelands. In the case that the convergence rate is significant and denudation is limited, then collapse-related
extension can be initiated during the late orogenic phase, still under boundary contractional forces. It is expected
that collapse-related extension takes course until the crust returns to normal thickness under contractional
boundary forces, and if tectonic quiescence comprises a short period, then collapse-related extension is easily
continued by a general rifting-related extension (Dewey, 1988). Conversely, if there is a balance between conver-
gence rate and denudation, the orogen will attain large gravitational potential by the end of crustal thickening,
so collapse-related extension will commence nearly synchronous to tectonic quiescence. This period of time
between the end of orogenic collapse and the onset of rifting (which here we call tectonic quiescence) is also
referred to as the”reactivation phase” (Peron-Pinvidic & Osmundsen, 2020). In Figure 1a we present approximate
periods of tectonic quiescence considering the end of the extensional collapse phase of each example, but they
could very well be longer periods of quiescence if one considers that the onset of extensional collapse took place
during the late syn-orogenic phase.
In this study, we present geodynamic thermo-mechanical numerical models that aimed to explore the role of
tectonic quiescence between a contractional tectonic phase with collisional orogen development and an exten-
sional phase of rifting that evolved into conjugate rifted margins. A recent study (Peron-Pinvidic et al., 2022) has
explored collision inheritance and long-term thermal weakening by modeling contraction followed by extension,
where a tectonic pause was tested between contraction and extension, as well as tectonic pauses during rifting,
which was then compared to the Mid-Norwegian rift system. We went further in modeling a full Wilson cycle
(i.e., extension-contraction-quiescence-extension) with the main focus on the discussion of the effects of varying
quiescence periods in the pre-rift lithospheric rheological regime and its influence in the final configuration of
the rifted margins, comparing them to examples of Atlantic-type rifted margins.

2. The Numerical Model


We used the finite element code Mandyoc (Sacek et al., 2022) to numerically solve the differential equations for
the conservation of mass, energy, and momentum for an incompressible non-Newtonian fluid (Zhong et al., 2007):

(1)
𝑢𝑢𝑖𝑖𝑖𝑖𝑖 = 0

(2)
𝜎𝜎𝑖𝑖𝑖𝑖𝑖𝑖𝑖 + 𝑔𝑔𝑖𝑖 𝜌𝜌 = 0

𝜕𝜕𝜕𝜕 𝐻𝐻 𝑇𝑇
(3) + 𝑢𝑢𝑖𝑖 𝑇𝑇,𝑖𝑖 = 𝜅𝜅𝜅𝜅,𝑖𝑖𝑖𝑖 + + 𝑢𝑢𝑖𝑖 𝑔𝑔𝑖𝑖 𝛼𝛼
𝜕𝜕𝜕𝜕 𝑐𝑐𝑝𝑝 𝑐𝑐𝑝𝑝

where

(4)
𝜎𝜎𝑖𝑖𝑖𝑖 = −𝑃𝑃 𝑃𝑃𝑖𝑖𝑖𝑖 + 𝜂𝜂(𝑢𝑢𝑖𝑖𝑖𝑖𝑖 + 𝑢𝑢𝑗𝑗𝑗𝑗𝑗 ),

(5)
𝜌𝜌 = 𝜌𝜌0 [1 − 𝛼𝛼(𝑇𝑇 − 𝑇𝑇0 )],

t is time, T is temperature, ui is the ith component of the velocity field, g is gravity, ρ0 is the reference rock density
at temperature T0 = 0°C, α is the coefficient of thermal expansion, κ is the thermal diffusivity, H heat production
per unit of mass, cp is specific heat capacity, P is the total pressure, η is the effective dynamic viscosity and δij is
the Kronecker delta. Repeated indexes mean summation and the indexes after the comma represent the partial
derivative of the respective coordinate.

In our thermomechanical model, the different layers of the interior of the Earth are simulated by the visco-plastic
rheology, where the effective viscosity η is a function of the non-linear power-law viscous rheology and a plastic

SALAZAR-MORA AND SACEK 3 of 19


Tectonics 10.1029/2022TC007492

Table 1
Physical Parameters of Each Layer
Parameter Air Sediments Déccolement Upper crust Lower crust Lithospheric mantle Asthenosphere
Flow law − Wet quartz 1
Wet quartz 1
Wet quartz 1
Wet quartz 1
Dry olivine 2
Wet olivine 2
n 1 4.0 4.0 4.0 4.0 3.5 3.0
C [4]
1 1 0.1 1 100 1 1
ρ0 (kg/m 3) 1 2,300 2,350 2,700 2,800 3,354 3,378
A [4] (Pa −n/s) 1.0 × 10 −18 8.574 × 10 −28 8.574 × 10 −28 8.574 × 10 −28 8.574 × 10 −10 2.4168 × 10 −15 1.393 × 10 −14
Ea (kJ/mol) 0 222 222 222 222 540 429
V (m /mol)3
0 0.0 0.0 0.0 0.0 25 × 10 −6
15 × 10 −6
H [3]
(W/kg) 0 4.63 × 10 −10
4.63 × 10 −10
4.63 × 10 −10
7.14 × 10 −11
9.0 × 10 −12
0.0
𝐴𝐴 𝐴𝐴0[4] (MPa) − 20 → 4 4 20 → 4 20 → 4 20 → 4 20 → 4
ϕ [4] − 15° → 2° 2° 15° → 2° 15° → 2° 15° → 2° 15° → 2°
Note. 1: Extracted from Gleason and Tullis (1995). 2: Extracted from Karato and Wu (1993). 3: Extracted from Hacker et al. (2015) for the upper and lower crust. 4:
Extracted from Salazar-Mora et al. (2018). The volumetric expansion coefficient α = 3.28 × 10 −5 K −1, the specific heat capacity cp = 1250 J/kg/K, and the thermal
diffusivity κ = κ0 = 10 −6 m 2/s were assumed constant in all the numerical simulation for simplicity. One exception is the thermal diffusivity at the lower portion of the
domain at temperatures above 1330°C. Between 1330 and 1340°C the thermal diffusivity linearly increases from κ0 to 50κ0. Above 1340°C, the diffusivity is constant
at 50κ0. This strategy is the same one used by Wolf et al. (2021), where the high thermal diffusivity in the asthenosphere mimics the effective convective heating in
the sublithospheric mantle, preserving the base of the thermal boundary layer close to the initial depth of the lithosphere even when the simulation takes hundreds of
millions of years.

yield criterion. The viscous component is a function of temperature, pressure, composition, and strain rate in the
rock, given by Huismans & Beaumont (2014):
1−𝑛 [ ]
(6) 𝐸 +𝑉𝑃
𝜂visc = 𝐶𝐴−1∕𝑛 𝜀̇ 𝐼𝐼𝑛 exp 𝑎
𝑛𝑅𝑇
( )1∕2
where C is a scale factor, A is the pre-exponential scale factor, n is the power law exponent,
𝐴𝐴 𝐴𝐴𝐴 𝐼𝐼𝐼𝐼 = 12 ̇𝜀𝜀′𝑖𝑖𝑖𝑖 ̇𝜀𝜀′𝑖𝑖𝑖𝑖
is the square root of the second invariant of deviatoric strain rate tensor, Ea is the activation energy, V is the
activation volume, and R is universal gas constant. The plastic yield limit follows the Drucker–Prager criterion:

𝜎𝑦𝑖𝑒𝑙𝑑 = 𝑐0 cos 𝜙 + 𝑃 sin 𝜙,


(7)

where ϕ is the internal angle of friction and c0 is the internal cohesion of the rock. ϕ and c0 vary linearly as a
function of cumulative strain ɛ, allowing the simulation of strain softening, and facilitating the deformation local-
ization (Huismans & Beaumont, 2003). Below and above the limits of strain softening presented in Table 1, c0
and ϕ are assumed constant (Salazar-Mora et al., 2018).

The combination of the plastic and viscous components results in the effective nonlinear viscosity, given by
Moresi & Solomatov (1998):
( )
𝜎𝜎𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦𝑦
(8)
( )
𝜂𝜂 = min 𝜂𝜂plast , 𝜂𝜂visc = min , 𝜂𝜂visc .
2𝜀𝜀̇ 𝐼𝐼𝐼𝐼

The topographic response due to extension is simulated by a sticky-air approach (Crameri et al., 2012) where a
layer of low density and low viscosity is inserted above the upper crust to emulate the air. We adopted Q1P0 finite
elements, which are bilinear in velocity and constant in pressure. Additional details on the numerical model are
presented in the Supporting Information S1.

3. Model Setup
The numerical domain consists of 1,600 × 300 km 2 subdivided into 800 × 150 square finite elements (Figure 2a).
Initially, the domain is composed of the asthenosphere, lithospheric mantle, lower crust, and upper crust. At the
top of the crust, two additional layers were introduced representing the preexistence of sedimentary layers: a

SALAZAR-MORA AND SACEK 4 of 19


Tectonics 10.1029/2022TC007492

Figure 2. Setup for all the numerical scenarios. (a) The geometry of the different layers, indicates the initial thermal structure
in red. The horizontal segments indicate the boundary conditions for velocity on the left and right sides of the model during
the onset of the numerical simulation, where vr represents the horizontal velocity between the two sides of the domain. (b)
Yield stress envelope for the lithosphere assuming the initial thermal structure and a strain rate of 10 −15 s −1. The continuous
dashed line represents strain-weakened lithosphere. (c) Variation of vr through time. The different numerical experiments
present a different quiescence interval of time Δtq.

1-km-thick décollement layer under a 3-km-thick sedimentary layer. The incorporation of the décollement layer
allows the simulation of thin-skinned tectonics of the sedimentary layer (Wolf et al., 2021), tectonism observed in
different modern foreland fold-thrust belts (e.g., DeCelles et al., 2001; Sommaruga, 1999). The weak décollement
layer was simulated by the decrease of the initial value of the cohesion and internal frictional angle (Table 1). To
simulate the free surface, a sticky-air layer was introduced at the top of the domain with 40 km with low viscosity
and compatible density with the atmospheric air.

The initial thermal structure is only depth-dependent (Figure 2a), with the temperature at the top of the model at
0°C and the base of the lithosphere at 1300°C, with the temperature of the lithosphere representing the solution
of the following differential equation.

𝜕𝜕 2 𝑇𝑇 (𝑧𝑧) 𝐻𝐻(𝑧𝑧)
(9)
𝜅𝜅 + =0
𝜕𝜕𝜕𝜕2 𝑐𝑐𝑝𝑝

where H(z) is the internal heat production of the different layers as presented in Table 1. The temperature in the
sublithospheric mantle is represented by an adiabatic increase until the bottom of the numerical domain:

𝑇 (𝑧) = 𝑇𝑝 exp(𝑔𝛼𝑧∕𝑐𝑝 )
(10)

where Tp = 1262°C is the potential temperature of the asthenospheric mantle.

The velocity boundary condition assumes a reference frame fixed on the lithospheric plate on the left portion of
the domain (Figure 2a), while the opposed side moves with a relative velocity vr. To preserve the conservation
of volume, we used a velocity profile on the left and right boundary, respectively vleft and vright, as follows (Silva
& Sacek, 2022):

SALAZAR-MORA AND SACEK 5 of 19


Tectonics 10.1029/2022TC007492

Table 2
List of Models and Varying Parameters
Model name Phase 1 (Myr) Phase 2 (Myr) Quiescence (Myr) Phase 3 (Myr) Total model time (Myr) Erosion Display
RM 25 50 0 50 125 No Figure 3, Movie S1
M15 25 50 15 75 165 No Figures 4a–4c, Movie S2
M15e 25 50 15 75 165 Yes Figures 4d–4f, Movie S3
M30 25 50 30 105 175 No Figures 5a–5c, Movie S4
M30e 25 50 30 105 175 Yes Figures 5d–5f, Movie S5
M60 25 50 60 135 205 No Movie S6
M60e 25 50 60 135 205 Yes Movie S7
M100 25 50 100 70 245 No Figures 6a–6c, Movie S8
M100e 25 50 100 70 245 Yes Figures 6d–6f, Movie S9

if 0 ≤ 𝑧 < ℎ1

⎪0,
if ℎ1 ≤ 𝑧 ≤ ℎ1 + ℎ2
𝑣𝑙𝑒𝑓 𝑡 (𝑧) = ⎨ 𝑧 − ℎ
(11) 1
⎪𝑣𝑏 ,
⎩ ℎ2

if 0 ≤ 𝑧 < ℎ1

⎪𝑣𝑟 ,
if ℎ1 ≤ 𝑧 ≤ ℎ1 + ℎ2
𝑣𝑟𝑖𝑔ℎ𝑡 (𝑧) = ⎨
(12) 𝑧 − ℎ1
⎪−𝑣𝑏 + 𝑣𝑟 ,
⎩ ℎ2

where

ℎ + ℎ2
(13)
𝑣𝑣𝑏𝑏 = 𝑣𝑣𝑟𝑟 1
ℎ2

with z as the vertical coordinate with the origin at the surface with h1 = 150 km and h2 = 110 km. These expres-
sions ensure conservation of volume, so that
ℎ1 +ℎ2 ℎ1 +ℎ2

∫ ∫
𝑣𝑣 ℎ
(14)
𝑣𝑣𝑙𝑙𝑙𝑙𝑙𝑙 𝑙𝑙 𝑑𝑑𝑑𝑑 = 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 𝑑𝑑𝑑𝑑 = 𝑏𝑏 2 .
2
0 0

In our numerical simulations, the relative velocity vr is a function of time (vr = vr(t), Figure 2c). Initially, all the
scenarios start with an extension phase that endures for Δtext1 = 25 Myr, with vr(0 ≤ t < Δtext1) = 1 cm/year. This
extension is followed by a contraction phase for Δtcomp = 50 Myr, with vr(Δtext1 ≤ t < Δtext1 + Δtcomp) = −1 cm/
year. After the contraction phase, we tested different quiescence periods Δtq, ranging from 0 to 100 Myr (see
Table 2), while the relative velocity between the plates is null (vr = 0). Each numerical model is indicated in the
study as a function of the quiescence period, where M15, M30, and M100 represent scenarios with Δtq = 15, 30,
and 100 Myr, respectively. The reference model, indicated in the text as RM, is the scenario with a null quiescence
period (Δtq = 0). After the quiescence phase, all the models underwent a second phase of extension (Figure 2c).

Additionally, we tested the impact of crustal thickness reduction due to erosive exhumation of the orogen. For
this, we performed numerical scenarios where we imposed a fixed erosion rate on top of the free surface at a
rate
𝐴𝐴 of 𝐴𝐴𝐴 = 10−4 m/year, for all points with altitude h > 200 m above the initial altitude. This simplified surface
processes model aims to compare scenarios with the same amount of crustal contraction but resulting in orogens
with different widths and crustal thicknesses (see further discussion in Supporting Information S1). The scenarios
with erosive exhumation are indicated with a letter “e”: for example, M15e represents the scenario with quies-
cence period Δtq = 15 Myr and includes erosive exhumation. All scenarios without the letter “e” do not take
surface processes into account.

In the scenarios with imposed erosion, the erosive exhumation starts at t = 50 Myr, in the middle of the contrac-
tion phase, and is preserved until the end of the numerical experiment. This delayed onset for the erosion ensures

SALAZAR-MORA AND SACEK 6 of 19


Tectonics 10.1029/2022TC007492

Figure 3. Reference model RM. (a) End of extension in Phase 1 after 25 Myr of divergence showing passive margin configuration. H refers to a crustal H-block. (b)
End of contraction in Phase 2 after 50 Myr of convergence showing the full orogenic structure. (c) End of extension in Phase 3 after 50 Myr of divergence showing the
final rifted passive margins. Note the mantle scar beneath the right rifted margin. Shades of gray indicate the magnitude of cumulative strain. Color-coding is the same
as Figure 2. See text for further details and Movie S1.

that both models, with and without surface processes, have the same geodynamic evolution until 50 Myr, facilitat-
ing the comparison of the different scenarios with the orogenic wedge at nearly the same position in the numerical
domain and the left portion of the model as the lower plate in all scenarios.

4. Modeling Results
4.1. Reference Model (RM)

During extension in Phase 1, two conjugate high angle extensional shear zones rooted in the lower continental
crust led to a nearly symmetric rift continental basin. Shortly after the establishment of the shear zones, one of
them randomly accumulated strain and connected with the continental lithospheric mantle. This lithospheric-scale
shear zone allowed for an asymmetric pattern of rifting, where the left rifted margin resulted in the crustal exten-
sion of ∼75 km, while the right conjugate rifted margin extended for 100–110 km (Figure 3a). Crustal extension
in the rifted margins is measured in our models from the furthest frictional-plastic shear zone in the proximal
domain toward the transition to oceanic lithosphere in the distal domain, which is summarized in Figure 7 b and
c. The lithospheric-scale shear zone is likely responsible for the greater amount of extension in the right margin,
which evolved into a low angle extensional shear zone bounding a crustal H-block (in the nomenclature of
Péron-Pinvidic & Manatschal, 2010) and enabled lithospheric mantle exhumation (Figure 3a).

SALAZAR-MORA AND SACEK 7 of 19


Tectonics 10.1029/2022TC007492

Figure 4. Models with 15 Myr of tectonic quiescence. Model M15 is shown from the end of tectonic quiescence (a), through early extension in Phase 3 (b) and final
rifted margin configuration (c). Panels (d), (e) and (f) show the same model evolution, now with surface processes (M15e). Shades of gray indicate the magnitude of
cumulative strain. Color-coding is the same as Figure 2. See text for further details and Movies S2 and S3.

In the following contraction Phase 2 (Figure 3b), the low angle extensional shear zone is entirely reactivated
to accommodate shortening in the more extended right margin, pushing back down the previously exhumated
lithospheric mantle and reassembling the crustal H-block, which then collided with the left margin. At this
stage, a central double-vergent pop-up structure comprising the upper crust and lithospheric mantle is uplifted
and controlled by reactivated inherited Phase 1 high angle extensional shear zones. The initial pop-structure
is carried onto the orogen retro-side and a first lithospheric-scale thrust sheet is transported in the direction of
the pro-side, after which propagating upper crustal-scale thrust sheets form in the development of the orogen,
accommodating shortening and permitting subduction of the lower continental crust along with the continental
lithospheric mantle to depths of 130 km (Figure 3b). At the contact between the lower and upper colliding plates,
the transition between pro-side crustal sheets into the retro-side is characterized by large-scale folding of the
thrust faults, which now transport crustal sheets toward the retro-side. On the surface, sediment layers are affected
by embryonic thin-skinned thrusting on both retro- and pro-sides. The resulting orogenic wedge is double vergent
and reaches more than 7 km in topography in the hinterland.

SALAZAR-MORA AND SACEK 8 of 19


Tectonics 10.1029/2022TC007492

Figure 5. Models with 30 Myr of tectonic quiescence. (a) Shows model M30 with the orogenic structure after 30 Myr of tectonic quiescence. (b) Model M30 during
the early extension in Phase 3 (13 Myr) showing lithospheric mantle break-up before continental crust break-up. (c) Shows the rifted margins configuration at final
stages of extension in Phase 3. Note that both margins were hyperextended. (d) Shows model M30e with the orogenic structure after 30 Myr of tectonic quiescence
and erosion. Note that it is colder than in (a). (e) Early extension in Phase 3 (13 Myr) showing the crust thinning before the lithospheric mantle. (f) Rifted margins
configuration of M30e at final stages of extension in Phase 3. Hyperextension is not observed. Shades of gray indicate the magnitude of cumulative strain. Color-coding
is the same as Figure 2. See text for further details and Movies S4 and S5.

As the transition between contraction Phase 2 and extension Phase 3 was immediate in the RM scenario, the
orogenic wedge remained relatively cold, which is shown by the lowered isotherms (red lines in Figure 3b). When
extension started, strain preferentially accumulated in the lithospheric-scale suture, which at first, connected with
the closer crustal-scale pro-side crustal sheets. Shortly after, the furthest pro-side crustal sheet is also reactivated
to accommodate the extension, which is followed by the reactivation of inherited contractional structures in the
retro-side. As the extension proceeded, the thickened orogenic crust in the pro-side is thinned with synchronous
eduction of the previously subducted lower continental crust and lithospheric mantle. During this thinning stage,
all the inherited contractional structures are no longer reactivated, and newly formed extensional shear zones
accommodated the extension of the pro-side. These new structures, in turn, are left unreactivate as their offset
with the suture or the necking zone increases (e.g., Salazar-Mora et al., 2018). Continental break-up is achieved
after nearly all the previously subducted lower crust and lithospheric mantle are educted, leaving a subduction
scar in the lithospheric mantle (Figure 3c).

SALAZAR-MORA AND SACEK 9 of 19


Tectonics 10.1029/2022TC007492

Figure 6. Models with 100 Myr of tectonic quiescence. (a) Shows the orogenic structure of model M100 at the end of contraction in Phase 2. (b) Shows the collapsed
orogenic structure after 100 Myr of tectonic quiescence. Note how the orogen has laterally spread in comparison to (a). (c) Shows the final rifted margins configuration
after full extension in Phase 3. Panels (d), (e) and (f) show the same spatio-temporal evolution, but for model M100e (with erosion). See text for further details and
Movies S8 and S9. Shades of gray indicate the magnitude of cumulative strain. Color-coding is the same as Figure 2.

The left margin, which developed mainly due to the extension of the orogenic pro-side, accommodated crustal
extension over ca. Forty0 km, preserving a relatively thick orogenic crust in the proximal domain. The necking
domain of this margin is characterized by the upward inflection of the Moho to converge with the top basement
and thoroughly reactivated contractional structures now connecting in a low-angle style at the base of the upper
crust. Within the distal domain, the highly strained and ductile continental crust is progressively thinned, at the
same time as educted lower continental crust occupies the base of the crust. Toward the ocean, the distal domain
comprises a block of the lower crust and mantle lithosphere inherited from previous phases 1 and 3 and an upper
crust H-block, which is bounded by lower continental crust toward the continent and by exhumed lithospheric
mantle toward the ocean.
The right conjugate rifted margin accommodated an extension over 150 km. In its proximal domain, a rela-
tively thick orogenic crust is preserved as well as a few reactivated contractional structures. The relative upward
inflection of the Moho is inherited from contraction Phase 2, not reflecting real thinning during rifting. This,
in turn, occurred in a very narrow necking domain (Figure 3c) with highly strained upper continental crust
and removal of lower continental crust. The distal domain comprises an H-block and transitions of lithospheric

SALAZAR-MORA AND SACEK 10 of 19


Tectonics 10.1029/2022TC007492

Figure 7. (a) Fraction of the subducted crust preserved along the fossil subduction zone as a function of the quiescence time
for each numerical scenario. The blue and red stripes indicate the tendency observed in the numerical scenarios, depending
on the orogenic style. (b) Conjugate margin widths derived from wide orogens (no erosion) as a function of the quiescence
time. Note that M30 and M60 are the most symmetrical margins. (c) Conjugate margin widths derived from narrow orogens
(with erosion) as a function of the quiescence time. Margins that show hyperextension are denoted with h and those that had
unreactivated suture during rifting are denoted with us.

mantle exhumation at the boundary with the oceanic lithosphere. Both of the margins are not hyperextended in
the sense of Unternehr et al. (2010), who characterized hyperextended margins as those less than 10 km thick
over a distance of more than 100 km.

4.2. Models With 15 Myr of Tectonic Quiescence (M15 and M15e)

Models M15 and M15e are characterized by a first extension Phase 1 and a contraction Phase 2, as in RM. After
the latter phase ceased (at 75 Myr), a quiescence phase of 15 Myr took place before the final extension Phase 3
commenced. Model M15 evolved during quiescence without surface processes, while M15e started the quies-
cence Phase 25 Myr after surface processes had begun. By the time quiescence ended (Figures 4a and 4d), the
orogenic wedge was rather hotter than the orogenic wedge in RM (Figure 3b), which is observed by the elevated
isotherms of 500 and 750°C. During quiescence, the uneroded orogenic wedge underwent orogenic collapse with
incipient lateral spreading and a decrease in topography from 7 to 5 km. On the other hand, the orogenic wedge
with surface processes rapidly attained 4 km, showing no sign of lateral spreading. The main effect of the efficient
surface processes is observed in the resulting thinner orogenic upper continental crust as well as a narrower and
less structured orogenic wedge (Figures 4a and 4d).

SALAZAR-MORA AND SACEK 11 of 19


Tectonics 10.1029/2022TC007492

This difference in crustal thickness directly affected the style of strain accumulation when the final Phase 3 of
extension began. During the first 10 Myr of extension in Phase 3, associated with the lower plate eduction, the
uneroded orogenic (Figure 4b) crust efficiently distributed strain from the suture into the crustal-scale contrac-
tional inherited structures, which lead to their reactivation in order to accommodate the extension. Differently, the
eroded orogenic crust was not successful in distributing strain throughout crustal inheritances and rather localized
strain in more ductile, highly strained portions of the crust devoid of frictional-plastic shear zones (Figure 4e).
This pattern of strain distribution strongly affected the final rifted margin configuration. Because the uneroded
orogenic pro-side efficiently localized and distributed strain during extension, a wider rifted margin was devel-
oped (extension accommodated over 425 km), in comparison to the model with erosion.
The left margin of M15 shows (Figure 4c) a proximal domain with preserved thick orogenic continental crust and
partially reactivated contractional shear zones, which do not extend toward the base of the upper crust. The neck-
ing domain is characterized by initial upward Moho inflection and reactivation of contractional shear zones now
connecting at the upper/lower crust transition. Toward the distal domain, the continental crust was progressively
thinned and stretched until it attained nearly 10 km of thickness oceanwards. The distal domain is characterized
by highly strained domains of the upper continental crust (darker shades) at the transition with the necking
domain as well as relatively strong blocks in the middle of the distal domain. Most of the lower continental crust
of this domain was emplaced after having been educted. The result of extension Phase 3 in model M15e shows
a left margin that accommodated extension over 385 km (Figure 4f). Its proximal domain preserved a relatively
thick orogenic crust with unreactivated contractional structures. The necking domain is narrower than in M15 and
its transition to the distal domain is rather abrupt. A large segment of highly strained and thinned upper continen-
tal crust and lower crust comprises the first half of the distal domain. The other half, oceanwards, contains highly
thinned upper continental crust and educted lower continental crust.
Continental break-up occurred after eduction of the previously underthrust continental lithosphere, which resulted
in a mantle scar beneath the right conjugate margin in both models. These mantle scars preserved relatively
lower continental crust than the scar observed in RM, although, as aforementioned, tracts of lower continental
crust were educted and emplaced in the right wide margins. The right conjugate margin of M15 accommodated
extension over 180–200 km, developing very narrow areas of highly thinned (<10 km) continental crust in the
distal domain and exhumation of lithospheric mantle toward the transition to the oceanic lithosphere. On the
other hand, M15e accommodated extension in its right margin over <100 km. H-blocks of both upper and lower
continental crust are formed in these margins.

4.3. Models With 30 Myr of Tectonic Quiescence (M30 and M30e)

By the end of contractional Phase 2, the orogenic structure of M30 and M30e is the same as the RM (Figure 3b).
With the following 30 Myr of tectonic quiescence, both orogenic wedges have their isotherms elevated due to
radiogenic heat production, although the uneroded orogenic wedge (Figure 5a) is rather hotter than the eroded
one (Figure 5d). In terms of orogenic collapse, the decrease in topography from 7 to 5 km is also observed in the
model without erosion with incipient lateral spreading, whereas topography decrease in the model with erosion
reached 3.5 km.
Similar to the models with 15 Myr of quiescence, when the final extension Phase 3 began, the uneroded orogenic
wedge first localized strain in the suture and then distributed to crustal inherited structures, while the eroded
orogenic wedge was not completely successful in distributing strain in discrete inheritances. During the eduction
of the underthrust continental lithosphere, in model M30, an asthenospheric upwelling causes lithospheric mantle
thermal necking, which leads to its break-up before the upper continental crust did so (Figure 5b). The mantle
necking was localized where the lithosphere was thickened and hot during the contraction Phase 2 and follow-
ing tectonic quiescence. This process of necking was also observed in the models with 15 Myr of quiescence,
but the thermal weakening of the lithospheric mantle was not sufficient to cause its breakup in the first attempt
(Figure 4e), so the necking migrated toward the orogenic suture, then allowing for the continental breakup (see
Movie S2).
With continued rifting, and with the lithospheric mantle already disrupted, the extension of M30 focused on
the upper continental crust, causing it to hyperextend until its final breakup (Figure 5c). In this case, the left
margin accommodated extension over ∼380 km, with wide (≥100 km) hyperextended crust portions toward the
margin distal domain. The right margin is significantly wider than those developed in RM and M15, accom-
modating extension over ∼420 km and also developing hyperextended margins toward the ocean. Due to the

SALAZAR-MORA AND SACEK 12 of 19


Tectonics 10.1029/2022TC007492

early lithospheric mantle necking, the final continental breakup did not take place near or at the orogenic suture.
Instead, the suture reactivation was abandoned and all the previously subducted lower continental crust remained
preserved there after the breakup. As a result, the distal domain of both left and right margins is completely
devoid of lower continental crust.
The evolution of model M30e during extension in Phase 3 was very similar to model M15e, where a thinner
crust did not allow for an efficient distribution of strain in inherited structures, causing strong crustal thinning
in highly ductile domains and also accommodating educted lower continental crust in the oceanward part of the
left margin distal domain (Figure 5f). The final configuration is a left margin that accommodated extension over
∼300 km and a right margin that accommodated over ∼50 km.

4.4. Models With 100 Myr of Tectonic Quiescence (M100 and M100e)

One of the striking differences between the models with 100 Myr of quiescence and the previous ones is the
orogenic collapse. During the first ∼5 Myr of tectonic quiescence, extensional collapse decreased the topography
from 7 to 5 km. After ∼30 Myr, the orogenic wedge was thermally weakened due to the heat producing elements
content, which allowed for the efficient ductile flow of the upper continental crust at the base of the orogenic
hinterland, permitting a large lateral spread (Figure 6b). As a result, by the end of quiescence, the orogen is much
wider and lower than the orogen before long-lived quiescence (Figure 6a). The collapsed uneroded orogenic
wedge decreased from 7 to 3 km in topography, while the collapsed eroded orogenic wedge (Figure 6e) decreased
to 2 km.
The orogenic lateral spreading of the models during quiescence developed a broader domain of thermally weak-
ened continental lithosphere by the end of quiescence (compare isotherms in Figures 5a and 6b). Upon extension
in Phase 3, this largely collapsed orogen developed an ultra-wide rifted margin on the left side, and a relatively
narrow one on the right side (Figure 6c). The left margin accommodated an extension of over ∼700 km, whereas
the right one accommodated over ∼100 km.
As shown in Figure 6c, the ultra-wide left margin preserved reactivated orogenic crustal shear zones in the
proximal domain, where the crust and lithospheric mantle still retain normal thicknesses. The transition from
thickened to thinner crust characterizes the necking zone. Rifting migrated oceanwards and the first lithospheric
mantle thermal necking (x = 600 km in Figure 6c) failed. This first mantle necking attempt allowed for associated
crustal thinning, preserved in the inner portion of the distal domain. Eduction was still efficient, and most of the
previously subducted lithospheric mantle and lower continental crust are educted and emplaced at the base of the
hyperextended portion of the left margin distal domain. After most of the previously subducted lower continental
crust is exhumed, the continental break-up was possible.
The model with surface processes did not show orogenic lateral spreading during and by the end of collapse
(Figure 6e), once the eroded orogenic wedge was thinner and colder (Figure 6d) than the uneroded orogenic
wedge (Figure 6a). Thus, upon extension in Phase 3, the thinning of the crust and mantle lithosphere occurred in
a narrower and colder region than in the uneroded wedge. During the development of the left margin, strain distri-
bution on the crustal-scale structures was limited, so the lithospheric mantle thermal necking actively and rapidly
forced the breakup of the continental crust before the mantle lithosphere. This, in turn, permitted the preservation
of the lower continental crust within the subduction scar. The resulting left margin accommodated an extension
of over ∼290 km, while the right margin was over <100 km (Figure 6f).
Figures 7b and 7c summarize the widths of the final conjugate rifted margins for all models, as well as some
features such as hyperextension and reactivation or not of the suture.

5. Discussion
5.1. Influence of the Quiescence Period on the Rifting Style: A Physical Explanation

In our numerical experiments, during the contraction phase, heat transport in the lithospheric mantle is mainly
guided by advection. This can be verified when we calculate the equivalent Péclet number PeL for the lithospheric
mantle:
𝐿𝐿𝐿𝐿𝑧𝑧
(15)
𝑃𝑃 𝑃𝑃𝐿𝐿 =
𝜅𝜅

SALAZAR-MORA AND SACEK 13 of 19


Tectonics 10.1029/2022TC007492

where L is the thickness of the lithospheric mantle and vz is the vertical component of the subduction velocity,
given by vz = v sin θs, where v is the relative velocity between the two plates and θs is the subduction angle. With
L = 100 km, v = 1 cm/year, and θs = 15°, the Péclet number is PeL ≈ 16. This high Péclet number implies that
the thermal structure of the subducting lithospheric mantle is not significantly modified by conduction during
continental convergence, preserving its rheological structure, and the isotherms are advected downward, follow-
ing the movement of subduction. In this case, when extension occurs immediately after the end of contraction
(reference model RM), the integrity of the lithosphere allows the restoration of the subducted plate close to the
surface without internal rupture (Figure 3c).

However, when the quiescence period is not null (Δtq > 0), the thermal conduction dominates the heat transport
during this interval of time, and the Péclet number is close to 0 (PeL → 0). In this quiescence stage, the subducted
plate is progressively heated by conduction and the isotherms are displaced upward (e.g., Figure 4a). This effect
is more expressive under the orogenic wedge, because the thickened crust concentrates a column of rock with
more radiogenic heat production, heating the lower crust and the underlying lithospheric mantle. For a quiescence
period of Δtq = 15 Myr, we observed that the decrease in rigidity due to vertical conductive heating occurs mainly
under the orogen. As a consequence, if this quiescence period is followed by extension, the necking of the litho-
spheric mantle occurs approximately under the orogenic wedge (Figure 4b), and a large fragment of the subducted
lithospheric mantle and lower crust is preserved in the orogenic suture, likely representing a fossil subducting
zone (Figure 4c). The weakening of the lithosphere under the orogenic wedge is more expressive in the scenarios
with a larger quiescence period (Δtq = 30 Myr, Figure 5), with warmer lithosphere under the orogen created by
a protracted quiescence period. In these scenarios, the necking under the orogen is more effective (Figure 5b)
and consequently, a larger fraction of the subducted lithosphere is preserved in depth after subsequent extension
(Figures 5c and 7a). In these scenarios with a relatively short quiescence period (Δtq ≤ 30 Myr), the erosion of the
orogen reduces the thickness of the main radiogenic heat production layer (Figures 4 and 5d). As a consequence,
the lithosphere under the orogen is colder than the equivalent scenarios without erosion. Therefore, the necking is
not efficient under the orogen during the final extensional phase (Figures 4e and 5e) and, consequently, a larger
fraction of the subducted plate is restored close to the surface (Figures 4f, 5f and 7a).

For larger quiescence periods (Δtq > 60 Myr), the preservation of subducted fragments of the lower crust is sensi-
tive to the orogen width and its crustal thickness, which in our numerical scenarios are dependent on the influence
of surface processes. In scenarios without erosion, the orogen evolved to a broad plateau with more than 300 km
in width (Figure 6a). In these scenarios with protracted quiescence time (Δtq > 60 Myr, Figure 7a), the horizontal
thermal gradient decreases with time and the thermal structure of the lithosphere is hot and laterally more homo-
geneous at the end of the quiescence period (Figure 6b). Consequently, during the extension in Phase 3, the litho-
spheric stretching is more diffuse in the subducting plate (Figure 6c). This occurs because the effective viscosity
of the entire slab in subduction is relatively low due to the heating induced by the thermal conduction, resulting
in a protracted rifting phase (>50 Myr). In this scenario, we observed that the ductile flow during lithospheric
eduction restored to the surface almost the entire slab previously subducted, preserving only a small fragment
of the subducting plate under the conjugate margin. However, in scenarios with efficient denudation, the orogen
is narrow and the crustal thickening is limited to a region with a width of less than 200 km (Figure 6d). In this
case, the thermal weakening of the lithosphere occurs localized under the orogen, favored by the increase in the
quiescence time (Figure 6e). In the extension of Phase 3, the necking of the subducted plate occurs essentially
under the orogen, resulting in the preservation of subducted lower crust under the conjugate margin (Figure 6f).

Therefore, we propose that, in scenarios with narrow orogens (<200 km), the preservation of fossil subduct-
ing crust increases with the quiescence time (Figure 7a). On the other hand, in scenarios with broad orogens
(300–400 km), the preservation of subducted lower crust is maximized when the quiescence time is of the order
of a few tens of millions of years. In these scenarios (e.g., M15, M30), the time interval of quiescence is suffi-
cient to heat mainly the lithospheric mantle under the orogenic wedge, turning this segment less rigid than the
surrounding portions of the plate.

5.2. Comparisons, Discussion, and Future Work

The models presented in this paper show very similar features in the crustal and lithospheric structure of observed
conjugate rifted margins. We chose two natural systems that evolved from a tectonic orogenic phase, through
a period of tectonic quiescence and then proceeded to conjugate rifted margin formation to draw first-order

SALAZAR-MORA AND SACEK 14 of 19


Tectonics 10.1029/2022TC007492

Figure 8. (a) Shows an example of the crustal structure of conjugate rifted margins in the North Atlantic, between Greenland and Scotland. The crustal structure of the
Greenland margin, Irminger Basin, Icelandic Basin, and Hatton Bank and Basin was drawn after seismic profiles in White and Smith (2009). The Scotland margin and
lithospheric mantle reflections (Flannan and W) were drawn after profiles in Flack and Warner (1990). (b) Shows model M30 for a comparison to the crustal structure
in (a). (c) Example of the crustal structure of a conjugate rifted margin pair in the South Atlantic, between Brazil and Africa. The crustal structure presented here was
drawn after seismic profiles in Araujo et al. (2022). (d) Shows model M100 for a comparison to the crustal structure in (c). See text for details.

similarities. The first example is portrayed in the North Atlantic rifted margins, between East Greenland and
Northwest Scotland. The time span between the end of orogenic collapse and initiation of rifting (i.e., tectonic
quiescence) in the North Atlantic is 30–40 Myr (Peron-Pinvidic & Osmundsen, 2020), as it is shown in Figure 1.
In Figure 8a we present the crustal structure of East Greenland through the Irminger Basin in the left rifted
margin, and the Icelandic Basin, Hatton Bank, and Basin in the right margin. These domains were drawn after
seismic profiles presented by White and Smith (2009) that were aligned at chron 22, showing a rather symmet-
ric pair of rifted margins. Toward the east, the crustal structure was drawn with the seismic profiles of Flack
and Warner (1990), where two seismic reflections appear in the lithospheric mantle, namely, Flannan and ‘W’
mantle reflections. By comparing the regional crustal structure of these conjugate margins to our model M30 in
Figure 8b, striking similarities can be observed. In the left margin, a proximal domain is characterized by a conti-
nental crust ∼30 km thick, which is then thinned in a necking domain and is finally stretched and hyperextended
toward its distal domain. The right margin is hyperextended in the distal domain (Icelandic Basin), presenting
crustal thickening inland (Hatton Bank onward). Still further inland, in the proximal domain, a good correla-
tion between the lithospheric-scale structures is observed. In this case, the Flannan reflector beneath Scotland's
continental crust matches the former subduction interface in model M30, which then underwent eduction during
the formation of the conjugate rifted margins. In the case of our model, nearly 50% of the previously subducted
continental crust did not return upon eduction (Figure 7a), so the subduction mantle scar beneath the upper plate
is enhanced (Figure 8b).

SALAZAR-MORA AND SACEK 15 of 19


Tectonics 10.1029/2022TC007492

Moving to a Central South Atlantic example, Figure 8c shows the crustal structure of the conjugate rifted
margins in east Brazil (i.e., Santos Basin) and west Africa (i.e., Benguela Basin). In this case example, the
period of tectonic quiescence is ∼300 Myr (Figure 1). These sections were drawn after the seismic profiles of
Araujo et al. (2022), which show an ultra-wide (>400 km) margin in Southeastern Brazil in contrast to a rather
narrow (∼100 Km) conjugate margin. The model that shows more similarities in terms of crustal structure is
M100, which simulated 100 Myr of tectonic quiescence. This model showed that after long-lived quiescence, the
former subduction zone was efficiently reactivated during rifting-related eduction, leaving very little amounts of
crust and lower plate related mantle lithosphere in the mantle scar beneath the narrow margin. The effectiveness
of eduction resulted in asymmetric conjugate rifted margins, where the educted lithosphere evolved into a very
wide rifted margin and the upper-plate lithosphere became a narrow margin (Figure 8d). The model also shows
a rather wide section in the left margin distal portion where extremely hyperextended (ultra-thinned) upper
continental crust is underlain by educted lower continental crust and mantle lithosphere (Figure 8d), which
is not observed in the seismic sections. What is also not observed in the seismic section, particularly beneath
the onshore Benguela Basin, is a mantle scar, as the model suggests. Once the model shows that most of what
was previously subducted returned to crustal levels upon eduction, then it is possible that this mantle scar is
not reflective enough to appear in seismic surveys. Another more likely possibility is the lack of deep seismic
surveys in this area.

Despite the agreement in overall crustal and lithospheric structure between models and natural examples, two
main limitations must be discussed, namely magmatism and erosion. Our numerical models did not consider the
role of magmatism during the orogenic phase nor during the final rifted margin formation. Nevertheless, both
natural examples that we presented in Figure 8 have an evolution that involved varying amounts of magmatism
both during orogeny and later rifting. Regarding the latter, particularly in the case of the Santos Basin, there is
still certain debate on whether the crust of the rifted margin is just (hyper-) extended crust (Araujo et al., 2022)
or if it is of magmatic origin (Karner et al., 2021). Magmatic additions in a thermally equilibrated continental
lithosphere (i.e., after tectonic quiescence) have substantial control on subsequent rifting, once mafic magmatic
bodies preclude the reactivation of crustal faults and suture zones (Chenin et al., 2019), which would likely result
in a different configuration than those shown in our models. Both natural examples shown above were compared
to the numerical models that did not take erosion into account. As it was discussed in Section 5.1, erosion
affected the orogenic structures in the way that the orogens did not evolve into thick, hot, and wide orogens, but
instead, behaved like cold, wedge-like ones. This in turn precluded the formation of symmetric hyperextended
margins (e.g., Figures 8a and 8b) or the development of asymmetric conjugates with one ultra-wide side and
a narrow counterpart (e.g., Figures 8c and 8d). We believe that the good agreement between our model results
and the crustal structures derived by seismic profiles could indicate that the pre-rift orogenic structure was not a
small cold orogen, but rather a transitional to large hot orogen (Jamieson & Beaumont, 2013). It is unlikely that
the evolution of the orogen was devoid of erosion, so the heat supply of the magmatic activity during orogeny
becomes important, even though an accurate quantification of how much heat comes from magmatic addition
or from thickening of the orogenic crust is still an open discussion, as well as erosion rates for Precambrian
orogens. Our models with no erosion might then represent a reasonable approximation of what could be a hot
pre-rift orogenic crust that led to the development of hyper-extended margins, balancing the lack of magmatic
heat supply.

In summary, our models highlight the importance of the tectonic quiescence in developing a rheological inher-
itance within the continental lithosphere of collapsed orogens that evolved into conjugate rifted margins. Our
results add to previous studies that show the importance of orogenic structural inheritance (Peron-Pinvidic
et al., 2022; Salazar-Mora et al., 2018), rheology and extension rates (Naliboff & Buiter, 2015; Tetreault &
Buiter, 2018) on rifted margin formation. Regarding the models presented by Peron-Pinvidic et al. (2022), which
followed a contraction-quiescence-extension approach, it is worth mentioning that we have achieved similar
results when 100 Myr of tectonic quiescence after the collision are considered, except that we showed that previ-
ously subducted lower continental crust was kept in the fossil subduction channel, while the upper crust moved
from the orogenic wedge to lateral spreading during the orogenic collapse. Comparing the models with no tectonic
quiescence, the difference is that our models do not show hyperextension. These differences could be attributed to
different input rheological parameters. Our contribution, which considered extension-contraction-quiescence-ex-
tension, highlights that periods of tectonic quiescence shorter than 100 Myr have different impacts on conjugate
margin symmetry and the formation of hyperextended domains (Figure 7).

SALAZAR-MORA AND SACEK 16 of 19


Tectonics 10.1029/2022TC007492

By understanding the limitations of our models, we enforce that future work should attempt to numerically imple-
ment magmatism, serpentinization, or hydration reactions and better-constrained erosion rates when considering
successive tectonic events.

6. Conclusions
In this study, we used thermo-mechanical numerical models to investigate the role of tectonic quiescence between
the end of an orogenic event and the onset of continental rifting. Periods of quiescence are usually overlooked
in both tectonic and geodynamics models that appraise successive tectonic events. Our main conclusions are the
following.
•  he duration of tectonic quiescence and orogenic size exert first-order controls on the rheological state of the
T
pre-rift continental lithosphere, which in turn influences the final conjugate rifted margins.
• With a relatively small period (≲15 Myr) of quiescence in wide orogenic wedges, thermal weakening was
inefficient. Therefore, during rifting, strain accumulated along the suture and caused continental breakup near
to it. Both previously subducted continental crust and mantle lithosphere were not completely exhumed to the
surface, and the final conjugate rifted margins configuration resulted asymmetric, with the educted plate (i.e.,
lower plate) generating a wider margin.
• During periods of ∼30–60 Myr of tectonic quiescence in wide orogenic wedges (>300 km), thermal weak-
ening was more efficient, although localized in the center of the orogen. As a result, the continental litho-
spheric mantle broke up before the continental crust, which underwent hyperextension and evolved into more
symmetric conjugate-rifted margins. Rifting in this case did not reactivate the suture efficiently, but rather
kept previously subducted lower crust and mantle lithosphere within the fossil subduction zone.
• When quiescence is long-lived (∼100 Myr), heating of the lithosphere under the wide orogenic wedge is
laterally and more efficiently distributed. As a result, a vigorously reactivated suture allowed for efficient
eduction of previously subducted continental crust and mantle lithosphere, which were emplaced within the
distal domain of the now ultra-wide rifted margin. The effectiveness of plate eduction is likely responsible for
the development of asymmetric conjugate rifted margins and ultra-wide domains.
• In the case of narrow orogens (<200 km), the duration of tectonic quiescence did not have major impacts on
the final conjugate rifted margins architecture, once all of them developed narrow asymmetric conjugates
with the wider margin being the educted plate. At the same time, increasing periods of quiescence showed
increasing amounts of the continental crust and mantle lithosphere being preserved in the fossil subduction
zone after the continental breakup.

Data Availability Statement


The numerical code (MANDYOC) used in this paper is available on GitHub repository (https://github.com/
ggciag/mandyoc). You can also find it through Sacek et al. (2022). Input files used in this paper are provided as
Supporting Information S1.

Acknowledgments References
This work was funded by grants from
Petrobras (project 2017/00461-9) and Araujo, M. N., Pérez-Gussinyé, M., & Muldashev, I. (2022). Oceanward rift migration during formation of santos–benguela ultra-wide rifted
Fundação de Amparo à Pesquisa do margins (p. 524). Geological SocietySpecial Publications. https://doi.org/10.1144/SP524-2021-123
Estado de São Paulo (FAPESP - projects Beaumont, C., Jamieson, R. A., Nguyen, M., & Lee, B. (2001). Himalayan tectonics explained by extrusion of a low-viscosity crustal channel
2017/24870-5 and 2017/10467-4). We coupled to focused surface denudation. Nature, 414(6865), 738–742.
would like to thank the constructive Caxito, F. A., Hartmann, L. A., Heilbron, M., Pedrosa-Soares, A. C., Bruno, H., Basei, M. A., & Chemale, F. (2022). Multi-proxy evidence for
comments and suggestions of John subduction of the neoproterozoic adamastor ocean and Wilson cycle tectonics in the south atlantic brasiliano orogenic system of Western
Naliboff and an anonymous reviewer that gondwana. Precambrian Research, 376, 106678. https://doi.org/10.1016/j.precamres.2022.106678
helped to improve our manuscript. We Chapman, T., Clarke, G. L., & Daczko, N. R. (2019). The role of buoyancy in the fate of ultra-high-pressure eclogite. Scientific Reports, 9(1), 1–9.
would also like to thank Rafael M. Silva Chenin, P., Jammes, S., Lavier, L. L., Manatschal, G., Picazo, S., Müntener, O., et al. (2019). Impact of mafic underplating and mantle depletion
for assistance with running benchmark on subsequent rifting: A numerical modeling study. Tectonics, 38(7), 2185–2207. https://doi.org/10.1029/2018TC005318
tests. Crameri, F., Schmeling, H., Golabek, G., Duretz, T., Orendt, R., Buiter, S., & Tackley, P. (2012). A comparison of numerical surface topography
calculations in geodynamic modelling: An evaluation of the ‘sticky air’method. Geophysical Journal International, 189(1), 38–54. https://
doi.org/10.1111/j.1365-246X.2012.05388.x
DeCelles, P. G., Robinson, D. M., Quade, J., Ojha, T., Garzione, C. N., Copeland, P., & Upreti, B. N. (2001). Stratigraphy, structure, and tectonic
evolution of the himalayan fold-thrust belt in Western Nepal. Tectonics, 20(4), 487–509.
Dewey, J. F. (1988). Extensional collapse of orogens. Tectonics, 7(6), 1123–1139. https://doi.org/10.1029/TC007i006p01123
Flack, C., & Warner, M. (1990). Three-dimensional mapping of seismic reflections from the crust and upper mantle, northwest of scotland.
Tectonophysics, 173(1), 469–481. https://doi.org/10.1016/0040-1951(90)90239-5

SALAZAR-MORA AND SACEK 17 of 19


Tectonics 10.1029/2022TC007492

Fossen, H. (2000). Extensional tectonics in the caledonides: Synorogenic or postorogenic? Tectonics, 19(2), 213–224. https://doi.org/10.1029/
1999TC900066
Fossen, H. (2010). Extensional tectonics in the North Atlantic caledonides: A regional review. In Continental Tectonics and Mountain Building:
The Legacy of Peach and Horne (Vol. 335, pp. 767–793). Geological Society of London. https://doi.org/10.1144/SP335.31
Gilotti, J. A., & McClelland, W. C. (2008). Geometry, kinematics, and timing of extensional faulting in the Greenland caledonides—A synthesis.
In The Greenland Caledonides: Evolution of the Northeast margin of Laurentia (pp. 251–271). Geological Society of America. https://doi.
org/10.1130/2008.1202(10
Gleason, G. C., & Tullis, J. (1995). A flow law for dislocation creep of quartz aggregates determined with the molten salt cell. Tectonophysics,
247(1–4), 1–23. https://doi.org/10.1016/0040-1951(95)00011-b
Gouiza, M., & Naliboff, J. (2021). Rheological inheritance controls the formation of segmented rifted margins in cratonic lithosphere. Nature
Communications, 12(1), 4653. https://doi.org/10.1038/s41467-021-24945-5
Hacker, B. R., Kelemen, P. B., & Behn, M. D. (2015). Continental lower crust. Annual Review of Earth and Planetary Sciences, 43, 167–205.
Hatcher, J., Robert, D., Bream, B. R., & Merschat, A. J. (2007). Tectonic map of the southern and central Appalachians: A tale of three orogens and
a complete Wilson cycle. In 4-D Framework of continental crust (pp. 595–632). Geological Society of America Memoir 200. https://doi.org
/10.1130/2007.1200(29
Huismans, R. S., & Beaumont, C. (2003). Symmetric and asymmetric lithospheric extension: Relative effects of frictional-plastic and viscous
strain softening. Journal of Geophysical Research, 108(B10).
Huismans, R. S., & Beaumont, C. (2014). Rifted continental margins: The case for depth-dependent extension. Earth and Planetary Science
Letters, 407, 148–162. https://doi.org/10.1016/j.epsl.2014.09.032
Jamieson, R., & Beaumont, C. (2013). On the origin of orogens. GSA Bulletin, 11125(11–12), 1671–1702. https://doi.org/10.1130/B30855.1
Karato, S.-I., & Wu, P. (1993). Rheology of the upper mantle: A synthesis. Science, 260(5109), 771–778. https://doi.org/10.1126/science.
260.5109.771
Karner, G., Johnson, C., Shoffner, J., Lawson, M., Sullivan, M., Sitgreaves, J., et al. (2021). Chapter 9: Tectono-magmatic development of the
Santos and Campos basins, Offshore Brazil. In Memoir 124: The Supergiant lower Cretaceous pre-Salt Petroleum systems of the Santos Basin
(pp. 215–256). AAPG. https://doi.org/10.1306/13722321MSB.9.1853
Ma, C., Foster, D. A., Hames, W. E., Mueller, P. A., & Steltenpohl, M. G. (2019). From the Alleghanian to the atlantic: Extensional collapse of
the southernmost Appalachian orogen. Geology, 47(4), 367–370. https://doi.org/10.1130/G46073.1
Manatschal, G., Lavier, L., & Chenin, P. (2015). The role of inheritance in structuring hyperextended rift systems: Some considerations based on
observations and numerical modeling. Gondwana Research, 27(1), 140–164. https://doi.org/10.1016/j.gr.2014.08.006
Moresi, L., & Solomatov, V. (1998). Mantle convection with a brittle lithosphere: Thoughts on the global tectonic styles of the Earth and Venus.
Geophysical Journal International, 133(3), 669–682. https://doi.org/10.1046/j.1365-246X.1998.00521.x
Moulin, M., Aslanian, D., & Unternehr, P. (2010). A new starting point for the south and equatorial Atlantic ocean. Earth-Science Reviews, 98(1),
1–37. https://doi.org/10.1016/j.earscirev.2009.08.001
Naliboff, J., & Buiter, S. J. (2015). Rift reactivation and migration during multiphase extension. Earth and Planetary Science Letters, 421, 58–67.
https://doi.org/10.1016/j.epsl.2015.03.050
Neuharth, D., Brune, S., Wrona, T., Glerum, A., Braun, J., & Yuan, X. (2022). Evolution of rift systems and their fault networks in response to
surface processes. Tectonics, 41(3), e2021TC007166. https://doi.org/10.1029/2021TC007166
Nürnberg, D., & Müller, R. (1991). The tectonic evolution of the south atlantic from late jurassic to present. Tectonophysics, 191(1), 27–53.
https://doi.org/10.1016/0040-1951(91)90231-G
Olive, J.-A., Malatesta, L. C., Behn, M. D., & Buck, W. R. (2022). Sensitivity of rift tectonics to global variability in the efficiency of river
erosion. Proceedings of the National Academy of Sciences, 119(13), e2115077119. https://doi.org/10.1073/pnas.2115077119
Peron-Pinvidic, G., Fourel, L., & Buiter, S. (2022). The influence of orogenic collision inheritance on rifted margin architecture: Insights from
comparing numerical experiments to the mid-Norwegian margin. Tectonophysics, 828, 229273. https://doi.org/10.1016/j.tecto.2022.229273
Péron-Pinvidic, G., & Manatschal, G. (2010). From microcontinents to extensional allochthons: Witnesses of how continents rift and break apart?
Petroleum Geoscience, 16, 189–197.
Peron-Pinvidic, G., & Osmundsen, P. T. (2020). From orogeny to rifting: Insights from the Norwegian ‘reactivation phase. Scientific Reports, 10,
14860. https://doi.org/10.1038/s41598-020-71893-z
Petersen, K. D., & Schiffer, C. (2016). Wilson cycle passive margins: Control of orogenic inheritance on continental breakup. Gondwana
Research, 39, 131–144. https://doi.org/10.1016/j.gr.2016.06.012
Sacek, V., Assunção, J., Pesce, A., & da Silva, R. M. (2022). Mandyoc: A finite element code to simulate thermochemical convection in parallel.
Journal of Open Source Software, 7(71), 4070.
Salazar-Mora, C. A., Huismans, R. S., Fossen, H., & Egydio-Silva, M. (2018). The Wilson cycle and effects of tectonic structural inheritance on
rifted passive margin formation. Tectonics, 37(9), 3085–3101. https://doi.org/10.1029/2018TC004962
Santiago, R., de Andrade Caxito, F., Neves, M. A., Dantas, E. L., de Medeiros Júnior, E. B., & Queiroga, G. N. (2020). Two generations of
mafic dyke swarms in the southeastern brazilian coast: Reactivation of structural lineaments during the gravitational collapse of the araçuaí-
ribeira orogen (500 ma) and west gondwana breakup (140 ma). Precambrian Research, 340, 105344. https://doi.org/10.1016/j.precamres.
2019.105344
Silva, R. M., & Sacek, V. (2022). Influence of surface processes on post-rift faulting during divergent margins evolution. Tectonics, 41,
e2021TC006808. https://doi.org/10.1029/2021TC006808
Slagstad, T., & Kirkland, C. L. (2018). Timing of collision initiation and location of the Scandian orogenic suture in the Scandinavian caledon-
ides. Terra Nova, 30(3), 179–188. https://doi.org/10.1111/ter.12324
Sommaruga, A. (1999). Décollement tectonics in the jura forelandfold-and-thrust belt. Marine and Petroleum Geology, 16(2), 111–134.
Tetreault, J., & Buiter, S. (2018). The influence of extension rate and crustal rheology on the evolution of passive margins from rifting to break-up.
Tectonophysics, 746, 155–172. https://doi.org/10.1016/j.tecto.2017.08.029
Thieulot, C. (2014). Elefant: A user-friendly multipurpose geodynamics code. Solid Earth Discussions, 6(2), 1949–2096.
Unternehr, P., Péron-Pinvidic, G., Manatschal, G., & Sutra, E. (2010). Hyper-extended crust in the South Atlantic: In search of a model. Petroleum
Geoscience, 16, 207–215. https://doi.org/10.1144/1354-079309-904
Vauchez, A., Barruol, G., & Tommasi, A. (1997). Why do continents break-up parallel to ancient orogenic belts? Terra Nova, 9(2), 62–66. http
s://doi.org/10.1111/j.1365-3121.1997.tb00003.x
White, R. S., & Smith, L. K. (2009). Crustal structure of the hatton and the conjugate east Greenland rifted volcanic continental margins, NE
Atlantic. Journal of Geophysical Research, 114(B2). https://doi.org/10.1029/2008JB005856

SALAZAR-MORA AND SACEK 18 of 19


Tectonics 10.1029/2022TC007492

Willett, S. D. (1999). Orogeny and orography: The effects of erosion on the structure of mountain belts. Journal of Geophysical Research,
104(B12), 28957–28981.
Wolf, S. G., Huismans, R. S., Braun, J., & Yuan, X. (2022). Topography of mountain belts controlled by rheology and surface processes. Nature,
606, 516–521. https://doi.org/10.1038/s41586-022-04700-6
Wolf, S. G., Huismans, R. S., Muñoz, J.-A., Curry, M. E., & van der Beek, P. (2021). Growth of collisional orogens from small and cold to
large and hot—Inferences from geodynamic models. Journal of Geophysical Research: Solid Earth, 126(2), e2020JB021168. https://doi.
org/10.1029/2020JB021168
Zhong, S., Yuen, D. A., Moresi, L. N., & Schubert, G. (2007). Numerical methods for mantle convection. Treatise on Geophysics, 7, 227–252.

SALAZAR-MORA AND SACEK 19 of 19

You might also like