Introduction

While the Andes serve as a neotectonic exemplar of continental arc magmatism and flat-slab (ridge/seamount-induced) subduction that has been useful to understand the geologic past1,2, another potential Andean neotectonic aspect—orographic-precipitation-induced aridification as has occurred in the Atacama desert—remains underexplored in Earth’s past. The initiation and termination of the late Palaeozoic ice age (340–250 Ma) is one of the most prominent climate change events in Earth history and coincided with the assembly of supercontinent Pangaea and global aridification3. Although the fall and rise of atmospheric CO2 has been considered as a leading factor controlling this severe climate change4, recent studies suggest that the uplift of the equatorial Hercynian Mountains and the assembly of Pangaea had potential effects on the tumultuous late Palaeozoic climate change5.

Accompanying Hercynian orogenesis in the core of Pangaea, the long-lived subduction of the Palaeo-Asian Ocean (PAO) gave birth to the world’s largest Palaeozoic accretionary orogen, the Central Asian Orogenic Belt (CAOB; a.k.a. the Altaids) in northeast Pangaea6,7, which is currently situated between the Baltica and Siberia cratons to the north and the Tarim and North China cratons to the south (Fig. 1a). The Xing’an–Inner Mongolia Orogen in the southeastern segment of the CAOB (Fig. 1a, b) was formed through the complex amalgamation between the South Mongolia arc-accretionary complex and the North China craton after the final closure of the PAO8,9.

Fig. 1: Tectonic map of the study area.
figure 1

a Regional tectonic map of the Central Asian Orogenic Belt showing the location of the study area (modified with permission from ref. 87, Elsevier). b Tectonic map of Inner Mongolia showing the major lithologic–tectonic units discussed in this study (modified with permission from ref. 9, Wiley).

One major manifestation of severe climate change during the assembly of northeast Pangaea was the Permian aridification of North China craton10 (hereafter referred to as simply North China). The aridification coincided with the largest extinction of biodiversity, of both regional-scale genera and species, ever documented in North China11. The orogenic processes of the CAOB potentially promoted this climate change as North China was the southern shoreline margin of the subducting PAO during the late Palaeozoic12,13. However, the timing of the closure of the PAO and the orogenic architecture of the southern CAOB during the Permian remains highly controversial14,15,16. Moreover, no quantification of late Palaeozoic surface topography has been obtained to date for the northern margin of North China. These ambiguities in temporal and spatial evolution of orogeny have led to much uncertainty in the palaeogeographic and palaeotopographic reconstruction of northeast Pangaea and the potential feedbacks between orogenic dynamics and climate change. The role of the tectonic boundary conditions in the Permian climatic evolution of North China remains unclear.

Sedimentary basins formed during orogenesis are the earliest and most sensitive archives that record orogenic processes and growth and erosion of topography17. Permian–Triassic sedimentary successions intermittently distributed along the northern margin of North China contain critical information of temporal and spatial variations in magmatic activity, exhumation and palaeoclimate change, and provide excellent recorders for tracing tectonic–climatic feedbacks of the adjacent convergent margin during the assembly of northeast Pangaea.

Here we present an integrated sedimentological and geochronologic study of Permian–Early Triassic sediments in northern North China to evaluate the provenance, basin development and tectonic setting. In addition, we compiled whole-rock La/Yb data for late Carboniferous to Permian magmatic rocks to estimate the crustal thickness and palaeoelevation of the northern margin of North China. The results indicate the development of a Permian–Early Triassic retroarc foreland system with an Andean-type orogenic plateau in northern North China, which revises the current understanding of the late Palaeozoic tectonic evolution of the southeastern CAOB and the mechanism of Permian aridification in North China.

Results

Stratigraphy and sedimentology

Permian to Early Triassic sedimentary rocks crop out in the Daqingshan area of North China. They unconformably overlie Ordovician carbonates or are in fault contact with the Proterozoic metamorphic basement (gneiss-schist complexes) of North China. We studied in detail the Wuyuan and Tumid sections where a continuous Permian to Early Triassic stratigraphy is well exposed (Supplementary Fig. 1a, b). Permian sedimentary rocks in the Dahongshan area, ~30 km north of Wuyuan county, are mapped as the Dahongshan Formation18, which rests unconformably on or is in fault contact with Mesoproterozoic metasedimentary rocks of the Zhaertai Group, and are intruded by Permian granitoids (Supplementary Fig. 1a).

The Dahongshan Formation has a total thickness of ~5000 m and consists of coal-bearing clastic rocks and volcaniclastic rocks in the lower and upper part, respectively18. It is subdivided into five rock members based on lithologic characteristics. The basal member (P1d1) consists of black thick-bedded conglomerate with interbedded pebbly sandstone, carbonaceous slate and limestone lens. The second member (P1d2) contains carbonaceous slate, tuff and pebbly sandstone, with local occurrences of 1–2-m thick coal seams (Supplementary Fig. 2a, b). Rhythmic structures are well developed in the tuff (Supplementary Fig. 2c). The widely-distributed third member (P1d3) mainly consists of conglomerate, sandstone, carbonaceous slate and tuff (Supplementary Fig. 2d–f). The conglomerates are thickly bedded with thickness reaching up to 150 m. They are sandy-matrix-supported and include moderately sorted, fine- to medium-grained (1–6 cm), subrounded pebbles dominated by quartzite and granitoid, with subordinate mudstone, tuff and metamorphic rocks. Several coal seams are present within the sandstone and slate. The fourth member (P1d4) contains primarily black conglomerate interbedded with grey feldspar quartz sandstone (Supplementary Fig. 2g, h) and minor marlstone (Supplementary Fig. 2i). Parallel and trough/planar cross beddings are preserved in the sandstone. The fifth member (P1d5) mainly consists of tuffaceous sandstone, tuff, carbonaceous slate and andesitic porphyrite. The lithofacies associations, including sandy-matrix-supported conglomerate, trough or planar cross-laminated sandstone, carbonaceous slate, marlstone and the development of coal seams, suggest that the Dahongshan Formation was deposited in the braided river delta plain.

Permian to Early Triassic sedimentary strata are well exposed ~10 km north of Tumid Youqi, Inner Mongolia (Supplementary Fig. 1b). These sediments rest unconformably on Cambrian–Ordovician shallow marine quartz sandstone and limestone, or are in fault contact with Archaean–Palaeoproterozoic metamorphic basement, and are unconformably overlain by Jurassic fluvial-lacustrine sediments. They have a total thickness of ~2000 m, and are folded into a major paired syncline–anticline. The Permian strata are divided into four formations (from bottom to top): the early Permian Shuanmazhuang Formation, the middle Permian Zahuaigou Formation and Shiyewan Formation and the late Permian Naobaogou Formation.

The early Permian Shuanmazhuang Formation (P1s) is the earliest nonmarine sedimentation resting unconformably on the Early Ordovician neritic facies sequence of the early Palaeozoic passive margin of North China. It mainly consists of grey conglomerate, pebbly sandstone and black-grey mudstone with the presence of coal seams reaching thicknesses of up to ~30 m (Supplementary Fig. 3a, b). The Shuangmazhuang Formation is folded into an anticline, and the duplex structure indicates northward thrusting (Supplementary Fig. 3b, c). Plant macrofossils of Neuropteris pseudovata and Lepidodendron szeianum were reported from this formation18. The field relationship and lithostratigraphy for the Permian strata that record information of climate change are summarised in Fig. 2. Contrasting lithologies and sedimentary facies occurred before and after the deposition of a thick-bedded conglomerate (Fig. 2a). The middle Permian Zahuaigou Formation (P2z) is composed of grey-white conglomerate, sandstone, argillaceous rock and coal seams (Fig. 2b). The middle Permian Shiyewan Formation (P2sy) contains thick-bedded conglomerate (Fig. 2c) with interbedded sandstone, siltstone and mudstone. The top of the Shiyewan Formation is conformably overlain by the red sandstone of the late Permian Naobaogou Formation (P3n) that has a thickness of ~1100 m (Supplementary Fig. 3d). The Naobgaogou Formation mainly consists of thick-bedded grayish purple polymictic conglomerate, purple medium bedded coarse-grained lithic feldspar sandstone, arkose and red siltstone (Fig. 2d, e and Supplementary Fig. 3e–g) and stands out in marked contrast to the underlying coal-bearing strata. Trough and tabular cross-stratification is well preserved within the sandstone (Supplementary Fig. 3e). The broadly tabular sandstone beds pinch out or thicken along strike, representing channelised sandbodies. Abundant nodular calcareous horizons, marl lens and gypsum crystals are observed within the red mudstone and siltstone. The conglomerate consists of sandy-matrix supported, 1–10 cm sized, pebbles of granites, porphyritic volcanic rocks, sandstones, cherts, schists, gneisses and quartzites (Supplementary Fig. 3e). The occurrence of mesoscale asymmetric folds indicates southward thrusting (Supplementary Fig. 3g). The Early Triassic Laowopu Formation (T1lw) rests conformably on the late Permian Naobaogou Formation (Supplementary Fig. 3h). It has a total thickness of >550 m and represents the core of the major syncline. This formation mainly consists of purple sandstone, pebbly sandstone and conglomerate. Large-scale tabular and trough cross-stratification with foreset heights of up to 60 cm is developed within the sandstones (Supplementary Fig. 3i).

Fig. 2: Stratigraphy and field geology of the Permian sedimentary sequence.
figure 2

a Simplified lithostratigraphic column of the Tumid section showing thick-bedded conglomerate between the coal-bearing sandstone and continental redbeds, indicating the change from humid to arid climatic conditions occurred after a major tectonic uplift event. b Sandstone with interlayered coal seams indicating a humid climate. c Thick-bedded conglomerate. d Continental redbeds indicating an arid climate. e Close-up of red sandstone.

The lithologies and field characteristics of the Permian to Lower Triassic strata show that they are all terrestrial deposits. The early–middle Permian sediments are dominated by multistorey sheet sandstone and conglomerate units with abundant scour surfaces and interbedded coal seams, suggesting a braided fluvial deposit in a coastal plain environment19. The presence of coal and abundant plant fossils implies a humid palaeoclimate during their deposition. The late Permian strata have sedimentary features with fining upwards and channelised nature of large sandbodies, pervasive tabular and trough cross-stratification, and abundant fine-grained siltstone and silty mudstone, which collectively indicate meandering fluvial and overbank deposition in a floodplain environment20. The development of large-scale trough and planar cross-stratification, abundant traction and scour structures, and lack of fine-grained material suggest that the Early Triassic Laowopu Formation was likely a braided fluvial delta deposit. The presence of strong red coloration, calcic palaeosols and gypsum deposits indicate that the late Permian and Early Triassic sedimentation developed in an arid climate under strong oxidising conditions (Fig. 2d, e and Supplementary Fig. 3e, g–i). A remarkable climate change therefore is reflected in the transition from the coal-bearing unit to the continental redbed unit15.

Sampling strategy

Several sandstone samples from both the Wuyuan and Tumid sections were collected for petrologic and detrital zircon U-Pb and Lu-Hf isotopic analyses to investigate the provenances and tectonic settings for the Permian to Lower Triassic sediments (Supplementary Fig. 1a, b). One granite and one tuff sample associated with the sedimentary sequence were also collected for zircon U-Pb dating and Lu-Hf isotopic studies. To reconstruct the orography accompanying the sedimentation, whole-rock geochemical data for latest Carboniferous–Permian igneous rocks were compiled to estimate the crustal thickness and palaeoelevation of the northern margin of North China.

Petrography

The tuff from the Dahongshan Formation consists mainly of microcrystalline quartz (Supplementary Fig. 4a), and the sandstones are composed of monocrystalline and polycrystalline quartz, plagioclase, potassium feldspar, and a variety of lithic fragments including schist, gneiss and granite (Supplementary Fig. 4b, c). Accessory minerals include biotite, hornblende, pyroxene, magnetite and zircon. The grains are mostly angular, suggesting they were eroded directly from the bedrock with a relatively short distance of transport before deposition. Sandstones from the Permian strata of the Tumid section show apparently similar petrographic characteristics being dominated by angular-shaped monocrystalline and polycrystalline quartz, plagioclase, potassium feldspar and lithic fragments. Lithic grain compositions include volcanic rocks, granitic rocks, schist and gneiss, with less siltstone and sandstone (Supplementary Fig. 4d, e). Accessory minerals include biotite, zircon, magnetite and pyroxene. Sandstones from the Early Triassic Laowopu Formation contain grain-supported angular monocrystalline and polycrystalline quartz, potassium feldspar, plagioclase and lithic fragments of volcanic and metamorphic rocks (Supplementary Fig. 4f).

Zircon U-Pb and Lu-Hf isotopes

The zircon U-Pb isotopic results of the 16 analysed samples are presented in Supplementary Data 1. The concordia plots for the two igneous samples and the histograms and kernel density estimation plots (KDE) for detrital zircons of each sedimentary samples are presented in Supplementary Fig. 5. Zircon Lu-Hf isotopic compositions are listed in Supplementary Data 2. The granite (21DHS01) and tuff (21DHS02) samples from the Wuyuan section have well-defined concordant ages yielding weighted mean values of 287 ± 1 Ma (n = 22, MSWD = 0.47) and 290 ± 1 Ma (n = 30, MSWD = 1.9), respectively (Supplementary Fig. 5a, b). They have average zircon εHf(t) values of −6.5 and −19.8, respectively. One hundred spots were analysed for each sandstone sample and a total of 1,400 detrital zircon U-Pb analyses were obtained, of which 1379 analyses yielded concordant ages. All samples of the Dahongshan Formation from the Wuyuan section yield three groups of detrital zircon U-Pb ages: Neoarchaean–early Palaeoproterozoic (2750–2200 Ma), late Palaeoproterozoic (2100–1600 Ma) and late Palaeozoic (400–270 Ma), although the relative proportions of these age populations vary from sample to sample (Supplementary Fig. 5c). For example, samples 21DHS03, 21DHS09 and 21DHS14 have relatively lower proportions of Palaeozoic age populations compared to samples 21DHS07, 21DHS08 and 21DHS16. When the ages of all the samples are combined, they yield three major age peaks at ~2515, 1756 and 297 Ma (Fig. 3a). The eight sandstone samples from the Tumid section also display comparable Neoarchaean–early Palaeoproterozoic (2715-2200 Ma), late Palaeoproterozoic (2100-1600 Ma) and late Palaeozoic (400-260 Ma) detrital zircon U-Pb age spectrum on the KDE plots (Supplementary Fig. 5d). They together yield three prominent age peaks at ~2513, 1845 and 304 Ma (Fig. 3a). A total of 162 spots with late Palaeozoic ages were further analysed for their Lu-Hf isotopic compositions. All the analysed spots yielded negative εHf(t) values from −2.7 to −30.2 except for three analyses with slightly positive εHf(t) values from 0.3 to 1.5 (Fig. 3b).

Fig. 3: Zircon U-Pb age distributions and zircon εHf(t) values.
figure 3

a Comparison of age populations of the detrital zircons from this study and potential source regions. Blue lines are normalised kernel density estimation plots (KDE) and rectangles are age histograms. b Comparison between zircon εHf(t) values of this study, Palaeozoic magmatism of the Solonker suture zone and regions to the north, and late Palaeozoic magmatism of the Inner Mongolia Palaeo-uplift (IMPU). Zircon U-Pb and Hf isotopic data for the Permian–Early Triassic samples are provided in Supplementary Data 1 and Supplementary Data 2, respectively. Compiled zircon U-Pb ages and εHf(t) values from other regions for comparison are provided in Supplementary Data 3.

To determine the provenance of the sedimentary rocks, we compiled 3,769 zircon U-Pb ages and 565 Hf isotopic data from the potential source regions for comparison (Fig. 3a and Supplementary Data 3).

Crustal thickness

Estimates of crustal thickness using whole-rock La/Yb from individual rock analyses range from 41 to 78 km, with individual uncertainties from 9 to 13 km (Supplementary Data 4). Average crustal thickness estimates for areas with ≥3 samples (where available) are labelled in Fig. 1b to provide a representation of data variability. No clear correlation between age, location and calculated crustal thickness was distinguishable in the data set. Therefore, we suggest that the best estimate of crustal thickness during latest Carboniferous–Permian is obtained by considering the data collectively. The crustal thickness estimate results for the entire data set are plotted as a histogram (Fig. 4), which shows that most of the crustal thickness values distribute between 47 and 68 km. The 57 analyses yield a mean value of 58 ± 11 km (Fig. 4), which constrains the average crustal thickness of northern North China during the latest Carboniferous–Permian.

Fig. 4: Crustal thickness of the Permian orogenic plateau of northern North China.
figure 4

Histogram and Gaussian distribution depicting the average crustal thickness estimate (58 ± 11 km) for the Permian orogenic plateau of northern North China. The range of crustal thickness estimated for the Late Cretaceous Nevadaplano (55–65 km, North American Cordillera)47 and for the modern Altiplano (60–70 km, Central Andes of the South American Cordillera)49 are indicated for comparison. Crustal thickness, geochemical and location information for individual magmatic rocks are provided in Supplementary Data 4.

Discussion

Provenance analyses

The overall similar detrital zircon U-Pb age spectrum on the KDE plots for the samples from the Wuyuan and Tumid sections indicates that they have similar provenance signatures (Fig. 3a). The late Palaeozoic zircons are euhedral to subhedral with oscillatory zoning in cathodoluminescence (CL) images (Supplementary Fig. 6), implying they are of first-cycled magmatic origins. The presence of tuffs indicates an important sedimentary supply from a nearby active magmatic source. These are consistent with the petrographic characteristics of these sedimentary rocks that contain different proportions of volcanic fragments. Nearly all the late Palaeozoic zircons in both the Wuyuan and Tumid sections have negative εHf(t) values (Fig. 3b), indicating their derivation from late Palaeozoic magmatic rocks that formed from the reworking of substantial ancient continental crust. Widespread late Palaeozoic magmatic rocks exposed in the southern CAOB and the Inner Mongolia Palaeo-uplift (IMPU) are potential source regions for the late Palaeozoic detritus. Palaeozoic magmatic rocks from the varies tectonic units of the southern CAOB, including the Bainaimiao arc, the Solonker suture zone, the Baolidao arc-accretion complex, and the Hegenshan ophiolite-arc-accretion complex exhibit mainly juvenile crustal signatures of intraoceanic subduction origins with little input from continental fragments7.

The Bainaimiao arc is characterised by magmatic and metamorphic rocks with zircon U-Pb ages between 520–410 Ma, dominated by a Late Ordovician–early Silurian (~445 Ma) peak21,22, and Precambrian detrital zircon ages from Archaean to Neoproterozoic23 (Fig. 3a). However, both early Palaeozoic and Mesoproterozoic–Neoproterozoic age populations are absent in the samples from the Wuyuan and Tumid sections (Fig. 3a). We therefore exclude the Bainaimiao arc as the source region for the Permian–Early Triassic sediments in this study. The Solonker suture zone is mainly composed of ophiolitic mélange, island arc tholeiite, mid-ocean-ridge basalt, tonalite and sanukitoid, with high εNd(t) values from −0.7 to 8.4 and zircon εHf(t) values from 14.7 to 19.1, which together record the evolution of a Permian (~299–260 Ma) intraoceanic arc-trench system within the PAO24,25. The Baolidao arc-accretion complex contains ~484–469 Ma Early Ordovician juvenile arc rocks with εNd(t) values from 1.5 to 5.3 and initial 87Sr/86Sr ratios from 0.7043 to 0.7066 (ref. 26). Amphibolites from the Xilingol complex yield zircon U-Pb ages of 382–327 Ma with εHf(t) values between −1.0 and 4.1 (ref. 27). Further north, the Hegenshan ophiolite arc-accretion complex contains late Palaeozoic ophiolitic mélanges and volcanic-plutonic complexes. Various Permian intrusions including adakites, alkaline granites, and A-type granites with zircon U-Pb ages of ~292–275 Ma and εHf(t) from 4.9 to 20.3 have been reported in the Erenhot–Hegenshan belt of the central-northern Inner Mongolia28,29,30. Therefore, the late Palaeozoic rocks from the southern CAOB have mainly positive zircon εHf(t) values, which are different from the late Palaeozoic detrital zircons in the Permian–Early Triassic sediments with mainly negative εHf(t) values (Fig. 3b). Thus, the various terranes north of the Bayan Obo–Chifeng fault also deemed as unlikely to have provided sedimentary sources for the Permian–Early Triassic strata in northern North China.

In contrast, the IMPU contains large-scale, late Palaeozoic magmatic rocks with negative whole-rock εNd(t) and zircon εHf(t) values. For example, granitoids with U-Pb ages of 324–274 Ma, εNd(t) of −17.4 to −9.3 and εHf(t) of −16.5 to 1.2 were reported from the eastern segment of the IMPU31,32. Late Carboniferous–early Permian mafic-ultramafic complexes are characterised by low initial 87Sr/86Sr ratios of 0.70521–0.70604, negative εNd(t) values from −14.1 to −9.3, and negative zircon εHf(t) values from −17.0 to −10.5, and were probably derived from metasomatized lithospheric mantle33. In the western IMPU, early Permian (286–279 Ma) mafic to felsic intrusions with εNd(t) of −13.9 to −4.67 and zircon εHf(t) from −11.5 to −6.7 were reported in the Guyang area, north of Baotou34. Early Permian (291 Ma) granitic plutons with low εNd(t) (−15.1) and high initial 87Sr/86Sr (0.7073) were reported in the Wulatezhongqi region north of Wuyuan35. In addition, along the Jining–Chicheng fault minor Devonian magmatism was reported, including an Early Devonian alkaline intrusive complex (whole-rock εNd(t) values of −12.7 to −17.9 and zircon εHf(t) values from −27.8 to −32.3; ref. 36), a Middle Devonian mafic complex33, and A-type granites (~387 Ma; εHf(t) = −11.5 to −8.4; TDM2 = 2820–2548 Ma; ref. 37). The εHf(t) values of the Palaeozoic zircons from the Permian–Early Triassic sedimentary successions are within the range defined by the Palaeozoic magmatism of the IMPU, but are distinct from those of the Solonker suture zone and the regions north of it (Fig. 3b). Therefore, both the age spectrum and εHf(t) values of the late Palaeozoic detrital zircons match well with those of the Palaeozoic magmatic rocks within the IMPU (Fig. 3a, b). Based on this synthesis, we suggest that the late Palaeozoic magmatic rocks within the IMPU were the sources for the late Palaeozoic detrital zircons in the Permian–Early Triassic sediments of this study.

The Neoarchaean–Palaeoproterozoic age populations of the sampled strata, with major peaks at ~2515 Ma and ~1845–1756 Ma, are comparable to age peaks of the Precambrian basement of North China (Fig. 3a), which generally yield zircon U-Pb ages between 3600 and 1600 Ma, with major peaks at 2800–2400 and 2100–1700 Ma (ref. 38). Archaean (2571–2441 Ma) TTG (tonalite-trondhjemite-granodiorite) gneisses and Palaeoproterozoic (1962–1812 Ma) metasedimentary rocks were reported in the Daqingshan area of the IMPU15. Voluminous Neoarchaean magmatic and metamorphic rocks dated at 2594–2502 Ma have been reported in the Guyang area in the western IMPU39,40. Although the Alxa Block also contains Neoarchaean–Palaeoproterozoic magmatic-metamorphic basement, the typical Mesoproterozoic–Neoproterozoic (1500–800 Ma) age signatures41 of the Alxa Block were not detected in this study, thus, we rule out the Alxa Block as a source region for the Permian–Early Triassic sediments. The Mesoproterozoic–Neoproterozoic age gap also excludes the southern Mongolian and Bainaimiao arcs as source regions as both of them contain numerous ages between 1200–750 Ma (refs. 23,38). The Neoarchaean–Palaeoproterozoic zircons for the Permian–Early Triassic sediments were therefore derived from North China (including the IMPU). In summary, zircon U-Pb and Hf isotopic data reveal that the sources of the Permian–Early Triassic sediments were the IMPU and North China craton, without any input from the juvenile materials from the southern CAOB north of the Bayan Obo–Chifeng fault.

Permian–Early Triassic retroarc foreland system in northern North China

The tectonic setting of the Permian–Early Triassic sedimentation in northern North China is critical for understanding the orogenic architecture of the southeast CAOB and the palaeogeography of northeast Pangaea. Sedimentological and provenance characteristics in this study support their deposition in a retroarc foreland basin. The Permian Dahongshan Formation in the Wuyuan area and the Permian–Early Triassic sediments in the Tumid area are thick, nonmarine siliciclastic successions. They unconformably overlie, or are in fault contact with, Palaeoproterozoic basement or early Palaeozoic passive margin littoral carbonate of the North China craton. The development of fold-and-thrust structures indicates a compressional tectonic environment for the Permian sedimentation (Supplementary Fig. 3b, c, g). The conglomerate of the Permian–Triassic sediments contains pebbles of granite, porphyrite, gneiss and quartzite (Supplementary Fig. 3f), and the sandstones contain variable amounts of quartz, feldspar and lithic fragments of granite, volcanic rocks, gneiss, schist and quartzite. The detrital zircon U-Pb age and Hf isotopic data further demonstrate that the IMPU and North China were the source regions. This dual provenance characterised by active magmatism juxtaposed with compressional deformation of the stable craton resulting in continental uplift is consistent with many well-studied retroarc foreland basins such as those developed in the Andes and throughout the Cordillera42,43.

The IMPU has been demonstrated to be a late Palaeozoic Andean-type continental arc on the northern margin of North China that was initiated by the southward subduction of the PAO in the late Carboniferous31,32. The main components of the continental arc are calc-alkaline or high-K calc-alkaline, metaluminous Permo–Carboniferous quartz diorite, diorite, granodiorite, tonalite, hornblende gabbro, I-type granites and adakites, with minor intermediate-acid volcanic lavas and tuffs32. The rock associations as well as geochemical and Sr-Nd-Hf isotopic data of the Permo–Carboniferous magmatism are consistent with an Andean-type continental arc32,44. The build-up of the magmatic arc generated a topographic load on the northern margin of North China, the subsidence of which accommodated the deposition of the early Permian sediments.

Available data indicate that the Permian–Early Triassic retroarc foreland system might have developed along the entire northern margin of North China. A Permian retroarc foreland basin has recently been recognised in the Langshan area of the northern margin of the Alxa Block to the west of the study area (Fig. 1a). Thrust-imbricated pyroclastic materials eroded from the late Carboniferous–Permian continental arc and sediments eroded from the Precambrian basement of the Alxa Block were deposited in the retroarc foreland basin45. Sedimentological and sandstone petrographic studies of the marginal marine to nonmarine alluvial Pyeongan Supergroup in Korea suggests that it was deposited in a retroarc foreland basin along the northern margin of North China between 320–250 Ma (ref. 46). These apparently isolated late Carboniferous to Early Triassic sedimentary successions might have once connected along-strike to form a continuous, >2000-km-long, east-west trending retroarc foreland system along the northern margin of North China ascribed to southward subduction of the PAO.

Permian Andean-type orogenic plateau in northern North China and its palaeoclimate implications

Before deposition of the early Permian Dahongshan and Shuanmazhuang Formations, the northern margin of North China was covered by Cambrian–Ordovician shallow marine deposits dominated by quartz sandstone and carbonate rocks, representing a north-facing passive continental margin. Silurian to early Carboniferous sedimentation is generally lacking across the northern margin of North China due to limited weathering of the then low-relief cratonic margin. Subduction of the PAO beneath North China since late Carboniferous generated a continental magmatic arc, namely, the IMPU32,44. The ~290 Ma Dahongshan and Shuanmazhuang Formations contain thick-bedded, coarse-grained conglomerates (Supplementary Fig. 2d, e) that were probably deposited in a proximal alluvial fan. This sedimentary change over time indicates that a progressive increase in relief of the nearby source areas was enough to generate debris flows during intense precipitation events in the early Permian. Detrital zircon U-Pb ages and Hf isotopic data indicate that the dominant sources for the early Permian sediments were derived from the IMPU. Therefore, the conglomerates of the Dahongshan and Shuanmazhuang Formations documented a phase of significant topographic growth and erosion of the IMPU by ~290 Ma. The uplift of the continental arc is also manifest in the abrupt increase in the proportions of late Palaeozoic detrital zircons in samples 21DHS07, 21DHS08 and 21DHS16 from the Dahongshan Formation (Supplementary Fig. 5c), which indicates accelerated erosion of the magmatic arc by ~290 Ma.

Our estimate of ~58 km for the average crustal thickness of northern North China during the latest Carboniferous–Permian is significantly greater than the average thickness of global continental crust (37 km), suggesting the existence of a high-elevation orogenic plateau (Fig. 4). This crustal thickness is comparable to that of the Late Cretaceous to early Palaeogene central United States Cordillera (Nevadaplano, 55–65 km; ref. 47) and the southern United States–northern Mexico Cordillera (57 km; ref. 48), and is only slightly lower than the average crustal thickness (60–70 km) beneath the highest elevations in the Altiplano of the Central Andes49. Assuming Airy isostatic compensation with an average crustal density of 2.77 g cm−3, an average mantle density of 3.27 g cm−3, and an average continental crust thickness of 37 km, our data indicate that the IMPU would have attained a palaeoelevation of 3.8 ± 0.7 km during the Permian. This palaeoelevation of the northern North China continental magmatic arc in northeast Pangaea is strikingly comparable to that of the late Carboniferous Variscan Belt (3.4 ± 0.7 km) of Western Europe within the core of Pangaea50. Hornblende thermobarometry on late Carboniferous granitic plutons also indicates that large-scale uplift and exhumation of the IMPU occurred from the late Carboniferous to Early Jurassic and at least 15 km of crust was eroded during this period51. A more complete picture of palaeotopography during the assembly of Pangaea is thus emerging, with implications for palaeoclimatic conditions across the supercontinent.

The shift in sedimentary facies documented here from early Permian coal-bearing paralic and alluvial fan deposits to late Permian–Early Triassic continental redbeds reveals a dramatic climate change in North China (Fig. 2). The transition from a humid to an arid climate of North China during the Permian has also been documented by several other geological and palaeoclimate records. The identification of abundant plant macrofossil assemblages with diverse Cathaysian flora, along with massive coal deposits, in the Taiyuan and Shansi Formations in central Shaanxi Province, suggests that North China had a rather humid climate52. A clear palaeoclimate drying tendency from the late Artinskian–early Kungurian onwards is indicated by the increase in xerophytic plants and decrease in coal deposits in the upper Shansi Formation and Lower Shihhotse Formation in the Ordos Basin53. In Shanxi Province, the Sunjiagou Formation, containing aeolian sandstone, gypsum and carbonate breccias with few plant fossils, rests unconformably on the Upper Shihhotse Formation palaeosols with abundant flora, which indicates a prominent shift from a subhumid environment to a more arid condition across the unconformity10. The gradual upward diminishing of black mudstone, along with the decrease of chemical index of alteration (CIA) values and total organic carbon (TOC) content over time, indicate long-term drying of North China from the early to middle Permian54.

Previous studies simply attributed the Permian aridification of North China to continental drift through subtropical horse latitudes or temperature arid climate zone55,56, without considering the possible impact of high topography created by accretionary/collisional orogenesis on the northern margin of North China. However, the sparse palaeomagnetic database for the Permian palaeolatitudes of North China suggests that it remained stable at ~5°N from the earliest Permian (~300 Ma) through late Permian (~260 Ma), and only eventually drifted northward into the 15–35˚N evaporite belt in the Triassic (~240 Ma; ref. 57). Therefore, available palaeomagnetic data do not support the idea that the northward drift of North China was responsible for the changes in sedimentary facies and climate described here as the tectonic transit critically post-dates the Permian climatic transition.

Alternatively, the Permian aridification was suggested to have resulted from an orographic rain shadow effect due to tectonic uplift associated with the collision between the Altaid arcs and North China15, although no detailed information on the scale and height of this topography was mentioned. This interpretation was based on the premise that the arc-continent collision between the Altaids and North China was underway during the Permian. However, this assumption is not supported by regional geological and provenance data. Firstly, all the samples in this study, irrespective of their depositional ages, display similar detrital zircon U-Pb age populations and Hf isotopic compositions, indicating that they have similar provenance characteristics. This observation is inconsistent with a Permian collision between the Altaid arcs and North China, in which a prominent provenance change would have occurred in the foreland sediments in response to arc-continent collision. The lack of early Palaeozoic and Mesoproterozoic–Neoproterozoic detrital zircon U-Pb ages implies that the terranes north of the Bayan Obo–Chifeng fault contributed no detritus to the Permian–Early Triassic sediments, thus indicating that the southern CAOB was probably separated from North China by an ocean (the PAO) during Permian–Early Triassic (Fig. 5). Such an interpretation is consistent with the provenance of late Palaeozoic to Mesozoic strata from the Xishan area near Beijing that shows detritus from the juvenile crust of the Xing’an-Inner Mongolia orogenic belt had not reached North China until the Late Triassic58. Secondly, the youngest blueschist facies metabasic rocks in Inner Mongolia have protolith ages of 239–235 Ma, indicating ongoing oceanic plate subduction until at least Middle Triassic59. Further west, the youngest sandstone matrix of the Kanguer accretionary mélange in the East Tianshan containing tectonic blocks of normal mid-ocean ridge basalts (NMORB) has a maximum depositional age of ~234 Ma (ref. 60), indicating that the PAO was still being subducted in the Triassic. Detrital zircon provenance data from the Muzitekexie fore-arc accretionary basin, integrated with the compilation of regional multi-disciplinary evidence, suggest that oceanic subduction was ongoing until the Early Triassic in the South Tianshan61. The provenance analyses of Triassic retroarc sediments in the foreland of the Alxa Block suggests that the PAO was finally closed during Middle-Late Triassic (240–232 Ma; ref. 14). Thirdly, Permian sedimentation in the Nomgon and Bulgan Uul basins in southern Mongolia62 and in the Yin’e Basin in northern Alxa Block in China63 is characterised by shallow marine turbidite fan deposition. Contrasting sedimentary facies existed on different sides of the Solonker suture zone until the latest Permian. Our provenance analysis in this study is consistent with regional geological data that suggest the PAO was continuously subducted beneath the northern margin of North China throughout the Permian–Early Triassic.

Fig. 5: Andean-like aridification model for Permian North China.
figure 5

a Global palaeogeographic reconstruction (~280 Ma) showing the tectonic position of North China, with modelled moisture transport directions indicated (adapted with permission from ref. 64, Elsevier). b Close-up view showing the northwest-southeast trending mountain ranges (adapted with permission from ref. 70, Annual Reviews) along the northern North China and the northeast wind. c The development of an Andean-type orogenic plateau with a palaeoelevation of ~3, 800 m blocked the moisture transported from the Palaeo-Asian Ocean and played an important role in the Permian aridification of North China. Topography is vertically exaggerated.

The Hercynian orogenesis resulted from the closure of the Rheic Ocean and subsequent collision between the Laurentia and Gondwanaland led to the assembly of the supercontinent Pangaea. Global climate studies based on sedimentary facies and palaeontological data indicate that the climate of Pangaea as a whole transitioned from icehouse to hothouse during the middle Permian55. The formation of the supercontinent modified the source of moisture advection from the ocean to the continents and resulted in the development of large arid areas in the west and central Pangaea64. The global aridification of Pangaea began in middle Pennsylvanian (~315 Ma) as represented by the change of terrestrial tropical wetland vegetation. The change did not encompass the entire Pangaea tropics, extending through Euramerica but not into China, where middle Pennsylvanian-type vegetation persisted well into the Permian in wetland environments65. Therefore, the Permian aridification of North China occurred well after that of the western Pangaea10. Moreover, the provenance data in this study and most palaeogeographic reconstructions based on palaeomagnetic data support that the North China was separated from the main Pangaea by the Palaeo-Tethys Ocean and Palaeo-Asian Ocean during the Permian57,66. The Permian aridification of North China was therefore unlikely induced by the uplift of the Hercynian Mountains.

An increase in atmospheric CO2 caused by Permian large igneous province (LIP) volcanism and associated Southern Hemisphere deglaciation has been considered the major driving force for the Permian drying in North China, presumably due to elevated surface temperature and evaporation that reduced soil moisture and the source of continental convective precipitation10,54. The long term increase in atmospheric CO2 may have led to long-term aridification in equatorial Pangaea as indicated by numeric climate model results64. However, on shorter timescales, the stratigraphic record may not wholly support this relationship; for example, continental drying in western equatorial Pangaea predates the rise in atmospheric CO2 (ref. 67). The lithologic, stratigraphic and field relationship evidence in this study indicates that the Permian aridification in North China might have occurred after a major tectonic uplift event as represented by the deposition of the thick-bedded conglomerate in between the opposite palaeoclimatic conditions (Fig. 2). A similar situation was also reported in the Permian stratigraphy in Shanxi Province where increased aridity occurred after the formation of a major unconformity10. These geological records demonstrate that the Permian aridification cannot be interpreted alone by the increase of atmospheric CO2. The role of topography growth during accretionary orogenesis must be taken into account when discussing the mechanism for the Permian aridification of North China.

Regional uplifts due to oceanic subduction and accretionary orogenesis along active continental margins, such as the Andes and elsewhere in the Cordillera, have been shown unequivocally to cause increasing aridity in the leeward slope68,69. Our provenance analyses combined with previous studies indicate that the northern margin of North China was an Andean-type active continental margin that was initiated in the late Carboniferous and culminated in the Permian, ascribed to the southward subduction of the PAO. The ~58 km thickness of continental crust and ~3,800 m palaeoelevation estimated here for northern North China covered at least the area between ~40°59´N and ~42°48´N latitude and ~106°19´E and ~119°42´E longitude (present coordinates; Fig. 1b). Palaeogeographic reconstructions have showed that North China trended northwest-southeast during the Permian57,70,71. The prevailing wind direction in North China was probably northeast64 as it was located at the lower latitude of the Northern Hemisphere, which was nearly perpendicular to the orientation of the mountain ranges along the northern margin of North China. All the evidence suggests that the orogenic plateau would have been high and big enough to act as a physiographic barrier to prevent moisture transported from the north-northeast64 reaching the foreland (Fig. 5). This would result in decreasing precipitation and increasing aridity in the leeward side. High-precision zircon U-Pb dating on tuffaceous rocks indicates that the aridification of North China was underway by the early Permian10. Based on our synthesis of available geological data, we propose that the Permian aridification of North China was dominantly the result of an orographic rain shadow effect due to the uplift of the northern active continental margin of North China (Fig. 5), similar to the role of the Andes in forming the Atacama Desert as well as other arid regions of the Cordillera.

Methods

Zircon U-Pb and Hf isotopic analyses

Zircon grains were separated by crushing, heavy-liquid and magnetic methods, and then were randomly (regardless of their grain size, shape, colour and degree of rounding) mounted in epoxy resin and polished to expose the interior. Cathodoluminescence (CL) imaging was conducted with a scanning electron microscope at the Institute of Geology and Geophysics, Chinese Academy of Sciences, to select suitable grains and optimal target sites (avoiding cracks and inclusions) for U-Pb dating and Lu-Hf isotopic analyses.

Zircon U-Pb dating was conducted using LA-ICPMS in Nanjing Hongchuang Geological Exploration Technology Service Co., Ltd. The Resolution SE model laser ablation system (Applied Spectra, USA) was equipped with ATL (ATLEX 300) excimer laser and a Two Volume S155 ablation cell. The laser ablation system was coupled to an Agilent 7900 ICPMS (Agilent, USA). Detailed tuning parameters can be found in Thompson et al.72. Pre-ablation was conducted for each spot analysis using 5 laser shots (~0.3 μm in depth) to remove potential surface contamination. The analysis was performed using 30 μm diametre spot at 5 Hz and a fluence of 2 J cm−2. Age calculations were conducted using IsoplotR (ref. 73). Zircon 91500 and GJ-1 was used as primary and secondary reference materials respectively. 91500 was analysed twice and GJ-1 analysed once every 10–12 analysis of the sample. Typically, 35-40 seconds of the sample signals were acquired after 20 seconds gas background measurement. NIST 610 and 91Zr were used to calibrate the trace element concentrations as external reference material and internal standard element respectively. During the seven batches of zircon U-Pb analysis, 151 spots on Zircon 91500 and 78 spots on Zircon GJ-1 were analysed respectively. The weighted mean 206Pb/238U ages of the Zircon 91500 reference materials for each batch are 1062.6 ± 3.4 Ma (n = 25, MSWD = 0.79), 1062.6 ± 2.5 Ma (n = 21, MSWD = 0.20), 1062.7 ± 2.7 Ma (n = 26, MSWD = 0.95), 1062.5 ± 3.5 Ma (n = 26, MSWD = 0.84), 1062.1 ± 2.6 Ma (n = 29, MSWD = 0.78), 1063.1 ± 5.5 Ma (n = 10, MSWD = 0.73) and 1062.5 ± 4.3 Ma (n = 13, MSWD = 0.19), respectively. The measured ages of the Zircon GJ-1 reference materials for each batch are 606.1 ± 2.6 Ma (n = 13, MSWD = 0.42), 608.8 ± 1.8 Ma (n = 11, MSWD = 1.2), 603.3 ± 2.0 Ma (n = 12, MSWD = 0.61), 605.7 ± 2.7 Ma (n = 13, MSWD = 0.45), 602.3 ± 1.9 Ma (n = 15, MSWD = 0.42), 604.0 ± 4.3 Ma (n = 5, MSWD = 1.3) and 606.2 ± 3.0 Ma (n = 8, MSWD = 0.82), respectively. These measured ages all agreed well with the reference value (1062.4 ± 0.4 Ma for 91500 and 600.7 ± 1.1 Ma for GJ-1, ref. 74) within uncertainty. A discordance filter of 20% was used for the U-Pb ages following ref. 75. The detrital zircon U-Pb ages are plotted as histograms and KDEs using DensityPlotter76. In this study, 206Pb/238U ages are quoted for ages younger than 1000 Ma; otherwise, the 207Pb/206Pb ages are presented.

Hf isotope measurements were performed using a Thermo Finnigan Neptune-plus MC-ICP-MS fitted with a J-100 femto-second laser ablation system. Analyses of standard zircons 91500 and Plešovice over the measurement period provide weighted mean 176Hf/177Hf values of 0.282297 ± 0.000021 (2σ) and 0.282480 ± 0.000015 (2σ), respectively, consistent with the recommended 176Hf/177Hf values of 0.282307 ± 0.000031 (2σ; ref. 77) and 0.282482 ± 0.000013 (2σ; ref. 78), respectively. The isotopic ratio of 176Hf/177Hf was normalised to 179Hf/177Hf = 0.7325 (ref. 79) for mass bias correction. 180Hf/177Hf ratio was calculated to monitor the accuracy of the date and yielded an average value of 1.88688 (MSWD = 1.5, n = 266) for the analytical session (standards and unknowns), which is within the range of values reported by ref. 80. The decay constant for 176Lu and the chondritic ratios of 176Hf/177Hf and 176Lu/177Hf used in calculations are 1.865 × 10−11year−1 (ref. 81), and 0.282772 and 0.0332 (ref. 82), respectively. Initial 176Hf/177Hf ratios and εHf(t) values are calculated by zircon crystallisation ages.

Crustal thickness estimation

Sr/Y and La/Yb are commonly used in igneous petrology to infer depths of magmatic diversification due to difference in the partition coefficient between these elements in various residual phases and intermediate melt83,84. Profeta et al.85 showed that whole-rock Sr/Y and La/Yb can be used as crustal thickness proxies in low magnesium intermediate calc-alkaline igneous rocks covering the compositional range of andesites and dacites and their intrusive equivalents. The empirical formula between whole-rock La/Yb and Moho depth for modern continental arcs85 was applied to estimate the crustal thickness of northern North China. The lower degree of scatter in global and regional correlations for La/Yb makes it a more reliable indicator of thicker crust in continental arcs compared to that of Sr/Y85. A compositional filter of SiO2 = 55–70 wt%, MgO = 1–4 wt% and Rb/Sr = 0.05−0.25 (ref. 48,85) was applied in this study, in order to select suitable rock samples for crustal thickness estimation. This data filtering method ensures that only intermediate-composition rocks were used in the calculation, while mafic rocks generated in the mantle and high silica rocks formed by melting of middle-upper continental crust or from highly fractionated melts were eliminated85. We compiled only unmineralized and minimally altered samples in this study. The filtered data set consisted of 57 whole-rock geochemical analyses that came from locations between 40°59′N and 42°48′N latitude and 106°19′E and 119°42′E longitude (present coordinates). The compiled samples ranged in crystallisation ages from 315 to 252 Ma (peak at ~280 Ma). Sample information, geochemical data and individual crustal thickness estimate are presented in Supplementary Data 4.

Bin size selection of histograms of occurrence data was guided by the crustal thickness data (n = 57; Fig. 4). According to Sturge’s rule86, the number of class intervals (bins), K, can be expressed as:

$$K=1+3.322\,{{{\log }}}_{n}$$
(1)

where n is the number of observations in the set. With 57 estimates of crustal thickness, then K = 6.8.