Next Article in Journal
Experimental Compressive Assessment of Different Stiffened Plate Welding Configurations
Next Article in Special Issue
Regional Geological Data on the Volturno Basin Filling and Its Relationship to the Massico Structure (Southern Tyrrhenian Sea, Italy)
Previous Article in Journal
Analysis of Characteristics and Main Controlling Factors of Shallow Geological Hazards in the Zhongsha Islands Region of the South China Sea
Previous Article in Special Issue
Atypical Linear Tectonic Block of the Intraplate Deformation Zone in the Central Indian Ocean Basin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Continental Weathering Triggered the Anoxia of Seawater and Mass Extinctions During the Late Ordovician

1
School of Geosciences, Yangtze University, Wuhan 430100, China
2
Hubei Key Laboratory of Marine Geological Resources, China University of Geosciences, Wuhan 430074, China
3
School of Earth Resources, China University of Geosciences, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2024, 12(12), 2237; https://doi.org/10.3390/jmse12122237
Submission received: 31 October 2024 / Revised: 4 December 2024 / Accepted: 4 December 2024 / Published: 5 December 2024

Abstract

:
During the Late Ordovician period, changes in climate and mass extinctions were observed; however, the factors influencing these phenomena have not been fully understood. In order to understand the relationships among redox water conditions, climates, and mass extinctions in the Late Ordovician, this study analyzes the chemical index of alteration (CIA) in shales and 87Sr/86Sr in carbonate leachates as proxies of changes in chemical weathering intensity and chemical weathering rate in the Late Ordovician (mainly from Katian to Hirnantian). The results show that an enhanced chemical weathering rate (increased 87Sr/86Sr ratios) and decreased chemical weathering intensity (decreased CIA values) characterized the late Katian, which might be attributed to the global orogenesis and enhanced precipitation/runoff under the warming climate (late-Boda warming). This enhanced chemical weathering rate contributed to the CO2 drawdown in the P. pacificus biozone, corresponding to the initiation of cooling and further glaciation. Meanwhile, the enhanced weathering-induced high primary productivity could have contributed to the expansion of anoxic seawater in the Katian, which further caused the Katian extinction. The Hirnantian Glaciation was characterized by high 87Sr/86Sr ratios in carbonates and extremely low CIA values in shales, which were likely related to the exposure of continents during low sea level and the glacial grinding of unweathered rocks. This study shows that the highest denudation rate and lowest chemical weathering intensity in the Hirnantian stage might have resulted in enhanced CO2 release and contributed to the end of glaciation.

1. Introduction

The Late Ordovician witnessed significant geological events involving climate changes, mass extinction, and biological recovery [1,2,3,4]. The former includes the warming of the climate (Boda warming event in late Katian) and the cooling of the climate (glaciation in Hirnantian), and the latter generally includes the Great Ordovician Biodiversification Event (GOBE) and Late Ordovician mass extinctions (LOMEs). Extensive studies have shown that Boda warming and Hirnantian Glaciation were related to changes in atmospheric CO2 and temperature caused by a combination of intense volcanic eruption, increases in continental weathering, and organic carbon burial [5,6,7,8,9]. Two phases of extinction marked the LOMEs in traditional understandings, and new studies have found that the LOMEs happened within the middle–late Katian (named as Katian extinction) [10,11]. A variety of extinction mechanisms have been proposed, mostly involving temperature, volcanism, and euxinia in seawater [12,13,14,15,16,17].
The chemical weathering of silicate rock entails the dissolution of silicate minerals and the formation of secondary minerals [18]. This process transforms atmospheric CO2 into dissolved HCO3 and releases significant amounts of nutrients (such as phosphorus and iron) into the ocean, thereby substantially influencing the paleoclimate, primary productivity, and oceanic hydrochemistry conditions [19,20,21]. Factors such as surface temperature, tectonic activity, volcanic eruptions, and land-plant colonization generally impact the rate of chemical weathering [22,23,24]. It is suggested that the intense tectonic activity during the Late Ordovician not only led to crustal uplift [5] but also facilitated the southward movement of the Appalachian orogenic belt relative to the Laurentian continent [8], both of which enhanced continental weathering. Moreover, the greenhouse climate resulting from significant volcanic activity in the late Katian and a plentiful supply of weathered protoliths further amplified continental weathering [25,26]. In contrast, the global cooling in the Hirnantian was characterized by weak chemical weathering intensity [27].
Silicate weathering mainly includes chemical and physical denudation rates (denudation rate = chemical weathering rate + physical weathering rate), with chemical weathering intensity representing the ratio of the chemical weathering rate to the denudation rate [28]. Generally, a linear array on the logarithmic plot of chemical weathering versus physical erosion rate could be observed under a low erosion rate (transport-limited), in which case chemical weathering intensity is enhanced due to thick soil (completely weathered) sequences [28]. In contrast, silicate weathering shows substantial scatter away from the trend at higher erosion rates (kinetically limited) when chemical weathering rates are higher [28]. Traditionally, weathering-controlled changes in climate mean that an enhanced chemical weathering rate, rather than increased chemical weathering intensity, contributes to the drawdown of atmospheric CO2 [18,19,20]. The relationship between climate and weathering is complicated, and several hypotheses have been proposed. For example, a high chemical weathering intensity triggered CO2 drawdown following a carbon release event caused by volcanism [29], a low chemical weathering intensity and high chemical weathering rate occurred under a warming climate [30,31], and the silicate weathering rate reached a maximum for intermediate denudation rates and further contributed to CO2 drawdown. In contrast, extremely low denudation rates (supply limited regime) and high denudation rates (kinetically limited regime) can decrease or even reverse CO2 sequestration [32].
The strontium (Sr) isotopic compositions in seawater are influenced by continental runoff and mantle contributions from hydrothermal systems [33,34,35,36,37,38]. As a result, seawater 87Sr/86Sr ratios could be used as significant indicators for continental weathering trends in geological periods because these 87Sr/86Sr ratios could be used to constrain continental weathering fluxes of dissolved Sr [33]. Weathering proxies in mudstones include the chemical alteration index (CIA), plagioclase alteration index (PIA), and chemical weathering index (CIW) [39,40,41]. All of these weathering proxies should reflect the changes in chemical weathering intensity. In south China, the redox water conditions, hydrodynamic conditions, volcanism, and evolution of biomass have been widely analyzed in previous studies [3,10,16,29]. Consequently, detailed investigations into the weathering trends (chemical weathering rate and intensity) of the Late Ordovician (using weathering indices in shales and 87Sr/86Sr in carbonate leachates) are essential to elucidate the critical relationships between continental weathering, paleoclimate, and the LOMEs.

2. Geological Setting

South China was located as a separate continent in the Ordovician (Figure 1A) [42,43], consisting of the Yangtze Block in northwest and the Cathaysia Block in the southeast. The collision of them happened in the early Neoproterozoic (Figure 1B), leading to the compression between the Yangtze Block and Cathaysia Block [44,45]. The graptolite in the Katian includes Dicellograptus complanatus, Dicellograptus complexus, Paraothograptus pacificus biozones, Metabolograptus extraodinarius, and Metabolograptus persculptus biozones in the Hirnantian and Akidograptus ascensus, Parakidograptus acuminatus, Cystograptus vesiculosus, and Coronograptus cyphus biozones in the Rhuddanian (Figure 1C) [46,47].
A shallow water section (Wuke section) was located in western Sichuan Province, south China (Figure 1B) [17]. The stratigraphy of the Wuke (WK) section is organized into the Linxiang (LX) Formation, the Tiezufeike (TZFK) Formation, and the Butuo (BT) Formation, in stratigraphic order. The LX Formation (~5 m) is primarily characterized by argillaceous limestones and calcareous shales. The Tiezufeike Formation (~15 m) is characterized predominantly by limestones and dolomitic limestones, with the upper part of Tiezufeike Formation containing a significant presence of shelly fauna, notably the Hirnantian fauna. The Butuo Formation (~>20 m) mainly consists of laminated limestones and calcareous mudstones, alternating with layers of argillaceous siltstones and mudstones.
The two boreholes (YD1 and DY3) analyzed in this study were located on the Upper Yangtze platform (Figure 1B). The boreholes are categorized into the Lingxiang Formation, Wufeng Formation, Guanyinqiao Bed, and Longmaxi Formation, arranged in stratigraphic order. The Wufeng and Longmaxi formations primarily consist of black shale, whereas the Guanyinqiao Bed comprises carbonaceous and muddy limestones containing abundant benthic fauna [16].
Environmental differences were noted between the western margin of the Yangtze Block, characterized by a deep shelf (shale dominated), and the southwestern Yangtze Block, characterized by a shallow shelf (carbonates dominated). In the deep shelf area, a widespread distribution of organic-rich shales of the Wufeng (WF) and Longmaxi (LMX) formations were observed [13,16,17]. The siliceous and argillaceous shales of the Wufeng and Longmaxi formations, dating from the latest Hirnantian to Aeronian, were deposited under anoxic water conditions [13,16,17]. The controlling mechanisms for this included intense upwelling, volcanism, and thermohaline stratification. In the shallow shelf areas, minor carbonate facies were present along the flanks of the Kangdian and Central Guizhou uplifts, where the waters likely became less anoxic and upwelling weakened compared to those in the deep-water shelf [41,43]. Also, the carbonates in the shallow shelf were deposited under more oxic water conditions, leading to less total organic matter in the rocks [17].

3. Methods

In this study, a total of 15 carbonate samples from the Wuke section and 29 shale samples from DY3 and YD1 boreholes were performed for geochemical analyses.

3.1. Major and Trace Elements

The melting method was used for the pretreatment of samples for major element analyses, with the cosolvent being lithium fluoride, lithium metaborate, and lithium tetraborate and the oxidant being ammonium nitrate and lithium bromide. The melting temperature was set to 1050 °C, and the melting time was 15 min. An X-ray fluorescence spectrometer (XRF) was used for the analysis of major elements, with the standard curve using Chinese National Standard Material (GBW07101-14) [48] and a relative standard deviation (RSD) of less than 2% [48]. The analytical precision was better than 10%.
The trace elements were analyzed by an Agilent 7700e ICP-MS at the Wuhan Sample Solution Analytical Technology Co., Ltd., Wuhan, China. First, fresh and fine-grained shale samples were powdered into 0.075 mm-diameter pellets, and these powders were placed in an oven at 105 °C for approximately 12 h. Second, approximately 50 mg of the sample was dissolved in HNO3 and HF. Third, HNO3, Milli-Q water, and an internal standard solution were added. Fourth, the final solution was transferred to a polyethylene bottle and diluted to 100 g by adding 2% HNO3. The analytical precision was better than 5%.
The CIA is calculated as follows: CIA = Al2O3/(Al2O3 + CaO* + Na2O + K2O) × 100 [49]. In the equation, CaO* represents the CaO content in the silicate, and thus, it is calculated following the method: CaO* is initially corrected using P2O5 contents (CaO − 10/3 × P2O5) [50]. If the calculated value exceeds the Na2O content, the CaO* value is set as the Na2O value; otherwise, the CaO* value is set as the CaO content. Due to the possible addition of K2O to weathered samples during K-metasomatism, a correction of CIA values (CIAcorr) has been proposed as follows: CIAcorr = Al2O3/(Al2O3 + CaO* + Na2O + K2O*) × 100 [51], where K2O* is calculated as follows: (K2O*) = {m × [(Al2O3)] + m × [(CaO*) + (Na2O)]}/(1 − m), with m = (K2O)/[(Al2O3) + (CaO*) + (Na2O) + (K2O)].
To eliminate the influence of K2O, Harnois [39] proposed the CIW proxy (CIW = Al2O3/(Al2O3 + CaO* +Na2O) × 100). Fedo et al. [51] defined the PIA proxy as a necessary weathering intensity, which is calculated as follows: PIA = (Al2O3 − K2O)/(Al2O3 − K2O + CaO* + Na2O) × 100. In addition, the compositional maturity of a rock (ICV) is expressed as ICV = (Fe2O3 + K2O + Na2O + CaO + MgO + MnO + TiO2)/A12O3 [52].

3.2. Elemental Concentrations and 87Sr/86Sr Analysis in Carbonates

About 100 mg of carbonate samples were rinsed three times with the Milli-Q water (18.2 MΩ). The carbonate samples were leached twice with 0.05 M acetic acid at room temperature for 24 h. Fractions of the carbonate leachates were retained in 2% HNO3 for elemental analysis using an Elan Quadrupole ICP-OES. The reference material JLs-1was analyzed [53] and indicates that accuracy and precision were less than ±7% for all elemental concentrations reported here.
Sr isotopic ratios were measured using the Neptune plus MC-ICP-MS at the Wuhan Sample Solution Analytical Technology Co., Ltd., Wuhan, China. The Sr isotope analysis of carbonate leachates involves three steps [54]. Sample digestion: about 50–200 mg of sample power (200 mesh) was placed in an oven at 105 °C for drying and was then dissolved using the HNO3 and HF. The solution was evaporated to dryness and was dissolved in 1.0 mL of 2.5 M HCl again. Column chemistry: the supernatant solution was loaded into an ion-exchange column packed with AG50W resin. The Sr fraction was eluted using 2.5 M HCl and gently evaporated to dryness prior to mass-spectrometric measurement. Sr isotope analyses were conducted at the Wuhan Sample Solution Analytical Technology Co., Ltd., Hubei, China, using a Neptune Plus MC-ICP-MS. International NIST 987 standard [54] was measured every seven samples analyzed, with the 87Sr/86Sr ratio of 0.710242 ± 14 (2SD, n = 345) having been achieved, which is identical within error to their published values 0.710248 ± 12 [55]. Additionally, the BCR-2 (basalt) and RGM-2 (rhyolite) yielded results of 0.705012 ± 22 (2SD, n = 63) and 0.704173 ± 20 (2SD, n = 20) for 87Sr/86Sr, respectively.

4. Results

4.1. Major Elements

The result of major element concentrations can be found in the Supplementary Files (Tables S1 and S2). Si, Al, and Ca are the primary elements and show high concentrations in the studied shales, while Fe, Mg, Na, K, Ti, and P are the minor elements and show low concentrations (Tables S1 and S2). Generally, Si, Al, Na, and K are linked with silicate minerals, and Ca and Mg are linked with carbonate minerals. The Si percentage varies from 35.7% to 70.0% in DY3 (except for one marlstone with low SiO2 contents (13.4%) in Guanyinqiao Member) and from 56.0% to 77.7% in YD1, Al varies from 3.2% to 11.5% in DY3 and from 4.7% to 15.7% in YD1, and Ca varies from 1.7% to 26.0% in DY3 and from 0.6% to 15.7% in YD1.
All the CIA, CIAcorr, CIW, and PIA profiles in DY3 and YD1 boreholes exhibit decreasing trends from the Katian to Hirnantian (Figure 2), with values ranging from 58.1 to 66.0, from 59.2 to 67.9, from 69.5 to 81.0, and from 62.1 to 75.4 in DY3 and from 60.4 to 68.3, from 61.1 to 70.5, from 69.3 to 83.5, and from 64.0 to 78.4 in YD1. The ICV shows variable values throughout the Katian to Hirnantian rocks, ranging from 0.98 to 14.4 in DY3 and from 0.90 to 4.43 in YD1.

4.2. Trace Elements

The trace elements in shales of DY3 borehole have been analyzed in this study (Table S1). Sc, Hf, Rb, La, and Th in the Katian show steady contents, varying from 8.40 μg/g to 11.2 μg/g, from 2.51 μg/g to 4.01 μg/g, from 104 μg/g to 148 μg/g, from 31.8 μg/g to 47.6 μg/g, and from 12.5 μg/g to 17.8 μg/g (Table S1). The Sc, Hf, Rb, Th in these rocks show relatively low contents compared to those in PAAS (Sc: 16, Hf: 5, Rb: 160), while the La and Th show more variables when compared with those in PAAS (La: 38, Th:14.6). Sc, Co, Rb, La, Th in the Hirnantian show variable contents, varying from 3.60 μg/g to 9.40 μg/g, from 1.59 μg/g to 4.17 μg/g, from 38.2 μg/g to 107 μg/g, from 15.5 μg/g to 69.2 μg/g, and from 4.85 μg/g to 13.8 μg/g (Table S1). Except for La, all other trace elements show lower contents relative to those in PAAS.

4.3. 87Sr/86Sr and Elements in Carbonate Leachates

The results of 87Sr/86Sr and elements in carbonate leachates can be found in the Supplementary Files (Table S3). 87Sr/86Sr have stratigraphic variations in the WK section, increasing from 0.708492 to 0.709022 in the lower Tiezufeike Formation and decreasing from 0.709022 to 0.708130 in the upper Tiezufeike Formation. In addition, the Sr and Mn in carbonate leachates show variable contents throughout the Tiezufeike Formation, ranging from 301.1 μg/g to 1446.6 μg/g, and ranging from 17.0 μg/g to 183.7 μg/g, respectively. The corresponding Mn/Sr and Sr/Ca ratios vary between 0.03 ppm/% and 0.50 ppm/% and between 29.91 ppm/% and 190.68 ppm/%, respectively.

5. Discussion

5.1. δ13 C Records and Climate Events in the Late Ordovician

Carbon isotopes are identified as a useful tool for determining local and global stratigraphic correlations. We compared the δ13C values in Late Ordovician sediments of the WK section [58] with those in other published studies [59,60,61,62]. In south China, the δ13Corg values are high in the D. complanatus Biozone (Figure 3; Linxiang Formation), and they shift toward low values within the D. complexus Biozones (lower Wufeng Formation in shale-dominated sections and lower Tiezufeike Formation in carbonate-dominated section). The δ13Corg values remained low throughout the lower-middle P. pacificus biozone; subsequently, the δ13Corg values began to increase slowly in the upper P. pacificus biozone and more rapidly in the lower Hirnantian (P. persculptus) zone and peaked in the upper part of the M. extraordinarius biozone. This peak in high δ13Corg (~−25‰) is concurrent with the occurrence of the Hirnantia–Dalmanitina fauna [58], predominantly a cold/cool-water fauna found in the upper Tiezufeike Formation in WK section and Guanyinqiao Member in the DY and YD boreholes (Figure 3).
Various climatic events, including the early and late-Boda warming, mid-Boda cooling, and Hirnantian Glaciation, have been delineated through petrological and geochemical evidence, with δ13C record correlation utilized to constrain the timing of these geological phenomena (Figure 3; [63]). δ13C values remained stable and high during the late Katian (from D. complanatus to D. complexus), aligning with the mid-Boda cooling [61]. Subsequently, δ13C values decrease in the P. pacificus biozone, reflecting intense volcanism and late-Boda warming [64]. An increase in δ13C is noted in the uppermost P. pacificus Biozone, equivalent to the upper Tiezufeike Formation in WK, the upper Wufeng Formation in Tianjiawan and Qiliao, and the Vinini Formation in Vinini Creek (Figure 3). A positive δ13C excursion is observed within the M. persculptus and M. persculptus biozones, alongside a significant regression and the most extensive Hirnantian glacial event. By the onset of the Silurian, δ13C values revert to lower levels similar to those before the Hirnantian Glaciation, marking the end of the glaciation and a transition to a warmer climate (Figure 3).

5.2. The Reliability of the Weathering Proxies in Shales and Carbonates

5.2.1. Evaluation of Weathering Proxies in Shales

The weathering proxies in shales are widely utilized to evaluate silicate weathering [65]. In this study, CIA, CIAcorr, CIW, and PIA show decreasing trends throughout the DY3 and YD1 boreholes (Figure 3; [66,67]) and positive correlations among these proxies could be observed (Figure 4A–C; [68,69]). A thorough evaluation of these proxies is essential because they might be affected by source rocks, hydrodynamic cycles, and post-depositional diagenesis (Figure 4A–F). In this case, this study uses the trace element concentrations and their ratios, Al2O3/SiO2, and K/Rb, to evaluate the influences of the above factors.
Traditionally, Zr, Ti, Al, and their ratios are used to identify the lithology of parental rocks; however, Zr is strongly controlled by the presence of zircons and hence the sedimentary rocks, as zircons tend to be enriched in the silt-size fraction. In addition, Ti abundances in shales generally co-vary with Al, because both elements are immobile during chemical weathering and associated with secondary clay minerals [68]. So, the granulometry of shales is likely the dominant parameter controlling Al2O3/TiO2 ratios in shales. Instead, we choose elements such as Hf, La, Th, and Sc in clastic rocks as effective indicators for determining the types of source rocks and tectonic settings [69]. The discrimination diagrams indicate that all the samples are near the regions of felsic to intermediate rocks (Figure 5A,B; [70,71,72]).
Additionally, A-CN-K diagrams are utilized to determine the types of source rocks. The deviation of our samples from the ideal weathering trend of source rocks (parallel to the A-CN axis) in the Al2O3− (CaO* + Na2O)− K2O (A-CN-K) ternary diagram (Figure 4F). The corrected CIA values (CIAcorr) are distributed along the ideal weathering trends of source rocks, and the intersection between the weathering trend line and feldspar join represents the source rock. In this study, the ternary diagrams show that the samples from two boreholes converge towards a point representing the source rocks (granodiorite or granite; Figure 4F). These comprehensive observations strongly support the conclusion that Late Ordovician shales derived from a consistent protolith source. Therefore, the observed variations in weathering indicators throughout the Katian and Hirnantian periods are unlikely to be attributable to changes in the types of source rocks.
Adding K2O during diagenesis can affect the accuracy of the CIA values because the samples are characterized by less weathering than their actual state. The K/Rb ratio can be used to assess the degree of K mobility due to metasomatism, with the rocks having lower K/Rb ratios (>220) in potassium metasomatism [66]. We found that our samples are characterized by slightly high K/Rb ratios (200–227), indicating minor influences of diagenesis (Figure 3). The CIAcorr values are calculated to avoid the influence of potassium metasomatism. In our study, a substantial positive correlation between CIAcorr and CIA indicates that the correction of the CIA preserves the temporal weathering trend (Figure 4A). Furthermore, the variations in weathering proxies might be related to multiple sedimentary cycles before burial, thus influencing the actual weathering experience of the protolith in source areas. Generally, the ICV is a reliable indicator for distinguishing mature and immature mudstones, with mature mudstones experiencing several weathering cycles and exhibiting low ICVs (<1). In contrast, immature mudstones exhibit high ICVs > 1 [67]. In our study, all the shales have high ICVs > 1, with just three samples showing slightly low values (0.9, 0.92, 0.98), suggesting that they experienced only one cycle of weathering (Figure 4D). The Al/Si ratio is a grain size proxy, with fine particles having high Al2O3/SiO2 ratios and coarse components having low Al/Si ratios. There is moderate correlation between CIAcorr and Al2O3/SiO2 in YD1, which could not be observed in DY3 (Figure 4E). We believe that there is a negligible influence of grain size on the variations in weathering proxies for the DY3. In YD1, the coarse part is characterized by low chemical weathering intensity, likely enhanced erosion of unweathered rocks with high size.

5.2.2. Evaluation of Carbonate 87Sr/86Sr

The Sr isotopic compositions in carbonates are susceptible to early diagenetic transformations and late diagenetic fluid reactions [73]; therefore, evaluating the reliability of carbonate 87Sr/86Sr before using it as the primary seawater record is important [74]. The seawater Sr is incorporated into carbonate minerals in the synsedimentary stage; however, the Sr concentrations in carbonate minerals could be modified during diagenesis [75]. Generally, 300 ppm Sr appears to be the lower threshold value, in which case the bulk carbonate could record the primary 87Sr/86Sr ratios [37]. All of the studied samples have high Sr contents (Figure 6) and there is no obvious relationship between Sr and 87Sr/86Sr (Figure 6), suggesting that primary seawater 87Sr/86Sr is potentially preserved in bulk carbonate.
Mn tends to be incorporated into carbonate minerals when bulk carbonate reacts with nonmarine diagenetic fluids, leading to higher Mn concentrations and radiogenic 87Sr/86Sr ratios in bulk carbonate [74]. As a result, Mn/Sr ratios are often used to discriminate diagenetic alterations, with unaltered bulk carbonates characterized by low Mn/Sr ratios [74]. In the WK section, the bulk carbonates have low ratios (Mn/Sr < 1; Figure 6), conforming to the geochemical criteria for Sr isotopes [74,75]. Moreover, bulk carbonate diagenesis should be characterized by low δ18O values and more radiogenic 87Sr/86Sr ratios than unaltered bulk carbonate, as diagenesis results in high temperatures. However, there are no correlations between Mn/Sr (Figure 7A), Sr/Ca (Figure 7B), or δ18O (Figure 7C) and 87Sr/86Sr, indicating that our samples were not altered by diagenesis.
Dolomitization is a possible process that can alter the geochemical composition (e.g., Mg/Ca, Mn/Sr, and 87Sr/86Sr ratios) of limestones [76]. However, the peak 87Sr/86Sr ratios do not correspond to the high dolomite contents in the WK section (Figure 6), and there is no correlation between Mg/Ca and 87Sr/86Sr (Figure 7D), suggesting that the observed variation in 87Sr/86Sr is not controlled by primary minerals or dolomitization during diagenesis.

5.3. Fluctuations in Weathering Proxies in the Late Ordovician

5.3.1. Fluctuations in CIAcorr Values in the Late Ordovician

Weathering proxies, including the CIA, CIAcorr, PIA, and CIW, have been used to analyze weathering trends throughout geological history. Previous studies have evaluated the limits of the chemical weathering intensity; however, there are some differences in the threshold values among strong, moderate, and weak [13,45,46]. In this study, the degree of chemical weathering intensity could be divided into intense (~75–100), intermediate (~65–75), and weak (~50–65) chemical weathering intensity [45,46]. It is broadly acknowledged that regions experiencing moderate to high chemical weathering intensity are distinguished by elevated CIA and CIAcorr values. In contrast, low CIA and CIAcorr values are indicative of areas with reduced chemical weathering intensity [13,16].
Within the Yangtze Block, CIA and CIAcorr values demonstrate a progressive decline from the Linxiang Formation to the Wufeng Formation [77,78], despite some variance in absolute values across different boreholes and sections (Figure 8). Similarly, in Scotland, CIAcorr and CIA values also exhibit a marked decrease, with the lower part of the Hartfell shales characterized by higher CIA values and the upper part by lower CIA values (Figure 8; [27]). Consequently, the observed weathering trends based on CIA and CIAcorr in the late Katian are deemed reliable, reflecting consistency across global boreholes and sections (Figure 8). This suggests a diminishing trend in chemical weathering intensity during the late Katian period (from D. complanatus to P. pacificus). Of course, the climatic zonality could influence the chemical weathering intensity, with high latitude areas generally showing low temperatures and low chemical weathering intensity and low latitude areas generally showing high temperatures and high chemical weathering intensity. In this case, the climatic zonality is another possibility influencing the differences in the CIA values between the China and Scotland sections, although this requires more exploration in the future.
In south China, CIA and CIAcorr exhibit lower values in the Hirnantian rocks (CIA average values for the DY and YD boreholes: 60 and 63; CIAcorr average values for the DY and YD boreholes: 61 and 64) than in the late Katian rocks (CIA average values for the DY and YD boreholes: 64 and 66; CIAcorr average values for the DY and YD boreholes: 69 and 67) (Table 1; Figure 8). This result is consistent with previous studies, suggesting that the rocks deposited during the Hirnantian Glaciation experienced weak chemical weathering intensity [16]. However, increases in CIA and CIAcorr could be observed in Scotland, in contrast to what has been observed in south China. This study attributes these differences in CIA trends to the different weathering regimes under the same climatic conditions. For example, the continental weathering in south China was characterized by the glacial grinding of unweathered rocks with low CIA values; however, a more incongruent weathering regime likely characterized the Scotland rocks based on the high lithium isotopes [27].

5.3.2. Fluctuations in Seawater 87Sr/86Sr in the Late Ordovician

Several seawater 87Sr/86Sr records from the Late Ordovician have been reported [79]. The overall published datasets from the mid–late Katian have shown relatively stable and lower seawater 87Sr/86Sr values (~0.7080), which are attributed to the enhanced weathering of fresh volcanic rocks [80]. However, all previous studies focus more on the Upper Ordovician carbonate successions, especially from the Tremadocian to Sandbian Stages [80]. In fact, there is still the lack of a detailed and high-resolution seawater 87Sr/86Sr curve from Katian to Hirnantian, which could provide important constraints on the changes in climates.
The overall 87Sr/86Sr ratios in Late Ordovician section from south China are higher than those in the southwestern Ontario borehole [81] and the Copenhagen Canyon section (Figure 9). However, the initial 87Sr/86Sr ratios in the Wuke section are similar to the published conodont apatite 87Sr/86Sr ratios in the late Sandbian and early Katian from the south China sections. It seems that the anomaly of high 87Sr/86Sr ratios in the late Katian might be caused by a restricted ocean, as is observed in the late Ediacaran West Gondwana basins [82] and the latest Ediacaran Sichuan Basins [83]. This condition of the restricted ocean in the Late Ordovician has also been observed by others, who suggest that moderate restricted conditions characterize the Yangtze shelf sea and likely influence the degree of water mass restriction [84,85]. However, we have re-analyzed the major and trace elements in the WK section to identify if our samples were deposited under a restricted basin [58]. The enrichments of Mo and U in sediments have widely been used to reconstruct the redox state of water columns, as well as the changes in open (seawater) versus enclosed (lake/lagoon) conditions [86]. The results show that a transition from suboxic to sulfidic water conditions could be observed from the early to late Katian, which is followed by suboxic water conditions in the Hirnantian (Figure 5C). Therefore, the majority of samples in our studied section were deposited in open marine settings, instead of restricted ocean. In addition, Co and Mn contents in marine sediments were effective proxies for distinguishing the restricted and upwelling settings due to the significant differences in the supply of these metals to the water column [72]. We found that the majority of samples have stable low Co×Mn values (<0.1), indicating open marine settings or upwelling settings (Figure 5D). As a result, we would like to highlight that the changes in carbonate 87Sr/86Sr ratios in the Wuke section still record the global trend of seawater 87Sr/86Sr curve. However, the reasons for the overall high values will be considered in future works.
Several similarities are found in the Late Ordovician 87Sr/86Sr curves between south China and other areas. First, both the 87Sr/86Sr curves in south China and southwestern Ontario show low Sr isotopes which is consistent with the overall ratios globally, likely resulting from enhanced continental basalt weathering, a sustainable increase in oceanic hydrothermal Sr flux, or a combination of both [87]. Second, the 87Sr/86Sr ratios become much more positive in the late Katian, reaching high Sr isotope ratios not attained until the Silurian found in previous studies [80]. Third, the prolonged high 87Sr/86Sr ratios in the Hirnantian have been observed in south China, southwestern Ontario, and the Copenhagen Canyon (Figure 9). These radiogenic 87Sr/86Sr ratios in Glaciation were caused by the enhanced mechanical erosion driven by sea-level fall or the preferential weathering of biotite in cooling climates [79]. Therefore, we still believe that our Sr isotope record has global-scale implications for the changes in silicate weathering, hydrothermal activity, and the weathering of fresh volcanic rocks.
The seawater 87Sr/86Sr is controlled by both river influx from continents and hydrothermal influx [36]. A major drop in marine 87Sr/86Sr records across the middle–late Ordovician boundary (Darriwilian-Sandbian) was attributed to enhanced continental basalt weathering [88], subduction-related volcanism, or a sustainable increase in oceanic hydrothermal Sr flux [87]. The initial low 87Sr/86Sr ratios in the lower most Tiezufeike Formation are consistent with the hypotheses about basalt weathering and hydrothermal input; therefore, the influence of these factors should be taken into account. First, A-CN-K diagrams are utilized to determine the types of source rocks, with the results showing that all the samples are sourced from intermediate-felsic igneous rocks, as a result, the effect of changes in the types of source rocks on the 87Sr/86Sr ratios is minor. Second, the contribution of hydrothermal activity to the deposition of Late Ordovician shales in south China is mainly based on the positive Eu anomaly (Eu/Eu*), because a reduction of Eu3+ to Eu2+ would occur under the extreme reduction in hydrothermal fluids [89,90]. The results show that weak hydrothermal activity (occasionally slightly positive Eu anomaly) exists on the northern and southwestern margins of the Upper Yangtze Platform, situated in an extensional tectonic background [85]; thus, the possibility of decreasing hydrothermal Sr flux could be excluded. In this case, we just retain the spreading rate or the hydrothermal Sr-input as a steady condition when considering that the spreading rate or the hydrothermal Sr-input is unknown at present. Anyway, we would like conduct more modeling work about the influences of spreading rate or the global hydrothermal Sr-input on the Late Ordovician Sr isotopes in seawater. Third, sea-level-fall-induced restricted seawater conditions are unlikely to lead to the higher 87Sr/86Sr ratios because the geochemical characteristics in bulk rocks above have identified open ocean settings. Overall, the increase in 87Sr/86Sr ratios has been attributed to the increasing continental crust erosion during the Taconic orogeny or the sea-level drop related to the onset of Hirnantian Glaciation.

5.4. The Changes in Continental Weathering in the Late Ordovician

Traditional studies have argued that continental weathering is controlled by plant coverage, temperature, runoff, and tectonic processes [91]. First, given that the first non-vascular land plants were only just evolving and colonizing the continents in the mid–late Ordovician, it is probable that the weathering pattern and corresponding clay types were different and less abundant [92]. For example, illites are thought to dominate prior to terrestrialization by plants [92]. In fact, the clay minerals of the Late Ordovician rocks in both south China and Scotland were dominated by illite [27,89]. In addition, the occurrence of non-vascular land plants would be suggested to induce the enhanced formation of clay minerals caused by a decreased denudation rate and increased chemical weathering intensity, leading to the preservation of cations in clays and decreased consumption of atmosphere CO2 [27]. However, both the changed types of clay and increased chemical weathering intensity were inconsistent with the observation in our geochemical records. As a result, the effect of missing plant coverage on the records of the weathering trends in this study might be minor.
Generally, the elevated temperature could enhance the feldspar dissolution, weathering rate, chemical weathering intensity (high CIA values), and further CO2 consumption [4]. Hydrologic regulation (precipitation and runoff) has also widely been proposed to control the silicate weathering, with increased runoff causing a high weathering rate in low-relief areas [91]. Some studies further suggested that the runoff-induced high denudation rate would contribute to the high chemical weathering and low chemical weathering intensity [31]. Tectonic uplift can enhance the chemical weathering rate by accelerating the rate of denudation [22]. In this case, high erosion and a rapid supply of fresh minerals would result in a short time to reach weathering equilibrium, so the weathering rate, solute concentrations, and fluxes are high [91].
The weathering flux to the oceans should be maximized when thermodynamic equilibrium between the dissolving and precipitating minerals is approached. Furthermore, peak weathering and CO2 drawdown occur with moderate erosion rates. In contrast, low and high erosion rates would decrease or reverse CO2 sequestration. In a simplified model, the low denudation rate would result in a supply limited regime (with thick soils isolating the bedrock from climatic conditions), while the high denudation rate would cause the kinetically limited regime (with rocks being eroded before undergoing significant hydrolysis) [32].
A certain proportion of sedimentary rocks in paleo-catchments will necessarily exaggerate the numerical estimate of the CIA. Therefore, the proposed weathering trends using CIA values of Late Ordovician shales need be checked in the future. Anyway, the declining CIA values from the D. complanatus to P. pacificus biozones might suggest diminishing chemical weathering intensity (Figure 8). However, oxygen isotope evidence indicates that a mid-Boda cooling climate happened within the D. complanatus and D. complexus biozones and a late-Boda warming climate happened in the P. pacificus biozone (Figure 10) [93]. Moreover, the CIA trend is contrary to the observation that increasingly radiogenic 87Sr/86Sr occur within the P. pacificus biozone and indicate an augmented weathering flux (Figure 10). Therefore, this study believes that the decrease in CIA values and chemical weathering intensity in late Katian should not just be controlled by temperature regulation. Instead, tectonic and hydrologic regulations likely account for the changes in continental weathering. During the mid-Boda cooling interval, the chemical weathering and denudation rates were much lower due to the low temperatures and precipitation/runoff. In this interval, the low denudation rate would result in more clay formation and thick soils isolating the bedrock from weathering, which is the supply limited regime (high W/D and chemical weathering intensity; Figure 11).
A warming climate and orogenesis likely contributed to reducing chemical weathering intensity within the P. pacificus biozone. The warming climate is caused by global volcanism and the injection of CO2 into the atmosphere (Figure 10; [94]), with the high temperatures in this interval leading to an increase in runoff of 6.8 to 2.3%/°C between high- and low-latitude rivers, respectively [95]. In this case, the rapid erosion of rocks and an exceptionally increased denudation rate would occur due to increased precipitation/runoff, as extensively reported in Early Silurian warming climates [17]. The orogeny in the late Katian has been reported globally, such as the Kwangsian Orogeny in south China, the Caledonian Orogen in Scotland, or the Taconic orogeny in the Appalachian margin of Laurentia, ensuring the continual exposure of fresh material [96,97,98], which tends to weather more rapidly. The intermediate denudation rate and enough fresh rocks would result in moderate erosion and intermediate weathering intensity (Figure 11).
Glaciation occurred within the M. extraordinarius and M. persculptus zones [93]. The weathering trends in Hirnantian Glaciation are subject to debate, with several perspectives reported. First, some studies propose that low temperatures and glacial ice cover could lower the chemical weathering rate, leading to a low chemical weathering intensity (low W/D ratio) [16]. Second, others suggest that glaciation enhances the denudation rate while reducing the chemical weathering intensity (low W/D ratio) [99]. However, the Yangtze Platform was located in subtropical regions during the Ordovician and hence was not subject to glacial weathering. Third, a cooling climate prolongs the continental residence time of water, allowing more clay formation, thereby reducing global weathering flux and promoting more incongruent weathering (intermediate W/D ratio) [27]. Given the high 87Sr/86Sr ratios and extremely low CIA values in Hirnantian Glaciation, this study suggests that the Hirnantian Glaciation was characterized by increased chemical weathering and denudation rates but low chemical weathering intensity (low W/D ratio). The amplified clastic fluxes from continents and increased continental material transported into the ocean are corroborated by lowεNd(t) values and high TiO2/Al2O3, Zr/Al2O3 ratios [100]. Increased erosion during low sea level in glaciation, the glacial grinding of unweathered rocks, or a combination of them, caused the kinetically limited regime, which contributed to the increased denudation rate. Furthermore, the high seawater 87Sr/86Sr ratios and low sulfate–sulfur isotopes could support the increased chemical weathering. The high 87Sr/86Sr ratios might be driven by glacial grinding and abrasion that produced fine-grained Rb-rich glacial till with minerals such as biotite [101], as well as the preferential weathering of biotite with radiogenic 87Sr/86Sr in glaciation [101]. A large negative sulfur excursion (δ34S in carbonate-associated sulfate) occurred throughout the Hirnantian Glaciation, and increasing continental weathering (especially sulfides) was required based on the geochemical box modeling [102]. Overall, the increased denudation rate exceeded the increased chemical weathering rate, leading to the lowest chemical weathering intensity in glaciation (Figure 11).

5.5. Implications for the Changes in Global Climates and LOMEs

The Late Ordovician and Early Silurian transition is characterized by an expansion of anoxia in the deeper oceans which resulted in the deposition of organic-rich shales which later became oil- and gas-source rocks [16,41,103]. Both the orbital forcing and paleoenvironmental changes have been explained for the deposition of those source rocks [9,104]. In the late Katian, a warming event (Boda event) was restricted to the upper part of the Ka4 time slice [105] or the late Katian P. pacificus biozone [63]. Previous findings imply that volcanism-induced paleoenvironmental repercussions in late Katian had long-term effects on the global system rather than short-term impacts on paleoenvironments and paleoclimates [106]. Therefore, volcanism likely contributed to the increase in pCO2, negative δ13C excursion, and the late-Boda warming event [26]. In the P. pacificus interval, higher continental runoff is expected to have occurred, due to warming and a humid climate caused by intense volcanism. Furthermore, high runoff accelerates chemical weathering and denudation. The enhanced chemical weathering rate increased the consumption of atmospheric CO2 and the nutrient influx into the ocean, leading to high primary productivity in surface waters [43,45]. The enhanced chemical weathering rate and weathering-induced high primary productivity contributed to the enrichment of organic matter, lower pCO2, and lower global temperatures. Therefore, this study suggests that the weathering rate reaches maximum values for intermediate denudation rates and weathering intensity (Figure 11), which further contributes to CO2 drawdown and cooling. Therefore, this study suggests that if intense volcanism can push the climate system out of balance (Boda warming), silicate weathering controlled by temperature and runoff would help to allow climatic recovery, even triggering glaciation [107]. We found that high chemical weathering and denudation rates, and the lowest chemical weathering intensity characterize the Hirnantian Glaciation. The high weathering rate seems to violate the fact that low temperatures and glacial cover characterized the Hirnantian. However, an increased chemical weathering rate in glaciations has also been proposed, with the factors being attributed to the enhanced erosion of silicate, and weathering of sulfides and carbonates [108]. In fact, the enhanced carbonate and sulfide weathering in the Hirnantian is supported by the high carbon isotopes and low sulfate-sulfur isotopes [102]. In this case, an increase in sulfide and carbonate weathering leads to a possible source of CO2 followed by a longer sink of CO2. This hypothesis is consistent with the observation of CO2 release with the highest weathering and denudation rates (Figure 11). Overall, the highest denudation rate and lowest chemical weathering intensity in Hirnantian should result in the CO2 release and termination of glaciation.
Traditional views suggested that the LOMEs were marked by two phases in the Late Ordovician (Figure 10). This study considered that the first phase corresponded to the start of glaciation and was within the boundary between Paraothograptus pacificus and Metabolograptus extraodinarius graptolite zones, and the second phase corresponded to the termination of glaciation and happened in the latest Hirnantian and earliest Rhuddanian periods (Figure 10; [4]). However, new studies found that the Late Ordovician extinctions happened within the middle–late Katian (named the Katian extinction) and also suggested a rapid global expansion of OMZs throughout the late Katian, based on trace elemental concentrations and Tl and U isotopes (Figure 10; [102]). This study suggests that weathering-induced high primary productivity and seawater anoxia likely played significant roles in controlling the LOMEs, given the coupling of high weathering rates and expanded anoxic seawater in the late Katian (Figure 10 and Figure 12). Previous studies suggested that oxygen level rather than temperature was the main factor controlling the Late Ordovician extinctions for as long as 35 million years [109]. The Late Ordovician crisis includes three rapid declines in species richness, with the first decline occurring in the late Katian, the second decline occurring in the earliest Hirnantian, and the third decline occurring in late Hirnantian (Figure 10; [110]). It is believed that weathering-induced seawater anoxia could explain the crisis in the late Katian, while the development of more intense euxinia and volcanism throughout the water column likely drove the second and third extinction in the Hirnantian (Figure 10; [14]).

6. Conclusions

The weathering trends during the Late Ordovician (late Katian-Hirnantian) have been evaluated by the chemical index of alteration (CIA) in shales and 87Sr/86Sr in carbonate leachates. The intervals within the D. complanatus and D. complexus biozones were marked by a low weathering rate (low 87Sr/86Sr ratios), and by high chemical weathering intensity (high CIA values). Conversely, the intervals within the P. pacificus biozone showed a high weathering rate (high 87Sr/86Sr ratios) and intermediate chemical weathering intensity (intermediate CIA values). This increase in weathering rate and reduction in chemical weathering intensity should be attributed to global orogenesis and enhanced precipitation/runoff under the warming climate. The Hirnantian Glaciation was characterized by high 87Sr/86Sr ratios and extremely low CIA values, which might be related to the exposure of continents during low sea level and the glacial grinding of unweathered rocks, with high 87Sr/86Sr ratios and low CIA values.
The changes in weathering had significant influences on the climatic changes and mass extinctions in the Late Ordovician. The intermediate weathering rate and intensity contributed to the CO2 drawdown in the P. pacificus biozone, corresponding to the initiation of cooling and further glaciation. Meanwhile, the enhanced weathering-induced high primary productivity resulted in the expansion of anoxic seawater in the Katian, which further caused Katian extinction. In the Hirnantian Glaciation, the highest denudation rate and lowest chemical weathering intensity likely resulted in enhanced CO2 release and contributed to the termination of glaciation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jmse12122237/s1, Table S1: Major and trace element concentrations in the DY3 borehore; Table S2: Major elements concentrations in the YD1 borehore; Table S3: Elemental concentrations and 87Sr/86Sr analysis in the Wuke section.

Author Contributions

Methodology, D.Y.; Formal analysis, X.Y.; Writing—original draft, P.T. All authors have read and agreed to the published version of the manuscript.

Funding

The project was supported by the Open Funds for Hubei Key Laboratory of Marine Geological Resources, China University of Geosciences, No. MGR202407, the National Natural Science Foundation of China (42402123, 41690131), and Beijing Nova Program (Z211100002121136).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Melchin, M.J.; Mitchell, C.E.; Holmden, C.; Štorch, P. Environmental changes in the Late Ordovician-early Silurian: Review and new insights from black shales and nitrogen isotopes. Geol. Soc. Am. Bull. 2013, 125, 1635–1670. [Google Scholar] [CrossRef]
  2. Harper, D.A.T.; Hammarlund, E.U.; Rasmussen, C.M.Ø. End Ordovician extinctions: A coincidence of causes. Gondwana Res. 2014, 25, 1294–1307. [Google Scholar] [CrossRef]
  3. Jin, C.S.; Liao, Z.W.; Lash, G.G. High-frequency redox variation across the Ordovician–Silurian transition, South China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2021, 566, 110218. [Google Scholar] [CrossRef]
  4. Deng, K.; Yang, S.Y.; Guo, Y.L. A global temperature control of silicate weathering intensity. Nat. Commun. 2022, 13, 1781. [Google Scholar] [CrossRef] [PubMed]
  5. Kump, L.R.; Arthur, M.A.; Patzkowsky, M.E.; Gibbs, M.T.; Pinkus, D.S.; Sheehan, P.M. A weathering hypothesis for glaciation at high atmospheric pCO2 during the Late Ordovician Interpreting carbon—Isotope excursions: Carbonates and organic matter. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1999, 16152, 18173–19887. [Google Scholar]
  6. Finlay, A.J.; Selby, D.; Gröcke, D.R. Tracking the Hirnantian glaciation using Os isotopes. Earth Planet. Sci. Lett. 2010, 293, 339–348. [Google Scholar] [CrossRef]
  7. Lenton, T.; Michael, C.; Martin, J.; Pires, N.; Dolan, L. First plants cooled the Ordovician. Nat. Geosci. 2012, 5, 86–89. [Google Scholar] [CrossRef]
  8. Swanson-Hysell, N.L.; Macdonald, F.A. Tropical weathering of the Taconic orogeny as a driver for Ordovician cooling. Geology 2017, 45, 719–722. [Google Scholar] [CrossRef]
  9. Sproson, A.D.; von Strandmann, P.A.E.; Selby, D.; Jarochowska, E.; Frýda, J.; Hladil, J.; Loydell, D.K.; Slavík, L.; Calner, M.; Maier, G.; et al. Osmium and lithium isotope evidence for weathering feedbacks linked to orbitally paced organic carbon burial and Silurian glaciations. Earth Planet. Sci. Lett. 2022, 577, 117260. [Google Scholar] [CrossRef]
  10. Deng, Y.; Fan, J.X.; Zhang, S.H.; Fang, X.; Chen, Z.Y.; Shi, T.K.; Wang, H.W.; Wang, X.B.; Yang, J.; Hou, X.D. Timing and patterns of the Great Ordovician Biodiversification Event and Late Ordovician mass extinction: Perspectives from South China. Earth Sci. Rev. 2021, 220, 103743. [Google Scholar] [CrossRef]
  11. Harper, D.A.T. Late Ordovician Mass Extinction: Earth, fire and ice. Natl. Sci. Rev. 2024, 11, nwad319. [Google Scholar] [CrossRef] [PubMed]
  12. Sheehan, P.M. The Late Ordovician mass extinction. Annu. Rev. Earth Planet. Sci. 2001, 29, 331–364. [Google Scholar] [CrossRef]
  13. Yan, D.T.; Chen, D.Z.; Wang, Q.C.; Wang, J.G. Large-scale climatic fluctuations in the latest Ordovician on the Yangtze block, South China. Geology 2010, 38, 599–602. [Google Scholar] [CrossRef]
  14. Hammarlund, E.U.; Dahl, T.W.; Harper, D.A.T.; Bond, D.P.G.; Nielsen, A.T.; Bjerrum, C.J.; Schovsbo, N.H.; Schönlaub, H.P.; Zalasiewicz, J.A.; Canfield, D.E. A sulfidic driver for the end-Ordovician mass extinction. Earth Planet. Sci. Lett. 2012, 331, 128–139. [Google Scholar] [CrossRef]
  15. Jones, D.; Martini, A.; Fike, D. A volcanic trigger for the Late Ordovician mass extinction? Mercury data from south China and Laurentia. Geology 2017, 45, 631–634. [Google Scholar] [CrossRef]
  16. Zou, C.N.; Qiu, Z.; Poulton, S.W.; Dong, D.Z.; Wang, H.Y.; Chen, D.Z.; Lu, B.; Shi, Z.S.; Tao, H.F. Ocean euxinia and climate change “double whammy” drove the Late Ordovician mass extinction. Geology 2018, 46, 535–538. [Google Scholar] [CrossRef]
  17. Liu, M.; Chen, D.Z.; Jiang, L.; Stockey, R.G.; Aseal, D.; Zhang, B.; Liu, K.; Yang, X.R.; Yan, D.T.; Planavsky, N.J. Oceanic anoxia and extinction in the latest Ordovician. Earth Planet. Sci. Lett. 2022, 588, 117553. [Google Scholar] [CrossRef]
  18. Wang, P.; Du, Y.S.; Yu, W.C.; Algeo, T.J.; Zhou, Q.; Xu, Y.; Qi, L.; Yuan, L.J.; Pan, W. The chemical index of alteration (CIA) as a proxy for climate change during glacial-interglacial transitions in Earth history. Earth Sci. Rev. 2020, 201, 103032. [Google Scholar] [CrossRef]
  19. Walker, J.C.G.; Hays, P.B.; Kasting, J.F. A negative feedback mechanism for the long-term stabilization of earth’s surface temperature. Geophys. Res. Lett. 1981, 86, 9776–9782. [Google Scholar] [CrossRef]
  20. Berner, R.A. Weathering, plants, and the long-term carbon cycle. Geochim. Et Cosmochim. Acta 1992, 56, 3225–3231. [Google Scholar] [CrossRef]
  21. Föllmi, K. The phosphorus cycle, phosphogenesis and marine phosphate-rich deposits. Earth Sci. Rev. 1996, 40, 55–124. [Google Scholar] [CrossRef]
  22. Raymo, M.E.; Ruddiman, W.F. Tectonic forcing of late Cenozoic climate. Nature 1992, 359, 117–122. [Google Scholar] [CrossRef]
  23. Caves, J.K.; Ibarra, D.E.; Von Blanckenburg, F. Neogene cooling driven by land surface reactivity rather than increased weathering fluxes. Nature 2019, 571, 99–102. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.Y.; Liang, K.; Xiao, Y.L.; Chen, B.; Shan, E.H.; Yang, T.Y.; Zhang, M.; Sun, H.; Gu, H.O.; Tong, F.T.; et al. Carbonate lithium isotope systematics indicate cooling triggered mass extinction during the Frasnian-Famennian transition. Glob. Planet. Chang. 2023, 230, 104284. [Google Scholar] [CrossRef]
  25. Buggisch, W.; Joachimski, M.M.; Lehnert, O.; Bergstrom, M.; Repetski, E.; Webers, F. Did intense volcanism trigger the first Late Ordovician icehouse? Geology 2010, 38, 327–330. [Google Scholar] [CrossRef]
  26. Lefebvre, V.; Servais, T.; Francois, L.; Averbuch, O. Did a Katian large igneous province trigger the Late Ordovician glaciation? A hypothesis tested with a carbon cycle model. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2010, 296, 310–319. [Google Scholar] [CrossRef]
  27. Pogge Von Strandmann, P.A.E.; Desrochers, A.; Murphy, M.J.; Finlay, A.J.; Selby, D.; Lenton, T.M. Global climate stabilisation by chemical weathering during the Hirnantian glaciation. Geochem. Perspect 2017, 3, 230–237. [Google Scholar] [CrossRef]
  28. West, A.J.; Galy, A.; Bickle, M. Tectonic and climatic controls on silicate weathering. Earth Planet. Sci. Lett. 2005, 235, 211–228. [Google Scholar] [CrossRef]
  29. Shen, J.; Yin, R.S.; Zhang, S.; Algeo, T.J.; Bottjer, D.J.; Yu, J.X.; Xu, G.Z.; Penman, D.; Wang, Y.D.; Li, L.Q.; et al. Intensified continental chemical weathering and carbon-cycle perturbations linked to volcanism during the Triassic–Jurassic transition. Nat. Commun. 2022, 13, 299. [Google Scholar] [CrossRef]
  30. Yang, J.H.; Cawood, P.A.; Condon, D.J.; Liu, J.Z.; Deng, X.S.; Wang, J.F.; Du, Y.S.; Yuan, D.X. Anomalous weathering trends indicate accelerated erosion of tropical basaltic landscapes during the Permo-Triassic warming. Earth Planet. Sci. Lett. 2022, 577, 117256. [Google Scholar] [CrossRef]
  31. Pogge von Strandmann, P.A.E.; Jones, M.T.; West, A.J.; Murphy, M.J.; Stokke, E.W.; Tarbuck, G.; Wilson, D.J.; Pearce, C.R.; Schmidt, D.N. Lithium isotope evidence for enhanced weathering and erosion during the Paleocene-Eocene Thermal Maximum. Sci. Adv. 2021, 7, eabh4224. [Google Scholar] [CrossRef] [PubMed]
  32. Bufe, A.; Rugenstein, J.K.C.; Hovius, N. CO2 drawdown from weathering is maximized at moderate erosion rates. Science 2024, 383, 1075–1080. [Google Scholar] [CrossRef] [PubMed]
  33. Veizer, J. Strontium isotopes in seawater through time. Annu. Rev. Earth Planet. Sci. 1989, 17, 141–167. [Google Scholar] [CrossRef]
  34. Dennison, R.E.; Koepnick, R.B.; Burke, W.H.; Hetherington, E.A. Construction of the Cambrian and Ordovician seawater 87Sr/86Sr curve. Chem. Geol. 1998, 152, 325–340. [Google Scholar] [CrossRef]
  35. McArthur, J.M.; Howarth, R.J.; Bailey, T.R. Strontium isotope stratigraphy: LOWESS version 3: Best fit to the marine isotope curve for 0–509 Ma and accompanying look-up table for deriving numerical age. J. Geol. 2001, 109, 155–170. [Google Scholar] [CrossRef]
  36. Saltzman, M.R.; Edwards, C.T.; Leslie, S.A.; Dwyer, G.S.; Bauer, J.A.; Repetski, J.; Harris, A.; Bergström, S.M. Calibration of a conodont apatite-based Ordovician 87Sr/86Sr curve to biostratigraphy and geochronology: Implications for stratigraphic resolution. Geol. Soc. Am. Bull. 2014, 126, 1551–1568. [Google Scholar] [CrossRef]
  37. Edwards, C.T.; Saltzma, M.R.; Leslie, S.A.; Bergstrm, S.M.; Sedlacek, A.R.C.; Howard, A.; Bauer, J.A.; Sweet, W.C.; Young, S.A. Strontium isotope (87Sr/86Sr) stratigraphy of Ordovician bulk carbonate: Implications for preservation of primary seawater values. Geol. Soc. Am. Bull. 2015, 127, 1275–1289. [Google Scholar] [CrossRef]
  38. Wang, W.Q.; Garbelli, C.; Zheng, Q.F.; Chen, J.; Liu, X.C.; Wang, W.; Shen, S.Z. Permian 87Sr/86Sr chemostratigraphy from carbonate sequences in South China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2018, 500, 84–94. [Google Scholar] [CrossRef]
  39. Harnois, L. The CIW index: A new chemical index of weathering. Sediment. Geol. 1988, 55, 319–322. [Google Scholar] [CrossRef]
  40. Price, J.R.; Velbel, M.A. Chemical weathering indices applied to weathering profiles developed on heterogeneous felsic metamorphic parent rocks. Chem. Geol. 2003, 202, 397–416. [Google Scholar] [CrossRef]
  41. Chen, W.Z.; Tian, J.C.; Lin, X.B.; Liang, Q.S.; Wang, X.; Yi, D.X.; Li, Y.Y. Climate fluctuations during the Ordovician-Silurian transition period in South China: Implications for paleoenvironmental evolution and organic matter enrichment. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2023, 613, 111411. [Google Scholar] [CrossRef]
  42. Zhang, G.; Guo, A.; Wang, Y.J.; Li, S.D.; Dong, Y.P.; Liu, S.F.; He, D.F.; Cheng, S.Y.; Lu, R.K.; Yao, A.P. Tectonics of South China continent and its implications. Sci. China Earth Sci. 2013, 56, 1804–1828. [Google Scholar] [CrossRef]
  43. Lu, Y.B.; Hao, F.; Yan, D.T.; Lu, Y.C. Volcanism-induced late Boda warming in the Late Ordovician: Evidence from the Upper Yangtze Platform, South China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2021, 578, 110579. [Google Scholar] [CrossRef]
  44. Chen, X. Influencece of the Late Ordovician glaciation on the basin configuration of the Yangtze Platform in China. Lethaia 1984, 17, 51–59. [Google Scholar]
  45. Chen, C.; Mu, C.; Zhou, K.K.; Liang, W.; Ge, X.Y.; Wang, X.P.; Wang, Q.Y.; Zheng, B.S. The geochemical characteristics and factors controlling the organic matter accumulation of the Late Ordovician-Early Silurian black shale in the Upper Yangtze Basin, South China. Mar. Pet. Geol. 2016, 76, 159–175. [Google Scholar] [CrossRef]
  46. Chen, X.; Rong, J.Y.; Fan, J.X.; Zhan, R.B.; Mitchell, C.E.; Harper, D.A.T.; Melchin, M.J.; Peng, P.A.; Finney, S.C.; Wang, X.F. The Global boundary Stratotype Section and Point (GSSP) for the base of the Hirnantian Stage (the uppermost of the Ordovician System). Epis. J. Int. Geosci. 2006, 29, 183–196. [Google Scholar] [CrossRef]
  47. Gradstein, F.M.; Ogg, J.G.; Schmitz, M.; Ogg, G. (Eds.) The Geologic Time Scale 2012; Elsevier: Amsterdam, The Netherlands, 2012; Volume 2. [Google Scholar]
  48. Ge, X.; Chen, D.; Zhang, G.; Huang, T.; Liu, M.; El-Shafeiy, M. Marine redox evolution and organic accumulation in an intrashelf basin, NE Sichuan Basin during the Late Permian. Mar. Pet. Geol. 2022, 140, 105633. [Google Scholar] [CrossRef]
  49. Nesbitt, H.W.; Young, G.M. Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 1982, 299, 715–717. [Google Scholar] [CrossRef]
  50. McLennan, S.M. Weathering and global denudation. J. Geol. 1993, 101, 295–303. [Google Scholar] [CrossRef]
  51. Fedo, C.M.; Nesbitt, H.W.; Young, G.M. Unraveling the effects of potassium metasomatism in sedimentary rocks and paleosols, with implications for paleoweathering conditions and provenance. Geology 1995, 23, 921–924. [Google Scholar] [CrossRef]
  52. Cox, R.; Low, D.R.; Cullers, R.L. The influence of sediment recycling and basement composition on evolution of mudrock chemistry in the southwestern United States. Geochim. Cosmochim. Acta 1995, 59, 2919–2940. [Google Scholar] [CrossRef]
  53. Yang, X.; Yan, D.; Chen, D.; Liu, M.; She, X.; Zhang, J.; Wei, X.; Lu, Z. Spatial variation of carbon isotopic compositions of carbonate and organic matter from the Late Ordovician sedimentary succession in the Yangtze Platform, South China: Implications for sea—Level eustacy and shoaling of marine chemocline. J. Asian Earth Sci. 2020, 202, 104540. [Google Scholar] [CrossRef]
  54. Xue, S.; Wang, Q.; Deng, J.; Wang, Y.; Peng, T. Mechanism of organic matter assimilation and its role in sulfde saturation of oxidized magmatic ore-forming system: Insights from C-S-Sr-Nd isotopes of the Tulaergen deposit in NW China. Miner. Depos. 2022, 57, 1123–1141. [Google Scholar] [CrossRef]
  55. Zhang, W.; Hu, Z.C. Estimation of isotopic reference values for pure materials and geological reference materials. At. Spectrosc. 2020, 41, 93–102. [Google Scholar] [CrossRef]
  56. Van de Kamp, P.C. Potassium distribution and metasomatism in pelites and schists: How and When, relation tpost-depositionalal events. J. Sediment Res. 2016, 86, 683–711. [Google Scholar] [CrossRef]
  57. Cullers, R.L.; Podkovyrov, V.N. Geochemistry of the Mesoproterozoic Lakhanda shales in southeastern Yakutia, Russia: Implications for mineralogical and provenance control and recycling. Precambrian Res. 2020, 104, 77–93. [Google Scholar] [CrossRef]
  58. Yang, X.R.; Yan, D.T.; Chen, D.Z.; Liu, M.; She, X.H.; Zhang, B.; Zhang, L.W.; Zhang, J.F. Spatiotemporal variations of sedimentary carbon and nitrogen isotopic compositions in the Yangtze Shelf Sea across the Ordovician-Silurian boundary. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2021, 567, 110257. [Google Scholar] [CrossRef]
  59. Chen, C. Research on Paleoceanography, Paleoclimate and Formation Mechanism of Source Rock During Geologic Transition Period from Late Ordovician to Early Silurian in Southern Sichuan Province—Northern Guizhou Province, South China; China University of Geosciences: Wuhan, China, 2018. [Google Scholar]
  60. Ahm, A.S.; Bjerrum, C.J.; Hammarlund, E.U. Disentangling the record of diagenesis, local redox conditions, and global seawater chemistry during the latest Ordovician glaciation. Earth Planet. Sci. Lett. 2017, 459, 145–156. [Google Scholar] [CrossRef]
  61. Myrow, P.M.; Fike, D.A.; Malmskog, E.; Leslie, S.A.; Zhang, T.R.; Singh, B.P.; Chaubey, R.S.; Prasad, S.K. Ordovician–Silurian boundary strata of the Indian Himalaya: Record of the latest Ordovician Boda event. Geol. Soc. Am. Bull. 2019, 131, 881–898. [Google Scholar] [CrossRef]
  62. Finnegan, S.; Bergmann, K.; Eiler, J.M.; Jones, D.S.; Fike, D.A.; Eisenman, I.; Hughes, N.C.; Tripati, A.K.; Fischer, W.W. The magnitude and duration of Late Ordovician-Early Silurian glaciation. Science 2011, 331, 903–906. [Google Scholar] [CrossRef]
  63. Armstrong, H.A.; Baldini, J.U.L.; Challands, T.J.; Grocke, D.R.; Owen, A.W. Response of the Intertropical Convergence Zone to Southern Hemisphere cooling during Upper Ordovician glaciation. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2009, 284, 227–236. [Google Scholar] [CrossRef]
  64. Wang, Y.; Tan, J.Q.; Wang, W.H. The influence of Late Ordovician volcanism on the marine environment based on high-resolution mercury data from South China. Geol. Soc. Am. Bull. 2023, 135, 787–798. [Google Scholar] [CrossRef]
  65. Basu, H.; Dandele, P.S.; Kumar, K.R.; Achar, K.K.; Umamaheswar, K. Geochemistry of black shales from the Mesoproterozoic Srisailam Formation, Cuddapah basin, India: Implications for provenance, palaeoweathering, tectonics, and timing of Columbia breakup. Chem. Erde. 2017, 77, 596–613. [Google Scholar] [CrossRef]
  66. Bayon, G.; Bindeman, I.N.N.; Trinquier, A.; Retallack, G.J.; Bekker, A. Long-term evolution of terrestrial weathering and its link to Earth’s oxygenation. Earth Planet. Sci. Lett. 2022, 584, 117490. [Google Scholar] [CrossRef]
  67. Perri, F.; Cavalcante, F.; Martín-Martín, M.; Sánchez-Navas, A.; Alcalá, F.J. Cenozoic tectono-sedimentary evolution of the external rif chain (Morocco) derived from mineralogical and geochemical analysis of mudrocks. Mar. Pet. Geol. 2024, 170, 107124. [Google Scholar] [CrossRef]
  68. Bayon, G.; Toucanne, S.; Skonieczny, C.; Andre, L.; Bermell, S.; Cheron, S.; Dennielou, B.; Etoubleau, J.; Freslon, N.; Gauchery, T.; et al. Rare earth elements and neodymium isotopes in world river sediments revisited. Geochim. Cosmochim. Acta 2015, 170, 17–38. [Google Scholar] [CrossRef]
  69. Bhatia, M.R.; Crook, K.A. Trace element characteristics of graywackes and tectonic setting discrimination of sedimentary basins. Contrib. Mineral. Petrol 1986, 92, 181–193. [Google Scholar] [CrossRef]
  70. Algeo, T.J.; Tribovillard, N. Environmental analysis of paleoceanographic systems based on molybdenum–uranium covariation. Chem. Geol. 2009, 268, 211–225. [Google Scholar] [CrossRef]
  71. Tribovillard, N.; Algeo, T.J.; Baudin, F.; Riboulleau, A. Analysis of marine environmental conditions based on molybdenum–uranium covariation—Applications to Mesozoic paleoceanography. Chem. Geol. 2012, 324–325, 46–58. [Google Scholar] [CrossRef]
  72. Sweere, T.; van den Boorn, S.; Dickson, A.J.; Reichart, G.-J. Definition of new trace metal proxies for the controls on organic matter enrichment in marine sediments based on Mn, Co, Mo and Cd concentrations. Chem. Geol. 2016, 441, 235–245. [Google Scholar] [CrossRef]
  73. Sawaki, Y.; Ohno, T.; Tahata, M.; Komiya, T.; Hirata, T.; Maruyama, S.; Windley, B.F.; Han, J.; Shu, D.G.; Li, Y. The Ediacaran radiogenic Sr isotope excursion in the Doushantuo Formation in the Three Gorges area, South China. Precambrian Res. 2010, 176, 46–64. [Google Scholar] [CrossRef]
  74. Jacobsen, S.B.; Kaufman, A.J. The Sr, C and O isotopic evolution of Neoproterozoic seawater. Chem. Geol. 1999, 161, 37–57. [Google Scholar] [CrossRef]
  75. Banner, J.L.; Hanson, G.N. Calculation of simultaneous isotopic and trace element variations during water-rock interaction with applications to carbonate diagenesis. Geochim. Cosmochim. Acta 1990, 54, 3123–3137. [Google Scholar] [CrossRef]
  76. Liu, C.; Wang, Z.; Raub, T.D.; Macdonald, F.A.; Evans, D.A.D. Neoproterozoic cap dolostone deposition in stratified glacial meltwater plume. Earth Planet. Sci. Lett. 2014, 404, 22–32. [Google Scholar] [CrossRef]
  77. He, L.; Wang, Y.P.; Chen, D.F. Geochemical features of sedimentary environment and paleoclimate during Late Ordovician to Early Silurian in southern Sichuan Basin. Geochemica 2019, 48, 555–566. [Google Scholar]
  78. Zhang, H. Analysis of Paleoclimate and Provenance Characteristics of Black Shale in Wufeng-Longmaxi Formation in Xianfeng Area, Western Hubei Province. Coal Geol. Chin. 2022, 34, 22–33, (In Chinese with English Abstract). [Google Scholar]
  79. Hu, D.P.; Zhang, X.L.; Zhou, L.; Finney, S.C.; Liu, Y.S.; Shen, D.; Shen, M.G.; Huang, W.; Shen, Y.N. 87Sr/86Sr evidence from the epeiric Martin Ridge Basin for enhanced carbonate weathering during the Hirnantian. Sci. Rep. 2017, 7, 11348. [Google Scholar] [CrossRef]
  80. Young, S.A.; Saltzman, M.R.; Foland, K.A.; Linder, J.S.; Kump, L.R. A major drop in seawater 87Sr/86Sr during the Middle Ordovician (Darriwilian): Links to volcanism and climate. Geology 2009, 37, 951–954. [Google Scholar] [CrossRef]
  81. Hannigan, R.; Brookfield, M.E.; Basu, A.R. A detailed 87Sr/86Sr isotopic curve for the mid-Cincinnatian (Upper Katian-Lower Hirnantian, Upper Ordovician), NE North American Shelf (Ontario, Canada) transition to the Hirnantian glaciation. Chem. Geol. 2010, 277, 336–344. [Google Scholar] [CrossRef]
  82. Guacaneme, C.; Babinski, M.; Bedoya-Rueda, C.; Paula-Santos, G.M.; Caetano-Filho, S.; Kuchenbecker, M.; Reis, H.L.S.; Trindade, R.I.F. Tectonically-induced strontium isotope changes in ancient restricted seas: The case of the Ediacaran-Cambrian Bambuí foreland basin system, east Brazil. Gondwana Res. 2021, 93, 275–290. [Google Scholar] [CrossRef]
  83. Zhang, X.; Zhou, G.; Zhang, P.; He, Y.; Wei, Z.; Wang, G.; Zhang, T.; He, W.; Ma, H.; Zhu, C.; et al. Strontium isotope and element constraints on the paleoenvironment of the latest Ediacaran in the Sichuan Basin, southeastern Tibetan Plateau. Front. Earth Sci. 2022, 10, 865709. [Google Scholar] [CrossRef]
  84. Lu, Y.B.; Jiang, S.; Lu, Y.C.; Xu, S.; Shu, Y.; Wang, Y.X. Productivity or preservation? The factors controlling the organic matter accumulation in the late Katian through Hirnantian Wufeng organic-rich shale, South China. Mar. Pet. Geol. 2019, 109, 22–35. [Google Scholar] [CrossRef]
  85. Qiu, Z.; Li, Y.F.; Xiong, W.; Fan, T.L.; Zhao, Q.; Zhang, Q.; Wang, Y.M.; Liu, W.; Liang, F.; Zhang, J.Q.; et al. Revisiting paleoenvironmental changes on the Upper Yangtze Blockduring the Ordovician-Silurian transition: New insights from elemental geochemistry. Sediment. Geol. 2023, 450, 106377. [Google Scholar] [CrossRef]
  86. Abdi, Z.; Rimmer, S.M.; Rowe, H.D.; Nordeng, S. Assessment of bi-metal paleo-redox proxies in the Bakken Formation black shales, Williston Basin, North Dakota, USA. Chem. Geol. 2024, 670, 122439. [Google Scholar] [CrossRef]
  87. Avila, T.D.; Saltzman, M.R.; Adiatma, Y.D.; Joachimski, M.M.; Griffith, E.M.; Olesik, J.W. Role of seafloor production versus continental basalt weathering in Middle to Late Ordovician seawater 87Sr/86Sr and climate. Earth Planet. Sci. Lett. 2022, 593, 117641. [Google Scholar] [CrossRef]
  88. Nardin, E.; Godderis, Y.; Donnadieu, Y.; Le Hir, G.; Blakey, R.C.; Puceat, E.; Aretz, M. Modeling the early Paleozoic long-term climatic trend. Geol. Soc. Am. Bull. 2011, 123, 1181–1192. [Google Scholar] [CrossRef]
  89. Ge, X.Y.; Mou, C.L.; Wang, C.S.; Men, X.; Chen, C.; Hou, Q. Mineralogical and geochemical characteristics of K-bentonites from the Late Ordovician to the Early Silurian in South China and their geological significance. Geol. J. 2019, 54, 514–528. [Google Scholar] [CrossRef]
  90. Olivarez, A.M.; Owen, R.M. The europium anomaly of seawater—Implications for fluvial versus hydrothermal REE inputs to the oceans. Chem. Geol. 1991, 92, 317–328. [Google Scholar]
  91. Maher, K.; Chamberlain, C.P. Hydrologic regulation of chemical weathering and the geologic carbon cycle. Science 2014, 343, 1502–1504. [Google Scholar] [CrossRef]
  92. Hazen, R.M.; Sverjensky, D.A.; Azzolini, D.; Bish, D.L.; Elmore, S.C.; Hinnov, L.; Milliken, R.E. Clay mineral evolution. Am. Mineral. 2013, 98, 2007–2029. [Google Scholar] [CrossRef]
  93. Trotter, J.A.; Williams, I.S.; Barnes, C.R.; Männik, P.; Simpson, A. New conodont δ18O records of Silurian climate change: Implications for environmental and biological events. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2016, 443, 34–48. [Google Scholar] [CrossRef]
  94. Hu, D.P.; Li, M.H.; Chen, J.B.; Luo, Q.Y.; Grasby, S.E.; Zhang, T.G.; Yuan, S.L.; Xu, Y.L.; Finney, S.C.; Sun, L.L.; et al. Major volcanic eruptions linked to the Late Ordovician mass extinction: Evidence from mercury enrichment and Hg isotopes. Glob. Planet. Chang. 2021, 196, 103374. [Google Scholar] [CrossRef]
  95. Chai, Y.F.; Berghuijs, W.R.; Naudts, K.; Janssen, T.A.J.; Yao, Y.; Dolman, H. Using precipitation sensitivity to temperature to adjust projected global runoff. Environ. Res. Lett. 2021, 16, 124032. [Google Scholar] [CrossRef]
  96. Van Staal, C.; Barr, S. Lithospheric architecture and tectonic evolution of the Canadian Appalachians and associated Atlantic margin. In Tectonic Styles in Canada: The Lithoprobe Perspective; Percival, J.A., Cook, F.A., Clowes, R.M., Eds.; Special Paper; Geological Association of Canada: St. John’s, NL, Canada, 2012; Volume 49, pp. 41–96. [Google Scholar]
  97. Xu, Y.; Cawood, P.; Du, Y. Intraplate orogenesis in response to Gondwana assembly: Kwangsian orogeny, South China. Am. J. Sci. 2016, 316, 329–362. [Google Scholar] [CrossRef]
  98. Longman, J.; Mills, B.J.W.; Manners, H.R.; Gernon, T.M.; Palmer, M.R. Late Ordovician climate change and extinctions driven by elevated volcanic nutrient supply. Nat. Geosci. 2021, 14, 924–929. [Google Scholar] [CrossRef]
  99. Schachtman, N.S.; Roering, J.J.; Marshall, J.A.; Gavin, D.; Granger, D.E. The interplay between physical and chemical erosion over glacial-interglacial cycles. Geology 2019, 47, 613–616. [Google Scholar] [CrossRef]
  100. Holmden, C.; Mitchell, C.E.; Laporte, D.F. Nd isotope records of late Ordovician sea-level change—Implications for glaciation frequency and global stratigraphic correlation. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2013, 386, 131–144. [Google Scholar] [CrossRef]
  101. Blum, J.D.; Erel, Y. A silicate weathering mechanism linking increases in marine 87Sr/86Sr with global glaciation. Nature 1995, 373, 415–418. [Google Scholar] [CrossRef]
  102. Kozik, N.P.; Gill, B.C.; Owens, J.D.; Lyons, T.W.; Young, S.A. Geochemical records reveal protracted and differential marine redox change associated with Late Ordovician climate and mass extinctions. AGU Adv. 2022, 3, e2021AV000563. [Google Scholar] [CrossRef]
  103. Farouk, S.; Lofty, N.M.; Qteishat, A.; Ahmad, F.; Shehata, A.M.; Al-Kahtany, K.; Hsu, C.S. Source and thermal maturity assessment of the Paleozoic-Mesozoic organic matter in the Risha gas field, Jordan. Fuel 2023, 335, 126998. [Google Scholar] [CrossRef]
  104. Farouk, S.; Abdeldaim, A.; Qteishat, A.; Ahmad, F.; Mohammed, I.; Al-Kahtany, K.; Powell, J.H. Orbital forcing and paleoenvironmental changes across the upper Ordovician glaciation-lower Silurian hot shale in the Risha gas field, northeast Jordan. Mar. Pet. Geol. 2024, 168, 107040. [Google Scholar] [CrossRef]
  105. Bergström, S.M.; Chen, X.; Gutiérrez-Marco, J.C.; Dronov, A. The new chronostratigraphic classifi cation of the Ordovician System and its relations to major regional series and stages and to δ13C chemostratigraphy. Lethaia 2009, 42, 97–107. [Google Scholar] [CrossRef]
  106. Shen, J.; Chen, J.B.; Algeo, T.J.; Feng, Q.L.; Yu, J.X.; Xu, Y.G.; Xu, G.Z.; Lei, Y.; Planavsky, N.J.; Xie, S.C. Mercury fluxes record regional volcanism in the South China craton prior to the end-Permian mass extinction. Geology 2021, 49, 452–456. [Google Scholar] [CrossRef]
  107. Conwell, C.T.; Saltzman, M.R.; Edwards, C.T.; Griffith, E.M.; Adiatma, Y.D. Nd isotopic evidence for enhanced mafic weathering leading to Ordovician cooling. Geology 2022, 50, 886–890. [Google Scholar] [CrossRef]
  108. Torres, M.A.; Moosdorf, N.; Hartmann, J.; Adkins, J.F.; West, A.J. Glacial weathering, sulfide oxidation, and global carbon cycle feedbacks. Proc. Natl. Acad. Sci. USA 2019, 114, 8716–8721. [Google Scholar] [CrossRef]
  109. Rasmussen, C.M.; Kröger, B.; Nielsen, M.L.; Colmenar, J. Cascading trend of Early Paleozoic marine radiations paused by Late Ordovician extinctions. Proc. Natl. Acad. Sci. USA 2019, 116, 7207–7213. [Google Scholar] [CrossRef]
  110. Rong, J.Y.; Harper, D.A.T.; Huang, B.; Li, R.Y.; Zhang, X.L.; Chen, D. The latest Ordovician Hirnantian brachiopod faunas: New global insights. Earth Sci. Rev 2020, 208, 103280. [Google Scholar] [CrossRef]
Figure 1. (A) Global paleogeography(~440 Ma, modified after http://deeptimemaps.com, accessed on 1 January 2023) during the Late Ordovician and Early Silurian period. 1: Dob’s Linn section, Scotland; 2: Vinini Creek section, Nevada; 3: Monitor Range, Nevada; 4: Holy Cross Mountains; 5: Parahio Valley India Himalaya section, India. (B) Early Silurian paleogeographic map showing the distribution of the lithofacies of the Yangtze area [43] showing the sites Wuke (WK), Shuanghe (SH), Tianjiawan (TJW), Mouchuangou (MCG), Qiliao (QL) sections and DY3, XY-1, YD1, Yihuang-1 (YH-1), Shenci-1 (SC-1) boreholes. (C) Time scale and graptolite biozones are from [46,47]; generic diversity across Late Ordovician to Early Silurian is from [10].
Figure 1. (A) Global paleogeography(~440 Ma, modified after http://deeptimemaps.com, accessed on 1 January 2023) during the Late Ordovician and Early Silurian period. 1: Dob’s Linn section, Scotland; 2: Vinini Creek section, Nevada; 3: Monitor Range, Nevada; 4: Holy Cross Mountains; 5: Parahio Valley India Himalaya section, India. (B) Early Silurian paleogeographic map showing the distribution of the lithofacies of the Yangtze area [43] showing the sites Wuke (WK), Shuanghe (SH), Tianjiawan (TJW), Mouchuangou (MCG), Qiliao (QL) sections and DY3, XY-1, YD1, Yihuang-1 (YH-1), Shenci-1 (SC-1) boreholes. (C) Time scale and graptolite biozones are from [46,47]; generic diversity across Late Ordovician to Early Silurian is from [10].
Jmse 12 02237 g001
Figure 2. Stratigraphic CIA, CIAcorr, CIW, PIA, K/Rb, ICV, and Al/Si of the Wufeng Formation and Guanyinqiao Member in DY3 and YD1 boreholes. The orange area represents the period of late-Boda warming climate, the black dotted lines represents 210 for K/Rb ratio and 1 for ICV [56,57], with high K/Rb ratios indicating minor potassium metasomatism and low ICV values indicating mature mudstones. Sil.: Silurian, Rhu.: Rhuddanian, LMX: Longmaxi Formation.
Figure 2. Stratigraphic CIA, CIAcorr, CIW, PIA, K/Rb, ICV, and Al/Si of the Wufeng Formation and Guanyinqiao Member in DY3 and YD1 boreholes. The orange area represents the period of late-Boda warming climate, the black dotted lines represents 210 for K/Rb ratio and 1 for ICV [56,57], with high K/Rb ratios indicating minor potassium metasomatism and low ICV values indicating mature mudstones. Sil.: Silurian, Rhu.: Rhuddanian, LMX: Longmaxi Formation.
Jmse 12 02237 g002
Figure 3. δ13C profiles of the Late Ordovician–Early Silurian strata in the Wuke, TJW, QL, Vinini Creek, and Parahio Valley India Himalaya [59,60,61]. The sea surface temperature is based on the clumped oxygen isotope data [62]. The pink and orange areas represent the periods of early and late-Boda warming climates, respectively. Sil.: Silurian, Rhu.: Rhuddanian, GYQ: Guanyinqiao Formation.
Figure 3. δ13C profiles of the Late Ordovician–Early Silurian strata in the Wuke, TJW, QL, Vinini Creek, and Parahio Valley India Himalaya [59,60,61]. The sea surface temperature is based on the clumped oxygen isotope data [62]. The pink and orange areas represent the periods of early and late-Boda warming climates, respectively. Sil.: Silurian, Rhu.: Rhuddanian, GYQ: Guanyinqiao Formation.
Jmse 12 02237 g003
Figure 4. Crossplots of CIAcorr versus (A) CIA, (B) PIA, and (C) CIW. (D) Plots of analyzed samples of DY3 and YD1 on the A-CN-K diagram [49,50,51]. A: Al2O3, CN: CaO* + Na2O, K: K2O, To: tonalite, Gd: granodiorite, Gr: granite, Kln: kaolinite, Gbs: gibbsite, Chl: chlorite, Ilt: Illite, Ms: muscovite, Kfs: K-feldspar. The arrows represent the weathering trends for YD1 and DY3 rocks. Crossplots of CIAcorr versus (E) ICV, (F) Al/Si. The black dotted line in (D) represents 1 for ICV [67], with low ICV values indicating mature mudstones.
Figure 4. Crossplots of CIAcorr versus (A) CIA, (B) PIA, and (C) CIW. (D) Plots of analyzed samples of DY3 and YD1 on the A-CN-K diagram [49,50,51]. A: Al2O3, CN: CaO* + Na2O, K: K2O, To: tonalite, Gd: granodiorite, Gr: granite, Kln: kaolinite, Gbs: gibbsite, Chl: chlorite, Ilt: Illite, Ms: muscovite, Kfs: K-feldspar. The arrows represent the weathering trends for YD1 and DY3 rocks. Crossplots of CIAcorr versus (E) ICV, (F) Al/Si. The black dotted line in (D) represents 1 for ICV [67], with low ICV values indicating mature mudstones.
Jmse 12 02237 g004
Figure 5. Crossplots of (A) Hf versus La/Th, (B) Sc/Th versus La/Sc, and (C) MoEF versus UEF; the SW means the ratio of MoEF/UEF in modern seawater, the solid lines reflect the changes in redox water conditions, restricted and upwelling settings [70,71]. (D) Mn × Co versus Al2O3; the dotted line represents the boundary value (0.4) between restricted and upwelling settings [72]. All the major and trace elements used above are analyzed for whole rocks from [58].
Figure 5. Crossplots of (A) Hf versus La/Th, (B) Sc/Th versus La/Sc, and (C) MoEF versus UEF; the SW means the ratio of MoEF/UEF in modern seawater, the solid lines reflect the changes in redox water conditions, restricted and upwelling settings [70,71]. (D) Mn × Co versus Al2O3; the dotted line represents the boundary value (0.4) between restricted and upwelling settings [72]. All the major and trace elements used above are analyzed for whole rocks from [58].
Jmse 12 02237 g005
Figure 6. Stratigraphic δ13Corg, δ13Ccarb, 87Sr/86Sr, Mn/Sr, Sr/Ca, and mineralogy of the Linxiang and Tiezufeike formations in WK section. The sea surface temperature is based on the clumped oxygen isotope data [62]. The pink and orange areas represent the periods of early and late-Boda warming climates, respectively. Sil.: Silurian, Rhu.: Rhuddanian.
Figure 6. Stratigraphic δ13Corg, δ13Ccarb, 87Sr/86Sr, Mn/Sr, Sr/Ca, and mineralogy of the Linxiang and Tiezufeike formations in WK section. The sea surface temperature is based on the clumped oxygen isotope data [62]. The pink and orange areas represent the periods of early and late-Boda warming climates, respectively. Sil.: Silurian, Rhu.: Rhuddanian.
Jmse 12 02237 g006
Figure 7. Crossplots of 87Sr/86Sr versus (A) Mn/Sr, (B) Sr/Ca, (C) δ18O, and (D) Mg/Ca.
Figure 7. Crossplots of 87Sr/86Sr versus (A) Mn/Sr, (B) Sr/Ca, (C) δ18O, and (D) Mg/Ca.
Jmse 12 02237 g007
Figure 8. Correlation of CIA and CIAcorr in the SD1, DY3 (this study), SH, QL, XY-1, YD1(this study), and Dob’s Linn [16,59]. The orange area represents the period of late-Boda warming climates. Sil.: Silurian, Rhu.: Rhuddanian, LMX: Longmaxi.
Figure 8. Correlation of CIA and CIAcorr in the SD1, DY3 (this study), SH, QL, XY-1, YD1(this study), and Dob’s Linn [16,59]. The orange area represents the period of late-Boda warming climates. Sil.: Silurian, Rhu.: Rhuddanian, LMX: Longmaxi.
Jmse 12 02237 g008
Figure 9. Stratigraphic 87Sr/86Sr of the Late Ordovician sediments in WK (this study), Swift Current core, Copenhagen Canyon sections [79,81]. The orange area represents the period of late-Boda warming climates. Sil.: Silurian, Rhu.: Rhuddanian, LMX: Longmaxi.
Figure 9. Stratigraphic 87Sr/86Sr of the Late Ordovician sediments in WK (this study), Swift Current core, Copenhagen Canyon sections [79,81]. The orange area represents the period of late-Boda warming climates. Sil.: Silurian, Rhu.: Rhuddanian, LMX: Longmaxi.
Jmse 12 02237 g009
Figure 10. Trends in CIAcorr, 87Sr/88Sr (this study), and δ34S in carbonate-associated sulfate (CAS). Generic diversity, Katian, and Hirnantian extinctions are based on [10]. Global anoxia event is based on [17]. The gray brand represents the CIA values for UCC (upper continental crust). Grap.: Graptolite Biozone, D. compla.-D.comple.: Dicellograptus complanatus-Dicellograptus complexus, M.e.-M. p.: Metabolograptus extraodinarius-Metabolograptus persculptus.
Figure 10. Trends in CIAcorr, 87Sr/88Sr (this study), and δ34S in carbonate-associated sulfate (CAS). Generic diversity, Katian, and Hirnantian extinctions are based on [10]. Global anoxia event is based on [17]. The gray brand represents the CIA values for UCC (upper continental crust). Grap.: Graptolite Biozone, D. compla.-D.comple.: Dicellograptus complanatus-Dicellograptus complexus, M.e.-M. p.: Metabolograptus extraodinarius-Metabolograptus persculptus.
Jmse 12 02237 g010
Figure 11. Relationship between denudation rate and chemical weathering rate [28,32]. The gray line shows a linear relation under mineral-supply limited regimes and a nonlinear relation in kinetically limited regimes. Peak weathering rate and CO2 drawdown occurs with moderate erosion rate; in contrast, low and high erosion rates would decrease or even reverses CO2 sequestration. The red square represents the early Kaitan, with low denudation rate and high chemical weathering intensity. The orange square represents the late Kaitan, with intermediate denudation rate and intermediate chemical weathering intensity. The blue square represents the Hirnantian with high denudation rate and low chemical weathering intensity.
Figure 11. Relationship between denudation rate and chemical weathering rate [28,32]. The gray line shows a linear relation under mineral-supply limited regimes and a nonlinear relation in kinetically limited regimes. Peak weathering rate and CO2 drawdown occurs with moderate erosion rate; in contrast, low and high erosion rates would decrease or even reverses CO2 sequestration. The red square represents the early Kaitan, with low denudation rate and high chemical weathering intensity. The orange square represents the late Kaitan, with intermediate denudation rate and intermediate chemical weathering intensity. The blue square represents the Hirnantian with high denudation rate and low chemical weathering intensity.
Jmse 12 02237 g011
Figure 12. Scheme illustrating redox dynamics during the Ordovician and Silurian transition; the data are from [14,16,59].
Figure 12. Scheme illustrating redox dynamics during the Ordovician and Silurian transition; the data are from [14,16,59].
Jmse 12 02237 g012
Table 1. The average values of CIA and CIAcorr in the rocks of SD1, DY3, Shuanghe, Qiliao, XY-1, YD1 in south China.
Table 1. The average values of CIA and CIAcorr in the rocks of SD1, DY3, Shuanghe, Qiliao, XY-1, YD1 in south China.
Borehole/SectionSD1DY3ShuangheQiliaoXY-1YD1
CIA or CIAcorr Value
Average CIA value in Hirnantian6360/595963
Average CIAcorr value in Hirnantian696159/6064
Average CIA value in late-Boda warming6664/666266
Average CIAcorr value late-Boda warming726965/6467
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

Tang, P.; Yang, X.; Yan, D. Enhanced Continental Weathering Triggered the Anoxia of Seawater and Mass Extinctions During the Late Ordovician. J. Mar. Sci. Eng. 2024, 12, 2237. https://doi.org/10.3390/jmse12122237

AMA Style

Tang P, Yang X, Yan D. Enhanced Continental Weathering Triggered the Anoxia of Seawater and Mass Extinctions During the Late Ordovician. Journal of Marine Science and Engineering. 2024; 12(12):2237. https://doi.org/10.3390/jmse12122237

Chicago/Turabian Style

Tang, Pan, Xiangrong Yang, and Detian Yan. 2024. "Enhanced Continental Weathering Triggered the Anoxia of Seawater and Mass Extinctions During the Late Ordovician" Journal of Marine Science and Engineering 12, no. 12: 2237. https://doi.org/10.3390/jmse12122237

APA Style

Tang, P., Yang, X., & Yan, D. (2024). Enhanced Continental Weathering Triggered the Anoxia of Seawater and Mass Extinctions During the Late Ordovician. Journal of Marine Science and Engineering, 12(12), 2237. https://doi.org/10.3390/jmse12122237

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