Geochemical characteristics of the late Eocene mudstones in the eastern Nima Basin, Xizang: implications for the paleoenvironment, provenance, and tectonic setting

Xinhang WANG, Siqi XIAO, Zhongpeng HAN, Yalin LI, Shuai LI, Jie DAI

Front. Earth Sci. ››

PDF(5061 KB)
Front. Earth Sci. All Journals
PDF(5061 KB)
Front. Earth Sci. ›› DOI: 10.1007/s11707-024-1123-0
RESEARCH ARTICLE

Geochemical characteristics of the late Eocene mudstones in the eastern Nima Basin, Xizang: implications for the paleoenvironment, provenance, and tectonic setting

Author information +
History +

Abstract

Paleoenvironmental reconstruction plays a pivotal role in providing insights into the uplift history of the Xizang Plateau during the Cenozoic. The Nima Basin, situated in the central Xizang Plateau, is crucial for studying the tectonic and geomorphic evolution of this region. The clastic composition and geochemical characteristics of the Niubao Formation hold considerable potential for unravelling the geological history and reconstructing depositional environments of central Xizang in the early Cenozoic. In this study, we present detailed geochemical characteristics to determine their provenance, paleoenvironmental conditions, and tectonic origins. The index of compositional variability (ICV > 1) of mudstones indicates that low compositional maturity sediments underwent weak sedimentary recycling. The chemical index of alteration (CIA: 59.8−72.9) reveals that parental rocks experienced a moderate chemical weathering degree. The paleoclimate indicators of the mudstones suggest an oxidizing and arid depositional environment, with a mean annual temperature (MAT) of 11.64°C ± 4.19°C. The geochemical evidence also demonstrates that the mudstones were derived from mixed felsic and intermediate igneous rocks that formed in a dominantly continental island arc tectonic setting. Similarities in the geochemical characteristics among the Niubao Formation and surrounding igneous rocks indicate that a continental-scale drainage system once drained westward in central Xizang. It is concluded that the central plateau experienced a cooler and drier climate coinciding with the presence of a large-scale drainage system during the late Eocene.

Graphical abstract

Keywords

late Eocene / geochemistry / Nima Basin / mudstone / paleoenvironment

Cite this article

Download citation ▾
Xinhang WANG, Siqi XIAO, Zhongpeng HAN, Yalin LI, Shuai LI, Jie DAI. Geochemical characteristics of the late Eocene mudstones in the eastern Nima Basin, Xizang: implications for the paleoenvironment, provenance, and tectonic setting. Front. Earth Sci., https://doi.org/10.1007/s11707-024-1123-0

1 Introduction

The early Cenozoic continental collision between the India and Asia accelerated the outward growth of the Xizang Plateau from central Xizang and significantly influenced the Cenozoic global climate and the development of the Asian monsoon system (Raymo and Ruddiman, 1992; Molnar et al., 1993, 2010; Cerling et al., 1997; Dupont-Nivet et al., 2007; Ding et al., 2014; Deng and Ding, 2015; Kapp and DeCelles, 2019; Ingalls et al., 2020; Yang et al., 2022; Feng et al., 2023). Nonetheless, the comprehension of the Cenozoic geomorphological and paleoenvironmental evolution in central Xizang remains a subject of intense debate, primarily due to methodological disparities among studies (Wang et al., 2014a; Su et al., 2019; Spicer et al., 2021a, 2021b; Ding et al., 2022). Several Cenozoic intermountain basins, developed in the Bangong-Nujiang Suture Zone (BNSZ) in central Xizang, accumulated up to 4 km of fluvio-lacustrine sediments that contain crucial paleogeographic evolution records. The isotopic compositions of carbonates obtained from those basins suggest arid and cold conditions during the Paleogene (Rowley and Currie, 2006; DeCelles et al., 2007b), whereas the same successions have yielded fossils of mammals (Deng et al., 2012), fish (Wu et al., 2017), plants (Jia et al., 2019; Jiang et al., 2019; Liu et al., 2019; Su et al., 2019; Tang et al., 2019), and biomarker remains (Jia et al., 2015) that reflect a warm and humid environment. The stable isotope analyses of paleosol carbonate in the Nima Basin point to an arid climate and high paleoelevation (4.5−5 km) by 26 Ma (Rowley and Currie, 2006; DeCelles et al., 2007a; Sun et al., 2014; Ingalls et al., 2020), whereas the paleontological evidence provided by the ecological signal of plant megafossils and fish suggests that tropical-subtropical warm and humid lowland habitats were present in the Xizang interior ca. 26 Ma, probably at an elevation of ~1 km (Wu et al., 2017). Additionally, reliable chronological frameworks (Han et al., 2019; Fang et al., 2020; Xiong et al., 2022) and numerical simulation models (Botsyun et al., 2019; Su et al., 2020) have resolved the disparities between paleontological and isotopic geochemistry investigations.
During the process of weathering, erosion, transport, deposition, and diagenesis, geochemical elements in source rocks exhibit stable chemical characteristics, resulting in minimal differentiation in fine-grained sediments (Nesbitt, 1979; Nesbitt and Markovics, 1997; Duddy, 1980; Bhatia, 1983; McLennan et al., 1983; Taylor and McLennan, 1985; Wronkiewicz and Condie, 1987; Crichton and Condie, 1993; Vital and Stattegger, 2000; Weltje and von Eynatten, 2004; Singh, 2009; Moradi et al., 2016; Caracciolo, 2020; Wang et al., 2020; Amorosi et al., 2022). The geochemistry of clastic sediments is an essential source of information for determining their provenance history (Girty et al., 1993; Cullers and Podkovyrov, 2000; Anderson et al., 2002; Wang et al., 2018a; Ramos-Vázquez and Armstrong-Altrin, 2019; Chen et al., 2020; Banerji et al., 2022), as well as their tectonic background (Roy et al., 2008; Gallala et al., 2009; Armitage et al., 2011; Bai et al., 2015; Wang et al., 2017a; Wang et al., 2018a; Lin et al., 2021; Sun et al., 2022).
The Paleocene to Eocene Niubao Formation is extensively distributed within the Bangong-Nujiang Suture Zone (BNSZ) and is one of the predominant Cenozoic strata in central Xizang. Although numerous paleontological identifications and hydrocarbon explorations have been conducted on the Niubao Formation, further geochemical studies are required to gain a comprehensive understanding of the paleoenvironment of the Niubao Formation. In this study, the geochemical characteristics of mudstones were analyzed in the Niubao Formation in the Nima Basin, with a comparative examination against neighboring igneous rocks. Additionally, the study applies an empirical formula to calculate the mean annual temperature (MAT) of the source area. The primary objective is to elucidate the origin and tectonic setting of the Niubao Formation and to provide a new perspective on the early Cenozoic paleogeomorphic pattern of central Xizang.

2 Geological background

The Qiangtang and Lhasa terranes are divided by the Bangong-Nujiang Suture Zone, which extends approximately 2000 km from east to west in central Xizang. The Nima Basin, one of the largest Cenozoic sedimentary basins within the BNSZ, spans an east–west-trending region with an area of 3000 km2 (Fig.1; DeCelles et al., 2007a, 2007b; Kapp et al., 2007; Chen et al., 2022). The Cenozoic terrestrial strata in the Nima Basin unconformably overlie deformed Mesozoic marine strata. The margins of the basin are strongly deformed by east–west basin-controlling faults along the northern and southern margins. Previous research has revealed that the Niubao Formation was deposited predominantly in a fluvial to deltaic environment during the Paleocene to Eocene. The Niubao Formation is distinguished by red upward-fining or upward-coarsening sequences of conglomerate, sandstone, siltstone, and mudstone, with occasional thin interlayers of marl and limestone that likely represent lacustrine facies. Recent investigations constructed a chronostratigraphic framework for the Niubao Formation using the ages of volcaniclastic interlayers as chronological anchors. Tuffaceous and paleomagnetic stratigraphic results in the Lunpola basin yielded an age range of 41.8−21.5 Ma for the middle to the upper part of the Niubao Formation (Fang et al., 2020). Su et al. (2020) obtained a zircon U-Pb age of 47.5 ± 0.7 Ma from a tephra-rich layer in the Niubao Formation in the Bangoin Basin. Xiong et al. (2022) established a depositional age range of 50−29 Ma for the Niubao Formation, extending from the base to the top of the formation.
Fig.1 (a) Primary terranes of the Xizang Plateau (modified from Li et al., 2015). (b) Geological map of the Nima Basin (modified from Chen et al., 2020).

Full size|PPT slide

In this study, a 410-m-thick stratigraphic section of the Niubao Formation was measured in the field on the south-eastern margin of the Nima Basin (31°57′04.30″N, 88°24′42.87″E; Fig.2). The sedimentary profile is dominated by red and grayish-green mudstones with a total thickness of approximately 200 m. The interbedded conglomerates, sandstones, and siltstones are characterized by kinds of sedimentary structure, including parallel or horizontal bedding, cross-bedding, and bioturbation. This section is near the Kanggale section (31°58′21.38″N, 88°26′10.19″E), where notable late Eocene fossils of Spittlebugs and Lestidae were discovered (Xia et al., 2022; Xu et al., 2022). The fossiliferous strata in Kanggale consist of similar grayish-green mudstones and calcareous shales, with interbedded mudstones, sandstones, and limestones within these layers (Xia et al., 2022). The striking similarity in lithologic composition and sedimentary features between the studied section and the Kanggale section indicates that the depositional age of the measured section in this study is assigned to the late Eocene.
Fig.2 The stratigraphic chart of the studied section and line chart depicting changes in paleoenvironmental and provenance indicators. The age of the section is determined through stratigraphic correlation (Xia et al., 2022; Wang et al., 2023).

Full size|PPT slide

3 Sampling and analytical methods

3.1 Sampling

Eighteen mudstone samples were systemically collected from different stratigraphic intervals for geochemical analysis. The sampling locations are shown in Fig.2. To minimize weathering effects, we selected unweathered samples and carefully removed weathering surfaces. All the samples were crushed in the laboratory before analysis.

3.2 Analytical methods

Major element, trace element, and rare earth element (REE) analyses were completed at Wuhan Sample Solution Analytical Technology Co., Ltd., Wuhan, China. Whole-rock major oxide concentrations were measured using an X-ray fluorescence spectrometer (XRF). The whole-rock major element analysis samples were prepared by the fusion bead method. For this procedure, samples were collected and pulverized to a size of less than 200 mesh using an agate mortar. Approximately 1 g of fine powdered sample was then fused with a 6 g flux, consisting of a well-balanced mixture of lithium tetraborate, lithium metaborate, and lithium fluoride at a ratio of 45:10:5, to form a uniform glass bead. Ammonium nitrate and lithium bromide were used as oxidant and release agents, respectively. The melting temperature was 1050°C and was maintained for 15 min. Measurements were made using a ZSX Primus II wavelength dispersive XRF instrument (RIGAKU, Japan) with a 4.0 kW Rh target X-ray tube at 50 kV and 60 mA. Quality control was ensured using physical and analysis replicates and the reference material GBW07101-14. The data were corrected by the theoretical α coefficient method, and the relative standard deviation (RSD) was less than 2%.
Whole-rock trace element analysis and REE analysis were conducted on an Agilent 7700e ICP-MS instrument. In the pretreatment phase for geochemical trace and REE analysis, the sample powder, with a 200 mesh size, was initially dried for 12 h at 105°C. A 50 mg portion of this powder was then accurately weighed and placed in a Teflon bomb, to which 1 mL of HNO3 and 1 mL of HF were added. This mixture was heated in a stainless steel jacket at 190°C for more than 24 h and then cooled. The contents were evaporated to near dryness, treated with HNO3, and evaporated again. A mixture of HNO3, Milli-Q water, and 1 mL of internal standard solution with 1 ppm In was added, and the Teflon bomb was resealed and placed in the oven for more than 12 h. The final solution was transferred to a polyethylene bottle and diluted with 2% HNO3 to a total of 100 g.
The analysis employed a GeolasPro laser ablation system integrated with a COMPexPro 102 ArF excimer laser, featuring a 193 nm wavelength and 200 mJ maximum energy, alongside a MicroLas optical system. Trace element compositions were determined using an Agilent 7700e ICP-MS instrument, with helium as the carrier gas and argon as the makeup gas, which was fed into the ICP through a T-connector. This system includes a “wire” signal smoothing device (Hu et al., 2015). The laser parameters were set to a 40 μm spot size and a 10 Hz frequency. Calibration was conducted using the reference materials NIST 610 and NIST 612. Each sample underwent a 20−30 s background acquisition followed by a 50 s data acquisition from the sample. The off-line selection, integration of the background and analytical signals, time-drift correction, and quantitative calibration were performed using the ICPMS DataCal 10.1 software (Lin et al., 2016).

4 Results

A total of 18 representative samples of the studied section were analyzed for major, trace, and rare earth elements. The results are given in Tab.1, Tab.2, Tab.3.
Tab.1 Composition of major elements, ICV, CIA, and MAT values in the mudstones of the Niubao Formation in the study area
SampleZP-01ZP-02ZP-03ZP-04ZP-05ZP-06ZP-07ZP-08ZP-09ZP-10ZP-11ZP-12ZP-13ZP-14ZP-15ZP-16ZP-17ZP-18
SiO254.5047.0550.7053.2458.5764.4567.8260.7073.1063.4250.0741.1254.8650.8954.0137.2352.6051.01
TiO20.690.660.690.720.680.780.610.730.520.640.760.440.760.650.690.530.580.60
Al2O319.3317.2717.0419.7014.6812.3410.0216.438.5712.1017.9114.5717.7519.1616.8412.807.839.50
TFe2O35.156.276.175.465.483.673.375.863.765.076.504.406.906.085.694.993.143.76
MnO0.040.070.120.070.120.110.150.060.080.110.090.110.060.050.070.160.120.11
MgO2.694.363.373.513.162.651.542.591.321.913.482.862.383.493.446.162.513.08
CaO3.637.386.964.085.055.366.473.014.835.835.7215.473.923.485.3614.1715.0513.33
Na2O1.040.811.320.951.251.401.281.011.381.822.190.890.871.340.850.951.131.25
K2O3.963.983.633.882.832.081.642.981.091.983.403.023.993.863.203.181.361.96
P2O50.150.130.170.150.150.160.110.100.080.120.150.110.130.120.090.140.120.12
LOI8.7411.339.827.918.107.387.366.565.707.289.4116.368.1310.239.8619.2615.7414.61
Total91.1887.9790.1891.7591.9593.0093.0093.4794.7292.9990.2682.9891.6389.1190.2480.3184.4484.70
ICV0.971.331.271.051.331.381.281.081.421.411.391.161.051.161.162.041.831.81
CIA71.4971.1667.3072.8367.1664.2162.5871.4460.0459.7962.2070.2171.1769.0372.9066.0460.1360.44
MAT14.7911.9913.792.718.1113.8916.3512.9115.4415.1714.304.3215.339.1415.425.5010.1810.44
Tab.2 Trace element concentration in mudstones of the Niubao Formation (unit in ppm)
SampleZP-01ZP-02ZP-03ZP-04ZP-05ZP-06ZP-07ZP-08ZP-09ZP-10ZP-11ZP-12ZP-13ZP-14ZP-15ZP-16ZP-17ZP-18
Sc17.016.415.917.213.511.68.814.57.210.617.812.216.316.415.311.88.110.1
V1231311401219583631005171120114114121104905872
Cr971121101058710668836673106631029892797180
Co11.116.917.013.214.913.510.213.47.212.317.812.014.015.014.214.39.512.2
Ni678061625358324930397246756962504156
Cu313429272719312514306225212925241819
Zn10090841098478579359708577958487684355
Ga23.623.322.225.619.415.512.121.410.414.622.719.123.125.021.616.79.512.1
Y32.325.925.430.730.627.429.731.323.730.024.523.331.527.430.420.422.623.1
Zr141131146153167200203172249217150100153138156107132120
Nb14.614.113.015.314.415.313.215.611.314.512.99.616.514.114.910.511.512.0
Hf3.893.563.824.084.395.225.274.636.375.654.182.954.263.794.302.983.433.23
Ta1.080.990.921.151.081.060.971.150.841.091.030.811.261.081.130.810.850.88
Th20.2714.9313.5618.5115.6912.3112.4317.7611.3515.8912.8416.6518.5117.3018.0311.809.5410.99
U6.775.714.044.893.543.242.633.612.743.212.165.492.784.153.222.272.042.23
Rb194.48177.27170.51193.46143.62100.9785.58154.3157.91107.34156.25155.66196.24192.18166.07137.5164.3990.66
Sr188293152144129108140150180150210322260432408387329244
Cs22.3919.1913.9115.2113.075.968.0915.054.4911.1619.0916.3720.8822.5117.2215.795.728.01
Ba289467255277244219739290173601256261304244360468131166
Pb32.623.729.420.030.312.018.325.719.134.922.127.324.727.528.422.118.419.8

4.1 Major elements

Major element analyses were carried out on 18 mudstone samples of the Niubao Formation of the Nima Basin, as detailed in Tab.1. The total content of major components in the mudstone ranges from 80.31 to 94.72 wt.%. The results revealed that SiO2, Al2O3, CaO, and Fe2O3 were the predominant components, followed closely by MgO and K2O. The concentrations of other components, such as TiO2, P2O5, and MnO, were generally less than 1.0%.
The mudstone samples exhibited noticeable variations in their elemental compositions. The SiO2 content varied widely, from 37.23 to 73.10 wt.%, averaging 54.74 wt.%. Al2O3 ranged from 7.83 to 19.70 wt.%, with a mean of 14.66 wt.%. The content of total Fe2O3 ranges from 3.14 to 6.90 wt.%, averaging at 5.10 wt.%. The TiO2 content varies between 0.44 and 0.78 wt.%, with a mean value of 0.65 wt.%. The CaO content spaned from 3.01 to 15.47 wt.%, averaging at 7.17 wt.%. The K2O content is 1.09−3.99 wt.% with an average of 2.89 wt.%. The Na2O content is in the range of 0.81−2.19 wt.% with an average of 1.21 wt.%. Additionally, the loss on ignition (LOI) ranged from 5.70 to 19.26 wt.%, potentially attributable to organic matter or volatile components.
The ratios of the concentrations of the major elements in the mudstones to those in the upper continental crust (UCC) are shown in Fig.3(a). In addition to some samples being enriched in CaO and K2O, all the samples exhibited apparent depletion in Na2O, with the remaining element concentrations plotting near those in the UCC.
Fig.3 (a) Upper continental crust (UCC)-normalized ratios for major elements. (b) Trace element compositions normalized against the upper continental crust (UCC) for the mudstones. (c) Chondrite-normalized rare earth elements (REEs) plot for the mudstone samples.

Full size|PPT slide

4.2 Trace and REEs

The trace element data of the mudstone samples are listed in Tab.2 and have been standardized in relation to the average UCC values (Taylor and McLennan, 1985; Fig.3(b)). The trace element analyses reveal that, compared with the UCC, the mudstone samples are significantly enriched in Ni, Zn, Rb, Cs, Th, and U and depleted in Co, Sr, Ba, Zr, and Hf.
The REE analysis results are presented in Tab.3 and data are plotted in the chondrite-normalized diagram (Fig.3(c)). The mudstones have total REE (ΣREE) ranging from 124.81 to 204.83 ppm, with an average of 162.34 ppm. the average value is slightly lower than the post-Archean Australian shale (PAAS) value of 211.76 ppm (Taylor and McLennan, 1985; McLennan et al., 2006). In terms of chondrite-normalized REE patterns, the mudstones closely align with PAAS. Light REEs (LREEs) in mudstones range from 109.76 to 183.43 ppm, while heavy REEs (HREEs) range from 13.86 to 21.40 ppm. Generally, the mudstones exhibit higher concentrations of LREEs than HREEs. The ratios of LREEs to HREEs in these samples vary from 7.29 to 9.24, with an overall average of 8.29, indicating relative enrichment of LREEs in the mudstones. Additionally, the mudstones exhibit consistent negative Eu anomalies, with values ranging from 0.59 to 0.70 and an average of 0.65.
Tab.3 REE contents in mudstones of the Niubao Formation (unit in ppm)
SampleZP-01ZP-02ZP-03ZP-04ZP-05ZP-06ZP-07ZP-08ZP-09ZP-10ZP-11ZP-12ZP-13ZP-14ZP-15ZP-16ZP-17ZP-18
Sc17.016.415.917.213.511.68.814.57.210.617.812.216.316.415.311.88.110.1
V1231311401219583631005171120114114121104905872
Cr971121101058710668836673106631029892797180
Co11.116.917.013.214.913.510.213.47.212.317.812.014.015.014.214.39.512.2
Ni678061625358324930397246756962504156
Cu313429272719312514306225212925241819
Zn10090841098478579359708577958487684355
Ga23.623.322.225.619.415.512.121.410.414.622.719.123.125.021.616.79.512.1
Y32.325.925.430.730.627.429.731.323.730.024.523.331.527.430.420.422.623.1
Zr141131146153167200203172249217150100153138156107132120
Nb14.614.113.015.314.415.313.215.611.314.512.99.616.514.114.910.511.512.0
Hf3.893.563.824.084.395.225.274.636.375.654.182.954.263.794.302.983.433.23
Ta1.080.990.921.151.081.060.971.150.841.091.030.811.261.081.130.810.850.88
Th20.2714.9313.5618.5115.6912.3112.4317.7611.3515.8912.8416.6518.5117.3018.0311.809.5410.99
U6.775.714.044.893.543.242.633.612.743.212.165.492.784.153.222.272.042.23
Rb194.48177.27170.51193.46143.62100.9785.58154.3157.91107.34156.25155.66196.24192.18166.07137.5164.3990.66
Sr188293152144129108140150180150210322260432408387329244
Cs22.3919.1913.9115.2113.075.968.0915.054.4911.1619.0916.3720.8822.5117.2215.795.728.01
Ba289467255277244219739290173601256261304244360468131166
Pb32.623.729.420.030.312.018.325.719.134.922.127.324.727.528.422.118.419.8

4.3 Mineral composition

The high SiO2 and low Al2O3 contents suggest that clay minerals are the main mineral component of the mudstone. The low Na2O concentrations suggest the minimal presence of plagioclase in the samples (Ross and Bustin, 2009; Zeng et al., 2020). The high levels of CaO are likely attributable to the presence of carbonate minerals, which are present in the formation of cement (Ghosh and Sarkar, 2010). Notably, the mudstones exhibit significant correlations between TFe2O3 and Al2O3 (R2 = 0.847) and K2O (R2 = 0.865; Tab.4), indicating the close association of Fe with clay minerals.
Tab.4 Correlation matrix of major and trace elements of the mudstones (* indicates p < 0.05; ** indicates p < 0.01, both suggesting statistically significant correlations)
Al2O3 SiO2 TiO2 TFe2O3 MnO CaO MgO Na2O K2O P2O5 Sc V Cr Co Ni
Al2O3 1
SiO2 −0.618** 1
TiO2 0.502* −0.089 1
TFe2O3 0.847** −0.612** 0.494* 1
MnO −0.516* −0.151 −0.391 −0.374 1
CaO −0.382 −0.465 −0.528* −0.315 0.692** 1
MgO 0.390 −0.832** 0.092 0.455 0.350 0.432 1
Na2O −0.216 0.206 0.116 −0.078 0.174 −0.075 −0.181 1
K2O 0.953** −0.769** 0.433 0.865** −0.347 −0.178 0.572* −0.318 1
P2O5 0.386 −0.540* 0.432 0.303 0.284 0.154 0.525* 0.105 0.488* 1
Sc 0.968** −0.540* 0.672** 0.860** −0.551* −0.462 0.332 −0.111 0.909** 0.424 1
V 0.945** −0.665** 0.474* 0.829** −0.396 −0.266 0.398 −0.197 0.936** 0.506* 0.934** 1
Cr 0.701** −0.317 0.826** 0.645** −0.427 −0.478* 0.292 −0.010 0.677** 0.563* 0.816** 0.754** 1
Co 0.685** −0.601** 0.623** 0.810** −0.055 −0.171 0.550* 0.149 0.712** 0.544* 0.760** 0.763** 0.758** 1
Ni 0.806** −0.558* 0.704** 0.773** −0.501* −0.288 0.410 −0.142 0.833** 0.454 0.877** 0.819** 0.881** 0.756** 1
Cu 0.476* −0.304 0.390 0.495* −0.094 −0.226 0.168 0.579* 0.392 0.278 0.558* 0.477* 0.400 0.610** 0.414
Zn 0.900** −0.327 0.576* 0.722** −0.637** −0.605** 0.159 −0.288 0.801** 0.297 0.897** 0.820** 0.695** 0.542* 0.697**
Ga 0.989** −0.535* 0.549* 0.853** −0.564* −0.475* 0.322 −0.227 0.925** 0.347 0.974** 0.937** 0.737** 0.698** 0.807**
Rb 0.983** −0.609** 0.484* 0.853** −0.507* −0.371 0.365 −0.327 0.963** 0.372 0.944** 0.931** 0.670** 0.649** 0.799**
Sr 0.208 −0.578* −0.224 0.268 0.021 0.438 0.536* −0.271 0.326 −0.159 0.098 0.144 −0.078 0.121 0.245
Nb 0.424 0.265 0.779** 0.371 −0.628** −0.809** −0.263 −0.159 0.296 0.017 0.530* 0.313 0.595** 0.299 0.478*
Cs 0.921** −0.645** 0.365 0.855** −0.480* −0.276 0.399 −0.206 0.921** 0.247 0.870** 0.835** 0.519* 0.595** 0.760**
Ba −0.047 0.089 −0.108 0.019 0.301 −0.102 0.025 −0.006 −0.012 −0.152 −0.085 −0.097 −0.168 0.030 −0.198
Th 0.766** −0.199 0.301 0.557* −0.681** −0.542* −0.063 −0.413 0.652** −0.024 0.695** 0.618** 0.326 0.220 0.459
U 0.610** −0.296 −0.015 0.272 −0.463 −0.214 0.050 −0.432 0.568* 0.204 0.512* 0.631** 0.309 0.128 0.381
Zr −0.458 0.913** −0.050 −0.395 −0.164 −0.589* −0.705** 0.304 −0.624** −0.471* −0.394 −0.519* −0.193 −0.420 −0.485*
Hf −0.402 0.894** −0.025 −0.343 −0.214 −0.625** −0.710** 0.304 −0.578* −0.491* −0.341 −0.477* −0.174 −0.392 −0.445
Yb 0.556* 0.197 0.564* 0.448 −0.635** −0.833** −0.322 −0.143 0.387 −0.024 0.602** 0.428 0.430 0.269 0.375
∑REE 0.766** −0.154 0.478* 0.539* −0.684** −0.617** −0.027 −0.475* 0.672** 0.121 0.745** 0.651** 0.507* 0.291 0.566*
LREE 0.786** −0.189 0.469* 0.557* −0.681** −0.598** 0.005 −0.489* 0.697** 0.131 0.758** 0.672** 0.516* 0.306 0.582*
HREE 0.471* 0.221 0.511* 0.284 −0.623** −0.730** −0.357 −0.271 0.328 −0.003 0.510* 0.345 0.357 0.099 0.330
Cu Zn Ga Rb Sr Nb Cs Ba Th U Zr Hf Yb ∑REE LREE HREE
Al2O3
SiO2
TiO2
TFe2O3
MnO
CaO
MgO
Na2O
K2O
P2O5
Sc
V
Cr
Co
Ni
Cu 1
Zn 0.333 1
Ga 0.455 0.932** 1
Rb 0.370 0.898** 0.977** 1
Sr −0.079 −0.125 0.131 0.212 1
Nb 0.080 0.647** 0.511* 0.465 −0.325 1
Cs 0.485* 0.748** 0.886** 0.927** 0.399 0.320 1
Ba 0.235 −0.036 −0.053 −0.012 −0.080 0.109 0.071 1
Th 0.122 0.845** 0.779** 0.811** 0.032 0.604** 0.742** 0.085 1
U 0.087 0.631** 0.594** 0.616** −0.039 0.181 0.532* 0.007 0.668** 1
Zr −0.139 −0.155 −0.373 −0.472* −0.593** 0.299 −0.514* 0.219 −0.117 −0.256 1
Hf −0.097 −0.099 −0.318 −0.417 −0.578* 0.330 −0.449 0.233 −0.054 −0.234 0.995** 1
Yb 0.235 0.759** 0.625** 0.594** −0.356 0.868** 0.487* 0.205 0.820** 0.393 0.279 0.328 1
∑REE 0.127 0.904** 0.799** 0.813** −0.124 0.735** 0.692** 0.085 0.941** 0.701** −0.095 −0.044 0.860** 1
LREE 0.128 0.914** 0.816** 0.831** −0.098 0.715** 0.710** 0.075 0.941** 0.714** −0.124 −0.072 0.840** 0.999** 1
HREE 0.096 0.687** 0.526* 0.529* −0.369 0.840** 0.417 0.179 0.809** 0.477* 0.211 0.248 0.952** 0.880** 0.857** 1
The K2O/Al2O3 ratio provides valuable insight into the primary mineral constituents within sediments. A ratio between 0.2 and 0.3 generally implies the dominance of illite in the sediment. Conversely, a ratio near 0 is indicative of sediment enriched with minerals such as kaolinite, smectite, and vermiculite (Cox et al., 1995; Zhou et al., 2015; Zeng et al., 2020). The K2O/Al2O3 ratios in mudstone vary from 0.13 to 0.25, averaging 0.19 (Tab.5). The highest value was obtained for the ZP-16 sample (0.25), and the lowest value was obtained for the ZP-09 sample (0.13). These data indicate that the ratio aligns with typical values for clay minerals, suggesting that illite is the predominant clay mineral present. Additionally, the notable associations between Sc, V, Zn, Ga, Rb, and both Al2O3 and K2O indicate that these elements are predominantly related to clay minerals (Tab.4).
Tab.5 Ratios table of selected major and trace elements of the mudstones
Sample ZP-01 ZP-02 ZP-03 ZP-04 ZP-05 ZP-06 ZP-07 ZP-08 ZP-09 ZP-10 ZP-11 ZP-12 ZP-13 ZP-14 ZP-15 ZP-16 ZP-17 ZP-18 Average Max Min
K2O/Al2O3 0.20 0.23 0.21 0.20 0.19 0.17 0.16 0.18 0.13 0.16 0.19 0.21 0.22 0.20 0.19 0.25 0.17 0.21 0.19 0.25 0.13
SiO2/Al2O3 2.82 2.73 2.98 2.70 3.99 5.22 6.77 3.69 8.53 5.24 2.80 2.82 3.09 2.66 3.21 2.91 6.71 5.37 4.12 8.53 2.66
K2O/Na2O 3.81 4.89 2.75 4.06 2.27 1.48 1.28 2.94 0.79 1.09 1.55 3.41 4.59 2.88 3.78 3.33 1.20 1.57 2.65 4.89 0.79
Al2O3/TiO2 27.98 26.36 24.70 27.20 21.75 15.92 16.40 22.57 16.42 18.82 23.50 32.89 23.27 29.71 24.44 24.15 13.55 15.84 22.53 32.89 13.55
Sr/Ba 0.65 0.63 0.60 0.52 0.53 0.49 0.19 0.52 1.04 0.25 0.82 1.23 0.86 1.77 1.13 0.83 2.52 1.47 0.89 2.52 0.19
Sr/Cu 6.01 8.65 5.16 5.35 4.71 5.57 4.49 5.93 13.04 4.95 3.38 12.85 12.38 14.74 16.51 16.04 18.71 12.89 9.52 18.71 3.38
V/Cr 1.27 1.17 1.27 1.15 1.10 0.79 0.93 1.22 0.77 0.98 1.14 1.81 1.13 1.23 1.13 1.14 0.82 0.89 1.11 1.81 0.77
Ni/Co 6.02 4.74 3.61 4.74 3.55 4.27 3.16 3.67 4.20 3.16 4.03 3.81 5.36 4.62 4.41 3.51 4.27 4.58 4.20 6.02 3.16
Th/Sc 1.19 0.91 0.85 1.08 1.16 1.06 1.41 1.23 1.58 1.50 0.72 1.36 1.14 1.05 1.18 1.00 1.18 1.09 1.15 1.58 0.72
Zr/Sc 8.31 7.96 9.14 8.90 12.42 17.19 22.92 11.86 34.72 20.43 8.43 8.19 9.39 8.38 10.23 9.11 16.40 11.86 13.10 34.72 7.96
Co/Th 0.55 1.13 1.25 0.71 0.95 1.10 0.82 0.75 0.64 0.77 1.38 0.72 0.76 0.87 0.79 1.21 1.00 1.11 0.92 1.38 0.55
La/Sc 2.50 2.33 2.10 2.45 2.72 2.87 3.46 2.75 3.78 3.08 1.70 2.80 2.46 2.24 2.48 2.57 3.18 2.76 2.68 3.78 1.70
Ti/Zr 32.22 34.17 31.52 30.97 26.30 24.95 19.44 27.18 13.27 19.15 33.77 32.00 32.67 31.49 29.27 36.93 31.01 35.47 28.99 36.93 13.27

5 Discussion

5.1 Sediment maturity and paleoweathering

The index of compositional variability (ICV) can determine the original character and maturity of sediments. Generally, ICV value < 1 indicates that the sample is intensively mature, rich in clay minerals, and may have undergone recycling or strong weathering during the first cycle of deposition, while ICV value > 1 indicates that the sample has a low maturity and a low clay mineral content (Barshad, 1966). For the studied samples, the ICV values of all the mudstones exhibit more than 1, except sample ZP17-01, which has an ICV value of 0.97 (Tab.1). Based on these values, it can be inferred that the rocks have a minimal clay content and the sediment is compositionally immature.
The SiO2/Al2O3 ratios depend on sediment recycling and weathering processes, indicating sediment maturity. The average ratio of SiO2/Al2O3 greater than 5.0 indicates progressive maturity sediment (Potter, 1978; Roser et al., 1996; Madukwe et al., 2016; Shekhar et al., 2018). The SiO2/Al2O3 ratios of our samples vary from 2.66 to 8.53 (Tab.5), indicating an intermediate maturity level, which can be compared with the ICV values.
Alternatively, the Th/Sc and Zr/Sc ratios are further used to infer the degree of sedimentary recycling (McLennan et al., 1995; Bai et al., 2015; Qiu et al., 2015; Chen et al., 2020). As recycling progresses, there is a notable increase in Zr/Sc, while Th/Sc remains stable (McLennan et al., 1993; Li et al., 2022a). The Th/Sc ratios in our samples range from 0.72 to 1.58 while the Zr/Sc ratios range from 7.96 to 34.72 (Tab.5). As shown in the Zr/Sc-Th/Sc diagram (Fig.4), the data spots corresponding to all mudstones are relatively concentrated and exhibit a weak sedimentary recycling trend, which suggests insignificant recycling with an initial depositional environment.
Fig.4 Diagram of Zr/Sc versus Th/Sc for the mudstone samples.

Full size|PPT slide

The chemical index of alteration (CIA) has been commonly used to estimate the intensity of chemical weathering in the source region (Nesbitt and Young, 1982). The formula can be shown as CIA = [Al2O3 / [Al2O3 + CaO* + Na2O + K2O] × 100, with all the elements expressed in molecular percentages. Here, CaO* represents the amount incorporated in the silicate fraction of the rock, defined as CaO* = CaO−10/3 × P2O5 (McLennan et al., 1993). The CIA values between 50 and 60 represent a low degree of chemical weathering in cold, dry climates. A range of 60−80 indicates moderate weathering, which is typical of warm, humid conditions. Finally, values ranging from 80 to 100 represent intense weathering observed in hot, humid tropical-subtropical climates, indicating that the parent rock has undergone substantial weathering (Nesbitt and Young, 1982; Fedo et al., 1995; Bock et al., 1998; Xu and Shao, 2018). The CIA values of the mudstone samples are given in Tab.1. They range between 59.79 and 72.90 with an average of 66.70, indicating that the Niubao Formation mudstones suffered from low to moderate weathering. As depicted on the A-CN-K (Al2O3-CaO + Na2O-K2O) ternary diagram, the mudstone samples show an evolution trend from the upper crust to kaolinite along the A-CN axis (Fig.5). Even though the samples also deviate from the ideal weathering trend (IWT) and are inclined toward illite, which could be attributed to K-metasomatism during diagenesis (Fedo et al., 1995; Michalopoulos and Aller, 1995), the subtle divergence of samples implies a weak diagenetic impact (Ma et al., 2015).
Fig.5 A-CN-K ternary plot and associated chemical index of alteration (CIA) variation. Dashed lines represent the ideal weathering trends of tonalite, granodiorite, and granite, respectively (Fedo et al., 1995).

Full size|PPT slide

5.2 Paleoenvironmental conditions

5.2.1 Paleosalinity and paleoclimate

The Sr/Ba ratios are a diagnostic indicator for determining paleosalinity and paleoclimate variations. High Sr/Ba ratios, typically above 1.0, signify water columns with high salinity prevalent in arid climatic settings, whereas values less than 0.5 represent water columns with low salinity associated with humid climatic conditions (Meng et al., 2012; Fu et al., 2016; Wang et al., 2017a). Among the analyzed samples, six samples had values more than 1, while two samples had values less than 0.5 (Tab.5; Fig.2), potentially highlighting variation in paleosalinity over time and the transient dry and humid climate change of the sedimentary environment. In addition, Sr/Cu ratios less than 5.0 are indicative of humid climate conditions, whereas those exceeding 5.0 are characteristic of arid climate conditions (Lerman, 1989; Meng et al., 2012; Jia et al., 2013; Liang et al., 2014; Cao et al., 2015; Fu et al., 2016, 2018). The Sr/Cu ratio was close to 5.0 for four samples, over 5 for all other samples, and greater than 12 for most of the samples, indicating relatively dry climate conditions (Tab.5).
Certain elements, such as V, Co, and Cu, are recognized as redox-sensitive and commonly accumulate in anoxic sediments (Algeo and Maynard, 2004; Tribovillard et al., 2006; Lézin et al., 2013; Ayinla et al., 2017). Key trace element ratios, such as V/Cr and Ni/Co ratios, are reliable indicators of redox conditions. The V/Cr ratios, which typically increase with decreasing oxygen content in a water column, can be used to classify environments: ratios less than 2.0 suggest oxic conditions, those between 2.0 and 4.25 indicate dysoxic conditions, and values above 4.25 signify anoxic conditions (Jones and Manning, 1994). The V/Cr ratios of the mudstones fluctuate from 0.77 to 1.81 (Tab.5; Fig.2), indicating oxic conditions. Similarly, Ni/Co ratio values less than 5.0 indicate an oxidizing environment, values between 5.0 and 7.0 suggest suboxic to dysoxic conditions, and values over 7.0 indicate reducing conditions (Jones and Manning, 1994). Our mudstone samples have Ni/Co ratios between 3.16 and 6.02 (Tab.5; Fig.2), further supporting oxic conditions during deposition.

5.2.2 Mean average temperature (MAT) calculation

Chemical weathering processes are significantly influenced by climatic factors (Nesbitt and Young, 1982). In the weathering of granitic bedrock soils, the sodium depletion fraction index (τNa) effectively indicates terrestrial chemical weathering intensity (Gaillardet et al., 1999; Anderson et al., 2002; Rasmussen et al., 2011; West, 2012; Garzanti et al., 2013). The intensity of chemical weathering in the topsoil of weathered granite bedrock shows a positive correlation with the mean annual surface temperature (MAT) in modern non-glacial areas (Yang et al., 2016). The empirical equation can express this relationship: MAT = −24.2τNa−0.9 (r2 = 0.84, P < 0.0001). Four specific criteria must be considered before employing this equation. First, annual precipitation should range between 400 mm and 4000 mm, avoiding extreme aridity or excessive humidity conditions. Secondly, the source rock should be felsic igneous rocks. Thirdly, the physical erosion rate should be 2−100 m/m.y. Fourthly, sorting and sedimentary recycling processes during transport need to be comparatively weak, and finally, diagenetic processes should be minimal.
Evidence from late Eocene fossils suggests the presence of tropical to subtropical climates (Fang et al., 2020). The diagram of Zr vs TiO2 for the samples points out that the mudstones predominantly originate from the felsic igneous rocks source area (Fig.6(b)). Furthermore, the erosion rate in the central regions of the plateau during the Eocene (~45 Ma) was below 50 m/m.y (Rohrmann et al., 2012). It is coupled with weak recycling effects (Fig.4) and the absence of diagenetic alteration minerals, which conform to the application conditions of the empirical formula. Using granite and granodiorite as the original rock (Chi, 2007), the τNa of 18 mudstone samples spans from 0.93 to 2.42 with an average of 1.34. These results suggest that the average land surface temperature during the late Eocene is estimated to be 11.64°C ± 4.19°C (Tab.1; Fig.2).
Fig.6 (a) Diagram of Al2O3 versus TiO2 for the mudstone samples. (b) Diagram of Zr versus TiO2 for the mudstone samples.

Full size|PPT slide

The carbonate formation temperature of paleosol carbonates derived from clumped isotope analysis can represent the surface temperature during formation (Ghosh et al., 2006). The carbonate formation temperatures in the upper member of the Niubao Formation range between 10.1°C and 12.9°C, with an average of 11.9°C and a standard deviation of 2.8°C (Xiong et al., 2022). Ingalls et al. (2020) interpreted the T(Δ47) values of ~30°C to 36°C in soil carbonate in the late Eocene to Oligocene upper Niubao Formation in the Lunpola Basin as the maximum possible surface temperature during summer. It has been shown that terrestrial soil carbonates recorded T(Δ47) values approximately 18°C warmer than MAT at midlatitudes during the Paleocene-Eocene (Snell et al., 2013). Thus, the findings suggest an annual average temperature of 12°C to 18°C during the late Eocene to Oligocene.
The numerical climate modeling and proxy thermal regimes yield consistent results, indicating a shift toward a drier and cooler climate in central Xizang during the mid-late Eocene (Su et al., 2019; Chen et al., 2021; Zhang et al., 2022). The presence of multiple animal fossils alongside a diverse grass community, palms, and a variety of other woody species suggests the existence of a subtropical open woodland ecosystem, with an estimated dry bulb mean annual temperature of ~15.6°C (Deng et al., 2012; Wu et al., 2017; Wang et al., 2018b; Mao et al., 2019; Del Rio et al., 2020; Xu et al., 2022; Zhang et al., 2022).
In summary, the paleotemperature inferred from the geochemical proxies in mudstones in the Nima Basin is in concordance with the results obtained from paleontological and isotopic investigations, providing support for the hypothesis that the central plateau exhibited a cooler climatic regime during the late Eocene.

5.3 Source lithotypes and tectonic setting

5.3.1 Source lithotypes

The REE parameters are relative to the sources of sedimentary rocks due to their high stability during weathering, erosion, and early diagenesis (McLennan et al., 1995; Bock et al., 1998; Asiedu et al., 2000; Cullers and Podkovyrov, 2000; Kasanzu et al., 2008; Armstrong-Altrin et al., 2009). Generally, negative Eu anomalies reflect felsic rocks, whereas minimal or no Eu anomalies mainly indicate mafic rocks (Roddaz et al., 2006; Kasanzu et al., 2008; Armstrong-Altrin et al., 2017). The distribution patterns of mudstone samples chondrite-normalized REEs are relatively consistent, indicating that the mudstones originated from the same source. The REE profiles of the mudstones also exhibit significant negative Eu anomalies (Fig.3(c)), strongly indicating that the mudstones were derived from felsic rocks. Additionally, the REE exhibit a negative correlation with CaO (R2 = 0.617; Tab.3), indicating that the REE concentrations in the mudstone samples are predominantly controlled by terrigenous material (Zeng et al., 2020).
The geochemical compositions of detrital deposits provide vital insights into their provenance (Wesolowski, 1992; Ayers and Watson, 1993; Cullers and Podkovyrov, 2000; Verma and Armstrong-Altrin, 2013, 2016; Armstrong-Altrin et al., 2013, 2015a, 2015b, 2017; Madhavaraju, 2015). The Al2O3/TiO2 ratio in clastic sediments is widely considered an effective indicator of the lithology of sedimentary source rocks. Mafic rocks have Al2O3/TiO2 ratios ranging from 3 to 8, intermediate rocks have ratios ranging from 8 to 21, and felsic rocks have ratios ranging from 21 to 70 (Hayashi et al., 1997; Armstrong-Altrin et al., 2015a, 2017; Zhou et al., 2015; Moradi et al., 2016; Verma and Armstrong-Altrin, 2016; Wang et al., 2017b). The ratios of our samples range from 13.55 to 32.89, averaging 22.53 (Tab.5; Fig.6(a)). Most samples exhibit values characteristic of felsic rocks, while only six samples display values associated with intermediate rocks. Similarly, the TiO2/Zr ratio can be used to investigate the origin of sediments. A TiO2/Zr ratio up to 200 suggests that the sediments originated from the erosion of mafic rocks. Alternatively, ratios between 55 and 200 are indicative of intermediate igneous rock sources, and ratios less than 55 are indicative of felsic igneous rock sources (Hayashi et al., 1997; Zeng et al., 2020; Kirubakaran et al., 2023). The Zr-TiO2 graph (Fig.6(b)) and the Al2O3 vs TiO2 (Fig.6(a)) diagram for the mudstone samples indicate a predominance of felsic igneous source rocks and subordinate intermediate igneous source rocks.
The Co/Th ratio is a distinctive indicator of felsic versus mafic sources, as felsic rocks typically exhibit lower Co/Th values than mafic rocks (Amorosi et al., 2002; Cullers, 2002). Our collected samples feature Co/Th values between 0.55 and 1.37, which are notably lower than the benchmarks of the UCC (1.65) and PAAS (1.37), suggesting a predominantly felsic source (Tab.5; Ma et al., 2015). Similarly, the La/Sc ratio has been recognized as a reliable source composition indicator (Floyd and Leveridge, 1987; Condie, 1993; McLennan et al., 1995; Gu et al., 2002; Ma et al., 2015). T he La/Sc-Co/Th plot (Fig.7(a)) reveals that the samples are predominantly concentrated between granodiorite and adamellite. Additionally, on the Zr/Sc-Th/Sc plot (Fig.7(b)), the majority of the data plot in the felsic source area.
Fig.7 (a) Diagram of La/Sc versus Co/Th for the mudstone samples. The various igneous rock data were extracted from Condie (1993). (b) Diagram of Hf‐versus‐La/Th for the mudstones.

Full size|PPT slide

In conclusion, geochemical results suggest that the mudstones of the Niubao Formation predominantly originated from the source area of felsic rocks with abundant felsic igneous rocks and minor intermediate igneous rocks.

5.3.2 Tectonic setting

The geochemical characteristics of sedimentary rocks could reveal the composition and tectonic setting of source rocks (Bhatia and Crook, 1986; Roser and Korsch, 1988). Clastic sediments from various tectonic environments can exhibit unique terrane-specific imprints, leading to diverse geochemical compositions (Roser et al., 1996; Bai et al., 2015).
The discriminant function diagram illustrates the mudstones plotted in the collision area (Fig.8; Verma and Armstrong-Altrin, 2013). The La-Th-Sc plot (Fig.9(a)) shows that most of mudstone samples cluster within and near the edges of the continental island arc field. In the Th-Sc-Zr/10 plot (Fig.9(b)), there is an observable spread of sample points, with most of them positioned in the continental island arc and a few scattered within the passive margin. The Th-Co-Zr/10 diagram (Fig.9(c)) displays some mudstone samples within the continental island arc field, while others gravitate toward the boundaries of the active continental margin. In addition, the mudstones have high ΣREE values (124.81−204.83 ppm), (La/Yb)N ratios (8.47−10.99), and negative Eu anomalies (Eu/Eu* values ranging from 0.59 to 0.70) reinforce the conclusion of the continental island arc (Bhatia and Crook, 1986; Tab.4).
Fig.8 Discriminant function diagram illustrating the tectonic setting of the Niubao Formation (Verma and Armstrong-Altrin, 2013). DF1 = (0.608 × ln(TiO2/SiO2)) + ( − 1.854 × ln(Al2O3/SiO2)) + (0.299 × ln(Fe2O3/SiO2)) + ( − 0.550 × ln(MnO/SiO2)) + (0.120 × ln(MgO/SiO2)) + (0.194 × ln(CaO/SiO2) + ( − 1.510 × ln(Na2O/SiO2) + (1.941 × ln(K2O/SiO2)) + (0.003 × ln(P2O5/SiO2))−0.294; and DF2 = ( − 0.554 × ln(TiO2/SiO2)) + ( − 0.995 × ln(Al2O3/SiO2)) + (1.765 × ln(Fe2O3/SiO2)) + ( − 1.391 × ln(MnO/SiO2)) + ( − 1.034 × ln(MgO/SiO2)) + (0.225 × ln(CaO/SiO2) + (0.713 × ln(Na2O/SiO2) + (0.330 × ln(K2O /SiO2)) + (0.637 × ln(P2O5/SiO2)) − 3.631.

Full size|PPT slide

Fig.9 (a) La-Th-Sc plot for mudstone samples. (b) Th-Sc-Zr/10 plot of mudstone samples. (c) Th-Co-Zr/10 plot for the mudstone samples. (A: Oceanic island arc; B: Continental island arc; C: Active continental margin; D: Passive margin. Bhatia and Crook, 1986; Basu et al., 2016).

Full size|PPT slide

5.4 Provenance interpretation

As an intermontane basin proposed by previous work, it is reasonable to speculate that the neighboring geological units fed the Niubao Formation in the Nima Basin, namely the southern Qiangtang, the northern Lhasa, and the BNSZ (Rowley et al., 2015; Mi et al., 2018; Han et al., 2022), which are characterized by widely distributed pre-Cenozoic felsic and intermediate igneous rocks, particularly Cretaceous and Jurassic in age (Chapman and Kapp, 2017). According to the detrital zircon U-Pb age distributions of the Niubao Formation, zircons originating from these igneous rocks are the predominant components of the distinct peaks (Kapp et al., 2007; Donaldson et al., 2013; Zhu et al., 2017; Li et al., 2022b; Liu et al., 2023). Several representative intrusive and volcanic rocks are compared to the trace elements and REE of the mudstone samples (Lee et al., 2012; Li et al., 2013; Wang et al., 2014b; Li et al., 2016; Yan et al., 2016; Chen et al., 2017; Hu et al., 2017; Liu et al., 2017). Fig.10(a) displays that the Cretaceous and Jurassic granodiorites from Amdo and the southern Qiangtang, along with the mudstone samples, show enrichment in Th and U elements relative to the upper crust. In contrast, the volcanic rocks on the west side of the Nima Basin do not show similar enrichment. In addition, the REE concentrations in the volcanic rocks from the west of the Nima Basin are lower than in the mudstone samples (Fig.10(b)), suggesting that the volcanic rocks on the west side of the Nima Basin are unlikely geochemically related to the mudstones. Beyond that, the abundance of limestone clasts in the conglomerates of the Niubao Formation is indicative that the extensively distributed Mesozoic marine strata of the southern Qiangtang and northern Lhasa terranes also contributed an enormous amount of debris to the Niubao Formation (Han et al., 2019; Kong et al., 2019; Liu et al., 2023).
Fig.10 (a) Diagram comparing the upper continental crust-normalized rare elements (REEs) of mudstones from the Nima Basin with those of their source igneous rock. (b) Diagram comparing the chondrite-normalized rare earth element of mudstones from the Nima Basin with those of their source igneous rock (data from Lee et al. (2012): the Cretaceous volcanic rocks in the northern Lhasa. Li et al. (2013): the Cretaceous Abushan volcanic rocks in the southern Qiangtang. Wang et al. (2014b): the Cretaceous andesites and dacites in the northern Lhasa. Li et al. (2016): the Late Jurassic granodiorites in Kangqiong. Yan et al. (2016): the Early-Middle Jurassic Calc-alkaline granodiorite in Amdo. Chen et al. (2017): the Late Cretaceous alkaline andesites in Amdo. Hu et al. (2017): the Early Cretaceous granodiorites in Zigetang Co. Liu et al. (2017): the Early Cretaceous mica granites in Amdo).

Full size|PPT slide

5.5 Palaeogeography of the drainage basin and implications for the landscape evolution

The increasing amount of quantitative paleoaltitude data sets based on various methods suggests a lowland valley with an elevation of 2−3 km during the Eocene, which ran between the highlands of the Gangdese arc (Currie et al., 2005; Ding et al., 2014; Currie et al., 2016) to the south and the central Qiangtang terrane to the north (Zhu et al., 2013, 2015; Wu et al., 2017; Su et al., 2019; Fang et al., 2020; Ding et al., 2022; Li et al., 2022c; Xiong et al., 2022; Bi et al., 2023). However, the west of central Xizang was near sea level during the late Eocene (Wei et al., 2016). Apatite fission track studies suggest that the Amdo area, a few hundred kilometres east of the Nima Basin, underwent rapid cooling in the early Cretaceous times (ca. 110 Ma; Bi et al., 2023; Lu et al., 2024). Low temperature thermochronological data demonstrated that the west of the Nima Basin did not experience rapid cooling until the late Cretaceous (ca. 70 Ma; Xue et al., 2022). On the basis of the differential cooling history, it is possible that the land surface uplift in the eastern portion of central Xizang commenced prior to that in the western portion, thereby establishing an elevation gradient from eastern highlands to western lowlands. The early Cenozoic sedimentary basins along the BNSZ accumulate several kilometres thick of fluvial-lacustrine sediments and show great similarities in sedimentary structural characteristics and sedimentary facies. It could be inferred that these basins may have been interconnected (e.g., Ma et al., 2015; Han et al., 2019). Furthermore, the westward direction of paleocurrents in the Lunpola and Bangoin basins suggests the possibility of a drainage system extending from east to west across the central part of the Xizang Plateau (He et al., 2012; Ma et al., 2015; Han et al., 2019). Hence, a westward drained source-to-sink system likely developed since the Cenozoic. In this geomorphologic pattern, it makes more sense that the mudstones of the Nima Basin have a stronger affinity for the igneous rocks to their east.

6 Conclusions

The geochemical analysis of the late Eocene of the Niubao Formation was undertaken to study the paleoclimate conditions, provenance, and tectonic settings. The main conclusions are summarized as follow.
1) The Niubao Formation mudstones are characterized by high contents of SiO2, Al2O3, and CaO and are primarily composed of clay minerals.
2) The mudstone samples have a total rare earth element (ΣREE) content of 124.81−204.83 ppm. They exhibit a higher proportion of light rare earth elements (LREEs) compared to heavy rare earth elements (HREEs) and display notable negative Eu anomalies.
3) The sediment is first recycled and compositionally immature. The CIA values (59.8−72.9) indicate that the source area of the Niubao Formation experienced moderate chemical weathering.
4) The mudstones were deposited in an oxidizing and arid depositional environment with a mean annual temperature (MAT) of 11.64°C ± 4.19°C, reflecting a cool environmental conditions in central Xizang during the late Eocene.
5) The mudstones were predominantly derived from felsic and intermediate igneous rocks that originated in a continental island arc setting. Igneous rocks from the north, south, and east of the Nima Basin contributed significantly to the composition of these mudstones.
6) Combining data of paleoaltimetry and low temperature thermochronology, the geochemical characteristics of the Niubao Formation suggest the possible development of a westward-draining source-to-sink system since the Cenozoic.

References

[1]
Algeo T J, Maynard J B (2004). Trace-element behavior and redox facies in core shales of Upper Pennsylvanian Kansas-type cyclothems.Chem Geol, 206(3−4): 289–318
CrossRef Google scholar
[2]
Amorosi A, Centineo M C, Dinelli E, Lucchini F, Tateo F (2002). Geochemical and mineralogical variations as indicators of provenance changes in Late Quaternary deposits of SE Po Plain.Sediment Geol, 151(3−4): 273–292
CrossRef Google scholar
[3]
Amorosi A, Sammartino I, Dinelli E, Campo B, Guercia T, Trincardi F, Pellegrini C (2022). Provenance and sediment dispersal in the Po-Adriatic source-to-sink system unraveled by bulk-sediment geochemistry and its linkage to catchment geology.Earth Sci Rev, 234: 104202
CrossRef Google scholar
[4]
Anderson S P, Dietrich W E, Brimhall G H Jr (2002). Weathering profiles, mass-balance analysis, and rates of solute loss: linkages between weathering and erosion in a small, steep catchment.Geol Soc Am Bull, 114(9): 1143–1158
CrossRef Google scholar
[5]
Armitage J J, Duller R A, Whittaker A C, Allen P A (2011). Transformation of tectonic and climatic signals from source to sedimentary archive.Nat Geosci, 4(4): 231–235
CrossRef Google scholar
[6]
Armstrong-Altrin J S, Lee Y I, Kasper-Zubillaga J J, Trejo-Ramírez E (2017). Mineralogy and geochemistry of sands along the Manzanillo and El Carrizal beach areas, southern Mexico: implications for palaeoweathering, provenance and tectonic setting.Geol J, 52(4): 559–582
CrossRef Google scholar
[7]
Armstrong-Altrin J S, Lee Y I, Verma S P, Worden R H (2009). Carbon, oxygen, and strontium isotope geochemistry of carbonate rocks of the upper Miocene Kudankulam Formation, southern India: implications for paleoenvironment and diagenesis.Chem Erde, 69(1): 45–60
CrossRef Google scholar
[8]
Armstrong-Altrin J S, Machain-Castillo M L, Rosales-Hoz L, Carranza-Edwards A, Sanchez-Cabeza J A, Ruíz-Fernández A C (2015a). Provenance and depositional history of continental slope sediments in the Southwestern Gulf of Mexico unraveled by geochemical analysis.Cont Shelf Res, 95: 15–26
CrossRef Google scholar
[9]
Armstrong-Altrin J S, Nagarajan R, Balaram V, Natalhy-Pineda O (2015b). Petrography and geochemistry of sands from the Chachalacas and Veracruz beach areas, western Gulf of Mexico, Mexico: constraints on provenance and tectonic setting.J S Am Earth Sci, 64: 199–216
CrossRef Google scholar
[10]
Armstrong-Altrin J S, Nagarajan R, Madhavaraju J, Rosalez-Hoz L, Lee Y I, Balaram V, Cruz-Martínez A, Avila-Ramírez G (2013). Geochemistry of the Jurassic and Upper Cretaceous shales from the Molango Region, Hidalgo, eastern Mexico: implications for source-area weathering, provenance, and tectonic setting.C R Geosci, 345(4): 185–202
CrossRef Google scholar
[11]
Asiedu D K, Suzuki S, Nogami K, Shibata T (2000). Geochemistry of Lower Cretaceous sediments, inner zone of Southwest Japan: constraints on provenance and tectonic environment.Geochem J, 34(2): 155–173
CrossRef Google scholar
[12]
Ayers J C, Watson E B (1993). Rutile solubility and mobility in supercritical aqueous fluids.Contrib Mineral Petrol, 114(3): 321–330
CrossRef Google scholar
[13]
Ayinla H A, Abdullah W H, Makeen Y M, Abubakar M B, Jauro A, Sarki Yandoka B M, Zainal Abidin N S (2017). Petrographic and geochemical characterization of the Upper Cretaceous coal and mudstones of Gombe Formation, Gongola sub-basin, northern Benue trough Nigeria: implication for organic matter preservation, paleodepositional environment and tectonic settings.Int J Coal Geol, 180: 67–82
CrossRef Google scholar
[14]
Bai Y Y, Liu Z J, Sun P C, Liu R, Hu X F, Zhao H Q, Xu Y B (2015). Rare earth and major element geochemistry of Eocene fine-grained sediments in oil shale-and coal-bearing layers of the Meihe Basin, Northeast China.J Asian Earth Sci, 97: 89–101
CrossRef Google scholar
[15]
Banerji U S, Dubey C P, Goswami V, Joshi K B (2022). Geochemical indicators in provenance estimation. In: Geochemical Treasures and Petrogenetic Processes. Singapore: Springer Nature Singapore, 95–121
[16]
Barshad I (1966). Factors affecting the frequency distribution of clay minerals in soils.Clays and Clay Minerals, 1966: 207
CrossRef Google scholar
[17]
Basu A, Bickford M E, Deasy R (2016). Inferring tectonic provenance of siliciclastic rocks from their chemical compositions: a dissent.Sediment Geol, 336: 26–35
CrossRef Google scholar
[18]
Bhatia M R (1983). Plate tectonics and geochemical composition of sandstones.J Geol, 91(6): 611–627
CrossRef Google scholar
[19]
Bhatia M R, Crook K A (1986). Trace element characteristics of graywackes and tectonic setting discrimination of sedimentary basins.Contrib Mineral Petrol, 92(2): 181–193
CrossRef Google scholar
[20]
Bi W J, Li Y L, Kamp P J J, Xu G Q, Zhang J W, Han Z P, Du L T, Wang C S, He H Y, Xu T K, Ma Z N (2023). Cretaceous–Cenozoic cooling history of the Qiangtang terrane and implications for Central Xizang formation.Geol Soc Am Bull, 135(5−6): 1587–1601
[21]
Bock B, McLennan S M, Hanson G N (1998). Geochemistry and provenance of the middle Ordovician Austin Glen member (Normanskill formation) and the Taconian orogeny in New England.Sedimentology, 45(4): 635–655
CrossRef Google scholar
[22]
Botsyun S, Sepulchre P, Donnadieu Y, Risi C, Licht A, Caves Rugenstein J K (2019). Revised paleoaltimetry data show low Xizang Plateau elevation during the Eocene.Science, 363(6430): eaaq1436
CrossRef Google scholar
[23]
Cao H S, Guo W, Shan X L, Ma L, Sun P C (2015). Paleolimnological environments and organic accumulation of the Nenjiang Formation in the Southeastern Songliao Basin, China.Oil Shale, 32(1): 5–24
CrossRef Google scholar
[24]
Caracciolo L (2020). Sediment generation and sediment routing systems from a quantitative provenance analysis perspective: review, application and future development.Earth Sci Rev, 209: 103226
CrossRef Google scholar
[25]
Cerling T E, Harris J M, MacFadden B J, Leakey M G, Quade J, Eisenmann V, Ehleringer J R (1997). Global vegetation change through the Miocene/Pliocene boundary.Nature, 389(6647): 153–158
CrossRef Google scholar
[26]
Chapman J B, Kapp P (2017). Xizang magmatism database.Geochem Geophys Geosyst, 18(11): 4229–4234
CrossRef Google scholar
[27]
Chen L, Chen X H, Zhang B M, Zhang G T, Li H, Lin W B, Chen P, Hu H Y, Tian W (2020). Provenance and palaeoenvironment of Upper Devonian Shetianqiao Formation mudstones in Shaoyang Sag, Xiangzhong Depression, Central China.Geol J, 55(1): 934–948
CrossRef Google scholar
[28]
Chen S S, Fan W M, Shi R D, Gong X H, Wu K (2017). Removal of deep lithosphere in ancient continental collisional orogens: a case study from central Xizang, China.Geochem Geophys Geosyst, 18(3): 1225–1243
CrossRef Google scholar
[29]
Chen X Q, Xing F C, Jiang S, Lu Y C, Liu Z R, Pan L, Hu H R (2021). Origin and formation model of Eocene dolomite in the upper Niubao Formation of the Lunpola Basin, Xizang Plateau.Interpretation (Tulsa), 9(3): SF11–SF22
CrossRef Google scholar
[30]
Chen Y, Yi H S, Xia G Q, Yi F, Tang W Q, Li G J, Wu X H, Shi Y X (2022). The organic geochemical characteristics from the Palaeogene lacustrine source rock in the Nyima Basin, Central Xizang, and their geological significance.Geol J, 57(3): 1186–1207
CrossRef Google scholar
[31]
Chi Q H (2007). Handbook of Applied Geochemical Element Abundance Data. Beijing: Geological Press
[32]
Condie K C (1993). Chemical composition and evolution of the upper continental crust: contrasting results from surface samples and shales.Chem Geol, 104(1−4): 1–37
CrossRef Google scholar
[33]
Cox R, Lowe D R, Cullers R L (1995). The influence of sediment recycling and basement composition on evolution of mudrock chemistry in the southwestern United States.Geochim Cosmochim Acta, 59(14): 2919–2940
CrossRef Google scholar
[34]
Crichton J G, Condie K C (1993). Trace elements as source indicators in cratonic sediments: a case study from the Early Proterozoic Libby Creek Group, southeastern Wyoming.J Geol, 101(3): 319–332
CrossRef Google scholar
[35]
Cullers R L (2002). Implications of elemental concentrations for provenance, redox conditions, and metamorphic studies of shales and limestones near Pueblo, CO, USA.Chem Geol, 191(4): 305–327
CrossRef Google scholar
[36]
Cullers R L, Podkovyrov V N (2000). Geochemistry of the Mesoproterozoic Lakhanda shales in southeastern Yakutia, Russia: implications for mineralogical and provenance control, and recycling.Precambrian Res, 104(1−2): 77–93
CrossRef Google scholar
[37]
Currie B S, Polissar P J, Rowley D B, Ingalls M, Li S, Olack G, Freeman K H (2016). Multiproxy paleoaltimetry of the late Oligocene-Pliocene Oiyug basin, southern Xizang.Am J Sci, 316(5): 401–436
CrossRef Google scholar
[38]
Currie B S, Rowley D B, Tabor N J (2005). Middle Miocene paleoaltimetry of southern Xizang: implications for the role of mantle thickening and delamination in the Himalayan orogen.Geology, 33(3): 181–184
CrossRef Google scholar
[39]
DeCelles P G, Kapp P, Ding L, Gehrels G E (2007a). Late Cretaceous to middle Tertiary basin evolution in the central Xizang Plateau: changing environments in response to tectonic partitioning, aridification, and regional elevation gain.Geol Soc Am Bull, 119(5−6): 654–680
CrossRef Google scholar
[40]
DeCelles P G, Quade J, Kapp P, Fan M, Dettman D L, Ding L (2007b). High and dry in central Xizang during the Late Oligocene.Earth Planet Sci Lett, 253(3−4): 389–401
CrossRef Google scholar
[41]
Del Rio C, Wang T X, Wu F X, Liang X Q, Spicer T E, Zhou Z K, Su T (2020). Fossil record of Ceratophyllum aff. muricatum Cham. (Ceratophyllaceae) from the middle Eocene of central Xizang Plateau, China.Rev Palaeobot Palynol, 281: 104284
CrossRef Google scholar
[42]
Deng T, Ding L (2015). Paleoaltimetry reconstructions of the Xizang Plateau: progress and contradictions.Natl Sci Rev, 2(4): 417–437
CrossRef Google scholar
[43]
Deng T, Wang S Q, Xie G P, Li Q, Hou S K, Sun B Y (2012). A mammalian fossil from the Dingqing Formation in the Lunpola Basin, northern Xizang, and its relevance to age and paleo-altimetry.Chin Sci Bull, 57(2−3): 261–269
CrossRef Google scholar
[44]
Ding L, Kapp P, Cai F L, Garzione C N, Xiong Z Y, Wang H Q, Wang C (2022). Timing and mechanisms of Xizang Plateau uplift.Nat Rev Earth Environ, 3(10): 652–667
CrossRef Google scholar
[45]
Ding L, Xu Q, Yue Y H, Wang H Q, Cai F L, Li S (2014). The andean-type gangdese mountains: paleoelevation record from the paleocene-eocene Linzhou Basin.Earth Planet Sci Lett, 392: 250–264
CrossRef Google scholar
[46]
Donaldson D G, Webb A A, Menold C A, Kylander-Clark A R, Hacker B R (2013). Petrochronology of Himalayan ultrahigh-pressure eclogite.Geology, 41(8): 835–838
CrossRef Google scholar
[47]
Duddy L R (1980). Redistribution and fractionation of rare-earth and other elements in a weathering profile.Chem Geol, 30(4): 363–381
CrossRef Google scholar
[48]
Dupont-Nivet G, Krijgsman W, Langereis C G, Abels H A, Dai S, Fang X (2007). Xizang plateau aridification linked to global cooling at the Eocene–Oligocene transition.Nature, 445(7128): 635–638
CrossRef Google scholar
[49]
Fang X M, Dupont-Nivet G, Wang C S, Song C H, Meng Q Q, Zhang W L, Nie J S, Zhang T, Mao Z Q, Chen Y (2020). Revised chronology of central Xizang uplift (Lunpola Basin).Sci Adv, 6(50): eaba7298
CrossRef Google scholar
[50]
Fedo C M, Wayne Nesbitt H, Young G M (1995). Unraveling the effects of potassium metasomatism in sedimentary rocks and paleosols, with implications for paleoweathering conditions and provenance.Geology, 23(10): 921–924
CrossRef Google scholar
[51]
Feng W, Meng Q, Song C, Fang X, Zhuang G, He P, Yang S, Zhang J, Chen Y, Zhang Y (2023). Constraints on the timing of the India‐Asia collision and unroofing history of the Himalayan orogen using detrital zircon U‐Pb‐Hf and whole‐rock Sr‐Nd isotopes in Cretaceous‐Miocene Lesser Himalayan sedimentary rocks.Basin Res, 35(3): 949–977
CrossRef Google scholar
[52]
Floyd P A, Leveridge B E (1987). Tectonic environment of the Devonian Gramscatho basin, south Cornwall: framework mode and geochemical evidence from turbiditic sandstones.J Geol Soc London, 144(4): 531–542
CrossRef Google scholar
[53]
Fu X G, Wang J, Chen W B, Feng X L, Wang D, Song C Y, Zeng S Q (2016). Elemental geochemistry of the early Jurassic black shales in the Qiangtang Basin, eastern Tethys: constraints for palaeoenvironment conditions.Geol J, 51(3): 443–454
CrossRef Google scholar
[54]
Fu X G, Wang J, Feng X L, Chen W B, Wang D, Song C, Zeng S Q (2018). Mineralogical compositions and elemental enrichment of shales in lacustrine rift basin: a case study in the Cenozoic Lunpola Basin, central Xizang.Geol J, 53(2): 439–457
CrossRef Google scholar
[55]
Gaillardet J, Dupré B, Allègre C J (1999). Geochemistry of large river suspended sediments: silicate weathering or recycling tracer.Geochim Cosmochim Acta, 63(23−24): 4037–4051
CrossRef Google scholar
[56]
Gallala W, Gaied M E, Montacer M (2009). Detrital mode, mineralogy and geochemistry of the Sidi Aïch Formation (Early Cretaceous) in central and southwestern Tunisia: implications for provenance, tectonic setting and paleoenvironment.J Afr Earth Sci, 53(4−5): 159–170
CrossRef Google scholar
[57]
Garzanti E, Padoan M, Setti M, Najman Y, Peruta L, Villa I M (2013). Weathering geochemistry and Sr‐Nd fingerprints of equatorial upper Nile and Congo muds.Geochem Geophys Geosyst, 14(2): 292–316
CrossRef Google scholar
[58]
Ghosh P, Adkins J, Affek H, Balta B, Guo W, Schauble E A, Schrag D, Eiler J M (2006). 13C–18O bonds in carbonate minerals: a new kind of paleothermometer.Geochim Cosmochim Acta, 70(6): 1439–1456
CrossRef Google scholar
[59]
Ghosh S, Sarkar S (2010). Geochemistry of Permo-Triassic mudstone of the Satpura Gondwana basin, central India: clues for provenance.Chem Geol, 277(1−2): 78–100
CrossRef Google scholar
[60]
Girty G H, Hanson A D, Yoshinobu A S, Knaack C, Johnson D (1993). Provenance of Paleozoic mudstones in a contact metamorphic aureole determined by rare earth element, Th, and Sc analyses, Sierra Nevada, California.Geology, 21(4): 363–366
CrossRef Google scholar
[61]
Gu X X, Liu J M, Zheng M H, Tang J X, Qi L (2002). Provenance and tectonic setting of the Proterozoic turbidites in Hunan, south China: geochemical evidence.J Sediment Res, 72(3): 393–407
CrossRef Google scholar
[62]
Han X, Dai J G, Lin J, Xu S, Liu B, Hu T, Zhang C, Wang C S (2022). An Oligocene-Miocene intermontane narrow lowland in the central Xizang Plateau: insights from provenance analysis and palynological record of a Cenozoic sedimentary succession.J Asian Earth Sci, 240: 105438
CrossRef Google scholar
[63]
Han Z P, Sinclair H D, Li Y L, Wang C S, Tao Z, Qian X Y, Ning Z J, Zhang J W, Wen Y X, Lin J, Zhang B S, Xu M, Dai J, Zhou A, Liang H, Cao S (2019). Internal drainage has sustained low‐relief Xizang landscapes since the early Miocene.Geophys Res Lett, 46(15): 8741–8752
CrossRef Google scholar
[64]
Hayashi K I, Fujisawa H, Holland H D, Ohmoto H (1997). Geochemistry of ∼1.9 Ga sedimentary rocks from northeastern Labrador, Canada.Geochim Cosmochim Acta, 61(19): 4115–4137
CrossRef Google scholar
[65]
He H Y, Sun J M, Li Q L, Zhu R X (2012). New age determination of the Cenozoic Lunpola basin, central Xizang.Geol Mag, 149(1): 141–145
CrossRef Google scholar
[66]
Hu P Y, Zhai Q G, Jahn B M, Wang J, Li C, Chung S L, Lee H Y, Tang S H (2017). Late Early Cretaceous magmatic rocks (118–113 Ma) in the middle segment of the Bangong–Nujiang suture zone, Xizang Plateau: evidence of lithospheric delamination.Gondwana Res, 44: 116–138
CrossRef Google scholar
[67]
Hu Z, Zhang W, Liu Y, Gao S, Li M, Zong K, Chen H, Hu S (2015). “Wave” signal-smoothing and mercury-removing device for laser ablation quadrupole and multiple collector ICPMS analysis: application to lead isotope analysis.Anal Chem, 87(2): 1152–1157
CrossRef Google scholar
[68]
Ingalls M, Rowley D B, Currie B S, Colman A S (2020). Reconsidering the uplift history and peneplanation of the northern Lhasa terrane, Xizang.Am J Sci, 320(6): 479–532
CrossRef Google scholar
[69]
Jia G D, Bai Y, Ma Y J, Sun J M, Peng P A (2015). Paleoelevation of Xizang Lunpola basin in the Oligocene-Miocene transition estimated from leaf wax lipid dual isotopes.Global Planet Change, 126: 14–22
CrossRef Google scholar
[70]
Jia J L, Liu Z J, Bechtel A, Strobl S A, Sun P C (2013). Tectonic and climate control of oil shale deposition in the Upper Cretaceous Qingshankou Formation (Songliao Basin, NE China).Int J Earth Sci, 102(6): 1717–1734
CrossRef Google scholar
[71]
Jia L B, Su T, Huang Y J, Wu F X, Deng T, Zhou Z K (2019). First fossil record of Cedrelospermum (Ulmaceae) from the Qinghai-Xizang Plateau: implications for morphological evolution and biogeography.J Syst Evol, 57(2): 94–104
CrossRef Google scholar
[72]
Jiang H, Su T, Wong W O, Wu F X, Huang J, Shi G (2019). Oligocene Koelreuteria (Sapindaceae) from the Lunpola Basin in central Xizang and its implication for early diversification of the genus.J Asian Earth Sci, 175: 99–108
CrossRef Google scholar
[73]
Jones B, Manning D A (1994). Comparison of geochemical indices used for the interpretation of palaeoredox conditions in ancient mudstones.Chem Geol, 111(1−4): 111–129
CrossRef Google scholar
[74]
Kapp P, DeCelles P G (2019). Mesozoic-Cenozoic geological evolution of the Himalayan-Xizang orogen and working tectonic hypotheses.Am J Sci, 319(3): 159–254
CrossRef Google scholar
[75]
Kapp P, DeCelles P G, Gehrels G E, Heizler M, Ding L (2007). Geological records of the Lhasa-Qiangtang and Indo-Asian collisions in the Nima area of central Xizang.Geol Soc Am Bull, 119(7−8): 917–933
CrossRef Google scholar
[76]
Kasanzu C, Maboko M A, Manya S (2008). Geochemistry of fine-grained clastic sedimentary rocks of the Neoproterozoic Ikorongo Group, NE Tanzania: implications for provenance and source rock weathering.Precambrian Res, 164(3−4): 201–213
CrossRef Google scholar
[77]
Kirubakaran N, Sridharan M, Senthil Nathan D, Rajamanickam M, Harikrishnan S (2023). Geochemical and sedimentological signatures of Ariyankuppam and Chunnambar Estuarine Sediments, Pondicherry, India: implications on weathering and provenance.J Geol Soc India, 99(9): 1275–1284
CrossRef Google scholar
[78]
Kong X, Mi W T, Zhu L D, Yang W G (2019). Zircon U-Pb chronology and provenance of the Paleogene sandstones in the Nima Basin, Xizang: implication for coeval paleogeography.Arab J Geosci, 12: 1–17
CrossRef Google scholar
[79]
Lee H Y, Chung S L, Ji J, Qian Q, Gallet S, Lo C H, Lee T Y, Zhang Q (2012). Geochemical and Sr-Nd isotopic constraints on the genesis of the Cenozoic Linzizong volcanic successions, southern Xizang.J Asian Earth Sci, 53: 96–114
CrossRef Google scholar
[80]
Lerman A (1989). Lakes Chemistry and Geology Physics. Texas: Geological Press
[81]
Lézin C, Andreu B, Pellenard P, Bouchez J L, Emmanuel L, Fauré P, Landrein P (2013). Geochemical disturbance and paleoenvironmental changes during the Early Toarcian in NW Europe.Chem Geol, 341: 1–15
CrossRef Google scholar
[82]
Li C, Zhao Z B, Lu H J, Li H B (2022c). Late Mesozoic-Cenozoic multistage exhumation of the central Bangong-Nujiang suture, central Xizang.Tectonophysics, 827: 229268
CrossRef Google scholar
[83]
Li L, Lu H J, Garzione C, Fan M J (2022b). Cenozoic paleoelevation history of the Lunpola Basin in Central Xizang: new evidence from volcanic glass hydrogen isotopes and a critical review.Earth Sci Rev, 231: 104068
CrossRef Google scholar
[84]
Li X L, Zhang X, Lin C M, Huang S Y, Li X (2022a). Overview of the application and prospect of common chemical weathering indices.Geol J China U, 28(1): 51
[85]
Li Y L, He J, Han Z P, Wang C S, Ma P F, Zhou A, Liu S A, Xu M (2016). Late Jurassic sodium-rich adakitic intrusive rocks in the southern Qiangtang terrane, central Xizang, and their implications for the Bangong–Nujiang Ocean subduction.Lithos, 245: 34–46
CrossRef Google scholar
[86]
Li Y L, He J, Wang C S, Santosh M, Dai J G, Zhang Y X, Wei Y S, Wang J G (2013). Late Cretaceous K-rich magmatism in central Xizang: evidence for early elevation of the Xizang plateau.Lithos, 160–161: 1–3
CrossRef Google scholar
[87]
Li Y L, Wang C S, Dai J G, Xu G Q, Hou Y L, Li X H (2015). Propagation of the deformation and growth of the Xizang-Himalayan orogen: a review.Earth Sci Rev, 143: 36–61
CrossRef Google scholar
[88]
Liang J L, Tang D Z, Xu H, Tao S, Li C C, Gou M F (2014). Formation conditions of Jimusaer oil shale at the northern foot of Bogda Mountain, China.Oil Shale, 31(1): 19
CrossRef Google scholar
[89]
Lin C M, Zhang X, Zhao X P, Li X, Huang S Y, Jiang K X (2021). Review of laboratory research methods for sedimentary petrology.Journal of Palaeogeography (Chinese Ed), 23(2): 223–244
[90]
Lin J, Liu Y, Yang Y, Hu Z (2016). Calibration and correction of LA-ICP-MS and LA-MC-ICP-MS analyses for element contents and isotopic ratios.Solid Earth Sci, 1(1): 5–27
CrossRef Google scholar
[91]
Liu D, Shi R, Ding L, Huang Q, Zhang X, Yue Y, Zhang L (2017). Zircon U–Pb age and Hf isotopic compositions of Mesozoic granitoids in southern Qiangtang, Xizang: implications for the subduction of the Bangong–Nujiang Tethyan Ocean.Gondwana Res, 41: 157–172
CrossRef Google scholar
[92]
Liu J, Su T, Spicer R A, Tang H, Wu F X, Srivastava G, Spicer T, Van Do T, Deng T, Zhou Z K (2019). Biotic interchange through lowlands of Xizang Plateau suture zones during Paleogene.Palaeogeogr Palaeoclimatol Palaeoecol, 524: 33–40
CrossRef Google scholar
[93]
Liu X, Gao R, Guo X, Ding L (2023). Detrital zircon U-Pb geochronology of the Lunpola basin strata constrains the Cenozoic tectonic evolution of central Xizang.Gondwana Res, 113: 179–193
CrossRef Google scholar
[94]
Lu L, Jin X, Yan L, Li W, Wei T, Shen Y (2024). Early uplift and exhumation of the Tanggula granitoid pluton since the Late Cretaceous: implications for the stepwise topographic growth model in the eastern Qiangtang terrane.Geol Soc Am Bull, 147–148: 22
CrossRef Google scholar
[95]
Ma P F, Wang L C, Wang C S, Wu X H, Wei Y S (2015). Organic-matter accumulation of the lacustrine Lunpola oil shale, central Xizang Plateau: controlled by the paleoclimate, provenance, and drainage system.Int J Coal Geol, 147: 58–70
CrossRef Google scholar
[96]
Madhavaraju J (2015). Geochemistry of late Cretaceous sedimentary rocks of the Cauvery Basin, south India: constraints on paleoweathering, provenance, and end Cretaceous environments. In: Ramkumar M, ed. Chemostratigraphy. Netherlands: Elsevier Science, 185–214
[97]
Madukwe H Y, Ayodele S O, Akinyemi S A, Adebayo O F (2016). Classification, maturity, provenance, tectonic setting, and source-area weathering of Ipole and Erin Ijesa stream sediments, south west Nigeria.Intern J Adv Sci Techn Res, 6(1): 232–255
[98]
Mao Z Q, Meng Q Q, Fang X M, Zhang T, Wu F L, Yang Y, Zhang W L, Zan J B, Tan M Q (2019). Recognition of tuffs in the middle-upper Dingqinghu Fm., Lunpola Basin, central Xizang Plateau: constraints on stratigraphic age and implications for paleoclimate. Palaeogeogr Palaeoclimatol Palaeoecol, 525: 44–56
[99]
McLennan S M, Hemming S R, Taylor S R, Eriksson K A (1995). Early Proterozoic crustal evolution: geochemical and Nd-Pb isotopic evidence from metasedimentary rocks, southwestern North America.Geochim Cosmochim Acta, 59(6): 1153–1177
CrossRef Google scholar
[100]
McLennan S M, Hemming S, McDaniel D K, Hanson G N (1993). Geochemical approaches to sedimentation, provenance, and tectonics. Special Papers-Geological Society of America, 21–21
[101]
McLennan S M, Taylor S R, Eriksson K A (1983). Geochemistry of Archean shales from the Pilbara Supergroup, western Australia.Geochim Cosmochim Acta, 47(7): 1211–1222
CrossRef Google scholar
[102]
McLennan S M, Taylor S R, Hemming S R (2006). Composition, Differentiation, and Evolution of Continental Crust: Constraints from Sedimentary Rocks and Heat Flow. Cambridge, UK: Cambridge University Press
[103]
Meng Q T, Liu Z J, Bruch A A, Liu R, Hu F (2012). Palaeoclimatic evolution during Eocene and its influence on oil shale mineralisation, Fushun basin, China.J Asian Earth Sci, 45: 95–105
CrossRef Google scholar
[104]
Mi W T, Yang W G, Zhu L D, Wu C G, Chen A Q (2018). Provenance of paleogene sediments in the south depression of Nima Basin, central Xizang and its geological implications.Geotectonica et Metallogenia, 42(1): 16
[105]
Michalopoulos P, Aller R C (1995). Rapid clay mineral formation in Amazon delta sediments: reverse weathering and oceanic elemental cycles.Science, 270(5236): 614–617
CrossRef Google scholar
[106]
Molnar P, Boos W R, Battisti D S (2010). Orographic controls on climate and paleoclimate of Asia: thermal and mechanical roles for the Xizang Plateau.Annu Rev Earth Planet Sci, 38(1): 77–102
CrossRef Google scholar
[107]
Molnar P, England P, Martinod J (1993). Mantle dynamics, uplift of the Xizang Plateau, and the Indian monsoon.Rev Geophys, 31(4): 357–396
CrossRef Google scholar
[108]
Moradi A V, Sarı A, Akkaya P (2016). Geochemistry of the Miocene oil shale (Hançili Formation) in the Çankırı-Çorum Basin, Central Turkey: implications for paleoclimate conditions, source–area weathering, provenance and tectonic setting.Sediment Geol, 341: 289–303
CrossRef Google scholar
[109]
Nesbitt H W (1979). Mobility and fractionation of rare earth elements during weathering of a granodiorite.Nature, 279(5710): 206–210
CrossRef Google scholar
[110]
Nesbitt H W, Markovics G (1997). Weathering of granodioritic crust, long-term storage of elements in weathering profiles, and petrogenesis of siliciclastic sediments.Geochim Cosmochim Acta, 61(8): 1653–1670
CrossRef Google scholar
[111]
Nesbitt H, Young G M (1982). Early Proterozoic climates and plate motions inferred from major element chemistry of lutites.Nature, 299(5885): 715–717
CrossRef Google scholar
[112]
Potter P E (1978). Petrology and chemistry of modern big river sands.J Geol, 86(4): 423–449
CrossRef Google scholar
[113]
Qiu X W, Liu C Y, Wang F F, Deng Y, Mao G Z (2015). Trace and rare earth element geochemistry of the Upper Triassic mudstones in the southern Ordos Basin, central China.Geol J, 50(4): 399–413
CrossRef Google scholar
[114]
Ramos-Vázquez M A, Armstrong-Altrin J S (2019). Sediment chemistry and detrital zircon record in the Bosque and Paseo del Mar coastal areas from the southwestern Gulf of Mexico.Mar Pet Geol, 110: 650–675
CrossRef Google scholar
[115]
Rasmussen C, Brantley S, Richter D D, Blum A, Dixon J, White A F (2011). Strong climate and tectonic control on plagioclase weathering in granitic terrain.Earth Planet Sci Lett, 301(3−4): 521–530
CrossRef Google scholar
[116]
Raymo M E, Ruddiman W F (1992). Tectonic forcing of late Cenozoic climate.Nature, 359(6391): 117–122
CrossRef Google scholar
[117]
Roddaz M, Viers J, Brusset S, Baby P, Boucayrand C, Hérail G (2006). Controls on weathering and provenance in the Amazonian foreland basin: insights from major and trace element geochemistry of Neogene Amazonian sediments.Chem Geol, 226(1−2): 31–65
CrossRef Google scholar
[118]
Rohrmann A, Kapp P, Carrapa B, Reiners P W, Guynn J, Ding L, Heizler M (2012). Thermochronologic evidence for plateau formation in central Xizang by 45 Ma.Geology, 40(2): 187–190
CrossRef Google scholar
[119]
Roser B P, Cooper R A, Nathan S, Tulloch A J (1996). Reconnaissance sandstone geochemistry, provenance, and tectonic setting of the lower Paleozoic terranes of the West Coast and Nelson, New Zealand.N Z J Geol Geophys, 39(1): 1–16
CrossRef Google scholar
[120]
Roser B P, Korsch R J (1988). Provenance signatures of sandstone-mudstone suites determined using discriminant function analysis of major-element data.Chem Geol, 67(1−2): 119–139
CrossRef Google scholar
[121]
Ross D J, Bustin R M (2009). Investigating the use of sedimentary geochemical proxies for paleoenvironment interpretation of thermally mature organic-rich strata: examples from the Devonian-Mississippian shales, Western Canadian Sedimentary Basin.Chem Geol, 260(1−2): 1–19
CrossRef Google scholar
[122]
Rowley D B, Currie B S (2006). Palaeo-altimetry of the late Eocene to Miocene Lunpola Basin, central Xizang.Nature, 439(7077): 677–681
CrossRef Google scholar
[123]
Rowley D B, Ingalls M, Colman A S, Currie B, Li S, Olack G, Lin D (2015). ~55 Ma aged high topography of the Lhasa Block From Stable and clumped isotope paleoaltimetry: implications for ~50±25% crustal mass deficit in the India-Asia collisional system. In AGU Fall Meeting Abstracts, Vol. 2015, T12B–06
[124]
Roy P D, Caballero M, Lozano R, Smykatz-Kloss W (2008). Geochemistry of late quaternary sediments from Tecocomulco lake, central Mexico: implication to chemical weathering and provenance.Chem Erde, 68(4): 383–393
CrossRef Google scholar
[125]
Shekhar S, Shukla A, Kumar P (2018). Geochemical and petrographic interpretation of Sandhan Formation: an insight into provenance, tectonics and paleoclimatic conditions.Earth Sci India, 11(3): 11
CrossRef Google scholar
[126]
Singh P (2009). Major, trace and REE geochemistry of the Ganga River sediments: influence of provenance and sedimentary processes.Chem Geol, 266(3−4): 242–255
CrossRef Google scholar
[127]
Snell K E, Thrasher B L, Eiler J M, Koch P L, Sloan L C, Tabor N J (2013). Hot summers in the Bighorn Basin during the early Paleogene.Geology, 41(1): 55–58
CrossRef Google scholar
[128]
Spicer R A, Su T, Valdes P J, Farnsworth A, Wu F X, Shi G, Spicer T E, Zhou Z (2021b). The topographic evolution of the Xizang Region as revealed by palaeontology.Palaeobiodivers Palaeoenviron, 101(1): 213–243
CrossRef Google scholar
[129]
Spicer R A, Su T, Valdes P J, Farnsworth A, Wu F X, Shi G, Spicer T E, Zhou Z (2021a). Why ‘the uplift of the Xizang Plateau’ is a myth.Nat Sci Rev, 8(1): 091
[130]
Su T, Farnsworth A, Spicer R A, Huang J, Wu F X, Liu J, Li S F, Xing Y W, Huang Y J, Deng W Y D, Tang H, Xu C L, Zhao F, Srivastava G, Valdes P J, Deng T, Zhou Z K (2019). No high Xizang Plateau until the Neogene.Sci Adv, 5(3): eaav2189
CrossRef Google scholar
[131]
Su T, Spicer R A, Wu F X, Farnsworth A, Huang J, Del Rio C, Deng T, Ding L, Deng W Y D, Huang Y J, Hughes A, Jia L B, Jin J H, Li S F, Liang S Q, Liu J, Liu X Y, Sherlock S, Spicer T, Srivastava G, Tang H, Valdes P, Wang T X, Widdowson M, Wu M X, Xing Y W, Xu C L, Yang J, Zhang C, Zhang S T, Zhang X W, Zhao F, Zhou Z K (2020). A Middle Eocene lowland humid subtropical “Shangri-La” ecosystem in central Xizang.Proc Natl Acad Sci USA, 117(52): 32989–32995
CrossRef Google scholar
[132]
Sun J M, Xu Q H, Liu W M, Zhang Z Q, Xue L, Zhao P (2014). Palynological evidence for the latest Oligocene−early Miocene paleoelevation estimate in the Lunpola Basin, central Xizang.Palaeogeogr Palaeoclimatol Palaeoecol, 399: 21–30
CrossRef Google scholar
[133]
Sun S, Chen A Q, Chen H D, Hou M C, Yang S, Xu S L, Wang F, Huang Z F, Ogg J G (2022). Early Permian chemical weathering indices and paleoclimate transition linked to the end of the coal-forming episode, Ordos Basin, North China Craton.Palaeogeogr Palaeoclimatol Palaeoecol, 585: 110743
CrossRef Google scholar
[134]
Tang H, Liu J, Wu F X, Spicer T, Spicer R A, Deng W Y D, Xu C L, Zhao F, Huang J, Li S F, Su T, Zhou Z K (2019). Extinct genus Lagokarpos reveals a biogeographic connection between Xizang and other regions in the Northern Hemisphere during the Paleogene.J Syst Evol, 57(6): 670–677
CrossRef Google scholar
[135]
Taylor S R, McLennan S M (1985). The Continental Crust: Its Composition and Evolution. Oxford: Blackwell Scientific, 312
[136]
Tribovillard N, Algeo T J, Lyons T, Riboulleau A (2006). Trace metals as paleoredox and paleoproductivity proxies: an update.Chem Geol, 232(1−2): 12–32
CrossRef Google scholar
[137]
Verma S P, Armstrong-Altrin J S (2013). New multi-dimensional diagrams for tectonic discrimination of siliciclastic sediments and their application to Precambrian basins.Chem Geol, 355: 117–133
CrossRef Google scholar
[138]
Verma S P, Armstrong-Altrin J S (2016). Geochemical discrimination of siliciclastic sediments from active and passive margin settings.Sediment Geol, 332: 1–12
CrossRef Google scholar
[139]
Vital H, Stattegger K (2000). Major and trace elements of stream sediments from the lowermost Amazon River.Chem Geol, 168(1−2): 151–168
CrossRef Google scholar
[140]
Wang C S, Dai J G, Zhao X X, Li Y L, Graham S A, He D F, Ran B, Meng J (2014a). Outward-growth of the Xizang Plateau during the Cenozoic: a review.Tectonophysics, 621: 1–43
CrossRef Google scholar
[141]
Wang H, Dutta S, Kelly R S, Rudra A, Li S, Zhang Q Q, Zhang Q Q, Wu Y X, Cao M Z, Wang B, Li J G, Zhang H C (2018b). Amber fossils reveal the Early Cenozoic dipterocarp rainforest in central Xizang.Palaeoworld (Amst), 27(4): 506–513
CrossRef Google scholar
[142]
Wang P, Du Y, Yu W, Algeo T J, Zhou Q, Xu Y, Qi L, Yuan L, Pan W (2020). The chemical index of alteration (CIA) as a proxy for climate change during glacial-interglacial transitions in Earth history.Earth Sci Rev, 201: 103032
CrossRef Google scholar
[143]
Wang Q, Zhu D C, Zhao Z D, Liu S A, Chung S L, Li S M, Liu D, Dai J G, Wang L Q, Mo X X (2014b). Origin of the ca. 90 Ma magnesia-rich volcanic rocks in SE Nyima, central Xizang: products of lithospheric delamination beneath the Lhasa-Qiangtang collision zone.Lithos, 198–199: 24–37
CrossRef Google scholar
[144]
Wang S F, Ch Y, Yi H S, Tang W Q, Zhou Y X, Cui R L, Wu X H, Bai R, Yang Y (2023). The characteristics of n-alkanes from the Palaeogene lacustrine oil shale in the Kanggale area, Nyima Basin, and their paleoenvironment and Paleoclimate significance.Sediment Geol Tethyan Geol, 43(3): 542–554
[145]
Wang Z W, Wang J, Fu X G, Feng X L, Wang D, Song C Y, Chen W B, Zeng S Q (2017a). Petrography and geochemistry of upper Triassic sandstones from the Tumengela Formation in the Woruo Mountain area, North Qiangtang Basin, Xizang: implications for provenance, source area weathering, and tectonic setting.Isl Arc, 26(4): e12191
CrossRef Google scholar
[146]
Wang Z W, Wang J, Fu X G, Zhan W Z, Armstrong-Altrin J S, Yu F, Feng X L, Song C Y, Zeng S Q (2018a). Geochemistry of the Upper Triassic black mudstones in the Qiangtang Basin, Xizang: implications for paleoenvironment, provenance, and tectonic setting.J Asian Earth Sci, 160: 118–135
CrossRef Google scholar
[147]
Wang Z W, Wang J, Fu X G, Zhan W Z, Yu F, Feng X, Song C L, Chen W B, Zeng S Q (2017b). Organic material accumulation of Carnian mudstones in the North Qiangtang Depression, eastern Tethys: controlled by the paleoclimate, paleoenvironment, and provenance.Mar Pet Geol, 88: 440–457
CrossRef Google scholar
[148]
Wei Y, Zhang K X, Garzione C N, Xu Y D, Song B W, Ji J L (2016). Low palaeoelevation of the northern Lhasa terrane during late Eocene: fossil foraminifera and stable isotope evidence from the Gerze Basin.Sci Rep, 6(1): 27508
CrossRef Google scholar
[149]
Weltje G J, von Eynatten H (2004). Quantitative provenance analysis of sediments: review and outlook.Sediment Geol, 171(1−4): 1–11
CrossRef Google scholar
[150]
Wesolowski D J (1992). Aluminum speciation and equilibria in aqueous solution: I. The solubility of gibbsite in the system Na-K-Cl-OH-Al (OH)4 from 0 to 100 °C.Geochim Cosmochim Acta, 56(3): 1065–1091
CrossRef Google scholar
[151]
West A J (2012). Thickness of the chemical weathering zone and implications for erosional and climatic drivers of weathering and for carbon-cycle feedbacks.Geology, 40(9): 811–814
CrossRef Google scholar
[152]
Wronkiewicz D J, Condie K C (1987). Geochemistry of Archean shales from the Witwatersrand Supergroup, South Africa: source-area weathering and provenance.Geochim Cosmochim Acta, 51(9): 2401–2416
CrossRef Google scholar
[153]
Wu F X, Miao D S, Chang M M, Shi G L, Wang N (2017). Fossil climbing perch and associated plant megafossils indicate a warm and wet central Xizang during the late Oligocene.Sci Rep, 7(1): 878
CrossRef Google scholar
[154]
Xia G Q, Zheng D R, Krieg-Jacquier R, Fan Q S, Chen Y, Nel A (2022). The oldest-known Lestidae (Odonata) from the late Eocene of Xizang: palaeoclimatic implications.Geol Mag, 159(4): 511–518
CrossRef Google scholar
[155]
Xiong Z Y, Liu X H, Ding L, Farnsworth A, Spicer R A, Xu Q, Valdes P, He S L, Zeng D, Wang C, Li Z Y, Guo X D, Su T, Zhao C Y, Wang H Q, Yue Y H (2022). The rise and demise of the Paleogene Central Xizang Valley.Sci Adv, 8(6): 0944
[156]
Xu X L, Shao L Y (2018). Limiting factors in utilization of chemical index of alteration of mudstones to quantify the degree of weathering in provenance.J Palaeogeogr (Chn Ed), 020(003): 515–522
[157]
Xu X T, Szwedo J, Huang D Y, Deng W Y D, Obroślak M, Wu F X, Su T (2022). A new genus of spittlebugs (hemiptera, cercopidae) from the eocene of central Xizang Plateau.Insects, 13(9): 770
CrossRef Google scholar
[158]
Xue W W, Najman Y, Hu X M, Persano C, Stuart F M, Li W, Ma A L, Wang Y (2022). Late Cretaceous to Late Eocene exhumation in the Nima area, central Xizang: implications for development of low relief topography of the Xizang Plateau.Tectonics, 41(3): e2021TC006989
CrossRef Google scholar
[159]
Yan H, Long X, Wang X C, Li J, Wang Q, Yuan C, Sun M (2016). Middle Jurassic MORB-type gabbro, high-Mg diorite, calc-alkaline diorite and granodiorite in the Ando area, central Xizang: evidence for a slab roll-back of the Bangong-Nujiang Ocean.Lithos, 264: 315–328
CrossRef Google scholar
[160]
Yang J H, Cawood P A, Du Y S, Li W Q, Yan J X (2016). Reconstructing Early Permian tropical climates from chemical weathering indices.Geol Soc Am Bull, 128(5−6): 739–751
CrossRef Google scholar
[161]
Yang Y B, Nie J S, Miao Y F, Wan S M, Jonell T N (2022). Xizang Plateau uplift and environmental impacts: new progress and perspectives.Front in Earth Sci, 10: 1020354
CrossRef Google scholar
[162]
Zeng S Q, Wang J, Chen W B, Fu X G, Feng X L, Song C Y, Wang D, Sun W (2020). Geochemical constraints on the provenance, paleoenvironment, and tectonic setting of late Triassic mudstones in the western Qiangtang Basin, Xizang.Lithosphere, 12(5): 631–648
[163]
Zhang X, Gélin U, Spicer R A, Wu F, Farnsworth A, Chen P, Del Rio C, Li S, Liu J, Huang J, Spicer T E, Tomlinson K W, Valdes P J, Xu X, Zhang S, Deng T, Zhou Z, Su T (2022). Rapid Eocene diversification of spiny plants in subtropical woodlands of central Xizang.Nat Commun, 13(1): 3787
CrossRef Google scholar
[164]
Zhou L, Friis H, Poulsen M L K (2015). Geochemical evaluation of the late Paleocene and early Eocene shales in Siri Canyon, Danish-Norwegian Basin.Mar Pet Geol, 61: 111–122
CrossRef Google scholar
[165]
Zhu D C, Wang Q, Cawood P A, Zhao Z D, Mo X X (2017). Raising the Gangdese mountains in southern Xizang.J Geophys Res Solid Earth, 122(1): 214–223
CrossRef Google scholar
[166]
Zhu D C, Wang Q, Zhao Z D, Chung S L, Cawood P A, Niu Y L, Liu S A, Wu F Y, Mo X X (2015). Magmatic record of India-Asia collision.Sci Rep, 5(1): 14289
CrossRef Google scholar
[167]
Zhu D C, Zhao Z D, Niu Y, Dilek Y, Hou Z Q, Mo X X (2013). The origin and pre-Cenozoic evolution of the Xizang Plateau.Gondwana Res, 23(4): 1429–1454
CrossRef Google scholar

Acknowledgments

This work was supported by the Second Xizang Plateau Scientific Expedition Program (No. 2019QZKK0803), National Natural Science Foundation of China (Grant Nos. 42172113 and 41802124), Fundamental Research Funds for the Central Universities (No. 265202138), “Deep-time Digital Earth” Science and Technology Leading Talents Team Funds for the Central Universities for the Frontiers Science Center for Deep-time Digital Earth, China University of Geosciences (Beijing) (No. 2652023001), and Key Laboratory of Sedimentary Basin and Oil and Gas Resources, Ministry of Natural Resources of the PRC (No. cdcgs2018006).

Competing interests

The authors declare that they have no competing interests.

RIGHTS & PERMISSIONS

2024 Higher Education Press
AI Summary AI Mindmap
PDF(5061 KB)

255

Accesses

0

Citations

Detail

Sections
Recommended

/