Next Article in Journal
Magmatic Redox Evolution and Porphyry–Skarn Transition in Multiphase Cu-Mo-W-Au Systems of the Eocene Tavşanlı Belt, NW Türkiye
Previous Article in Journal
Editorial for Special Issue “Adsorption Properties and Environmental Applications of Clay Minerals”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sedimentary Facies and Geochemical Signatures of the Khewra Sandstone: Reconstructing Cambrian Paleoclimates and Paleoweathering in the Salt Range, Pakistan

1
Department of Earth Science, Quaid-i-Azam University, Islamabad 45320, Pakistan
2
Department of Geology and Mines, Faculty of Science, Nangarhar University, Jalalabad 2601, Afghanistan
3
Department of Geology, University of Vienna, 1090 Vienna, Austria
4
Department of Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, China
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(8), 789; https://doi.org/10.3390/min15080789
Submission received: 2 June 2025 / Revised: 11 July 2025 / Accepted: 22 July 2025 / Published: 28 July 2025
(This article belongs to the Section Mineral Geochemistry and Geochronology)

Abstract

Red sandstones of the Cambrian age are globally distributed and represent an important sedimentation phase during this critical time interval. Their sedimentology and geochemistry can provide key information about the sedimentation style, paleoclimatic conditions, and weathering trends during the Cambrian. In the Salt Range of Pakistan, the Khewra Sandstone constitutes the Lower Cambrian strata and consists of red–maroon sandstones with minor siltstone and shale in the basal part. Cross-bedding, graded bedding, ripple marks, parallel laminations, load casts, ball and pillows, desiccation cracks, and bioturbation are the common sedimentary features of the formation. The sandstones are fine to medium to coarse-grained with subangular to subrounded morphology and display an overall coarsening upward trend. Petrographic analysis indicates that the sandstones are sub-arkose and sub-lithic arenites, and dolomite and calcite are common cementing materials. X-ray Diffraction (XRD) analysis indicates that the main minerals in the formation are quartz, feldspars, kaolinite, illite, mica, hematite, dolomite, and calcite. Geochemical analysis indicates that SiO2 is the major component at a range of 53.3 to 88% (averaging 70.4%), Al2O3 ranges from 3.1 to 19.2% (averaging 9.2%), CaO ranges from 0.4 to 25.3% (averaging 7.4%), K2O ranges from 1.2 to 7.4% (averaging 4.8%), MgO ranges from 0.2 to 7.4% (averaging 3.5%), and Na2O ranges from 0.1 to 0.9% (averaging 0.4%), respectively. The results of the combined proxies indicate that the sedimentation occurred in fluvial–deltaic settings under overall arid to semi-arid paleoclimatic conditions with poor to moderate chemical weathering. The Khewra Sandstone represents the red Cambrian sandstones on the NW Indian Plate margin of Gondwana and can be correlated with contemporaneous red sandstones in the USA, Europe, Africa, Iran, and Turkey (Türkiye).

1. Introduction

The Cambrian Period represents a pivotal time interval in the Earth’s geological history. Climate change, continental drift, the emergence of new life forms, and the catastrophic extinction of a significant percentage of species were the numerous events that occurred during this period [1]. The Earth’s surface temperature increased dramatically (from an average of 12 °C to around 22 °C), changing the global ice age that preceded this time. Polar or high-altitude glaciers hardly existed during the Cambrian [2], and the eustatic sea level was relatively high [3,4]. The continents were mostly low-lying deserts and alluvial plains, and during the Sauk transgression, the rising Cambrian sea encroached upon these regions, creating huge epicontinental seas [5,6,7].
Early Cambrian red sandstones, which are widely distributed across the globe (e.g., in India, Iran, Turkey (Türkiye), the Arabian Peninsula, Egypt, Poland, Sweden, the United States, and Canada), serve as valuable sedimentary records for reconstructing sedimentation styles, paleoclimatic conditions, and weathering patterns during this pivotal geological interval (e.g., [1,2,3]). As such, these have been the focus of numerous studies exploring Cambrian sea-level fluctuations [3,4,5], continental configurations, and biotic evolution [7,8], as well as sediment provenance, basin development, and chemical weathering trends [9,10,11,12,13,14,15]. These investigations generally indicate that red Cambrian sandstones were deposited in fluvio–deltaic settings under arid to semi-arid paleoclimatic conditions, characterized by low to moderate degrees of chemical weathering. Details on Cambrian correlation are provided in Chapter 7 (Regional and Global Correlation) to avoid repetition.
In Pakistan, the Salt Range is a crucial region for understanding the geological and paleoclimatic changes that occurred on the Indian Plate during the Late Neoproterozoic to Early Cambrian [8,9]. The first strata from the Indian Plate, identified as Cambrian, are found in the Salt Range [10]. The Khewra Sandstone is the lowermost unit of the Cambrian Jhelum Group and shows disconformable contact with the overlying Kussak Formation [9,10,11]. The formation consists of red–maroon sandstones with some shale in the lower part and is distributed throughout the Salt Range and the Khisor Range [12]. Despite being an essential reservoir rock in the area, the sedimentation style, paleoclimates, and weathering trend of the formation have not been focused on in detail. A greater understanding of the sedimentation style and weathering trend of the Khewra Sandstone may provide valuable information for regional and global reconstruction of the Cambrian paleoclimates and weathering trend. Hence, the present study focuses on the sedimentation style, the paleoclimate, and the weathering trend during the deposition of the Khewra Sandstone, in the Salt Range, Pakistan. This study uses outcrop data, petrography, bulk mineralogy (XRD), and bulk rock geochemistry data to contribute to our understanding of the sedimentation style, the paleoclimatic variation, and the weathering trend during the Cambrian.

2. Paleogeography and General Geology

During the Early Cambrian, Gondwana was the dominant landmass that formed following the breakup of the supercontinent Rodinia. It included present-day Africa, South America, Antarctica, Australia, Europe, India, and parts of Asia (Figure 1a) [13]. Gondwana rotated counter-clockwise, causing the eastern margin to drift northward [14,15]. By the end of the Cambrian, this continental mass had migrated southward again, likely reaching between 10° and 15° south latitude [6]. The Salt Range [16,17,18] and the northwestern margin of the Indian Plate [19,20] occupied a high-latitude paleogeographic position during the Cambrian, likely between 30° and 60° south of the equator [21,22,23,24]. At that time, the region lay along the northern passive margin of Gondwana and was influenced by arid to semi-arid climatic conditions, which favored the deposition of extensive evaporite and siliciclastic sequences [21,22,23,24,25]. The Khewra Sandstone was deposited unconformably over the salt and gypsum-rich evaporites of the underlying Precambrian Salt Range Formation [16,17], reflecting a transition from restricted evaporitic to more open marine and fluvial environments. The Indian Plate, as part of Gondwana, contributed to this paleogeographic setting [18,19,20,21,22].
Today, the Khewra Sandstone is part of the Himalayan foreland basin (Kohat–Potwar Plateau). The underlying Salt Range Formation plays a critical structural role as a detachment surface that accommodates thrusting associated with the Himalayan orogeny [10,17]. Some studies have noted an unconformity between the Khewra Sandstone and the Salt Range Formation, potentially marked by a magmatic intrusion known as the “Khewra Trap”. The “Khewra Trap” was named as “Khewrite” [27]. It has a purple–green color, is 6 m thick, and consists of highly decomposed radiating needles, probably composed of pyroxene [9]. However, the nature and significance of this intrusion remain insufficiently documented and warrant further clarification.
The Salt Range is prominently exposed along the east–west trending Salt Range Thrust (SRT), extending between the Jhelum River in the east and the Indus River in the west (Figure 1b,c) [28]. The Cambrian strata in this area, collectively referred to as the Jhelum Group, begin with the Khewra Sandstone, which is overlain sequentially by the Kussak, Jutana, and Baghanwala formations [8]. The pioneering study on the Khewra Sandstone was conducted by Wynne [29], who used the term “Purple sandstone series” for the formation. Noetling [30] used the term “Khewra Group” for this rock unit, and it was formally revised to the “Khewra Sandstone” by the Stratigraphic Committee of Pakistan in 1973 [31]. The Khewra Gorge near Khewra town in the eastern Salt Range is the type locality of the formation. Excellent exposures are found in the Nilawahan Gorge and Vasnal Village near Kallar Kahar in the central Salt Range. Khan Zaman Nala in the western Salt Range serves as the principal reference section of the formation [11].

3. Materials and Methods

3.1. Fieldwork

Stratigraphic sections of the Khewra Sandstone were studied at four locations, including the Khewra Gorge (KG), Nilawahan (NW), Vasnal (VS), and Khan Zaman Nala (KZ) in an east-to-west transect covering the eastern, central, and western Salt Range (Figure 1 and Figure 2). One hundred and seventy-seven (177) samples were collected, encompassing all facies variations (Table 1).
Standard sedimentological procedures, e.g., [32], were followed to study the sedimentary structures and features, including observations of the lateral and vertical facies variations. Various types of cross-beds, parallel lamination, beds, deformation structures, ripple marks, erosional surfaces, and oxidized zones were identified in the field. Important features and structures were described and photographed, and lithological samples were collected from essential beds.

3.2. Petrography

Based on sedimentological variation, 100 samples were selected from four studied outcrops, and thin sections were prepared in the Department of Geology, Bacha Khan University, Charsadda, Pakistan. The grain size, shape, texture, and composition of each sample were studied using a Nikon ECLIPSE LV100ND microscope with a Nikon ECLIPSE 5MP HD camera (Nikon, Tokyo, Japan) in the sedimentology lab, National Centre of Excellence in Geology (NCEG), University of Peshawar, Pakistan. Point counting [32,33,34] was the primary focus of the petrographic studies, and 500 grains were counted for each sample (thin section). The quartz grains were petrographically subdivided into monocrystalline quartz (Qm) and polycrystalline quartz (Qp). Further, Qm was subdivided into monocrystalline quartz with unit extinction (Qmue) and monocrystalline quartz with undulose extinction (Qmuu). Likewise, polycrystalline quartz was subdivided into polycrystalline quartz with two to three quartz crystals (Qpq(2–3)) and polycrystalline quartz with more than three quartz crystals (Qpq>3). Chert was included in Qp [35]. Feldspars (F) were categorized as alkali feldspars (K) and plagioclase feldspars (P). Lithic fragments (L) have the same meaning as rock fragments (R) [36] and include sedimentary lithics (Ls), metamorphic lithics (Lm), volcanic lithics (Lv), and metavolcanic lithics (Lvm). Plutonic lithic fragments were subdivided into constituent grains (visible under a microscope) following the Gazzi–Dickenson point counting method [35]. The total lithic fragments (Lt) are the sum of Ls, Lv, Lm, and Lvm [36]. The data obtained were plotted and interpreted following the practices of the existing literature (various QFL plots) [32,37,38,39,40,41,42].

3.3. X-Ray Diffraction (XRD) Analysis

Based on field observation and petrographic studies, 40 samples from four sections were selected for XRD analysis (Table 1). The XRD was applied to identify various mineral species and to determine the quantitative mineralogical composition of the Khewra Sandstone. The XRD analysis was conducted in the chemistry lab of the TAIYUAN University of Technology, China, using a HAOYUAN DX-2700B X-ray diffractometer (XRD) (Haoyuan Instrument Co., Ltd., Dandong, China), a Cu Kα radiation source (40 KV, 30 mA) at a scanning rate of 10°/min and 2θ ranging from 5° to 70°.

3.4. Geochemistry

Geochemical proxies are key tools in the paleoclimatic and paleoenvironmental interpretation of sedimentary rocks [26,43,44,45,46,47,48,49,50,51]. Therefore, geochemical analysis of the bulk sedimentary composition of 60 selected samples was conducted using X-ray fluorescence (XRF) spectroscopy at a commercial lab. Each sample was powdered, homogenized, and then analyzed for 90 to 120 s. The analysis included major and minor oxides and trace elements.
Various geochemical proxies were used to interpret paleoclimates and paleoenvironments of the Khewra Sandstone. These included the Chemical Index of Alteration (CIA) to estimate the intensity of chemical weathering and paleoclimate reconstruction [47], C-values to describe the paleoclimate [51], Al2O3 vs. CIAmolar and K2O/Na2O vs. CIAmolar plots to distinguish between the arid, subtropical, and tropical paleoclimate types [48], a Al2O3, CaO* + Na2O, and K2O ternary plot called the A-CN-K plot [48,52,53], and a binary plot of SiO2 against the total Al2O3 + K2O + Na2O [37], as the quantity of Al, Si, K, and Na in sediments is strongly controlled by compositional variation [54,55]. Other proxies, such as Sr/Cu, Rb/Sr, and Sr/Ba, were also used to evaluate the paleoclimatic conditions during sedimentation [56,57]. Similarly, the V/Cr and Cu/Zn proxies were used to interpret the paleo-oxygenation level [58,59,60].
The combined analyses of field data, petrographic studies, XRD, and sediment geochemistry yielded reliable results, which helped us to interpret the depositional paleoenvironments and paleoclimate of the Khewra Sandstone.

4. Results

4.1. Outcrop Data and Lithofacies

Outcrop observations during the fieldwork (Figure 3a–k) identified five distinct lithofacies in the four studied sections of the Khewra Sandstone (Table 2). The characteristics of each lithofacies are discussed below:

4.1.1. Channel Margin and Floodplain Shale Lithofacies (CMFShL)

This lithofacies consists of olive green and maroon shale (Figure 3a). The maximum thickness of the shale is 75 cm in the KZ section, and the minimum observed thickness (4 cm) is in the KG section. The Channel Margin and Floodplain Shale Lithofacies (CMFShL) occurs in the lower part of the KG, NW, and KZ sections and is repeated three times in the KG outcrop, four times in the KZ outcrop, and is also observed in the uppermost part of the NW section, while the lithofacies was not observed in the vs. section. The shales are generally laminated and are overlain by channel margin sandstones.

4.1.2. Channel Margin Sandstone Lithofacies (CMSL)

The Channel Margin Sandstone Lithofacies (CMSL) comprises maroon colored, fine-grained, moderately well-sorted sandstones (Figure 3b). Thin-bedded, fine-grained sandstones are present in the lower portions of the KG, NW, and KZ sections. In contrast, sandstones in the vs. section occur in the upper part. The maximum thickness of these fine sandstones is ca. 85 cm in the KG section and ca. 80 cm in the KZ section. Ripple marks (mostly asymmetrical), including in-phase catenary ripples, occur in the KG section (Figure 3c). However, climbing ripples are present in the middle part of the KG outcrop. Desiccation cracks are present in the middle parts of the KG and KZ sections. The upper part of the CMSL displays load and cast structures (Figure 3d). Deformational sedimentary structures, including balls and pillows, slumps, and convolute beds, are seen in the lower parts of the KG and NW sections (Figure 3e).

4.1.3. Delta Lobe/Delta Plain Lithofacies (DLDPL)

The Delta Lobe/Delta Plain Lithofacies (DLDPL) contains the interbedding of fine-grained maroon sandstone and shale, which is present in the lower parts of all four studied sections (Figure 3f). The thickness of the lithofacies is generally 60–80 cm. The alternating sandstone beds that underlie the shale are thicker in the KZ section compared to other sections. The maximum thickness of alternating thicker sandstone layers with thinner shale layers from the KZ section is 13.5 m. The lithofacies is repeated four times in the KG, two times in the NW, once in the vs., and three times in the KZ section.

4.1.4. Channel Belt Sandstone Lithofacies (CBSL)

The Channel Belt Sandstone Lithofacies (CBSL) consists of medium to coarse and very coarse-grained sandstone with maroon and light grey colors. The sandstones display tabular cross-bedding, low-angle and tangential cross-bedding, and herringbone cross-stratification. Tabular cross-bedding occurs in all the studied sections. Low-angle and tangential cross-beds are restricted to the middle part of the KG section. The angle between these cross-beds is less than 20 degrees. Planar bedding in medium to coarse-grained sandstone is observed in the middle parts of all four studied sections and is repeated two times in the upper portion of the KG, NW, and KZ sections and three times in the vs. section. These vary in thickness from 60 cm (VS) to 8 m (KG). Planar cross-beds are present in the lower parts of the KG, NW, and KZ sections. Trough cross-beds are found only in the lower part of the KG section (Figure 3g), where the thickness of the trough cross-bedded strata is 8.54 m. Herringbone cross-stratifications occur in the middle parts of the KG, NW, and KZ sections (Figure 3h).

4.1.5. Channel Belt Conglomerate Lithofacies (CBCL)

The Channel Belt Conglomerate Lithofacies (CBCL) is a clast-supported conglomerate layer at the top of the Khewra Sandstone that marks contact with the overlying Kusak Formation (Figure 3i,j). This lithofacies is present in the KG and vs. sections and could not be observed in the NW and KZ sections. The clast size varies from coarse to very coarse pebbles (10 cm in diameter). The conglomerate mainly comprises quartz pebbles (Figure 3i) with some sedimentary, metamorphic, and volcanic rock fragments.

4.2. Petrography

Petrographic studies of the Khewra Sandstone indicate that quartz is the dominant component of sandstone, followed by feldspar and lithics (Figure 4 and Figure 5; Supplementary Material Table S1). Muscovite, biotite, rutile, zircon, and minor opaque minerals also occur (Figure 4a–f). Dolomite, iron oxide, and calcite are the dominant cementing materials (Figure 4b). Quartz averages 68.1% of the total framework mineralogical composition, ranging from 57.0 to 84.6%, and Qm predominates, averaging 63.7%, while Qp content averages 4.4%. The grains vary in size from fine to medium to coarse, and texturally they are angular, sub-angular to sub-rounded, while well-rounded and elongated quartz grains were also noticed. Generally, the grains are moderate to well-sorted (Figure 4d). The grain size in the NW and vs. sections is typically medium to coarse-grained, while in the KG and KZ sections, it starts from fine-grained sandstone or siltstone and coarsens in the upper parts. Point/tangential contacts are most common, while long-linear, concavo–convex, and suture contacts are also observed (Figure 4c). Feldspars are relatively fresh, and F in the framework grain averages 6%, ranging from 3.1 to 9.8%. K predominates, averaging 4.6%, while P averages 1.4% of the total framework composition (Figure 4c). The L contents range from 1.8 to 6.6% (averaging 3.7%) and include Ls = 2.0%, Lm = 0.6%, and Lv = 1.1% (Supplementary Material Table S1).

4.3. XRD Data

The XRD analysis of the Khewra Sandstone indicates that quartz is the dominant mineral (67.3%), followed by hematite (6.8%) and calcite (6.6%). Feldspar (both k-feldspar and plagioclase), illite, mica (mostly muscovite), kaolinite, and dolomite are the other common minerals (Figure 6). Quartz increases in the upper section of the KG, VS, and KZ sections while staying stable in the NW section. In contrast, dolomite decreases in the upper parts of the KG and KZ sections. Hematite is high in the shale samples from the middle part of the KG section. Feldspars are comparatively low in the NW section and relatively decreased in the upper part of the KZ and vs. sections. Illite, smectite, and chlorite are abundant in the associated shale samples. Kaolinite is also found in minor amounts throughout the formation.

4.4. Geochemistry Data

Geochemical analysis indicates that SiO2 is the most significant component, ranging from 53.3 to 88.04% (averaging 70.40%), followed by Al2O3 ranging from 3.05 to 19.17% (averaging 9.27%), CaO ranging from 0.4 to 25.26% (averaging 7.38%), K2O ranging from 1.23 to 7.42% (averaging 4.79%), MgO ranging from 0.2 to 7.42% (averaging 3.53%), and Na2O ranging from 0.14 to 0.87% (averaging 0.45%), respectively (Supplementary Material Table S2). Comparatively, the average SiO2 content is faintly higher in the NW section, averaging 72.91%. Meanwhile, Al2O3 shows the highest average in the KG section (10.20%).

4.4.1. Chemical Index of Alteration (CIA)

The average CIA value is 62.2, ranging from 50.7 to 71.8 in all four sections. The average CIA value for the KG section is 62.6, and it has a range of 50.8–71.7 (Supplementary Material Table S2). The average CIA values for the NW, VS, and KZ sections are 59.4 (ranging from 51.5 to 69.8), 58.8 (ranging from 50.7 to 65.1), and 62.2 (ranging from 56.1 to 71.5), respectively. The KG section shows a relative drop in the uppermost part, with a CIA value of (50.8). However, the CIA value remains generally low throughout the formation with minimal variation. Meanwhile, the KG section appears to have the highest CIA value in the shale (71.7). In general, the shales of the formation have continuously higher CIA values than their coarse-grained counterparts (siltstones and sandstones). In that order, the shale and fine-grained sandstone lithofacies such as the CMFShL, the CMSL, and the DLDPL show relatively high CIA values, and the CBSL of the middle and upper portions and CBCL show low CIA values, respectively.

4.4.2. The Al2O3 vs. CIAmolar and K2O/Na2O vs. CIAmolar Plots

The Al2O3 vs. CIAmolar and K2O/Na2O vs. CIAmolar plots (Figure 7a,b) show samples clustering in the range of (1–3) for CIAmolar and in the ranges of 3–19 and 1–39 for Al2O3 and K2O/Na2O, respectively. The Al2O3 values are comparatively high in the lower parts and slowly decrease in the upper parts of the formation. The highest values of Al2O3 and CIAmolar are plotted for the shales and siltstones. The lowest Al2O3 values are plotted for the coarser sandstone. The fine-grained sandstones/siltstones with desiccation cracks show Al2O3 in the 10%–13% range.
The K2O/Na2O ratio displays a significantly wide range (2.61–37.35) among all four sections. Its values decrease towards the upper part of the KG and KZ sections, while there are no major changes in the NW and vs. sections. The higher ratio (37.35) of K2O/Na2O occurs in the lower part of the formation, and the lowest value (2.61) relates to the upper part of the formation. Although the Na2O content consistently remains below 0.9%, the larger fluctuations in the K2O/Na2O ratio are attributable to a sudden drop in Na2O content.

4.4.3. The C-Values

The C-value (C = {(∑Fe + Mn + Cr + Ni + V + Co) + (∑Ca + Mg + Sr + Ba + K + Na)} (Figure 7c) plot is used to interpret the paleoclimate. The average C-value for the present sandstones is 0.19, ranging between 0.05 and 0.65, and the average for the shales is 0.67, ranging from 0.32 to 1.2. The highest C-value is seen in the shales from the middle portion of the KG section.

4.4.4. Chemical Maturity Plot (SiO2 vs. Al2O3 + K2O + Na2O)

The chemical maturity plot (SiO2 vs. Al2O3 + K2O + Na2O) displays samples clustering between 50 and 90 on the y-axis and 5 and 30 on the x-axis (Figure 7d). The SiO2 content ranges from 53.3 to 85.5 throughout the formation, where in the KG it ranges between 53.3 and 85.5, in the NW ranges from 66.1 to 78.0, in the vs. range from 61.1 to 74.3, and in the KZ range from 58.2 to 84.3. The Al2O3 + K2O + Na2O ranges from 4.7 to 24.9 in the studied sections, where in the KG, it ranges between 4.7 and 24.9, in the NW ranges from 9.9 to 21.4, in the vs. range from 11.9 to 23.04, and in KZ ranges from 8.5 to 23.8. The SiO2 (quartz) increases upward in all sections in contrast to feldspars; hence, the highest SiO2 concentrations are associated with lower Al2O3 + K2O + Na2O values. The shale from all four sections contains < 63% SiO2 with more than 22% Al2O3 + K2O + Na2O.

4.4.5. A-CN-K (Al2O3-CaO*+ Na2O-K2O) and A-N-K (Al2O3-Na2O-K2O) Plots

In both the A-CN-K and A-N-K plots, the samples cluster near the A-vertex along the A-K side in the muscovite position (>90%) (Figure 8a,b). Three shale samples from the KG section and one sample each from the NW, VS, and KZ sections cluster near the illite position (<10%). Fine-grained sandstone clusters near muscovite, while medium- and coarse-grained sandstones cluster close to muscovite and shift slightly toward the K-feldspar position, respectively. The Al2O3 and K2O display a wide range for the KG, VS, and KZ sections, while in the NW section, these display uniformity in concentration. Na2O indicates stable levels within the vs. section and low to moderate variability in the KZ, NW, and KZ sections, respectively. The CIA values for the samples are plotted to the left side of the A-CN-K plot to reflect the chemical weathering intensity corresponding to the composition evolution of the samples.

4.4.6. Al2O3/Na2O vs. CIA and Al2O3/K2O vs. CIA Plots

The Al2O3/Na2O vs. CIA and Al2O3/K2O vs. CIA plots were used to evaluate the paleoclimatic conditions and weathering trend during the deposition of the Khewra Sandstone (Figure 9a,b). The Al2O3/Na2O ratio averages 27.4 and varies between 6 and 77.4. The highest Al2O3/Na2O value is marked in the CMSL of the lower part of the KG section. The average Al2O3/K2O ratio in all four sections is 2.1, ranging from 0.8 to 2.8, with slightly higher values corresponding to shale. However, the lowest Al2O3/K2O ratios occur in the DLDPFL.

4.4.7. Weathering Index of Parker (WIP) vs. CIA Plot

The WIP values are plotted on a logarithmic scale along the Y-axis. The WIP vs. CIA plot shows samples clustering between 10 and 100 on the Y-axis (Figure 10). All The highest WIP values are obtained from the shale samples (CMFShL) of the lower part of the KZ section. In contrast, the lowest WIP values are obtained from the coarse-grained sandstone (CBSL) of the uppermost part of the KG section. In the KG, VS, and KZ sections, the WIP values increase upwards except in the shale samples, while they stay more or less the same in the NW section.

4.4.8. K/Na, Sr/Ba, Rb/Sr, V/Cr, and Cu/Zn Ratios

The K/Na, Sr/Ba and Rb/Sr ratios are used for paleoclimate and paleo-weathering trends, and the V/Cr and Cu/Zn ratios are used to evaluate paleo-oxygenation levels. Presently, data for the KG section (type locality) is included (Figure 11). The K/Na ratio is relatively high (>10) in the lower part of the KG section, particularly in the CMFShL and CMSL lithofacies, reaching higher values (>30) in KG-13 (41.79) and KG-16 (38.04). The middle and upper parts display low values (<10), with the lowest value of 2.92 in KG-51. The Sr/Ba ratio yielded low values (<0.3) throughout the formation, with a peak value (0.31) in the shale samples (e.g., KG-34) of the middle part. The Rb/Sr ratio generally indicates low values (<1) in the formation, and the lowest value of 0.27 corresponds to KG-51 in the upper part. The lower part of the formation yielded comparatively higher values (Figure 11, Supplementary Material Table S2). The V/Cr ratio of the formation is always below 2. The lowest ratio of 0.1 corresponds to KG-16. The Cu/Zn ratios for the formation always remain below 3, and the lowest value of 0.01 occurs in KG-34. The highest value of 2.92 occurs in KG-54 (the uppermost part).

5. Discussion and Interpretation

5.1. Depositional Lithofacies and Lithofacies Associations

5.1.1. Channel Margin and Floodplain Lithofacies Associations (CMFLAs)

The presence of the CMFShL that features laminated shale with an olive-green and maroon color is indicative of a fluvio–deltaic setting in low-energy, channel margin, or overbank-floodplain areas [70,71]. This reflects subaerial episodic sedimentation in channel margin and the overbank-floodplain regions of fluvio–deltaic settings during flood/high water levels [72,73,74,75], a characteristic of shallow deltaic or marginal marine settings [76]. The laminations indicate low bioturbation and undisturbed sedimentation in protected channel margins or deltaic overbank settings [77,78]. Such laminations also reflect water column stability, where minimal mixing occurs, supporting a shallow marine or marginal marine interpretation [78,79]. The vertical up-section repetition of the CMFShL in the studied sections suggests cyclic depositional events, likely driven by the combined interplay of flood pulses, channel migration, and sea-level fluctuations typical of fluvio–deltaic settings [80]. The absence of the CMFShL in the vs. section may suggest relatively higher energy, possibly a more proximal setting, where fine sediments could not settle [81,82,83].
Fine sandstones of the CMSL with thin beddings suggest deposition in low-energy environments such as abandoned distributary channels or delta fronts [84,85]. Ripple marks generally represent deposition in shallow, subaerial fluvio–deltaic environments characterized by fluvial and tidal processes [86]. Asymmetrical ripples indicate a unidirectional transport (generally a fluvial process), while in-phase catenary ripples suggest regular tidal action [87,88]. Thus, the CMSL supports sedimentation in the delta front, mouth bars, or channel margin/distributary channels, where fluvial currents dominated but were modulated by tidal influences [89]. The presence of load and cast structures suggests rapid sedimentation events, potentially linked to flooding or high-energy conditions (on channel margins), causing the deformation of water-saturated sediments [90]. The desiccation cracks in the KG and KZ sections imply a fluvio–deltaic setting with alternating wet and dry periods [91,92,93]. Deformational sedimentary structures reflect soft-sediment deformation due to rapid sedimentation and liquefaction in deltaic, fluvial, or shallow marine environments [94]. Ball and pillow structures, slumps, and convolute beds indicate instability from gravitational forces or seismic activity [95]. Deformation occurs when rapidly deposited, water-saturated sediments are loaded or destabilized, leading to slumping or internal sediment deformation before lithification [96]. Thus, the closely associated depositional features allow the CMFShL and CMSL of the Khewra Sandstone to be grouped into the CMFLA.

5.1.2. Delta Plain Lithofacies Association (DPLA)

The interbedded sandstone and shale of the DLDPL indicate cyclic changes in flow energy and sediment supply, which may suggest deposition in channel margin and delta plain settings [61]. Similar deposition can also occur in distributary channels and inter-distributary areas with periods of alternate calm (shale) and high-energy events (sandstone) [97,98]. The alternation signifies episodic sedimentation, with fluctuations between energetic flow (coarser sandstone deposition) and quiescent periods (fine-grained shale deposition), typical of deltaic environments where sediment supply varies with fluvial processes and water level changes [99,100]. The presence of alternate thicker sandstone beds with thinner shale beds suggests progradation and aggradational conditions in delta lobe systems of the fluvio–deltaic settings [101,102,103,104]. Therefore, the DLDPL represents the DPLA within the Khewra Sandstone.

5.1.3. Channel Belt Lithofacies Associations (CBLAs)

The presence of tabular cross-beds in the CBSL highlights an active fluvial system where medium to coarse-grained, moderately to well-sorted sandstones were deposited in channel belt/bar setting [55,105,106]. Planar bedding reflects low-energy water flow and steady sedimentation typical of fluvio–deltaic environments [107]. Low-angle and tangential foreset cross-beds imply deposition in lower-energy areas, such as river mouths or the delta front, where sediment accumulates as the water flow slows [108]. The planar bedding in the medium to coarse-grained sandstones indicates relatively high energy conditions and upper flow regimes in channel belt depositional settings [109].
The variation from light grey to maroon colors suggests periods of oxidation and reduction, with maroon suggesting more oxidizing conditions [110]. Thickness variations (60 cm to 8 m) reflect sedimentation rates and accommodation space changes, with thicker beds indicating periods of higher sediment influx and/or more significant accommodation space [111]. Planar cross-beds represent deposition by unidirectional currents in fluvial, deltaic, or shallow marine settings [84,112]. Trough cross-beds indicate deposition in moderate-energy environments with unidirectional currents, typical of fluvial or shallow marine settings, particularly within channels [113,114]. Herringbone cross-stratification forms by alternating cross-beds inclined in opposing directions, indicating fluctuating flow directions influenced by tidal and variable fluvial forces [115].
The occurrence of pebbly beds of the CBCL in the upper part of the formation suggests a high-energy fluvial environment, likely channel belt deposits [86]. The moderate sorting indicates periods of fluctuating energy conditions during high-energy events (e.g., floods), where larger pebbles were mobilized, while lower-energy conditions allowed the deposition of finer sediments in the pore spaces, resulting in a bimodal sedimentary fabric [116,117]. Clast-supported conglomerates suggest dynamic sediment transport conditions typical of high-energy environments [118]. Very coarse pebbles (10 cm in diameter) imply energetic sediment transport during periods of increased fluvial discharge [119,120,121]. The uppermost conglomerate bed of the formation may indicate localized deposition, likely reflecting a high-energy flood event [112,122]. Therefore, the CBCL indicates a transition zone influenced by high-energy flow conditions typical of fluvial channels or delta fronts [123,124]. The channel belt features of the CBSL and CBCL allow these to be placed in the CBLAs.

5.1.4. Vertical Facies Architecture

Shifts in the depositional environment can be interpreted by the vertical arrangement of the lithology and grain size variations [125,126,127,128]. As per “Walter’s law of facies association”, the coarsening upward succession observed in the formation signifies regression and progradation of the delta and/or sea-level fall [129,130]. Therefore, the presence of shale in the lower part of the Khewra Sandstone indicates a sea-level high stand (transgression) and deposition in relatively deeper water [131,132,133,134]. The overlying medium to coarse-grained sandstones of the CMSL and the CBSL support deposition during progressive regression and sea-level low-stand [135]. Therefore, the overlying coarser-grained delta front and delta plain sandstones indicate delta progradation [136,137]. Similarly, the apparent upward thickening of the sandstone beds of the CBSL in all four sections and the occurrence of CBCL in the uppermost parts of the formation also suggest an increase in sediment supply and sea-level fall, thereby supporting the delta progradation throughout the deposition of the Khewra Sandstone [138,139].

5.2. Petrography

The framework grain composition, size, sorting, roundness, and matrix are the petrographic features of sandstone that reveal several key aspects of the depositional environment, paleoclimate, and weathering trend during sedimentation [140,141,142]. The present petrographic results suggest a mixed depositional environment system dominated by fluvio–deltaic and deltaic shallow-marine settings. The high amount of quartz with a sub-angular to sub-rounded texture and moderate to well-sorted grains signifies fluvio–deltaic sedimentation [143,144,145,146,147,148,149,150,151].
Subangular to subrounded coarse-grained sandstones are characteristic of fluvio–deltaic systems, possibly distributary channels, where higher energy promotes the transportation and deposition of coarse material [149,150,151,152,153,154,155]. In contrast, a matrix of fine-grained clays indicates deposition in quiet-water, much-protected deltaic environments such as back swamps, tidal flats, or lagoons [156,157,158]. The upward coarsening sequence points to deltaic progradation [159]. The present sandstones are classified as sub-arkose and sub-litharenite (Figure 5a), suggesting a transitional depositional environment, typically between continental and marine settings. These sandstones are commonly associated with fluvial, deltaic, and shallow-marine settings where sediments are derived from uplifted areas and transported toward a basin [160,161,162].
The composition of the framework grains, especially the amount of quartz and the common occurrence of feldspar and lithics, aligns with arid conditions [163,164]. The relatively low alteration of feldspars and the presence of fresh feldspars suggest an arid to semi-arid paleoclimate during deposition [165,166]. Quartz is highly resistant to both chemical dissolution and mechanical disintegration, and its dominance, along with the occurrence of fresh, unaltered feldspars, suggests that mechanical weathering rather than chemical weathering played a more significant role in sediment production [167,168]. In tropical or humid climates, feldspars tend to weather rapidly into clay minerals such as kaolinite due to chemical weathering [169,170]. However, in the present case, the fresh and unaltered feldspars suggest limited chemical weathering, resulting in the prevalence of an arid climate lacking moisture and hydrolysis [171,172,173,174].
The paleoclimate, paleo-weathering, and their control on sandstone mineralogy can be obtained by QFL, QFR, binary log/log plots of QP/(F + R) vs. Qt/(F + R), and ln(Q/F) vs. ln(Q/R) plots [33,38,39]. Presently, the samples are clustered in the sub-arkose arenite field along the QF side of the plot, suggesting a semi-humid climate during deposition (Figure 5a,b). The binary log/log diagram (Figure 5c) is a sensitive plot for the indication of paleoclimatic conditions [40]. This plot indicates that the Khewra Sandstone was deposited in semi-arid to semi-humid climatic conditions. The ln(Q/F) vs. ln(Q/R) plot (Figure 5d), which is generally used to represent the Cumulative Chemical Weathering Index (CCWI), suggests that the weathering intensity in the region was low to moderate, as determined by the equation Iw = C × R, following Weltje’s methodology [40].
Overall, the framework composition plots (Figure 5a–d) generally do not support the presence of a hot and humid paleoclimate during the deposition of the Khewra Sandstone. Instead, these indicate semi-arid to semi-humid paleoclimatic conditions. However, since the QFL plots consider only three parameters (quartz, feldspars, and lithic fragments) and are limited to sand-sized grains, their effectiveness in distinguishing between semi-arid and semi-humid climates is constrained [40]. This limitation in paleoclimatic interpretation based solely on framework mineralogy was addressed in the present study by incorporating comprehensive bulk sediment geochemistry. This approach not only includes fine-grained sediments but also accounts for a broader range of geochemical variables simultaneously [40,48,52]. The significance of bulk sediment geochemistry is further elaborated in Section 5.4.

5.3. XDR Data

The common occurrence of feldspars in the present samples suggests mechanical denudation [167,175,176,177]. This supports the deposition of fine-grained sandstone/siltstone and shales, with higher feldspar content, under low chemical weathering conditions [162,178,179]. The relative drop of feldspars in the upper parts of the KZ and vs. sections suggests increased weathering, possibly linked to climatic shifts from arid toward more humid, warm conditions [180].
Dolomite and calcite indicate carbonate cementation, and their presence, particularly the high dolomite content in the lower part of the KG section, suggests conditions conducive to carbonate precipitation linked to phreatic zone diagenesis, rich in saline/marine groundwater [181,182]. The decrease in dolomite in the upper parts of the KG and KZ sections could reflect a shift towards more clastic-dominated environments. This is possibly tied to changing sea-level or reduced carbonate productivity and the presence of fresh groundwater in the pore spaces of the sandstones during the diagenesis [183]. This, together with the occurrence of iron oxide (hematite), is typically associated with oxidative conditions, potentially tied to subaerial exposure and vadose zone diagenesis or oxygenated shallow marine conditions during sedimentation [184,185].
Clay minerals (kaolinite, illite, smectite, and chlorite) generally occur in low quantities, implying hydraulic sorting and differential removal of the fine sediments during deposition [186,187]. Kaolinite typically forms under warm, humid conditions with intense chemical weathering [188,189]. Its low average throughout the formation suggests semiarid to subtropical conditions [190]. In contrast, illite, smectite, and chlorite are usually associated with lower-temperature diagenetic processes and show low chemical weathering during their deposition [191,192]. Their presence may indicate colder and arid climatic conditions or limited chemical weathering, reflecting climatic variability during the deposition of the Khewra Sandstone [193,194,195,196].

5.4. Bulk Geochemistry

This study employs various geochemical proxies derived from major oxides and trace elements to interpret the intensity of chemical weathering and paleoclimatic conditions during the deposition of the Khewra Sandstone (Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11). To maintain clarity and conciseness, a synthesized interpretation and discussion of these proxies is presented below.

5.4.1. Chemical Weathering Intensity

The CIA is a reliable proxy for interpreting the intensity of chemical weathering, with higher values reflecting progressively intense chemical weathering [64,197], and is used for paleoclimatic reconstructions [42]. The low average CIA value (61.2) for the Khewra Sandstone suggests poor chemical weathering and arid climatic conditions (Figure 8, Figure 9, Figure 10 and Figure 11). Even the high CIA values (up to 71.8) in fine-grained shale samples are still considered moderate, implying semi-arid conditions [198,199,200,201]. The minimal variation in the CIA throughout the studied sections suggests an overall regional arid to semi-arid paleoclimate characterized by poor to intermediate chemical weathering [202,203,204].
As poor chemical weathering is generally linked to arid and semi-arid climatic conditions, the mola CIA plots (Al2O3 vs. CIAmolar and K2O/Na2O vs. CIAmolar plots) [50] can confirm this interpretation. Presently, these plots indicate an arid climate for the formation (Figure 7a,b). The high amount of Al2O3 in the lower parts and its relative decrease upward indicate a decrease in clay/shale content within the formation in the vertical direction [205,206,207]. The fluctuation in the K2O/Na2O ratio depends on feldspar alteration. Na and K are mobile elements found mostly in plagioclase and k-felspar, respectively. Thus, the sudden drop in Na2O in the KG-13 and KG-16 samples of the CMSL lithofacies in the lower part of the KG section may be related to plagioclase alteration. Moreover, the low K2O/Na2O ratio suggests poor chemical weathering, potentially tied to an arid climate [208,209,210].
The C-values [50] are used for paleoclimatic interpretation and were used here to validate the molar CIA results. Higher C-values (>0.8) indicate humid conditions, values of 0.4–0.8 indicate semi-humid conditions, while lower values (<0.4) suggest semi-arid to arid paleoclimatic conditions [211]. Presently, the average C-value for the Khewra Sandstone is 0.25 (Figure 7c), indicating deposition under arid paleoclimatic conditions. The majority of the samples (63%) with C-values less than 0.2 indicate deposition under an arid climate. Of the remaining, 23% have C-values of 0.2–0.4, indicating semi-arid conditions. The remaining samples (dominantly shales in the lower part) have values ranging from 0.4 to 0.8, indicating semi-arid to semi-humid paleoclimate, and may be related to increased weathering in wetter climates [212]. The higher Fe2O3 in the shales may be an alternative reason for slightly higher C-values in the shale, rather than intense weathering in a humid paleoclimate [211].
The chemical maturity plot (SiO2 vs. Al2O3 + K2O + Na2O) is a relationship between silica (and thus quartz), alumina (Al2O3), and alkali elements (K2O + Na2O) (and thus feldspars) that provides insight into paleo-weathering and paleoclimatic conditions [65]. The clustering of samples in the semi-arid to semi-humid fields (Figure 7d) indicates that the Khewra Sandstone was deposited under poor to moderate weathering and, hence, an arid to semi-arid paleoclimate [26,213]. The high SiO2 content of the sandstones forces these samples to be plotted close to the humid field, higher on the y-axis (50–90), in all the sections, and may yield a misleading interpretation for a humid paleoclimate. Therefore, shales/clays, the weathering products of coarse-grained rocks, are the more reliable indicators of weathering conditions [49,64]. Presently, shale samples from all four reflect semi-arid to arid paleoclimatic conditions (Figure 7d).
The A-CN-K and A-N-K ternary plots indicate paleoclimatic conditions and paleo-weathering trends and include reference minerals and standards for correlation [213]. The cluster of samples near the illite–muscovite position along the A-K side indicates low to moderate chemical weathering that is likely in arid to semi-arid settings (Figure 8a,b). The low CIA values (plotted to the left side of the triangles) support this interpretation [63,214].
The debate that Na2O, K2O, and Al2O3 may be related to plagioclase and k-feldspar’s alteration allows the use of Al2O3/Na2O vs. CIA and Al2O3/K2O vs. CIA plots to interpret paleoclimate and paleo-weathering linked to such alterations [215]. Generally, the low ratios of Al2O3/Na2O and Al2O3/K2O show less alteration of plagioclases and k-feldspar, respectively, while their high ratios suggest their intense alteration [216]. Presently, both plots show low to moderate chemical weathering under an arid to semi-arid paleoclimate (Figure 9a,b) [50]. The slightly high Al2O3/K2O in the shale samples from the lower and middle parts of all four sections may indicate k-feldspar alteration [217]. In addition, the Al2O3/K2O ratio in shales can also be influenced by their compositional characteristics, as the clay minerals (illite, kaolinite, and chlorite) are typically associated with shales [218,219]. However, the highest ratio of Al2O3/Na2O in KG-13 (77.39) and KG-16 (69.40) in the CMSL’s lower part of the KG section suggests plagioclase alteration in moderate chemical weathering [220,221,222].
The WIP vs. CIA plot can also validate the weathering conditions indicated by the CIA [66]. The cluster of samples on the WIP vs. CIA plot indicates a weathering trend between low and moderate, likely under arid to semi-arid climatic conditions (Figure 10). The CMFShL and fine-grained CMSL lithofacies, which exhibit high WIP values, indicate relatively moderate weathering, while samples from the DLDPL lithofacies show comparatively low WIP values, suggesting mild or weak chemical weathering. The sample’s population on the plot unanimously negates strong weathering and rules out hot and humid paleoclimatic conditions [220,221].

5.4.2. Paleoclimate, Depositional Setting, and Paleo-Oxygenation Level

The K/Na, Sr/Ba, Rb/Sr, V/Cr, and Cu/Zn ratios help interpret the paleoclimate, the weathering intensity, and the paleo-oxygenation levels of the depositional environment [60,223,224]. Higher K/Na ratios indicate the chemical weathering trend of plagioclase and k-feldspar (discussed in the previous section). Under arid to semi-arid climatic conditions, the plagioclase content (and hence Na content) is not differentially removed due to poor chemical weathering in these settings, resulting in low K/Na ratios [26,169]. Presently, the K/Na ratio shows low values (<20), thereby indicating poor chemical weathering and deposition under arid to semi-arid paleoclimate (Figure 11). The increase in the K/Na ratios in KG-13 (41.79) and KG-16 (38.04) shows an increased alteration of plagioclase and possibly a climate shift [225]. An alternative mechanism for high K/Na ratios in these samples may be the lack of plagioclase in the parent rocks. However, sediment source switching for such a short period is highly unlikely to have occurred. In general, the high K/Na ratios in the shale can be attributed to the presence of illite-rich clay minerals, as revealed in the XRD (Figure 6).
Terrestrial sediments are generally rich in Ba and depleted in Sr, while marine sediments are usually the other way around [226,227], thereby making the Sr/Ba ratio a good proxy for identifying a shift from terrestrial to marine environments [228,229,230]. Generally, freshwaters (and thus fluvial and terrestrial settings) have low Sr/Ba ratios (<1.0), brackish waters (e.g., delta front settings) have Sr/Ba ratios of 1.0–3.0 that increase to 3.0–8.0 in saltwater (e.g., prodelta), and normal marine settings have Sr/Ba ratios of > 8.0 [225]. In the present case, the Sr/Ba ratio is always <0.5 (Figure 11), strongly supporting deposition in fluvial/terrestrial settings. Sr and Ba behave differently in response to chemical weathering and leaching. Sr is more easily dissolved and carried away by water than Ba. However, in arid to semi-arid climatic conditions where there is less rainfall, and surface water is limited and Sr can still be leached (removed) by any available water, while Ba tends to stay in place because it does not dissolve as easily [231]. This means sandstones retain more Ba and lose more Sr over time, leading to a lower Sr/Ba ratio in the rock (Figure 11).
Rubidium (Rb) is easily adsorbed during weathering processes by illitic clays, while Sr is leached in the process [232]. This makes the Rb/Sr ratio a good indicator of the intensity of chemical weathering in paleoclimatic interpretations [128]. In the present case, the Rb/Sr ratios have a low range of 0.27 to 1.17, indicating poor chemical weathering and, thus, arid to semiarid paleoclimatic conditions during the deposition of the Khewra Sandstone (Figure 11).
The interpretation of the weathering trend in the absence of paleo-oxygenation data can be misleading. The V/Cr ratio is used to interpret paleo-oxygenation levels of the depositional environments [233]. A V/Cr ratio of >4–5 is generally considered high and indicates reducing, often anoxic conditions, typically in marine or organic-rich depositional environments. A V/Cr ratio of <2 is generally considered low and indicates oxidizing conditions, common in continental or terrestrial depositional settings [233,234]. In the present scenario, the V/Cr ratios are <2, suggesting oxic conditions in continental/terrestrial environments (Figure 11) [235,236].
The V/Cr ratios as paleo-oxygenation level proxies can be verified by the Cu/Zu ratio. The Cu/Zn ratio in sandstones depends on the depositional conditions. In oxic environments, Zn is often more mobile compared to Cu, and in reducing conditions, Cu tends to be more stable and enriched compared to Zn [237,238]. Therefore, a lower Cu/Zn ratio suggests more oxygen-rich conditions, and a higher Cu/Zn ratio might indicate reducing conditions [239]. In the present case, the Cu/Zn ratios are generally <2 (Figure 11), indicating deposition of the Khewra Sandstone under oxic conditions.

6. Paleoclimate and Sedimentation

6.1. Paleoclimate and Paleo-Weathering Trend

The lithofacies and lithofacies association (Table 2), petrographic analysis (Figure 5), XRD analysis (Figure 6), and geochemical proxies (Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11) used in this study provide coherent evidence for an arid to semi-arid paleoclimate with low to moderate chemical weathering during the deposition of the Khewra Sandstone. The vertical grain-size increase observed in the formation indicates the gradual onset of fluvial conditions and an associated sea-level fall.
The angular to subrounded, poor to moderate sorting of the quartz grains and the fresh to weakly altered nature of the minerals like feldspar and lithics (Figure 4) suggest more mechanical weathering than chemical disintegration and align with arid conditions [206,240,241]. The QFL and the QP/(F + RF) vs. Qt/(F + RF) plots (Figure 5a–c) suggest semi-arid to semi-humid climatic conditions. The ln(Q/F) vs. ln(Q/R) plot (Figure 5d) shows that the weathering intensity in the region was low to moderate.
The common occurrence of feldspars and clay minerals like illite, smectite, and chlorite, as observed in the XRD analysis (Figure 6), implies low chemical weathering under arid to semi-arid conditions. The low average of kaolinite negates intense leaching and dissolution, negating tropical conditions [242,243]. Illite, smectite, and chlorite are generally the products of mechanical/physical weathering under colder and arid climatic conditions [244,245] and thus reflect deposition of the Khewra Sandstone in arid to semi-arid conditions.
The geochemical proxies, including the low to moderate CIA value (<72) and the positions of samples in the arid fields of Al2O3 vs. CIA and K2O/Na2O vs. CIA plots, indicate arid conditions of deposition (Figure 7a,b). Similarly, the C-values and the chemical maturity plot (SiO2 vs. Al2O3 + Na2O + K2O) further support arid to semi-arid deposition conditions for the Khewra Sandstone (Figure 7c,d). The A-CN-K and A-N-K ternary plots confirm the low-to-moderate chemical weathering intensity. The bivariate plots of CIA vs. Al2O3/K2O and CIA vs. Al2O3/Na2O (Figure 9a,b) and the WIP vs. CIA plot (Figure 10) provide further supporting evidence for low to moderate weathering conditions and, hence, arid to semi-arid conditions during the Khewra Sandstone’s deposition. The elemental ratios (K/Na, Sr/Ba, and Rb/Sr) also support weathering under arid to semi-arid climatic conditions (Figure 11).

6.2. Depositional Environments and Sedimentation Style

The sedimentary structures and features observed in the outcrop, lithofacies, and lithofacies associations indicate deposition of the Khewra Sandstone in fluvio–deltaic and deltaic to shallow marine sub-environments. The CMFShL indicates a fluvio–deltaic environment with episodic sedimentation in channel margin and overbank/floodplain areas. Similarly, the occurrence of the CMSL suggests a fluvio–deltaic setting with low-to-moderate energy conditions in the channel margin to delta-top areas. The presence of DLDPL points to a cyclic deposition, indicating fluctuating flow energy in a delta-plane environment. The prominent sandstones of the CBSL suggest high-energy depositional settings, such as channels fill, point bars, or deltaic plains, and the presence of the CBCL indicates high-energy fluvial settings marking transitions to deltaic conditions. The coarsening upward sequence points to delta progradation, where the depositional environment shifts from a deeper shallow marine to fluvial settings over time. The Khewra Sandstone’s thickness variation across the eastern, central, and western Salt Range may highlight local tectonic influences or differences in subsidence rates influenced by the accommodation space available for sedimentation. The elemental proxies, including Sr/Ba, Rb/Sr, V/Cr, Ni/Co, and Cu/Zn (Figure 11), indicate deposition in well-oxygenated conditions. Therefore, considering all the available evidence, the depositional system of the Khewra Sandstone can be interpreted as a fluvio–deltaic to deltaic shallow marine environment, where sedimentation occurred under well-oxygenated floodplains, distributary channels, mouth bars, the delta front, and the delta plain (Figure 12).

7. Regional and Global Correlation

Cambrian red sandstones have a global scale distribution (Figure 13) and have remained the subject of several studies in various localities, including the Lalun Formation in the Shirgesht area of Central Iran, the Zabuk Formation in south-eastern Turkey (Türkiye), the Mt. Simon Sandstone in western Ohio, Midcontinent, and the Hickory Sandstone in Texas, USA, the Sandstone succession in the Podlasie region of East Poland, the Araba Formation in the east Sinai of Egypt, the Hardeberga Formation in Scania of southern Sweden, and sandstone from the Anti-Atlas region (Morocco), exhibit similarities with the Khewra Sandstone in the lithological, depositional environment, and paleoclimate characteristics.

7.1. Correlation with Iran and Turkey (Türkiye)

The Lalun Formation in the Shirgesht area, Central Iran, consists of felsic-rich, quartzose sandstones that suggest a mixed shallow-marine, fluvio–deltaic, and tide-dominated coastal depositional system. The texturally and compositionally immature strata of the formation indicate poor chemical weathering and thus arid to semi-arid conditions [246]. The formation represents Cambrian siliciclastics within a peri-Gondwana epicontinental basin situated along the northeast margin of the Arabian Shield [247,248]. Thus, both the Khewra Sandstone and the Lalun Formation were deposited on similar Gondwanan margins and represent fluvio–deltaic to shallow marine deposition under arid to semi-arid climates, with sediment influx from craton sources [247].
The Zabuk Formation of Turkey (Türkiye) is a thick succession of Early Cambrian, mixed, fluvial–shallow marine siliciclastic that was deposited on the northern margin of Gondwana. The detrital assemblage of the formation implies poor chemical weathering, suggesting an arid to semi-arid paleoclimate [249]. Thus, the formation represents an Early Cambrian fluvial–shallow marine mixed siliciclastic system on the northwestern margin of Gondwana (the northern margin of the Arabian Plate) similar to the Khewra Sandstone.

7.2. Correlation with the USA

The Cambrian Mt. Simon Sandstone in western Ohio represents siliciclastic sedimentation under tidally influenced transgressive barrier–fluvial systems with tidal channels and mixed flats. The formation is the product of broad continental interior deposition with poor chemical weathering under semi-arid to perihumid settings [250,251]. Similarly, the Hickory Sandstone is a siliciclastic Cambrian succession in Texas, USA, and consists of medium to coarse-grained sandstones, conglomerates, and siltstones. Cross-beds, graded beds, and ripple marks are the common features in the formation, and the formation was deposited in fluvial to deltaic shallow marine settings [252,253]. These formations indicate an Early Cambrian continental to marginal marine depositional system with coastal depositional energy and arid to semi-arid paleoclimatic conditions like those interpreted for the Khewra Sandstone in the present work.

7.3. Correlation with Egypt and Morocco

In eastern Sinai, Egypt, the Early Cambrian Araba Formation consists of sandstones, siltstone, and shale. Cross-bedding, ripple marks, and graded bedding are common sedimentary features. The textural and compositional immaturity of the formation indicates its deposition in a continental-shallow marine setting characterized by poor chemical weathering under an arid paleoclimate [254,255,256]. The formation presents lithological similarities with the Khewra Sandstone and was deposited on the Gondwanan margin with sediments sourced from the Arabian–Nubian Shield.
Similarly, the Cambrian sandstones from the Anti-Atlas region, Morocco, consist of poorly sorted quartz and unaltered feldspars. The formation represents fluvial–deltaic and sheet sands deposition under an arid to semi-arid Gondwanan setting [257]. The formation shares a similar provenance to the Khewra Sandstone, linked to Pan-African orogeny, and was deposited under a similar climatic regime.

7.4. Correlation with Sweden, Poland, and Norway

The Cambrian Hardeberga Formation in Scania, southern Sweden, displays sedimentary structures such as hummocky cross-stratification, ripples, and tabular cross-beddings. It is interpreted to have been deposited as a fluvio–deltaic shallow marine depositional unit under an arid to semi-arid paleoclimate on the Baltica continental margin [258]. Similarly, the Cambrian succession in the Podlasie region, East Poland, consists of conglomerate, sandstone, heterolith, and mudstone. Sedimentary features include massive, parallel lamination, cross-lamination, ripple cross-lamination, flaser and wavy lamination, and bioturbations, and indicate a fluvio–deltaic shallow marine environment. The presence of angular quartz and fresh feldspars indicates poor chemical weathering and sedimentation under an arid to semi-arid paleoclimate at the Baltic’s western rifted margin [259,260]. Thus, the Hardeberga Formation and Podlasie successions share striking similarities in depositional settings, weathering trends, and paleoclimatic conditions with the Khewra Sandstone. This allows correlation between the Early Cambrian sedimentation on the Baltica continental margin and the northwestern Indian Plate margin in a Gondwanan setting.
The Breidvika Formation from Northern Norway is dated as Lower Cambrian and comprises mudstone, siltstone, and sandstone. It includes sedimentary structures such as ripples, cross-bedding, and hummocky cross-stratification. The formation is suggested to have been deposited in a deltaic environment, showing cyclic coarsening upward sequences under well-oxygenated conditions [261]. Thus, the formation represents a deltaic-shallow marine Baltica equivalent of the Khewra Sandstone.
The Nagaur Sandstone from Rajasthan, India, is dated to the Lower Cambrian period [9]. It primarily consists of silty shale and sandstone. The formation is interpreted to have been deposited in a deltaic environment, and the region likely experienced a warm, arid to semi-arid climate during sedimentation. Similarly, the Amin Formation from North Oman spans from the Lower to Middle Cambrian and comprises sandstone, siltstone, shale, and conglomerate. The depositional environment for this formation is illustrated as an alluvial fan and aeolian, and the deposition occurred under arid to semi-arid climatic conditions [262,263]. Likewise, in Jordan, the Salib Formation from the Lower Cambrian strata includes sandstone, shale, and conglomerate. It was deposited as a fining upward sequence in braided stream–alluvial depositional settings under arid to semi-arid conditions [264]. These formations represent stratigraphic equivalents of the Khewra Sandstone in the nearby Gondwanan terrains and indicate sedimentary provenance associated with the Pan-African orogeny.
The lower Cambrian Cerros San Francisco Formation in Uruguay consists mainly of sandstones and siltstones. The depositional setting for this formation is described as a shallow marine environment with a fining upward sequence [265]. Also, the Billy Creek Formation, from the Lower to Middle Cambrian strata in South Australia, comprises claystone, siltstone, and salt pseudomorphs. Deposited in a fluvio–deltaic environment, this formation displays a coarsening upward trend [188,212]. Furthermore, Namibia’s Lower Cambrian Stockdale Formation consists of shale, sandstone, and calcareous nodules. The formation shows a fluvio–deltaic deposit with a coarsening upward sequence [266,267]. All these stratigraphic units reflect textural angularity and compositional immaturity that is attributed to poor chemical weathering under arid to semi-arid paleoclimatic conditions, similar to those that prevailed during the deposition of the Khewra Sandstone.
Figure 13. Correlation of Khewra Sandstone with regional and global red sandstones [9,187,246,249,250,251,254,257,259,261,263,264,265,266].
Figure 13. Correlation of Khewra Sandstone with regional and global red sandstones [9,187,246,249,250,251,254,257,259,261,263,264,265,266].
Minerals 15 00789 g013

8. Conclusions

The Lower Cambrian Khewra Sandstone (Salt Range, Pakistan) primarily consists of maroon–red sandstone with minor shale and siltstone in the lower sections.
  • The formation exhibits a coarsening upward succession with moderate to well-sorted, angular to sub-rounded, fine to coarse-grained, cross-bedded sandstones.
  • Lithofacies analysis and lithofacies associations support deposition environments that include fluvio–deltaic to shallow marine settings, encompassing distributary channels, channel margins, floodplains, delta plains, and mouth bars.
  • Framework mineralogy classifies the sandstones as sub-arkoses and sub-litharenites, indicating limited chemical weathering and deposition in arid to semi-arid paleoclimates.
  • XRD analysis advocates for sedimentation under arid to semi-arid climatic conditions.
  • Geochemical proxies support deposition in oxic, arid to semi-arid environments with poor to moderate chemical weathering intensity.
  • Lithological and environmental attributes, along with paleoclimate and weathering trends, align the Khewra Sandstone with globally recognized Cambrian red sandstone successions from the USA, Europe, Africa, Iran, Turkey, Arabia, and India.
These findings enhance the understanding of Early Cambrian sedimentary systems along the Gondwanan margin. Further research, including single-grain provenance studies, isotopic analysis, and geochronology, is recommended to clarify the formation’s relation to the Pan-African orogeny and the broader Gondwana Assembly.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min15080789/s1, Table S1: Percentage framework composition of the selected samples from all studied sections. Q = Quartz, Qm = Monocrystalline quartz, Qp = Polycrystalline quartz, Qt = Total quartz, Lt = Total lithics, Ls = Sedimentary lithics, Lv = Volcanic lithics, Lm = Metamorphic lithics, F = Feldspar, P = Plagioclase, K = Potassium feldspar, Ch = Chert, O = Others, C = Cements. Quartz-feldspar-lithics (QFL) values used in triangular plots are also included. Please note that 500 counts per thin section were used in the present study and percentage distribution has been presented in the present table. Table S2: Geochemical composition of selected Khewra Sandstone samples from different stratigraphic sections. All major oxides are in percentage (%), while trace elements are in Parts Per Million (ppm).

Author Contributions

Conceptualization: A.B.Q., S.I. and M.W.; Data curation: A.B.Q.; Formal analysis: A.B.Q., S.I., A.H.K. and M.I.; Funding acquisition: S.I., B.S. and M.W.; Investigation: A.B.Q. and M.I.; Methodology: A.B.Q., S.I., A.H.K. and M.I.; Project administration: S.I. and M.W.; Resources: S.I., A.H.K., B.S. and M.W.; Software: A.B.Q., S.I. and M.I.; Supervision: S.I.; Validation: A.B.Q., S.I. and B.S.; Visualization: S.I. and M.W.; Writing—original draft: A.B.Q.; Writing—review and editing: A.B.Q., S.I., B.S. and M.W. All authors have read and agreed to the published version of the manuscript.

Funding

Fieldwork Funding by Higher Education Commission (HEC) of Pakistan, Geochemical analysis Funding by Austrian Academy of Sciences through the project “NOMAJU”, under the framework of UNESCO-IGCP projects IGCP-710 and IGCP-732, Open Access Funding by the University of Vienna.

Data Availability Statement

Data have been provided as Supplementary Materials.

Acknowledgments

This research forms part of the PhD work of the corresponding author (Abdul Bari Qanit) and was supported by the Higher Education Commission (HEC) of Pakistan under the Allama Iqbal Scholarship Program for Afghan students. The logistical and technical assistance provided by the Department of Earth Sciences, Quaid-i-Azam University, Islamabad, Pakistan, and the Department of Geology, University of Vienna, Austria, is gratefully acknowledged. The insightful comments and suggestions from the editor and three anonymous reviewers significantly enhanced the quality of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
KGKhewra Gorge
NWNilawahan Gorge
VSVasnal Village
KZKhan Zaman Nala
XRDX-ray Diffraction
XRFX-ray Fluorescence
CMFShLChannel Margin and Floodplain Shale Lithofacies
CMSLChannel Margin Sandstone Lithofacies
DLDPLDelta Lobe/Delta Plain Lithofacies
CBSLChannel Belt Sandstone Lithofacies
CBCLChannel Belt Conglomerate Lithofacies
CMFLAChannel Margin and Floodplain Lithofacies Associations
DPLADelta Plain Lithofacies Association
CBLAsChannel Belt Lithofacies Associations
CIAChemical Index of Alteration
WIPWeathering Index of Parker

References

  1. Peng, S.C.; Babcock, L.E.; Ahlberg, P. The Cambrian period. Geol. Time Scale 2020 2020, 2, 565–629. [Google Scholar] [CrossRef]
  2. Babcock, L.E.; Peng, S.C.; Brett, C.E.; Zhu, M.Y.; Ahlberg, P.; Bevis, M.; Robison, R.A. Global climate, sea level cycles, and biotic events in the Cambrian Period. Palaeoworld 2015, 24, 5–15. [Google Scholar] [CrossRef]
  3. Babcock, L.; Peng, S.; Brett, C.E.; Zhu, M.; Ahlberg, P.; Bevis, M. Evidence of Global Climatic and Sea Level Cycles in the Cambrian; IGCP Project 591 Field Workshop; Nanjing University Press: Kunming, China, 2014; pp. 9–11. [Google Scholar]
  4. Haq, B.U.; Schutter, S.R. A chronology of Paleozoic sea-level changes. Science 2008, 322, 64–68. [Google Scholar] [CrossRef]
  5. Algeo, T.J.; Seslavinsky, K.B.; Wilkinson, B.H. The Paleozoic world: Continental flooding, hypsometry, and sea level. Am. J. Sci. 1995, 295, 787–822. [Google Scholar] [CrossRef]
  6. McKerrow, W.S.; Scotese, C.R.; Brasier, M.D. Early Cambrian continental reconstructions. J. Geol. Soc. 1992, 149, 599–606. [Google Scholar] [CrossRef]
  7. Maruyama, S.; Sawaki, Y.; Ebisuzaki, T.; Ikoma, M.; Omori, S.; Komabayashi, T. Initiation of leaking Earth: An ultimate trigger of the Cambrian explosion. Gondwana Res. 2014, 25, 910–944. [Google Scholar] [CrossRef]
  8. Shah, S.M.I. Stratigraphy of Pakistan (memoirs of the geological survey of Pakistan). Geol. Surv. Pak. 2009, 22, 93–114. [Google Scholar]
  9. Khan, S.H.; Sheng, Y.-M.; Mughal, M.S.; Singh, B.P.; Khan, M.R.; Zhang, C. Provenance of the Lower Cambrian Khewra Sandstone: Implications for Pan-African Orogeny. Sediment. Geol. 2022, 438, 106197. [Google Scholar] [CrossRef]
  10. Hughes, N.C.; Myrow, P.M.; Ghazi, S.; McKenzie, N.R.; Stockli, D.F.; DiPietro, J.A. Cambrian geology of the Salt Range of Pakistan: Linking the Himalayan margin to the Indian craton: Reply. Bulletin 2020, 132, 446–448. [Google Scholar] [CrossRef]
  11. Jehangiri, M.; Hanif, M.; Arif, M.; Jan, I.U.; Ahmad, S. The early Cambrian Khewra sandstone, salt range, Pakistan: Endorsing southern Indian provenance. Arab. J. Geosci. 2015, 8, 6169–6187. [Google Scholar] [CrossRef]
  12. Kazmi, A.H.; Abbasi, I.A. Stratigraphy & Historical Geology of Pakistan; Department and National Centre of Excellence in Geology: Peshawar, Pakistan, 2008; p. 524. [Google Scholar]
  13. Bassett, M.G. Early Palaeozoic peri-Gondwana terranes: New insights from tectonics and biogeography. Geol. Soc. Lond. Spéc. Publ. 2009, 325, 1–2. [Google Scholar] [CrossRef]
  14. Pankhurst, R.J.; Vaughan, A.P.M. The tectonic context of the Early Palaeozoic southern margin of Gondwana. Geol. Soc. Lond. Spec. Publ. 2009, 325, 171–176. [Google Scholar] [CrossRef]
  15. Keppie, D.F.; Keppie, J.D.; Landing, E. A tectonic solution for the Early Cambrian palaeogeographical enigma. Geol. Soc. Lond. Spec. Publ. 2024, 542, 167–177. [Google Scholar] [CrossRef]
  16. Ghazi, S.; Sharif, S.; Zafar, T.; Riaz, M.; Haider, R.; Hanif, T. Sedimentology and Stratigraphic evolution of the early Eocene Nammal Formation, Salt Range, Pakistan. Stratigr. Geol. Correl. 2020, 28, 745–764. [Google Scholar] [CrossRef]
  17. Richards, L.; Jourdan, F.; Collins, A.S.; King, R.C. Deformation recorded in polyhalite from evaporite detachments revealed by 40Ar/39Ar dating. Geochronology 2021, 3, 545–559. [Google Scholar] [CrossRef]
  18. Bordet, P. The western border of the Indian plate: Implications for Himalayan geology. Tectonophysics 1978, 51, T71–T76. [Google Scholar] [CrossRef]
  19. Myrow, P.M.; Snell, K.E.; Hughes, N.C.; Paulsen, T.S.; Heim, N.A.; Parcha, S.K. Cambrian depositional history of the Zanskar Valley region of the Indian Himalaya: Tectonic implications. J. Sediment. Res. 2006, 76, 364–381. [Google Scholar] [CrossRef]
  20. Hussain, S.A.; Han, F.-Q.; Han, J.; Khan, H.; Widory, D. Chlorine isotopes unravel conditions of formation of the Neoproterozoic rock salts from the Salt Range Formation, Pakistan. Can. J. Earth Sci. 2020, 57, 698–708. [Google Scholar] [CrossRef]
  21. Chatterjee, S. India’s northward drift from Gondwana to Asia during the Late Cretaceous-Eocene. Proc. Indian Natl. Sci. Acad. 2016, 82, 479–487. [Google Scholar] [CrossRef]
  22. Liu, Q.; Tsunogae, T.; Zhao, G.; Uthup, S.; Takahashi, K.; Yao, J.; Wu, Y.; Han, Y.; Ikehata, K. Early Cambrian high pressure/low temperature metamorphism in the southeastern Tarim craton in response to circum-Gondwana cold subduction. Geosci. Front. 2023, 14, 101561. [Google Scholar] [CrossRef]
  23. Witzke, B.J. Palaeoclimatic constraints for Palaeozoic palaeolatitudes of Laurentia and Euramerica. Geol. Soc. Lond. Mem. 1990, 12, 57–73. [Google Scholar] [CrossRef]
  24. Chumakov, N.M. Climates and climate zonality of the Vendian: Geological evidence. Geol. Soc. Lond. Spec. Publ. 2007, 286, 15–26. [Google Scholar] [CrossRef]
  25. Daraei, M.; Bayet-Goll, A.; Geyer, G.; Bahrami, N. Late Cambrian climate change recorded by a shift from an arid carbonate platform to a storm-dominated cool-water platform at the Gondwana margin (Alborz Zone, Iran). Geol. J. 2023, 58, 795–824. [Google Scholar] [CrossRef]
  26. Bhattacharjee, J.; Ghosh, K.K.; Bhattacharya, B. Petrography and geochemistry of sandstone--mudstone from Barakar Formation (early Permian), Raniganj Basin, India: Implications for provenance, weathering and marine depositional conditions during Lower Gondwana sedimentation. Geol. J. 2018, 53, 1102–1122. [Google Scholar] [CrossRef]
  27. Mosebach, R. Khewarite vom Khewra Gorge, Pakistan, ein neucr types kalircicher efiusiugesteine. Neues Jahrb. Mineral. Geol. Palaeont. 1956, 89, 182–207. [Google Scholar]
  28. Fatmi, A.N. Lithostratigraphic Units of the Kohat-Potwar Province, Indus Basin, Pakistan; Memoirs of the Geological Survey of Pakistan; The Geological Survey of Pakistan: Karachi, Pakistan, 1974; 10, pp. 1–80.
  29. Wynne, A.B. On the geology of the Salt Range. Punjab. Ibid. Mem. 1878, 14, 313. [Google Scholar]
  30. Noetling, F. On the Cambrian formation ofthe eastem Salt Range. Geol. Surv. India Rec. 1894, 27 Pt 4, 71–86. [Google Scholar]
  31. Nichols, G. Sedimentology and Stratigraphy Blackwell Publishing; John Wiley & Sons, Ltd. Publication: London, UK, 1999; p. 411. [Google Scholar]
  32. Gazzi, P. Le arenarie del flysch sopracretaceo dell’Appennio modenese: Correlazioni con il flysch di Monghidero. Mineral. Petrogr. Acta 1966, 12, 69–97. [Google Scholar] [CrossRef]
  33. Dickinson, W.R. Interpreting detrital modes of graywacke and arkose. J. Sediment. Res. 1970, 40, 695–707. [Google Scholar] [CrossRef]
  34. Graham, S.A.; Ingersoll, R.V.; Dickinson, W.R. Common provenance for lithic grains in Carboniferous sandstones from Ouachita Mountains and Black Warrior Basin. J. Sediment. Res. 1976, 46, 620–632. [Google Scholar] [CrossRef]
  35. Dickinson, W.R.; Suczek, C.A. Plate tectonics and sandstone compositions. Am. Assoc. Pet. Geol. Bull. 1979, 63, 2164–2182. [Google Scholar] [CrossRef]
  36. Sares, S.W. The Effect of Grain Size on Detrital Modes. J. Sediment. Res. 1984, 54, 103–106. [Google Scholar] [CrossRef]
  37. Gazzi, P.; Zuffa, G.G.; Gandolfi, G.; Paganelli, L. Provenienza e dispersione litoranea delle sabbie delle spiagge adriatiche fra le foci dell’Isonzo e del Foglia: Inquadramento regionale. Mem. Della Soc. Geol. Ital. 1973, 12, 1–37. [Google Scholar]
  38. Suttner, L.J.; Basu, A.; Mack, G.H. Climate and the origin of quartz arenites. J. Sediment. Res. 1981, 51, 1235–1246. [Google Scholar] [CrossRef]
  39. Suttner, L.J.; Dutta, P.K. Alluvial Sandstone Composition and Paleoclimate; I, Framework Mineralogy. J. Sediment. Petrol. 1986, 56, 329–345. [Google Scholar] [CrossRef]
  40. Weltje, G.J. Provenance and dispersal of sand-sized sediment: Reconstruction of dispersal patterns and sources of sand-sized sediments by means of inverse modelling techniques. Geol. Ultraiectina 1994, 121, 1–208. [Google Scholar]
  41. Adhikari, S.K.; Sakai, T. Petrography of the Neogene Siwalik Group sandstones, Khutia Khola section, Nepal Himalaya: Implications for provenance, paleoclimate and tectonic setting. J. Nepal. Geol. Soc. 2017, 53, 17–30. [Google Scholar] [CrossRef]
  42. Iqbal, S.; Wagreich, M.; Jan, U.I.; Kuerschner, W.M.; Gier, S.; Bibi, M. Hot-house climate during the Triassic/Jurassic transition: The evidence of climate change from the southern hemisphere (Salt Range, Pakistan). Glob. Planet. Change 2019, 172, 15–32. [Google Scholar] [CrossRef]
  43. Ogbahon, O.A.; Olujinmi, O.B. Geochemistry of Maastrichtian clastic sedimentary rocks from Western flank of Anambra Basin, Nigeria: Implications for provenance, tectonic setting, paleoclimate and depositional paleoenvironment. Int. J. Geosci. 2019, 10, 91–118. [Google Scholar] [CrossRef]
  44. Borgohain, P.; Hussain, M.F.; Bezbaruah, D.; Vanthangliana, V.; Phukan, P.P.; Gogoi, M.P.; Bharali, B. Petrography and whole-rock geochemistry of Oligocene Barail Sandstones of Surma basin: Implications for tectono-provenance and paleoclimatic condition. J. Earth Syst. Sci. 2020, 129, 1–26. [Google Scholar] [CrossRef]
  45. Alqahtani, F.; Khalil, M. Geochemical analysis for evaluating the paleoweathering, paleoclimate, and depositional environments of the siliciclastic Miocene-Pliocene sequence at Al-Rehaili area, Northern Jeddah, Saudi Arabia. Arab. J. Geosci. 2021, 14, 239. [Google Scholar] [CrossRef]
  46. Kairouani, H.; Abbassi, A.; Zaghloul, M.N.; El Mourabet, M.; Micheletti, F.; Fornelli, A.; Mongelli, G.; Critelli, S. The Jurassic climate change in the northwest Gondwana (External Rif, Morocco): Evidence from geochemistry and implication for paleoclimate evolution. Mar. Pet. Geol. 2024, 163, 106762. [Google Scholar] [CrossRef]
  47. Nesbitt, H.W.; Young, G.M. Formation and diagenesis of weathering profiles. J. Geol. 1989, 97, 129–147. [Google Scholar] [CrossRef]
  48. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Blackwell Scientific Publications: Oxford, UK, 1985; pp. 1–328. [Google Scholar] [CrossRef]
  49. Zhao, Z.Y.; Zhao, J.H.; Wang, H.J.; Liao, J.D.; Liu, C.M. Distribution characteristics and applications of trace elements in Junggar Basin. Nat. Gas Explor. Dev. 2007, 30, 30–32. [Google Scholar] [CrossRef]
  50. Goldberg, K.; Humayun, M. The applicability of the Chemical Index of Alteration as a paleoclimatic indicator: An example from the Permian of the Paraná Basin, Brazil. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2010, 293, 175–183. [Google Scholar] [CrossRef]
  51. Basu, A. Petrology of Holocene fluvial sand derived from plutonic source rocks; implications to paleoclimatic interpretation. J. Sediment. Res. 1976, 46, 694–709. [Google Scholar] [CrossRef]
  52. Nesbitt, H.W.; Young, G.M.; McLennan, S.M.; Keays, R.R. Effects of chemical weathering and sorting on the petrogenesis of siliciclastic sediments, with implications for provenance studies. J. Geol. 1996, 104, 525–542. [Google Scholar] [CrossRef]
  53. El Bilali, L.; Rasmussen, P.E.; Hall, G.E.M.; Fortin, D. Role of sediment composition in trace metal distribution in lake sediments. Appl. Geochem. 2002, 17, 1171–1181. [Google Scholar] [CrossRef]
  54. Chen, L.; Lu, Y.C.; Wu, J.Y.; Xing, F.C.; Liu, L.; Ma, Y.Q.; Rao, D.; Peng, L. Sedimentary facies and depositional model of shallow water delta dominated by fluvial for Chang 8 oil-bearing group of Yanchang Formation in southwestern Ordos Basin, China. J. Cent. South Univ. 2015, 22, 4749–4763. [Google Scholar] [CrossRef]
  55. Lerman, A.; Imboden, D.M.; Gat, J.R.; Chou, L. Physics and Chemistry of Lakes; Springer: Berlin/Heidelberg, Germany, 1995; p. 138. [Google Scholar]
  56. Zuo, X.; Li, C.; Zhang, J.; Ma, G.; Chen, P. Geochemical characteristics and depositional environment of the Shahejie Formation in the Binnan Oilfield, China. J. Geophys. Eng. 2020, 17, 539–551. [Google Scholar] [CrossRef]
  57. Ernst, W.G.; Wai, C.M. Mössbauer, infrared, X-ray and optical study of cation ordering and dehydrogenation in natural and heat-treated sodic amphiboles. Am. Mineral. J. Earth Planet. Mater. 1970, 55, 1226–1258. [Google Scholar] [CrossRef]
  58. Rimmer, S.M. Geochemical paleoredox indicators in Devonian–Mississippian black shales, central Appalachian Basin (USA). Chem. Geol. 2004, 206, 373–391. [Google Scholar] [CrossRef]
  59. Cai, G.Q.; Guo, F.; Liu, X.T.; Sui, S.L. Carbon and oxygen isotope characteristics and palaeoenvironmental implications of lacustrine carbonate rocks from the Shahejie Formation in the Dongying Sag. Earth Environ. 2009, 37, 347–354. [Google Scholar]
  60. Fielding, C.R. Fluvial channel and overbank deposits from the Westphalian of the Durham coalfield, NE England. Sedimentology 1986, 33, 119–140. [Google Scholar] [CrossRef]
  61. Wu, C.; Ullah, M.S.; Lu, J.; Bhattacharya, J.P. Formation of point bars through rising and falling flood stages: Evidence from bar morphology, sediment transport and bed shear stress. Sedimentology 2016, 63, 1458–1473. [Google Scholar] [CrossRef]
  62. Walker, R.G. Facies Model: Response to Sea Level Change; The Geological Association of Canada: St. John’s, NL, Canada, 1992; pp. 119–139. [Google Scholar]
  63. Bilobe, J.A.; Feist-Burkhardt, S.; Eyong, J.T.; Samankassou, E. Biostratigraphy of Cretaceous-Neogene sedimentary infill of the Mamfe basin, southwest Cameroon: Paleoclimate implication. J. Afr. Earth Sci. 2021, 182, 104279. [Google Scholar] [CrossRef]
  64. McLennan, S.M.; Hemming, S.; McDaniel, D.K.; Hanson, G.N. Geochemical Approaches to Sedimentation, Provenance, and Tectonics. In Processes Controlling the Composition of Clastic Sediments; Johnsson, M.J., Basu, A., Eds.; Geological Society of America Special Publications: Boulder, CO, USA, 1993; Volume 284, pp. 21–40. [Google Scholar] [CrossRef]
  65. Bibi, M.; Wagreich, M.; Iqbal, S.; Gier, S.; Jan, I.U. Sedimentation and glaciations during the Pleistocene: Palaeoclimate reconstruction in the Peshawar Basin, Pakistan. Geol. J. 2020, 55, 671–693. [Google Scholar] [CrossRef]
  66. Jun, C.; Yongjin, W.; Yang, C.; Lianwen, L.I.U.; Junfeng, J.I.; Huayu, L.U. Rb and Sr geochemical characterization of the Chinese loess stratigraphy and its implications for palaeomonsoon climate. Acta Geol. Sin. Ed. 2000, 74, 279–288. [Google Scholar] [CrossRef]
  67. Lerman, A.; Hull, A.B. Background aspects of lake restoration: Water balance, heavy metal content, phosphorus homeostasis. Swiss J. Hydrol. 1987, 49, 148–169. [Google Scholar] [CrossRef]
  68. Jones, C.E.; Jenkyns, H.C.; Coe, A.L.; Stephen, H.P. Strontium isotopic variations in Jurassic and Cretaceous seawater. Geochim. Cosmochim. Acta 1994, 58, 3061–3074. [Google Scholar] [CrossRef]
  69. Postma, G. Depositional Architecture and Facies of River and Fan Deltas: A Synthesis. Coarse-Grained Deltas; John Wiley & Sons: Hoboken, NJ, USA, 1990; pp. 13–27. [Google Scholar] [CrossRef]
  70. Fielding, C.R.; Trueman, J.D.; Alexander, J. Sharp-Based, Flood-Dominated Mouth Bar Sands from the Burdekin River Delta of Northeastern Australia: Extending the Spectrum of Mouth-Bar Facies, Geometry, and Stacking Patterns. J. Sediment. Res. 2005, 75, 55–66. [Google Scholar] [CrossRef]
  71. Kendall, B.; Creaser, R.A.; Reinhard, C.T.; Lyons, T.W.; Anbar, A.D. Transient episodes of mild environmental oxygenation and oxidative continental weathering during the late Archean. Sci. Adv. 2015, 1, e1500777. [Google Scholar] [CrossRef]
  72. Ahmad, K.; Hussain, S.A.; Al-Obaidi, M. Paleoceanographic Reconstruction of Upper Cretaceous, Black Shale Succession, Northeastern Iraq Using Geochemical Proxies to Indicate Paleoredox and Paleoenvironment Conditions. Diyala J. Pure Sci. 2018, 14, 237–264. [Google Scholar] [CrossRef]
  73. Scibiorski, J.; Peyrot, D.; Lang, S.C.; Payenberg, T.H.D.; Payenberg, T.H.D.; Charles, A. Depositional settings and palynofacies assemblages of the Upper Triassic fluvio-deltaic Mungaroo Formation, northern Carnarvon Basin, Western Australia. J. Sediment. Res. 2020, 90, 403–428. [Google Scholar] [CrossRef]
  74. Mou, C.; Wang, X.-P.; Wang, Q.-Y.; Ge, X.; Zan, B.; Zhou, K.-K. Lithofacies Paleogeography and Geological Survey of Shale Gas; Springer Nature: Dordrecht, The Netherlands, 2023. [Google Scholar] [CrossRef]
  75. Yang, R.; He, S.; Wang, X.; Hu, Q.; Hu, D.; Yi, J. Paleo-ocean redox environments of the Upper Ordovician Wufeng and the first member in lower Silurian Longmaxi formations in the Jiaoshiba area, Sichuan Basin. Can. J. Earth Sci. 2016, 53, 426–440. [Google Scholar] [CrossRef]
  76. Woo, J.; Lee, H.S.; Ozyer, C.A.; Rhee, C.W. Effect of Lamination on Shale Reservoir Properties: Case Study of the Montney Formation, Canada. Geofluids 2021, 2021, 1–14. [Google Scholar] [CrossRef]
  77. Zhao, X.; Xie, D.; Jina, F.; Pub, X.; Hanb, W.; Shib, Z.; Zhangb, W.; Dong, J. Laminae Features, Genesis of Deep Lacustrine Shales and the Implications for Shale Oil Mobility in the Paleogene Kongdian Formation of Cangdong Sag; Bohai Bay Basin, China. SSRN 2023, 1–34. [Google Scholar] [CrossRef]
  78. O’Brien, N.R. Shale lamination and sedimentary processes. Geol. Soc. Lond. Spec. Publ. 1996, 116, 23–36. [Google Scholar] [CrossRef]
  79. Leonowicz, P. Nearshore transgressive black shale from the Middle Jurassic shallow-marine succession from southern Poland. Facies 2016, 62, 16. [Google Scholar] [CrossRef]
  80. Einsele, G. Depositional Rhythms and Cyclic Sequences; Springer: Berlin/Heidelberg, Germany, 1992; pp. 271–310. [Google Scholar] [CrossRef]
  81. Gorsline, D.S. A review of fine-grained sediment origins, characteristics, transport and deposition. Geol. Soc. Lond. Spec. Publ. 1984, 15, 17–34. [Google Scholar] [CrossRef]
  82. Gugliotta, M.; Saito, Y.; Saito, Y.; Nguyen, V.L.; Ta, T.K.O.; Tamura, T. Sediment distribution and depositional processes along the fluvial to marine transition zone of the Mekong River delta, Vietnam. Sedimentology 2019, 66, 146–164. [Google Scholar] [CrossRef]
  83. Scholle, P.A.; Spearing, D.R. Sandstone Depositional Environments; AAPG Memoir 31 (No. 31); AAPG: London, UK, 1982; p. 405. [Google Scholar]
  84. Oyanyan, R.O.; Oti, M.N. Sedimentology and ichnology of late Oligocene delta front reservoir sandstone deposit, greater ughelli depobelt, Niger delta. Am. J. Geosci. 2015, 5, 12. [Google Scholar] [CrossRef]
  85. Breckenridge, J.; Maravelis, A.G.; Catuneanu, O.; Ruming, K.; Holmes, E.; Collins, W.J. Outcrop analysis and facies model of an Upper Permian tidally influenced fluvio-deltaic system: Northern Sydney Basin, SE Australia. Geol. Mag. 2019, 156, 1715–1741. [Google Scholar] [CrossRef]
  86. Boon, J.D. Tidal discharge asymmetry in a salt marsh drainage system. Limnol. Oceanogr. 1975, 20, 71–80. [Google Scholar] [CrossRef]
  87. Hoitink, A.J.F.; Hoekstra, P.; van Maren, D.S. Flow asymmetry associated with astronomical tides: Implications for the residual transport of sediment. J. Geophys. Res. 2003, 108, 3315. [Google Scholar] [CrossRef]
  88. Huang, H.; Chen, C.; Blanton, J.O.; Andrade, F.H. A numerical study of tidal asymmetry in Okatee Creek, South Carolina. Estuar. Coast. Shelf Sci. 2008, 78, 190–202. [Google Scholar] [CrossRef]
  89. Owen, G. Load structures: Gravity-driven sediment mobilization in the shallow subsurface. Geol. Soc. Lond. Spec. Publ. 2003, 216, 21–34. [Google Scholar] [CrossRef]
  90. Duquesne, A.; Carozza, J.M. Improving Grain Size Analysis to Characterize Sedimentary Processes in a Low-Energy River: A Case Study of the Charente River (Southwest France). Appl. Sci. 2023, 13, 8061. [Google Scholar] [CrossRef]
  91. Boukhalfa, K.; Soussi, M.; Reynaud, J.Y.; Banerjee, S. Sandstones and red-beds in the Lower Cretaceous Sidi Aich Formation, Chotts basin (Southern Tunisia): Facies, architecture and depositional environments. Geol. Soc. Lond. Spec. Publ. 2023, 545, 293–323. [Google Scholar] [CrossRef]
  92. Shah, K.S.; Hashim, M.H.B.M.; Rehman, H.; Ariffin, K.S. Effect of Wet-Dry Cycling on the Microstructure of Various Weathering Grade Sandstone. Appl. Mech. Mater. 2024, 920, 183–187. [Google Scholar] [CrossRef]
  93. Yu, W.; Liang, Q.L.; Tian, J.; Han, Y.; Wang, F.; Zhao, M. Sedimentary Responses of Late Triassic Soft-Sedimentary Deformation to Paleoearthquake Events in the Southwestern North China Plate. Minerals 2022, 12, 1044. [Google Scholar] [CrossRef]
  94. Al-Saqarat, B.S.; Abbas, M.; Al Hseinat, M.; Qutishat, T.A.; Shammar, D.; AlShamaileh, E. Interpreting Soft-Sediment Deformation Structures: Insights into Earthquake History and Depositional Processes in the Dead Sea, Jordan. Geosciences 2024, 14, 217. [Google Scholar] [CrossRef]
  95. Oman, G. Soft-sediment deformation structures in the Late Messinian Abu Madi Formation, onshore Nile Delta, Egypt: Triggers and tectonostratigraphic implications. Geol. J. 2022, 57, 2302–2320. [Google Scholar] [CrossRef]
  96. Reading, H.G. Sedimentary Environments: Processes, Facies and Stratigraphy; John Wiley & Sons: Hoboken, NJ, USA, 2009; p. 669. [Google Scholar]
  97. Ghinassi, M.; Moody, J.A.; Martin, D. Influence of extreme and annual floods on point-bar sedimentation: Inferences from Powder River, Montana, USA. Geol. Soc. Am. Bull. 2019, 131, 71–83. [Google Scholar] [CrossRef]
  98. Willis, B.J.; Tang, H. Three-Dimensional Connectivity of Point-Bar Deposits. J. Sediment. Res. 2010, 80, 440–454. [Google Scholar] [CrossRef]
  99. Jablonski, B.V.J.; Dalrymple, R.W. Recognition of strong seasonality and climatic cyclicity in an ancient, fluvially dominated, tidally influenced point bar: Middle McMurray Formation, Lower Steepbank River, north-eastern Alberta, Canada. Sedimentology 2016, 63, 552–585. [Google Scholar] [CrossRef]
  100. Norwood, E.M.; Holland, D.S. Lithofacies Mapping a Descriptive Tool for Ancient Delta Systems of the Louisiana Outer Continental Shelf. GCAGS Trans. 1974, 24, 175–188. [Google Scholar]
  101. Swenson, J.B.; Paola, C.; Pratson, L.F.; Voller, V.R.; Murray, A.B. Fluvial and marine controls on combined subaerial and subaqueous delta progradation: Morphodynamic modeling of compound-clinoform development. J. Geophys. Res. 2005, 110. [Google Scholar] [CrossRef]
  102. La Fontaine, N.M.; Hofmann, M.H. Quantifying the stratigraphic and spatial facies distribution in an ancient mixed-influence delta. Mt. Geol. 2019, 56, 19–44. [Google Scholar] [CrossRef]
  103. Jones, R.; Marcelissen, R.; Fralick, P. Sedimentology and Stratigraphy of a Large, Pre-Vegetation Deltaic Complex. Front. Earth Sci. 2022, 10, 875838. [Google Scholar] [CrossRef]
  104. Jacks, C.; Kamola, D.L. Variability in river-dominated deltaic systems: Facies models and drivers based on the Late Cretaceous Panther Tongue Sandstone. Central Utah. In Proceedings of the Second International Meeting for Applied Geoscience & Energy, Houston, TX, USA, 28 August–1 September 2022. [Google Scholar] [CrossRef]
  105. Guala, M. On the scaling and growth limit of fluvial dunes. J. Geophys. Res. Earth Surf. 2023, 128. [Google Scholar] [CrossRef]
  106. Innocent, E.O.; Anthony, A.I.; Etobro, A.A.I. Depositional Setting of Sandstones from the Oligocene-Miocene Ogwashi-Asaba Formation, Niger Delta Basin, Nigeria: Evidence from Grain Size Analysis and Geochemistry. Univers. J. Geosci. 2015, 3, 71–82. [Google Scholar] [CrossRef]
  107. O’Brien, K.C. Stratigraphic Architecture of a Shallow-Water Delta Deposited in a Coastal-Plain Setting: Neslen Formation, Floy Canyon, Utah; Colorado School of Mines: Golden, CO, USA, 2015; p. 1602532. [Google Scholar] [CrossRef]
  108. Alexander, J.; Fielding, C.R. Coarse-Grained Floodplain Deposits in the Seasonal Tropics: Towards a Better Facies Model. J. Sediment. Res. 2006, 76, 539–556. [Google Scholar] [CrossRef]
  109. Ahmad, M.Z.; Singh, P. Geochemistry of core sediments from Cauvery delta South-East India: Inferences on weathering and paleo-redox conditions. Quaternaire 2023, 34, 1–25. [Google Scholar] [CrossRef]
  110. Yuan, X.; Guerit, L.; Braun, J.; Rouby, D.; Shobe, C.M. Thickness of Fluvial Deposits Records Climate Oscillations. J. Geophys. Res. Solid Earth 2022, 127, e2021JB023510. [Google Scholar] [CrossRef]
  111. Cardenas, B.T.; Stacey, K.; Baran, Z.J. Relationships between fluvial dune cross-set thickness, planview width, and trough geometry. Geology 2023, 51, 1163–1167. [Google Scholar] [CrossRef]
  112. De Alvaro, M.M. Architecture and Origin of Fluvial Cross-Bedding Based on Flume Experiments and Geological Examples Field Case Studies: Rillo de Gallo, Spain and Northumberland, UK. PhD Thesis, University of East Anglia, Norwich, UK, 2015. [Google Scholar]
  113. Khan, Z.A.; Tewari, R.C. Lithofacies correlation in Early Permian fluvial Gondwana stratigraphy of southeastern India using cross-association statistics. J. Earth Syst. Sci. 2019, 128, 3. [Google Scholar] [CrossRef]
  114. Alam, M.M.; Crook, K.A.W.; Taylor, G. Fluvial herring-bone cross-stratification in a modern tributary mouth bar, Coonamble, New South Wales, Australia. Sedimentology 1985, 32, 235–244. [Google Scholar] [CrossRef]
  115. Malmon, D.V.; Reneau, S.L.; Dunne, T. Sediment sorting and transport by flash floods. J. Geophys. Res. 2004, 109. [Google Scholar] [CrossRef]
  116. Ikhane, P.R.; Oladipo, O.V.; Oyebolu, O.O. Predicting the depositional environments and transportation mechanisms of sediments using granulometric parameters, bivariate and multivariate analyses. Geosci. Eng. 2019, 65, 11–28. [Google Scholar] [CrossRef]
  117. Mader, D. Depositional Mechanisms Controlling Formation of Coarse Fluvial Conglomerates in the Lower Triassic Continental Red Beds of Middle Europe; Springer: Berlin/Heidelberg, Germany, 1985; pp. 251–280. [Google Scholar] [CrossRef]
  118. Restrepo, J.C.; Schrottke, K.; Traini, C.; Ortíz, J.C.; Orejarena, A.; Otero, L.; Higgins, A.; Marriaga, L. Sediment Transport and Geomorphological Change in a High-Discharge Tropical Delta (Magdalena River, Colombia): Insights from a Period of Intense Change and Human Intervention (1990–2010). J. Coast. Res. 2016, 32, 575–589. [Google Scholar] [CrossRef]
  119. Goswami, P.K.; Deopa, T. Lithofacies characters and depositional processes of a Middle Miocene Lower Siwalik fluvial system of the Himalayan foreland basin, India. J. Asian Earth Sci. 2017, 162, 41–53. [Google Scholar] [CrossRef]
  120. Colombera, L.; Mountney, N.P. The lithofacies organization of fluvial channel deposits: A meta-analysis of modern rivers. Sediment. Geol. 2019, 383, 16–40. [Google Scholar] [CrossRef]
  121. Gani, M.R.; Bhattacharya, J.P. Basic Building Blocks and Process Variability of a Cretaceous Delta: Internal Facies Architecture Reveals a More Dynamic Interaction of River, Wave, and Tidal Processes Than Is Indicated by External Shape. J. Sediment. Res. 2007, 77, 284–302. [Google Scholar] [CrossRef]
  122. Gwinn, V.E. Deduction of flow regime from bedding character in conglomerates and sandstones. J. Sediment. Res. 1964, 34, 656–658. [Google Scholar] [CrossRef]
  123. Evans, J.E.; Holm-Denoma, C.S. Processes and facies relationships in a Lower(?) Devonian rocky shoreline depositional environment, East Lime Creek Conglomerate, south-western Colorado, USA. Depos. Rec. 2018, 4, 133–156. [Google Scholar] [CrossRef]
  124. Boulay, S.; Colin, C.; Trentesaux, A.; Clain, S.; Liu, Z.; Lauer-Leredde, C. Sedimentary responses to the Pleistocene climatic variations recorded in the South China Sea. Quat. Res. 2007, 68, 162–172. [Google Scholar] [CrossRef]
  125. Yi, L.; Yu, H.; Ortiz, J.D.; Xu, X.; Qiang, X.; Huang, H.; Shi, X.; Deng, C. A reconstruction of late Pleistocene relative sea level in the south Bohai Sea, China, based on sediment grain-size analysis. Sediment. Geol. 2012, 281, 88–100. [Google Scholar] [CrossRef]
  126. Ning, W.; Tang, J.; Filipsson, H.L. Long-term coastal openness variation and its impact on sediment grain-size distribution: A case study from the Baltic Sea. Earth Surf. Dyn. 2016, 4, 773–780. [Google Scholar] [CrossRef]
  127. Li, M.; Ouyang, T.; Tian, C.; Zhu, Z.; Peng, S.; Tang, Z.; Qiu, Y.; Zhong, H.; Peng, X. Sedimentary responses to the East Asian monsoon and sea level variations recorded in the northern South China Sea over the past 36 kyr. J. Asian Earth Sci. 2019, 171, 213–224. [Google Scholar] [CrossRef]
  128. Hay, M.J.; Plint, A.G. High-frequency sequences within a retrogradational deltaic succession: Upper Cenomanian Dunvegan Formation, Western Canada Foreland Basin. Depos. Rec. 2020, 6, 524–551. [Google Scholar] [CrossRef]
  129. Leila, M.; Moscariello, A.; Kora, M.; Mohamed, A.; Samankassou, E. Sedimentology and reservoir quality of a Messinian mixed siliciclastic-carbonate succession, onshore Nile Delta, Egypt. Mar. Pet. Geol. 2020, 112, 104076. [Google Scholar] [CrossRef]
  130. Adnan, A.; Shukla, U.K. A case of normal regression with sea level transgression: Example from the Ganurgarh shale, Vindhyan basin, Maihar area, M.P. India. J. Geol. Soc. India 2014, 84, 406–416. [Google Scholar] [CrossRef]
  131. Okoro, A.U.; Igwe, E.O. Lithofacies and depositional environment of the Amasiri Sandstone, southern Benue Trough, Nigeria. J. Afr. Earth Sci. 2014, 100, 179–190. [Google Scholar] [CrossRef]
  132. Mode, A.W. Sequence stratigraphy and depositional environments of Middle-Late Miocene sediments in the eastern part of the Coastal Swamp depobelt, Niger Delta Basin, Nigeria. Arab. J. Geosci. 2015, 8, 9815–9827. [Google Scholar] [CrossRef]
  133. Blanco-ferrera, S.; Sanz-lópez, J.; Domínguez-cuesta, M.J.; López-fernández, C.; Alberto, L.; Martos, E. Fundamentos Conceptaules y Didácticos Transgresiones, regresiones y fósiles Transgressions, regressions and fossils. Enseñ. Cienc. Rev. Investig. Exp. Didáct. 2019, 27, 18–30. [Google Scholar]
  134. Sweet, D.E.; Soreghan, G.S. Estimating magnitudes of relative sea-level change in a coarse-grained fan delta system: Implications for Pennsylvanian glacioeustasy. Geology 2012, 40, 979–982. [Google Scholar] [CrossRef]
  135. Skene, K.I.; Piper, D.J.W.; Aksu, A.E.; Syvitski, J.P.M. Evaluation of the global oxygen isotope curve as a proxy for Quaternary sea level by modeling of delta progradation. J. Sediment. Res. 1998, 68, 1077–1092. [Google Scholar] [CrossRef]
  136. Liu, J.; Saito, Y.; Kong, X.; Wang, H.; Wen, C.; Yang, Z.; Nakashima, R. Delta development and channel incision during marine isotope stages 3 and 2 in the western South Yellow Sea. Mar. Geol. 2010, 278, 54–76. [Google Scholar] [CrossRef]
  137. van Cappelle, M.; Ravnås, R.; Hampson, G.J.; Johnson, H.D. Depositional evolution of a progradational to aggradational, mixed-influenced deltaic succession: Jurassic Tofte and Ile formations, southern Halten Terrace, offshore Norway. Mar. Pet. Geol. 2017, 80, 1–22. [Google Scholar] [CrossRef]
  138. Obi, I.S.; Onuoha, K.M.; Dim, C.I.P. Highlighting relationships between sand thicknesses, reservoir-seal pairs and paleobathymetry from a sequence stratigraphic perspective: An example from Tortonian Serravallian deposits, onshore Niger Delta Basin. Energy Geosci. 2024, 5, 100215. [Google Scholar] [CrossRef]
  139. Pettijohn, F.J.; Potter, P.E.; Siever, R. Sand and Sandstone; Springer Science and Business Media: Berlin/Heidelberg, Germany, 2012; p. 552. [Google Scholar]
  140. Odumoso, S.E.; Oloto, I.N.; Omoboriowo, A.O. Sedimentological and depositional enviroment of the Mid-Maastrichtian Ajali Sandstone, Anambra Basin, Southern Nigeria. Int. J. Sci. Technol. 2013, 3, 26–33. [Google Scholar]
  141. Tucker, M.E.; Jones, S.J. Sedimentary Petrology; John Wiley & Sons: New York, NY, USA, 2023; p. 418. [Google Scholar]
  142. Olivarius, M.; Rasmussen, E.S.; Siersma, V.; Knudsen, C.; Pedersen, G.K. Distinguishing fluvio-deltaic facies by bulk geochemistry and heavy minerals: An example from the Miocene of Denmark. Sedimentology 2011, 58, 1155–1179. [Google Scholar] [CrossRef]
  143. Canale, N.; Ponce, J.J.; Carmona, N.B.; Drittanti, D.I.; Olivera, D.E.; Martínez, M.A.; Bournod, C.N. Sedimentología e icnología de deltas fluvio-dominados afectados por descargas hiperpícnicas de la formación lajas (Jurásico medio), cuenca neuquina, Argentina. Andean Geol. 2015, 42, 114–138. [Google Scholar] [CrossRef]
  144. Machado, G.M.V.; Albino, J.; Leal, A.P.; Bastos, A.C. Quartz grain assessment for reconstructing the coastal palaeoenvironment. J. S. Am. Earth Sci. 2016, 70, 353–367. [Google Scholar] [CrossRef]
  145. Shchepetkina, A.; Gingras, M.K.; Mángano, M.G.; Buatois, L.A. Fluvio-tidal transition zone: Terminology, sedimentological and ichnological characteristics, and significance. Earth-Sci. Rev. 2019, 192, 214–235. [Google Scholar] [CrossRef]
  146. Mazumdar, P.; Mukhopadhyay, A.; Banerjee, T.; Thorie, A.; Eriksson, P.G. Process variations of a Neoproterozoic delta system in response to the influence of mixed-energy systems and sea-level fluctuation: An insight from Simla Group, Lesser Himalaya, India. Arab. J. Geosci. 2021, 14, 596. [Google Scholar] [CrossRef]
  147. Ogbe, O.B. Reservoir sandstone grain-size distributions: Implications for sequence stratigraphic and reservoir depositional modelling in Otovwe field, onshore Niger Delta Basin, Nigeria. J. Pet. Sci. Eng. 2021, 203, 108639. [Google Scholar] [CrossRef]
  148. Gresina, F.; Farkas, B.; Fábián, S.Á.; Szalai, Z.; Varga, G. Comparison of fluvial and aeolian sedimentary environments based on morphological analysis of their mineral components. In Proceedings of the EGU General Assembly Conference Abstracts, Vienna, Austria, 23–28 April 2023; p. EGU-13637. [Google Scholar]
  149. Kasim, S.A.; Ismail, M.S.; Ahmed, N. Grain size statistics and morphometric analysis of Kluang-Niyor, Layang-Layang, and Kampung Durian Chondong tertiary sediments, onshore peninsular Malaysia: Implications for paleoenvironment and depositional processes. J. King Saud. Univ. 2023, 35, 102481. [Google Scholar] [CrossRef]
  150. Liang, P.; Yang, X. Grain Shape Evolution of Sand-Sized Sediments During Transport From Mountains to Dune Fields. J. Geophys. Res. Earth Surf. 2023, 128, e2022JF006930. [Google Scholar] [CrossRef]
  151. Gaafar, G.R.; Altunbay, M.; Bal, A.; Anuar, N.B. Ascendancy of continuous profiles of grain-size distribution for depositional environment studies. In Proceedings of the International Petroleum Technology Conference, Kuala Lumpur, Malaysia, 10–12 December 2014; p. IPTC-17754. [Google Scholar]
  152. Prodger, S.; Russell, P.; Davidson, M. Grain-size distributions on high-energy sandy beaches and their relation to wave dissipation. Sedimentology 2017, 64, 1289–1302. [Google Scholar] [CrossRef]
  153. Reynolds, T. ‘Grain-size bookkeeping,’ a new aid for siliciclastic systems with examples from paralic environments. J. Sediment. Res. 2019, 89, 976–1016. [Google Scholar] [CrossRef]
  154. Dong, D.T.; Qiu, L.W.; Ma, P.J.; Yu, G.D.; Wang, Y.Z.; Zhou, S.B.; Yang, B.L.; Huang, H.Q.; Yang, Y.Q.; Li, X. Initiation and evolution of coarse-grained deposits in the Late Quaternary Lake Chenghai source-to-sink system: From subaqueous colluvial apron (subaqueous fans) to Gilbert-type delta. J. Palaeogeogr. 2022, 11, 194–221. [Google Scholar] [CrossRef]
  155. Davies, D.K.; Ethridge, F.G. Sandstone composition and depositional environment. Am. Assoc. Pet. Geol. Bull. 1975, 59, 239–264. [Google Scholar] [CrossRef]
  156. Amaral, E.J.; Pryor, W.A. Depositional environment of the St. Peter sandstone deduced by textural analysis. J. Sediment. Res. 1977, 47, 32–52. [Google Scholar] [CrossRef]
  157. Cox, R.; Lowe, D.R. Quantification of the effects of secondary matrix on the analysis of sandstone composition, and a petrographic-chemical technique for retrieving original framework grain modes of altered sandstones. J. Sediment. Res. 1996, 66, 548–558. [Google Scholar] [CrossRef]
  158. Baioumy, H.M.; Boulis, S.N. Glauconites from the Bahariya Oasis: An evidence for Cenomanian marine transgression in Egypt. J. Afr. Earth Sci. 2012, 70, 1–7. [Google Scholar] [CrossRef]
  159. Seraine, M.; Campos, J.E.G.; Martins-Ferreira, M.A.C.; Giorgioni, M.; Angelo, T.V. Tectonic significance of abrupt immature sedimentation in a shallow cratonic margin basin: The Arkose Level, Mesoproterozoic Paranoá Group. J. S. Am. Earth Sci. 2020, 97, 102397. [Google Scholar] [CrossRef]
  160. Akinlotan, O.O.; Adepehin, E.J.; Rogers, G.H.; Drumm, E.C. Provenance, palaeoclimate and palaeoenvironments of a non-marine Lower Cretaceous facies: Petrographic evidence from the Wealden Succession. Sediment. Geol. 2021, 415, 105848. [Google Scholar] [CrossRef]
  161. Boboye, O.A.; Jaiyeoba, O.K.; Okon, E.E. Sedimentological characteristics and mineralogical studies of some Cretaceous sandstones in Nigeria: Implications for depositional environment and provenance. J. Sediment. Environ. 2021, 6, 531–550. [Google Scholar] [CrossRef]
  162. Zhang, M.; Zhang, Y.; Li, X.; Wang, J.; Meng, G.; Shi, B. Mineral compositions of soil in the arid and semiarid region and their environmental significance. J.-Lanzhou Univ. Nat. Sci. 2007, 43, 1. [Google Scholar]
  163. Pszonka; Wendorff, M. Cathodoluminescence-revealed diagenesis of carbonates and feldspars in Cergowa sandstones (Oligocene), Outer Carpathians. Gospod. Surowcami Miner.-Miner. Resour. Manag. 2014, 30, 21–36. [Google Scholar] [CrossRef]
  164. Pszonka; Wendorff, M. Carbonate cements and grains in submarine fan sandstones—The Cergowa Beds (Oligocene, Carpathians of Poland)recorded by cathodoluminescence. Int. J. Earth Sci. 2017, 64, 28–53. [Google Scholar] [CrossRef]
  165. Ding, Z.; Gong, S.; Xiao, G.; Wang, Y.; Yuan, W.; Zhang, J.; Wang, J.; Lai, Z. Episodic sediment accumulation linked to global change in the endorheic Qaidam Basin of the Tibetan Plateau revealed by feldspar luminescence dating. Quat. Geochronol. 2024, 81, 101522. [Google Scholar] [CrossRef]
  166. Madhavaraju, J.; Armstrong-Altrin, J.S.; James, R.A.; Hussain, S.M. Palaeoenvironment and provenance signatures inferred from quartz grain surface features: A case study from Huatabampo and Altata beaches, Gulf of California, Mexico. J. S. Am. Earth Sci. 2021, 111, 103441. [Google Scholar] [CrossRef]
  167. Babu, V.; Rai, S.K.; Bhan, U.; Devaraju, J. Chemical weathering of sediment (CWS): A web-based application for chemical weathering study of rocks and sediment. Mater. Today Proc. 2022, 68, 979–985. [Google Scholar] [CrossRef]
  168. Iqbal, S.; Wagreich, M.; Bibi, M.; Jan, I.U.; Gier, S. Multi-proxy provenance analyses of the Kingriali and Datta formations (Triassic—Jurassic transition): Evidence for westward extension of the neo-Tethys passive margin from the salt range (Pakistan). Minerals 2021, 11, 573. [Google Scholar] [CrossRef]
  169. Iqbal, S.; Bibi, M.; Wagreich, M. Geochemistry of the Triassic--Jurassic lateritic bauxites of the Salt Range: Implications for eastward extension of the Tethyan bauxite deposits into Pakistan. Int. J. Earth Sci. 2023, 112, 1527–1552. [Google Scholar] [CrossRef]
  170. Basu, A.; Mookherjee, M.; Clapp, S.; Chariton, S.; Prakapenka, V.B. High-pressure Raman scattering and X-ray diffraction study of kaolinite, Al2Si2O5(OH)4. Appl. Clay Sci. 2023, 245, 107144. [Google Scholar] [CrossRef]
  171. Fagel, N.; Israde-Alcántara, I.; Safaierad, R.; Rantala, M.; Schmidt, S.; Lepoint, G.; Pellenard, P.; Mattielli, N.; Metcalfe, S. Environmental significance of kaolinite variability over the last centuries in crater lake sediments from Central Mexico. Appl. Clay Sci. 2024, 247, 107211. [Google Scholar] [CrossRef]
  172. Guo, L.; Wu, J.; Chen, Y.; Xiong, S.; Cui, J.; Ding, Z. Modern silicate weathering regimes across China revealed by geochemical records from surface soils. J. Geophys. Res. Earth Surf. 2022, 127, e2022JF006728. [Google Scholar] [CrossRef]
  173. Xu, G.; Shen, J.; Algeo, T.J.; Yu, J.; Feng, Q.; Frank, T.D.; Fielding, C.R.; Yan, J.; Deconink, J.F.; Lei, Y. Limited change in silicate chemical weathering intensity during the Permian--Triassic transition indicates ineffective climate regulation by weathering feedbacks. Earth Planet. Sci. Lett. 2023, 616, 118235. [Google Scholar] [CrossRef]
  174. Yuan, G.; Cao, Y.; Schulz, H.M.; Hao, F.; Gluyas, J.; Liu, K.; Yang, T.; Wang, Y.; Xi, K.; Li, F. A review of feldspar alteration and its geological significance in sedimentary basins: From shallow aquifers to deep hydrocarbon reservoirs. Earth-Sci. Rev. 2019, 191, 114–140. [Google Scholar] [CrossRef]
  175. Bartz, M.; Peña, J.; Grand, S.; King, G.E. Potential impacts of chemical weathering on feldspar luminescence dating properties. Geochronol. Discuss. 2023, 5, 51–64. [Google Scholar] [CrossRef]
  176. Yhasnara, M.; Costa, P.J.M.; Dourado, F.; Martins, M.V.A.; Feist, L.; Bellanova, P.; Reicherter, K. Microtextural signatures in quartz grains and foraminifera from tsunami deposits of the Portuguese shelf. Geo-Marine Lett. 2023, 43, 5. [Google Scholar] [CrossRef]
  177. Marra, K.R.; Madden, M.E.E.; Soreghan, G.S.; Hall, B.L. Chemical weathering trends in fine-grained ephemeral stream sediments of the McMurdo Dry Valleys, Antarctica. Geomorphology 2017, 281, 13–30. [Google Scholar] [CrossRef]
  178. Uren, A.L.; Laukamp, C.; George, A.D.; Occhipinti, S.A.; Aitken, A.R.A. Inferring sandstone grain size using spectral datasets: An example from the Bresnahan Group, Western Australia. Remote Sens. Environ. 2021, 252, 112109. [Google Scholar] [CrossRef]
  179. Wu, K.; Liu, S.; Shi, X.; Colin, C.; Bassinot, F.; Lou, Z.; Zhang, H.; Zhu, A.; Fang, X.; Mohamed, C.A.R. The Effect of Size Distribution on the Geochemistry and Mineralogy of Tropical River Sediments and Its Implications regarding Chemical Weathering and Fractionation of Alkali Elements. Lithosphere 2022, 2022, 8425818. [Google Scholar] [CrossRef]
  180. Janssen, M.; Caracciolo, L.; Bonnell, L.M.; Lander, R.H.; Munnecke, A.; Beltrán-Triviño, A.; Muto, F.; Stollhofen, H. Climatic, depositional and environmental controls on early carbonate cementation in fluvial and shallow marine sandstones. Mar. Pet. Geol. 2023, 156, 106433. [Google Scholar] [CrossRef]
  181. Drummond, J.B.R.; Pufahl, P.K.; James, N.P.; Layton-Matthews, D.; Kyser, T.K. Sedimentary dolomite in Western Australia and the dolomite problem: Genesis of channel and playa uranium deposits. Sedimentology 2024, 71, 2248–2289. [Google Scholar] [CrossRef]
  182. Hu, Z.; Bialik, O.M.; Hohl, S.V.; Xia, Z.; Waldmann, N.D.; Liu, C.; Li, W. Response of Mg isotopes to dolomitization during fluctuations in sea level: Constraints on the hydrological conditions of massive dolomitization systems. Sediment. Geol. 2021, 420, 105922. [Google Scholar] [CrossRef]
  183. Raiswell, R.; Hardisty, D.S.; Lyons, T.W.; Canfield, D.E.; Owens, J.D.; Planavsky, N.J.; Poulton, S.W.; Reinhard, C.T. The iron paleoredox proxies: A guide to the pitfalls, problems and proper practice. Am. J. Sci. 2018, 318, 491–526. [Google Scholar] [CrossRef]
  184. Sindol, G.P.; Babechuk, M.G.; Conliffe, J.; Slack, J.F.; Rosca, C.; Schoenberg, R. Shallow-ocean and atmospheric redox signatures preserved in the ca. 1.88 Ga Sokoman iron formation, Labrador Trough, Canada. Precambrian Res. 2022, 379, 106750. [Google Scholar] [CrossRef]
  185. Song, Q.; Hong, H.; Algeo, T.J.; Fang, Q.; Zhao, C.; Liu, C.; Xu, Y. Clay mineralogy mediated by pH and chemical weathering intensity of Permian--Triassic boundary K-bentonites at Dongpan (Guangxi, South China). Chem. Geol. 2023, 617, 121262. [Google Scholar] [CrossRef]
  186. Ling, C.; Liu, Z.; Yu, X.; Zhao, Y.; Siringan, F.P.; Le, K.P.; Sathiamurthy, E.; You, C.F.; Chen, K. Clay minerals control silicon isotope variations of fine-grained river sediments: Implication for the trade-off between physical erosion and chemical weathering. Chem. Geol. 2024, 662, 122249. [Google Scholar] [CrossRef]
  187. Moore, P.S. Stratigraphy and Sedimentology of the Billy Creek Formation (Cambrian, Flinders Ranges) and Its Equivalents on the Northeast Coast of Kangaroo Island, South Australia. Ph.D. Thesis, University of Adelaide, Adelaide, Australia, 1979. Available online: https://digital.library.adelaide.edu.au/items/60329f31-468a-47e5-aa6b-679280bacec4 (accessed on 21 July 2025).
  188. Singer, A.; Galan, E. Palygorskite-Sepiolite: Occurrences, Genesis and Uses; Elsevier: Amsterdam, The Netherlands, 2000; p. 339. [Google Scholar]
  189. Papoulis, D.; Tsolis-Katagas, P.; Katagas, C. Progressive stages in the formation of kaolin minerals of different morphologies in the weathering of plagioclase. Clays Clay Miner. 2004, 52, 275–286. [Google Scholar] [CrossRef]
  190. Deon, F.; van Ruitenbeek, F.; van der Werff, H.; van der Meijde, M.; Marcatelli, C. Detection of interlayered Illite/smectite clay minerals with XRD, SEM analyses and reflectance spectroscopy. Sensors 2022, 22, 3602. [Google Scholar] [CrossRef]
  191. Azzam, F.; Blaise, T.; Patrier, P.; Beaufort, D.; Barbarand, J.; Elmola, A.A.; Brigaud, B.; Portier, E.; Clerc, S. Impact of sediment provenance and depositional setting on chlorite content in Cretaceous turbiditic sandstones, Norway. Basin Res. 2024, 36, e12867. [Google Scholar] [CrossRef]
  192. Šegvić, B.; Benvenuti, A.; Moscariello, A. Illite-smectite-rich clay parageneses from quaternary tunnel valley sediments of the Dutch Southern North Sea—Mineral origin and paleoenvironment implications. Clays Clay Miner. 2016, 64, 608–627. [Google Scholar] [CrossRef]
  193. Rong, K.; Zeng, Z.; Yin, X.; Chen, S.; Wang, X.; Qi, H.; Ma, Y. Smectite formation in metalliferous sediments near the East Pacific Rise at 13 N. Acta Oceanol. Sin. 2018, 37, 67–81. [Google Scholar] [CrossRef]
  194. Yin, K.; Hong, H.; Churchman, G.J.; Li, Z.; Fang, Q. Mixed-layer illite-vermiculite as a paleoclimatic indicator in the Pleistocene red soil sediments in Jiujiang, southern China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2018, 510, 140–151. [Google Scholar] [CrossRef]
  195. Guo, N.; Guo, W.; Shi, W.; Huang, Y.; Guo, Y.; Lian, D. Characterization of Illite Clays associated with the Sinongduo low sulfidation epithermal deposit, Central Tibet using field SWIR spectrometry. Ore Geol. Rev. 2020, 120, 103228. [Google Scholar] [CrossRef]
  196. Nesbitt, H.W.; Young, G.M. Prediction of some weathering trends of plutonic and volcanic rocks based on thermodynamic and kinetic considerations. Geochim. Cosmochim. Acta 1984, 48, 1523–1534. [Google Scholar] [CrossRef]
  197. Aswad, M. Calculation of Mineralogical and Chemical Weathering Indices (Xd, MIA and CIA) and their Significance in Soils at Selected Areas in Northern Iraq. Iraqi Natl. J. Earth Sci. 2022, 22, 1–14. [Google Scholar] [CrossRef]
  198. Wang, C.; Ding, L.; Cai, F.; Zhang, L.; Li, Z.; Yue, Y. Rifting of the Indian passive continental margin: Insights from the Langjiexue basalts in the central Tethyan Himalaya, southern Tibet. Bulletin 2022, 134, 2633–2648. [Google Scholar] [CrossRef]
  199. Bahlburg, H.; Dobrzinski, N. Chapter 6 A review of the Chemical Index of Alteration (CIA) and its application to the study of Neoproterozoic glacial deposits and climate transitions. Geol. Soc. Lond. Mem. 2011, 36, 81–92. [Google Scholar] [CrossRef]
  200. Shao, J.; Yang, S. Does chemical index of alteration (CIA) reflect silicate weathering and monsoonal climate in the Changjiang River basin? Chin. Sci. Bull. 2012, 57, 1178–1187. [Google Scholar] [CrossRef]
  201. Partridge, T.C. Paleoclimates of the Arid and Semi-Arid Zones of Southern Africa During the Last Climatic Cycle; Société géologique de France: Paris, France, 1995; pp. 73–83. [Google Scholar]
  202. Hoelzmann, P.; Gasse, F.; Dupont, L.M.; Salzmann, U.; Staubwasser, M.; Leuschner, D.C.; Sirocko, F. Palaeoenvironmental Changes in the Arid and Sub Arid Belt (Sahara-Sahel-Arabian Peninsula) from 150 kyr to Present; Springer: Berlin/Heidelberg, Germany, 2004; pp. 219–256. [Google Scholar] [CrossRef]
  203. Cheng, H.; Zhang, P.Z.; Spotl, C.; Edwards, R.L.; Cai, Y.J.; Zhang, D.Z.; Sang, W.C.; Tan, M.; An, Z.S. The climatic cyclicity in semiarid-arid central Asia over the past 500,000 years. Geophys. Res. Lett. 2012, 39, 1–5. [Google Scholar] [CrossRef]
  204. Spötl, C.; Houseknecht, D.W.; Jaques, R. Clay Mineralogy and illite crystallinity of the Atoka Formation, Arkoma Basin, and frontal Ouachita Mountains. Clays Clay Miner. 1993, 41, 745–754. [Google Scholar] [CrossRef]
  205. Modi, A.L. Processes Controlling the Composition of First-Cycle Sediments Deposited in an Arid-Climate, with Implications for Provenance Reconstruction Studies; University of Tennessee–Knoxville: Knoxville, TN, USA, 2011. [Google Scholar]
  206. Warke, P. Weathering in Arid Regions; Academic Press: Cambridge, MA, USA, 2013; pp. 197–227. [Google Scholar] [CrossRef]
  207. Zhu, B.; Yang, X. Chemical Weathering of Detrital Sediments in the Taklamakan Desert, Northwestern China. Geogr. Res. 2009, 47, 57–70. [Google Scholar] [CrossRef]
  208. Schoonejans, J.; Vanacker, V.; Opfergelt, S.; Ameijeiras-Mariño, Y.; Christl, M. Kinetically limited weathering at low denudation rates in semiarid climatic conditions. J. Geophys. Res. 2016, 121, 336–350. [Google Scholar] [CrossRef]
  209. Kalinin, P.I.; Kudrevatykh, I.Y.; Malyshev, V.V.; Pilguy, L.S.; Buhonov, A.V.; Mitenko, G.V.; Alekseev, A.O. Chemical weathering in semi-arid soils of the Russian plain. Catena 2021, 206, 105554. [Google Scholar] [CrossRef]
  210. Yuan, W.; Liu, X.; Fu, Y. Study on deterioration of strength parameters of sandstone under the action of dry-wet cycles in acid and alkaline environment. Arab. J. Sci. Eng. 2018, 43, 335–348. [Google Scholar] [CrossRef]
  211. Retallack, G.J. Cambrian paleosols and landscapes of South Australia. Aust. J. Earth Sci. 2008, 55, 1083–1106. [Google Scholar] [CrossRef]
  212. Nesbitt; Young, G.M. Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 1982, 299, 715–717. [Google Scholar] [CrossRef]
  213. Paikaray, S.; Banerjee, S.; Mukherji, S. Geochemistry of shales from the Paleoproterozoic to Neoproterozoic Vindhyan Supergroup: Implications on provenance, tectonics and paleoweathering. J. Asian Earth Sci. 2008, 32, 34–48. [Google Scholar] [CrossRef]
  214. El Mourabet, M.; Barakat, A.; Zaghloul, M.N.; El Baghdadi, M. Geochemistry of the Miocene Zoumi flysch thrust-top basin (External Rif, Morocco): New constraints on source area weathering, recycling processes, and paleoclimate conditions. Arab. J. Geosci. 2018, 11, 126. [Google Scholar] [CrossRef]
  215. Usui, Y.; Yamazaki, T. K-feldspar enrichment in the Pacific pelagic sediments before Miocene. Prog. Earth Planet. Sci. 2023, 10, 1–12. [Google Scholar] [CrossRef]
  216. Furlan, S.; Clauer, N.; Chaudhuri, S.; Sommer, F. K transfer during burial diagenesis in the mahakam delta basin (kalimantan, Indonesia). Clays Clay Miner. 1996, 44, 157–169. [Google Scholar] [CrossRef]
  217. Tank, R. Clay mineral composition of the Tipton shale member of the Green River Formation (Eocene) of Wyoming. J. Sediment. Res. 1969, 39, 1593–1595. [Google Scholar] [CrossRef]
  218. Okunlola, O.A.; Egbulem, C. Geological Setting, Compositional and Economic Appraisal of Clay-Shale Occurrence in Itu-Mbonuso/Iwere Area, South-Eastern Nigeria. J. Geogr. Geol. 2015, 7, 85. [Google Scholar] [CrossRef]
  219. Graham, E.R. The plagioclase feldspars as an index to soil weathering. Soil. Sci. Soc. Am. J. 1950, 14, 300–302. [Google Scholar] [CrossRef]
  220. Howell, J.A.; Mountney, N.P. Climatic cyclicity and accommodation space in arid to semi-arid depositional systems: An example from the Rotliegend Group of the UK southern North Sea. Geol. Soc. Lond. Spec. Publ. 1997, 123, 63–86. [Google Scholar] [CrossRef]
  221. Tunçay, T.; Dengiz, O.; Bayramin, İ.; Kilic, S.; Başkan, O. Chemical weathering indices applied to soils developed on old lake sediments in a semi-arid region of Turkey. Eurasian J. Soil Sci. 2019, 8, 60–72. [Google Scholar] [CrossRef]
  222. Garzanti, E.; Padoan, M.; Setti, M.; Najman, Y.; Peruta, L.; Villa, I.M. Weathering geochemistry and Sr-Nd fingerprints of equatorial upper Nile and Congo muds. Geochem. Geophys. Geosyst. 2013, 14, 292–316. [Google Scholar] [CrossRef]
  223. Dai, A.; Song, M. Little influence of Arctic amplification on mid-latitude climate. Nat. Clim. Change 2020, 10, 231–237. [Google Scholar] [CrossRef]
  224. Roy, P.D.; Caballero, M.; Lozano, R.; Smykatz-Kloss, W. Geochemistry of late quaternary sediments from Tecocomulco lake, central Mexico: Implication to chemical weathering and provenance. Geochemistry 2008, 68, 383–393. [Google Scholar] [CrossRef]
  225. Wang, Y.Y.; Guo, W.Y.; Zhang, G. Application of some geochemical indicators in determining of sedimentary environment of the Funing Group (Paleogene), Jin-Hu Depression, Kiangsu Province. J. Tongji Univ. 1979, 7, 51–60. [Google Scholar]
  226. Thorpe, C.L.; Lloyd, J.R.; Law, G.T.; Burke, I.T.; Shaw, S.; Bryan, N.D.; Morris, K. Strontium sorption and precipitation behaviour during bioreduction in nitrate impacted sediments. Chem. Geol. 2012, 306, 114–122. [Google Scholar] [CrossRef]
  227. Liu, B.J. Sedimentary Petrology; Geological Publishing House: Beijing, China, 1980. [Google Scholar]
  228. Liu, Y.J.; Cao, L.M.; Li, Z.; Wang, H.N.; Chu, T.Q.; Zhang, J.R. Element Geochemistry; Science China Press: Beijing, China, 1984; p. 365. [Google Scholar]
  229. Qian, L.J.; Chen, H.D.; Lin, L.B.; Xu, S.L.; Ou, L.H. Geochemical characteristics and environmental implications of Middle Jurassic Shaximiao Formation, western margin of Sichuan Basin. Acta Sedimentol. Sin. 2012, 30, 1061–1071. [Google Scholar]
  230. Kessler, L.G. Diagenetic sequence in ancient sandstones deposited under desert climatic conditions. J. Geol. Soc. 1978, 135, 41–49. [Google Scholar] [CrossRef]
  231. Fei, G.; Li, B.; Yang, J.; Chen, X.; Luo, W.; Li, Y.; Tang, W.; Gu, C.; Zhong, W.; Yang, G. Geology, Fluid Inclusion Characteristics and H-O-C Isotopes of Large Lijiagou Pegmatite Spodumene Deposit in Songpan-Garze Fold Belt, Eastern Tibet: Implications for ore Genesis. Resour. Geol. 2018, 68, 37–50. [Google Scholar] [CrossRef]
  232. Huang, S.; Lin, C.; Zhang, X.; Zhang, N. Controls of diagenesis on the quality of shallowly buried terrestrial coarse-grained clastic reservoirs: A case study of the Eocene Shahejie Formation in the Damintun Sag, Bohai Bay Basin, Eastern China. J. Asian Earth Sci. 2021, 221, 104950. [Google Scholar] [CrossRef]
  233. Sricharoenvech, P.; Siebecker, M.G.; Tappero, R.; Landrot, G.; Fischel, M.H.; Sparks, D.L. Chromium speciation and mobility in contaminated coastal urban soils affected by water salinity and redox conditions. J. Hazard. Mater. 2023, 462, 132661. [Google Scholar] [CrossRef]
  234. Babu, S.S.; Rao, V.P.; Satyasree, N.; Ramana, R.V.; Mohan, M.R.; Sawant, S.S. Mineralogy and geochemistry of the sediments in rivers along the east coast of India: Inferences on weathering and provenance. J. Earth Syst. Sci. 2021, 130, 1–24. [Google Scholar] [CrossRef]
  235. Rahman, M.M.; Hasan, M.F.; Hasan, A.S.M.M.; Alam, M.S.; Biswas, P.K.; Zaman, M.N. Chemical weathering, provenance, and tectonic setting inferred from recently deposited sediments of Dharla River, Bangladesh. J. Sediment. Environ. 2021, 6, 73–91. [Google Scholar] [CrossRef]
  236. Gablina, I.F.; Malinovskii, Y.M. Periodicity of copper accumulation in the Earth’s sedimentary shell. Lithol. Miner. Resour. 2008, 43, 136–153. [Google Scholar] [CrossRef]
  237. Volkov, A.V.; Ras, G.; Novikov, I.A.; Razumovsky, A.A.; Murashov, K.Y.; Sidorova, N.V. Geochemical features and formation conditions of the cupriferous sandstones of the Orenburg Pre-Urals. Lithosphere 2018, 18, 593–606. [Google Scholar] [CrossRef]
  238. Holse, C.; Elkjær, C.F.; Nierhoff, A.; Sehested, J.; Chorkendorff, I.; Helveg, S.; Nielsen, J.H. Dynamic Behavior of CuZn Nanoparticles under Oxidizing and Reducing Conditions. J. Phys. Chem. C 2014, 119, 2804–2812. [Google Scholar] [CrossRef]
  239. Shanmugam, G. Significance of framework dissolution in interpreting sandstone provenance. Am. Assoc. Pet. Geol. Bull. 1985, 69, 146. [Google Scholar] [CrossRef]
  240. Sharma, R.P.; Raja, P.; Bhaskar, B.P. Pedogenesis and mineralogy of alluvial soils from semi-arid southeastern part of Rajasthan in Aravalli range, India. J. Geol. Soc. India 2020, 95, 59–66. [Google Scholar] [CrossRef]
  241. Polzer, W.L.; Hem, J.D. The dissolution of kaolinite. J. Geophys. Res. 1965, 70, 6233–6240. [Google Scholar] [CrossRef]
  242. Dur, J.C.; Wiriyakitnateekul, W.; Lesturgez, G.; Elsass, F.; Pernes, M.; Hartmann, C.; Tessier, D. Clay mineral dissolution following intensive cultivation in a tropical sandy soil. Laser 2005, 350, 9.5–13.1. Available online: https://www.fao.org/4/ag125e/AG125E12.htm (accessed on 21 July 2025).
  243. Saha, S.; Reza, A.H.M.S.; Roy, M.K. Illite crystallinity index an indicator of physical weathering of the Sediments of the Tista River, Rangpur, Bangladesh. Int. J. Adv. Geosci. 2020, 8, 27–32. [Google Scholar] [CrossRef]
  244. Sun, Y.; Li, Y.; Zhang, C.; Qin, X.; Shen, J.; He, H.; Pan, Y. Weathering of Chlorite Illite Deposits in the Hyperarid Qaidam Basin: Implications to Post-Depositional Alteration on Martian Clay Minerals. Front. Astron. Sp. Sci. 2022, 9, 875547. [Google Scholar] [CrossRef]
  245. Poursoltani, M.R. Architectural analysis of an Early Cambrian braided-river system on the north Gondwana margin: The lower sandstone of the Lalun Formation in the Shirgesht area, central Iran. J. Afr. Earth Sci. 2020, 171, 103935. [Google Scholar] [CrossRef]
  246. Bayet-Goll, A.; Geyer, G.; Daraei, M. Tectonic and eustatic controls on the spatial distribution and stratigraphic architecture of late early Cambrian successions at the northern Gondwana margin: The siliciclastic-carbonate successions of the Lalun Formation in central Iran. Mar. Pet. Geol. 2018, 98, 199–228. [Google Scholar] [CrossRef]
  247. Poursoltani, M.R.; Gibling, M.R.; Pe-Piper, G. Diagenesis, burial history, and hydrocarbon potential of Cambrian sandstone in the northern continental margin of Gondwana: A case study of the Lalun Formation of central Iran. J. Asian Earth Sci. 2019, 172, 143–169. [Google Scholar] [CrossRef]
  248. Mergl, M.; Hoşgör, İ.; Yilmaz, I.O.; Zamora, S.; Colmenar, J. Divaricate patterns in Cambro-Ordovician obolid brachiopods from Divaricate patterns in Cambro-Ordovician obolid brachiopods from Gondwana. Hist. Biol. 2017, 30, 1015–1029. [Google Scholar] [CrossRef]
  249. Gürsu, S.; Köksal, S.; Möller, A.; Kamenov, G.D.; Göncüoğlu, M.C.; Hefferan, K.; Mueller, P.A.; Kozlu, H. Combined U-Pb ages and Lu-Hf systematics of detrital zircons from Early Cambrian Gondwanan siliciclastic rocks in S Turkey: Provenance and correlations with coeval successions in peri-Gondwanan terranes. Gondwana Res. 2022, 107, 423–450. [Google Scholar] [CrossRef]
  250. Saeed, A.; Evans, J.E. Subsurface facies analysis of the late Cambrian Mt. Simon Sandstone in western Ohio (midcontinent North America). Open J. Geol. 2012, 2, 35–47. [Google Scholar] [CrossRef]
  251. Cornish, F.G. Tidally influenced deposits of the Hickory Sandstone, Cambrian, Central Texas; Texas Scholar Works; University of Texas Libraries: Austin, TX, USA, 1975; Available online: http://hdl.handle.net/2152/20401 (accessed on 21 July 2025).
  252. Teran, I.A.P. Stratal Architecture and Sedimentology of a Portion of the Upper Cambrian Hickory Sandstone, Central Texas, USA.; Texas A & M University: Bizzell St, TX, USA, 2010. [Google Scholar]
  253. El-Araby, A.; Abdel-Motelib, A. Depositional facies of the Cambrian Araba Formation in the Taba region, east Sinai, Egypt. J. Afr. Earth Sci. 1999, 29, 429–447. [Google Scholar] [CrossRef]
  254. Khalifa, M.A.; Soliman, H.E.; Wanas, H.A. The Cambrian Araba Formation in northeastern Egypt: Facies and depositional environments. J. Asian Earth Sci. 2006, 27, 873–884. [Google Scholar] [CrossRef]
  255. Tawfik, H.A.; Ghandour, I.M.; Maejima, W.; Abdel-Hameed, A.T. Petrography and geochemistry of the Lower Paleozoic Araba Formation, northern Eastern Desert, Egypt: Implications for provenance, tectonic setting and weathering signature. J. Geosci. Osaka City Univ. 2011, 54, 1–16. [Google Scholar] [CrossRef]
  256. Benssaou, M.; Hamoumi, N. The western Anti-Atlas of Morocco: Sedimentological and palaeogeographical formation studies in the Early Cambrian. J. Afr. Earth Sci. 2001, 32, 351–372. [Google Scholar] [CrossRef]
  257. Hamberg, L. Tidal and Seasonal Cycles in a Lower Cambrian Shallow Marine Sandstone (Hardeberga Fm.); Mem. 16, 255-273; CSPG Special Publication: Scania, Sweden, 1991. [Google Scholar]
  258. Golonka, J.; Porębski, S.J.; Barmuta, J.; Papiernik, B.; Bębenek, S.; Barmuta, M.; Botor, D.; Pietsch, K.; Słomka, T. Palaeozoic palaeogeography of the East European Craton (Poland) in the framework of global plate tectonics. Ann. Soc. Geol. Pol. 2019, 87, 381–403. [Google Scholar] [CrossRef]
  259. Wendorff, M. Facies architecture of the Cambrian succession at the western margin of Baltica in the Podlasie region (E Poland). Ann. Soc. Geol. Pol. 2019, 89, 453–469. [Google Scholar] [CrossRef]
  260. McIlroy, D.; Brasier, M.D. Ichnological evidence for the Cambrian explosion in the Ediacaran to Cambrian succession of Tanafjord, Finnmark, northern Norway. Geol. Soc. Lond. Spéc. Publ. 2016, 448, 351–368. [Google Scholar] [CrossRef]
  261. Droste, H.H.J. Stratigraphy of the lower Paleozoic Haima supergroup of Oman. GeoArabia 1997, 2, 419–472. [Google Scholar] [CrossRef]
  262. Al-Husseini, M.I. Early Cambrian Asfar Sequence. GeoArabia 2010, 15, 137–160. [Google Scholar] [CrossRef]
  263. Powell, J.H.; Abed, A.M.; Le Nindre, Y.-M. Cambrian stratigraphy of Jordan. GeoArabia 2014, 19, 81–134. [Google Scholar] [CrossRef]
  264. Blanco, G.; Rajesh, H.M.; Gaucher, C.; Germs, G.J.B.; Chemale, F., Jr. Provenance of the Arroyo del Soldado Group (Ediacaran to Cambrian, Uruguay): Implications for the paleogeographic evolution of southwestern Gondwana. Precambrian Res. 2009, 171, 57–73. [Google Scholar] [CrossRef]
  265. Geyer, G. The Fish River Subgroup in Namibia: Stratigraphy, depositional environments and the Proterozoic--Cambrian boundary problem revisited. Geol. Mag. 2005, 142, 465–498. [Google Scholar] [CrossRef]
  266. Blanco, G.; Germs, G.J.B.; Rajesh, H.M.; Chemale, F., Jr.; Dussin, I.A.; Justino, D. Provenance and paleogeography of the Nama Group (Ediacaran to early Palaeozoic, Namibia): Petrography, geochemistry and U--Pb detrital zircon geochronology. Precambrian Res. 2011, 187, 15–32. [Google Scholar] [CrossRef]
  267. Selvaraj, K.; Chen, C.T.A. Moderate chemical weathering of subtropical Taiwan: Constraints from solid-phase geochemistry of sediments and sedimentary rocks. J. Geol. 2006, 114, 101–116. [Google Scholar] [CrossRef]
Figure 1. The paleogeographic and present-day location of the Salt Range, Pakistan. (a) A paleogeographic reconstruction of the Indian Plate and the Salt Range during the Early Cambrian period; (b) a structural map of Pakistan highlighting the position of the Salt Range; and (c) a geological map of the Salt Range indicating the locations of the studied stratigraphic sections (yellow stars) [11,15,26].
Figure 1. The paleogeographic and present-day location of the Salt Range, Pakistan. (a) A paleogeographic reconstruction of the Indian Plate and the Salt Range during the Early Cambrian period; (b) a structural map of Pakistan highlighting the position of the Salt Range; and (c) a geological map of the Salt Range indicating the locations of the studied stratigraphic sections (yellow stars) [11,15,26].
Minerals 15 00789 g001
Figure 2. Lithological logs of the Khewra Sandstone from the studied stratigraphic sections in the Salt Range, Pakistan. The logs illustrate vertical variations in lithology and sedimentary structures across the study area.
Figure 2. Lithological logs of the Khewra Sandstone from the studied stratigraphic sections in the Salt Range, Pakistan. The logs illustrate vertical variations in lithology and sedimentary structures across the study area.
Minerals 15 00789 g002
Figure 3. Representative outcrop features of the Khewra Sandstone in the Salt Range. (a) Olive-green and light-maroon shale observed in the lower part of the formation in the KG section; (b) fine-grained, moderately to well-sorted sandstone; (c) catenary ripple marks; (d) load casts from the lower part of the KG section; (e) ball-and-pillow structures in the KG section; (f) Interbedded sandstone and shale in the lower portion of the KG outcrop; (g) trough cross-bedded sandstone (green arrow) and climbing ripples (red arrow); (h) herringbone cross-bedding; (i) pebbly sandstone bed; (j) conglomeratic bed marking the contact between the Khewra Sandstone and the overlying Kussak Formation; and (k) lower and upper contacts of the Khewra Sandstone in the KZ section.
Figure 3. Representative outcrop features of the Khewra Sandstone in the Salt Range. (a) Olive-green and light-maroon shale observed in the lower part of the formation in the KG section; (b) fine-grained, moderately to well-sorted sandstone; (c) catenary ripple marks; (d) load casts from the lower part of the KG section; (e) ball-and-pillow structures in the KG section; (f) Interbedded sandstone and shale in the lower portion of the KG outcrop; (g) trough cross-bedded sandstone (green arrow) and climbing ripples (red arrow); (h) herringbone cross-bedding; (i) pebbly sandstone bed; (j) conglomeratic bed marking the contact between the Khewra Sandstone and the overlying Kussak Formation; and (k) lower and upper contacts of the Khewra Sandstone in the KZ section.
Minerals 15 00789 g003
Figure 4. Photomicrographs of representative sandstone samples from the Khewra Sandstone. (a) A microcline, elongated quartz, and lithic fragment; (b) monocrystalline quartz; (c) point and triple junction grain contacts, polycrystalline quartz, and biotite; (d) moderately sorted detrital grains; (e) quartzite lithic fragment with silica cement; and (f) calcite cement and muscovite grains.
Figure 4. Photomicrographs of representative sandstone samples from the Khewra Sandstone. (a) A microcline, elongated quartz, and lithic fragment; (b) monocrystalline quartz; (c) point and triple junction grain contacts, polycrystalline quartz, and biotite; (d) moderately sorted detrital grains; (e) quartzite lithic fragment with silica cement; and (f) calcite cement and muscovite grains.
Minerals 15 00789 g004
Figure 5. Petrographic data of the Khewra Sandstone plotted using various classification and interpretive diagrams; (a) the QFL ternary plot shows that the sandstones are classified as sub-arkose and sub-lithic arenite; (b) the QFL plot shows the paleoclimate of the Khewra Sandstone as sub-humid, along with the continental block provenance to the formation; (c) the Qp/F + R vs. Qt/F + R bivariate diagram shows a semi-arid to semi-humid paleoclimate [63]; and (d) ln(Q/F) vs. ln(Q/R) plot almost plain relief with low to moderate weathering conditions [41].
Figure 5. Petrographic data of the Khewra Sandstone plotted using various classification and interpretive diagrams; (a) the QFL ternary plot shows that the sandstones are classified as sub-arkose and sub-lithic arenite; (b) the QFL plot shows the paleoclimate of the Khewra Sandstone as sub-humid, along with the continental block provenance to the formation; (c) the Qp/F + R vs. Qt/F + R bivariate diagram shows a semi-arid to semi-humid paleoclimate [63]; and (d) ln(Q/F) vs. ln(Q/R) plot almost plain relief with low to moderate weathering conditions [41].
Minerals 15 00789 g005
Figure 6. X-ray Diffraction (XRD) diffractograms of selected samples from the Khewra Sandstone across the studied sections. (a) KG-7: a sandstone sample from the lower part of the formation in the KG section; (b) KG-34: a shale sample from the middle part of the KG section; (c) KG-47: a sandstone sample from the upper part of the KG section; (d) NW-38: a sandstone sample from the middle part of the NW section; and (e) VS-19: a sandstone sample from the upper part of the vs. section; (f) KZ-37: a sandstone sample from the upper part of the KZ section. (g) A pie chart and (h) statistical bar chart summarizing the overall mineralogical composition of the Khewra Sandstone across the four studied sections.
Figure 6. X-ray Diffraction (XRD) diffractograms of selected samples from the Khewra Sandstone across the studied sections. (a) KG-7: a sandstone sample from the lower part of the formation in the KG section; (b) KG-34: a shale sample from the middle part of the KG section; (c) KG-47: a sandstone sample from the upper part of the KG section; (d) NW-38: a sandstone sample from the middle part of the NW section; and (e) VS-19: a sandstone sample from the upper part of the vs. section; (f) KZ-37: a sandstone sample from the upper part of the KZ section. (g) A pie chart and (h) statistical bar chart summarizing the overall mineralogical composition of the Khewra Sandstone across the four studied sections.
Minerals 15 00789 g006
Figure 7. (a,b) K2O/Na2O vs. CIA(Molar) and Al2O3% vs. CIA(Molar) plots representing deposition under arid paleoclimate [50]; (c) C-values indicating sedimentation and weathering in arid to semi-arid conditions [51]; and (d) SiO2 vs. Al2O3 + Na2O + K2O negating hot and humid paleoclimatic conditions during the deposition of the Khewra Sandstone [40].
Figure 7. (a,b) K2O/Na2O vs. CIA(Molar) and Al2O3% vs. CIA(Molar) plots representing deposition under arid paleoclimate [50]; (c) C-values indicating sedimentation and weathering in arid to semi-arid conditions [51]; and (d) SiO2 vs. Al2O3 + Na2O + K2O negating hot and humid paleoclimatic conditions during the deposition of the Khewra Sandstone [40].
Minerals 15 00789 g007
Figure 8. (a) An A-CN-K (Al2O3-Ca2O + Na2O-K2O) ternary plot [64]; and (b) an A-N-K (Al2O3-Na2O-K2O) ternary plot [65], where both plots show low to moderate chemical weathering.
Figure 8. (a) An A-CN-K (Al2O3-Ca2O + Na2O-K2O) ternary plot [64]; and (b) an A-N-K (Al2O3-Na2O-K2O) ternary plot [65], where both plots show low to moderate chemical weathering.
Minerals 15 00789 g008
Figure 9. (a) Al2O3/Na2O vs. CIA; (b) Al2O3/K2O vs. CIA; both plots show low to moderate weathering for the source rocks of the Khewra Sandstone [50].
Figure 9. (a) Al2O3/Na2O vs. CIA; (b) Al2O3/K2O vs. CIA; both plots show low to moderate weathering for the source rocks of the Khewra Sandstone [50].
Minerals 15 00789 g009
Figure 10. WIP vs. CIA plot showing low to moderate weathering [66,67].
Figure 10. WIP vs. CIA plot showing low to moderate weathering [66,67].
Minerals 15 00789 g010
Figure 11. Geochemical proxies illustrating paleoclimatic, paleo-weathering, and paleo-redox conditions of the Khewra Sandstone. Combined plots of the Chemical Index of Alteration (CIA), Al2O3, K/Na, Sr/Ba, Rb/Sr, V/Cr, and Cu/Zn ratios from the representative Khewra Gorge stratigraphic section suggest an arid to semi-arid paleoclimate, low to moderate intensity of chemical weathering, and predominantly oxic to sub-oxic depositional conditions [47,48,49,56,57,58,59,60,68,69].
Figure 11. Geochemical proxies illustrating paleoclimatic, paleo-weathering, and paleo-redox conditions of the Khewra Sandstone. Combined plots of the Chemical Index of Alteration (CIA), Al2O3, K/Na, Sr/Ba, Rb/Sr, V/Cr, and Cu/Zn ratios from the representative Khewra Gorge stratigraphic section suggest an arid to semi-arid paleoclimate, low to moderate intensity of chemical weathering, and predominantly oxic to sub-oxic depositional conditions [47,48,49,56,57,58,59,60,68,69].
Minerals 15 00789 g011
Figure 12. (ac). A schematic model illustrating the sedimentary depositional environments of the Khewra Sandstone. (a) The paleogeographic position of the Indian Plate during the Cambrian; (b) a generalized geological map of the Salt Range and the NW-Indian Plate margin; and (c) a conceptual depositional model for the Khewra Sandstone.
Figure 12. (ac). A schematic model illustrating the sedimentary depositional environments of the Khewra Sandstone. (a) The paleogeographic position of the Indian Plate during the Cambrian; (b) a generalized geological map of the Salt Range and the NW-Indian Plate margin; and (c) a conceptual depositional model for the Khewra Sandstone.
Minerals 15 00789 g012
Table 1. Geographic locations and sample details of the studied sections of the Khewra Sandstone.
Table 1. Geographic locations and sample details of the studied sections of the Khewra Sandstone.
S/NoSectionLocationThickness (m)Samples
TotalSandstoneShalePetrographyXRDGeochemistry
1Khewra Gorge
(KG)
32°40′5.26″ N
73°0′13.68″ E
14355496331520
2Nilawahan Gorge
(NW)
32°38′13.19″ N
72°35′22.80″ E
785348526916
3Vasnal Village
(VS)
32°42′54.13″ N
72°32′5.49″ E
22191721248
4Khan Zaman Nala
(KZ)
32°32′6.88″ N
71°48′5.02″ E
6750428291216
Table 2. A summary of the lithofacies and lithofacies associations identified within the Khewra Sandstone based on [61,62]. The table outlines key sedimentological characteristics, including the lithology, sedimentary structures, grain size, and interpreted depositional processes and environments for each facies and facies association.
Table 2. A summary of the lithofacies and lithofacies associations identified within the Khewra Sandstone based on [61,62]. The table outlines key sedimentological characteristics, including the lithology, sedimentary structures, grain size, and interpreted depositional processes and environments for each facies and facies association.
CodeLithofaciesKey FeaturesVertical and
Lateral Distribution
Lithofacies
Association
Depositional Settings
CMFShLChannel Margin and Floodplain Shale LithofaciesLaminated, maroon and olive-green shale, low bioturbationLower parts, repeated in middle and upper parts of KG (3 cycles), KZ (4 cycles), NW (1 cycle); absent in VSChannel margin, overbankFluvio–deltaic; channel margin and overbank areas
CMSLChannel Margin Sandstone LithofaciesFine-grained sandstone, ripples, desiccation cracks, cross bedding, soft sedimentary deformationLower parts, some occurrences in middle and upper sections of KG, KZ, and NWChannel margin, distributary channelsFluvio–deltaic with low- to moderate-energy conditions; channel margin, distributary channels, delta front, mouth bars
DLDPLDelta Lobe/Delta Plain LithofaciesInterbedded fine sandstone and shale, cyclic variations in thicknessPresent in all sections with interbedding cyclesDelta lobe, floodplainCyclic deposition indicating fluctuating flow energy; delta plain environment
CBSLChannel Belt Sandstone LithofaciesMedium to coarse-grained sandstone, tabular and tangential cross-beds, planar lamination, herringbone stratificationMiddle parts, tabular cross-beds widespread in KG, NW, and KZ; herringbone cross-beds in KG and NWChannel beltsHigh-energy depositional settings, such as fluvial and delta distributary channels and delta fronts
CBCLChannel Belt Conglomerate LithofaciesClast-supported quartz conglomerateContact with the Kussak Formation, found at the top in KG and vs. onlyHigh-energy river discharge, transition zoneHigh-energy fluvial settings marking transitions to deltaic conditions
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Qanit, A.B.; Iqbal, S.; Kamran, A.H.; Idrees, M.; Sames, B.; Wagreich, M. Sedimentary Facies and Geochemical Signatures of the Khewra Sandstone: Reconstructing Cambrian Paleoclimates and Paleoweathering in the Salt Range, Pakistan. Minerals 2025, 15, 789. https://doi.org/10.3390/min15080789

AMA Style

Qanit AB, Iqbal S, Kamran AH, Idrees M, Sames B, Wagreich M. Sedimentary Facies and Geochemical Signatures of the Khewra Sandstone: Reconstructing Cambrian Paleoclimates and Paleoweathering in the Salt Range, Pakistan. Minerals. 2025; 15(8):789. https://doi.org/10.3390/min15080789

Chicago/Turabian Style

Qanit, Abdul Bari, Shahid Iqbal, Azharul Haq Kamran, Muhammad Idrees, Benjamin Sames, and Michael Wagreich. 2025. "Sedimentary Facies and Geochemical Signatures of the Khewra Sandstone: Reconstructing Cambrian Paleoclimates and Paleoweathering in the Salt Range, Pakistan" Minerals 15, no. 8: 789. https://doi.org/10.3390/min15080789

APA Style

Qanit, A. B., Iqbal, S., Kamran, A. H., Idrees, M., Sames, B., & Wagreich, M. (2025). Sedimentary Facies and Geochemical Signatures of the Khewra Sandstone: Reconstructing Cambrian Paleoclimates and Paleoweathering in the Salt Range, Pakistan. Minerals, 15(8), 789. https://doi.org/10.3390/min15080789

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop