1 s2.0 S0024493722003620 Main

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

LITHOS 436-437 (2023) 106953

Contents lists available at ScienceDirect

LITHOS
journal homepage: www.elsevier.com/locate/lithos

The Chicxulub impact structure reveals the first in-situ Jurassic magmatic
intrusions of the Yucatán Peninsula, Mexico
Sietze J. de Graaff a, b, *, Catherine H. Ross c, Jean-Guillaume Feignon d, Pim Kaskes a, b,
Sean P.S. Gulick c, e, Steven Goderis a, Thomas Déhais a, b, Vinciane Debaille b,
Ludovic Ferrière d, f, Christian Koeberl d, Nadine Mattielli b, Daniel F. Stockli c, Philippe Claeys a
a
Research Unit: AMGC, Department of Chemistry, Vrije Universiteit Brussel, Pleinlaan 2, 1050 Brussels, Belgium
b
Laboratoire G-Time, Université Libre de Bruxelles, ULB, Av. F.D. Roosevelt 50, 1050 Brussels, Belgium
c
Institute for Geophysics & Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, TX 78758, USA
d
Department of Lithospheric Research, University of Vienna, Althanstrasse 14, A-1090 Vienna, Austria
e
Center for Planetary Systems Habitability, University of Texas at Austin, Austin, TX 78712, USA
f
Natural History Museum Vienna, Burgring 7, A-1010 Vienna, Austria

A R T I C L E I N F O A B S T R A C T

Keywords: Impact events that create complex craters excavate mid- to lower-crustal rocks, offering a unique perspective on
Chicxulub impact structure the interior composition and internal dynamics of planetary bodies. On the Yucatán Peninsula, Mexico, the
Gulf of Mexico opening surface geology mainly consists of ~3 km thick sedimentary rocks, with a lack of exposure of crystalline base­
Impact Melt Rock
ment in many areas. Consequently, current understanding of the Yucatán subsurface is largely based on impact
Jurassic pre-impact magmatic dikes
ejecta and drill cores recovered from the 180–200-km-diameter Chicxulub impact structure. In this study, we
Geochronology
Petrography present the first apatite and titanite U–Pb ages for pre-impact dacitic, doleritic, and felsitic magmatic dikes
preserved in Chicxulub’s peak ring sampled during the 2016 IODP-ICDP Expedition 364. Dating yielded two age
groups, with Carboniferous dacites (328–318 Ma) and a felsite (330± 9 Ma) overlapping in age with most of the
granitoid basement sampled in the Expedition 364 drill core, as well as Jurassic dolerites (169–159 Ma) and a
felsite (158 ± 19 Ma) that represent the first in situ sampling of Jurassic-age magmatic intrusions for the Yucatán
Peninsula. Further investigation of the Nd, Sr, and Hf isotopic compositions of these pre-impact lithologies and
impact melt rocks from the peak ring structure suggest that dolerites generally contributed up to ~10 vol% of the
Chicxulub impact melt rock sampled in the peak ring. This percentage implies that the dolerites comprised a
large part of the Yucatán subsurface by volume, representing a hitherto unsampled pervasive Jurassic magmatic
phase. We interpret this magmatic phase to be related to the opening of the Gulf of Mexico, representing the first
physical sampling of lithologies associated with the southern extension of the opening of the Gulf of Mexico and
likely constraining its onset to the Late Middle Jurassic.

1. Introduction numerical modelling has shown that large impact events (structures
with a rim-to-rim diameter exceeding ~200 km) can bring mid- to
Impact events are ubiquitous processes on planetary bodies, having lower-crustal rocks closer to the surface (Ivanov, 2005; Kring et al.,
important roles in the evolution of planets (Melosh, 1989). Sufficiently 2016; Morgan et al., 2016). In the case of the ~320-km-diameter
large impact events can instigate planet-wide events, such as major bi­ Schrödinger basin on the Moon, modelling suggests the presence of li­
otic turnover (e.g., Gulick et al., 2019; Schulte et al., 2010), climatic thologies in its peak ring structure that were possibly uplifted by ~30
repercussions (e.g., Vellekoop et al., 2014) and/or increased volcanic km (Kring et al., 2016), revealing components of the interior lunar crust.
output (e.g., Byrnes and Karlstrom, 2018). The predominantly destruc­ Consequently, impact structures naturally excavate crustal material and
tive nature of impact events often overshadows their efficacy in are potentially valuable in locations where the basement is poorly
furthering geological knowledge of their target rocks. For instance, exposed or in areas where deep drilling is not viable. On the Yucatán

* Corresponding author at: Research Unit: AMGC, Department of Chemistry, Vrije Universiteit Brussel, Pleinlaan 2, 1050 Brussels, Belgium.
E-mail address: [email protected] (S.J. de Graaff).

https://doi.org/10.1016/j.lithos.2022.106953
Received 12 June 2022; Received in revised form 6 November 2022; Accepted 19 November 2022
Available online 23 November 2022
0024-4937/© 2022 Elsevier B.V. All rights reserved.
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

nt
Carboniferous igneous, high-
grade metamorphic rocks ta Fro
i
ach Five Islands Salt domes
Jurassic aged lithologies Ou Suwanee Block
Famatian (400 - 500 Ma)

Pan-African (500 - 650 Ma)

Grenvillian (0.9 - 1.3 Ga) Gulf of Mexico


Mesozoic - Cenozoic accreted
terranes Aserradero/Novillo

Oa
Quaternary volcanics DSDP Leg 77

xaq
uia
Grenville /Appalachian
(1.3 Ga - 440 Ma M0077A
Peri-Gondwanan Huiznopala Guaniguanico massif
(~490 - 400, 420 - 350 &
330 - 270 Ma
Permian Chiapas Massif
Fig. 2
tns
Trans-Mexican Maya Block yaM
VolcanicBelt Guichicovi Ch Ma
iap es
Oa

as atan
xaq

Acatlan M um Z
Pacific Ocean as uch PF
uia

Complex sif C
Alt
os M N
Gulf ofTehuantepec Chortis
0 500km

Fig. 1. Terrane map showing the Gulf of Mexico region modified from Ross et al. (2022), with locations of known Jurassic crystalline lithologies shown: the
Louisiana salt dome xenoliths (Stern et al., 2011), NW-Cuba (Allibon et al., 2008) and DSDP leg 77 (Dallmeyer, 1984). Red square denotes area shown in Fig. 2. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Peninsula, Mexico, crystalline basement is not exposed at the surface, International Continental Scientific Drilling Program (ICDP) Expedition
with the (near-) surface geology mainly consisting of ~3 km of nearly 364 sampled the well-preserved peak ring of the impact structure,
horizontal strata of Lower Cretaceous to Quaternary carbonates and recovering ~830 m of near-continuous core from Hole M0077A, sam­
evaporites (Hildebrand et al., 1991; Lopez Ramos, 1975). The closest pling pre-, syn-, and post-impact lithologies (Morgan et al., 2017). The
exposed basement rocks are >500 km from the northern tip of the pre-impact lithologies entail crystalline basement that originated from
peninsula (Fig. 1). Consequently, current understanding of the Yucatán the mid-crustal level (8- to 10-km depth; Morgan et al., 2016), which
basement is mostly based on ejecta material (e.g., Belza et al., 2015) and hosts a variety of pre-impact magmatic dikes and impactite material (i.
limited samples recovered from earlier drilling efforts within the Chic­ e., lithologies formed because of the impact). The M0077A core thus
xulub impact structure (e.g., Kring, 2005; Sharpton et al., 1996). In provides access to a diverse array of lithologies and constitutes, for the
2016, the International Ocean Discovery Program (IODP) and first time, a unique window into the upper to middle crust of the Yucatán

Fig. 2. Surface geological map of the Northern


Yucatán Peninsula adapted from Kaskes et al. (2022),
summarizing ages obtained for samples from this
study and other boreholes. Borehole locations from
Lopez Ramos (1975); Urrutia-Fucugauchi et al.
(1996); and Morgan et al. (2017). Ages for M0077A:
This study (marked with an asterisk), Zhao et al.
(2020) and Ross et al. (2022); Yax-1: Keppie et al.
(2011) and Schmieder et al. (2018); C1: recalculated
data from Swisher et al. (1992) by Renne et al. (2013);
Y1: Lopez Ramos (1975) and Marton and Buffler
(1994); Y6: Krogh et al. (1993b) and Kettrup and
Deutsch (2003). Yax-1 = Yaxcopoil-1; C1 = Chicxu­
lub-1; Y1 = Yucatán-1; Y6 = Yucatán-6; U5 = UNAM-
5; U7 = UNAM-7.

2
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

Fig. 3. Hole M0077A core overview (adapted from de Graaff et al., 2022) with pre-impact lithologies show. UIM = upper impact melt rock unit; LIMB = lower
impact melt rock-bearing unit.

crystalline basement. 2. Geological setting


In this study we present new Sr, Nd, and Hf isotopic data for a subset
of samples of 22 pre-impact crystalline basement rocks and 10 impactite At the site of the Chicxulub impact structure, the crystalline basement
lithologies cored at Hole M0077A, previously characterized for major rocks are part of the Maya Block, the southeasternmost Mexican geological
and trace element compositions in de Graaff et al. (2022) and Feignon terrane (Fig. 1; Keppie et al., 2011; Weber et al., 2018). The Maya Block is
et al. (2021). Moreover, U–Pb ages for apatite and/or titanite in a set of widely viewed as a peri-Gondwanan terrane that forms the pre-Mesozoic
seven pre-impact dike samples that intrude the sampled crystalline basement of the Yucatán Peninsula, its Gulf of Mexico shelf, Chiapas,
basement were obtained in situ on 30 μm thin sections through depth- and north-central Guatemala (Martens et al., 2010; Fig. 1).
profile analyses to provide age constraints on the basement lithol­ Basement outcrops of the SW Maya Block are dominated by late
ogies. Through this combined geochemical approach, the nature of the Permian igneous and metamorphic rocks (270–250 Ma; Weber et al.,
relatively unknown Yucatán subsurface is better constrained, while also 2007) that are mostly exposed in the Chiapas Massif Complex (Fig. 1).
further demonstrating the value of large impact structures as windows The massif also contains Mesoproterozoic metamorphic and plutonic
into the deep crust. basement rocks that were intruded by Ordovician-age granites (Weber

3
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

et al., 2018, and references therein). Further exposures of the southern impactite that stayed close to the surface during crater development (de
Maya Block are located in the Altos Cuchumatanes in Guatemala and in Graaff et al., 2022; Fig. 3), and the lower impact melt rock-bearing unit
the Maya Mountains in central Belize. In the Altos Cuchumatanes, which is thought to represent impact melt rock injected into the crystalline
Pennsylvanian (312–317 Ma) to Middle Ordovician (461 Ma) granodi­ basement during the compression/excavation stage of cratering (de Graaff
orite intrudes ~1 Ga medium- to high-grade gneiss (Fig. 1; Weber et al., et al., 2022; Fig. 3). The sample set from de Graaff et al. (2022) was
2018). Whereas the basement of the Maya Mountains have Silurian characterized using a Nu Plasma II multi-collector inductively coupled
U–Pb ages of 420–405 Ma (Fig. 1; Steiner and Walker, 1996), that is plasma–mass spectrometer (HR-MC-ICP-MS) at Université Libre des
overlain by rhyolite interbedded with conglomerates, yielding U–Pb Bruxelles (ULB) (Laboratoire G-Time). Reference materials yielded NBS
ages of 406 + 7/− 6 Ma (Martens et al., 2010). In contrast, the northern 987 87Sr/86Sr = 0.710247 ± 0.000031 (2 standard deviation (SD), n = 12,
portion of the Maya Block underlying the Chicxulub structure appears to in agreement with the accepted value of 0.710248, Weis et al., 2006,),
mainly consist of ~550–545 Ma Pan-African granitoid basement based Rennes in-house standard 143Nd/144Nd = 0.511962 ± 0.000011 (2SD, n =
on zircon U–Pb analysis of distal impact ejecta (i.e., single grains within 16, in agreement with the accepted value of 0.511961; Chauvel and
the Cretaceous–Paleogene (K–Pg) boundary deposits) and suevite clasts Blichert-Toft, 2001) and JMC475 176Hf/177Hf = 0.282165 ± 0.000019
from the Yucatán-6 and Yaxcopoil-1 cores (Krogh et al., 1993a, 1993b; (2SD, n = 13, in agreement with the accepted value of 0.282163, Blichert-
Kamo and Krogh, 1995; Kettrup and Deutsch, 2003; Kamo et al., 2011; Toft, 2001). Procedural reference material BHVO-2 yielded 87Sr/86Sr =
Keppie et al., 2011; Fig. 2). Rubidium-Sr whole rock analyses from 0.703520 ± 0.000017 (2SD, n = 3), 143Nd/144Nd = 0.512978 ± 0.000000
rhyolite in the Yucatán-1 core (Fig. 2) yielded Early Devonian ages (410 (2SD, n = 3) and 176Hf/177Hf = 0.283107 ± 0.000011 (2SD, N = 6)
Ma; Lopez Ramos, 1975), while peak-ring granitoid samples recovered overlapping with known uncertainties of BHVO-2 consensus values
in the Hole M0077A core have thus far yielded Carboniferous U–Pb (0.703479 ± 0.00002, 0.512984 ± 0.000011, and 0.283105 ± 0.000011
ages ranging from 307 to 341 Ma (Fig. 2; Zhao et al., 2020; Ross et al., respectively; Weis et al., 2006, 2007). Total procedural blanks yield <500
2022). Paleomagnetic reconstruction of the Gulf of Mexico region in­ pg for Sr, < 26 pg for Nd and < 20 pg for Hf. For the samples from Feignon
dicates that the Maya Block behaved as a cohesive unit since the Late et al. (2021), Sr and Nd isotopic data were determined using a Thermo-
Triassic (ca. 230 Ma, e.g., Steiner, 2005). During the breakup of Pangea Finnigan Triton thermal ionization mass spectrometer (TIMS) at the Uni­
at the end of the Triassic, the Maya Block rifted away from the southern versity of Vienna (Austria). Mass fractionation was corrected for 88Sr/86Sr
margin of Laurentia to open the Gulf of Mexico between 230 and 150 Ma = 8.3752 and 146Nd/144Nd = 0.7219, respectively. A mean 87Sr/86Sr of
(e.g., Dickinson and Lawton, 2001; Mann et al., 2007; Pindell and 0.710257 ± 0.000006 (2SD, n = 5) was determined for NBS987 (Sr) and a
Dewey, 1982; Steiner, 2005; Stern and Dickinson, 2010). It reached its mean 143Nd/144Nd ratios of 0.511841 ± 0.000002 (2SD, n = 5) was ob­
present paleolatitude relative to North America during the Late Jurassic tained for the La Jolla (Nd) international standards during the period of
(ca. 150 Ma; Molina-Garza et al., 1992), after which it remained investigation. Total procedural blanks yield <1 ng for Sr and 40 pg for Nd.
geologically stable until the Chicxulub impact event at ~66.05 Ma (e.g., Methods for sample preparation and purification for the Feignon et al.
Renne et al., 2013; Sprain et al., 2018). (2021) and de Graaff et al. (2022) samples can be found in Appendix A.
Age determinations on dacitic, doleritic and felsitic dikes were car­
3. IODP-ICDP Expedition 364 ried out via in situ apatite and titanite U–Pb laser ablation inductively
coupled plasma mass spectrometry (LA-ICP-MS) analysis geochronology
The IODP-ICDP Expedition 364 Hole M0077A core (Fig. 3) sampled at the University of Texas at Austin. Prior to in situ LA-ICP-MS analyses,
the peak ring of the Chicxulub impact structure. The peak ring is pri­ uncovered thin sections (30 μm thickness) of the various dikes were
marily composed of coarse-grained granitoid that is overlain by syn- imaged using a JEOL 6460 low-vacuum Scanning Electron-Microprobe
impact melt rocks and pervasively intruded by pre-impact magmatic (SEM) at the Dept. Geological Sciences Geomaterials Characterization
dikes, varying from medium- to fine-grained dolerite, medium- to fine- and Imaging Facility at the University of Texas at Austin in back-
grained monzonitic felsite, and plagioclase-phyric dacite (Fig. 3), and scattered electron (BSE) and energy-dispersive X-ray (EDS) modes to
dikes of syn-impact melt rock and impact melt rock–bearing breccia (de document igneous rock textures and locate the accessory phases for LA-
Graaff et al., 2022; Morgan et al., 2017). In terms of syn-impact melt ICP-MS analysis. For apatite and titanite, we employed EDS elemental
rocks, the upper impact melt rock unit, overlying the peak-ring, is CaP and CaTi mapping of the thin sections, respectively. Subsequent to
clearly distinct from the lower impact melt rock-bearing unit, which EDS elemental mapping, we used the LA-ICP-MS system to raster over
intrudes the granitoid at different intervals throughout the peak ring the targeted apatite and titanite crystals, firing short-bursts (<10 shots)
structure (e.g., Morgan et al., 2017; de Graaff et al., 2022; Fig. 3). This while monitoring U and Pb counts, to refine the locations of the small
distinction is made based on core depth, petrography, and geochemistry (<50 μm) apatite and titanite phases. After spatial optimization, quan­
(see de Graaff et al., 2022 for an in-depth description of the different titative 100-shot LA-ICP-MS spot analyses were carried out using a
melt rock units). The granitoids, the most ubiquitous lithology within PhotonMachines Analyte G.2 Excimer laser with a large-volume Helex
the core, are considered the host rock into which the dolerites, felsites, sample cell and a Thermo Element2 ICP-MS. For each sample, we carried
dacites, and lower impact melt rock-bearing unit intruded. The dolerites out 50–100 quantitative spot analyses measuring 206Pb, 207Pb, 208Pb,
232
are the most prevalent intrusive pre-impact lithology in this section of Th, 235U, and 238U. The laser used an energy of 4 mJ, an energy
the peak-ring, whereas the felsites and dacites are comparatively rare, density of 1.37 J/cm2, a pulse rate of 10 Hz, and a 40 μm spot size.
with only four and three dikes observed in the entire core, respectively For mafic dolerite samples apatite was dated in situ. LA-ICP-MS an­
(Fig. 3). Pervasive alteration of both pre- and postimpact lithologies is alyses of samples were bracketed by Madagascar Fractionation reference
apparent throughout the Hole M0077A core (Kring et al., 2020). (MAD) apatite as the primary isotopic reference standard (in-house
TIMS ages of 472.4 ± 0.7 Ma; Thomson et al., 2012) and McClure
4. Sampling and methods Mountain (523.5 ± 1.5 Ma; Schoene and Bowring, 2006) as a secondary
standard. For intermediate dacite and felsic rhyolite dikes, either apatite
Strontium, Nd, and Hf isotopic compositions were determined on the (see above) or titanite were dated by in situ LA-ICP-MS analysis. For
same powder aliquots of a subset of 22 pre-impact crystalline basement titanite U–Pb analysis, CaTi elemental EDS maps were used to target
rock, sampling granitoid, dolerites, felsites and dacites, and 10 impact melt accessory titanite and if necessary spatial rastering was employed to
rock samples, previously characterized for major and trace element com­ optimize U and Pb counts prior to quantitative analysis. For titanite,
positions in Feignon et al. (2021) and de Graaff et al. (2022). The impact OTL-1 was used as primary titanite reference material (1015 ± 2 Ma;
melt rocks investigated in this study stem from five samples of the upper Kennedy et al., 2010) and bracketed every 4 samples for elemental and
impact melt rock unit, which is interpreted to represent a melt-bearing depth-dependent fractionation. BLR-1 (1047 ± 0.4 Ma, Aleinikoff et al.,

4
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

Ages of the pre-impact magmatic dikes. Core Section refers to the specific core sampling interval of the sample denoted as <Core>_<Section>_<Top>_<Bottom>. All errors are 2 sigma. Tera-Wasserburg Concordia
0.514
A Y-6 Gneiss (Kerup and Deutsch, 2003) Granitoids

0.8285
Y-6 Amphibolite (Kerup and Deutsch, 2003) Dolerites

Pbc

0.8
Y-6 Granite (Kerup and Deutsch, 2003) Dacites






MORB NW-Cuba dolerites (Allibon et al., 2008) Felsites
OI UIM
IA

MSWD
M LIMB
an
0.513 tle
Seawater

41
15
ar
i
143Nd/144Nd






ra
y

Uncertainty (2σ)
Co
Ocean-floor
Offset towards seawater sediments
0.512

1
2





Titanite Age (Ma)
0.511
0.701 0.703 0.705 0.707 0.709 0.711 0.713 0.715
87Sr/ 86Sr
i
B Atlantic MORB

321
318
Iceland
24






Indian MORB
DMM

0.8309

0.8232
18

0.828
Hawaii

0.84
0.84

0.84
Pbc


12
Society EMII

diagrams are shown in Appendix A. Ap = Apatite; Ttn = Titanite; TW-Windows = Manually selected windows used in Tera-Wasserburg diagrams.
MSWD
εHfi

Pitcairn

1.7
2.1
1.8
4.9
2.2
4.4
6


0
HIMU

Uncertainty (2σ)
-6 EMI
Kerguelen Xenoliths (Stern et al., 2011)
-12

23
18
23
19
9
9
-8 -4 0 4 8 12


εNdi

Apatite Age (Ma)


Fig. 4. (A) 143Nd/144Ndi vs 87Sr/86Sri (66 Ma) with important magmatic do­
mains from fig. 13.11 in Philpotts and Ague (2009) are shown. MORB = Mid-
ocean ridge basalt (MORB); OI = oceanic islands; AI = island-arc rocks; Co =
continental flood basalts (Co). (B) εHfi vs εNdi (66 Ma) with present-day mantle

169
159
162
158
330
328

sources from Stern et al. (2011). Error bars are smaller than symbols. Xenoliths
refer to the Louisiana salt dome xenoliths from Stern et al. (2011). UIM = upper

84 (Ap) 80 (Ttn)
impact melt rock unit; LIMB = lower impact melt rock-bearing unit; EMI and
TW-Windows

EMII = enriched mantle types; MORB = mid ocean ridge basalt; HIMU = high
238
U/204Pb ratio; DMM = depleted mantle MORB. 133
238
2002; Bonamici et al., 2015) was used as a secondary standard for
78
53
47

54
quality control. The primary standards were used for both down-hole
64 (Ap) 80 (Ttn)

elemental and isotopic mass-fractionation correction. Data were


Grains analyzed

reduced using Iolite v3.5 (Paton et al., 2011) data reduction software
and VizualAge UcomPbine (Chew et al., 2014).
U–Pb data for all titanite and apatite phases were determined from
50
49
92
97
56

64

multiple whole-grain (bulk) integrations or by selecting multiple inter­


nal zones from individual grains that exhibit heterogeneous common Pb
Lithology

Dolerite
Dolerite
Dolerite

compositions. Tera-Wasserburg concordia lower intercept ages were


Felsite
Felsite
Dacite
Dacite

calculated from arrays of whole-grain or intra-grain zone data, forming


common Pb (Pbc) and radiogenic Pb (Pb*) discordia mixing-line arrays
(Tera and Wasserburg, 1972). No common 204Pb corrections were
Core Depth (mbsf)

applied due to the interference with 204Hg in the He and Ar carrier and
sample gases, but a 207Pb correction was applied to the primary refer­
1124.16
1138.15

1161.41

ence material. Tera-Wasserburg Concordia diagrams were plotted and


854.74
865.48
934.84

920.37

lower intercept 206Pb/238U ages were calculated using IsoplotR (Ver­


meesch, 2018). All results are presented as 2-sigma with internal and
external uncertainty propagated. All Tera-Wasserburg Concordia dia­
234_2_101.5_106
239_1_105_107
164_3_14_16.5

grams are presented in Appendix A.


Core Section

140_2_19_21
144_1_19_21
169_2_97_99

247_2_21_23

5. Results

For the granitoids, felsites, and dacites, the 87Sr/86Sr isotope ratios
Sample_ID

range from 0.7054 to 0.7088 reflecting an offset towards seawater


STS_11

STS_13
Table 1

values (Fig. 4, Appendix B), and crustal Nd and Hf values (Fig. 4, Ap­
STS_6
BD_2
432

517

757

pendix B; εNd between − 3.33 and − 2.09, εHf between − 1.97 and +

5
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

0.23). The dolerites show a comparable offset towards more radiogenic clast recovered from DSDP Leg 77 Hole 538A (~190 Ma; Dallmeyer,
87
Sr/86Sr of 0.705726 to 0.706900, similar to those of the other crys­ 1984; Fig. 1). Essentially, there is a severe lack of in situ Jurassic rocks
talline lithologies (Fig. 4), but are characterized by radiogenic εHf along the south, west, and northern coast of the Gulf of Mexico, pre­
values between +12.37 and + 12.73 and less radiogenic Nd values, with venting a well-constrained age determination for the opening of the Gulf
εNd varying from +7.6 to − 2.7, which potentially reflect the signature of Mexico. This marks the dolerites presented in this study as the first in-
of an enriched mantle source. situ Jurassic crystalline lithologies to have been sampled in the Yucatán
The impact melt rock units exhibit a similar trend towards more Peninsula, and the southern Gulf of Mexico region in general.
radiogenic Sr values at near identical Nd values as observed for the
crystalline basement material (Fig. 4; 87Sr/86Sr of 0.70724 to 0.70893, 6.2. The excavated lithologies of the Hole M0077A core
likely indicating a hydrothermal overprint recorded in all the presented
lithologies, as is also suggested based on petrographic and trace element Two distinct ages of magmatism were identified crosscutting the
compositions by Kring et al. (2020) and de Graaff et al. (2022). granitoid of the Hole M0077A core, with the dacites presenting a
Furthermore, both types of impact melt rocks show crustal Nd and Hf Carboniferous phase and the dolerites indicating a Jurassic phase, with
values comparable to the felsic crystalline basement for most samples, the felsites overlapping in age with both (Table 1). Notably, the isotopic
except three samples with Nd and Hf values displaying more radiogenic composition of the dacites and felsites are nearly identical, suggesting,
compositions, with εNd varying from − 1.92 to +0.04 and εHf from to a certain extent, that they likely originated from the same geodynamic
+3.16 to +11.36, respectively (Fig. 4). environment (Fig. 4). On the basis of petrography and major and trace
The ages for the pre-impact lithologies of the M0077A core are element compositions, de Graaff et al. (2022) suggested that the dacites
presented in Table 1, with all Tera-Wasserburg Concordia diagrams originated from the same magmatic source as the granitoids, likely
shown in Appendix A. The dacites yield Carboniferous ages (~328–318 having undergone less fractional crystallization, while the felsites reflect
Ma) that overlap with known ages of the Carboniferous peak ring intracrustal partial melts of the granitoid crust. These interpretations
granitoid (Ross et al., 2022; Zhao et al., 2020). The dolerites yield agree with the isotopic compositions presented here, but are inconsis­
Jurassic ages (~169–159 Ma), whereas the felsites vary significantly in tent with the two age groups found for the felsites (Table 1). The fact
age, with one sample yielding a Carboniferous age of 330 ± 9 Ma, while that the felsites overlap in age with both Carboniferous granitoid and
another displays a Jurassic age of 158 ± 19 Ma (Table 1). dacite or Jurassic dolerite appears to indicate that felsites formed in
response to partial melting during both the emplacement of voluminous
6. Discussion Carboniferous felsic plutonic rocks as well as the emplacement of
Jurassic mafic dikes.
6.1. The sampling age gap of the Yucatán Peninsula The enigmatic felsites notwithstanding, the dolerites present a
distinct phase of magmatism intruding into the Carboniferous granitoid.
In current literature, the central, eastern, and southern Mexican crust The isotopic composition of the dolerites shows a mantle Nd composi­
comprises late Mesoproterozoic protoliths (1.25–1.0 Ga) that underwent tion (Fig. 4), but with a distinct offset towards more radiogenic Sr
granulite-facies metamorphism (~ 0.99 Ga; e.g., Ortega-Gutierrez et al., (Fig. 4A) and less radiogenic Hf crustal values (Fig. 4B). These isotopic
1995; Ortega-Gutiérrez et al., 2018; Fig. 1A). Ediacaran sedimentary unit data need to be interpreted with care, however, as the pervasive hy­
deposition and rift-related mafic dyke swarm emplacement occurred in NE drothermal alteration of the Chicxulub impact structure likely over­
Mexico and Chiapas after the onset of the formation of the Iapetus Ocean printed pristine isotopic compositions (Kring et al., 2020; Simpson et al.,
and the final fragmentation of Rodinia (~650 Ma; González-Guzmán et al., 2020; Feignon et al., 2021; de Graaff et al., 2022). Alteration notwith­
2016; Weber et al., 2019, 2020). Ordovician (~480–450 Ma) magmatism standing, these isotopic values imply a crustal component incorporated
and related crustal anatexis, as recorded in Chiapas, Altos Cuchumatanes, into the source region of the dolerite, rather than typical mid-oceanic
and Rabinal, were linked to the northern continuation of the Famatinian ridge basalt (MORB) or oceanic island basalt (OIB) derived dolerites.
arc along western Gondwana (Ortega-Obregón et al., 2008, 2009; Estrada- When comparing the composition of the dolerites to the other lithologies
Carmona et al., 2012; Weber et al., 2018; Alemán-Gallardo et al., 2019; of Jurassic age sampled in the Gulf of Mexico region, we observe that the
Juárez-Zúñiga et al., 2019). By the Devonian, the Rheic Ocean was closing, M0077A dolerites are comparable to the Louisiana salt dome xenoliths
leading up to the formation of the Pangean supercontinent in the late in Hf isotopic composition and N-MORB-normalized REE content (Stern
Paleozoic (e.g., Nance and Linnemann, 2008). Along the NW margin of et al., 2011; Figs. 4B and 5), potentially indicating that they originated
Gondwana, this convergence culminated in a laterally-diachronous colli­ from a similar source. The NW Cuba dolerites, however, show an
sion and suturing of Gondwana with Laurentia during the Ouachita- N-MORB normalized trace element pattern more akin to spreading
Marathon-Appalachian Orogeny in the latest Carboniferous and early center type tholeiites (Fig. 5; Allibon et al., 2008), with a mantle-like Nd
Permian (e.g., Dickinson and Lawton, 2001). The recent studies on the isotope composition (Fig. 4A). Based on paleogeographic re­
Hole M0077A core support the occurrence of a pre-collisional continental constructions around the time of dolerite formation (~160 Ma; Fig. 6)
volcanic arc located along the northern margin of the Maya Block before the plate configuration would suggest that the Louisiana salt dome xe­
NW Gondwana collided with Laurentia, closing the Rheic Ocean (Ross noliths (Stern et al., 2011) and M0077A dolerites (this study) are con­
et al., 2022). Age constraints on the younger geological history of the jugate margins during the opening of the Gulf of Mexico, while the
Yucatán Peninsula, and, most notably, the opening of the Gulf of Mexico Cuban dolerites are closer to the proto-Caribbean spreading center
are only based on seismic imaging and paleomagnetic and plate tectonic (Fig. 6; Pindell, 1985). These observations strongly indicate that the salt
reconstructions (e.g., Hudec et al., 2013; Pindell et al., 2016; Pindell and dome xenoliths and Hole M0077A dolerites are related to the opening of
Kennan, 2009; Steiner, 2005). the Gulf of Mexico. On the other hand, the distinction between those
Despite this general understanding about when and how it opened, dolerites and the NW Cuba dolerites highlights a difference in source
the Gulf of Mexico is a rare example of where the formation of a sizable between the Gulf of Mexico region north and south of the Maya Block
oceanic basin is still unclear, as the thick blanketing sediments (locally as during the Middle to Late Jurassic.
thick as 16 km; e.g., Muehlberger, 1992) cover any lithological evidence
of its opening (e.g., Stern et al., 2011). However, age constraints from 6.3. Implications of the excavated dolerites
that period do exist, with Jurassic ages found on the northern margin of
the Gulf of Mexico in xenoliths sampled from salt diapirs offshore Loui­ With the majority of the Hole M0077A core consisting of Carbonif­
siana (160–158 Ma; Stern et al., 2011; Fig. 1), doleritic dikes in northwest erous granitoids (e.g., Morgan et al., 2016), their presence indicates they
Cuba (165–160 Ma; Allibon et al., 2008; Fig. 1) and in a single dolerite represent a pervasive magmatic phase in the geodynamic history of the

6
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

Fig. 5. N-MORB–normalized trace-element concentrations, with normalization values from (Sun and McDonough, 1989). NW-Cuba dolerites ages and trace element
data from Allibon et al. (2008); Louisiana salt dome xenolith ages and trace element data from Stern et al. (2011). Hole M0077A core trace element data from de
Graaff et al. (2022), ages from this study.

Yucatán Peninsula (e.g., Ross et al., 2022; Zhao et al., 2020). In contrast, impactites in general, are by majority of andesitic composition,
the prevalence of the dolerites remains uncertain as the core only pro­ requiring the melting and/or mixing of both felsic and mafic compo­
vides a spatially restricted sample set. Essentially, while the dolerites nents to explain its geochemical composition (Claeys et al., 2003; de
yield the first in situ Jurassic age of the Yucatán Peninsula, it remains to Graaff et al., 2022; Kettrup and Deutsch, 2003; Kring and Boynton,
be determined whether we can unequivocally link them to the opening 1992; Tuchscherer et al., 2004). Given the large amount of granitoid
of the Gulf of Mexico. For this reason, we examine the impact melt rock present in the Hole M0077A core, and thus by extension the peak ring
units of the Hole M0077A core. As impact melt rock is essentially the (see also Morgan et al., 2016), a large mafic component is thus necessary
molten amalgamation of the available target rock, it can be used to to explain the overall andesitic composition of the Chicxulub impactites
constrain the nature and volume of specific pre-impact lithologies in the and the Hole M0077A impact melt rocks (de Graaff et al., 2022;
crustal make-up of the target rocks within impact craters (Grieve et al., Tuchscherer et al., 2004). However, overall, mafic lithologies are noted
1977). The Hole M0077A impact melt rock units, and Chicxulub to be rare in the Chicxulub impactite suite (e.g., Tuchscherer et al.,

Fig. 6. Schematic Late-Middle Jurassic


s (~160 Ma) paleogeography of the Gulf
North America nlap ch
ian region, modified from Pindell (1985).
eo us o Mississippi
Embayment ala During this time seafloor spreading
etac Ap
p
f Cr Ouachita ridges develop in the Gulf, heralding the
it o er
n
Lim th opening of the Gulf of Mexico, while
u
So Gulf-wide salt is being deposited. High­
lighted with yellow stars are the loca­
MU
a lco

SU CMS tions of the M0077A dolerites (this


Liano
study), Northwestern Cuba samples (SG;
ia-T

ETS NLS WA Suwanee Allibon et al., 2008) and Louisiana salt


Mex

Marathon
Block
Blake dome xenoliths (NLS; Stern et al., 2011).
Platea Paleo 10 N CMS = Central Mississippi salt basin;
u
- GOM ETS = East Texas salt basin; FSB =
C Proto Florida Straits Block; MU = Monroe
Bl oah
oc uil uplift; NLS = North Louisiana salt basin;
tic
rly

k a
PS = Gulf of Paria salt basin; SG = Sierra
an

Guaniguanico; SU = Sabine uplift; TA =


Ea

TA
M0077A
FSB Tampa Arch; WA = Wiggins Arch. Strike
Atl

Ya
Bl qui SG and dip symbols are schematic for rifting
oc
k
a
May ck (Pindell, 1985). (For interpretation of
Blo the references to colour in this figure
to legend, the reader is referred to the web
Salt provinces Pro ean version of this article.)
NW Cuba
a r ibb
Sample
C PS
location
Spreading
center South America
Thrust 500 km
front

7
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

2004), stressing the importance of constraining the role of the dolerites not dealing with pristine samples of homogenized impact melt, some­
in the formation of impact melt rock. thing already concluded for major and trace element composition (de
The Chicxulub impact melt rock studied here generally shows Nd, Sr, Graaff et al., 2022). Importantly, however, no significant trends towards
and Hf isotopic compositions very similar to those of the felsic crystal­ other potential endmembers (Y-6 gneisses or Y-6 granite) are observed
line basement (i.e., granitoid, dacites and felsites) in most samples. (Fig. 4A) in the Hole M0077A core, except for one impact melt rock
However, most samples are offset towards slightly higher Hf and lower sample effectively overlapping in Nd and Sr isotopic composition with
Nd ratios (best observed in Fig. 4B), and three samples define a trend amphibolite, a lithology previously proposed as the mafic component of
towards higher Hf and Nd ratios (Fig. 4B). Based on mixing lines the Chicxulub impact melt (Tuchscherer et al., 2004; Fig. 4A). However,
calculated between mean values of the dolerite and granitoid datasets based on their comparable isotopic values, this would require this
for Nd and Sr, the impact melt rock samples indicate a potential doleritic sample to be almost entirely composed of amphibolite melt, which is not
contribution of up to ~45 vol% in one sample (Fig. 7). This doleritic realistic based on the sample’s major and trace element compositions
contribution is also present in the Nd and Hf isotope systematics, with a (see de Graaff et al., 2022; sample 194_3_19_21). Additionally, am­
possible admixture of ~10 vol% of dolerite that in some samples reaches phibolites have only been documented as rare clasts in impact breccias
up to ~40 vol%, and in a single sample potentially even 90 vol% (Fig. 7). (Tuchscherer et al., 2004), which contrasts to the multitude of doleritic
The samples that suggest a larger vol% of dolerite admixture are likely a dikes sampled in the single Hole M0077A core.
consequence of impact melt rock heterogeneity. A similar suggestion has In a study of the major and trace elemental compositions of the
been made for impact melt rock samples showing an enriched trace impact melt rock, de Graaff et al. (2022) suggested that the dolerites are
element signature, which likely reflect a potential majority felsitic indeed the major mafic component that contributed to the impact melt
component (de Graaff et al., 2022). These large variations in impact melt rocks of the peak ring structure. These observations are further sub­
rock compositions in the Hole M0077A core strongly implies that we are stantiated by the isotopic compositions presented here, moreover,

Fig. 7. (A) 143Nd/144Ndi vs 87Sr/86Sri (66 Ma) with mixing lines calculated between averaged dolerite and granitoid. (B) εHfi vs εNdi (66 Ma) with mixing lines
calculated between specific end-members. UIM = upper impact melt rock unit; LIMB = lower impact melt rock-bearing unit

8
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

Distance from center of impact (km) Fig. 8. Schematic cross section of the pre-impact
NW SE target rock of the Yucatán Peninsula modified from
90 45 0 45 90
0 Kring (2005), highlighting the target rocks affected
Carbonate and evaporite platform 200 GPa Sandstone layers
by the impact, based on results presented in this
100 GPa
Future 80 GPa study. Depth of pressure domains, future peak-ring
5 Pan Peak ring
and transient cavity size from Morgan et al. (2016)
60 GPa
African Upper crust
basement Middle crust and Artemieva et al. (2017). Lithologies above 60
10 (550 - Carboniferous arc GPa are subjected to melting, ejection, and/or
465 Ma) magmatism Transient vaporization. The basement is composed of pre-
Depth (km)

cavity
15 Cambrian metamorphic basement (Kettrup and
Deutsch, 2003) that is overlain or intruded by felsic
20 dacites-diorites-granodiorites and granites of Pan-
African age (550–465 Ma) and Carboniferous age
??
?? ?? (345–317 Ma) (Ross et al., 2022). These are in turn
25 ?? intruded by Jurassic (169–159 Ma; this study) mag­
??
matism. Upper – Mid crustal boundary from Ver­
30 Precambrian metamorphic ?? Jurassic
meesch and Morgan (2008).
basement (1.4 - 0.7 Ga) magmatism
??
Upper Mantle
35

paired with the macroscopic observation of the prevalent occurrence of 7. Conclusions


dolerites in the M0077A core (Morgan et al., 2017), the dolerites have
the potential to represent the major mafic source end-member that This study presents the first apatite and titanite U–Pb ages and Nd,
contributed to the impact melt rocks of the Chicxulub impact structure, Sr and Hf isotopic compositions for pre-impact dacitic, doleritic, and
rather than the earlier suggested comparatively rare amphibolites (e.g., felsitic magmatic dikes preserved in Chicxulub’s peak ring sampled
Tuchscherer et al., 2004). However, based on the heterogeneity of the during the 2016 IODP-ICDP Expedition 364. Apatite and titanite U–Pb
impact melt rocks presented here, it is likely that the melt rock units ages yielded two age groups, with Carboniferous dacites (328–318 Ma)
sampled in the Hole M0077A core are not fully representative of the and a felsite (330 ± 9 Ma) overlapping in age with most of the granitoid
entire ~2.5 km thick Chicxulub melt sheet (see also de Graaff et al., basement, as well as Jurassic dolerites (169–159 Ma) and a felsite (158
2022). Nonetheless, the impactites originating from the Chicxulub ± 19 Ma) that represent the first in situ sampling of Jurassic-age crys­
impact structure are by majority andesitic (e.g., Belza et al., 2015; talline lithologies for the Yucatán Peninsula. The Nd, Sr, and Hf isotopic
Koeberl, 1993; Tuchscherer et al., 2005). With the dolerites being the compositions of the pre-impact lithologies and impact melt rocks pre­
first major mafic magmatic sequence sampled in the Yucatán subsurface, sented here indicate that the doleritic magmatic sequence contributed to
overlapping in age and composition with similar mafic lithologies on the the Hole M0077A impact melt rock units, with some samples reflecting
northern part of the Gulf of Mexico (Fig. 5), we suggest the doleritic up to ~40 vol% of dolerite incorporation. This implies that the dolerites
magmatic sequence as the most likely candidate for the Chicxulub are likely sourced from a magmatic parent body that comprises a large
impactite suite mafic end-member. part of the Yucatán subsurface by volume, representing a hitherto
Consequently, the dolerites potentially constrain parts of a pervasive unsampled pervasive Jurassic magmatic phase. The proposed preva­
magmatic sequence that constitutes a key component of the northern lence of this magmatic phase in the Yucatán Peninsula’s subsurface
Yucatán subsurface that volumetrically accounts for the andesitic implies they were part of a significant magmatic event occurring be­
composition of the Chicxulub impact melt rocks, and impactite suite. tween the Early Triassic and the Late Jurassic. As such, we interpret
Paired with the Jurassic ages of the dolerites, these lithologies likely these dolerites as the first physical sampling of lithologies associated
represent the first physical sampling of lithologies associated with the with the southern extension of the opening of the Gulf of Mexico, likely
southern extension of the opening of the Gulf of Mexico, potentially constraining its onset to the Late Middle Jurassic. These new insights
constraining its onset to the Middle to Late Jurassic (Fig. 6). into the geodynamic history of the Yucatán Peninsula disclosed from a
single core within a peak-ring structure reinforce that medium-to-large
6.4. Impact structures, a window into the subsurface complex impact structures are useful natural probes of deep-basement
lithologies on terrestrial planets, providing unique windows into the
With the recognition of the first in situ Jurassic crystalline lithologies crust.
of the Yucatán Peninsula this study confirms the value of impact events
to excavate deeper or otherwise inaccessible crustal rock (e.g., Kring Declaration of Competing Interest
et al., 2016, Figs. 2, 3). In the case of Chicxulub, a near-complete tec­
tonic history of the Yucatán is preserved in a single core in the peak ring The authors declare that they have no known competing financial
structure despite a lack of exposure in the surrounding area (Fig. 1). As interests or personal relationships that could have appeared to influence
impact cratering is ubiquitous on rocky bodies in our solar system, using the work reported in this paper.
geochemical and geochronological tools on lithologies formed and
excavated during impact leads to critical information on a planet’s Data availability
deeper crustal composition, especially when inaccessible through other
means. Following these assertions, we present a schematic cross-section All data presented and used in this manuscript can be found in
of the pre-impact subsurface of the Yucatán Peninsula (Fig. 8), high­ Table 1 and Appendix A and B, moreover the U-Pb data will be made
lighting the excavated portions of the subsurface as a result of the available at www.geochron.org. Data sets from other studies are refer­
Chicxulub impact and the apparent pervasive nature of the Jurassic enced in the figure captions of Figs. 1, 2, 4 and 5. Samples of the
dolerites. Expedition 364 Hole M0077A core can be requested from https://iodp.
tamu.edu/curation/samples.html.

9
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

Acknowledgements southeastern Gulf of Mexico: Tectonic implications: College Station, Texas. Deep Sea
Drill. Prog. Initial Rep. 77, 497–504.
de Graaff, S.J., Kaskes, P., Déhais, T., Goderis, S., Debaille, V., Ross, C.H., Gulick, S.P.S.,
We thank Wendy Debouge, Sabrina Cauchies and Jeroen de Jong Feignon, J.G., Ferrière, L., Koeberl, C., Smit, J., Matielli, N., Claeys, P., 2022. New
(ULB); Monika Horschinegg and Wencke Wegner (University of Vienna); insights into the formation and emplacement of impact melt rocks within the
and Phil Orlandini and Lisa Stockli (UT) for their assistance with sample Chicxulub impact structure, following the 2016 IODP-ICDP Expedition 364: Geol.
Soc. Am. Bull. 134 (1-2), 293–315. https://doi.org/10.1130/B35795.1.
preparation and analytical procedures. The VUB team is supported by Dickinson, W.R., Lawton, T.F., 2001. Carboniferous to cretaceous assembly and
FWO; project G0A6517N and BELSPO; PK is an FWO Ph.D. fellow fragmentation of Mexico. Geol. Soc. Am. Bull. 113, 1142–1160 doi:10.1130/.
(project 11E6621N). PC thanks the VUB Strategic Program for support. Estrada-Carmona, J., Weber, B., Martens, U., López-Martínez, M., 2012. Petrogenesis of
Ordovician magmatic rocks in the southern Chiapas Massif complex: Relations with
VD acknowledges the ERC Starting Grant “ISo-SyC” and Fund for Sci­ the early Palaeozoic magmatic belts of northwestern Gondwana. Int. Geol. Rev. 54,
entific Research (FRS-FNRS), for funding. VD, NM, PC and SG thank the 1918–1943. https://doi.org/10.1080/00 206814.2012.685553.
Excellence of Science project “ET-HoME” for support. Partial funding Feignon, J.G., de Graaff, S.J., Ferrière, L., Kaskes, P., Déhais, T., Goderis, S., Claeys, P.,
Koeberl, C., 2021. Chicxulub impact structure, IODP-ICDP Expedition 364 drill core:
was provided by the University of Vienna doctoral school IK-1045. This Geochemistry of the granite basement. Meteorit. Planet. Sci. 56, 1243–1273. https://
research was supported by the U.S. Science Support Program and Na­ doi.org/10.1111/maps.13705.
tional Science Foundation grant OCE-1737351. DFS acknowledges González-Guzmán, R., Weber, B., Manjarrez-Juárez, R., Cisneros de León, A., Hecht, L.,
Herguera-García, J.C., 2016. Provenance, age constraints and metamorphism of
support from the University of Texas Chevron Gulf Centennial Profes­ Ediacaran metasedimentary rocks from the El Triunfo complex (SE Chiapas, México):
sorship. SPSG was supported by NASA 19-HW19_2-0055. Greg Shellnutt evidence for Rodinia breakup and Iapetus active margin. Int. Geol. Rev. 58 (16),
is thanked for editorial handling, and we thank John Spray and Martin 2065–2091 doi.org/10.1080/00206814.2016.1207208.
Grieve, R.A., Dence, M.R., Robertson, P.B., 1977. Cratering processes-as interpreted from
Schmieder for their constructive feedback on the manuscript. This
the occurrence of impact melts. In: Roddy, D.J., Pepin, R.O., Merril, R.B. (Eds.),
research used samples provided by the International Ocean Discovery Impact and Explosion Cratering: Planetary and Terrestrial Implications. Pergamon
Program. This is University of Texas Institute for Geophysics Contribu­ Press, New York, pp. 791–814.
tion #3886 and Center for Planetary Systems Habitability Contribution Gulick, S.P.S., et al., 2019. The first day of the Cenozoic. Proc. Natl. Acad. Sci. U. S. A.
116, 19342–19351. https://doi.org/10.1073/pnas.1909479116.
#0060. Hildebrand, A.R., Penfield, G.T., Kring, D.A., Pilkington, M., Camargo, Z.A., Jacobsen, S.
B., Boynton, W.V., 1991. Chicxulub Crater: a possible cretaceous/Tertiary boundary
Appendix A. Supplementary data impact crater on the Yucatán Peninsula, Mexico. Geology 19, 867–871. https://doi.
org/10.1130/0091-7613(1991)019<0867:CCAPCT>2.3.CO;2.
Hudec, M.R., Norton, I.O., Jackson, M.P., Peel, F.J., 2013. Jurassic evolution of the Gulf
The online supplementary data consists of Appendix A, containing of Mexico salt basin. AAPG Bull. 97 (10), 1683–1710.
the methodology for sample preparation for all the presented Ivanov, B.A., 2005. Numerical modeling of the largest terrestrial meteorite craters. Sol.
Syst. Res. 39 (5), 381–409.
geochemical results and all Tera-Wasserburg Concordia diagrams of the Juárez-Zúñiga, S., Solari, L.A., Ortega-Obregón, C., 2019. Ordovician to Silurian igneous
ages presented in Table 1; and Appendix B, containing all presented rocks in southern Mexico and Central America: Geochronologic and isotopic
geochemical data and specific core depths of all samples. Supplementary constraints on paleogeographic models. J. S. Am. Earth Sci. 93, 462–479. https://
doi.org/10.1016/j.jsames.2019.04.023.
data to this article can be found online at https://doi.org/10.1016/j.lith Kamo, S.L., Krogh, T.E., 1995. Chicxulub crater source for shocked zircon crystals from
os.2022.106953 the Cretaceous-Tertiary boundary layer, Saskatchewan: evidence from new U-Pb
data. Geology 23, 281–284. https://doi.org/10.1130/0091-7613(1995)023<0281:
CCSFSZ>2.3.CO;2.
References
Kamo, S.L., Lana, C., Morgan, J.V., 2011. U–Pb ages of shocked zircon grains link distal
K–Pg boundary sites in Spain and Italy with the Chicxulub impact. Earth Planet. Sci.
Aleinikoff, J., Wintsch, R., Fanning, C., Dorais, M., 2002. U–Pb geochronology of zircon Lett. 310, 401–408. https://doi.org/10.1016/j.epsl.2011.08.031.
and polygenetic titanite from the Glastonbury Complex, Connecticut, USA: an Kaskes, P., et al., 2022. Formation of the crater suevite sequence from the Chicxulub peak
integrated SEM, EMPA, TIMS, and SHRIMP study. Chem. Geol. 188, 125–147. ring: a petrographic, geochemical, and sedimentological characterization. Geol. Soc.
https://doi.org/10.1016/S0009-2541(02)00076-1. Am. Bull. 134 (3–4), 895–927. https://doi.org/10.1130/B36020.1.
Alemán-Gallardo, E.A., Ramírez-Fernández, J.A., Rodríguez-Díaz, A.A., Velasco- Kennedy, A.K., Kamo, S.L., Nasdala, L., Timms, N.E., 2010. Grenville skarn titanite:
Tapia, F., Jenchen, U., Cruz- Gámez, E.M., León-Barragán, L.D., León, I.N.-D., 2019. potential reference material for SIMS U–Th–Pb analysis. Can. Mineral. 48 (6),
Evidence for an Ordovician continental arc in the pre-Mesozoic basement of the 1423–1443.
Huizachal–Peregrina Anticlinorium, Sierra Madre Oriental, Mexico: Peregrina Keppie, J.D., Dostal, J., Norman, M., Urrutia-Fucugauchi, J., Grajales-Nishimura, M.,
Tonalite. Mineral. Petrol. 113, 505–525. https://doi.org/10.1007/s00710-019- 2011. Study of melt and a clast of 546 Ma magmatic arc rocks in the 65 Ma
00660-4. Chicxulub bolide breccia, northern Maya block, Mexico: western limit of Ediacaran
Allibon, J., Lapierre, H., Bussy, F., Tardy, M., Cruz Gàmez, E.M., Senebier, F., 2008. Late arc peripheral to northern Gondwana. Int. Geol. Rev. 53 (10), 1180–1193. https://
Jurassic continental flood basalt doleritic dykes in northwestern Cuba: remnants of doi.org/10.1080/00206810903545527.
the Gulf of Mexico opening. Bull. Soc. Géol. France 179 (5), 445–452. Kettrup, B., Deutsch, A., 2003. Geochemical variability of the Yucatàn basement:
Artemieva, N., Morgan, J., Expedition 364 Science Party, 2017. Quantifying the release Constraints from crystalline clasts in Chicxulub impactites. Meteorit. Planet. Sci. 38
of climate-active gases by large meteorite impacts with a case study of Chicxulub. (7), 1079–1092. https://doi.org/10.1111/j.1945-5100.2003.tb00299.x.
Geophys. Res. Lett. 44 (20), 10–180. https://doi.org/10.1002/2017GL074879. Koeberl, C., 1993. Chicxulub crater, Yucatán: Tektites, impact glasses, and the
Belza, J., Goderis, S., Smit, J., Vanhaecke, F., Baert, K., Terryn, H., Claeys, Ph., 2015. geochemistry of target rocks and breccias. Geology 21 (3), 211–214 doi:10.1130/
High spatial resolution geochemistry and textural characteristics of ‘microtektite’ 0091-7613(1993)021<0211:CCYTIG>2.3.CO;2.
glass spherules in proximal Cretaceous–Paleogene sections: insights into glass Kring, D.A., 2005. Hypervelocity collisions into continental crust composed of sediments
alteration patterns and precursor melt lithologies. Geochim. Cosmochim. Acta 152, and an underlying crystalline basement: comparing the Ries (~ 24 km) and
1–38. https://doi.org/10.1016/j.gca.2014.12.013. Chicxulub (~ 180 km) impact craters. Geochemistry 65 (1), 1–46. https://doi.org/
Blichert-Toft, J., 2001. On the Lu-Hf isotope geochemistry of silicate rocks. Geostand. 10.1016/j.chemer.2004.10.003.
Newslett. 25 (1), 41–56. https://doi.org/10.1111/j.1751-908X.2001.tb00786.x. Kring, D.A., Boynton, W.V., 1992. Petrogenesis of an augite-bearing melt rock in the
Bonamici, C.E., Fanning, C.M., Kozdon, R., Fournelle, J.H., Valley, J.W., 2015. Combined Chicxulub structure and its relationship to K/T impact spherules in Haiti. Nature 358
oxygen-isotope and U-Pb zoning studies of titanite: new criteria for age preservation. (6382), 141–144.
Chem. Geol. 398, 70–84. Kring, D.A., Kramer, G.Y., Collins, G.S., Potter, R.W., Chandnani, M., 2016. Peak-ring
Byrnes, J.S., Karlstrom, L., 2018. Anomalous K-Pg–aged seafloor attributed to impact- structure and kinematics from a multi-disciplinary study of the Schrödinger impact
induced mid-ocean ridge magmatism: Science. Advances 4 (2), eaao2994. https:// basin. Nat. Commun. 7 (1), 1–10. https://doi.org/10.1038/ncomms13161.
doi.org/10.1126/sciadv.aao2994. Kring, D.A., et al., 2020. Probing the hydrothermal system of the Chicxulub impact
Chauvel, C., Blichert-Toft, J., 2001. A hafnium isotope and trace element perspective on crater. Sci. Adv. 6 (22) https://doi.org/10.1126/sciadv.aaz3053.
melting of the depleted mantle. Earth Planet. Sci. Lett. 190, 137–151. https://doi. Krogh, T.E., Kamo, S.L., Bohor, B.A., 1993a. Fingerprinting the K/T impact site and
org/10.1016/S0012-821X(01)00379-X. determining the time of impact by U-Pb dating of single shocked zircons from distal
Chew, D.M., Petrus, J.A., Kamber, B.S., 2014. U–Pb LA–ICPMS dating using accessory ejecta. Earth Planet. Sci. Lett. 119, 425–429. https://doi.org/10.1016/0012-821X
mineral standards with variable common Pb. Chem. Geol. 363, 185–199. (93)90150-8.
Claeys, Ph., Heuschkel, S., Lounejeva-Baturina, E., Sanchez- Rubio, G., Stöffler, D., 2003. Krogh, T.E., Kamo, S.L., Sharpton, V.L., Marin, L.E., Hildebrand, A.R., 1993b. U–Pb ages
The suevite of drill hole Yucatàn 6 in the Chicxulub impact crater. Meteorit. Planet. of single shocked zircons linking distal K/T ejecta to the Chicxulub crater. Nature
Sci. 38, 1299–1317. https://doi.org/10.1111/j.1945-5100.2003.tb00315.x. 366, 731–734. https://doi.org/10.1038/366731a0.
Dallmeyer, R.D., 1984. 40Ar/39Ar ages from a pre-Mesozoic crystalline basement Lopez Ramos, E., 1975. Geological Summary of the Yucatan Peninsula. In: Nairn, A.E.M.,
penetrated at Holes 537 and 538a of the Deep Sea Drilling Project Leg 77, Stehli, F.G. (Eds.), The Gulf of Mexico and the Caribbean. Springer, Boston, MA,
pp. 257–282. https://doi.org/10.1007/978-1-4684-8535-6_7.

10
S.J. de Graaff et al. LITHOS 436-437 (2023) 106953

Mann, P., Rogers, R.D., Gahagan, L., 2007. Overview of plate tectonic history and its Sprain, C.J., Renne, P.R., Clemens, W.A., Wilson, G.P., 2018. Calibration of chron C29r:
unresolved tectonic problems. In: Buncdschud, J. (Ed.), Central America: Geology, New high-precision geochronologic and paleomagnetic constraints from the Hell
Resources and Hazards, Vol. 1. Balkema Publishers, pp. 201–237. Creek region, Montana. GSA Bull. 130 (9–10), 1615–1644.
Martens, U., Weber, B., Valencia, V.A., 2010. U/Pb geochronology of Devonian and older Steiner, M.B., 2005. Pangean reconstruction of the Yucatan block: Its Permian, Triassic,
Paleozoic beds in the southeastern Maya Block, Central America: its affinity with and Jurassic geologic and tectonic history. In: Anderson, T.H., Nourse, J.A.,
peri-Gondwanan terranes. Geol. Soc. Am. Bull. 122, 815–829. https://doi.org/ McKee, J.W., Steiner, M.B. (Eds.), The Mojave-Sonora Megashear Hypothesis:
10.1130/B26405.1. Development, Assessment, and Alternatives: Geological Society of America Special
Marton, G., Buffler, R.T., 1994. Jurassic reconstruction of the Gulf of Mexico Basin. Int. Paper, vol. 393, pp. 457–480. https://doi.org/10.1130/0-8137-2393-0.457.
Geol. Rev. 36, 545–586. https://doi.org/10.1080/00206819409465475. Steiner, M.B., Walker, J.D., 1996. Late Silurian plutons in Yucatan. J. Geophys. Res. Solid
Melosh, H.J., 1989. Impact cratering: A geologic process. In: New York: Oxford Earth 101, 17727–17735. https://doi.org/10.1029/96JB00174.
University Press. Clarendon Press, Oxford, p. 245. Stern, R.J., Dickinson, W.R., 2010. The Gulf of Mexico is a Jurassic back-arc basin.
Molina-Garza, R.S., Van der Voo, R., Urrutia-Fucugauchi, J., 1992. Paleomagnetism of Geosphere 6, 739–754. https://doi.org/10.1130/GES00585.1.
the Chiapas Massif, southern Mexico: evidence for rotation of the Maya block and Stern, R.J., Anthony, E.Y., Ren, M., Lock, B.E., Norton, I., Kimura, J.I., Miyazaki, T.,
implications for the opening of the Gulf of Mexico. Geol. Soc. Am. Bull. 104 (9), Hanyu, T., Chang, Q., Hirahara, Y., 2011. Southern Louisiana salt dome xenoliths:
1156–1168 doi:10.1130/0016-7606(1992)104<1156:POTCMS>2.3.CO;2. first glimpse of Jurassic (ca. 160 Ma) Gulf of Mexico crust. Geology 39, 315–318.
Morgan, J.V., Gulick, S.P.S., et al., 2016. The formation of peak rings in large impact https://doi.org/10.1130/G31635.1.
craters. Science 354 (6314), 878–882. https://doi.org/10.1126/science.aah6561. Sun, S, McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
Morgan, J.V., Gulick, S.P.S., Mellett, C.L., et al., 2017. Chicxulub: Drilling the K-Pg Implications for mantle composition and processes. In: Saunders, A.D., Norry, M.J.
Impact Crater: Proceedings of the International Ocean Discovery Program, volume (Eds.), Magmatism in the Ocean Basins, 42. Geological Society [London] Special
364: College Station, Texas, International Ocean Discovery Program. https://doi. Publication, pp. 313–345. https://doi.org/10.1144/GSL.SP.1989.042.01.19.
org/10.14379/iodp.proc.364.2017. Swisher, C.C., Grajales-Nishimura, J.M., Montanari, A., Margolis, S.V., Claeys, P.,
Muehlberger, W.R., 1992. Tectonic map of North America, Southeast and Southwest Alvarez, W., Renne, P., Cedillo-Pardoa, E., Maurrasse, F.J.R., Curtis, G.H., Smit, J.,
sheets: Tulsa, Oklahoma. Am. Assoc. Petrol. Geol. Scale 1, 5,000,000. 1992. Coeval 40Ar/39Ar ages of 65.0 million years ago from Chicxulub crater melt
Nance, R.D., Linnemann, U., 2008. The Rheic Ocean: origin, evolution, and significance. rock and Cretaceous-Tertiary boundary tektites. Science 257 (5072), 954–958.
GSA Today 18 (12), 4–12 doi.org/10.1130/GSATG24A.1. Tera, F., Wasserburg, G.J., 1972. U-Th-Pb systematics in lunar highland samples from the
Ortega-Gutierrez, F., Ruiz, J., Centeno-Garcia, E., 1995. Oaxaquia, a Proterozoic Luna 20 and Apollo 16 missions. Earth Planet. Sci. Lett. 17 (1), 36–51.
microcontinent accreted to North America during the late Paleozoic. Geology 23, Tuchscherer, M.G., Reimold, W.U., Koeberl, C., Gibson, R.L., de Bruin, D., 2004. First
1127–1130 doi:10.1130/0091-7613(1995)023<1127:OAPMAT>2.3.CO;2. petrographic results on impactites from the Yaxcopoil-1 borehole, Chicxulub
Ortega-Gutiérrez, F., Elías-Herrera, M., Morán-Zenteno, D.J., Solari, L., Weber, B., Luna- structure, Mexico. Meteorit. Planet. Sci. 39, 899–930. https://doi.org/10.1111/
González, L., 2018. The pre-Mesozoic metamorphic basement of Mexico, 1.5 billion j.1945-5100.2004.tb00937.x.
years of crustal evolution. Earth Sci. Rev. 183, 2–37. https://doi.org/10.1016/j. Thomson, S.N., Gehrels, G.E., Ruiz, J., Buchwaldt, R., 2012. Routine low-damage apatite
earscirev.2018.03.006. U-Pb dating using laser ablation–multicollector–ICPMS. Geochemistry, Geophysics,
Ortega-Obregón, C., Solari, L.A., Keppie, J.D., Ortega- Gutiérrez, F., Solé, J., Morán- Geosystems 13.
Ical, S., 2008. Middle-Late Ordovician magmatism and Late Cretaceous collision in Tuchscherer, M.G., Reimold, W.U., Koeberl, C., Gibson, R.L., 2005. Geochemical and
the southern Maya block, Rabinal-Salamá area, central Guatemala: implications for petrographic characteristics of impactites and cretaceous target rocks from the
North America–Caribbean plate tectonics. Geol. Soc. Am. Bull. 120 (5–6), 556–570. Yaxcopoil-1 borehole, Chicxulub impact structure, Mexico: Implications for target
https://doi.org/10.1130/B26238.1. composition. Meteorit. Planet. Sci. 40 (9–10), 1513–1536. https://doi.org/10.1111/
Ortega-Obregón, C., Keppie, J.D., Murphy, J.B., Lee, J.K.W., Ortega-Rivera, A., 2009. j.1945-5100.2005.tb00415.x.
Geology and geochronology of Paleozoic rocks in western Acatlán complex, southern Urrutia-Fucugauchi, J., Marin, L., Trejo-Garcia, A., 1996. UNAM scientific drilling
Mexico: evidence for contiguity across an extruded high-pressure belt and program of Chicxulub impact structure-evidence for a 300 kilometer crater diameter.
constraints on Paleozoic reconstructions. Geol. Soc. Am. Bull. 121, 1678–1694. Geophys. Res. Lett. 23 (13), 1565–1568. https://doi.org/10.1029/96GL01566.
https://doi.org/10.1130/B26597.1. Vellekoop, J., Sluijs, A., Smit, J., Schouten, S., Weijers, J.W., Damsté, J.S.S.,
Paton, C., Hellstrom, J., Paul, B., Woodhead, J.H., Hergt, J., 2011. Iolite: freeware for the Brinkhuis, H., 2014. Rapid short-term cooling following the Chicxulub impact at the
visualization and processing of mass spectrometry data. J. Anal. At. Spectrom. 26, Cretaceous–Paleogene boundary. Proc. Natl. Acad. Sci. 111 (21), 7537–7541.
2508–2518. Vermeesch, P., 2018. IsoplotR: a free and open toolbox for geochronology. Geosci. Front.
Philpotts, A., Ague, J., 2009. Principles of Igneous and Metamorphic Petrology. 9 (5), 1479–1493.
Cambridge University Press. Vermeesch, P.M., Morgan, J.V., 2008. Structural uplift beneath the Chicxulub impact
Pindell, J.L., 1985. Alleghenian reconstruction and subsequent evolution of the Gulf of structure. J. Geophys. Res. Solid Earth 113, B07103. https://doi.org/10.1029/
Mexico, Bahamas, and Proto-Caribbean. Tectonics 4 (1), 1–39. 2007JB005393.
Pindell, J., Dewey, J.F., 1982. Permo-Triassic reconstruction of western Pangea and the Weber, B., Iriondo, A., Premo, W.R., Hecht, L., Schaaf, P., 2007. New insights into the
evolution of the Gulf of Mexico/Caribbean region. Tectonics 1, 179–211. https://doi. history and origin of the southern Maya Block, SE México: U–Pb–SHRIMP zircon
org/10.1029/TC001i002p00179. geochronology from metamorphic rocks of the Chiapas massif. Int. J. Earth Sci. 96,
Pindell, J.L., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and 253–269. https://doi.org/10.1007/s00531-006-0093-7.
northern South America in the mantle reference frame: an update. Geol. Soc. Lond., Weber, B., González-Guzmán, R., Manjarrez-Juárez, R., de León, A.C., Martens, U.,
Spec. Publ. 328 (1), 1–55. Solari, L., Hecht, L., Valencia, V., 2018. Late Mesoproterozoic to early Paleozoic
Pindell, J.L., Miranda, C.E., Cerón, A., Hernandez, L., 2016. Aeromagnetic map history of metamorphic basement from the southeastern Chiapas Massif complex,
constrains Jurassic–Early Cretaceous synrift, break up, and rotational seafloor Mexico, and implications for the evolution of NW Gondwana. Lithos 300–301,
spreading history in the Gulf of Mexico. In: Mesozoic of the Gulf Rim and beyond: 177–199 doi:10.1016/j.lithos.2017.12.009.
New progress in science and exploration of the Gulf of Mexico Basin: SEPM Society Weber, B., Schmitt, A.K., Cisneros de León, A., González-Guzmán, R., 2019. Coeval early
for Sedimentary Geology, 35, pp. 123–153. Ediacaran breakup of Amazonia, Baltica, and Laurentia: evidence from micro-
Renne, P.R., et al., 2013. Time scales of critical events around the Cretaceous-Paleogene baddeleyite dating of dykes from the Novillo Canyon, Mexico. Geophys. Res. Lett. 46
boundary. Science 339, 684–687. https://doi.org/10.1126/science.1230492. (4), 2003–2011. https://doi.org/10.1029/2018GL079976.
Ross, C.H., et al., 2022. Evidence of Carboniferous arc magmatism preserved in the Weber, B., Schmitt, A.K., de León, A.C., González-Guzmán, R., Gerdes, A., 2020.
Chicxulub impact structure. Geol. Soc. Am. Bull. 134 (1–2), 241–260. https://doi. Neoproterozoic extension and the Central Iapetus Magmatic Province in southern
org/10.1130/B35831.1. Mexico–New U-Pb ages, Hf-O isotopes and trace element data of zircon from the
Schmieder, M., Shaulis, B.J., Lapen, T.J., Kring, D.A., 2018. U–Th–Pb systematics in Chiapas Massif Complex. Gondwana Res. 88, 1–20. https://doi.org/10.1016/j.
zircon and apatite from the Chicxulub impact crater, Yucatán, Mexico. Geol. Mag. gr.2020.06.022.
155, 1330–1350. https://doi.org/10.1017/S0016756817000255. Weis, D., Kieffer, B., Maerschalk, C., Barling, J., de Jong, J., Williams, G.A., Hanano, D.,
Schoene, B., Bowring, S.A., 2006. U–Pb systematics of the McClure Mountain syenite: Pretorius, W., Mattielli, N., Scoates, J.S., Goolaerts, A., Friedman, R.M., Mahoney, J.
thermochronological constraints on the age of the 40Ar/39Ar standard MMhb. B., 2006. High-precision isotopic characterization of USGS reference materials by
Contrib. Mineral. Petrol. 151 (5), 615–630. TIMS and MC-ICP-MS. Geochem. Geophys. Geosyst. 7, 1–30. https://doi.org/
Schulte, P., et al., 2010. The Chicxulub asteroid impact and mass extinction at the 10.1029/2006GC001283.
Cretaceous-Paleogene boundary. Science 327, 1214–1218. https://doi.org/10.1126/ Weis, D., Kieffer, B., Hanano, D., Silva, I.N., Barling, J., Pretorius, W., Maerschalk, C.,
science.1177265. Mattielli, N., 2007. Hf isotope compositions of US Geological Survey reference
Sharpton, V.L., Lee, S., Ryder, G., Schuraytz, B.C., 1996. A model of the Chicxulub impact materials. Geochem. Geophys. Geosyst. 8 (6), 1–15. https://doi.org/10.1029/
basin based on evaluation of geophysical data, well logs, and drill core samples. GSA 2006GC001473.
Spec. Pap. 307, 55–74. Zhao, J., et al., 2020. Geochemistry, geochronology and petrogenesis of Maya Block
Simpson, S.L., Osinski, G.R., Longstaffe, F.J., Schmieder, M., Kring, D.A., 2020. granitoids and dikes from the Chicxulub Impact Crater, Gulf of México: Implications
Hydrothermal alteration associated with the Chicxulub impact crater upper peak- for the assembly of Pangea. Gondwana Res. 82, 128–150. https://doi.org/10.1016/j.
ring breccias. Earth and Planetary Science Letters 547, 116425. gr.2019.12.003.

11

You might also like