Next Article in Journal
Genesis of Fe-Ti-(V) Oxide-Rich Rocks by Open-System Evolution of Mafic Alkaline Magmas: The Case of the Ponte Nova Massif, SE Brazil
Previous Article in Journal
Bismuth White (Bismuth Oxychloride) and Its Use in Portrait Miniatures Painted by George Engleheart
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Petrogenesis of the Early Jurassic–Early Cretaceous Adakite-like Rocks in the Erguna Block, NE China: Implications for the Tectonic Evolution of the Mongol–Okhotsk Ocean

1
Research Center of Applied Geology, China Geological Survey, Chengdu 610036, China
2
College of Geography and Planning, Chengdu University of Technology, Chengdu 610059, China
3
MRL Key Laboratory of Metallogeny and Mineral Assessment, Institute of Mineral Resources, Chinese Academy of Geological Sciences, Beijing 100037, China
4
Institute of Geology, Chinese Academy of Geological Sciences, Beijing 100037, China
*
Authors to whom correspondence should be addressed.
Minerals 2024, 14(7), 725; https://doi.org/10.3390/min14070725
Submission received: 17 June 2024 / Revised: 15 July 2024 / Accepted: 16 July 2024 / Published: 19 July 2024

Abstract

:
The petrogenesis and geodynamic setting of the Mesozoic magmatic rocks in the Erguna Block, NE China remains controversial, especially the relationship between magmatism and the subduction history of the Mongol–Okhotsk oceanic plate. Here we present data for the Early Jurassic–Early Cretaceous adakite-like magmatic rocks from Chaoman Farm in the northeastern part of the Erguna Block. Zircon U-Pb dating reveals that the syenogranites crystallized at around 190–180 Ma, while the monzonites, quartz diorite porphyries, and quartz monzonite porphyries were emplaced at around 147–143 Ma. The syenogranites, monzonites, quartz diorite porphyries, and quartz monzonite porphyries are adakite-like rocks. The syenogranites and quartz monzonite porphyries were produced by the partial melting of a thickened ancient mafic lower continental crust and a thickened juvenile lower crust, respectively. Meanwhile, the monzonites and quartz diorite porphyries were formed as a result of partial melting of the oceanic crust. In conclusion, the occurrence of these Early Jurassic magmatic rocks was closely linked to the process of southward subduction of the Mongol–Okhotsk oceanic plate. On the contrary, the Late Jurassic to early Early Cretaceous magmatism (147–143 Ma) occurred in an extensional environment, and was probably triggered by upwelling of the asthenosphere.

1. Introduction

Early Mesozoic granitoids and late Mesozoic volcanic rocks are widespread in the Erguna Block and its surrounding regions (Figure 1) [1,2,3,4,5,6,7,8,9]. These rocks are linked with the southward subduction of the Mongol–Okhotsk Ocean, which had a strong influence on the Mesozoic tectonic evolution and geodynamic process of the Great Xing’an Range [4,5,10,11]. The southward subduction of the Mongol–Okhotsk oceanic plate not only triggered intense magmatic activity, but also facilitated the formation of porphyry deposits during the Early Mesozoic [4,5]. Therefore, understanding the petrogenesis and geodynamics of Mesozoic magmatism can offer crucial insights into the formation and evolutionary history of the Erguna Block.
Up to now, the petrogenesis and the geodynamic setting of the Mesozoic magmatic rocks in the Erguna Block and the related tectonic transition from the Mongol–Okhotsk Ocean subduction to the continent collision have been intensely debated [2,4,8,9,12,13]. The Mesozoic magmatic suite consists of mafic to intermediate–felsic rocks [2,4,5,8,12]. The mafic rocks have been debated to derive from partial melting of depleted mantle modified by subduction-related fluids [5], or upwelling of the asthenosphere in the oceanic ridge [14]. In contrast, the intermediate–felsic rocks, including granitic and syenitic rocks, are closely related to the Cu–Mo deposits [1,2,4,5,8,9,13,14,15,16,17], and their petrogenesis remains controversial. For example, their sources have been speculated as the depleted and heterogeneous lower crust [5], the oceanic slab melts [8,9,14,18], the thickened mafic lower crust [4,8], the delaminated mafic lower crust [4], or mixing and mingling of lower crustal and upper mantle materials in the crust–mantle transition zone [5,19]. Moreover, the final closure time of the Mongol–Okhotsk Ocean [20,21,22] is controversial, ranging from the Middle Jurassic [2,5,7,11,23], the Late Jurassic–Early Cretaceous [8,9,13,14,24,25,26,27], to the Middle Cretaceous [14,18].
The Erguna Block is situated southeast of the Mongol–Okhotsk suture zone in Chinese territory, and is an ideal locus for elucidating the history of the southward subduction of the Mongol–Okhotsk oceanic plate. In this paper, we present a combined study of zircon U–Pb dating and Hf isotopes, whole-rock major and trace elemental compositions, and Sr–Nd–Pb isotope compositions for newly documented Early Jurassic–Early Cretaceous magmatic rocks from the Chaoman Forest Farm polymetallic exploration area in the northeastern Erguna Block. By integrating available geochemical data, we examine the source-region characteristics and petrogenesis of these rocks, proposing a petrogenetic model linked to different components of the lower crust in the northeastern Erguna Block. The results not only clarify the timing and petrogenesis of the Mesozoic magmatism in the Erguna Block, but also enhance our understanding of the subduction processes of the Mongol–Okhotsk oceanic plate in NE China.
Figure 1. (a) Simplified geological sketch map of the CAOB (the Central Asian Orogenic Belt) showing the main tectonic subdivisions [28]. (b) Tectonic sketch map of NE China [29].
Figure 1. (a) Simplified geological sketch map of the CAOB (the Central Asian Orogenic Belt) showing the main tectonic subdivisions [28]. (b) Tectonic sketch map of NE China [29].
Minerals 14 00725 g001

2. Geologic Setting and Petrography

The Great Xing’an Range is situated in the northeast of the Central Asian Orogenic Belt, and consists of the Erguna, Xing’an, and Songnen blocks from west to east (Figure 1) [30,31]. The Erguna Block is bounded by the Mongol–Okhotsk suture zone to the north and the Xinlin–Xiguitu suture zone to the east (Figure 1a) [32]. It is generally accepted that the tectonic evolution history of the Erguna Block was dominated by the subduction of the Paleo-Asian oceanic slab during the Paleozoic and Mongol–Okhotsk oceanic slab during the Mesozoic [3,4,5,6,7,12]. Around 500 Ma, the Erguna Block collided with the Xing’an Block along the Xinlin–Xiguitu suture zone [33,34,35] and later amalgamated with the central Mongolia blocks during the early Paleozoic [36,37,38]. The basement of the Erguna Block consists of the Xinghuadukou, Ergunahe, Luomahu, and Jiageda Groups [39], which include Precambrian metamorphic supracrustal rocks and sporadic Paleoproterozoic and Neoproterozoic intrusions (Figure 1b) [40,41,42]. These are overlain by Paleozoic marine sedimentary rocks, as well as Mesozoic volcanic rocks and terrestrial clastic rocks [12,43].
The Erguna Block is characterized by extensive Paleozoic and Mesozoic intrusive and Mesozoic volcanic rocks resulting from multistage oceanic plate subduction and continent–continent collision during the Paleozoic–Mesozoic (Figure 1b) [2,3,4,5,7,13,44]. Mesozoic magmatic rocks are the most widespread throughout the Erguna Block [1,4,8,9,13,33,34]. The Mesozoic intrusive rocks in the Erguna Block comprise various lithologies and are dominated by calc-alkaline I-type granitoids that are accompanied by subordinate intermediate–mafic rocks generated during the Triassic–Early Cretaceous (250–110 Ma) [3,4,5,7,8,9,13]. The Late Jurassic to Early Cretaceous magmatism is characterized by voluminous volcanic and subvolcanic rocks in the Erguna Block [4]. This complex Mesozoic magmatism displays a close temporal, spatial, and genetic relation to the formation of porphyry Cu–Mo deposits and hydrothermal Ag and Pb–Zn deposits in the Erguna Block [45], such as the Badaguan Cu–Mo deposit [46], Taipingchuan Cu–Mo deposit [47], Wunugetushan Cu–Mo deposit [48], Fukeshan Cu–Mo deposit [8], Erentaolegai Ag deposit [49], and Jiawula Ag–Pb-Zn [50].
The Chaoman Forest Farm polymetallic exploration area is situated at the northeastern Erguna Block in the northern Great Xing’an Range, ~50 km southeast of Mohe country (Figure 2). The study area is well covered by dense primordial forest with poor Mesozoic magmatic rock outcrops along the roads (Figure 2), so we took 35 core samples from drillholes ZK0401 and ZK21001 in the study area. The monzonite and quartz diorite porphyry samples were collected from drillhole ZK0401 at a depth of 505 m; syenogranite and quartz monzonite porphyry samples were from drillhole ZK21001 at a depth of 712 m (Figure 2). Rock types, mineral assemblages, crystallization ages, and the main geochemical features are summarized in Table 1 and representative photographs of the rocks are shown in Figure 3.
The syenogranite is coarse-grained and contains 20–30 vol.% quartz, 15–20 vol.% plagioclase, 30–35 vol.% K-feldspar, and 1–3 vol.% biotite, along with accessory magnetite, zirconm and apatite (Figure 3a,b). The monzonite is massive in structure and contains alkali feldspar (33–35 vol.%), plagioclase (14–21 vol.%), amphibole (20–27 vol.%), biotite (12–18 vol.%), quartz (~3 vol.%), and accessory zircon and apatite (4–5 vol.%; Figure 3c). The quartz diorite porphyry shows porphyritic textures (Figure 3d,e). Phenocrysts of medium- to fine-grained amphibole make up 20 vol.% of the rock and plagioclase phenocrysts make up 15 vol.%; there are a few biotite phenocrysts (Figure 3d,e). Secondary minerals are sericite, quartz, magnetite, chalcopyrite, and pyrite. The quartz monzonite porphyry is porphyritic in texture, and has a microgranular groundmass. Phenocrysts in the quartz monzonite porphyry are mainly medium- to fine-grained stumpy plagioclases and biotite (Figure 3f).

3. Sample Preparation and Analytical Methods

3.1. Zircon U–Pb Dating and Hf Isotopic Analyses

The zircon grains were separated from crushed samples using conventional heavy liquid and magnetic techniques at the Langfang Regional Geological Survey (Langfang, China). Cathodoluminescence (CL) images of analyzed zircon grains were acquired using a FEI NOVA NanoSEM 450 scanning electron microscope outfitted with a Matan Mono CL4 cathodoluminescence system at the Institute of Geology, Chinese Academy of Geological Sciences (CAGS), Beijing, China. The CL image conditions include 2 min scanning time, 15 kV accelerating voltage, and 120 nA beam current. The analysis spots (free from cracks or inclusions) were selected after transmitted light optical microscopy and CL images. The zircon U–Pb analysis used a Neptune laser ablation multiple accepting inductively coupled plasma mass spectrometer (LA–MC-ICP-MS) at the Institute of Mineral Resources, CAGS, Beijing, China. The zircon GJ-1 and the NIST SRM610 were used as external standards for the measurements of the U–Pb isotope ratios and the trace elements, respectively. The spot diameter of the laser ablation pits was 30 μm. Detailed operating conditions for instruments and analytical procedures are the same as described by Hou et al., (2009) [51]. The ICPMSDataCal [52] and Isoplot [53] programs were used for data reduction. Common Pb was corrected following the method outlined by Andersen (2002) [54]. Errors on pooled ages are quoted at the 95% (2σ) confidence level.
Zircon Hf isotopic analyses were performed on the same zircon grains as the U–Pb isotope analyses stated above. In situ zircon Hf isotopic analyses were also performed at the Institute of Mineral Resources, CAGS, Beijing, China, using a Neptune MC-LA-ICP-MS with a 193 nm excimer ArF laser ablation system (GeoLas Plus). All zircon grains were analyzed using a single-spot ablation mode with a spot size of 44 μm and a laser repetition rate of 8 Hz. The GJ-1 zircon standard was used for external standardization. The weighted mean 176Hf/177Hf isotopic ratio of the GJ-1 zircon grains was 0.282007 ± 0.000011 (2σ, n = 36). Details of the analytical methods are given by Hou et al., (2007) [55]. The measured 176Lu/177Hf ratios and the 176Lu decay constant of 1.867 × 10−11 yr−1 [56] were used to calculate initial 176Hf/177Hf ratios. The chondritic values of 176Lu/177Hf = 0.0336 and 176Hf/177Hf = 0.282785 reported by Bouvier et al., (2008) were used for the calculation of the εHf values [57]. The depleted mantle Hf model ages (TDM) were calculated using the measured 176Lu/177Hf ratios based on the assumption that the depleted mantle reservoir has a linear isotopic growth from 176Hf/177Hf = 0.279718 at 4.55 Ma to 0.283250 at present, with 176Lu/177Hf = 0.0384 [58]. Two-stage model ages (TDM2) were calculated by assuming that the parental magma was produced from an average continental crust (176Lu/177Hf = 0.015) [59].

3.2. Whole-Rock Major and Trace Elemental Analyses

We selected 18 fresh samples for whole-rock geochemical analyses after petrographic examination and removal of the altered/weathered surface. All samples were crushed and powdered to ∼200 mesh in an agate mill. Chemical analyses were performed at the Nuclear Industry Beijing Institute of Geology Research and Testing Research Center. Major element compositions were determined by X-ray fluorescence (MEXRF26) using fused glass disks, with analytical precision and accuracy better than 5%. The trace element concentrations were determined using ICP–MS, which gives a precision better than 10% for most of the elements analyzed.

3.3. Whole-Rock Sr–Nd–Pb Isotopic Analyses

Whole-rock Sr–Nd–Pb isotopic composition was undertaken with a Finnigan MAT-262 mass spectrometer at the Laboratory for Radiogenic Isotope Geochemistry, University of Science and Technology of China, Hefei. For Nd–Sr isotope analyses, Rb–Sr and light rare-earth elements were isolated on quartz columns by conventional ion exchange chromatography with a 5 mL resin bed of AG 50W-X12 (200–400 mesh). Nd and Sm were separated from other rare-earth elements on quartz columns using 1.7 mL Teflon powder as the cation exchange medium. For isotopic measurements, Sr was loaded with a Ta–HF activator on pre-conditioned W-filaments and was measured in single-filament mode. Nd was loaded on Re-filaments and measurements were performed in a double-filament configuration. Measured Sr and Nd isotopic ratios were normalized for mass fractionation using 86Sr/88Sr = 0.1194 and 146Nd/144Nd = 0.7219, respectively. For Pb isotope determinations, 200 mg of the sample material was dissolved in HF–HNO3 acid at 120 °C for 7 days using Teflon vials. Pb was separated by anion exchange chromatography with diluted HBr acid as the elutant. The precision of 87Rb/86Sr and 147Sm/144Nd ratios is better than 0.5%. The precision of the measured Pb isotopic ratios is better than 0.01%. Analytical precisions are stated as 2σ standard errors. Further details on analytical techniques are given in [60,61].

4. Results

Zircon U–Pb geochronology, major and trace element compositions of the samples, zircon Hf isotope data, and whole-rock Sr–Nd–Pb isotope data are listed in Tables S1–S4, respectively.

4.1. Zircon U–Pb Geochronology

Zircon grains from 12 representative samples were chosen for LA–ICP-MS U–Pb dating. The zircon U–Pb isotopic data are given in Table S1. All the zircons from these intermediate–felsic rocks display well-developed oscillatory growth zoning in the CL images (Figure 4), and have Th/U ratios of 0.23–3.34 (mostly between 0.4 and 1.0) (Table S1), indicative of a magmatic origin [62]. They are euhedral with a long prismatic shape, with average crystal lengths of 80–120 μm and length-to-width ratios from 2:1 to 3:1 (Figure 4). Zircons from the diabase are 80–100 μm long with aspect ratios of 1:1 to 3:1, and show patchy or banded zoning (Figure 4). The U–Pb concordia diagrams are shown in Figure 5.
Sixteen U–Pb spots on 16 zircon grains from sample ZK21001-46, 10 U–Pb spots on 10 zircon grains from sample ZK21001-53 and 18 U–Pb spots on 18 zircon grains from sample ZK21001-59 of the syenogranite yielded weighted mean 206Pb/238U ages of 181 ± 1.5 Ma (MSWD = 1.08), 190 ± 1.2 Ma (MSWD = 0.0328) and 189 ± 0.94 Ma (MSWD = 0.2), respectively.
Eight U–Pb spots on eight zircon grains from sample ZK0401-10-1, 16 U–Pb spots on 16 zircon grains from the sample ZK0401-13-2 and eight U–Pb spots on eight zircon grains from sample ZK0401-17 of the monzonite yielded weighted mean 206Pb/238U ages of 147 ± 2.7 Ma (MSWD = 1.90), 147 ± 1.2 Ma (MSWD = 0.63) and 147 ± 1.50 Ma (MSWD = 0.86), respectively.
For quartz diorite porphyry sample ZK0401-3, the 206Pb/238U ages from 20 analyses ranged from 144 to 143 Ma (Table S1), yielding a weighted mean 206Pb/238U age of 143 ± 0.9 Ma (MSWD = 0.10; Figure 5g), which represents the crystallization age of the quartz diorite porphyry. For quartz monzonite porphyry sample ZK21001-53, the 206Pb/238U ages from five analyses ranged from 140 to 150 Ma (Table S1), yielding a weighted mean 206Pb/238U age of 143 ± 3.7 Ma (MSWD = 1.3) (Figure 5h), which represents the crystallization age of the quartz diorite porphyry.
In summary, these zircon U–Pb ages indicate that the magmatic rocks from the Chaoman Forest Farm polymetallic exploration area were emplaced in two stages, during the Early Jurassic, and the Late Jurassic–early Early Cretaceous, similar to magmatic rocks (~144–193 Ma) associated with the Fukeshan deposit near the Chaoman Forest Farm (Figure 2) [8].

4.2. Whole-Rock Major and Trace Elemental Compositions

Whole-rock major and trace elemental compositions of the magmatic rocks from the Chaoman Forest Farm are listed in Table S2. On a total alkalis (Na2O + K2O) versus SiO2 (TAS) diagram (Figure 6a) [63], most samples are classified as monzogabbro, monzodiorite, monzonite, quartz monzonite, or with some plotting as granodiorite or granite (Figure 6a). According to the K2O versus SiO2 (Figure 6b) [64] diagrams, the intermediate–felsic rocks belong to the high-K calc-alkaline series. In addition, the syenogranites are peraluminous rocks with A/CNK [molar Al2O3/(CaO + K2O + Na2O)] ratios of 1.0–1.2 (Figure 6c and Table S2) [65], while the monzonites, quartz diorite porphyries and quartz monzonite porphyries are metaluminous with A/CNK ratios of 0.8–1.0 (Figure 6c). The intermediate–felsic rocks are transitional I-type granite to A-type granite on the (Na2O + K2O) versus 10,000 Ga/Al diagrams (Figure 6d). In addition, syenogranites, monzonites, quartz diorite porphyries, and quartz monzonite porphyries have relatively low heavy rare-earth element (HREE) (Yb = 0.57–1.84 ppm) and Y concentrations (9.40–19.0 ppm), as well as high Sr contents (345–1168 ppm) and Sr/Y ratios (27.3–111). Therefore, these rocks have an adakitic character according to the definition of Defant and Drummond (1990, 1993) [66,67], which is also revealed by the use of the Sr/Y versus Y and (La/Yb)N versus (Yb)N discrimination diagrams (Figure 7). In the PRIMA-normalized spidergram (Figure 8a), monzonite and quartz diorite porphyry show strong enrichments in large ion lithophile elements (LILEs) and pronounced negative anomalies in Nb, Ta, and Ti. Furthermore, they display slightly negative Eu anomalies and have inclined chondrite-normalized rare-earth element (REE) patterns (Figure 8b). The syenogranite and quartz monzonite porphyry samples have low total rare trace elements (REEs) with slightly negative Eu anomalies, and show slight LREE enrichment (Figure 8c). On the primitive mantle-normalized trace element spidergram (Figure 8d), the samples exhibit notable negative Nb, Ta, and Ti, along with positive Sr anomalies.

4.3. In Situ Zircon Lu–Hf Isotopes

The Table S3 lists the in situ Lu–Hf isotopic data for the representative zircon grains from the nine samples.
A total of 20 Hf isotopic analyses on the zircon grains from two samples of the syenogranite yielded initial 176Hf/177Hf ratios ranging from 0.282517 to 0.282692 and gave εHf(t) values of –5.3 to 0.9 and TDMC Hf crustal model ages of 1.8–1.3 Ga (Figure 9a). Zircon grains from the three samples of monzonite yielded initial 176Hf/177Hf ratios ranging from 0.282687 to 0.282799 with consistent εHf(t) values of –0.2 to 3.7 (n = 30) and TDMC Hf model ages of 1.4 to 1.09 Ga (Figure 9a). Ten Hf isotopic spot analyses on the zircon grains from the quartz diorite porphyry sample yielded initial 176Hf/177Hf ratios ranging from 0.282764 to 0.282842 with corresponding εHf(t) values of 2.3 to 5.1 and TDMC Hf model ages of 1199–984 Ma (Figure 9a). Four Hf isotopic spot analyses on the zircon grains from sample of the quartz monzonite porphyry yielded a narrow range of εHf(t) values of 0.7 to 3.5 and TDMC Hf model ages of 1326 to 1107 Ma (Figure 9a).

4.4. Whole-Rock Sr–Nd–Pb Isotopes

Rb–Sr, Sm–Nd, and Pb isotopic data of the magmatic rocks from the Chaoman Forest Farm are given in Table S4. All rocks show a limited variation in their radiogenic Sr, Nd, and Pb isotope composition when calculated based on the crystallization ages presented above.
The syenogranites have high (87Sr/86Sr)i (0.7063–0.7086) and low εNd(t) (−1.7 to −8.5) (Figure 9b) and show slightly lower radiogenic Pb isotopic compositions, with (206Pb/204Pb)i = 18.292–18.467, (207Pb/204Pb)i = 15.622–15.642 and (208Pb/204Pb)i = 38.388–38.532 (Figure 10). The monzonites, quartz diorite porphyries, and quartz monzonite porphyries have low (87Sr/86Sr)i (0.7053–0.7066) and high εNd(t) (−3.5 to 0.5) (Figure 9b) and show slightly less radiogenic Pb isotopic compositions, with (206Pb/204Pb)i = 18.406–18.486, (207Pb/204Pb)i = 15.587–15.624 and (208Pb/204Pb)i = 38.344–38.482, respectively (Figure 10). The initial 207Pb/204Pb and 208Pb/204Pb ratios plot above the Northern Hemisphere Reference Line (NHRL) (Figure 10) [79].

5. Discussion

5.1. The Episodic Magmatism in the Erguna Block during the Mesozoic

Our age data of this study, together with literature zircon U–Pb age data, reveal that the Mesozoic magmatism in the Erguna Block is dominated by granitoids and rhyolites with minor mafic rocks and can be subdivided into five episodes: Early-Middle Triassic (249–237 Ma), Late Triassic (229–201 Ma), Early-Middle Jurassic (199–171 Ma), Late Jurassic (155–145 Ma), and Early Cretaceous (145–119 Ma) [1,3,4,5,7,8,9,12,13,82]. The Early-Middle Triassic (249–237 Ma) intrusive rocks form a suite of gabbro–diorites, diorites, quartz diorites, granodiorites, monzogranites, and syenogranites [1,3]. The Late Triassic (229–201 Ma) intrusive rocks consist of a suite of granodiorites, syenogranites, and gabbro–diorites, and are widespread in the Erguna Block [2,5,83,84]. The Early-Middle Jurassic (199–171 Ma) intrusive rocks in the Erguna Block consist of monzogranites, syenogranites, quartz porphyries, and granite porphyries [1,2,83,85,86] and show close relationships with porphyry Cu–Mo deposits, such as the Taipingchuan [74] and Wunugetushan porphyry Cu–Mo deposit [19,75]. The Late Jurassic to Early Cretaceous intrusive rocks contain syenogranites, monzonites and quartz monzonites, diorite porphyries, quartz monzonite porphyries and granodiorite porphyries, and have close relationships with Cu–Mo mineralization, such as Fukeshan [8] and Xiaokele [9]. The Early Cretaceous (145–119 Ma) magmatic rocks are characterized by large-scale volcanic rocks, and minor granitoids in the Erguna Block [4,7,12,23].
In this study, we have obtained new zircon U–Pb ages from magmatic rocks in the northeastern Erguna Block, which include ~190 Ma and ~136 Ma to ~140 Ma. The magmatic rocks from the Chaoman Forest Farm were produced during two periods of activity: (1) Early Jurassic (190–180 Ma), when abundant granitoids were emplaced, including syenogranites; and (2) Late Jurassic to early Early Cretaceous (147–143 Ma), when the monzonite, quartz diorite porphyry and quartz monzonite porphyry were emplaced.
According to the definition of adakite [66,87], adakite-like rocks are those with SiO2 ≥ 56 wt. %, Al2O3 ≥ 15 wt. %, Sr/Y ≥ 20, Y ≤ 18 ppm and Yb ≤ 1.9 ppm. In this study, the Early Jurassic syenogranites and Late Jurassic to early Early Cretaceous monzonites, quartz diorite porphyries, and quartz monzonite porphyries belong to the adakite-like group.

5.2. Origin of the Adakite-like Rocks

As previously described, the Early Jurassic syenogranites and Late Jurassic to early Early Cretaceous monzonite, quartz diorite porphyry and quartz monzonite porphyry have high SiO2 contents (up to 70.37 wt.%), Sr (345–1168 ppm) and Sr/Y (27–111) ratios, low Y (9.40–19.0 ppm) and Yb (0.57–1.84 ppm) concentrations and slightly negative Eu anomalies (Figure 7 and Figure 8) (Table S2), typical of adakites [66]. As initially defined by Defant and Drummond (1990), adakites originated from the partial melting of a young subducted oceanic crust exhibiting distinctive geochemical signatures, including high Sr and low Y and Yb contents, resulting in high Sr/Y ratios [66]. In addition, later studies have proposed different models to interpret the genesis of adakitic rocks, including (1) magma mixing between felsic and basaltic magmas [88], (2) partial melting of thickened or underplated lower crust [89,90,91] or delaminated continental crust [92], and (3) fractional crystallization (FC) or crustal assimilation and fractional crystallization (AFC) from basaltic parent magmas [87,93,94].
We propose that the studied Chaoman Forest Farm Late Jurassic–early Early Cretaceous monzonites and quartz diorite porphyries probably formed by partial melting of subducting oceanic crust based on the following reasons: (1) The Chaoman Forest Farm monzonites and quartz diorite porphyries have relatively low K2O/Na2O (0.43–0.97) and high CaO/Al2O3 (0.21–0.36) values, which are similar to those observed in adakitic rocks formed from partial melting of subducted oceanic crust (with K2O/Na2O < 0.71 and CaO/Al2O3 > 0.2) [95], just like the Late Jurassic slab-derived adakites from the Fukeshan deposit in its neighboring areas (Figure 6) [8]. Furthermore, the monzonites and quartz diorite porphyries mostly plot in the field of adakitic rocks derived from melting of subducting oceanic crust in the Ni, Cr, Mg# and MgO discrimination diagrams (Figure 11a–e). The high MgO and Mg# values may indicate the interaction of a slab-derived adakitic melt and a mantle-derived magma, which is also supported by the presence of the contemporaneous mafic rocks [96,97].
The SiO2 versus Mg# diagram (Figure 11b) illustrates that mantle AFC modelling [102] suggests that the adakitic melts probably mixed with a minor fraction (<10%) of mantle-derived magma [103]; (2) The HREE and Y contents, and the Sr/Y, (La/Yb)N and (Dy/Yb)N ratios of the monzonites and quartz diorite porphyries are generally consistent with those of the Late Jurassic adakites derived from subducted oceanic crust in the Erguna Block. Their comparatively low (La/Yb)N ratios could also be attributed to the interaction with a mantle-derived melt. Furthermore, their low Th and Th/La values are comparable to those of typical MORB and the Cretaceous slab-derived adakites in the southern Lhasa terrane (Figure 11f), indicating that their formations are closely associated with the subducted basaltic oceanic crust (Figure 11f) [104,105]; and (3) The monzonites and quartz diorite porphyries display relatively low (87Sr/86Sr)i (0.7062–0.7066) and slightly enriched to depleted εHf(t) values (–0.2 to 5.1) and εNd(t) values (–3.49 to –1.13), similar to the Early Jurassic oceanic slab-derived adakitic rocks associated with the Taipingchuan porphyry Cu–Mo deposit, as well as the Late Jurassic to Early Cretaceous slab-derived adakitic rocks linked to the Fukeshan porphyry Cu–Mo deposit (Figure 9) [8,74]. Furthermore, most of the εHf(t) values plot between the “new continental crust” evolutionary line and chondrite line [78] (Figure 9a), implying a significant involvement of newly formed juvenile crustal material. The weakly negative εHf(t) and εNd(t) values indicate that the monzonites and quartz diorite porphyries were derived from an oceanic slab with minor contamination of old continental crustal materials [8,74].
Notably, the monzonites show higher Sr/Y and (La/Yb)N ratios than the quartz diorite porphyries (Figure 9 and Figure 10). The batch melting model shown on the (La/Yb)N versus (Yb)N diagram indicates that the monzonites could have been derived from ~20% melting of a presumed 10% garnet-bearing amphibolite source, and that the quartz diorite porphyries was probably derived from the partial melting of a 3% garnet-bearing amphibolite source (Figure 7). In summary, the Chaoman Forest Farm monzonites and quartz diorite porphyries were likely generated by partial melting of subducting oceanic crust. As these slab-derived melts ascended, they mixed with mantle-derived magma, resulting in increased MgO and compatible elements contents, followed by minor crustal contamination.
The adakitic Early Jurassic syenogranites and Early Cretaceous quartz monzonite porphyries have high SiO2 and K2O and low A/NK ratios, MgO (0.51–1.10 wt.%) and Mg# (32–39) values similar to those of experimentally-derived metabasaltic and eclogitic melts at high pressures (1.0–4.0 GPa) and adakitic rocks originating from a thickened lower crust (Figure 6b,c and Figure 11a,b). Melting experiments indicate that primary melts from basaltic rocks exhibit low MgO and Mg# values [106], akin to those observed in adakitic rocks formed through partial melting of thickened lower crust [107]. In the Cr versus SiO2, Ni versus SiO2, and Ni versus Cr diagrams (Figure 11c–e), all syenogranites and quartz monzonite porphyries plot within the adakitic rocks field, indicating derivation through partial melting of the thickened lower crust, consistent with their very low concentrations of compatible elements (Cr = 1.89–2.91 ppm, Ni = 1.47–2.68 ppm) (Figure 11e). Therefore, we suggest that the Early Jurassic syenogranites and Early Cretaceous quartz monzonite porphyries were probably derived by partial melting of the thickened lower crust.
Prior research has suggested that adakites derived from the lower crust originate from eclogite or garnet amphibolite source rocks [91,108,109]. Nevertheless, garnet exhibits significance enrichment in HREE, whereas the amphibole is relatively enriched in MREE [110]. The concave-upward Dy–Ho–Er–Tm-depleted patterns (Figure 8d) observed in the syenogranites and quartz monzonite porphyries suggest that their formation involved partial melting of the basaltic lower crust under variable water fugacities, producing a residue in garnet and amphibole [11,69]. The syenogranites have higher La/Sm and Sm/Yb ratios than the quartz monzonite porphyries, suggesting that the syenogranites originated from a thickened lower crust with either an LREE-enriched source or a garnet-bearing, strongly metasomatized source typical of continental arc settings [96]. According to the (La/Yb)N vs. (Yb)N diagram derived from batch-melting modelling (Figure 7), the syenogranites may have formed through less than 20% melting of a presumed source composed of 15% garnet-bearing amphibolite. In contrast, the quartz monzonite porphyry samples show 10%–30% partial melting of a 5% garnet-bearing amphibolite source (Figure 7).
In addition, the syenogranites show markedly distinct initial Sr–Nd–Pb-Hf isotopic compositions from the quartz monzonite porphyries (Figure 9 and Figure 10), there is a disparity in their geochemical characters indicating different magma sources. The syenogranites have lower εHf(t) (−5.3 to 0.9), εNd(t) (−8.46 to −8.45) values, high (87Sr/86Sr)i (0.7085–0.7086) and slightly low radiogenic Pb isotopic compositions with (206Pb/204Pb)i = 18.247–18.292 relative to those of the quartz monzonite porphyries (Figure 9 and Figure 10). All syenogranite samples plot in the evolutionary zone of the Paleoproterozoic–Mesoproterozoic crust below the chondrite evolutionary line (Figure 9a), implying that the syenogranites were probably formed by reworking of Paleoproterozoic–Mesoproterozoic mafic lower crust. However, the quartz monzonite porphyries have high Hf isotopic compositions close to those of the monzonite and quartz diorite porphyries (Figure 9a), indicating that they were the mixing products between crustal and mantle-derived melts rather than a single Mesoproterozoic juvenile mafic lower crust.
Therefore, we conclude that the Early Jurassic syenogranites resulted from the partial melting of a thickened Paleoproterozoic–Mesoproterozoic mafic lower rocks (15% garnet-bearing amphibolite). The Early Cretaceous quartz monzonite porphyries were probably derived by partial melting of a thickened Mesoproterozoic juvenile mafic lower crust (5% garnet-bearing amphibolite) with variable contributions of mantle-derived magma.

5.3. Spatial-Temporal Variations of Nd and Hf Isotopes of the Mesozoic Magmatic Rocks in the Erguna Block

The integration of isotopic data from the Mesozoic magmatic rocks in the Erguna Block reveals distinctive geochemical characteristics in zircon εHf(t) and whole-rock εNd(t) values across various locations [5,19,74,75,82]. The zircon εHf(t) and whole-rock εNd(t) values of the Mesozoic granitoids in the Erguna Block change with longitude and latitude, specifically. The lower continental crust shows a progressive increase in the incorporation of ancient crustal material from southwest to northeast [5,82]. For example, the Early Jurassic granitoids from the western part of the Erguna Block have higher zircon εHf(t) and whole-rock εNd(t) values than those of the rocks in the eastern part of the Erguna Block [5,19,74,75,82]. Examining the correlation between zircon ages, TDM2 ages, and the variation in εHf(t) values across latitudinal gradients, Sun et al., 2017 suggested that the Early-Middle Triassic granitoids in the northeast of the Erguna Block were derived from Mesoproterozoic source rocks; the Late Jurassic granitoids in the central Erguna Block were derived from late Mesoproterozoic source rocks; and the Late Jurassic and Early Cretaceous granitoids in the southwest Erguna Block were derived from Neoproterozoic source rocks [82]. Notably, most of the Early Jurassic felsic rocks in the southwest part of the Erguna Block have higher εHf(t) values than those of contemporaneous mafic rocks in the Erguna Block (Figure 12) [5]. Thus, the mantle-derived melt additions are not a cause of the relatively high zircon εHf(t) values of the Early Jurassic felsic rocks in the southwest. This implies that the spatial variation of isotopic compositions of the Mesozoic rocks is a response to a heterogeneity of the lower continental crust of the Erguna Block during the Mesozoic [5,82]. This interpretation is supported by the presence of outcrops of Neoproterozoic and Paleoproterozoic granitoids in the northeast of the Erguna Block, which have obvious negative εHf(t) values (−3.9 to −8.5) with corresponding Archaeozoic to Paleoproterozoic Hf model ages, indicating that the Erguna Block is underlain by a Paleoproterozoic and even an Archaeozoic basement [42,83,111,112,113,114].
Furthermore, based on the statistics of zircon εHf(t) and whole-rock εNd(t) values of the Mesozoic magmatic rocks in the Erguna Block, these magmatic rocks have formed from the end of the Permian at least to the end of the Mesozoic and also show temporal variation in Nd-Hf isotopic composition (Figure 12). In terms of Nd–Hf isotopic compositions (Figure 12), it is found that their zircon εHf(t) and whole-rock εNd(t) values gradually increase with a decrease in their ages (Figure 12). As discussed above, the Early Jurassic granitoids of the Chaoman Forest Farm show negative whole-rock εNd(t) and zircon εHf(t) values, suggesting their predominant derivation from mature or recycled continental crust materials. In contrast, the Late Jurassic to Early Cretaceous granitoids from the Chaoman Forest Farm mainly show negative to near-zero εNd(t) and positive εHf(t) values (Figure 12), implying that they were probably derived from the juvenile crust. The relationship between Nd-Hf isotopic data and the timing of magmatic rock formation indicates a shift in magma sources from ancient crustal melting to juvenile crust during the Mesozoic [82,114]. The temporal variations observed are equal in scale to spatial geochemical trends persistent throughout the Mesozoic across the Erguna Block and such spatial–temporal variations of isotopic compositions are a common feature of magmatic rocks generated in continental magmatic arcs with different basement components [91,115,116].
Figure 12. Diagrams of whole-rock εNd(t) (a) and zircon εHf(t) (b) against U–Pb ages of the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: Mesozoic mafic rocks in the Erguna Block [5,19,111]; Mesozoic intermediate–felsic rocks in the Erguna Block [2,4,5,8,10,13,15,16,17,74,75,117,118,119,120,121].
Figure 12. Diagrams of whole-rock εNd(t) (a) and zircon εHf(t) (b) against U–Pb ages of the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: Mesozoic mafic rocks in the Erguna Block [5,19,111]; Mesozoic intermediate–felsic rocks in the Erguna Block [2,4,5,8,10,13,15,16,17,74,75,117,118,119,120,121].
Minerals 14 00725 g012
It is widely accepted that the Mongol–Okhotsk Oceanic plate was subducted southward beneath the Erguna Block during the Late Permian to Early Jurassic [3,4,5,82,114]. Consequently, the north margin of the Erguna Block formed a continental magmatic arc and the Late Permian to Early Jurassic juvenile mafic rocks underplated into the lower crust of the magmatic arc. As a result, the Late Jurassic and Early Cretaceous granitoids in the west derived from the juvenile mafic lower crust of the Permian–Jurassic magmatic arc have relatively high zircon εHf(t) and whole-rock εNd(t) values, whereas the Early Jurassic granitoids in the east derived from the ancient mafic rocks in the lower crust of the Erguna Block have relatively low zircon εHf(t) and whole-rock εNd(t) values (Figure 12b). Overall, the Mesozoic granitoids in the Erguna Block were derived from the source rocks with different ages in different parts of the lower continental crust due to crustal accretion and reworking process [5,82,114]. The findings suggest a gradual replacement of the ancient lower crust in the Erguna Block by mantle-derived juvenile materials from the Early Jurassic to Early Cretaceous, with the underplating of arc basaltic magmas playing a pivotal role in vertical crustal growth.

5.4. Tectonic Setting of Mesozoic Magmatism in the Erguna Block

5.4.1. Early Jurassic Active Continental Margin

After the final closure of the Paleo-Asian Ocean during the end of the Permian to beginning of the Early Triassic [1,86,122,123], the tectonic evolution of NE China was characterized predominantly by the Paleo-Asian and Paleo-Pacific tectonic regimes during the Mesozoic [3,5,13]. It has previously been thought that the Mongol–Okhotsk Oceanic plate was subducted southward beneath the Erguna Block between the Late Permian and Early Jurassic, as supported by the occurrence of calc-alkaline volcanic belts and granitoid intrusions along the southern margins of the Mongol–Okhotsk orogenic belt [1,3,5,45,85,86,124]. The northern margin of the Erguna Block is generally considered as an Andean-type continental arc during the early Mesozoic, i.e., Late Triassic-Early Jurassic magmatism in the Erguna Block was associated with the southward subduction of the Mongol–Okhotsk Ocean [3,6,13]. The existence of subduction-related porphyry Cu–Mo deposits [16] (e.g., Wunugetushan, ~180 Ma) indicates an active continental margin setting for the Erguna Block. In addition, calc-alkaline geochemical characteristics, enrichment in LILE, and depletion in HFSE also clearly point to a subduction-related origin for the Early Jurassic magmas [5,8,13].
The Early Jurassic syenogranite samples are characterized by significant enrichment in LILE and LREE elements, along with depletion in HFSE and HREE elements. They fall within the volcanic arc field (Figure 13), showcasing geochemical features akin to subduction-related magmas [125]. This indicates that the Mongol–Okhotsk Oceanic plate was undergoing persistent southwards subduction during the Early Jurassic (Figure 14a) [8]. The Early Jurassic syenogranites in the Chaoman Forest Farm are adakitic and formed from primary magmas generated by the partial melting of a thickened lower crust. Earlier research has demonstrated that during the Early Jurassic, the Erguna Block underwent crustal shortening and thickening [11,86]. The compressional active continental margin setting during the Early Jurassic might be triggered by a low subduction angle of the Mongol–Okhotsk Oceanic plate, similar to an Andean-type active continental margin [126].

5.4.2. Middle Jurassic–Early Early Cretaceous Extension

In the Badaguan area of the northern Euguna Block, the syenogranite with an age of ~155 Ma has strongly negative Eu anomalies and high alkaline contents, suggesting its affinity to A-type granite [4]. The volcanic rocks of the Tamulangou Formation in the Manzhouli area of the central Euguna Block (Figure 1), such as basaltic trachyandesites, trachyandesites, and trachytes, dated between 158–166 Ma (with a peak age of 162 Ma), exhibit geochemical characteristics transitioning from alkaline to sub-alkaline compositions [12,128]. Thus, the Erguna Block could have undergone a transition from compression to extension during the late Middle Jurassic. This process can also be demonstrated by the developmental history of the Mohe Basin. During the Jurassic, against the backdrop of the gradual west-to-east subduction of the Mongol–Okhotsk Ocean, the Mohe Basin underwent four distinct stages: initial basin formation, compression and subsequent uplift, formation of volcanic rift basins, and subsequent uplift and contraction. The western part of the basin displays thrust-over structures, while the central and eastern portions exhibit characteristics of extensional rifting [129,130]. A new paleomagnetic study suggests that the Mongol–Okhotsk Ocean was still ~3000 km wide in its eastern part at ca. 155 Ma [25], so the extensional setting is unlikely to have resulted from a post-collisional setting. Instead, a more plausible explanation is that it was triggered by the rollback of the subducted Mongol–Okhotsk Oceanic plate.
However, the geodynamic mechanism related to the late Late Jurassic–early Early Cretaceous (155–143 Ma) magmatism in the Euguna Block has been the subject of intense debate. One of the most controversial issues regarding the tectonic evolution is the timing of the final closure of the Mongol–Okhotsk Ocean. Although some authors have suggested that the Mongol–Okhotsk Ocean closed during the Middle Jurassic, to the northwest of the Erguna Block [2,11,22,131,132,133,134,135], a larger number of studies have argued that the southwards subduction of the Mongol–Okhotsk Oceanic plate lasted until the Late Jurassic–Early Cretaceous on the basis of the Late Jurassic–Early Cretaceous subduction-related igneous rocks associated with the newly discovered Fukeshan porphyry Cu–Mo deposit and paleomagnetic data [24,25]. In addition, Yang et al., (2015) suggested that the rapid closure of the Mongol–Okhotsk Ocean during the latest Jurassic–earliest Cretaceous led to the development of a large fold-and-thrust belt in northern China and Mongolia [26]. Guo et al., (2017b) reviewed the stratigraphic and sedimentological characteristics for the Mohe–Upper Amur Basin and proposed that the closure of the eastern Mongol–Okhotsk Ocean should have occurred in the latest Jurassic–early Cretaceous [119].
In this study, we explore these questions by analyzing the associations of magmatic rocks and investigating spatial and temporal variations in the late Mesozoic magmatism within the Erguna Block. The Early Jurassic to late Early Cretaceous magmatic rocks are composed of a series of basic to acid rocks with different geochemical characteristics. As discussed above, the late Late Jurassic to early Early Cretaceous monzonites, quartz diorite porphyries, and quartz monzonite porphyries in the Erguna Block are adakite-like rocks. The monzonites and quartz diorite porphyries were formed by partial melting of the subducting oceanic crust, and the quartz monzonite porphyries were formed by partial melting of a thickened lower crust. In addition, the trace element distribution patterns of the studied samples exhibit enriched LREE, LILE, and fluid mobile elements and depleted HFSE (Nb-Ta-Ti), as observed in subduction-related arc magmatic rocks [136,137]. In both Nb versus Y and Rb versus (Y + Nb) diagrams [127], the felsic samples plot within the volcanic arc field (Figure 13). All these features suggest that the tectonic setting of the Late Jurassic to early Early Cretaceous magmatic rocks might belong to an active continental margin, related to the subduction of the Mongol–Okhotsk Oceanic plate.
Experimental results and thermal modeling indicate that partial melting of the subducted oceanic crust only occurs primarily in hot subduction zone settings where temperatures range between 800 and 1000 °C at depths of 70–80 km [97,138,139], including mid-ocean ridge subduction [104,140] and oceanic slab roll-back regions [69,105]. Thus, a simple normal-angle (cold) subduction model cannot easily explain shallow slab-melting (i.e., oceanic slab-derived adakite-like rocks such as the studied monzonites and quartz diorite porphyries). Deng et al., (2019a) attributed the late Late Jurassic adakite-like magmas to the subduction of a mid-ocean ridge [8]. If this model were correct, mid-ocean ridge subduction would produce a high temperature and relatively H2O-poor magmas above the slab window. This is inconsistent with requirement of the amphibole crystallization as noted above. Here, we suggest that asthenospheric upwelling—probably a result of slab roll-back—provided the required high thermal regime necessary for partial melting of the subducting Mongol–Okhotsk Oceanic crust and thus was responsible for the formation of the Chaoman oceanic slab-derived monzonites and quartz diorite porphyries (Figure 14b). The slab roll-back would induce the invasion of deep asthenosphere into the mantle wedge and the development of an extensional environment in the overlying lithosphere mantle, which was further verified by the presence of the simultaneously A-type granites [4]. The high thermal anomaly resulted in the partial melting of subducted oceanic crust, forming the adakitic monzonites and quartz diorite porphyries (Figure 14b). Furthermore, the thickening and reworking of the ancient crust of the Erguna Block, formed the coexisting thickened lower crust-derived adakitic quartz monzonite porphyries (Figure 14b). Thus, the Erguna Block may still be an active continental margin attributed to the southward subduction of the Mongol–Okhotsk Oceanic plate before ca. 143 Ma.

6. Conclusions

(1) The syenogranites have consistent ages of ca. 190 Ma and were emplaced in the Early Jurassic. Intrusions of the monzonites, quartz diorite porphyries, and quartz monzonite porphyries occurred during the Late Jurassic to early Early Cretaceous at ca. 147–143 Ma. These rocks exhibit typical adakitic geochemical characteristics.
(2) The syenogranites and quartz monzonite porphyries originated from the thickened ancient mafic lower continental crust and thickened juvenile lower crust, respectively. In contrast, the monzonites and quartz diorite porphyries potentially formed because of melting associated with an oceanic slab.
(3) There is a heterogeneity of the lower continental crust in the Erguna Block during the Mesozoic and during the Early Jurassic to Early Cretaceous, and a gradual replacement of the ancient lower crust by mantle-derived juvenile materials occurred in the Erguna Block. The Erguna Block may still be an active continental margin attributed to the southward subduction of the Mongol–Okhotsk Oceanic plate before ca. 143 Ma.
(4) The Early Jurassic magmatic rocks were formed in an active continental margin setting related to the southward subduction of the Mongol–Okhotsk oceanic plate. The Late Jurassic to early Early Cretaceous magmatism in the Erguna Block occurred in an extensional environment that was probably triggered by the asthenosphere upwelling as a result of the rollback of the subducted Mongol–Okhotsk Oceanic plate, whereas the late Early Cretaceous magmatism occurred in an extensional environment that was related to the upwelling of asthenospheric material as a result of delamination of the previously thickened continental lithosphere.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/min14070725/s1, Table S1: LA-ICP-MS zircon U–Pb isotopic dating results of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block; Table S2: Whole rock major and trace element data of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block; Table S3: Zircon Hf isotope of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block; Table S4: Whole-rock Sr-Nd-Pb isotope of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block.

Author Contributions

Y.W.: Conceptualization, Data curation, Investigation, Visualization, Writing—original draft. Y.Z.: Conceptualization, Project administration, Funding acquisition, Investigation, Resources, Supervision, Writing—review & editing. X.S.: Conceptualization, Writing—review & editing. Z.T.: Conceptualization, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the National Key Research and Development Program of China (Grant No.2017YFC0601303), and the Program of the China Geological Survey (Grant No.DD20243088).

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article and its Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wu, F.Y.; Sun, D.Y.; Ge, W.C.; Zhang, Y.B.; Grant, M.L.; Wilde, S.A.; Jahn, B.M. Geochronology of the Phanerozoic granitoids in northeastern China. J. Asian Earth Sci. 2011, 41, 1–30. [Google Scholar] [CrossRef]
  2. Sun, D.Y.; Gou, J.; Wang, T.H.; Ren, Y.S.; Liu, Y.J.; Guo, H.Y.; Liu, X.M.; Hu, Z.C. Geochronological and geochemical constraints on the Erguna massif basement, NE China-subduction history of the Mongol–Okhotsk oceanic crust. Int. Geol. Rev. 2013, 55, 1801–1816. [Google Scholar] [CrossRef]
  3. Tang, J.; Xu, W.L.; Wang, F.; Wang, W.; Xu, M.J.; Zhang, Y.H. Geochronology and geochemistry of Early-Middle Triassic magmatism in the Erguna Massif, NE China: Constraints on the tectonic evolution of the Mongol-Okhotsk Ocean. Lithos 2014, 184, 1–16. [Google Scholar] [CrossRef]
  4. Tang, J.; Xu, W.L.; Wang, F.; Zhao, S.; Li, Y. Geochronology, geochemistry, and deformation history of late Jurassic-Early Cretaceous intrusive rocks in the Erguna Massif, NE China: Constraints on the late Mesozoic tectonic evolution of the Mongol-Okhotsk orogenic belt. Tectonophysics 2015, 658, 91–110. [Google Scholar] [CrossRef]
  5. Tang, J.; Xu, W.L.; Wang, F.; Zhao, S.; Wang, W. Early Mesozoic southward subduction history of the Mongol–Okhotsk oceanic plate: Evidence from geochronology and geochemistry of Early Mesozoic intrusive rocks in the Erguna Massif, NE China. Gondwana Res. 2016, 31, 218–240. [Google Scholar] [CrossRef]
  6. Zhang, Y.T.; Sun, F.Y.; Wang, S.; Xin, W. Geochronology and geochemistry of Late Jurassic to Early Cretaceous granitoids in the northern Great Xing’an Range, NE China: Petrogenesis and implications for late Mesozoic tectonic evolution. Lithos 2018, 312, 171–185. [Google Scholar] [CrossRef]
  7. Gou, J.; Sun, D.Y.; Qin, Z. Late Jurassic–Early Cretaceous tectonic evolution of the Great Xing’an Range: Geochronological and geochemical evidence from granitoids and volcanic rocks in the Erguna Block, NE China. Int. Geol. Rev. 2019, 61, 1842–1863. [Google Scholar] [CrossRef]
  8. Deng, C.Z.; Sun, D.Y.; Han, J.S.; Chen, H.Y.; Li, G.H.; Xiao, B.; Li, R.C.; Feng, Y.Z.; Li, C.L.; Lu, S. Late-stage southwards subduction of the Mongol-Okhotsk oceanic slab and implications for porphyry Cu-Mo mineralization: Constraints from igneous rocks associated with the Fukeshan deposit, NE China. Lithos 2019, 326–327, 341–357. [Google Scholar] [CrossRef]
  9. Deng, C.Z.; Sun, D.Y.; Han, J.S.; Li, G.H.; Feng, Y.Z.; Xiao, B.; Li, R.C.; Shi, H.L.; Xu, G.Z.; Yang, D.G. Ages and petrogenesis of the Late Mesozoic igneous rocks associated with the Xiaokele porphyry Cu–Mo deposit, NE China and their geodynamic implications. Ore Geol. Rev. 2019, 107, 417–433. [Google Scholar] [CrossRef]
  10. Wu, G.; Chen, Y.J.; Sun, F.Y.; Li, J.C.; Li, Z.T.; Wang, X.J. Geochemistry of the late Jurassic granitoids in the northern end area of Da Hinggan Mountains and their geological and prospecting implications. Acta Petrol. Sin. 2008, 24, 899–910, (In Chinese with English Abstract). [Google Scholar]
  11. Li, Y.; Xu, W.L.; Tang, J.; Pei, F.P.; Wang, F.; Sun, C.Y. Geochronology and geochemistry of Mesozoic intrusive rocks in the Xing’an Massif of NE China: Implications for the evolution and spatial extent of the Mongol-Okhotsk tectonic regime. Lithos 2018, 304–307, 57–73. [Google Scholar] [CrossRef]
  12. Xu, W.L.; Pei, F.P.; Wang, F.; Meng, E.; Ji, W.Q.; Yang, D.B.; Wang, W. Spatial-temporal relationships of Mesozoic volcanic rocks in NE China: Constraints on tectonic overprinting and transformations between multiple tectonic regimes. J. Asian Earth Sci. 2013, 74, 167–193. [Google Scholar] [CrossRef]
  13. Gou, J.; Sun, D.Y.; Ren, Y.S.; Hou, X.G.; Yang, D.G. Geochemical and Hf isotopic compositions of Late Triassic–Early Jurassic intrusions of the Erguna Block, Northeast China: Petrogenesis and tectonic implications. Int. Geol. Rev. 2017, 59, 347–367. [Google Scholar] [CrossRef]
  14. Zhang, K.J. Genesis of the Late Mesozoic Great Xing’an Range Large Igneous Province in eastern central Asia: A Mongol-Okhotsk slab window model. Int. Geol. Rev. 2014, 56, 1557–1583. [Google Scholar] [CrossRef]
  15. Mi, K.F.; Liu, Z.J.; Li, C.F.; Liu, R.B.; Wang, J.P.; Peng, R.M. Origin of the Badaguan porphyry Cu-Mo deposit, Inner Mongolia, northeast China: Constraints from geology, isotope geochemistry and geochronology. Ore Geol. Rev. 2017, 81, 154–172. [Google Scholar] [CrossRef]
  16. Mi, K.F.; Liu, Z.J.; Liu, R.B.; Li, C.F.; Wang, J.P.; Peng, R.M. U–Pb zircon, geochemical and Sr–Nd–Hf isotopic constraints on age and origin of the intrusions from Wunugetushan porphyry deposit, Northeast China: Implication for Triassic–Jurassic Cu–Mo mineralization in Mongolia–Erguna metallogenic belt. Int. Geol. Rev. 2018, 60, 496–512. [Google Scholar] [CrossRef]
  17. Kang, Y.J.; She, H.Q.; Lai, Y.; Wang, Z.Q.; Li, J.W.; Zhang, Z.H.; Xiang, A.P.; Jiang, Z.S. Evolution of Middle-Late Triassic granitic intrusions from the Badaguan Cu-Mo deposit, Inner Mongolia: Constraints from zircon U-Pb dating, geochemistry and Hf isotopes. Ore Geol. Rev. 2018, 95, 195–215. [Google Scholar] [CrossRef]
  18. Zhang, K.J.; Yan, L.L.; Ji, C. Switch of NE Asia from extension to contraction at the mid-Cretaceous: A tale of the Okhotsk oceanic plateau from initiation by the Perm Anomaly to extrusion in the Mongol–Okhotsk ocean? Earth-Sci. Rev. 2019, 198, 102941. [Google Scholar] [CrossRef]
  19. Chen, Z.G.; Zhang, L.C.; Wan, B.; Wu, H.Y.; Cleven, N. 2011, Geochronology and geochemistry of the Wunugetushan porphyry Cu–Mo deposit in NE China, and their geological significance. Ore Geol. Rev. 2011, 43, 92–105. [Google Scholar]
  20. Kuzmin, M.I.; Fillipova, I.B. The history of the Mongol–Okhotsk belt in the Middle–Late Paleozoic and Mesozoic. Lithospheric Plate Struct. 1979, 8, 189–226. (In Russian) [Google Scholar]
  21. Zorin, Y.A. Geodynamics of the western part of the Mongolia-Okhotsk collisional belt, Trans-Baikal region (Russia) and Mongolia. Tectonophysics 1999, 306, 189–226. [Google Scholar] [CrossRef]
  22. Tomurtogoo, O.; Windley, B.F.; Kröner, A.; Badarch, G.; Liu, D.Y. Zircon age and occurrence of the Adaatsag ophiolite and Muron shear zone, Central Mongolia: Constraints on the evolution of the Mongol-Okhotsk Ocean, suture and orogen. J. Geol. Soc. 2005, 162, 125–134. [Google Scholar] [CrossRef]
  23. Ying, J.F.; Zhou, X.H.; Zhang, L.C.; Wang, F.; Zhang, Y.T. Geochronological and geochemical investigation of the late Mesozoic volcanic rocks from the Northern Great Xing’an Range and their tectonic implications. Int. J. Earth Sci. 2010, 99, 357–378. [Google Scholar] [CrossRef]
  24. Metelkin, D.V.; Vernikovsky, V.A.; Kazansky, A.Y.; Wingate, M.T.D. Late Mesozoic tectonics of Central Asia based on paleomagnetic evidence. Gondwana Res. 2010, 18, 400–419. [Google Scholar] [CrossRef]
  25. Pei, J.L.; Sun, Z.M.; Liu, J.; Liu, J.; Wang, X.S.; Yang, Z.Y.; Zhao, Y.; Li, H.B. A paleomagnetic study from the late Jurassic volcanics (155 Ma), North China: Implications for the width of Mongol-Okhotsk Ocean. Tectonophysics 2011, 510, 370–380. [Google Scholar]
  26. Yang, Y.T.; Guo, Z.X.; Song, C.C.; Li, X.B.; He, S. A short-lived but significant Mongol-Okhotsk collisional orogeny in latest Jurassic-earliest cretaceous. Gondwana Res. 2015, 28, 1096–1116. [Google Scholar] [CrossRef]
  27. Ren, Q.; Zhang, S.H.; Wu, Y.Q.; Yang, T.S.; Gao, Y.J.; Turbold, S.; Zhao, H.Q.; Wu, H.C.; Li, H.Y.; Fu, H.R.; et al. New Late Jurassic to Early Cretaceous Paleomagnetic Results From North China and Southern Mongolia and Their Implications for the Evolution of the Mongol-Okhotsk Suture. J. Geophys. Res. -Solid Earth 2018, 123, 10–370. [Google Scholar] [CrossRef]
  28. Jahn, B.M.; Wu, F.Y.; Chen, B. Massive granitoid generation in central Asia: Nd isotopic evidence and implication for continental growth in the Phanerozoic. Episodes 2000, 23, 82–92. [Google Scholar] [CrossRef] [PubMed]
  29. Xu, B.; Zhao, P.; Bao, Q.Z.; Zhou, Y.H.; Wang, Y.Y.; Luo, Z.W. Preliminary study on the Pre-Mesozoic tectonic unit division of the Xing-Meng Orogenic Belt (XMOB). Acta Petrol. Sin. 2014, 30, 184–1857. [Google Scholar]
  30. Zhou, J.B.; Cao, J.L.; Wilde, S.A.; Zhao, G.C.; Zhang, J.J.; Wang, B. Paleo-Pacific subduction-accretion: Evidence from Geochemical and U-Pb zircon dating of the Nadanhada accretionary complex, NE China. Tectonics 2014, 33, 2444–2466. [Google Scholar] [CrossRef]
  31. Zhou, J.B.; Li, L. The Mesozoic accretionary complex in Northeast China: Evidence for the accretion history of Paleo-pacific subduction. J. Asian Earth Sci. 2017, 145, 91–100. [Google Scholar] [CrossRef]
  32. Sengör, A.M.C.; Natal’in, B.A. Paleotectonics of Asia: Fragments of a Synthesis. In The Tectonic Evolution of Asia; Yin, A., Harrison, M., Eds.; Cambridge University Press: Cambridge, UK, 1996; pp. 486–640. [Google Scholar]
  33. Ge, W.C.; Wu, F.Y.; Zhou, C.Y.; Rahman, A.A.A. Emplacement age of the Tahe granite and its constraints on the tectonic nature of the Ergun block in the northern part of the Da Hinggan Range. Chin. Sci. Bull. 2005, 50, 2097–2105. [Google Scholar]
  34. Wu, G.; Sun, F.Y.; Zhao, C.S.; Li, Z.T.; Zhao, A.L.; Pang, Q.B.; Li, G.Y. Discovery of the Early Paleozoic post-collisional granites in northern margin of the Erguna massif and its geological significance. Chin. Sci. Bull. 2005, 50, 2733–2743. [Google Scholar] [CrossRef]
  35. Feng, Z.Q.; Liu, Y.J.; Li, L.; She, H.Q.; Jiang, L.W.; Du, B.Y.; Liu, Y.W.; Li, W.M.; Wen, Q.B.; Liang, C.Y. Subduction, accretion, and collision during the Neoproterozoic–Cambrian orogeny in the Great Xing’an Range, NE China: Insights from geochemistry and geochronology of the Ali River ophiolitic mélange and arc-type granodiorites. Precambrian Res. 2018, 311, 117–135. [Google Scholar] [CrossRef]
  36. Sorokin, A.A.; Kudryashov, N.M.; Li, J.Y.; Zhuravlev, D.Z.; Yan, P.; Sun, G.H.; Gao, L.M. Early Paleozoic granitoids in the eastern margin of the Argun’ terrane, Amur area: First geochemical and geochronologic data. Petrology 2004, 12, 367–376. [Google Scholar]
  37. Wu, G.; Chen, Y.C.; Chen, Y.J.; Zeng, Q.T. Zircon U–Pb ages of the metamorphic supracrustal rocks of the Xinghuadukou Group and granitic complexes in the Argun massif of the northern Great Hinggan Range, NE China, and their tectonic implications. J. Asian Earth Sci. 2012, 49, 214–233. [Google Scholar] [CrossRef]
  38. Lu, L.; Qin, Y.; Zhang, K.J.; Han, C.Y.; Wei, T.; Li, Z.F.; Qu, Z.H. Provenance and tectonic settings of the Late Paleozoic sandstones in central Inner Mongolia, NE China: Constraints on the evolution of the southeastern Central Asian Orogenic Belt. Gondwana Res. 2020, 77, 111–135. [Google Scholar] [CrossRef]
  39. Heilongjiang Bureau of Geology and Mineral Resources. Regional Geology of Heilongjiang Province; Geological Publishing House: Beijing, China, 1993; (In Chinese with English Abstract). [Google Scholar]
  40. Miao, L.C.; Fan, W.M.; Zhang, F.Q.; Liu, D.Y.; Jian, P.; Shi, G.H.; Tao, H.; Shi, Y.R. Zircon SHRIMP geochronology of the Xinkailing-Kele complex in the northwestern Lesser Xing’an Range, and its geological implications. Chin. Sci. Bull. 2004, 49, 201–209. [Google Scholar] [CrossRef]
  41. Gou, J.; Sun, D.Y.; Liu, Y.J.; Ren, Y.S.; Zhao, Z.H.; Liu, X.M. Geochronology, petrogenesis, and tectonic setting of Mesozoic volcanic rocks, southern Manzhouli area, Inner Mongolia. Int. Geol. Rev. 2013, 55, 1029–1048. [Google Scholar] [CrossRef]
  42. Sun, L.X.; Ren, B.F.; Zhao, F.Q.; Ji, S.P.; Geng, J.Z. Late Paleoproterozoic magmatic records in the Erguna massif: Evidences from the zircon U-Pb dating of granitic gneisses. Geol Bull China 2013, 32, 341–352, (In Chinese with English Abstract). [Google Scholar]
  43. Lu, L.; Qin, Y.; Zhang, K.J.; Han, C.Y.; Li, Z.F.; Wei, T.; Qu, Z.H.; Xu, Y.Z. Petrogenesis of a post-collisional magmatism at the Carboniferous–Permian boundary in central Inner Mongolia, NE China: Insights into the closure timing of the Hegenshan Ocean? Int. Geol. Rev. 2019, 62, 2013–2038. [Google Scholar] [CrossRef]
  44. Gou, J.; Sun, D.Y.; Zhao, Z.H.; Ren, Y.S.; Zhang, X.Y.; Fu, C.L.; Wang, X.; Wei, H.Y. Zircon LA-ICPMS U-Pb dating and petrogenesis of rhyolites in Baiyingaolao Formation from the southern Manzhouli, Inner-Mongolia. Acta Petrol. Sin. 2010, 26, 333–344, (In Chinese with English Abstract). [Google Scholar]
  45. Wu, G.; Chen, Y.J.; Sun, F.Y.; Zhang, Z.; Liu, A.K.; Li, Z.T. Geochemistry and genesis of the Late Jurassic granitoids at the northern Great Hinggan Range: Implications for exploration. Acta Geol. Sin. 2010, 84, 321–332. [Google Scholar] [CrossRef]
  46. Gao, B.Y.; Zhang, L.C.; Jin, X.D.; Li, W.J.; Chen, Z.G.; Zhu, M.T. Geochronology and geochemistry of the Badaguan porphyry Cu–Mo deposit in Derbugan metallogenic belt of the NE China, and their geological significances. Int. J. Earth Sci. 2016, 105, 507–519. [Google Scholar] [CrossRef]
  47. Zhang, L.C.; Gao, B.Y.; Li, W.J.; Chen, Z.G.; Sakyi, P.A.; Jin, X.D. Early Mesozoic tectonomagmatic activity and mineralization in Northeast China: Evidence from Re–Os to U–Pb studies of the Taipingchuan porphyry Cu–Mo deposit in the Derbugan metallogenic belt. Int. Geol. Rev. 2014, 56, 1837–1851. [Google Scholar] [CrossRef]
  48. Wang, Y.H.; Zhao, C.B.; Zhang, F.F.; Liu, J.J.; Wang, J.P.; Peng, R.M.; Liu, B. SIMS zircon U-Pb and molybdenite Re–Os geochronology, Hf isotope, and whole-rock geochemistry of the Wunugetushan porphyry Cu–Mo deposit and granitoids in NE China and their geological significance. Gondwana Res. 2015, 28, 1228–1245. [Google Scholar] [CrossRef]
  49. Xu, L.Q.; Liu, C.; Deng, J.F.; Li, N.; Dai, M.; Bai, L.B. Geochemical characteristics and zircon U-Pb SHRIMP age of igneous rocks in Erentaolegai silver deposit, Inner Mongolia. Acta Petrol. Sin. 2014, 30, 3203–3212, (In Chinese with English Abstract). [Google Scholar]
  50. Li, T.G.; Wu, G.; Liu, J.; Hu, Y.Q.; Zhang, Y.F.; Luo, D.F.; Mao, Z.H. Fluid inclusions and isotopic characteristics of the Jiawula Pb-Zn-Ag deposit, Inner Mongolia, China. J. Asian Earth Sci. 2015, 103, 305–320. [Google Scholar] [CrossRef]
  51. Hou, K.J.; Li, Y.H.; Tian, Y.R. In situ U-Pb zircon dating using laser ablation-multi ion counting-ICP-MS. Miner. Depos. 2009, 28, 481–492. [Google Scholar]
  52. Liu, Y.S.; Gao, S.; Hu, Z.C.; Gao, C.G.; Zong, K.Q.; Wang, D.B. Continental and oceanic crust recycling-induced melt-peridotite interactions in the Trans-North China Orogen: U-Pb dating, Hf isotopes and trace elements in zircons of mantle xenoliths. J. Petrol. 2010, 51, 537–571. [Google Scholar] [CrossRef]
  53. Ludwig, K. Sqiud 2: A User’s Manual; Berkeley Geochron Center Special Publication: Berkeley, CA, USA, 2009; pp. 1–110. [Google Scholar]
  54. Andersen, T. Correction of common lead in U–Pb analyses that do not report 204 Pb. Chem. Geol. 2002, 192, 59–79. [Google Scholar] [CrossRef]
  55. Hou, K.J.; Li, Y.H.; Zou, T.R.; Qü, X.M.; Shi, Y.R.; Xie, G.Q. Laser ablation MC-ICP-MS technique for Hf isotope microanalysis of zircon and its geological applications. Acta Petrol. Sin. 2007, 23, 2595–2604, (In Chinese with English Abstract). [Google Scholar]
  56. Söderlund, U.; Patchett, P.J.; Vervoort, J.D.; Isachsen, C.E. The 176Lu decay constant determined by Lu–Hf and U–Pb isotope systematics of Precambrian mafic intrusions. Earth Planet. Sci. Lett. 2004, 219, 311–324. [Google Scholar] [CrossRef]
  57. Bouvier, A.; Vervoort, J.D.; Patchett, P.J. The Lu–Hf and Sm–Nd isotopic composition of CHUR: Constraints from unequilibrated chondrites and implications for the bulk composition of terrestrial planets. Earth Planet. Sci. Lett. 2008, 273, 133–147. [Google Scholar] [CrossRef]
  58. Griffin, W.L.; Pearson, N.J.; Belousova, E.; Jackson, S.E.; van Achterbergh, E.; O’Reilly, S.Y.; Shee, S.R. The Hf isotope composition of cratonic mantle: LAM-MC-ICPMS analysis of zircon megacrysts in kimberlites. Geochim. Cosmochim. Acta 2000, 64, 133–147. [Google Scholar] [CrossRef]
  59. Griffin, W.L.; Wang, X.; Jackson, S.E.; Pearson, N.J.; O’Reilly, S.Y.; Xu, X.; Zhou, X. Zircon chemistry and magma mixing, SE China: In-situ analysis of Hf isotopes, Tonglu and Pingtan igneous complexes. Lithos 2002, 61, 237–269. [Google Scholar] [CrossRef]
  60. Chen, F.K.; Siebel, W.; Satir, M.; Terzioglu, M.N.; Saka, K. Geochronology of the Karadere basement (NWTurkey) and implications for the geological evolution of the Istanbul zone. Int. J. Earth Sci. 2002, 91, 469–481. [Google Scholar] [CrossRef]
  61. Chen, F.K.; Li, X.H.; Wang, X.L.; Li, Q.L.; Siebel, W. Zircon age and Nd-Hf isotopic composition of the Yunnan Tethyan belt, southwestern China. Int. J. Earth Sci. 2007, 96, 1179–1194. [Google Scholar] [CrossRef]
  62. Hoskin, P.W.O.; Schaltegger, U. The composition of zircon and igneous and metamorphic petrogenesis. Rev. Mineral. Geochem. 2003, 53, 27–62. [Google Scholar] [CrossRef]
  63. Middlemost, E.A.K. Naming materials in the magma/igneous rock system. Earth-Sci. Rev. 1994, 37, 215–224. [Google Scholar] [CrossRef]
  64. Rollinson, H.R. Using Geochemical Data: Evaluation, Presentation, Interpretation; Longman Group UK: London, UK, 1993; pp. 1–352. [Google Scholar]
  65. Maniar, P.D.; Piccoli, P.M. Tectonic discrimination of granitoids. Geol. Soc. Am. Bull. 1989, 101, 635–643. [Google Scholar] [CrossRef]
  66. Defant, M.J.; Drummond, M.S. Derivation of some modern arc magmas by melting of young subducted lithosphere. Nature 1990, 347, 662–665. [Google Scholar] [CrossRef]
  67. Defant, M.J.; Drummond, M.S. Mount St. Helens: Potential example of the partial melting of the subducted lithosphere in a volcanic arc. Geology 1993, 21, 547–550. [Google Scholar] [CrossRef]
  68. Whalen, J.B.; Currie, K.L.; Chappell, B.W. A-type granites: Geochemical characteristics, discrimination and petrogenesis. Contrib. Mineral. Petrol. 1987, 95, 407–419. [Google Scholar] [CrossRef]
  69. Shui, X.F.; He, Z.Y.; Klemd, R.; Zhang, Z.M.; Lu, T.Y.; Yan, L.L. Early Jurassic adakitic rocks in the southern Lhasa sub-terrane, southern Tibet: Petrogenesis and geodynamic implications. Geol. Mag. 2018, 155, 132–148. [Google Scholar] [CrossRef]
  70. Drummond, M.S.; Defant, M.J.; Kepezhinskas, P.K. Petrogenesis of slab-derived trondhjemite–tonalite–dacite/adakite magmas. Transactions of the Royal Society of Edinburgh. Earth Sci. 1996, 87, 205–215. [Google Scholar]
  71. Sun, S.S.; McDonough, W.F. Chemical and isotopic systematics of oceanic basalts: Implications for mantle composition and process. Geol. Soc. Spec. Publ. 1989, 42, 313–345. [Google Scholar] [CrossRef]
  72. Dokuz, A.; Tanyolu, E.; Genç, S. A mantle-and a lower crust-derived bimodal suite in the Yusufeli (Artvin) area, NE Turkey: Trace element and REE evidence for subduction-related rift origin of early Jurassic Demirkent intrusive complex. Int. J. Earth Sci. 2006, 95, 370–394. [Google Scholar] [CrossRef]
  73. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Blackwell Scientific: Hoboken, NJ, USA, 1985. [Google Scholar]
  74. Chen, Z.G.; Zhang, L.C.; Lu, B.Z.; Li, Z.L.; Wu, H.Y.; Xiang, P.; Huang, S.W. Geochronology and geochemistry of the Taipingchuan copper–molybdenum deposit in Inner Mongolia, and its geological significances. Acta Petrol. Sin. 2010, 26, 1437–1449, (In Chinese with English Abstract). [Google Scholar] [CrossRef]
  75. Zhang, F.F.; Wang, Y.H.; Liu, J.J.; Wang, J.P.; Zhao, C.B.; Song, Z.W. Origin of the Wunugetushan porphyry Cu-Mo deposit, Inner Mongolia, NE China: Constraints from geology, geochronology, geochemistry, and isotopic compositions. J. Asian Earth Sci. 2016, 117, 208–224. [Google Scholar] [CrossRef]
  76. Vervoort, J.D.; Patchett, P.J.; Blichert-Toft, J.; Albarede, F. Relationships between Lu-Hf and Sm-Nb isotopic systems in the global sedimentary system. Earth Planet. Sci. Lett. 1999, 168, 79–99. [Google Scholar] [CrossRef]
  77. Hou, Z.Q.; Zhang, H.; Pan, X.; Yang, Z. Porphyry Cu (-Mo-Au) deposits related to melting of thickened mafic lower crust: Examples from the eastern Tethyan metallogenic domain. Ore Geol. Rev. 2011, 39, 21–45. [Google Scholar] [CrossRef]
  78. Dhuime, B.; Hawkesworth, C.; Cawood, P. When continents formed. Science 2011, 331, 154–155. [Google Scholar] [CrossRef] [PubMed]
  79. Hart, S.R. A large-scale isotope anomaly in the southern hemisphere mantle. Nature 1984, 309, 753–757. [Google Scholar] [CrossRef]
  80. Jiang, Y.H.; Jiang, S.Y.; Ling, H.F.; Dai, B.Z. Low-degree melting of a metasomatized lithospheric mantle for the origin of Cenozoic Yulong monzogranite-porphyry, east Tibet: Geochemical and Sr–Nd–Pb–Hf isotopic constraints. Earth Planet. Sci. Lett. 2006, 241, 617–633. [Google Scholar] [CrossRef]
  81. Zindler, A.; Hart, S.R. Chemical geodynamics. Annu. Rev. Earth Planet. Sci. 1986, 14, 493–571. [Google Scholar] [CrossRef]
  82. Sun, C.Y.; Tang, J.; Xu, W.L.; Li, Y.; Zhao, S. Crustal Accretion and Reworking Processes of Micro-continental Massifs within Orogenic Belt: A Case Study of the Erguna Massif, NE China. Sci. China Earth Sci. 2017, 60, 1256–1267. [Google Scholar] [CrossRef]
  83. She, H.Q.; Li, J.W.; Xiang, A.P.; Guan, J.D.; Yang, Y.C.; Zhang, D.Q.; Tan, G.; Zhang, B. U–Pb ages of the zircons from primary rocks in middle-northern Daxinganling and its implications to geotectonic evolution. Acta Petrol. Sin. 2012, 28, 571–594, (In Chinese with English Abstract). [Google Scholar]
  84. Wang, Q.; Li, X.H.; Jia, X.H.; Wyman, D.; Tang, G.J.; Li, Z.X.; Ma, L.; Yang, Y.H.; Jiang, Z.Q.; Gou, G.N. Late Early Cretaceous adakitic granitoids and associated magnesian and potassium-rich mafic enclaves and dikes in the Tunchang–Fengmu area, Hainan Province (South China): Partial melting of lower crust and mantle, and magma hybridization. Chem. Geol. 2012, 328, 222–243. [Google Scholar] [CrossRef]
  85. Wang, W.; Xu, W.L.; Wang, F.; Meng, E. Zircon U–Pb chronology and assemblages of Mesozoic granitoids in the Manzhouli-Erguna area, NE China: Constraints on the regional tectonic evolution. Geol. J. China Univ. 2012, 18, 88–105, (In Chinese with English Abstract). [Google Scholar]
  86. Wang, T.; Guo, L.; Zhang, L.; Yang, Q.D.; Zhang, J.J.; Tong, Y.; Ye, K. Timing and evolution of Jurassic-cretaceous granitoid magmatisms in the Mongol-Okhotsk belt and adjacent areas, NE Asia: Implications for transition from contractional crustal thickening to extensional thinning and geodynamic settings. J. Asian Earth Sci. 2015, 97, 365–392. [Google Scholar] [CrossRef]
  87. Richards, J.P.; Kerrich, R. Adakite-like rocks: Their diverse origins and questionable role in metallogenesis. Econ. Geol. 2007, 102, 537–576. [Google Scholar] [CrossRef]
  88. Streck, M.J.; Leeman, W.P.; Chesley, J. High-magnesian andesite from Mount Shasta: A product of magma mixing and contamination, not a primitive mantle melt. Geology 2007, 35, 351–354. [Google Scholar] [CrossRef]
  89. Atherton, M.P.; Petford, N. Generation of sodium-rich magmas from newly underplated basaltic crust. Nature 1993, 362, 144–146. [Google Scholar] [CrossRef]
  90. Kay, R.W.; Kay, S.M. Delamination and delamination magmatism. Tectonophysics 1993, 219, 177–189. [Google Scholar] [CrossRef]
  91. Ding, H.X.; Hou, Q.Y.; Zhang, Z.M. Petrogenesis and tectonic significance of the Eocene adakite-like rocks in western Yunnan, southeastern Tibetan Plateau. Lithos 2016, 245, 161–173. [Google Scholar] [CrossRef]
  92. Wang, Q.; Xu, J.F.; Jian, P.; Bao, Z.W.; Zhao, Z.H.; Li, C.F.; Xiong, X.L.; Ma, J.L. Petrogenesis of adakitic porphyries in an extensional tectonic setting, Dexing, South China: Implications for the genesis of porphyry copper mineralization. J. Petrol. 2006, 47, 119–144. [Google Scholar] [CrossRef]
  93. Castillo, P.R.; Janney, P.E.; Solidum, R.U. Petrology and geochemistry of Camiguin Island, southern Philippines: Insights to the source of adakites and other lavas in a complex arc setting. Contrib. Mineral. Petrol. 1999, 134, 33–51. [Google Scholar] [CrossRef]
  94. Macpherson, C.G.; Dreher, S.T.; Thirlwall, M.F. Adakites without slab melting: High pressure differentiation of island arc magma, Mindanao, the Philippines. Earth Planet. Sci. Lett. 2006, 243, 581–593. [Google Scholar] [CrossRef]
  95. Stern, C.R.; Kilian, R. Role of the subduction slab, mantle wedge and continental crust in the generation of adakites from the Andean Austral Volcanic Zone. Contrib. Mineral. Petrol. 1996, 123, 263–281. [Google Scholar] [CrossRef]
  96. Rapp, P.R.; Shimizu, N.; Norman, M.D. Applegate, reaction between slab-derived melt and peridotite in the mantle wedge: Experimental constrains at 3.8 Gpa. Chem. Geol. 1999, 160, 335–356. [Google Scholar]
  97. Sen, C.; Dunn, T. Experimental modal metasomatism of a spinel lherzolite and the production of amphibole-bearing peridotite. Contrib. Mineral. Petrol. 1994, 119, 422–432. [Google Scholar] [CrossRef]
  98. Huang, X.L.; Xu, Y.G.; Lan, J.B.; Yang, Q.J.; Luo, Z.Y. Neoproterozoic adakitic rocks from Mopanshan in the western Yangtze craton: Partial melts of a thickened lower crust. Lithos 2009, 112, 367–381. [Google Scholar] [CrossRef]
  99. Plank, T. Constraints from Thorium/Lanthanum on sediment recycling at subduction zones and the evolution of the continents. J. Petrol. 2005, 46, 921–944. [Google Scholar] [CrossRef]
  100. Plank, T.; Langmuir, C.H. The chemical composition of subducting sediment and its consequences for the crust and mantle. Chem. Geol. 1998, 145, 325–394. [Google Scholar] [CrossRef]
  101. Niu, Y.L.; Batiza, R. Trace element evidence from seamounts for recycled oceanic crust in the eastern equatorial Pacific mantle. Earth Planet. Sci. Lett. 1997, 148, 471–484. [Google Scholar] [CrossRef]
  102. DePaolo, D.J. Trace element and isotopic effects of combined wallrock assimilation and fractional crystallization. Earth Planet. Sci. Lett. 1981, 53, 189–202. [Google Scholar] [CrossRef]
  103. Hofmann, A.W. Chemical differentiation of the earth: The relationship between mantle, continental crust, and oceanic crust. Earth Planet. Sci. Lett. 1988, 90, 297–314. [Google Scholar] [CrossRef]
  104. Zhang, Z.M.; Zhao, G.C.; Santosh, M.; Wang, J.L.; Dong, X.; Shen, K. Late Cretaceous charnockite with adakitic affinities from the Gangdese batholith, southeastern Tibet: Evidence for Neo-Tethyan mid-ocean ridge subduction? Gondwana Res. 2010, 17, 615–631. [Google Scholar] [CrossRef]
  105. Ma, L.; Wang, Q.; Wyman, D.A.; Jiang, Z.Q.; Yang, J.H.; Li, Q.L.; Gou, G.N.; Guo, H.F. Late Cretaceous crustal growth in the Gangdese area, southern Tibet: Petrological and Sr–Nd–Hf–O isotopic evidence from Zhengga diorite–gabbro. Chem. Geol. 2013, 349, 54–70. [Google Scholar] [CrossRef]
  106. Rapp, R.P.; Watson, E.B. Dehydration melting of metabasalt at 8–32kbar: Implications for continental growth and crust–mantle recycling. J. Petrol. 1995, 36, 891–931. [Google Scholar] [CrossRef]
  107. Castillo, P.R. Adakite petrogenesis. Lithos 2012, 134, 304–316. [Google Scholar] [CrossRef]
  108. Chung, S.L.; Liu, D.Y.; Ji, J.Q.; Chu, M.F.; Lee, H.Y.; Wen, D.J.; Lo, C.H.; Lee, T.Y.; Qian, Q.; Zhang, Q. Adakites from continental collision zones: Melting of thickened lower crust beneath southern Tibet. Geology 2003, 31, 1021–1024. [Google Scholar] [CrossRef]
  109. Karsli, O.; Dokuz, A.; Uysal, İ.; Aydin, F.; Kandemir, R.; Wijbrans, J. Generation of the Early Cenozoic adakitic volcanism by partial melting of mafic lower crust, Eastern Turkey: Implications for crustal thickening to delamination. Lithos 2010, 114, 109–120. [Google Scholar] [CrossRef]
  110. Green, T.H. Experimental studies of trace-element partitioning applicable to igneous petrogenesis-Sedona 16 years later. Chem. Geol. 1994, 117, 1–36. [Google Scholar] [CrossRef]
  111. Shao, J.; Li, Y.F.; Zhou, Y.H.; Wang, H.B.; Zhang, J. Neo-Archaean magmatic event in Erguna Massif of northeast China: Evidence from the zircon LA-ICP-MS dating of the gneissic monzogranite from the drill. J. Jilin Univ. Earth Sci. Ed. 2015, 45, 364–373, (In Chinese with English Abstract). [Google Scholar]
  112. Zhao, S.; Xu, W.L.; Tang, J.; Li, Y.; Guo, P. Timing of formation and tectonic nature of the purportedly Neoproterozoic Jiageda Formation of the Erguna Massif, NE China: Constraints from field geology and U-Pb geochronology of detrital and magmatic zircons. Precambrian Res. 2016, 281, 585–601. [Google Scholar] [CrossRef]
  113. Zhao, S.; Xu, W.L.; Tang, J.; Li, Y.; Guo, P. Neoproterozoic magmatic events and tectonic attribution of the Erguna Massif: Constraints from geochronological, geochemical and Hf isotopic data of intrusive rocks. Earth Sci. 2016, 41, 1803–1829, (In Chinese with English Abstract). [Google Scholar]
  114. Sun, C.Y.; Xu, W.L.; Cawood, P.A.; Tang, J.; Zhao, S.; Li, Y.; Zhao, X.M. Crustal growth and reworking: A case study from the Erguna Massif, eastern central Asian orogenic Belt. Sci. Rep. 2019, 9, 1–8. [Google Scholar] [CrossRef]
  115. Zhu, D.C.; Zhao, Z.D.; Niu, Y.; Mo, X.X.; Chung, S.L.; Hou, Z.Q. The Lhasa Terrane: Record of a microcontinent and its histories of drift and growth. Earth Planet. Sci. Lett. 2011, 301, 241–255. [Google Scholar] [CrossRef]
  116. Attia, S.; Cottle, J.M.; Paterson, S.R. Erupted zircon record of continental crust formation during mantle driven arc flare-ups. Geology 2020, 48, 446–451. [Google Scholar] [CrossRef]
  117. Zhang, L.C.; Zhou, X.H.; Ying, J.F.; Wang, F.; Guo, F.; Wan, B.; Chen, Z.G. Geochemistry and Sr–Nd–Pb–Hf isotopes of Early Cretaceous basalts from the Great Xinggan Range, ne china: Implications for their origin and mantle source characteristics. Chem. Geol. 2008, 256, 12–23. [Google Scholar] [CrossRef]
  118. Sun, D.Y.; Gou, J.; Ren, Y.S.; Fu, C.L.; Wang, X.; Liu, X.M. Zircon U-Pb dating and study on geochemistry of volcanic rocks in Manitu Formation from southern Manchuria, Inner Mongolia. Acta Petrol. Sin. 2011, 27, 3083–3094. [Google Scholar]
  119. Guo, Z.X.; Yang, Y.T.; Zyabrev, S.; Hou, Z.H. Tectonostratigraphic Evolution of the Mohe-Upper Amur Basin Reflects the Final Closure of the Mongol-Okhotsk Ocean in the Latest Jurassic–Earliest Cretaceous. J. Asian Earth Sci. 2017, 145, 494–511. [Google Scholar] [CrossRef]
  120. Gou, J.; Sun, D.Y.; Hou, X.G.; Yang, D.G. Geochronology, geochemistry, and zircon Hf isotopes of the late permian–early Triassic Wuma intrusions in the Erguna Block, northeast China: Petrogenesis and implications for tectonic setting and crustal growth. Geol. J. 2018, 53, 1906–1920. [Google Scholar] [CrossRef]
  121. Wang, W.; Tang, J.; Xu, W.L.; Wang, F. Geochronology and geochemistry of Early Jurassic volcanic rocks in the Erguna Massif, northeast China: Petrogenesis and implications for the tectonic evolution of the Mongol-Okhotsk suture belt. Lithos 2015, 218–219, 73–86. [Google Scholar] [CrossRef]
  122. Sun, D.Y.; Wu, F.Y.; Zhang, Y.B.; Gao, S. The final closing time of the west Lamulun River-Changchun-Yanji plate suture zone. J. Jilin Univ. 2004, 34, 174–181, (In Chinese with English Abstract). [Google Scholar]
  123. Cao, H.H.; Xu, W.L.; Pei, F.P.; Guo, P.Y.; Wang, F. Permian tectonic evolution of the eastern section of the northern margin of the North China Plate: Constraints from zircon U-Pb geochronology and geochemistry of the volcanic rocks. Acta Petrol. Sin. 2012, 28, 2733–2750, (In Chinese with English Abstract). [Google Scholar]
  124. Li, Y.; Xu, W.L.; Wang, F.; Tang, J.; Zhao, S.; Guo, P. Geochronology and geochemistry of late Paleozoic–early Mesozoic igneous rocks of the Erguna Massif, NE China: Implications for the early evolution of the Mongol–Okhotsk tectonic regime. J. Asian Earth Sci. 2017, 144, 205–224. [Google Scholar] [CrossRef]
  125. Roberts, M.P.; Clemens, J.D. Origin of high-potassium, talc-alkaline, I-type granitoids. Geology 1993, 21, 825–828. [Google Scholar] [CrossRef]
  126. Conrad, C.P.; Bilek, S.; Lithgow-Bertelloni, C. Great earthquakes and slab pull: Interaction between seismic coupling and plate-slab coupling. Earth Planet. Sci. Lett. 2004, 218, 109–122. [Google Scholar] [CrossRef]
  127. Pearce, J.A. Sources and settings of granitic rocks. Episodes 1996, 19, 120–125. [Google Scholar] [CrossRef]
  128. Meng, E.; Xu, W.L.; Yang, D.B.; Qiu, K.F.; Li, C.H.; Zhu, H.T. Zircon U-Pb chronology, geochemistry of Mesozoic volcanic rocks from the Lingquan basin in Manzhouli area, and its tectonic implications. Acta Petrol. Sin. 2011, 27, 1209–1226, (In Chinese with English Abstract). [Google Scholar]
  129. Zhang, S.; Lin, C.M.; Wu, C.D.; Yang, J.G. Tectonic characteristics and basin evolution of the Mohe basin, Heilongjiang Province. Geol. J. China Univ. 2003, 9, 411–419, (In Chinese with English Abstract). [Google Scholar]
  130. Wang, Y.C.; Zhao, Y.Y.; Liu, C.H.; Shui, X.F.; Cheng, X.D.; Gong, X.; Liu, X.; Tan, W.; Hong, J.N. Detrital zircon geochronology, geochemistry and geological significance of sandstone in the Ershierzhan Formation of the Mohe Basin. Acta Geol. Sin. 2020, 94, 869–887, (In Chinese with English Abstract). [Google Scholar]
  131. Arzhannikova, A.V.; Demonterova, E.I.; Jolivet, M.; Mikheeva, E.A.; Ivanov, A.V.; Arzhannikov, S.G.; Khubanov, V.B.; Kamenetsky, V.S. Segmental closure of the Mongol-Okhotsk Ocean: Insight from detrital geochronology in the East Transbaikalia Basin. Geosci. Front. 2022, 13, 1–13. [Google Scholar] [CrossRef]
  132. Demonterova, E.I.; Ivanov, A.V.; Mikheeva, E.A.; Arzhannikova, A.V.; Frolov, A.O.; Arzhannikov, S.G.; Bryanskiy, N.V.; Pavlova, L.A. Early to Middle Jurassic history of the southern Siberian continent (Transbaikalia) recorded in sediments of the Siberian Craton: Sm-Nd and U-Pb provenance study. Bull. Soc. Géol. Fr. 2017, 188, 1–29. [Google Scholar] [CrossRef]
  133. Sorokin, A.A.; Zaika, V.A.; Kovach, V.P.; Kotov, A.B.; Xu, W.; Yang, H. Timing of closure of the eastern Mongol–Okhotsk Ocean: Constraints from U–Pb and Hf isotopic data of detrital zircons from metasediments along the Dzhagdy Transect. Gondwana Res. 2020, 81, 58–78. [Google Scholar] [CrossRef]
  134. Kravchinsky, V.A.; Cogné, J.P.; Harbert, W.P.; Kuzmin, M. Evolution of the Mongol–Okhotsk Ocean as constrained by new palaeomagnetic data from the Mongol–Okhotsk suture zone, Siberia. Geophys. J. R. Astron. Soc. 2002, 148, 34–57. [Google Scholar] [CrossRef]
  135. Li, Y.; Ding, L.L.; Xu, W.L.; Wang, F.; Tang, J.; Zhao, S.; Wang, Z.J. Geochronology and geochemistry of muscovite granite in Sunwu area, NE China: Implications for the timing of closure of the Mongol-Okhotsk Ocean. Acta Petrol. Sin. 2015, 31, 56–66, (In Chinese with English Abstract). [Google Scholar]
  136. Hawkesworth, C.J. Mantle and Slab Contribution in Arc Magmas. Annu. Rev. Earth Planet. Sci. 1993, 21, 175–204. [Google Scholar] [CrossRef]
  137. Kelemen, P.B.; Hanghoj, K.; Greene, A. One view of the geochemistry of subduction-related magmatic arcs, with an emphasis on primitive andesite and lower crust: Treatise on. Geochemistry 2003, 3, 593–659. [Google Scholar]
  138. Peacock, S.M.; Rushmer, T.; Thompson, A.B. Partial melting of subducting oceanic-crust. Earth Planet. Sci. Lett. 1994, 121, 227–244. [Google Scholar] [CrossRef]
  139. Van Keken, P.E.; Hacker, B.R.; Syracuse EMAbers, G.A. Subduction factory: 4. Depth-dependent flux of H2O from subducting slabs worldwide. J. Geophys. Res. Solid Earth 2011, 116, 1–15. [Google Scholar] [CrossRef]
  140. Lagabrielle, Y.; Guivel, C.; Maury, R.C.; Bourgois, J.; Fourcade, S.; Martin, H. Magmatic–tectonic effects of high thermal regime at the site of active ridge subduction: The Chile triple junction model. Tectonophysics 2000, 326, 255–268. [Google Scholar] [CrossRef]
Figure 2. Geological map of the Chaoman Forest Farm polymetallic exploration area in the northeastern part of the Erguna Block.
Figure 2. Geological map of the Chaoman Forest Farm polymetallic exploration area in the northeastern part of the Erguna Block.
Minerals 14 00725 g002
Figure 3. Representative photomicrographs of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block; (a) ZK21001-55 Syenogranite; (b) ZK21001-59 Syenogranite; (c) ZK0401-9 Monzonite; (d) ZK0401-1 Quartz diorite porphyry; (e) ZK0401-3 Quartz diorite porphyry; (f) ZK21001-43 Quartz monzonite porphyry; Bi = biotite; Amp = amphibole; Kfs = K-feldspar; Pl = pla-gioclase; Qtz = quartz.
Figure 3. Representative photomicrographs of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block; (a) ZK21001-55 Syenogranite; (b) ZK21001-59 Syenogranite; (c) ZK0401-9 Monzonite; (d) ZK0401-1 Quartz diorite porphyry; (e) ZK0401-3 Quartz diorite porphyry; (f) ZK21001-43 Quartz monzonite porphyry; Bi = biotite; Amp = amphibole; Kfs = K-feldspar; Pl = pla-gioclase; Qtz = quartz.
Minerals 14 00725 g003
Figure 4. CL images of zircon grains of the Chaoman Forest Farm Mesozoic magmatic rocks. Solid and dashed circles indicate the locations of U–Pb dating and Hf isotope analyses, respectively (ah).
Figure 4. CL images of zircon grains of the Chaoman Forest Farm Mesozoic magmatic rocks. Solid and dashed circles indicate the locations of U–Pb dating and Hf isotope analyses, respectively (ah).
Minerals 14 00725 g004
Figure 5. U–Pb concordia diagrams of the Chaoman Forest Farm Mesozoic magmatic rocks. 190 ± 1.4 Ma (MSWD = 0.03) (a); 190 ± 1.2 Ma (MSWD = 0.28) (b); 189 ± 0.94 Ma (MSWD = 0.2) (c); 146 ± 1.5 Ma (MSWD = 0.08) (d); 144 ± 1.2 Ma (MSWD = 0.08) (e); 150 ± 1.5 Ma (MSWD = 0.02) (f); 143 ± 0.9 Ma (MSWD = 0.1) (g); 190 ± 3.7 Ma (MSWD = 1.3) (h).
Figure 5. U–Pb concordia diagrams of the Chaoman Forest Farm Mesozoic magmatic rocks. 190 ± 1.4 Ma (MSWD = 0.03) (a); 190 ± 1.2 Ma (MSWD = 0.28) (b); 189 ± 0.94 Ma (MSWD = 0.2) (c); 146 ± 1.5 Ma (MSWD = 0.08) (d); 144 ± 1.2 Ma (MSWD = 0.08) (e); 150 ± 1.5 Ma (MSWD = 0.02) (f); 143 ± 0.9 Ma (MSWD = 0.1) (g); 190 ± 3.7 Ma (MSWD = 1.3) (h).
Minerals 14 00725 g005
Figure 6. (a) Total alkalis vs. silica diagram [63]; (b) K2O vs. SiO2 diagram [64]; (c) A/NK vs. A/CNK diagram [65]; (d) Na2O + K2O vs. 10,000 Ga/Al discrimination diagram [68]. Data sources: Lower crust-derived adakites in the Lhasa terrane and Subducted oceanic crust-derived adakites in the Lhasa terrane [69].
Figure 6. (a) Total alkalis vs. silica diagram [63]; (b) K2O vs. SiO2 diagram [64]; (c) A/NK vs. A/CNK diagram [65]; (d) Na2O + K2O vs. 10,000 Ga/Al discrimination diagram [68]. Data sources: Lower crust-derived adakites in the Lhasa terrane and Subducted oceanic crust-derived adakites in the Lhasa terrane [69].
Minerals 14 00725 g006
Figure 7. (a) Sr/Y vs. Y diagrams [66] and (b) diagram of batch-melting modeling of chondrite-normalized (La/Yb)N ratios vs. (Yb)N [70], where N means normalized to chondrite [71]. An Eastern Pontides gabbro (G518) [72] is used as the source rock for the REE modeling under amphibolite and eclogite conditions, with varying garnet contents and respective partition coefficients (I–VI).
Figure 7. (a) Sr/Y vs. Y diagrams [66] and (b) diagram of batch-melting modeling of chondrite-normalized (La/Yb)N ratios vs. (Yb)N [70], where N means normalized to chondrite [71]. An Eastern Pontides gabbro (G518) [72] is used as the source rock for the REE modeling under amphibolite and eclogite conditions, with varying garnet contents and respective partition coefficients (I–VI).
Minerals 14 00725 g007
Figure 8. Chondrite-normalized REE pattern (a,c) [73] and spider diagrams (b,d) [71] for the Chaoman Forest Farm Mesozoic magmatic rocks.
Figure 8. Chondrite-normalized REE pattern (a,c) [73] and spider diagrams (b,d) [71] for the Chaoman Forest Farm Mesozoic magmatic rocks.
Minerals 14 00725 g008
Figure 9. (a) Plot of zircon εHf(t) values vs. U–Pb ages, (b) diagrams of εNd(t) vs. (87Sr/86Sr)i for the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: Taipingchuan igneous rocks [74]; Fukeshan igneous rocks [8]; Wunugetushan igneous rocks [19,75]; Badaguan igneous rocks [15]. The fields for the Erguna Block are from Deng et al., (2019a) [8]. The fields for MORB (Mid-Oceanic Ridge Basalt), OIB (Ocean Island Basalt) and IAB (Island Arc Basalt) are from Vervoort et al., (1999) [76]. EMI and EMII represent two types of mantle end-members [77]. The new continental crust (island arc) evolutionary line is defined by isotopic growth from 176Hf/177Hf = 0.279703 at 4.55 Ga to 0.283145 at present, with 176Lu/177Hf = 0.0375 [78].
Figure 9. (a) Plot of zircon εHf(t) values vs. U–Pb ages, (b) diagrams of εNd(t) vs. (87Sr/86Sr)i for the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: Taipingchuan igneous rocks [74]; Fukeshan igneous rocks [8]; Wunugetushan igneous rocks [19,75]; Badaguan igneous rocks [15]. The fields for the Erguna Block are from Deng et al., (2019a) [8]. The fields for MORB (Mid-Oceanic Ridge Basalt), OIB (Ocean Island Basalt) and IAB (Island Arc Basalt) are from Vervoort et al., (1999) [76]. EMI and EMII represent two types of mantle end-members [77]. The new continental crust (island arc) evolutionary line is defined by isotopic growth from 176Hf/177Hf = 0.279703 at 4.55 Ga to 0.283145 at present, with 176Lu/177Hf = 0.0375 [78].
Minerals 14 00725 g009
Figure 10. (a) (207Pb/204Pb)i vs. (206Pb/204Pb)i and (b) (208Pb/204Pb)i vs. (206Pb/204Pb)i for the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: subducted oceanic slab-derived adakites and MORB [80]; Northern Hemisphere Reference Line (NHRL) [79]; mantle source reservoirs BSE, DMM, EM I and EM II [81].
Figure 10. (a) (207Pb/204Pb)i vs. (206Pb/204Pb)i and (b) (208Pb/204Pb)i vs. (206Pb/204Pb)i for the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: subducted oceanic slab-derived adakites and MORB [80]; Northern Hemisphere Reference Line (NHRL) [79]; mantle source reservoirs BSE, DMM, EM I and EM II [81].
Minerals 14 00725 g010
Figure 11. Discrimination diagrams for the Chaoman Forest Farm Mesozoic magmatic rocks. (a) MgO vs. SiO2 diagram [98]. Data for metabasaltic and eclogite experimental melts (1–4 GPa) are from Rapp et al., (1999) and references therein [96]; (b) Mg# vs. SiO2 diagram [84]. Mantle AFC curves are after Rapp et al., (1999) (Curve 1); the proportion of assimilated peridotite is also shown. The crustal AFC curve is after Stern and Kilian (1996) (Curve 2) [95]; (c) Ni vs. SiO2 diagram [98]; (d) Cr vs. SiO2 diagram [98]; (e) Ni versus Cr diagram [69]; (f) Th/La versus Th diagram. The data for upper continental crust are from Plank (2005) and references therein [99]. The data for marine sediments are from Plank and Langmuir (1998) and for MORB are from Niu and Batiza (1997) [100,101].
Figure 11. Discrimination diagrams for the Chaoman Forest Farm Mesozoic magmatic rocks. (a) MgO vs. SiO2 diagram [98]. Data for metabasaltic and eclogite experimental melts (1–4 GPa) are from Rapp et al., (1999) and references therein [96]; (b) Mg# vs. SiO2 diagram [84]. Mantle AFC curves are after Rapp et al., (1999) (Curve 1); the proportion of assimilated peridotite is also shown. The crustal AFC curve is after Stern and Kilian (1996) (Curve 2) [95]; (c) Ni vs. SiO2 diagram [98]; (d) Cr vs. SiO2 diagram [98]; (e) Ni versus Cr diagram [69]; (f) Th/La versus Th diagram. The data for upper continental crust are from Plank (2005) and references therein [99]. The data for marine sediments are from Plank and Langmuir (1998) and for MORB are from Niu and Batiza (1997) [100,101].
Minerals 14 00725 g011
Figure 13. (a) Nb vs. Y and (b) Rb vs. (Y + Nb) diagrams [127]. Abbreviations: WPG: Within-plate granitoid; VAG: volcanic arc granitoid; Syn-COLG: syn-collision granitoid; ORG: ocean ridge granitoid.
Figure 13. (a) Nb vs. Y and (b) Rb vs. (Y + Nb) diagrams [127]. Abbreviations: WPG: Within-plate granitoid; VAG: volcanic arc granitoid; Syn-COLG: syn-collision granitoid; ORG: ocean ridge granitoid.
Minerals 14 00725 g013
Figure 14. Conceptual diagram illustrating the proposed tectonic model and magma genesis of the Εarly Jurassic–late Early Cretaceous Mesozoic magmatic rocks in the Erguna Block. (a) Εarly Jurassic, (b) Middle Jurassic–early Cretaceous.
Figure 14. Conceptual diagram illustrating the proposed tectonic model and magma genesis of the Εarly Jurassic–late Early Cretaceous Mesozoic magmatic rocks in the Erguna Block. (a) Εarly Jurassic, (b) Middle Jurassic–early Cretaceous.
Minerals 14 00725 g014
Table 1. Lithologies, major mineral assemblages, and geochemical features of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block.
Table 1. Lithologies, major mineral assemblages, and geochemical features of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block.
SampleRock TypeAge ± 1σ Mineral AssemblageSiO2MgOMg#εHf (t)(87Sr/86Sr)iεNd (t)
ZK21001-53Syenogranite190 ± 1.2 MaQtz + Kfs + Pl + Bi69.310.5734−4.2~−0.70.7085−8.5
ZK21001-55Syenogranite Qtz + Kfs + Pl + Bi68.710.5332
ZK21001-56Syenogranite Qtz + Kfs + Pl + Bi70.370.5239
ZK21001-57Syenogranite Qtz + Kfs + Pl + Bi68.290.5535
ZK21001-59Syenogranite189 ± 0.9 MaQtz + Kfs + Pl + Bi70.270.5134−5.3~0.90.7086−8.5
ZK0401-9-2Monzonite Pl + Kfs + Amp56.474.1556
ZK0401-10-1Monzonite146 ± 1.5 MaPl + Kfs + Amp56.534.1255−0.2~3.7
ZK0401-13-2Monzonite144 ± 1.2 MaPl + Kfs + Amp55.783.9254−0.1~3.40.70658−3.4
ZK0401-13-3Monzonite Pl + Kfs + Amp55.673.8455
ZK0401-14Monzonite Pl + Kfs + Amp56.443.8654
ZK0401-15-2Monzonite Pl + Kfs + Amp58.962.8052
ZK0401-16-1Monzonite Pl + Kfs + Amp57.143.8655
ZK0401-17Monzonite150 ± 1.5 MaPl + Kfs + Amp57.683.87550.4~3.00.70656−3.5
ZK0401-1Quartz diorite porphyry Qtz + Pl + Amp63.901.9352
ZK0401-2Quartz diorite porphyry Qtz + Pl + Amp64.311.8951
ZK0401-3Quartz diorite porphyry143 ± 0.9 MaQtz +Pl + Amp58.243.74562.3~5.10.70617−1.1
ZK21001-43Quartz monzonite porphyry143 ± 3.7 MaQtz + Pl + Kfs64.891.10380.7~3.50.705280.0
ZK21001-44Quartz monzonite porphyry Qtz + Pl + Kfs65.191.1038
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Y.; Zhao, Y.; Shui, X.; Tao, Z. Petrogenesis of the Early Jurassic–Early Cretaceous Adakite-like Rocks in the Erguna Block, NE China: Implications for the Tectonic Evolution of the Mongol–Okhotsk Ocean. Minerals 2024, 14, 725. https://doi.org/10.3390/min14070725

AMA Style

Wang Y, Zhao Y, Shui X, Tao Z. Petrogenesis of the Early Jurassic–Early Cretaceous Adakite-like Rocks in the Erguna Block, NE China: Implications for the Tectonic Evolution of the Mongol–Okhotsk Ocean. Minerals. 2024; 14(7):725. https://doi.org/10.3390/min14070725

Chicago/Turabian Style

Wang, Yuanchao, Yuanyi Zhao, Xinfang Shui, and Zaili Tao. 2024. "Petrogenesis of the Early Jurassic–Early Cretaceous Adakite-like Rocks in the Erguna Block, NE China: Implications for the Tectonic Evolution of the Mongol–Okhotsk Ocean" Minerals 14, no. 7: 725. https://doi.org/10.3390/min14070725

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

Article Metrics

Back to TopTop