Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Applied Geochemistry 141 (2022) 105321

Contents lists available at ScienceDirect

Applied Geochemistry
journal homepage: www.elsevier.com/locate/apgeochem

Geochemistry of marine sediments adjacent to the Los Tuxtlas Volcanic


Complex, Gulf of Mexico: Constraints on weathering and provenance
John S. Armstrong-Altrin a, *, Mayla A. Ramos-Vázquez b, Jayagopal Madhavaraju c,
Mario Emilio Marca-Castillo d, María Luisa Machain-Castillo a, Antonio Zoilo Márquez-García e
a
Unidad de Procesos Oceánicos y Costeros, Instituto de Ciencias del Mar y Limnología, Universidad Nacional Autónoma de México, Ciudad Universitaria, Ciudad de
México, 04510, CDMX, Mexico
b
División de Geociencias Aplicadas, Instituto Potosino de Investigación Científica y Tecnológica (IPICYT), Camino a la Presa San José 2055, San Luis Potosí, 78216,
Mexico
c
Estación Regional del Noroeste, Instituto de Geología, Universidad Nacional Autónoma de México, Hermosillo, Sonora, Mexico
d
Posgrado en Ciencias del Mar y Limnología, Universidad Nacional Autónoma de México, Ciudad Universitaria, Ciudad de México, 04510, CDMX, Mexico
e
Universidad Autónoma Metropolitana, Unidad Iztapalapa, Iztapalapa, 09340, CDMX, Mexico

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

Editorial handling by Dr M Liotta The provenance of GoM45 and GoM63 core sediments recovered in the deep-sea area (~1666 m–1672 m water
depth, respectively), in front of the Los Tuxtlas volcanic field, Gulf of Mexico was analyzed based on the textural,
Keywords: mineralogical, and geochemical variations. The radiocarbon age data of Globorotalia menardii group foraminiferal
Deep-sea sediments species varied from 17,709 to 20,711 calibrated years B.P for the sediment core GoM45, which represent
Continental shelf
Pleistocene/Holocene boundary. The core sediments were characterized by detrital components with enriched
Radiocarbon ages
SiO2, Al2O3, and Fe2O3 concentrations. The weathering indices like Chemical Index of Alteration (CIA), Chemical
Drilling
Mineralogy Index of Weathering (CIW), and Plagioclase Index of Alteration (PIA) revealed a low intensity of weathering for
the core sediments.
The V/Cr, Ni/Co, Cu/Zn, and V/Cr elemental ratios indicated an oxic depositional environment for the Gulf of
Mexico deep sea area. In addition, the geochemistry data indicated a passive margin setting, which is consistent
with the geology of the Gulf of Mexico. The trace and rare earth element concentrations, and Eu anomaly
indicated that the sediments were derived from the natural erosion of andesitic and basaltic rocks in the Los
Tuxtlas region, SW Gulf of Mexico. Los Tuxtlas region is partly influenced by the Papaloapan and Coatzacoalcos
Rivers, which discharges into the SW Gulf of Mexico.

1. Introduction processes (Li et al., 2016). The tectonic, climatic, and environmental
changes, affect the source-to-sink sediment transport pathway, which
The provenance of deep-sea sediments can be analyzed through eventually influence the depositional basin (Mason et al., 2017;
textural, mineralogical, and geochemical compositions (Madhavaraju Schwestermann et al., 2021). Recent studies revealed the importance of
et al., 2016a, 2021; Basu, 2020; Kasper-Zubillaga et al., 2021). Tracing sediment composition to infer the provenance as well as to reconstruct
and quantifying the provenance of deep-sea sediments is particularly the paleoclimate of a particular region (Verma and Armstrong-Altrin,
useful to constraint the erosional history and palaeoenvironment (Liu 2013; Anthony et al., 2019; Banerjee et al., 2020; Jamwal et al., 2020;
et al., 2020; Roddaz et al., 2021). The provenance and source-to-sink Krishnan et al., 2020; Chougong et al., 2021; Yang et al., 2021). The
sediment dispersal systems are controlled by the intrabasinal and deep-sea sediments are commonly classified as lithogenous, cosmoge­
extrabasinal sources, lithology, source rock types, and sediment nous, and biogenous. The redox state of both sea water and deep-sea
pathway (Wang et al., 2020; Razum et al., 2021). Knowledge of sedi­ sediments is most important to understand the marine depositional
ment recycling and sediment transfer mechanism from land to sea pro­ environment. Chemical changes in deep-sea sediments are controlled by
vide important constraints on continental weathering and earth surface the oxygen concentration of the sea water. Similarly, the concentration

* Corresponding author.
E-mail addresses: armstrong@cmarl.unam.mx, john_arms@yahoo.com (J.S. Armstrong-Altrin).

https://doi.org/10.1016/j.apgeochem.2022.105321
Received 17 December 2021; Received in revised form 25 April 2022; Accepted 26 April 2022
Available online 29 April 2022
0883-2927/© 2022 Elsevier Ltd. All rights reserved.
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Fig. 1. Map showing core locations in the SW Gulf of Mexico (modified after Anaya-Gregorio et al., 2018).

of metals in deep-sea sediments are controlled by redox conditions of sea sediments recovered from the southern part of the Gulf of Mexico and
water and the availability of trace metals in the marine depositional investigated the provenance and tectonic environment. In addition,
environment (Sato et al., 2012; Meinhardt et al., 2016; Ekoa Bessa et al., heavy metal contamination of estuary (Vera-Mendoza and Sal­
2018; Swain et al., 2021; Rasul and Al-Jaleel, 2022). Chemical compo­ as-de-León, 2014; Botello et al., 2015; Rosales-Hoz et al., 2015; Arm­
sition of deep-sea sediments can be evaluated based on the major and strong-Altrin et al., 2019) and marine sediments near the Coatzacoalcos
trace element concentrations, because compositional changes are area, Gulf of Mexico were investigated by few researchers (Arm­
controlled by the relative proportions of sediment-forming minerals strong-Altrin and Machain-Castillo, 2016; Ramos-Vázquez et al., 2018;
related to their parent rocks. The immobile trace elements like rare earth Ramos-Vázquez and Armstrong-Altrin, 2021a,b; Ayala-Pérez et al.,
elements (REE), Cr, Hf, Nb, Ni, Th, and V are highly reliable indicators to 2021; Mapel-Hernández et al., 2021).
infer sediment provenance due to their resistance against geological Sediments are a potential source of trace metals in marine environ­
processes such as weathering, recycling, and sorting during transport ments and play an important role in transmission of metals (Hill et al.,
(Cullers 2000; Zaid et al., 2015; Madhavaraju et al., 2017; Velmurugan 2013; LeMonte et al., 2017; Horasan et al., 2020). Intermittent rivers
et al., 2019; Subin et al., 2022). In addition, chondrite normalized REE and their flow cessation occur naturally and spatiotemporally in these
patterns and differences in Eu anomalies are extensively utilized in systems. The majority of trace metals originate from natural sources
various studies to infer source rock characteristics (Armstrong-Altrin such as volcanic eruptions, rock weathering, erosion, and atmosphere.
et al., 2017; Madhavaraju et al., 2019; Chaudhuri et al., 2020; Ramír­ Similarly, anthropogenic activities such as agriculture, industry, and
ez-Montoya et al., 2021). extensive urban development in a coastal environment are considered as
There are few studies, which discussed the sediment composition of a threat to the human population (Sundararajan et al., 2017; Jha et al.,
deep-sea sediments in the Gulf of Mexico (Armstrong-Altrin, 2015; 2019; Liu et al., 2019; Shaheen et al., 2019; Tehna et al., 2019; Zhan
Armstrong-Altrin et al., 2015; Ramos-Vázquez et al., 2017; Anaya-­ et al., 2020; Andráš et al., 2021; Armstrong-Altrin et al., 2021a).
Gregorio et al., 2018). These studies analyzed the geochemistry of This study provides a comprehensive major and trace element

2
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

geochemistry data integrated with mineralogy and granulometric


analysis of two sediment cores GoM45 and GoM63 recovered in the SW
Gulf of Mexico. The objective of this study is to investigate the weath­
ering, provenance, and sedimentation rate.

2. Study area

Two sediment cores, GoM45 and GoM63 (~1666 m–1672 m water


depth, respectively; ~5 m long) were recovered 80 km off the coast in
front of Los Tuxtlas volcano, SW Gulf of Mexico (Fig. 1; 19◦ 10′ 12′′ N
94◦ 45′ 64′′ W and 19◦ 09′ 70′′ N 95◦ 09′ 57′′ W, respectively).
The core recovered areas were dominantly influenced by sediments
supplied by the Coatzacoalcos and Papaloapan Rivers. The Coat­
zacoalcos River is one of the major rivers supplying sediments to the
southern part of the Gulf of Mexico, drains between 17◦ 46′ -18◦ 10′ N and
92◦ 25′ -94◦ 31′ W. It is 325 km in length with a catchment area of 17,369
km2, and a mean surface runoff of 28,093 Hm3, which is located in the
coastal plain of the Gulf of Mexico, mostly in the Tabasco and Veracruz
states. The Papaloapan River basin (17◦ 19◦ N and 95◦ –97◦ 40′ W) is the
second largest hydrological basin in Mexico, drains 39,189 km2 of
catchment area before reaching the Gulf of Mexico with an annual
volume of discharge of 39,175 km3. The Papaloapan River flows into the
Gulf of Mexico through the Alvarado lagoon, Veracruz State. The
Papaloapan basin is divided into upper (above 1000 m.a.s.l.) and lower
Papaloapan. Los Tuxtlas area in Veracruz, forms part of the lower
Papaloapan, even though it includes an area higher than 1000 m.a.s.l.
(Tamayo, 1991). The Veracruz state is considered as one of the intensive
industrialized areas, second in the manufacture of iron and steel, and it
is fourth in the extraction of oil and gas (Jiménez-Orocio et al., 2015;
Rosales-Hoz et al., 2015; Cancino-Solórzano et al., 2016). There are
river tributaries originating from Tuxtla, Santa Marta, and San Martín
volcanoes, which spill their waters directly into the SW Gulf of Mexico.
Fig. 2. Photograph showing GoM45 core section recovered from the SW Gulf
2.1. Geology of Mexico.

Los Tuxtlas Volcanic Complex consists of various monogenetic


3. Methodology
erupted cones in the region between San Martín and Catemaco, Veracruz
State (Fig. 1). The volcanoes are younger in age, mostly less than 4000
The GoM45 and GoM63 sediment cores (~1666 m–1672 m water
years old (Ruiz-Mendoza et al., 2021; Verma et al., 2021). The area is
depth, respectively; ~5 m long) were retrieved on-board the research
almost completely covered by pyroclastic deposits and lava flows, and
vessel “Justo Sierra” on June 2017, which belongs to National Autono­
with a low proportion of marine sediments of Tertiary age. Other than
mous University of Mexico (UNAM). The recovered sediment cores were
the volcanic rocks, sedimentary rocks, such as limestone, clay and
sub-sampled at every cm for various analyses and kept in the Sedi­
sandstone are also present in the Los Tuxtlas area (Verma, 2000) (Fig. 1).
mentology Laboratory, ICML, UNAM (Fig. 2). The samples were dried
In general, the outcrops of the river draining areas in the central part of
for 24 h at 40 ◦ C and powdered using agate mortar. The sediment
the western Gulf of Mexico are composed of 1) alluvium and soils
samples were analyzed for 14C planktonic foraminifera, particle size
(Quaternary age), 2) volcanic rocks of mafic and intermediate types
distribution, mineralogy, and geochemistry including major, trace, and
(Cenozoic age), 3) clastic and calcareous sedimentary rocks (Cenozoic
rare earth element concentrations.
and Mesozoic age), and 4) metamorphic rocks comprising schists and
gneisses (Paleozoic and Precambrian age). Volcanic rocks are dominated
3.1. Foraminiferal analysis
by Miocene-Pliocene andesites of sub-alkaline, mafic, and intermediate
compositions (Verma, 2001; Armstrong-Altrin et al., 2021b) (Fig. 1).
Samples for foraminiferal analysis were sieved through an American
The volcanic units of the study areas belong to the overlap region of the
Society for Testing Materials (ASTM) sieve No. 230 (<62 μm), then
Mexican Volcanic Belt (MVB) and the Eastern Alkaline Province (Verma,
washed with running water, dried and picked under a stereoscopic mi­
2000).
croscope. All planktonic species were determined and counted based on
the methodology proposed by Martin and Liddell (1989). For biostrati­
2.2. Hydrology and oceanography
graphic analysis Globorotalia menardii group graphics were constructed.
For 14C analysis three samples (3 mg mixed planktonic foraminifera
The anticyclonic eddies that are separated episodically from the Loop
each) were sent to Beta Analytic Testing Laboratory for Accelerator Mass
current in the eastern Gulf of Mexico travel to the western limit of the
Spectrometer (AMS) dating. The conversion of 14C radiocarbon ages into
Gulf of Mexico, interacting with numerous accompanying cyclones and
calendar years BP was performed using the CALIB software (Stuiver and
significantly contributing to the variability of the upper layer (above
Reimer, 1993, version 5.0.2; http://calib.qub.ac.uk/calib). After cali­
500–1000 m depth) in the central and western regions. During summer,
bration, ages between dated sections were obtained by linear interpo­
when the winds over the Tamaulipas and Veracruz platforms have a
lation. The sedimentation rate in the SW Gulf of Mexico deep sea area
northward component, the currents are predominantly in that direction
was reported as ~0.0166 cm/year (Ramos-Vázquez et al., 2018).
and confluence occurs near the border with the United States (Mon­
real-Gómez et al., 1992).

3
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

14
Fig. 3. C ages of Globorotalia menardii planktonic foraminifera in the GoM45 and GoM63 sediment cores from the SW Gulf of Mexico.

3.2. Granulometric analysis 1000 ◦ C for 2 h to determine loss on ignition (LOI). Lithium tetraborate
was mixed with the samples and heated to 1000 ◦ C to form a fused
The textural characteristics were analyzed by Beckman Coulter sample for XRF analysis with a Rigaku RIX-3000 equipped with a Rh
particle size analyzer (Coulter LS230), at the Sedimentology Laboratory, tube. Calibration curves were prepared by using international reference
ICML, UNAM (number of samples n = 20). Grain sizes were measured materials and the geochemical standard JGB1 (GSJ) was used to cali­
based on the classification scale proposed by Wentworth (1922), as well brate data quality. The analytical accuracy was better than 5%.
as described by Rząsa and Owczarzak (2013). The textural parameters:
grain size, sorting, skewness, and kurtosis were calculated by the 3.4.2. Trace and rare earth element concentrations
equations proposed by Folk and Ward (1957). The trace and rare earth element concentrations were quantified by
an Ultratrace - total acid digestion method using ICP-OES/ICP-MS at the
3.3. Mineralogy Activation Laboratories Ltd., Canada. All samples were determined in
triplicate and the analytical precision was described by relative standard
In order to detect the mineralogical composition, 5 samples from deviation (RSD), with a range of 5%–10%. The standard reference ma­
GoM45 and 5 from GoM63 were selected. The mineralogy of bulk sed­ terials used were SY-2 (syenite CCRMP, Canada) and GSR-4 (sediment
iments was determined by a Siemens D5000 X-ray Diffractometer IGGE, China). The REE data were normalized against chondrite
(XRD). In addition, the chemistry of selected minerals was detected by normalization factors listed in McDonough and Sun (1995) and Gromet
the PHILLIPS XL-30 scanning electron microscope (SEM) equipped with et al. (1984). The upper continental crust (UCC) values are from Taylor
EDAX spectrometer (EDS) system, located at the Institute of Geophysics, and McLennan (1985). Eu anomaly (Eu/Eu*) was calculated based on
UNAM, Mexico. the equation Eu/Eu* = EuCN/[(SmCN)*(GdCN)]1/2, where CN represents
the chondrite normalized values.
Chemical index of alteration (CIA; Nesbitt and Young, 1982),
3.4. Geochemistry Chemical Index of Weathering (CIW; Harnois, 1988), and Plagioclase
Index of Alteration (PIA; Fedo et al., 1995) are calculated based on the
For major, trace and REE geochemistry, 20 sediment samples were following equations. CIA = [Al2O3/(Al2O3 + CaO* + Na2O* + K2O)] *
selected at various intervals, i.e. 10 samples from GoM45 and 10 from 100, CIW = [Al2O3/(Al2O3 + CaO* + Na2O*)] * 100, PIA =
GoM63 cores. [Al2O3–K2O/(Al2O3 + CaO* + Na2O*)] * 100, CaO* and Na2O* = Ca
and Na in silicate fractions.
3.4.1. Major element concentrations
For major element analysis, 20 g of sediment samples were powdered
by an agate mortar and were sieved through an American Society for
Testing Materials (ASTM) sieve No. 230 (<62 μm). Powdered samples
were heated to 110 ◦ C for 6 h followed by heating in a muffle furnace at

4
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Table 1 4. Results
Textural parameters with mean grain size (Mz) for the GoM45 and GoM63
sediment cores from the Gulf of Mexico (after Folk and Ward, 1957). 4.1. Age assessment
Core GoM45

Interval (cm) Sand Silt Clay Mz (φ) (φ) SkI KG


For AMS 14C dates, CALIB software (Stuiver and Reimer, 1993,
version 5.0.2) was used to convert to calendar years BP. GoM45 reached
9–10 0.03 52.95 47.02 7.94 1.60 0.03 0.94
the Holocene/Pleistocene boundary around 180 cms, 20,711 cal yrs BP
40–41 0.00 46.50 53.50 8.19 1.52 0.02 0.92
110–111 0.00 49.85 50.15 8.06 1.57 0.02 0.90 at 330 cm, and 17,709 cal yrs BP at 478 cm; which reveals a block of
160–161 1.05 61.77 37.18 7.44 1.81 0.00 0.98 older sediments between 478 cm and 210 cm.
230–231 0.00 49.09 50.91 8.03 1.66 − 0.03 0.01 Armstrong-Altrin et al. (2015) reported an age of 26,500 years BP for
269–270 0.10 56.89 43.01 7.66 1.78 − 0.04 0.93 the nearby 5 m long core (GoM61). Similar age was also reported for
329–330 0.00 55.76 44.24 7.78 1.65 0.03 0.89
369–370 0.11 54.44 45.45 7.76 1.80 − 0.06 0.93
cores collected at the southwestern Gulf of Mexico (Anaya-Gregorio
440–441 0.00 48.30 51.70 8.07 1.64 − 0.02 0.89 et al., 2018; Machain-Castillo et al., 2020). Stratigraphic correlation
460–461 0.96 57.98 41.06 7.42 2.01 − 0.06 0.84 with Globorotalia menardii group for the 3 cores (Fig. 3) showed their
Core GoM63 base within the Upper Pleistocene and the Pleistocene/Holocene
Interval (cm) Sand Silt Clay Mz (φ) (φ) SkI KG
boundary at 150 cm, 190 cm, and 180 cm (for cores GoM61, GoM63, and
9–10 0.29 64.39 35.31 7.42 1.72 0.07 0.95
40–41 0.00 54.52 45.48 7.88 1.62 0.05 0.91 GoM45 respectively).
110–111 0.00 47.61 52.39 8.14 1.55 0.01 0.91
160–161 0.00 46.11 53.89 8.14 1.62 − 0.06 0.94
4.2. Granulometry
230–231 0.00 51.58 48.42 7.91 1.71 − 0.02 0.83
269–270 0.83 57.92 41.25 7.49 1.93 − 0.07 0.91
329–330 0.00 52.24 47.76 7.91 1.65 0.01 0.87 The textural parameters are calculated based on the methodology
380–381 0.47 55.79 43.73 7.68 1.83 − 0.04 0.92 proposed by Folk and Ward (1957) (Table 1). The sediments are very
440–441 0.00 50.61 49.39 7.56 1.69 − 0.02 0.86
fine (clay and silt; < 63 μm) in both cores and the amount of fine fraction
460–461 0.54 56.41 43.05 7.59 1.92 − 0.06 0.89
in most of the samples is >80%. The standard deviation in phi (φ) scale
varies from ~7.4 to 8.2 and ~7.4 to 8.1 for the GoM45 and GoM63 core
sediments, respectively. The variations in skewness and kurtosis

Fig. 4. SEM-EDS analysis for GoM45 core sediments. A) quartz and feldspar, B) barite, C) and D) magnetite, E) ilmenite, and F) cassiterite.

5
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Fig. 5. SEM-EDS analysis for GoM63 core sediments. A) quartz, B) barite, C) zircon, and D) kaolinite.

Fig. 7. Major element concentrations of GoM45 and GoM63 core sediments


normalized against average upper continental crust values (Taylor and
Fig. 6. Geochemical classification of GoM45 an GoM63 core sediments based McLennan 1985).
on the log (SiO2/Al2O3) and log(Fe2O3/K2O) ratios (Herron, 1988).

4.4. Geochemistry
between the two core sediments are not significant (Table 1).
The major element concentrations of GoM45 and GoM63 sediment
4.3. Mineralogy
cores are listed in Table 1S (supplementary material). The SiO2 content
ranges from ~42.8 to 46.5 wt% and ~43.6 to 45.5 wt% in GoM45 and
At intervals between 269-270 cm and 460–461 cm, biogenic and
GoM63, respectively. The average Al2O3 content in GoM45 and GoM63
angular lithic grains are identified. In core GoM45, SEM-EDS reveals an
are similar (the mean with a standard deviation value is 13 ± 0.4 and 13
enrichment in Si, K, Na, Ca, Ba, Fe, Cr, Ti, and Sn contents (Fig. 4A–F).
± 0.3, respectively). Also, the variation in the distribution of Fe2O3,
This enrichment indicates the presence of minerals like quartz (Fig. 4A),
CaO, K2O, Na2O, and MgO in both cores are not significant (Table 1S).
feldspar (Fig. 4A), barite (Fig. 4B), magnetite (Fig. 4C and D), ilmenite
Based on the geochemical classification diagram of Herron (1988) the
(Fig. 4E), and cassiterite (Fig. 4F). Similarly, GoM63 sediments are
GoM45 and GoM63 core sediments are classified as shale (Fig. 6).
elevated with Si, K, Na, Ca, Al, Fe, Ba, and Zr contents, indicating the
The major element contents in core sediments are normalized against
presence of minerals such as quartz (Fig. 5A), barite (Fig. 5B), zircon
the average UCC values (Fig. 7). The concentration of MnO is higher in
(Fig. 5C), and kaolinite (Fig. 5D).
the core sediments than UCC, which is probably due to the concentration
of manganosite. Numerous studies addressed the occurrence of man­
ganosite in the Molango Upper Jurassic carbonate platform, located near
to the coastal region of the SW Gulf of Mexico (Okita, 1992; Okita and

6
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Fig. 8. Trace element concentrations of GoM45 and GoM63 core sediments


normalized against average upper continental crust values (Taylor and
McLennan 1985).

Fig. 10. Al2O3 - (CaO + Na2O) – K2O (A–CN–K, in molecular proportion;


Nesbitt and Young, 1982) ternary diagram to infer the intensity of weathering.
Average concentration of upper continental crust (UCC; Taylor and McLennan,
1985) is plotted for comparison.

REE patterns of the core sediments are enriched in LREE and depleted in
HREE (Fig. 9). The Eu/Eu* ratios in the GoM45 and GoM63 core sedi­
ments vary between ~ 0.63–0.76 and ~ 0.64–0.74, respectively.

4.5. Elemental variations

The correlations between SiO2 versus K2O and Na2O in GoM45 (r =


0.76 and − 0.07, respectively) and GoM63 cores (r = 0.68 and r = 0.60,
respectively) are statistically significant, indicating the predominance of
aluminosilicates. The correlation of Al2O3 versus Rb and Ba for GoM45
(r = − 0.32 and 0.04, respectively, n = 10) and GoM63 (r = − 0.42 and r
= − 0.42, respectively, n = 10) are not statistically significant, indicating
Fig. 9. Chondrite-normalized REE patterns for GoM45 and GoM63 core sedi­ that these elements are related to accessory phases (Armstrong-Altrin,
ments (Taylor and McLennan, 1985). 1 this study; 2 UCC (Taylor and McLennan, 2020; Omar et al., 2021). Similarly, the correlation between Th and Zr is
1985). n = number of samples. statistically significant, but is not significant for Th versus Al2O3 for
GoM45 (r = 0.42 and r = − 0.59, respectively) and GoM63 (r = 0.77 and
Shanks, 1992; Johnson et al., 2016; Anaya-Gregorio et al., 2018). Slight r = − 0.46, respectively), which implies that its distribution is associated
enrichment in TiO2, Al2O3, Fe2O3, MgO, and CaO contents relative to with accessory phases such as zircon (Nagarajan et al., 2015; Ekoa Bessa
UCC is also noted (Fig. 7). et al., 2021a; b, c). A negative correlation of Al2O3 versus Zr, Y, and Nb
The trace element concentrations in the GoM45 and GoM63 core for GoM45 (r = − 0.43, − 0.61 and 0.49, respectively) and GoM63 (r =
sediments are listed in Table 2S (supplementary material). The trace − 0.50, − 0.56 and − 0.15, respectively) also suggests that these elements
element patterns normalized against UCC are similar between GoM45 are integrated in accessory minerals (Bertrand et al., 2011; Hernán­
and GoM63 core sediments, except an enrichment in Cu content in two dez-Hinojosa et al., 2018; Janpou et al., 2022). The correlation of Al2O3
GoM45 samples (Fig. 8). High field strength elements (HFSE) like Zr and against V, Sc, and Cr for GoM45 (r = 0.47, − 0.44, and 0.26, respectively)
Hf in GoM45 (92.3–127.7 ppm and 2.2–3.4 ppm, respectively) and and GoM63 (r = 0.26, − 0.38, and 0.03, respectively) indicates that these
GoM63 (94.5–152.8 ppm and 2.2–3.7 ppm) are lower than in UCC (190 elements are not completely associated with phyllosilicates. Similarly,
and 5.8 ppm, respectively). Y content in a few samples from GoM63 are the correlation between Al2O3 and Ba and Sr for GoM45 (r = 0.17 and
slightly enriched compared with UCC. The large ion lithophile elements − 0.54, respectively) and GoM63 (r = − 0.42 and − 0.84, respectively)
(LILE; Ba, Rb, and Sr) in GoM45 (299–417 ppm, 81.4–107 ppm, and are not significant, suggesting their concentrations are not controlled by
315–561 ppm, respectively) and GoM63 (290–429 ppm, 72.6–104 ppm, clay minerals.
and 245–586 ppm, respectively) are almost similar. In addition, Sr The ΣREE content in GoM45 and GoM63 show a statistically signif­
content in few samples are slightly enriched compared with UCC. icant correlation against Zr (r = 0.65 and r = 0.88, respectively; n = 10),
Compared with UCC, the transitional trace elements (TTE) such as Cr whereas it is not significant against Al2O3 (r = − 0.56 and r = − 0.51,
and Ni contents are generally lower (<100 ppm), whereas V, Cu, Co, and respectively, n = 10), which suggests that REEs are linked with detrital
Sc contents in GoM45 and GoM63 are slightly enriched in few samples minerals like zircon rather than clay minerals (Yu et al., 2016; Wang
(Fig. 8; Table 2S). et al., 2018). Similarly, a significant correlation observed between REE
The REE content in GoM45 and GoM63 core sediments are listed in and Y for the GoM45 (r = 0.94) and GoM63 (r = 0.96) also imply that
Table 3S (supplementary material). The variations in REE content be­ REEs are probably associated with accessory minerals. In general, the
tween GoM45 and GoM63 are not significant. The chondrite normalized elemental variations in sediments are indicating the association of major
and trace element concentrations with accessory detrital phases rather

7
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Fig. 11. Major element based discriminant function diagram to infer sediment provenance (Roser and Korsch, 1988). The discriminant functions are:
Discriminant Function 1 = (− 1.773*TiO2) + (0.607*Al2O3) + (0.760*Fe2O3) + (− 1.500*MgO) + (0.616*CaO) + (0.509*Na2O) + (− 1.224*K2O) + (− 9.090);
Discriminant Function 2 = (0.445*TiO2) + (0.070*Al2O3) + (− 0.250*Fe2O3) + (− 1.142*MgO) + (0.438*CaO) + (1.475*Na2O) + (1.426*K2O) + (− 6.861) Data of
Papaloapan River sediments are from Armstrong-Altrin et al. (2015).

than phyllosilicates. The SiO2/Al2O3, Al2O3/Na2O, and K2O/Na2O ratios are used in
various studies to infer the textural maturity of sediments and high
5. Discussion values (>6, >5, and >1, respectively) indicate a high compositional
maturity (Cox et al., 1995; Madhavaraju et al., 2020; Al-Juboury et al.,
5.1. Sediment weathering and recycling 2021). The low SiO2/Al2O3, Al2O3/Na2O, and K2O/Na2O values in
GoM45 (~3.2–3.5, ~4.9–6.1 and ~0.65–0.95, respectively) and GoM63
The CIA and CIW values in GoM45 (~42–48 and ~44–53, respec­ (~3.3–3.5, ~4.8–6.1, and ~0.66–0.87, respectively) are also indicating
tively) and GoM63 (~42–47 and ~44–51, respectively) indicate low a low compositional maturity.
intensity of weathering in the source area. CIA values for the GoM45 and The accumulation of zircon in sediments increase the HREE content
GoM63 core samples are also plotted in the Al2O3 – (CaO + Na2O) – K2O in relation to LREE and decrease the LaNASC/YbNASC ratio less than 2
(A–CN–K) ternary diagram (Fig. 10), which permits to understand the (Tawfik et al., 2018). The correlation between Zr and LaNASC/YbNASC is
compositional variations (Fedo et al., 1995). On the A-C-NK diagram, statistically not significant for GoM45 (r = − 0.16) and GoM63 (r =
the samples are plotted together, indicating the intensity of weathering − 0.30), which indicates a moderate degree of sorting during sediment
is similar for the GoM45 and GoM63 core samples. The PIA is another deposition. In addition, concentration of minerals such as monazite and
method to measure the intensity of weathering in clastic sediments. Less allanite in sediments increase the P2O5 and LREE contents, which can
weathered sediments have a value < 50, and depending on the intensity indicate the degree of sorting and sediment recycling. Concentration of
of weathering the PIA value will increase (Fedo et al., 1995). The PIA these minerals in sediments will significantly increase the (Gd/Yb)NASC
values of GoM45 and GoM63 vary from ~40.4 to 48.06 and ~40.37 to ratio, higher than the NASC (>2). The (Gd/Yb)NASC ratio in GoM45
46.4, respectively, which reveals a low intensity of weathering.

Fig. 13. Chondrite-normalized rare earth element (REE) patterns of core sed­
iments are compared with probable source rocks from the southern Gulf of
Mexico. 1 this study; 2,3,4 andesite, basaltic andesite, and basalt are compiled
Fig. 12. Ni–Th*10-V ternary diagram (Bracciali et al., 2007) for the GoM45 from Rosales-Lagarde et al. (2005), Verma (2001), Torres-Sánchez et al. (2019),
and GoM63 core sediments. Average compositions of dacite, andesite, and Verma et al. (2020, 2021); 5 Taylor and McLennan (1985). n = number
basalt from the adjacent area are also plotted for comparison. of samples.

8
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Fig. 14. Major element (M)-based multidimensional tectonic discriminant function diagram for the discrimination of active (A) and passive (P) margin settings
(Verma and Armstrong-Altrin, 2016). The function (DF(A-P) M) is to be calculated from the equation DF(AP)M = (3.0005 * ilr1TiM) + (− 2.8243 * ilr2AlM) + (− 1.0596 *
ilr3FeM) + (− 0.7056 * ilr4MnM) + (− 0.3044 * ilr5MgM) + (0.6277 * ilr6CaM) + (− 1.1838 * ilr7NaM) + (1.5915 * ilr8KM) + (0.1526 * ilr9PM) − 5.9948.

varies from ~1.31 to 1.90 and for GoM63 it varies from ~1.29 to 1.19, considered as a redox indicator, the V/Cr ratio >4.25 represents an
indicating moderate sorting and recycling. anoxic depositional condition, while <2 are indicative of a more
oxidizing condition. Jones and Manning (1994) also suggested that
5.2. Provenance Ni/Co ratio <5 indicate oxic, whereas ratios between 5.0 and 7.0 suggest
dysoxic conditions, and ratios above 7.0 indicative of suboxic and anoxic
Textural classification reveals that the GoM45 and GoM63 sediments environments. Similarly, Cu/Zn ratios increase from oxic (<1) to redox
are mostly silt and clayey silt types. The major element-based discrim­ (>1) depositional conditions (Hallberg, 1976). The V/Cr, Ni/Co, and
inant function diagram of Roser and Korsch (1988) is widely used in Cu/Zn ratios from GoM45 (1.63 ± 0.20, 2.26 ± 1.19 and 0.59 ± 0.60,
various studies to infer sediment provenance (Madhavaraju et al., respectively) and GoM63 cores (1.75 ± 0.32, 1.41 ± 0.99 and 0.65 ±
2016a, b; Ramos-Vázquez et al., 2018; Ngueutchoua et al., 2019; Arm­ 0.31, respectively) indicate an oxic depositional condition.
strong-Altrin et al., 2019, 2021a; Venu and Velmayil, 2021). This dia­
gram indicated an intermediate rock provenance for the GoM45 and 6. Conclusions
GoM63 core sediments (Fig. 11). On the other hand, the Al2O3/TiO2
ratio is widely used in various studies to infer source rock composition The mineralogy and geochemistry of the GoM45 and GoM63 core
(Armstrong-Altrin et al., 2018). According to Hayashi et al. (1997) the sediments in the SW Gulf of Mexico are analyzed. The core sediments are
Al2O3/TiO2 ratio varies from ~3 to 8 for mafic, ~ 8–21 for intermediate dominated by silt and clay. The XRD data indicates that the distribution
and ~21–70 for felsic igneous rocks. The Al2O3/TiO2 ratio in GoM45 of minerals in the GoM45 and GoM63 are similar and contains minerals
and GoM63 varies from 19 to 21, which indicates that the core sedi­ such as barite, gibbsite, kaolinite, magnetite, plagioclase, and chlorite.
ments were derived from intermediate source rock like andesite. The CIA, CIW, and PIA values of the core sediments indicate low
The Cr, Ni, and V contents are very low, which could be the result of weathering in the source area. The major and trace element concen­
the contribution of sediments derived from the intermediate source trations indicate the intermediate igneous rocks such as andesite and
rocks, probably andesite (supplementary Table 2S). This inference is basaltic andesite are the source rocks for the deep-sea sediments. The
further supported by the Ni–Th*10-V ternary diagram (Fig. 12; Bracciali trace elements and their elemental ratios suggest that the deep-sea
et al., 2007). Similarly, the Th/Co vs La/Sc diagram also suggests an sediments are deposited in an oxic environment. The chondrite
intermediate rock as a possible source rock for the core sediments. normalized REE patterns and Eu anomaly also suggest that the sedi­
Furthermore, mafic rocks usually contain high LREE/HREE ratio with ments were derived from the andesite and basaltic andesitic rocks
positive Eu anomaly, whereas felsic rocks usually contain low located along the SW Gulf of Mexico coastal region. Major element
LREE/HREE ratio with high negative Eu anomaly (Cullers, 2000). Low concentrations indicate deposition of sediments in a passive margin,
LREE/HREE ratio with a small negative Eu anomaly is observed in the which is consistent with the geology of the Gulf of Mexico.
GoM45 and GoM63 core sediments, indicating that the sediments were
mostly derived from the intermediate igneous rocks. In addition, to infer CRediT authorship contribution statement
the provenance, the REE data of andesite, basalt, and basaltic andesites
exposed along the coastal areas of the southern Gulf of Mexico are John S. Armstrong-Altrin: Writing - original draft, Formal analysis,
compiled from literatures (Rosales-Lagarde et al., 2005; Verma, 2001; Resources, Funding acquisition. Mayla A. Ramos-Vázquez: Investiga­
Torres-Sánchez et al., 2019; Verma et al., 2020, 2021). The chondrite tion, Writing - review and editing. Jayagopal Madhavaraju: Methodol­
normalized REE patterns and Eu anomaly of core sediments are com­ ogy, Formal analysis, editing. Mario Emilio Marca-Castillo: Logistic
parable with andesite and basaltic andesites (Fig. 13). This suggests that fieldwork, Data curation, Methodology, Validation, Formal analysis.
the intermediate igneous rocks like andesite and basaltic andesite are María Luisa Machain-Castillo: Logistic fieldwork, Methodology, Formal
the probable source rocks for the GoM45 and GoM63 core sediments. analysis, review and editing. Antonio Z. Márquez-García: Methodology,
The geochemical composition of detrital sediments is reliable to infer Formal analysis.
the tectonic setting of the source area (Madhavaraju et al., 2020; Critelli
et al., 2021; Roddaz et al., 2021). The tectonic discriminant function Declaration of competing interest
diagrams based on major oxides proposed by Verma and
Armstrong-Altrin (2016) is applied to determine the tectonic setting. On The authors declare that they have no known competing financial
this diagram the samples are plotted in the passive margin field, indi­ interests or personal relationships that could have appeared to influence
cating a passive margin setting for the SW Gulf of Mexico sediments the work reported in this paper.
(Fig. 14).
Acknowledgement
5.3. Depositional environment
Ramos-Vázquez is thankful to CONACyT for the postdoctoral schol­
Trace element contents in clastic sediments, particular elemental arship (CVU: 595593). Armstrong-Altrin is grateful to “Consejo Nacional
ratios like V/Cr, Ni/Co, and Cu/Zn are highly useful to infer the depo­ de Ciencia y Technología” (CONACyT; no: A1-S-21287) and PAPIIT
sitional condition (Jones and Manning, 1994). The V/Cr ratio is projects (no: IN107020) for financial assistance. We extend our sincere

9
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

thanks to Patricia Girón García, Carlos Linares-López, Laura E. Gómez Johnsson, M.J. (Eds.), Sedimentary Provenance and Petrogenesis: Perspectives from
Petrography and Geochemistry, vol. 420, pp. 73–93. Geol. Soc. Amer. Special Paper.
Lizárraga, and Teodoro Hernández Treviño for analytical help. We are
Cancino-Solórzano, Y., Paredes-Sánchez, J.P., Gutiérrez-Trashorras, A.J., Xiberta-
grateful to our laboratory assistants Ricardo M. Domínguez, Eduardo Bernat, J., 2016. The development of renewable energy resources in the State of
Morales la Garza, Susana Santiago, and Arturo Ronquillo Arvizu for Veracruz, Mexico. Util. Pol. 39, 1–4.
powdering sediment samples. We are also grateful to the ICML Institu­ Chaudhuri, A., Banerjee, S., Prabhakar, N., Das, A., 2020. The use of heavy mineral
chemistry in reconstructing provenance: a case study from Mesozoic sandstones of
tional (no. 616) and FACIES-PEMEX-PEP (no: 420401851) projects for Kutch Basin, India. Geol. J. 55 (12), 7807–7817.
financial support for sample collection. Chougong, D.T., Bessa, A.Z.E., Ngueutchoua, G., Yongue, R.F., Ntyam, S.C., Armstrong-
Altrin, J.S., 2021. Mineralogy and geochemistry of Lobé River sediments, SW
Cameroon: implications for provenance and weathering. J. Afr. Earth Sci. 183,
Appendix A. Supplementary data 104320, 1-19.
Cox, R., Lowe, D.R., Cullers, R.L., 1995. The influence of sediment recycling and
Supplementary data to this article can be found online at https://doi. basement composition on evolution of mudrock chemistry in the southwestern
United States. Geochem. Cosmochim. Acta 59, 2919–2940.
org/10.1016/j.apgeochem.2022.105321. Critelli, S., Martín-Martín, M., Capobianco, W., Perri, F., 2021. Sedimentary history and
palaeogeography of the cenozoic clastic wedges of the malaguide comple, internal
References betic cordillera, southern Spain. Mar. Petrol. Geol. 124, 104775.
Cullers, R.L., 2000. The geochemistry of shales, siltstones and sandstones of
Pennsylvanian-Permian age, Colorado, U.S.A.: implications for provenance and
Al-Juboury, A., Hussain, S.H., Al-Lhaebi, S.H., 2021. Geochemistry and mineralogy of the
metamorphic studies. Lithos 51, 305–327.
Silurian Akkas Formation, Iraqi western desert: implications for palaeoweathering,
Ekoa Bessa, A.Z., El-Amier, Y.A., Doumo, E.P.E., Ngueutchoua, G., 2018. Assessment of
provenance and tectonic setting. Arabian J. Geosci. 14, 760.
Sediments Pollution by Trace Metals in the Moloundou Swamp, Southeast
Anaya-Gregorio, A., Armstrong-Altrin, J.S., Machain-Castillo, M.L., Montiel-García, P.C.,
Cameroon. Annu. Res. Revue Biol. 30, 1–13.
Ramos-Vázquez, M.A., 2018. Textural and geochemical characteristics of late
Ekoa Bessa, A.Z., Nguetchoua, G., Janpou, A.K., El-Amier, Y.A., Nguetnga, O.N.N.M.,
Pleistocene to Holocene fine-grained deep-sea sediment cores (GM6 and GM7),
Kayou, U.R., Bisse, S.B., Mapuna, E.C.N., Armstrong-Altrin, J.S., 2021a. Heavy metal
recovered from southwestern Gulf of Mexico. J. Palaeogeogr. 7 (3), 253–271.
contamination and its ecological risks in the beach sediments along the Atlantic
Andráš, P., Midula, P., Matos, J.X., Buccheri, G., Drímal, M., Dirner, V., Melichová, Z.,
Ocean (Limbe coastal fringes, Cameron). Earth Sys. Environ. 5, 433–444.
Ingrid, T., 2021. Comparison of soil contamination at the selected European copper
Ekoa Bessa, A.Z., Paul-Désiré, N., Fuh, G.C., Armstrong-Altrin, J.S., Betsi, T.B., 2021b.
mines. Carpathian J. Earth Environ. Sci. 65 (1), 163–174.
Mineralogy and geochemistry of the Ossa lake Complex sediments, Southern
Anthony, E.J., Almar, R., Besset, M., Reyns, J., Laibi, R., Ranasinghe, R., Abessolo
Cameroon: Implications for paleoweathering and provenance. Arabian J. Geosci. 14.
Ondoua, G., Vacchi, M., 2019. Response of the Bright of Benin (Gulf of Guinea, West
Article no. 322.
Africa) coastline to anthropogenic and natural forcing, Part 2: sources and patterns
Ekoa Bessa, A.Z., Armstrong-Altrin, J.S., Fuh, G.C., Betsi, T.B., Kelepile, T., Ndjigui, P.-D.,
of sediment supply, sediment cells, and recent shoreline change. Continent. Shelf
2021c. Mineralogy and geochemistry of the Ngaoundaba Crater Lake sediments,
Res. 173, 93–103.
northern Cameroon: implications for provenance and trace metals status. Acta
Armstrong-Altrin, J.S., 2015. Evaluation of two multi-dimensional discrimination
Geochim 40, 718–738.
diagrams from beach and deep-sea sediments from the Gulf of Mexico and their
Fedo, C.M., Wayne, N.H., Young, G.M., 1995. Unraveling the effects of potassium
application to Precambrian clastic sedimentary rocks. Int. Geol. Rev. 57 (11–12),
metasomatism in sedimentary rocks and paleosols, with implications for
1446–1461.
paleoweathering conditions and provenance. Geology 23, 921–924.
Armstrong-Altrin, J.S., 2020. Detrital zircon U-Pb geochronology and geochemistry of
Folk, R.L., Ward, W.C., 1957. Brazos River bar, a study in the significance of grain-size
the Riachuelos and Palma Sola beach sediments, Veracruz State, Gulf of Mexico: a
parameters. J. Sediment. Petrol. 27, 3–26.
new insight on palaeoenvironment. J. Palaeogeogr. 9 (4), 28.
Gromet, L.P., Haskin, L.A., Korotev, R.L., Dymek, K.F., 1984. The “north American shale
Armstrong-Altrin, J.S., Machain-Castillo, M.L., 2016. Mineralogy, geochemistry, and
composite”: Its compilation, major and trace element characteristics. Geochem.
radiocarbon ages of deep-sea sediments from the Gulf of Mexico, Mexico. J. S. Am.
Cosmochim. Acta 48 (12), 2469–2482.
Earth Sci. 71, 182–200.
Hallberg, R.O., 1976. A geochemical method for investigation of paleoredox conditions
Armstrong-Altrin, J.S., Machain-Castillo, M.L., Rosales-Hoz, L., Carranza-Edwards, A.,
in sediments. Ambio Spec. Rep. 139–147.
Sanchez-Cabeza, J.A., Ruíz-Fernández, A.C., 2015. Provenance and depositional
Harnois, L., 1988. The CIW index: A new chemical index of weathering. Sediment. Geol.
history of continental slope sediments in the Southwestern Gulf of Mexico unraveled
55 (3–4), 319–322.
by geochemical analysis. Continent. Shelf Res. 95, 15–26.
Hayashi, K., Fujisawa, H., Holland, H., Ohmoto, H., 1997. Geochemistry of ~1.9 Ga
Armstrong-Altrin, J.S., Lee, Y.I., Kasper-Zubillaga, J.J., Trejo-Ramírez, E., 2017.
sedimentary rocks from northeastern Labrador, Canada. Geochem. Cosmochim. Acta
Mineralogy and geochemistry of sands along the Manzanillo and El Carrizal beach
61 (9), 4115–4137.
areas, southern Mexico: implications for palaeoweathering, provenance, and
Hernández-Hinojosa, V., Montiel-García, P.C., Armstrong-Altrin, J.S., Nagarajan, R.,
tectonic setting. Geol. J. 52 (4), 559–582.
Kasper-Zubillaga, J.J., 2018. Textural and geochemical characteristics of beach
Armstrong-Altrin, J.S., Ramos-Vázquez, M.A., Zavala-León, A.C., Montiel-García, P.C.,
sands along the western Gulf of Mexico, Mexico. Carpathian J. Earth Environ. Sci. 13
2018. Provenance discrimination between Atasta and Alvarado beach sands, western
(1), 161–174.
Gulf of Mexico, Mexico: constraints from detrital zircon chemistry and U-Pb
Herron, M.M., 1988. Geochemical classification of terrigenous sands and shales from
geochronology. Geol. J. 53 (6), 2824–2848.
core or log data. J. Sediment. Res. 58, 820–829.
Armstrong-Altrin, J.S., Botello, A.V., Villanueva, S.F., Soto, L.A., 2019. Geochemistry of
Hill, N.A., Simpson, S.L., Johnston, E.L., 2013. Beyond the bed: effects of metal
surface sediments from the northwestern Gulf of Mexico: implications for
contamination on recruitment to bedded sediments and overlying substrata.
provenance and heavy metal contamination. Geol. Q. 63 (3), 522–538.
Environ. Pol. 173, 182–191.
Armstrong-Altrin, J.S., Madhavaraju, J., Vega-Bautista, F., Ramos-Vázquez, M.A., Pérez-
Horasan, B.Y., Ozturk, A., Unal, Y., 2020. Geochemical and anthropogenic factors
Alvarado, B.Y., Kasper-Zubillaga, J.J., Ekoa Bessa, A.Z., 2021a. Mineralogy and
controlling the heavy metal accumulation in the soils of Sarayonu Ladik link roads.
geochemistry of tecolutla and Coatzacoalcos beach sediments, SW Gulf of Mexico.
Carpathian J. Earth Environ. Sci. 15 (1), 145–156.
Appl. Geochem. 134, 105103.
Jamwal, M., Pandita, S.K., Sharma, M., Bhat, G.M., 2020. Petrography, provenance and
Armstrong-Altrin, J.S., Ramos-Vázquez, M.A., Hermenegildo-Ruiz, N.Y.,
diagenesis of Murree Group of rocks exposed along Basohli – Bani road, Kathua,
Madhavaraju, J., 2021b. Microtexture and U-Pb geochronology of detrital zircon
Jammu. J. Indian Assoc. Sediment. 37 (2), 15–26.
grains in the Chachalacas beach, Veracruz State, Gulf of Mexico. Geol. J. 56 (5),
Janpou, A.K., Ngueutchoua, G., Ekoa Bessa, A.Z., Armstrong-Altrin, J.S., Kayou, U.R.K.,
2418–2438.
Odilia-Alexandra, N.N.M.N., Njanko, T., Bela, V.A., Tiotsop, M.S.K., Tankou, J.G.,
Ayala-Pérez, M.P., Armstrong-Altrin, J.S., Machain-Castillo, M.L., 2021. Heavy metal
2022. Composition, weathering, and provenance of beach sands adjacent to volcanic
contamination and provenance of sediments recovered at the Grijalva River delta,
rocks in the northern Gulf of Guinea, SW Cameroon. J. Afr. Earth Sci. 188, 104473.
southern Gulf of Mexico. J. Earth Syst. Sci. 130, 88.
Jha, D.K., Ratnam, K., Rajaguru, S., Dharani, G., Devi, M.P., Kirubagaran, R., 2019.
Banerjee, S., Choudhury, T.R., Saraswati, P.K., Khanolkar, S., 2020. The formation of
Evaluation of trace metals in seawater, sediments, and bivalves of Nellore, southeast
authigenic deposits during Paleogene warm climatic intervals: a review.
coast of India, by using multivariate and ecological tool. Mar. Pollut. Bull. 146, 1–10.
J. Palaeogeogr. 9, 27.
Jiménez-Orocio, O., Espejel, I., Martínez, M.L., 2015. La investigación científica sobre
Basu, A., 2020. Chemical weathering, first cycle quartz sand, and its bearing on quartz
dunas costeras de México: origen, evolución y retos. Rev. Mex. Biodivers. 86 (2),
arenite. J. Indian Assoc. Sediment. 37 (2), 3–14.
486–507.
Bertrand, S., Hughen, K.A., Sepúlveda, J., Pantoja, S., 2011. Geochemistry of surface
Johnson, J.E., Webb, S.M., Fischer, C., Ma, W.W., 2016. Manganese mineralogy and
sediments from the fjords of Northern Chilean Patagonia (44–47S): spatial
diagenesis in the sedimentary rock record. Geochem. Cosmochim. Acta 173,
variability and implications for paleoclimate reconstructions. Geochem. Cosmochim.
210–231.
Acta 76, 125–146.
Jones, B., Manning, D.C., 1994. Comparison of geochemical indices used for the
Botello, A.V., Soto, L.A., Ponce-Vélez, G., Villanueva, S.F., 2015. Baseline for PAHs and
interpretation of paleo-redox conditions in Ancient mudstones. Chem. Geol. 111,
metals in NW Gulf of Mexico related to the deep water horizon oil spill. Estuar. Coast
111–129.
Shelf Sci. 156, 124–133.
Kasper-Zubillaga, J.J., Martínez-Serrano, R.G., Arellano-Torres, E., Álvarez-Sánchez, L.
Bracciali, L., Marroni, M., Pandolfi, L., Rocchi, S., 2007. Geochemistry and petrography
F., Andrade, D.P., Bermúdez, A.G., Carlos-Delgado, L., 2021. Petrographic and
of western Tethys Cretaceous sedimentary covers (Corsica and Northern Apennines):
geochemical analysis of dune sands from southeastern, Oaxaca, Mexico. Geol. J. 56
from source areas to configuration of margins. In: Arribas, J., Critelli, S.,
(6), 3012–3034.

10
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Krishnan, G., Nagendra, R., Elango, L., Narasimha, K.N.P., Vybhav, K., Mujumdar, M., Ramírez-Montoya, E., Madhavaraju, J., Monreal, R., 2021. Geochemistry of the
2020. Spatial and Temporal Variations in Geochemistry of Cauvery River Sediments sedimentary rocks from the Antimonio and Rio Asuncion Formations, Sonora,
(Tamilnadu, India): Indicators of Provenance and Weathering. J. Indian Assoc. Mexico: Implications for weathering, provenance and chemostratigraphy. J. South
Sediment. 37 (2), 49–60. Amer. Earth Sci. 106, 103035.
LeMonte, J.J., Stuckey, J.W., Sanchez, J.Z., Tappero, R., Rinklebe, J., Sparks, D.L., 2017. Ramos-Vázquez, M.A., Armstrong-Altrin, J.S., 2021a. Provenance of sediments from
Sea level rise induced arsenic release from historically contaminated coastal soils. Barra del Tordo and Tesoro beaches, Tamaulipas State, northwestern Gulf of Mexico.
Environ. Sci. Technol. 51, 5913–5922. J. Palaeogeogr. 10. Article no. 20.
Li, C., Yang, S., Zhao, J.-X., Dosseto, A., Bi, L., Clark, T.R., 2016. The time scale of river Ramos-Vázquez, M.A., Armstrong-Altrin, J.S., 2021b. Microtextures on quartz and zircon
sediment source-to-sink processes in East Asia. Chem. Geol. 446, 138–146. grain surfaces in the Barra del Tordo and Tesoro beaches, northwestern Gulf of
Liu, R., Guo, L., Men, C., Wang, Q., Miao, Y., Shen, Z., 2019. Spatial-temporal variation Mexico. Arabian J. Geosci. 14. Article no. 949.
of heavy metals sources in the surface sediments of the Yangtze River estuary. Mar. Ramos-Vázquez, M.A., Armstrong-Altrin, J.S., Rosales-Hoz, L., Machain-Castillo, M.L.,
Pollut. Bull. 138, 526–533. Carranza-Edwards, A., 2017. Geochemistry of deep-sea sediments in two cores
Liu, D., Bertrand, S., Villaseñor, T., Dijck, T.V., Fagel, N., Mattielli, N., 2020. Provenance retrieved at the mouth of the Coatzacoalcos river delta, Western Gulf of Mexico,
of northwestern Patagonian river sediments (44-48◦ S): A critical evaluation of Mexico. Arabian J. Geosci. 10, 148.
mineralogical, geochemical and isotopic tracers. Sediment. Geol. 408, 105744. Ramos-Vázquez, M.A., Armstrong-Altrin, J.S., Machain-Castillo, M.L., Gío-Argáez, F.R.,
Machain-Castillo, M.L., Ruiz-Fernández, A.C., Alonso-Rodríguez, R., Sanchez-Cabeza, J. 2018. Foraminiferal assemblages, 14C ages, and compositional variations in two
A., Gío-Argáez, F.R., Rodríguez-Ramírez, A., Villegas-Hernández, R., Mora-García, A. sediment cores in the western Gulf of Mexico. J. South Amer 88, 480–496. Earth Sci.
I., Fuentes-Sánchez, A.P., Cardoso-Mohedano, J.G., Hernández-Becerril, D.U., Rasul, A., Al-Jaleel, H., 2022. Assessment of Heavy Metal Pollution in Clay Fraction of
Esqueda-Lara, K., Santiago-Pérez, S., Gómez-Ponce, M.A., Pérez-Bernal, L.H., 2020. the Euphrates River Sediments, AlQaim, Haditha Area, Western Desert Iraq. Iraqi
Anthropogenic and natural impacts in the marine area of influence of the Grijalva - Geol. J. 55 (1A), 139–157.
Usumacinta River (Southern Gulf of Mexico) during the last 45 years. Mar. Pollut. Razum, I., Lužar-Oberiter, B., Zaccarini, F., Babić, L., Miko, S., Hasan, O., Ilijanić, N.,
Bull. 156, 111245. Beqiraj, E., Pawlowsky-Glahn, V., 2021. New sediment provenance approach based
Madhavaraju, J., Ramírez-Montoya, E., Monreal, R., González-León, C.M., Pi-Puig, T., on orthonormal log ratio transformation of geochemical and heavy mineral data:
Espinoza-Maldonado, I.G., Grijalva-Noriega, F.J., 2016a. Paleoclimate, Sources of eolian sands from the southeastern Adriatic archipelago. Chem. Geol. 583,
paleoweathering and paleoredox conditions of Lower Cretaceous shales from the 120451.
Mural Limestone, Tuape section, northern Sonora, Mexico: Constraints from clay Roddaz, M., Dera, G., Mourlot, Y., Calvès, G., Kim, J.-H., Chaboureau, A.-C., Mounic, S.,
mineralogy and geochemistry. Rev. Mex. Ciencias Geol. 33 (1), 34–48. Raisson, F., 2021. Provenance constraints on the Cretaceous-Paleocene erosional
Madhavaraju, J., Tom, M., Lee, Y.I., Balaram, V., Ramasamy, S., Carranza-Edwards, A., history of the Guiana Shield as determined from the geochemistry of clay-size
Ramachandran, A., 2016b. Provenance and tectonic settings of sands from Puerto fraction of sediments from the Arapaima-1 well (Guyana-Suriname basin). Mar.
Peñasco, Desemboque and Bahia Kino beaches, Gulf of California, Sonora, Mexico. Geol. 434, 106433.
J. South Amer. Earth Sci. 71, 262–275. Rosales-Hoz, L., Carranza-Edwards, A., Martínez-Serrano, R., Alatorre, M.A., Armstrong-
Madhavaraju, J., Pacheco-Olivas, S.A., Gonzalez-Leon, C.M., Espinoza-Maldonado, I.G., Altrin, J.S., 2015. Textural and geochemical characteristics of continental margin
Sanchez-Medrano, P.A., Villanueva-Amadoz, U., Monreal, R., Pi-Puig, T., Ramirez- sediments in the SW Gulf of Mexico: implications for source and seasonal change.
Montoya, E., Grijalva-Noriega, F.J., 2017. Clay Mineralogy and geochemistry of the Environ. Monit. Assess. 187 (205), 1–19.
Lower Cretaceous siliciclastic rocks of the Morita Formation, Sierra San José section, Rosales-Lagarde, L., Centeno-García, E., Dostal, J., Sour-Tovar, F., Ochoa-Camarillo, H.,
Sonora, Mexico. J. South Amer 76, 397–411. Earth Sci. Quiroz-Barroso, S., 2005. The Tuzancoa Formation: Evidence of an Early Permian
Madhavaraju, J., Saucedo-Samaniego, J.C., Löser, H., Espinoza-Maldonado, I.C., submarine continental Arc in East-Central Mexico. Int. Geol. Rev. 47 (9), 901–919.
Solari, L., Monreal, R., Grijalva-Noriega, F.J., Jaques-Ayala, C., 2019. Detrital zircon Roser, B.P., Korsch, R.J., 1988. Provenance signatures of sandstone-mudstone suites
record of Mesozoic volcanic arcs in the Lower Cretaceous Mural Limestone, determined using discrimination function analysis of major element data. Chem.
northwestern Mexico. Geol. J. 54, 2621–2645. Geol. 67, 119–139.
Madhavaraju, J., Rajendra, S.P., Lee, Y.I., Montoya, E.R., Ramasamy, S., SantaCruz, R.F., Ruiz-Mendoza, D.V., Verma, S.K., Torres-Sánchez, D., Barry, T.L., Moreno, J.A., Torres-
2020. Mineralogy and geochemistry of clastic sediments of the Terani Formation, Hernández, J.R., 2021. Geochemistry and geochronology of intermediate volcanic
Cauvery Basin, southern India: implications for paleoweathering, provenance and rocks from the Compostela area, Nayarit, Mexico: Implications for petrogenesis and
tectonic setting. Geosci. J. 24, 651–667. tectonic setting. Geol. J. 56, 4401–4428.
Madhavaraju, J., Armstrong-Altrin, J.S., Pillai, R.B., Pi-Puig, T., 2021. Geochemistry of Rząsa, S., Owczarzak, W., 2013. Methods for the granulometric analysis of soil for science
sands from the Huatabampo and Altata beaches. Gulf of California, Mexico. Geol. J. and practice. Pol. J. Soil Sci. XLVI/I, 1–50.
56, 2398–2417. Sato, H., Hayashi, K.-I., Ogawa, Y., Kawamura, K., 2012. Geochemistry of deep sea
Mapel-Hernández, M.D., Armstrong-Altrin, J.S., Botello, A.V., Lango-Reynoso, F., 2021. sediments at cold seep sites in the Nankai Trough: Insights into the effect of
Bioavailability of Cd and Pb in sediments of the National Park Veracruz Reef System, anaerobic oxidation of methane. Mar. Geol. 323–325, 47–55.
Gulf of Mexico. Appl. Geochem. 133 article no. 105085. Schwestermann, T., Eglinton, T.I., Haghipour, N., McNichol, A.P., Ikehara, K.,
Martin, R.E., Liddell, W.D., 1989. Relation of counting methods to taphonomic gradients Strasser, M., 2021. Event-dominated transport, provenance, and burial of organic
and biofacies zonation of foraminiferal sediment assemblages. Mar. Micropaleontol. carbon in the Japan Trench. Earth Planet Sci. Lett. 563, 116870.
15 (1–2), 67–89. Shaheen, S.M., Abdelrazek, M.A.S., Elthoth, M., Moghanm, F.S., Mohamed, R.,
Mason, C.C., Fildani, A., Gerber, T., Blum, M.D., Clark, J.D., Dykstra, M., 2017. Climatic Hamza, A., El-Habashi, N., Wang, J.X., Rinklebe, J., 2019. Potentially toxic elements
and anthropogenic influences on sediment mixing in the Mississippi source-to-sink in saltmarsh sediments and common reed (Phragmites australis) of Burullus coastal
system using detrital zircons: Late Pleistocene to recent. Earth Planet Sci. Lett. 466, lagoon at North Nile Delta, Egypt: a survey and risk assessment. Sci. Total Environ.
70–79. 649, 1237–1249.
McDonough, W.F., Sun, S.-S., 1995. The composition of the Earth. Chem. Geol. 120, Stuiver, M., Reimer, P.J., 1993. Extended 14C database and revised CALIB radiocarbon
223–253. calibration program. Radiocarbon 35, 215–230.
Meinhardt, A.K., März, C., Schuth, S., Lettmann, K.A., Schnetger, B., Wolff, J.O., Subin, P.R., Ramasamy, S., Armstrong-Altrin, J.S., Chandrasekar, T., 2022. The
Brumsack, H.J., 2016. Diagenetic regimes in Arctic Ocean sediments: Implications petrography and geochemistry of clastic rocks from the Upper Cretaceous Terani
for sediment geochemistry and core correlation. Geochem. Cosmochim. Acta 188, Formation of the Cauvery Basin, Southern India. Geol. Carpathica 73 (1), 63–79.
125–146. Sundararajan, S., Khadanga, M.K., Kumar, J.P.P.J., Raghumaran, S., Vijaya, R., Jena, B.
Monreal-Gómez, A., Salas-de-León, D., Padilla-Pilotze, A., Alatorre-Mendieta, M., 1992. K., 2017. Ecological risk assessment of trace metal accumulation in sediments of
Hydrography and estimation of density currents in the southern part of the Bay of Veraval Harbor, Gujarat, Arabian Sea. Mar. Pollut. Bull. 114, 592–601.
Campeche, Mexico. Cienc. Mar. 18 (4), 115–133. Swain, A., Singh, S.K., Mohapatra, K.K., Patra, A., 2021. Sewage sludge amendment
Nagarajan, R., Armstrong-Altrin, J.S., Kessler, F.L., Hidalgo-Moral, E.L., Dodge-Wan, D., affects spinach yield, heavy metal bioaccumulation, and soil pollution indexes.
Taib, N.I., 2015. Provenance and tectonic setting of Miocene siliciclastic sediments, Arabian J. Geosci. 14. Article no. 171.
Sibuti Formation, northwestern Borneo. Arabian J. Geosci. 8, 8549–8565. Tamayo, J.L., 1991. Geografía Moderna de México, eleventh ed. Trillas, Mexico City.
Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred Tawfik, H.A., Salah, M.K., Maejima, W., Armstrong-Altrin, J.S., Abdel-Hameed, A.-M.T.,
from major element chemistry of lutites. Nature 299, 715–717. Ghandour, M.M.E., 2018. Petrography and geochemistry of the Lower Miocene
Ngueutchoua, G., Ekoa Bessa, A.Z., Eyong, J.T., Zandjio, D.D., Djaoro, H.B., Nfada, L.T., Moghra sandstones, Qattara Depression, north Western Desert, Egypt. Geol. J. 53,
2019. Geochemistry of cretaceous fine-grained siliciclastic rocks from Upper 1938–1953.
Mundeck and Logbadjeck Formations, Douala sub-basin, SW Cameroon: implications Taylor, S.R., McLennan, S.M., 1985. The Continental Crust: its Composition and
for weathering intensity, provenance, paleoclimate, redox condition, and tectonic Evolution. Blackwell, Oxford, UK.
setting. J. Afr. Earth Sci. 152, 215–236. Tehna, N., Sababa, E., Ekoa Bessa, A.Z., Etame, J., 2019. Mine waste and heavy metal
Okita, P.M., 1992. Manganese carbonate mineralization in the Molango District, Mexico. pollution in Betare-Oya Mining Area (Eastern Cameroon). Environ. Earth Sci. Res. J.
Econ. Geol. 87 (5), 1345–1366. 6 (4), 167–176.
Okita, P.M., Shanks III, W.C., 1992. Origin of stratiform sediment-hosted manganese Torres-Sánchez, D., Verma, S.K., Verma, S.P., Velasco-Tapia, F., Torres-Hernández, J.R.,
carbonate ore deposits: Examples from Molango, Mexico, and TaoJiang, China. 2019. Petrogenetic and tectonic implications of Oligocene-Miocene volcanic rocks
Chem. Geol. 99 (1–3), 139–163. from the Sierra de San Miguelito complex, central Mexico. J. South Amer. Earth Sci.
Omar, N., McCann, T., Al-Juboury, A.I., Suárez-Ruiz, I., 2021. Solid bitumen in shales 95, 102311.
from the Middle to Upper Jurassic Sargelu and Naokelekan Formations of Velmurugan, K., Madhavaraju, J., Balaram, V., Ramasamy, S., Ramachandran, A.,
northernmost Iraq: implication for reservoir characterization. Arabian J. Geosci. 14, Ramirez-Montoya, E., Saucedo-Samaniego, J.C., 2019. In: Mondal, M.E.A.,
755. Armstrong-Altrin, J.S., Singh, S.P. (Eds.), Provenance and Tectonic Setting of the
Clastic Rocks of the Kerur Formation, Badami Group, Mohare Area, Karnataka, India.

11
J.S. Armstrong-Altrin et al. Applied Geochemistry 141 (2022) 105321

Geological Evolution of the Precambrian Indian Shield, Book Publication under SES volcanic rocks from the Mesa Virgen-Calerilla, Zacatecas, Mexico: Implications for
Series by. Springer-Verlag, pp. 239–269. the magma source and tectonic setting. Geol. J. 56 (7), 3771–3790.
Venu, U.A., Velmayil, P., 2021. Texture, mineralogy and geochemistry of Teri sediments Wang, Z., Wang, J., Fu, X., Zhan, W., Armstrong-Altrin, J.S., Yu, F., Feng, X., Song, C.,
from the Kuthiraimozhi deposit, southern Tamilnadu, India: implications on Zeng, S., 2018. Geochemistry of the Upper Triassic black mudstones in the
provenance, weathering and palaeoclimate. Arabian J. Geosci. 14, 364. Qiangtang Basin, Tibet: Implications for paleoenvironment, provenance, and
Vera-Mendoza, R., Salas-de-León, D.A., 2014. Effect of environmental factors on tectonic setting. J. Asian Earth Sci. 160, 118–135.
zooplankton abundance and distribution in river discharge influence areas in the Wang, L., MacLennan, S.A., Cheng, F., 2020. From a proximal-deposition-dominated
southern Gulf of Mexico. In: Amezcua, F., Bellgraph, B. (Eds.), Fisheries Management basin sink to a significant sediment source to the Chinese Loess Plateau: Insight from
of Mexican and Central American Estuaries. Estuaries of the World. Springer, the quantitative provenance analysis on the Cenozoic sediments in the Qaidam
Netherlands, pp. 93–112. basin, northern Tibetan Plateau Palaeogeo. Palaeoclimatol. Palaeoecol. 556,
Verma, S.P., 2000. Geochemical evidence for a lithospheric source for magmas from Los 109883.
Humeros caldera, Puebla, Mexico. Chem. Geol. 164 (1–2), 35–60. Wentworth, C.K., 1922. A scale of grade and class terms for clastic sediments. J. Geol. 30,
Verma, S.P., 2001. Geochemical evidence for a Rift-Related Origin of bimodal volcanism 377–392.
at Meseta Río San Juan, North-Central Mexican Volcanic Belt. Int. Geol. Rev. 43, Yang, W., Peng, S., Wang, M., Zhang, H., 2021. Provenance of upper Permian-Triassic
475–493. sediments in the south of North China: Implications for the Qinling orogeny and
Verma, S.P., Armstrong-Altrin, J.S., 2013. New multi-dimensional diagrams for tectonic basin evolution. Sediment. Geol. 424, 106002.
discrimination of siliciclastic sediments and their application to Precambrian basins. Yu, L., Zou, S., Cai, J., Xu, D., Zou, F., Wang, Z., Wu, C., Liu, M., 2016. Geochemical and
Chem. Geol. 355 (4), 117–133. Nd isotopic constraints on provenance and depositional setting of the Shihuiding
Verma, S.P., Armstrong-Altrin, J.S., 2016. Geochemical discrimination of siliciclastic Formation in the Shilu Fe-Co-Cu ore district, Hainan Province, south China. J. Asian
sediments from active and passive margin settings. Sediment. Geol. 332, 1–12. Earth Sci. 119, 100–117.
Verma, S.K., Fimbres, K.G.A., Torres-Sánchez, D., Hernández, J.R.T., Torres-Sánchez, S. Zaid, S.M., Elbadry, O., Ramadan, F., Mohamed, M., 2015. Petrography and
A., López-Loera, H., 2020. Geochemistry and petrogenesis of Oligocene felsic geochemistry of Pharaonic sandstone monuments in Tall San Al Hagr, Al Sharqiya
volcanic rocks from the Pinos Volcanic Complex, Mesa Central, Mexico. J. South Governorate, Egypt: Implications for provenance and tectonic setting. Turk. J. Earth
Amer 102, 102704. Earth Sci. Sci. 24, 344–364.
Verma, S.K., Torres-Sánchez, D., Hernández-Martínez, K.R., Malviya, V.P., Singh, P.K., Zhan, S., Wu, J., Wang, J., Jing, M., 2020. Distribution characteristics, sources
Torres-Hernández, J.R., Rivera-Escota, B.A., 2021. Geochemistry of Eocene felsic identification and risk assessment of n-alkanes and heavy metals in surface
sediments, Tajikistan, Central Asia. Sci. Total Environ. 709, 136278.

12

You might also like