Next Article in Journal
Secondary Minerals from Minothem Environments in Fragnè Mine (Turin, Italy): Preliminary Results
Next Article in Special Issue
Two-Stage Orogenic Cycle of the Eastern Paleo-Asian Ocean from Early Palaeozoic to Early Triassic: Constraints from Magmatic Rocks of the Southeastern Central Asian Orogenic Belt
Previous Article in Journal
Research on the Stability Mechanism of the Surrounding Rock of Gob-Side Entry Retaining by Roof Cutting in Dianping Coal Mine
Previous Article in Special Issue
Investigation of the 176Yb Interference Correction during Determination of the 176Hf/177Hf Ratio by Laser Ablation and Solution Analysis on the Neoma MC-ICP-MS
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geochronology, Geochemistry, and Lu-Hf Isotopic Compositions of Monzogranite Intrusion from the Chang’anpu Mo Deposit, NE China: Implications for Tectonic Setting and Mineralization

1
College of Earth Sciences, Jilin University, Changchun 130061, China
2
Jilin Nonferrous Metal Geological Exploration Bureau, Team 607, Jilin 132000, China
3
Xinjiang Research Centre for Mineral Resources, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi 830011, China
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(8), 967; https://doi.org/10.3390/min12080967
Submission received: 20 June 2022 / Revised: 26 July 2022 / Accepted: 27 July 2022 / Published: 29 July 2022

Abstract

:
The Chang’anpu Molybdenum deposit occurs in the monzogranite intrusions in the Lesser Khingan Mountains-Zhangguangcai Mountains metallogenic belt. Previous work focused on the study of deposits, including geological characteristics, mineralization time, S-Pb isotope, etc. However, systematic petrogeochemical study of monzogranite intrusion and comparative analysis with other porphyry deposits in the region are lacking. Three monzogranite dating samples yield LA-ICP-MS zircon weighted mean 206Pb/238U ages of 174.7 ± 1.3 Ma, 174.9 ± 1.4 Ma, and 174.3 ± 1.8 Ma, respectively, indicating that the magmatism occurred in the middle Jurassic of Mesozoic. The 14 monzogranite samples show alkali rich and relatively high silica content (up to 84.39%) with the differentiation index (DI) ranges from 86 to 96, showing that monzogranite have been subjected to fractional crystallization during its evolution; the depletion of Ba, Sr, P, Nb, Ti, and Eu also indicates that the rock has undergone crystallization fractionation, the monzogranite belong to the highly fractionated I-type. Positive εHf(t) values (6.72–8.85) and young TDM2 (551–673 Ma) of the monzogranite indicate that the formation of Chang’anpu monzogranite intrusion is related to the partial melting of juvenile lower crust, originated from the Mesoproterozoic depleted mantle. The magmatism and related Mo mineralization in the Chang’anpu deposit occurred in an active continental margin setting associated with westward subduction of the Paleo-Pacific plate beneath the Eurasian plate.

1. Introduction

Porphyry deposits have become the main source of copper and molybdenum resources in the world, and the main target of exploration due to their large reserves and wide distribution. Globally, economically important porphyry molybdenum deposits are mainly distributed in the Alpine fold belt, such as the Western Cordillera in North America, British Columbia in Canada, Southeast Alaska, and the Western United States. The three largest porphyry molybdenum deposits in the world, Climax, Urad-Hendeson, and Mt.Emmons, all occur in the Colorado molybdenum belt. Secondly, the Circum-Pacific metallogenic belt and Central Asian Orogenic Belt (CAOB) [1,2,3,4,5,6]. In China, porphyry molybdenum deposits are mainly distributed in East Qinling, Lesser Khingan Mountains-Zhangguangcai Mountains and Yanliao molybdenum belts. The molybdenum deposits in China are mainly the products of the late Indosinian to Yanshanian magmatism, which is different from the late Cretaceo-Oligocene molybdenum deposits in foreign countries [7,8,9]. The study of geochronology and petrogeochemistry of intrusions which the porphyry deposits occurred in is not only helpful to understand the process of diagenesis and mineralization [10,11,12,13,14,15], but can also reflect the tectonic evolution process controlling the formation of intrusions and ore deposits [16,17,18,19,20].
Most porphyry deposits in both oceanic (e.g., Philippines, Indonesia) and continental (Andes) magmatic arcs, as well as accretionary belts of different age, are related one way or another to adakitic (high Sr/Y) magmatic rocks. These mainly include: Aginsky, Asachinsky, Ametistovy epithermal gold deposits in Cenozoic Kamchatka arc, Russian Far East; Santo Tomas II, Didipio porphyry-type copper-gold deposits in the Philippines; Kelian epithermal gold deposit in Indonesia; Escondida, Chuquicamata copper porphyry cluster in Andean magmatic arc; Pebble porphyry Cu–Au–Mo deposit in Southwestern Alaska, belong to the Circum-Pacific magmatic arcs; Erdenet and Oyu Tolgoi Cu–Au porphyry systems in Mongolia; and Bystrinsky Cu–Fe–Mo–Au deposit in the Trans-Baikal region, Central Asian Orogenic Belt [21]. Several porphyry molybdenum deposits, including Dongfengbeishan, Daheishan, Xingshan, Dashihe, Fu’anpu, Jidetun, Houdaomu, Liushengdian, Huojihe, Culing, and Luming etc., have been exploited in the Lesser Khingan Mountains-Zhangguangcai Mountains in NE China in recent years, making it an important Mo province in China [22,23,24,25,26,27,28,29,30,31,32,33] (Figure 1, Table 1). The age of the ore-forming rocks is mainly concentrated in the early-middle Jurassic due to the extensive development of Yanshanian fault structures and granitoid intrusions in Eastern China, such as Daheishan (granodiorite porphyry, 166.6 ± 4 Ma), Fu’anpu (porphyritic monzogranite, 166.9 ± 6.7 Ma), Jidetun (granodiorite, 180.2 ± 0.8), Xingshan (granodiorite porphyry, 171.7 ± 2.2 Ma), Huojihe (granodiorite, 181.0 ± 1.9 Ma), Luming (monzogranite porphyry, 183.2 ± 1.9 Ma), and Cuiling (quratz monzonite, 178.0 ± 0.7 Ma) (Figure 1, Table 1). The Cretaceous deposits are relatively few, represented by Xiaodonggou (138 ± 3 Ma). Through study of the geochemical characteristics of porphyry, metallogenic porphyry is mainly granodiorite, quartz monzonite, and quartz diorite, with a small amount of granite [34,35]. On the whole, the ore-forming rocks are mainly medium-acid calc-alkaline rocks, while the basic rocks and granites with high acidity are rarely reported. In terms of tectonic environment, the ore-forming porphyry bodies in a continental-margin-arc environment are mainly calc-alkaline series, while in an island-arc and orogenic belt they are characterized by high potassium, which can be divided into high-K calc-alkaline series and shoshonite series; the lithology is mainly granodiorite and monzogranite [24,25,26,27,28].
The Chang’anpu deposit is a newly discovered porphyry molybdenum deposit in Lesser Khingan Mountains-Zhangguangcai Mountains. Previous work focused on the study of this and similar deposits in the region, including geological characteristics, mineralization time, S-Pb isotope, etc., [28,36,37,38,39,40]; however, systematic petrogeochemical study of monzogranite intrusion and comparative analysis with other porphyry deposits in the region are lacking. In this study, we conducted detailed research of the monzogranite, the host rock of the Chang’anpu deposit, including U-Pb dating, geochemical analyses, and Hf isotopic compositional analyses, to systematically elucidate the geological age, petrogenetic type, magma source, and evolution of Chang’anpu monzogranite, as well as tectonic setting during the formation of regional large-scale metallogenic porphyry.

2. Regional Geological Setting

The Central Asian Orogenic Belt (CAOB) is the largest accumulation of accreted terranes (island arcs, back-arc basins, oceanic plateaus and islands, micro-continents, etc.) on our planet, located between the East European and Siberian Cratons, and the North China-Tarim Craton. It was formed by the merging of different Paleozoic terranes under the progressive subduction of the Paleo-Asian Ocean plate [41,42,43,44,45]. Characteristics of the CAOB include the presence of juvenile crust and granites with relatively young Nd model ages [42,44,46,47,48,49]. The CAOB is richly endowed with Cu–Au–Mo porphyry and epithermal Au deposits of various size from the Paleozoic to Mesozoic age, many of which are associated with adakite magmatism [21,50,51,52]. The Porphyry deposits in CAOB are widely distributed, starting from Russia’s Ural Mountains in the west, through Central Asia, China’s Xinjiang and Inner Mongolia, to NE China in the east, spanning thousands of kilometers, and almost running through the whole Eurasia continent [32,53,54]. Most of these deposits in the CAOB were formed in subduction-accretion complex after the late Neoproterozoic. NE China, located at the eastern end of CAOB, was mainly formed during the Neoproterozoic to Late Paleozoic age. Its formation is related to the subduction of Paleo-Asian plate, the collision between the Siberian craton and North China Craton (NCC), and the amalgamation of microcontinental terranes, or massifs, such as Erguna, Hingan, Songliao, and Jimusi-Khanka (from west to east) [55,56,57,58,59,60] (Figure 1A). NE China has been modified and overprinted by the Paleo-Asian oceanic and the circum-Pacific tectonic system [61]; multiphasic tectonism had produced favorable space for the formation of porphyry intrusions and the clustering of porphyry deposits.
The Lesser Khingan Mountains-Zhangguangcai Mountains metallogenic belt belongs to the Songnen Massif, the whole metallogenic belt extend from Heilongjiang in the north, to the Jilin in the south in an arc shape of NNW-NS, with a total length of nearly 700 km and a maximum width of up to 100 km (Figure 1A).
The area has been subjected to intense sedimentation, folding and metamorphism at the Precambrian-Phanerozoic transition, followed by large-scale Paleozoic gabbro- granodiorite- granite magmatism, with predominant high-K calc-alkaline and shoshonitic chemistry [62]. During Mesozoic, it was involved in the development of multiple subduction zones and active continental margins related to the closure of the Mongol-Okhotsk Ocean, and the subduction of the Paleo-Pacific Ocean, and was marked by a calc-alkaline to shoshonitic volcanic series and associated granitoids [47,63,64,65,66].
The mining area has relatively simple lithology and undeveloped strata, all of which are Mesozoic granite intrusions; from west to east there are plagiogranite, monzogranite, and alkali-feldspar granite, and the Chang’anpu deposit is wholly hosted in monzogranite (Figure 2A). The orebody is stratified in space, with an E-W direction (Figure 2A,B). The ore is mainly disseminated. Metallic minerals are molybdenite, pyrite and a small amount of chalcopyrite, pyrrhotite, galena, and sphalerite. Faults that develop in this region are mainly compressional faults; the tectonic framework in this area is composed of the Xinan-Ermu fault in a NW direction and the Huoshidingzi- Chang’anpu fault in a NE direction. The Chang’anpu deposit is at the intersection of the two; fracture zone, argillic alteration and structural breccia are common in the deposit. An EW compressional fault, F1, has developed in the deposit, which is about 2 km long and 100 m to 200 m wide. It consists of a series of parallel fault zones, including fracture zone, gouge, scratch, tectonic breccia, and later, intrusive quartz vein. The F1 fault is a secondary structure of the NW-trending Xin’an-Ermu fault, and is the main rock-ore-controlling tectonic in this deposit, controlling the spatial morphology and distribution of ore bodies [36,39,40].

3. Petrological Characteristics

As the deposit is fully hosted in monzogranite intrusion, and the spatiotemporal distribution of monzogranite bodies has a large range, the samples of monzogranite in different locations and layers were selected to conduct geochronology and geochemical analyses. Three U–Pb dating samples and 14 geochemical samples were taken from positions near the surface (CAP-N1), deep through the ore body (CAP-N2, taken from 650 m of drill hole ZK10-3), and far away from the deposit (CAP-N3); the sample (CAP-N1) was used for Hf isotopic analysis. All samples were selected from unaltered or slightly altered monzogranite. Tthe sampling locations are shown in Figure 2A.
Monzogranite show a grayish-red, granular texture, and massive structure. The overall modal composition is: K-feldspars, xenomorphic granular textures, particle size 1.0–2.5 mm, content of about 30%–35% (in volume); Plagioclase grains, euhedral to subhedral, are partially altered to sericite, particle size 1.0–3.0 mm, content 25%–30%; quartz, anhedral granular texture, particle size 0.5–2.5 mm, content 25%–30%; Biotite is about 5%, partially altered to chlorite (Figure 3). The content of K-feldspars in the surrounding rock of the orebody was increased, and some parts of the rock were broken and re-cemented under the tectonic stress, which is of brecciform structure with quartz veins inserted into it.

4. Analytical Methods

4.1. Major and Trace Element Determinations

The samples were analyzed in Hebei Institute of Geology and Mineral Resources, Langfang, China. Before testing, all the samples were pulverized into granules smaller than 200 mesh in agate mortars. The concentration of major oxides in the samples was analyzed by X-ray Fluorescence spectrometer (XRF) except FeO (wet chemical method). Inductively coupled plasma-mass spectrometry (ICP-MS) was selected to measure trace elements. Analysis precision of major and trace elements was better than 5% and 10%, respectively.

4.2. Zircon U–Pb Dating

After crushing the collected samples, conventional methods using flotation and magnetic separation were used. Samples were selected with binoculars, looking for a crystal shape and transparency, and being free of cracks of zircon grains. Using epoxy resin, and after grinding and polishing, reflected light and cathodoluminescence (CL) of the image was acquired. The transmitted and reflected light images of zircon were collected at the Comprehensive Identification Laboratory of Rock and Mineral Resources, College of Earth Sciences, Jilin University, while CL images and LA-ICP-MS U-Pb isotope analysis of zircon were collected at MLR Key Laboratory of Mineral Resources Evaluation in Northeast Asia, Jilin University, Changchun, China. Through the reflected light and CL images of zircon, the regions without cracks and inclusions, and with uniform absorption, were selected and analyzed using standard determination procedures on an Agilent 7500AICP-MS instrument. In the experiment, high purity He was used as the carrier gas of the denuded material; the standard reference substance NISTSRM610 of synthetic silicate glass, developed by the National Institute of Standards and Technology, was used to optimize the instrument. Zircon standard samples were Plesovice (age 337 ± 0.37 Ma) [67]. NIST SRM 612 was used as the component standard, and 29Si was used as the internal standard element. The laser spot diameter was 32 μm and data processing was performed using ICPMSDataCal [68,69]; general Pb calibration was carried out by Andersen (2002) [70]; age calculation harmony and plot drawing were completed by international standard program Isoplot (v3.0) [71]; the given isotope ratio and age error were all at the level of 1σ.

4.3. Lu–Hf Isotopic Analysis

Based on LA-ICP-MS zircon U-Pb dating, Hf isotopic compositions of zircon microregions were determined. In situ Lu-Hf isotopic analysis of zircons was performed, using laser-denuded multi-receiver plasma mass spectrometry (LA-MC-ICP-MS) at Northwestern University, Xi’an, China. MC-ICP-MS is all Neptune. For the 193 nm laser, helium gas was used as the carrier gas under given instrument conditions to improve signal sensitivity. In the single point ablation mode, the patches were fixed to 60 μm and 44 μm, respectively. For detailed instrument operating conditions and analysis methods, refer to Wu et al. (2006) [72]. In εHf (t) and model age calculation, the present 176Lu/177Hf and 176Hf/177Hf of the chondritic, and depleted mantle, are 0.0332, 0.282772 and 0.0384, 0.28325, respectively [73]. When calculating the two-stage model age (TDM2), 176Lu/177Hf = 0.015 was taken as the mean continental crust [74]. All zircons had low 176Lu/177Hf ratios, almost all of which were less than 0.002, indicating that zircons have low accumulation of radiological Hf after their formation. Offline processing of the analytical data, including sample and blank signal selection and isotope mass fractionation correction, was performed using software ICPMSDataCal [69].

5. Results

5.1. Major and Trace Element Geochemistry

The major and trace element contents of the monzogranite in Chang’anpu deposit are shown on Table 2, and plotted in Figure 4. According to the CIPW-NORM ratio of quartz to feldspar, the lithology is mainly monzogranite. The results of sample analysis showed high concentrations of SiO2 (69.50–84.39 wt.%), Al2O3 (6.35–15.36 wt.%), and Na2O + K2O (4.39–9.81 wt.%), and relatively low CaO (0.47–2.00 wt.%), MgO (0.13–0.67 wt.%), TiO2 (0.06–0.37 wt.%), and P2O5 (0.01–0.15 wt.%) concentrations. The trace elements composition of mineralized monzonite is shown in Table 2; the REE content varies widely from 75.97 to 145.83 ppm, with a strong light rare earth element (LREE)/heavy rare earth element (HREE) fractionation ((La/Yb)N = 13.44–22.54). The chondrite-normalized REE patterns of monzogranite show a negative Eu anomaly (Eu/Eu* = 0.39–0.92) and LRRE enrichment over the HREE (Figure 4A). In the primitive-mantle-normalized trace elements diagram (Figure 4B), the monzogranite is enriched in Rb, Th, and K and relatively depleted in Nb, Ba, P, Sr, and Ti.

5.2. Zircon U–Pb Geochronology

Zircons from samples CAP-1, CAP-2, and CAP-3 are colorless or light brown, transparent and euhedral. Most of the zircons are 80–110 μm in length, with an aspect ratio of 1:1–3:1. Zircons generally have low CL brightness, obvious oscillation zone and dark rims (Figure 5), combined with a high Th/U ratio (0.38–1.25; Table 3), indicating a magmatic origin. The zircon U-Pb dating results are shown in Table 3. Except for individual data, most of the analysis results were nearly concordant (Figure 5). The monzogranite sample CAP-1, CAP-2, and CAP-3 yielded 206Pb/238U ages 172 to 178 Ma, 174 to 177 Ma, 165 to 181 Ma, with a weighted mean 206Pb/238U age 174.7 ± 1.3 Ma (n = 16, mean squared weighted deviation (MSWD) = 0.56); 174.9 ± 1.4 Ma (n = 12, MSWD = 0.19); 174.3 ± 1.8 Ma (n = 17, MSWD = 0.71), respectively (Figure 5). These zircon U–Pb ages of monzogranite are consistent with the molybdenite Re-Os isochron age of 167.3–168.7 Ma obtained by Song et al. (2016) [76]. The emplacement age of intrusions and the formation age of mineralization are close to each other and their sequence is reasonable, which suggest a causal relationship between mineralization and monzogranite intrusion.

5.3. Zircon Hf Isotopic

The Hf isotopic composition of 15 zircons from the CAP-1 of Chang’anpu monzogranite were analyzed at the same sites, with the U–Pb dating. The data were presented in Table 4. The initial 176Hf/177Hf ratio was corrected by the U–Pb age previously obtained. 176Lu/177Hf ratios of zircon range from 0.000520–0.000984. εHf(t) values range from +6.72 to +8.85 (t = 171 Ma); the average is +7.60. Two-stage Hf model ages (TDM2) of zircons range from 551 to 673 Ma. All analysis results were plotted between the evolutionary reference lines of the Chondrite Uniform Reservoir (CHUR) and depleted mantle in the εHf(t) vs. age diagram (Figure 6; Table 3).

6. Discussion

6.1. Petrogenesis and Sources of the Monzogranite Intrusion

6.1.1. Genetic Type of the Monzogranite

The genetic types of granitic can reflect the tectonic setting and source of magmatic rocks [78,79,80,81], which can genetically be divided into I-, S-, A-, and M- types [82,83,84,85]. The Chang’anpu monzogranite can be excluded as M-type granite, firstly because M-type granite has high Na/K (>1) and is mainly composed of plagioclase [42]. Many studies have shown that the content of P2O5 is negatively correlated with SiO2 in I-type, while P2O5 increases or remains unchanged with the increase of SiO2 in S-type [86,87]. The P2O5 content of monzogranite in Chang’anpu deposit ranges from 0.01% to 0.15%, and is negatively correlated with SiO2 (Figure 7G). In general, A/CNK = 1.1 can be used as a boundary between S-type and I-type, while peraluminous granites are generally believed to be S-type [82]. The Chang’anpu monzogranite were weaky peraluminous with A/CNK < 1.1 (0.96–1.07). These two points show that the monzogranite is not S-type. Generally, alkaline dark minerals are common in A-type granite, which are enriched in REE, Nb, Y, and other high field strength elements (HFSE), while depleted in Eu, Ba, and Sr. However, the content of mafic alkaline minerals in monzogranite is very low, and Nb, Ta, Ti, and Y elements are not enriched. Furthermore, the content of (Zr + Nb + Ce + Y) in samples ranged from 112.6 to 280.8 ppm, which was lower than that of A-type (≥350 ppm). Thus, Chang’anpu monzogranite is not an A-type, but an I-type.
Except for the N3-Q3 sample with abnormally high SiO2 content (84.39%) due to silicification alteration, the SiO2 in Figure 7A–G shows a linear correlation with other elements, indicating that the magma contains a certain content of differentiated minerals, such as feldspar and hornblende [88]. The data obtained from monzogranite samples show alkali rich and relatively high silica content (maximum 84.39%, most 69.50%–75.42%), with a high differentiation index (86–96) and the depletion of P, Nb, Ti and Eu elements in the samples indicating that the rock has undergone crystallization fractionation during magma evolution. The fractional crystallization of plagioclase and K-feldspar causes negative Eu anomalies, while the separation of Ti-bearing minerals and apatite usually results in the depletion of Nb, Ti, and P. It is well known that in the process of magmatic fractionation crystallization, Ba and Sr are enriched in plagioclase, while Rb tends to residual magma, making it possible to trace the fractional crystallization by Rb/Ba and Rb/Sr ratios [42]. The contents of Sr and Ba show they are negatively correlated with SiO2, whereas Rb/Ba and Rb/Sr are positively correlated with SiO2 (Figure 7H,I), indicating that monzogranite was formed by fractional crystallization. Data from samples in the Th/Ti vs. Th diagram define a curving array [89] (Figure 8A), and are distributed along the fractional crystallization line in the La/Sm vs. La diagram (Figure 8B). Fractional crystallization and partial melting may be regarded as antipathetic and seemingly reversible processes, their manifestation in the Earth’s crust leads to markedly different outcomes; large volumes of peraluminous melt is formed during partial melting, while limited volumes of peraluminous melt is formed during fractional crystallization. Although fractional crystallization was the dominant controlling factor for the evolution of monzogranite, the existence of a large amount of peraluminous I-type melt from Archaean through to the present indicates that partial melting plays an important role in the long-term evolution of the crust [90]. In SiO2 vs. K2O diagram (Figure 8C), samples plot in high-K calc-alkaline field and are weakly peraluminous (Figure 8D). Based on the above, it suggests that the Chang’anpu monzogranite intrusion have been subjected to significant fractional crystallization, belong to high-K calc-alkaline highly fractionated I-type granite. Monzogranite exhibited high SiO2 and Sr, relatively lower MgO, Y, Yb content and high Sr/Y ratio, and weak Eu anomaly, indicating that it has the attributes of adakite. On the Sr/Y − Y discriminant diagram (Figure 9), except for a few samples, most samples fall into the adakite area [21,91,92,93], indicating that monzogranite belongs to adakite rock.

6.1.2. Petrogenesis and Source of the Magma

As a common accessory mineral in granites, zircon is often used to identify the source of magma [68,97,98]. The value of 176Hf/177Hf varies greatly among zircons from different magmatic sources. The Lu-Hf fractionation results in high Lu/Hf ratios in mantle residue, but low Lu/Hf ratios in basaltic magmas during partial melting. The depleted mantle usually has positive radiogenic εHf(t) values, while the enriched mantle, ancient crust show negative εHf(t) values. The magma derived from the juvenile crust usually has a positive εHf(t) value as it has not had enough time to be derived from depleted mantle. To summarize, chondrites, depleted mantle, and juvenile crust have a high 176Hf/177Hf ratio, which is usually positive, while the 176Hf/177Hf ratio is relatively low, and often negative, in enriched mantle and pre-existing crust.
The monzogranite of the Chang’anpu deposit contain zircons with the 206Pb/238U ages 163 to 179 Ma, the εHf(t) values of +6.72 to +8.85, and the TDM2 ages range from 551 to 673 Ma, plot between the evolutionary reference lines of the CHUR and depleted mantle (Figure 6). However, granite derived directly from mantle can be ruled out, as few granites distributed in this area have mantle-source characteristics [23]. This indicates that lower crust materials derived from Mesoproterozoic depleted mantle, or subducted slab, may had contributed to the magmatic source. Similar to the most porphyry deposits in the Lesser Khingan Mountains–Zhangguangcai Mountains, Chang’anpu Mo deposits have δ34S values broadly consistent with those of mantle magma (0‰ ± 3‰). The Pb content is a mixture of the Pb from mantle-derived magma and the radiogenic Pb from country rocks [22,37,38]. Both of the S-Pb isotope compositions are similar to subduction zone magmatism, indicating that the diagenetic and ore-forming materials have a mixed source of crust and mantle. Wu et al. (2000) [42] pointed out that the magmatic sources of Mesozoic granites in NE China usually involve juvenile lower crustal materials that have been partial melted, and are characterized by positive εNd(t) value, low 87Sr/86Sr ratio, and younger Nd model age. Comparing Chang’anpu monzogranite with other Mesozoic granites in NE China, most of them have positive εHf(t) values and younger TDM ages, which proves that crust beneath NE China had a significant growth episode during the Meso-Neoproterozoic age, and also indicating their source was depleted mantle or subducted slab [23,24]. In conclusion, the monzogranite in Chang’anpu deposit has the same source and tectonic setting as other porphyry deposits distributed in Lesser Khingan Mountains–Zhangguangcai Mountains; the partial melting and subsequent fractional crystallization of the juvenile lower crust derived from Mesoproterozoic depleted mantle, or subducted slab, played a key role in the formation of the monzogranite intrusion.

6.2. Temporal and Spatial Distribution of Regional Metallogenic Granitic Intrusions

Zircon U-Pb dating shows that ore-forming granites in the Lesser Khingan Mountains-Zhangguangcai Mountains were mainly formed in the early-middle Jurassic-late early Cretaceous: the Luming monzogranite (197.1 ± 1.6 Ma) and monzogranite porphyry (183.2 ± 1.9 Ma) were formed in the early Jurassic, and mineralization is mainly related to late monzogranite porphyry; the Fu’anpu porphyritic monzogranite (166.9 ± 6.7 Ma) and dark microgranular enclave (quartz diorite) (168.6 ± 1.9 Ma) were formed in the middle Jurassic; both Jidetun monzogranite (170.8 ± 1.9 Ma) and granodiorite (180.2 ± 0.8 Ma) were formed in the middle Jurassic; the Baoshan granodiorite (164.4 ± 1.9 Ma) and quartz diorite (165.2 ± 1.7 Ma) were formed in the middle Jurassic; the age of the Kanchuangou granodiorite is 111.8 ± 1.4 Ma, which was formed in the late early Cretaceous; and the Chang’anpu monzogranite (174.3 ± 1.8 Ma −174.9 ± 1.4 Ma) was formed in the middle Jurassic (this study). Based on published data and the results of this study, it is found that the ore-forming granites in the early Jurassic are mainly distributed in the Lesser Hinggan Mountains and the Northern Yanbian area, and are mostly monzonite (porphyry)-granodiorite-quartz monzonite. The ore-forming granites of middle Jurassic are mainly distributed in the Zhangguangcai Mountains and Southern Yanbian, and the lithology is monzonitic-granodiorite (porphyry). Early-middle Jurassic granites are the most widely distributed and longest lasting magmatic activity in the region; most ore-forming rocks of the deposit were formed in this period. In spatial morphology, it is distributed in a NNE direction and often occurs as batholith, stock, and compound rock mass. The rock types are mainly calc-alkaline and high-K calc-alkaline monzogranite-granodiorite, which is considered to be the product of the subduction of the Pacific plate beneath the Eurasian continent [47,99].

6.3. Diagenetic and Metallogenic Model

In the early Jurassic, the Pacific Ocean plate subducted beneath the Eurasian continental plate, and continuous subduction resulted in the regional compression environment, thickening of the continental crust, partial melting of the thickened lower crust material and dehydration and melting of the subducted slab-formed mineral-rich granitic magma. Under the compression environment, magma cannot easily pass directly through the upper crust to form volcanic rocks, and can only continuously accumulate to form a larger magma chamber, which accumulates more crust-derived ore-forming materials [100]. Magma in the shallow magma chamber was unable to erupt and began fractional crystallization during evolution under pressure, forming porphyry [3,101,102]. As molybdenum is an incompatible element, the continuous evolution of crystallization and fractionation resulted in the enrichment of molybdenum in the late residual magma melt. When the magma is deeply emplaced, pressure is higher and the solubility of volatiles increases correspondingly, no boiling occurs at the beginning of emplacement, and no independent fluid phase is formed. With the slow evolution of magma to the late stage, volatiles are saturated, forming large-scale low-medium salinity supercritical fluid under high fluid pressure [103,104,105]. The ore bearing fluid is exsolved from the magma chamber, immiscibility leads to phase separation of the fluid, and boiling causes large-scale molybdenum precipitation, forming large-scale porphyry molybdenum deposits related to plutonic intrusions represented by Daheishan, Luming, Chang’anpu, Fu’anpu, and Jidetun, etc. From the perspective of the formation and evolution of porphyry molybdenum deposit, the compressive tectonic environment is more conducive to plutonic magma in forming porphyry in the shallow part, providing space and pressure conditions for the evolution of molybdenum-bearing melt, and the accumulation and precipitation of molybdenum elements.

6.4. Geodynamic Setting

Numerous Jurassic porphyry molybdenum deposits have been discovered and mined in NE China over the last few years; however, there is still some controversy overtheir tectonic setting. Some researchers attribute metallogenesis to basin-and range tectonic setting [106]; closure of the Asian ocean [107]; and the mantle plume model [108], whereas others believe that the post-orogenic extension is the metallogenic tectonic setting [109]. The samples of Chang’anpu monzogranite is characterized by enriched LILEs and incompatible elements, but relatively depleted in HFSEs, which conforms to the granitic magmas that derived from the active continental margin tectonic setting that have undergone extensive fractional crystallization during evolution. All samples plotted in the volcanic arc and syn-collision field in Nb-Y and Rb- (Y + Nb) discrimination diagrams suggesting an active continental margin setting (Figure 10A,B). Similarly, the monzogranite samples show a continental arc setting in the Zr/Al2O3 versus TiO2/Al2O3, and Ce/P2O5 versus Zr/TiO2 discrimination diagrams (Figure 10C,D). The Th/Ta ratio of sample ranges from 10.9 to 21.1 belongs to the active continental margin igneous rocks [42]. By comparing the tectonic setting of giant porphyry deposits around the world, it shows that the subduction-related island- and continental-arc settings are relatively common tectonic settings for most porphyry deposits [47]. Most of the worldwide porphyry deposits occur in active convergent plate boundaries, and Mesozoic-Cenozoic orogenic belts with subduction-related tectonic setting [3,4]. The models of adakite petrogenesis include the melting of thickened and delaminated mafic lower crust, basalt underplating of the continental crust, and high-pressure fractionation (amphibole-garnet) of mantle-derived, hydrous mafic melts [91]. As the Chang’anpu monzogranite intrusion was developed in a continental arc setting, we suggest that the adakites were related to the oceanic slab melting and then being producing through the northeastward subduction of the Paleo-Pacific ocean crust, beneath the continental terranes in Eastern China in the Jurassic.
The Eastern part of China underwent Circum-Pacific tectonics to varying degrees, and NE China suffered from the subduction of Paleo-Pacific since the Jurassic [112]. The initial subduction time of Pacific plate was determined by Dongning-Wangqing-Hunchun calc-alkaline volcanic associations, which was found in the Heilongjiang-Jilin area, with ages of 173 to 190 Ma [47]. The deformation age of Heilongjiang complex also provides key markers for the duration of the subduction of Paleo-Pacific plate in Early Jurassic. According to [113] the westward subduction of Pacific plate had already affected Lesser Khingan Mountains-Zhangguangcai Mountains in the Early Jurassic by analyzing the geochronology and geochemistry data of granitoids. The NNE direction distribution of Jurassic granitoids in the Lesser Khingan Mountains-Zhangguangcai Mountains also suggests that the early Yanshanian large-scale diagenesis and mineralization should be related to the subduction of the Pacific plate; the subduction also formed a tectonic transition period conducive to the formation of porphyry deposits [6,42,47,114]. By comparing the geotectonic background of the Lesser Khingan Mountains-Zhangguangcai Mountains in the Early-Middle Jurassic, and Early Cretaceous, it can be found that the region was mainly under the background of lithospheric compression related to the subduction of the Pacific plate in Early-Middle Jurassic. In the region, granites are distributed in a NNE direction, and often occurs as batholith, stock, and compound rock mass, with rock types mainly being calc-alkaline and high-K calc-alkaline monzogranite- granodiorite, suggesting that its formation may be related to the continuous subduction of the Pacific plate [46,99]. In addition, the molybdenum deposits formed are mainly of stock and fine-vein type, related to plutonic intrusions. The geodynamic in Early Cretaceous for this region, and in Eastern China, is that continental breakup and rapid plate motion, including Pacific plate subduction, resulted in large scale lithospheric delamination, leading to a mantle avalanche, asthenosphere upwelling, and subsequent crustal melting in an extensional setting [115], forming mainly porphyry type and hydrothermal vein type copper gold deposits. From the above diagenetic and metallogenic mechanism, plutonic magmatism under a compression environment is more conducive to the formation and evolution of porphyry bodies and the migration, accumulation and precipitation of molybdenum elements, to form molybdenum ore bodies. Therefore, most of the porphyry Mo deposits in China were formed under the compression background of the Early-Middle Jurassic. The subduction of the Pacific plate in the Early-Middle Jurassic thickened the lithosphere, which accelerated the interaction between crust-mantle materials. The ocean crust subducted to a certain depth and began to dehydrate, then removed CO2, which, as it left the ocean crust, accelerated the partial melting of the lithospheric mantle to form basaltic magma. The underplating of basaltic magma melted the crust and formed ore-bearing granitic magma, which is the reason magma in the Early-Middle Jurassic has the characteristics of crust-mantle mixed source. In conclusion, we consider that the formation of the monzogranite intrusion and Chang’anpu deposit was closely related to the westward subduction of Pacific plate beneath Eurasian Plate, in Early-Middle Jurassic. During the same geological period, the subduction-related magmatism also formed a series of non-ferrous deposits in Eastern China (Figure 11).

7. Conclusions

  • LA–ICP–MS zircon U–Pb dating reveals that the emplacement age of monzogranite intrusions in the Chang’anpu Mo deposit was 174.3 ± 1.8 to 174.9 ± 1.4 Ma. The emplacement age of intrusions, and the formation age of mineralization (168.0 ± 1.0 Ma), are close to each other and their sequence is reasonable, which suggests the causal relationship between mineralization and monzogranite intrusion.
  • The Chang’anpu monzogranite intrusion has been subjected to significant fractional crystallization, belonging to high-K calc-alkaline highly fractionated I-type granite.
  • Positive εHf(t) values (6.72–8.85) and young TDM2 (551–673 Ma) of monzogranite indicate that the formation of intrusion and deposit are related to the partial melting, and subsequent fractional crystallization, of the juvenile lower crust derived from Mesoproterozoic depleted mantle.
  • Formation of the monzogranite intrusion and Chang’anpu deposit was closely related to the magmatism triggered by the westward subduction of Pacific plate beneath Eurasian Plate, in Early-Middle Jurassic.

Author Contributions

Conceptualization, J.Z.; methodology, Y.Y.; software, W.W.; investigation, J.Z., Y.Y., S.H. and W.W.; data curation, S.H.; writing-original draft preparation, J.Z.; writing-review and editing, J.Z. and Y.Y.; visualization, W.W.; supervision, Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data provided in this study can be obtained from the Figures and Tables in the article.

Acknowledgments

The authors are most grateful to the staff of the Laboratory of Hebei Institute of Geology and Mineral Resources, Langfang, as well as the State Key Laboratory of Continental Dynamics, Department of Geology, Northwest University, during the zircon U-Pb dating, geochemical analyses and Lu-Hf isotopic analyses. The author also deeply appreciates the editor-in-chief and two anonymous reviewers for their kind and critical constructive comments and suggestions, which greatly improved the quality of our manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cooke, D.R.; Hollings, P.; Walshe, J.L. Giant porphyry deposits: Characteristics, distribution, and tectonic controls. Econ. Geol. 2005, 100, 801–818. [Google Scholar] [CrossRef]
  2. Goldfarb, R.J.; Groves, D.I.; Gardoll, S. Orogenic gold and geological time: A global synthesis. Ore Geol. Rev. 2001, 18, 1–75. [Google Scholar] [CrossRef]
  3. Sillitoe, R.H. A plate tectonic model for the origin of porphyry copper deposits. Econ. Geol. 1972, 67, 184–197. [Google Scholar] [CrossRef]
  4. Sillitoe, R.H. Characteristics and controls of the largest porphyry copper-gold and epithermal gold deposits in the circum-Pacific region. Aust. J. Earth Sci. 1997, 44, 373–388. [Google Scholar] [CrossRef]
  5. Kerrich, R.; Goldfarb, R.; Groves, D.; Garwin, S.; Jia, Y.F. The characteristics, origins and geodynamic settings of supergiant gold metallogenic provinces. Sci. China Ser. 2000, 43, 1–68. [Google Scholar] [CrossRef]
  6. Mitchell, A.H.G. Metallogenic belts and angle of dip of Benioff zones. Nature 1973, 245, 49–52. [Google Scholar] [CrossRef]
  7. He, X.H.; Deng, X.H.; Pirajno, F.; Zhang, J.; Li, C.; Chen, S.B.; Sun, H.W. The genesis of the granitic rocks associated with the Mo-mineralization at the Hongling deposit, eastern Tianshan, NW China: Constraints from geology, geochronology, geochemistry, and Sr-Nd-Hf isotopes. Ore Geol. Rev. 2022, 146, 104947. [Google Scholar] [CrossRef]
  8. Li, J.; Li, C.-Y.; Liang, J.-L.; Song, M.-C.; Zhang, L.-P.; Song, Y.-X. Mineralization of the Shangjiazhuang Mo deposit in the Jiaodong peninsula, China: Constraints from S–H–O isotopes and fluid inclusions. Solid Earth Sci. 2021, 6, 370–384. [Google Scholar] [CrossRef]
  9. Hou, X.-G.; Sun, D.-Y.; Gou, J.; Yang, D.-G. The origin of variable-delta 18O zircons in Jurassic and Cretaceous Mo-bearing granitoids in the eastern Xing-Meng Orogenic Belt, Northeast China. Int. Geol. Rev. 2019, 61, 129–149. [Google Scholar] [CrossRef]
  10. Guo, X.; Zhou, T.; Jia, Q.; Li, J.; Kong, H. Highly differentiated felsic granites linked to Mo mineralization in the East Kunlun Orogenic Belt, NW China: Constrains from geochemistry, and Sr-Nd-Hf isotopes of the Duolongqiarou porphyry Mo deposit. Ore Geol. Rev. 2022, 145, 104891. [Google Scholar] [CrossRef]
  11. Gao, Y.; Yang, Y.-C.; Han, S.-J.; Meng, F. Geochemistry of zircon and apatite from the Mo ore-forming granites in the Dabie Mo belt, East China: Implications for petrogenesis and mineralization. Ore Geol. Rev. 2020, 126, 103733. [Google Scholar] [CrossRef]
  12. Shen, P.; Hattori, K.; Pan, H.D. Oxidation Condition and Metal Fertility of Granitic Magmas: Zircon Trace-Element Data from Porphyry Cu Deposits in the Central Asian Orogenic Belt. Econ. Geol. 2016, 110, 1861–1878. [Google Scholar] [CrossRef]
  13. Mi, K.; Lü, Z.; Yan, T.; Yao, X.; Ma, Y.; Lin, C. Zircon geochronological and geochemical study of the Baogaigou Tin deposits, southern Great Xing’an Range, Northeast China: Implications for the timing of mineralization and ore genesis. Geol. J. 2020, 55, 5062–5081. [Google Scholar] [CrossRef]
  14. Mi, K.-F.; Lü, Z.-C.; Yan, T.-J.; Zhao, S.-J.; Yu, H.-Y. SHRIMP U-Pb zircon geochronology and Hf isotope analyses of Middle Permian-Early Triassic intrusions in southern Manzhouli area, Northeast China: Implications for the subduction of Mongol-Okhotsk plate beneath the Erguna massif. Int. Geol. Rev. 2020, 62, 549–567. [Google Scholar] [CrossRef]
  15. Wan, L.; Lu, C.; Zeng, Z.; Mohammed, A.S.; Liu, Z.; Dai, Q.; Chen, K. Nature and significance of the late Mesozoic granitoids in the southern Great Xing’an range, eastern Central Asian Orogenic Belt. Int. Geol. Rev. 2019, 61, 584–606. [Google Scholar] [CrossRef]
  16. Chen, P.; Liu, B.; Long, Z.; Zhou, L.; Fu, Y.; Zeng, Q. Ore genesis of the Sadaigoumen porphyry Mo deposit, North China Craton: Constraints from pyrite trace element and lead isotope analyses. Ore Geol. Rev. 2022, 142, 104698. [Google Scholar] [CrossRef]
  17. Cao, M.; Qin, K.; Li, G.; Evans, N.; Hollings, P.; Jin, L. Genesis of ilmenite-series I-type granitoids at the Baogutu reduced porphyry Cu deposit, western Junggar, NW-China. Lithos 2016, 246–247, 13–30. [Google Scholar] [CrossRef]
  18. Zhang, X.-Z.; Wang, Q.; Dong, Y.-S. High-pressure granulite facies overprinting during the exhumation of eclogites in the Bangong-Nujiang suture zone, central Tibet: Link to flat-slab subduction. Tectonics 2017, 36, 2918–2935. [Google Scholar] [CrossRef]
  19. Zhai, D.; Liu, J.; Tombros, S.; Williams-Jones, A.E. The genesis of the Hashitu porphyry molybdenum deposit, Inner Mongolia, NE China: Constraints from mineralogical, fluid inclusion, and multiple isotope (H, O, S, Mo, Pb) studies. Miner. Depos. 2018, 53, 377–397. [Google Scholar] [CrossRef] [Green Version]
  20. Zheng, S.-H.; Gu, X.-X.; Zhang, Y.-M.; Wang, J.-L.; Peng, Y.-W.; Xu, J.-C.; Lv, X. Temporal and spatial separation mechanisms of the Cu and Mo mineralization in the Dabate porphyry deposit, Western Tianshan, Xinjiang, China. Ore Geol. Rev. 2022, 146, 104924. [Google Scholar] [CrossRef]
  21. Kepezhinskas, P.; Berdnikov, N.; Kepezhinskas, N.; Konovalova, N. Adakites, high-Nb basalts and copper-gold deposits in magmatic arcs and collisional orogens: An overview. Geosciences 2022, 12, 29. [Google Scholar] [CrossRef]
  22. Guo, Y.; Zeng, Q.; Yang, J.; Guo, F.; Guo, W.; Liu, J. Zircon U-Pb geochronology and geochemistry of Early-Middle Jurassic intrusions in the Daheishan ore district, NE China: Petrogenesis and implications for Mo mineralization. J. Asian Earth Sci. 2018, 165, 59–78. [Google Scholar] [CrossRef]
  23. Zhou, L.-L.; Zeng, Q.-D.; Liu, J.-M.; Friis, H.; Zhang, Z.-L.; Duan, X.-X. Geochronology of the Xingshan molybdenum deposit, Jilin Province, NE China, and its Hf isotope significance. J. Asian Earth Sci. 2013, 75, 58–70. [Google Scholar] [CrossRef]
  24. Zhou, L.-L.; Zeng, Q.-D.; Liu, J.-M.; Friis, H.; Zhang, Z.-L.; Duan, X.-X.; Lan, T.-G. Geochronology of magmatism and mineralization of the Daheishan giant porphyry molybdenum deposit, Jilin Province, Northeast China: Constraints on ore genesis and implications for geodynamic setting. Int. Geol. Rev. 2014, 56, 929–953. [Google Scholar] [CrossRef]
  25. Ju, N.; Ren, Y.-S.; Wang, C.; Wang, H.; Zhao, H.-L.; Qu, W.-J. Ore genesis and molybdenite Re-Os dating of Dashihe molybdenum deposit in Dunhua, Jilin. Glob. Geol. 2012, 31, 68–76. [Google Scholar]
  26. Li, L.-X.; Song, Q.-H.; Wang, D.-H.; Wang, C.-H.; Qu, W.-J.; Wang, Z.-G.; Bi, S.-Y.; Yu, C. Re-Os isotopic dating of molybdenite from the Fuanpu Molybdenum deposit of Jilin Province and discussion on Its metallogenesis. Rock Miner. Anal. 2009, 28, 283–287. [Google Scholar]
  27. Lu, Z.-Q.; Li, X.-J.; Qiu, C.; Liang, B.-S. Geology, geochemistry and geochronology of ore-bearing intrusions in Jidetun molybdenum deposit in mid-east Jilin Province. Miner. Depos. 2016, 35, 349–364. [Google Scholar]
  28. Zhang, Y. Research on Characteristics of Geology, Geochemistry and Metallogenic Mechanism of the Jurassic Molybdenum Deposits in the Mid-East Area of Jilin; Jilin University: Changchun, China, 2013; pp. 1–144. [Google Scholar]
  29. Wang, Q.; Li, Z.-X.; Chung, S.-L.; Wyman, D.A.; Sun, Y.-L.; Zhao, Z.-H.; Zhu, Y.-T.; Qiu, H.-N. Late Triassic high-Mg andesite/dacite suites from northern Hohxil, North Tibet: Geochronology, geochemical characteristics, petrogenetic processes and tectonic implications. Lithos 2011, 126, 54–67. [Google Scholar] [CrossRef]
  30. Tan, H.-Y.; Lü, J.-C.; Zhang, S.; Kou, L.-L. LA—ICP—MS zircon U-Pb andmolybdenite Re-Os dating for the luming large-scale molybdenum deposit inXiao Hinggan mountains and its geological implication. Jilin Univ. 2012, 42, 1757–1770. [Google Scholar]
  31. Chen, J.; Sun, F.-Y.; Pan, T.; Wang, J.; Huo, L. Geological features of Huojihe Molybdenum deposit in Heilongjiang, and geochronology and geochemistry of mineralized granodiorite. Jilin Univ. 2012, 42, 207–215. [Google Scholar]
  32. Yang, Y.-C.; Han, S.-J.; Sun, D.-Y.; Guo, J.; Zhang, S.-J. Geological and geochemical features and geochronology of porphyry molybdenum deposits in the Lesser Xing’an Range–Zhangguangcai Range metallogenic belt. Acta Pet. Sin. 2012, 28, 379–390. [Google Scholar]
  33. Shu, Q.; Chang, Z.; Lai, Y.; Hu, X.; Wu, H.; Zhang, Y.; Wang, P.; Zhai, D.; Zhang, C. Zircon trace elements and magma fertility: Insights from porphyry (-skarn) Mo deposits in NE China. Miner. Depos. 2019, 54, 645–656. [Google Scholar] [CrossRef]
  34. Misra, K.C. Understanding Mineral Deposits; Kluwer Academic Publishers: Alphen aan den Rijn, The Netherlands, 2000; pp. 353–413. [Google Scholar]
  35. Singer, D.A.; Berger, V.I.; Menzie, W.D. Porphyry copper deposit density. Econ. Geol. 2005, 100, 491–514. [Google Scholar] [CrossRef]
  36. Ju, N.; Zhang, S.; Zhang, D.; Kou, L.-L. A study on metallogenic epoch and geological characteristic of Chang’anbu Cu-Mo deposit, Shulan City of Jilin Province. Acta Geol. Sin. 2015, 89, 148–149. [Google Scholar]
  37. Ju, N.; Ren, Y.; Zhang, S.; Kou, L.; Zhang, D.; Gu, Y.; Yang, Q.; Wang, H.; Shi, L.; Sun, Q. The Early Jurassic Chang’anbu porphyry Cu–Mo deposit in Northeastern China: Constraints from zircon U-Pb geochronology and H-O-S-Pb stable isotopes. Geol. J. 2017, 53, 2437–2448. [Google Scholar] [CrossRef]
  38. Ju, N.; Zhang, S.; Kou, L.-L.; Wang, H.-P.; Zhang, D.; Gu, Y.-C.; Wu, T. Source and Tectonic Setting of Porphyry Mo Deposits in Shulan, Jilin Province, China. Minerals 2019, 9, 657. [Google Scholar] [CrossRef] [Green Version]
  39. Zhang, D.; Ju, N.; Zhang, S.; Kou, L.-L.; Gu, Y.-C. Geochemical characteristics and geological implication of the ore-forming rocks of the Changanpu Mo(Cu) deposit in Jilin Province. Geol. Resour. 2016, 26, 253–259. [Google Scholar]
  40. Zhou, Y.; Song, Q.; Zhang, Y.; Wang, Y.; Yu, C. Zircon U-Pb ages and Hf isotope composition of the ore-bearing intrusion from the Chang’anpu Mo-Cu Deposit, Jilin Province. Gold 2016, 37, 25–29. [Google Scholar]
  41. Han, S.-J.; Sun, J.-G.; Bai, L.-A.; Xing, S.-W.; Chai, P.; Zhang, Y.; Yang, F.; Men, L.-J.; Li, Y.-X. Geology and ages of porphyry and medium-to high-sulphidation epithermal gold deposits of the continental margin of Northeast China. Int. Geol. Rev. 2013, 55, 287–310. [Google Scholar] [CrossRef]
  42. Wu, F.-Y.; Jahn, B.-M.; Wilde, S.; Sun, D.-Y. Phanerozoic crustal growth: U-Pb and Sr-Nd isotopic evidence from the granites in northeastern China. Tectonophysics 2000, 328, 89–113. [Google Scholar] [CrossRef]
  43. Jahn, B.-M.; Wu, F.; Chen, B. Massive granitoid generation in Central Asia: Nd isotope evidence and implication for continental growth in the Phanerozoic. Episodes 2000, 23, 82–92. [Google Scholar] [CrossRef] [Green Version]
  44. Jahn, B.-M.; Capdevila, R.; Liu, D.; Vernon, A.; Badarch, G. Sources of Phanerozoic granitoids in the transect Bayanhongor-Ulaan Baatar, Mongolia: Geochemical and Nd isotopic evidence, and implications for Phanerozoic crustal growth. Asian Earth Sci. 2004, 23, 629–653. [Google Scholar] [CrossRef]
  45. Pei, Q.-M.; Zhang, S.-T.; Hayashi, K.-I.; Cao, H.-W.; Li, D.; Tang, L.; Hu, X.-K.; Li, H.-X.; Fang, D.-R. Permo-Triassic granitoids of the Xing’an-Mongolia segment of the Central Asian Orogenic Belt, Northeast China: Age, composition, and tectonic implications. Int. Geol. Rev. 2018, 60, 1172–1194. [Google Scholar] [CrossRef]
  46. Wu, F.-Y.; Sun, D.-Y.; Li, H.; Jahn, B.-M.; Wilde, S. A-type granites in northeastern China: Age and geochemical constraints on their petrogenesis. Chem. Geol. 2002, 187, 143–173. [Google Scholar] [CrossRef]
  47. 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 granites in northeastern China. J. Asian Earth Sci. 2011, 41, 1–30. [Google Scholar] [CrossRef] [Green Version]
  48. Jahn, B.-M.; Griffin, W.L.; Windley, B. Continental growth in the Phanerozoic: Evidence from Central Asia. Tectonophysics 2000, 328, 7–10. [Google Scholar] [CrossRef]
  49. Jahn, B.-M.; Wu, F.; Capdevila, R.; Fourcade, S.; Wang, Y.; Zhao, Z.; Wang, Y. Highly evolved juvenile granites with tetrad REE patterns: The Woduhe and Baerzhe granites from the Great Xing’an (Khingan) Mountains in NE China. Lithos 2001, 59, 171–198. [Google Scholar] [CrossRef]
  50. Shen, P.; Pan, H.; Hattori, K.; Cooke, D.R.; Seitmuratova, E. Large Paleozoic and Mesozoic porphyry deposits in the Central Asian Orogenic Belt: Geodynamic settings, magmatic sources, and genetic models. Gondwana Res. 2018, 58, 161–194. [Google Scholar] [CrossRef]
  51. Cao, M.-J.; Li, G.-M.; Qin, K.-Z.; Evans, N.J.; Seitmuratova, E.Y. Assessing the magmatic affinity and petrogenesis of granitoids at the giant Aktogai porphyry Cu deposit, Central Kazakhstan. Am. J. Sci. 2016, 316, 614–688. [Google Scholar] [CrossRef]
  52. Zhang, L.; Xiao, W.; Qin, K.; Zhang, Q. The adakite connection of the Tuwu-Yandong copper porphyry belt, eastern Tianshan, NW China: Trace element and Sr-Nd-Pb isotope geochemistry. Miner. Depos. 2006, 41, 188–200. [Google Scholar] [CrossRef]
  53. Goldfarb, R.J.; Taylor, R.D.; Collins, G.S.; Goryachev, N.A.; Orlandini, O.F. Phanerozoic continental growth and gold metallogeny of Asia. Gondwana Res. 2014, 25, 48–102. [Google Scholar] [CrossRef]
  54. Seltmann, R.; Porter, T.M.; Pirajno, F. Geodynamics and metallogeny of the central Eurasian porphyry and related epithermal mineral systems: A review. J. Asian Earth Sci. 2014, 79, 810–841. [Google Scholar] [CrossRef]
  55. Natal’in, B.A.; Borukayev, C.B. Mesozoic sutures in the southern Far East of USSR. Geotectonics 1991, 25, 64–74. [Google Scholar]
  56. Badarch, G.; Cunningham, W.D.; Windley, B.F. A new terrane subdivision for Mongolia: Implications for the Phanerozoic crustal growth of Central Asia. Asian Earth Sci. 2002, 21, 87–110. [Google Scholar] [CrossRef]
  57. 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]
  58. Xiao, W.; Windley, B.F.; Hao, J.; Zhai, M. Accretion leading to collision and the Permian Solonker Suture, Inner Mongolia, China: Termination of the central Asian orogenic belt. Tectonics 2003, 22, 1069–1090. [Google Scholar] [CrossRef] [Green Version]
  59. Wu, G.; Chen, Y.; Chen, Y.; Zeng, Q. 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. Asian Earth Sci. 2012, 49, 214–233. [Google Scholar] [CrossRef]
  60. Chen, Y.-J.; Zhang, C.; Wang, P.; Pirajno, F.; Li, N. The Mo deposits of Northeast China: A powerful indicator of tectonic settings and associated evolutionary trends. Ore Geol. Rev. 2017, 81, 602–640. [Google Scholar] [CrossRef]
  61. Ge, W.-C.; Lin, Q.; Sun, D.-Y.; Wu, F.-Y.; Won, C.K.; Won, L.M.; Jim, Y.S.; Yun, S.-H. Geochemical characteristics of the Mesozoic basalts in Da Hinggan Ling: Evidence of the mantle-crust interaction. Acta Pet. Sin. 1999, 15, 397–407. [Google Scholar]
  62. Berdnikov, N.V.; Nevstruev, V.G.; Saksin, B.G. Genetic aspects of the noble metal mineralization at the Poperechnoe deposit, lesser Khingan, Russia. Pacific Geol. 2017, 11, 421–435. [Google Scholar] [CrossRef]
  63. Didenko, A.N.; Kaplin, V.B.; Malyshev, Y.F.; Shevchenko, B.F. Lithospheric structure and Mesozoic geodynamics of the eastern Central Asian orogen. Geol. Geophys. 2010, 51, 492–506. [Google Scholar] [CrossRef]
  64. Khanchuk, A.I.; Didenko, A.N.; Popeko, L.I.; Sorokin, A.A.; Shevchenko, B.F. Structure and evolution of the Mongol-Okhotsk orogenic belt. In The Central Asian Orogenic Belt: Geology, Evolution, Tectonics, and Models; Kröner, A., Ed.; Borntraeger Science Publishers: Stüttgart, Germany, 2015; pp. 211–234. [Google Scholar]
  65. Berdnikov, N.V.; Nevstruev, V.G.; Kepezhinskas, P.K.; Mochalov, A.G.; Yakubovich, O.V. PGE mineralization in andesite explosive breccias associated with the Poperechnoye iron-manganese deposit (Lesser Khingan, Far East Russia): Whole-rock geochemical, 190Pt-4He isotopic, and mineralogical evidence. Ore Geol. Rev. 2020, 118, 3352. [Google Scholar] [CrossRef]
  66. Berdnikov, N.V.; Nevstruev, V.G.; Kepezhinskas, P.K.; Astapov, I.; Konovalova, N. Gold in mineralized volcanic systems in the Lesser Khingan Range (Russian Far East): Textural types, composition and possible origins. Geosciences 2021, 11, 103. [Google Scholar] [CrossRef]
  67. Slama, J.; Kosler, J.; Condon, D.J.; Crowley, J.L.; Gerdes, A.; Hanchar, J.M.; Horstwood, S.A.; Morris, G.A. Plesovice Zircon: A new natural reference material for U-Pb and Hf isotopic microanalysis. Chem. Geol. 2008, 249, 1–35. [Google Scholar] [CrossRef]
  68. Liu, Y.-S.; Hu, Z.-C.; Gao, S. In Situ Analysis of Major and Trace Elements of Anhydrous Minerals by LA-ICP-MS without Applying an Internal Standard. Chem. Geol. 2008, 257, 34–43. [Google Scholar] [CrossRef]
  69. Liu, Y.; Hu, Z.; Zong, K.; Gao, C.; Gao, S.; Xu, J.; Chen, H. Reappraisement and refinement of zircon U–Pb isotope and trace element analyses by LA-ICP-MS. Chin. Sci. Bull. 2010, 55, 1535–1546. [Google Scholar] [CrossRef]
  70. Anderson, T. Correction of common lead in U-Pb analyses that do not report 204Pb. Chem. Geol. 2002, 192, 59–79. [Google Scholar] [CrossRef]
  71. Ludwig, K.R. User’s Manual for Isoplot 3.0: A Geochronological Toolkit for Microsoft Excel; Berkeley Geochron Center Special Publication: Berkeley, CA, USA, 2003; pp. 1–71. [Google Scholar]
  72. Wu, F.-Y.; Yang, Y.-H.; Xie, L.-W.; Yang, J.-H.; Xu, P. Hf isotopic compositions of the standard zircons and baddeleyites used in U-Pb geochronology. Chem. Geol. 2006, 234, 105–126. [Google Scholar] [CrossRef]
  73. Blichert-Toft, J.; Albarede, F. The Lu-Hf isotope geochemistry of chondrites and the evolution of the mantle-crust system. Earth Planet. Sci. Lett. 1997, 148, 243–258. [Google Scholar] [CrossRef]
  74. Griffin, W.L.; Wang, X.; Jackson, S.E.; Pearson, N.J.; O’Reilly, S.Y.; Xu, X.-S.; Zhou, X.-M. 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]
  75. Sun, S.-S.; McDonough, W.F. Chemical and isotopic systematics of oceanic basalts: Implications for mantle composition and processes. Geol. Soc. Lond. Spec. Publ. 1989, 42, 313–345. [Google Scholar] [CrossRef]
  76. Song, Q.-H.; Xing, S.-W.; Zhang, Y.; Li, C.; Wang, Y.; Yu, C. Origin and Geochronology of Chang’anpu Mo-Cu Deposit in Jilin Province: Constraints from Molybdenite Re-Os Isotope Systematics. Rock Mineral. Anal. 2016, 35, 550–557. [Google Scholar]
  77. 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]
  78. Pearce, J.A.; Harris, N.B.W.; Tindle, A.G. Trace element discrimination diagrams for the tectonic interpretation of granitic rocks. J. Pet. 1984, 25, 956–983. [Google Scholar] [CrossRef] [Green Version]
  79. Barbarin, B. A review of the relationships between granitoid types, their origins and their geodynamic environments. Lithos 1999, 46, 605–626. [Google Scholar] [CrossRef]
  80. Sylvester, P.J. Post-collisional alkaline granites. J. Geol. 1989, 97, 261–280. [Google Scholar] [CrossRef]
  81. Whalen, J.B.; Currie, K.L.; Chappell, B.W. A-type granites: Geochemical characteristics, discrimination and petrogenesis. Contrib. Mineral. Pet. 1987, 95, 407–419. [Google Scholar] [CrossRef]
  82. Chappell, B.W.; White, A.J.R. Two contrasting granite type. Pac. Geol. 1974, 8, 173–174. [Google Scholar]
  83. Pitcher, W.S. Granite Type and Tectonic Environment; Mountain Building Processes; Academic Press: London, UK, 1982; pp. 19–40. [Google Scholar]
  84. Pitcher, W.S. The Nature and Origin of Granite Blackie; Academic and Professional: London, UK, 1993; p. 311. [Google Scholar]
  85. Wu, F.-Y.; Jahn, B.-M.; Wilde, S.A.; Lo, C.-H.; Yui, T.-F.; Lin, Q.; Ge, W.-C.; Sun, D.-Y. Highly fractionated I-type granites in NE China (II): Isotopic geochemistry and implications for crustal growth in the Phanerozoic. Lithos 2003, 67, 191–204. [Google Scholar] [CrossRef]
  86. Wu, F.-Y.; Li, X.-H.; Zheng, Y.-F.; Gao, S. Lu-Hf isotopic systematics and their application in petrology. Acta Pet. Sin. 2007, 23, 185–220. [Google Scholar]
  87. Cheng, Y.-B.; Mao, J.-W. Age and geochemistry of granites in Gejiu area, Yunnan province, SW China: Constraints on their petrogenesis and tectonic setting. Lithos 2010, 120, 258–276. [Google Scholar]
  88. 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. Pet. 1999, 134, 33–51. [Google Scholar] [CrossRef]
  89. Schiano, P.; Monzier, M.; Eissen, J.P.; Martin, H.; Koga, K.T. Simple mixing as the major control of the evolution of volcanic suites in the Ecuadorian Andes. Contrib. Mineral. Petr. 2010, 160, 297–312. [Google Scholar] [CrossRef]
  90. Chappell, B.W.; Bryant, C.J.; Wyborn, D. Peraluminous I-type granites. Lithos 2012, 153, 142–153. [Google Scholar] [CrossRef]
  91. Castillo, P.R. Adakite petrogenesis. Lithos 2012, 135, 304–316. [Google Scholar] [CrossRef]
  92. Defant, M.J.; Xu, J.-F.; Kepezhinskas, P.; Wang, Q.; Zhang, Q.; Xiao, L. Adakites: Some variations on a theme. Acta Pet. Sin. 2002, 18, 128–142. [Google Scholar]
  93. Drummond, M.S.; Defant, M.J.; Kepezhinskas, P.K. Petrogenesis of slab-derived trondhjemite- tonalite- dacite/adakite magmas. Earth Environ. Sci. Trans. R. Soc. Edinburgh. 1996, 87, 205–215. [Google Scholar] [CrossRef] [Green Version]
  94. Cocherine, A. Systematic use of trace element distribution patterns in log-log diagram for plutonic suites. Geochim. Et. Cosmochim. Acta 1986, 1, 2517–2522. [Google Scholar] [CrossRef]
  95. Peccerillo, A.; Taylor, A.R. Geochemistry of Eocene calc–alkaline volcanic rocks from the Kastamonu area, Northern Turkey. Contr. Miner. Petro. 1976, 58, 63–81. [Google Scholar] [CrossRef]
  96. Maniar, P.D.; Piccoli, P.M. Tectonic discrimination of granitoids. Geol. Soc. Am. Bull. 1989, 101, 635–643. [Google Scholar] [CrossRef]
  97. Vervoort, J.D.; Blichert-Toft, J. Evolution of the depleted mantle: Hf isotope evidence from juvenile rocks through time. Geochim. Et. Cosmochim. Acta 1999, 63, 533–556. [Google Scholar] [CrossRef]
  98. Kinny, P.D.; Maas, R. Lu-Hf and Sm-Nd isotope systems in zircon. Rev. Mineral. Geochem. 2003, 53, 327–341. [Google Scholar] [CrossRef]
  99. Zhang, Y.; Wu, F.; Wiled, S.A.; Zhai, M.; Lu, X.; Sun, D. Zircon U-Pb ages and tectonic implications of Early Paleozoic granitoids at Yanbian, Jilin Province, northeast China. Isl. Arc. 2004, 13, 484–505. [Google Scholar] [CrossRef]
  100. Wilkinson, J.J. Triggers for the formation of porphyry ore deposits in magmatic arcs: Predisposition or perfect storm? Nat. Geosci. Rev. Artic. 2013, 10, 1038–1940. [Google Scholar]
  101. Solomon, M. Subduction, arc reversal, and the origin of porphyry copper-gold deposits in island arcs. Geology 1990, 18, 630–633. [Google Scholar] [CrossRef]
  102. Richards, J.P. Tectono-magmatic precursors for porphyry Cu-(Mo-Au) deposit formation. Econ. Geol. 2003, 98, 1515–1533. [Google Scholar] [CrossRef]
  103. Audétat, A. Source and evolution of molybdenum in the porphyry Mo (-Nb) deposit at Cave Peak, Texas. J. Pet. 2010, 51, 1739–1760. [Google Scholar] [CrossRef]
  104. Audétat, A.; Dolejš, D.; Lowenstern, J.B. Molybdenite saturation in silicic magmas: Occurrence and petrological implications. J. Pet. 2011, 52, 891–904. [Google Scholar] [CrossRef] [Green Version]
  105. Li, Z.-Z.; Qin, K.-Z.; Li, G.-M.; Ishihara, S.; Jin, L.-Y.; Song, G.-X.; Meng, Z.-J. Formation of the giant Chalukou porphyry Mo deposit in northern Great Xing’an Range, NE China: Partial melting of the juvenite lower crust in untra-plate extensional environment. Lithos 2014, 202–203, 138–1156. [Google Scholar] [CrossRef]
  106. Fan, W.-M.; Guo, F.; Wang, Y.-J.; Lin, G. Late Mesozoic calc-alkaline volcanism of post-orogenic extension in the northern Da Hinggan Mountains, northeastern China. J. Volcanol. Geotherm. Res. 2003, 121, 115–135. [Google Scholar] [CrossRef]
  107. Shao, J.-A.; Zang, S.-X.; Mou, B.-L.; Li, X.-B.; Wang, B. Extensional tectonics and asthenospheric upwelling in the orogenic belt: A case study from Hinggan—Mongolia orogenic belt. Chin. Sci. Bull. 1994, 39, 533–537. [Google Scholar]
  108. Lin, Q.; Ge, W.-C.; Sun, D.-Y.; Wu, F.-Y. Geomechanical significance of the Mesozoic volcanics in Northeast Asia. Chin. J. Geophys. Chin. Ed. 1999, 42, 75–84. [Google Scholar]
  109. Mao, J.-W.; Wang, Z.-L. A preliminary study on time limits and geodynamic setting of large-scale metallogeny in East China. Miner. Depos. 2000, 19, 289–296. [Google Scholar]
  110. Muller, D.; Groves, D.I. Potassic igneous rocks and associated gold-copper mineralization. Miner. Resour. Rev. 2019, 53, 1235–1236. [Google Scholar]
  111. Pearce, J.A. Sources and settings of granitic rock. Episodes 1996, 19, 120–125. [Google Scholar] [CrossRef] [Green Version]
  112. Wakita, K.; Metcalfe, I. Ocean plate stratigraphy in East and Southeast Asia. J. Asian Earth Sci. 2005, 24, 679–702. [Google Scholar] [CrossRef]
  113. Yu, J.-J.; Wang, F.; Xu, W.-L.; Gao, F.-H.; Pei, F.-P. Early Jurassic mafic magmatism in the Lesser Xing’an—Zhangguangcai Range, NE China, and its tectonic implications: Constraints from zircon U-Pb chronology and geochemistry. Lithos 2012, 142, 256–266. [Google Scholar] [CrossRef]
  114. Xiao, W.; Song, D.; Windley, B.F.; Li, J.; Han, C.; Wan, B.; Zhang, J.; Ao, S.; Zhang, Z. Accretionary processes and metallogenesis of the Central Asian Orogenic Belt: Advances and perspectives. Sci. China Earth Sci. 2020, 63, 329–361. [Google Scholar] [CrossRef] [Green Version]
  115. Wu, F.-Y.; Lin, J.-Q.; Wilde, S.A.; Zhang, X.-O.; Yang, J.-H. Nature and significance of the Early Cretaceous giant igneous event in eastern China. Earth Planet. Sci. Lett. 2005, 233, 103–119. [Google Scholar] [CrossRef]
  116. Guo, F.; Li, H.; Fan, W.; Li, J.; Zhao, L.; Huang, M.; Xu, W. Early Jurassic subduction of the Paleo-Pacific Ocean in NE China: Petrologic and geochemical evidence from the Tumen mafic intrusive complex. Lithos 2015, 224, 46–60. [Google Scholar] [CrossRef]
Figure 1. Simplified tectonic map (A), and geological map (B), of the Chang’anpu Mo deposit and adjacent areas (Modified after Han et al., 2013) [40].
Figure 1. Simplified tectonic map (A), and geological map (B), of the Chang’anpu Mo deposit and adjacent areas (Modified after Han et al., 2013) [40].
Minerals 12 00967 g001
Figure 2. Regional geological and metallogenic map (A), and Exploration section map (B) in the Chang’anpu Mo deposit.
Figure 2. Regional geological and metallogenic map (A), and Exploration section map (B) in the Chang’anpu Mo deposit.
Minerals 12 00967 g002
Figure 3. Microphotographs and photographs of monzogranite in the Chang’anpu area. (A) monzogranite (hand specimen); (B) brecciated monzogranite (hand specimen); (C) photomicrograph of monzogranite adjacent to the ore body; (D) photomicrograph of monzogranite (cross-polarized light). Abbreviations: Qtz—quartz; Pl—plagioclase; Bt—biotite; Kfs—potash feldspar.
Figure 3. Microphotographs and photographs of monzogranite in the Chang’anpu area. (A) monzogranite (hand specimen); (B) brecciated monzogranite (hand specimen); (C) photomicrograph of monzogranite adjacent to the ore body; (D) photomicrograph of monzogranite (cross-polarized light). Abbreviations: Qtz—quartz; Pl—plagioclase; Bt—biotite; Kfs—potash feldspar.
Minerals 12 00967 g003
Figure 4. (A) Chondrite normalized REE pattern; (B) primitive mantle normalized trace element spider diagram of monzogranite in the Chang’anpu porphyry Mo deposit (Sun and McDonough, 1989) [75].
Figure 4. (A) Chondrite normalized REE pattern; (B) primitive mantle normalized trace element spider diagram of monzogranite in the Chang’anpu porphyry Mo deposit (Sun and McDonough, 1989) [75].
Minerals 12 00967 g004
Figure 5. Cathodoluminescent images and LA-ICP-MS zircon U–Pb concordia diagrams for the monzogranite. (A) CAP-N1, (B) CAP-N2, (C) CAP-N3.
Figure 5. Cathodoluminescent images and LA-ICP-MS zircon U–Pb concordia diagrams for the monzogranite. (A) CAP-N1, (B) CAP-N2, (C) CAP-N3.
Minerals 12 00967 g005
Figure 6. Zircon Hf isotopic characteristics of monzogranite. CHUR: bulk earth chondritic uniform reservoir. (A) The characteristic of Hf isotopic composition, (B) εHf(t) versus T (Ma) diagram of Mo deposits in NE China.
Figure 6. Zircon Hf isotopic characteristics of monzogranite. CHUR: bulk earth chondritic uniform reservoir. (A) The characteristic of Hf isotopic composition, (B) εHf(t) versus T (Ma) diagram of Mo deposits in NE China.
Minerals 12 00967 g006
Figure 7. Harker variation diagrams for the monzogranite. TiO2 (A), K2O (B), Na2O (C), Al2O3 (D), MgO (E), CaO (F), P2O5 (G), Rb/Ba (H), and Rb/Sr (I) versus SiO2 diagrams.
Figure 7. Harker variation diagrams for the monzogranite. TiO2 (A), K2O (B), Na2O (C), Al2O3 (D), MgO (E), CaO (F), P2O5 (G), Rb/Ba (H), and Rb/Sr (I) versus SiO2 diagrams.
Minerals 12 00967 g007
Figure 8. Petrogenetic diagrams for the monzogranite. (A) Th vs. Th/Ti diagram (Schiano et al., 2010) [89]. (B) La/Sm vs. La diagram [94]. The inset is aschematic CI versus CI/CC diagram (CI, incompatible element concentration, and CC, compatible element concentrations). (C) Diagrams of SiO2 vs. K2O. (D) A/CNK vs. A/NK. The boundary lines in the SiO2 vs. K2O and A/CNK vs. A/NK diagrams are from Peccerillo and Taylor (1976), and Maniar and Piccoli (1989) [95,96], respectively. The symbols are the same as in Figure 7.
Figure 8. Petrogenetic diagrams for the monzogranite. (A) Th vs. Th/Ti diagram (Schiano et al., 2010) [89]. (B) La/Sm vs. La diagram [94]. The inset is aschematic CI versus CI/CC diagram (CI, incompatible element concentration, and CC, compatible element concentrations). (C) Diagrams of SiO2 vs. K2O. (D) A/CNK vs. A/NK. The boundary lines in the SiO2 vs. K2O and A/CNK vs. A/NK diagrams are from Peccerillo and Taylor (1976), and Maniar and Piccoli (1989) [95,96], respectively. The symbols are the same as in Figure 7.
Minerals 12 00967 g008
Figure 9. Variations of Sr/Y versus Y (ppm) in monzogranite associated with Chang’anpu porphyry Mo deposit (adapted from Kepezhinskas al., 2022) [21].
Figure 9. Variations of Sr/Y versus Y (ppm) in monzogranite associated with Chang’anpu porphyry Mo deposit (adapted from Kepezhinskas al., 2022) [21].
Minerals 12 00967 g009
Figure 10. (A) Rb– (Y + Nb), (B) Nb–Y, (C) Zr/Al2O3—TiO2/Al2O3 and (D) Ce/P2O5—Zr/TiO2 discrimination diagrams of the monzogranite (adapted from Pearce et al., 1984; Muller and Groves, 1994; Pearce, 1996) [78,110,111]. VAG: volcanic arc granites; ORG: ocean ridge granites; WPG: within-plate granites; syn-COLG: syn-collision granites; post-COLG: post-collision granites. The symbols are the same as in Figure 7.
Figure 10. (A) Rb– (Y + Nb), (B) Nb–Y, (C) Zr/Al2O3—TiO2/Al2O3 and (D) Ce/P2O5—Zr/TiO2 discrimination diagrams of the monzogranite (adapted from Pearce et al., 1984; Muller and Groves, 1994; Pearce, 1996) [78,110,111]. VAG: volcanic arc granites; ORG: ocean ridge granites; WPG: within-plate granites; syn-COLG: syn-collision granites; post-COLG: post-collision granites. The symbols are the same as in Figure 7.
Minerals 12 00967 g010
Figure 11. Diagenetic and metallogenic model of Early–Middle Jurassic porphyry deposit (modified after Guo et al. (2015) [116].
Figure 11. Diagenetic and metallogenic model of Early–Middle Jurassic porphyry deposit (modified after Guo et al. (2015) [116].
Minerals 12 00967 g011
Table 1. Geochronology and tectonic setting of metallogenic porphyry in Early–Middle Jurassic nonferrous metal deposits in NE China.
Table 1. Geochronology and tectonic setting of metallogenic porphyry in Early–Middle Jurassic nonferrous metal deposits in NE China.
No.Typical DepositsMineral TypeMetallogenic RockAges (Ma)Tectonic SettingReferences
1DongfengbeishanMoPorphyritic monzogranite195.6 ± 1.9Extensional tectonic setting caused by subduction of the Pacific plate[22]
2DaheishanMoGranodiorite porphyry166.6 ± 4Extensional tectonic setting caused by subduction of the Pacific plate[23]
3XingshanMoGranodiorite porphyry171.7 ± 2.2Active continental margin; related to subduction of the Pacific plate[24]
4DashiheMoPorphyritic granodiorite186.7 ± 5Extensional tectonic setting caused by subduction of the Pacific plate[25]
5Fu’anpuMoPorphyritic monzogranite166.9 ± 6.7Extensional tectonic setting caused by subduction of the Pacific plate[26]
6JidetunMoGranodiorite180.2 ± 0.8Active continental margin; related to subduction of the Pacific plate[27]
7Chang’anpuMoMonzogranite174.3–174.9Extensional tectonic setting caused by subduction of the Pacific plateThis study
8HoudaomuMoYanshanian granite167.5 ± 1.2Extensional tectonic setting caused by subduction of the Pacific plate[28]
9LiushengdianMoMonzogranite porphyry169.4 ± 1.0Volcanic arc setting; related to subduction of the Pacific plate[29]
10LumingMoMonzogranite porphyry183.2 ± 1.9Volcanic arc setting; related to subduction of the Pacific plate[30]
11HuojiheMoGranodiorite181.0 ± 1.9Volcanic arc setting; related to subduction of the Pacific plate[31]
12CuilingMoQuratz monzonite178.0 ± 0.7Volcanic arc setting; related to subduction of the Pacific plate[32]
Table 2. Major oxides (wt.%) and trace elements (ppm) compositions for the Chang’anpu monzogranite.
Table 2. Major oxides (wt.%) and trace elements (ppm) compositions for the Chang’anpu monzogranite.
Sample No.N1-Q1N1-Q2N1-Q3N1-Q4N1-Q5N1-Q6N2-Q1N2-Q2N2-Q3N3-Q1N3-Q2N3-Q3N3-Q4N3-Q5
SiO269.8170.2270.1170.3670.0369.5073.4875.4270.1969.7672.1684.3974.7871.33
Al2O315.2415.2015.3614.9315.3015.0213.9313.0815.2115.0014.286.3511.9613.12
TFe2O32.562.172.342.242.262.300.681.002.262.242.051.221.021.95
MgO0.670.630.660.560.590.660.130.140.600.670.500.150.340.60
CaO1.861.801.821.861.941.770.740.471.941.781.440.741.612.00
Na2O4.254.424.424.444.584.043.313.024.554.304.100.162.130.21
K2O3.583.773.693.733.704.106.505.883.634.134.224.235.577.19
MnO0.020.020.020.020.020.030.020.010.020.020.020.020.020.03
TiO20.370.340.350.330.340.370.060.070.340.350.290.190.190.16
P2O50.140.120.130.120.120.13<0.01<0.010.120.140.100.150.040.03
LOI 10001.240.901.000.910.941.300.840.660.901.581.341.361.983.04
Mg#0.340.360.360.320.330.360.270.220.340.370.320.190.400.38
La33.231.132.931.931.630.517.128.032.324.720.514.920.724.0
Ce67.961.367.864.162.861.133.953.064.958.747.535.846.052.7
Pr7.376.427.216.996.656.663.425.087.067.205.694.014.795.35
Nd25.622.324.724.523.524.110.715.124.723.018.013.413.915.5
Sm4.083.463.993.773.704.141.732.383.803.983.092.472.432.76
Eu0.910.910.870.850.920.930.260.290.890.800.720.490.480.56
Gd2.772.342.762.542.442.671.362.082.372.632.041.681.742.09
Tb0.330.320.320.310.310.320.210.320.300.340.270.230.280.28
Dy1.681.431.541.431.501.581.181.791.491.671.191.231.611.56
Ho0.290.260.290.250.260.250.260.380.260.300.240.230.340.29
Er0.810.720.770.640.690.660.821.360.710.830.600.661.020.83
Tm0.120.100.110.090.100.090.150.250.100.120.100.100.170.13
Yb0.660.640.660.590.660.631.071.750.620.760.560.671.210.87
Lu0.110.110.110.090.100.100.190.340.100.120.100.100.200.15
Y7.77.07.57.07.06.87.211.77.17.15.85.36.17.5
ΣREE145.83131.41144.03138.05135.23133.7372.35112.12139.60125.15100.6075.9794.87107.07
LREE139.06125.49137.47132.11129.17127.4367.11103.85133.65118.3895.5071.0788.30100.87
HREE6.775.926.565.946.066.305.248.275.956.775.104.906.576.20
LREE/HREE20.5421.2020.9622.2421.3220.2312.8112.5622.4617.4918.7314.5013.4416.27
LaN/YbN33.9132.7633.6136.4532.2832.6410.7710.7935.1221.9124.6814.9911.5318.60
δEu0.780.920.760.790.880.800.500.390.850.710.830.700.680.69
δCe1.000.991.020.990.990.991.010.990.991.051.041.091.071.08
Rb104.597.4103.594.394.2118.521420093.5113.0118.0129.5195.0272
Ba710741729758757870751327736710690360430480
Th11.507.7112.108.948.258.9833.137.78.456.586.5916.4532.136.2
U4.893.093.875.634.642.696.6316.103.751.72.88.08.120.2
K29,10031,10031,30031,10029,50033,00052,00047,90028,50032,90033,20033,10044,40056,700
Ta0.960.690.800.710.840.581.822.230.690.360.550.641.110.85
Nb9.27.99.28.68.57.64.29.07.74.76.95.54.23.8
Sr551575578571604529175.0134.057448839979.4125.0119.0
P6305906105505505902030560600480620170160
Zr196.0183.0190.0190.0193.0195.082.985.6168.0178158667979
Ti186018501820184018201800260430182014001640890750620
Table 3. Results of LA-ICPMS U–Pb dating for the single-grain zircon from the monzogranite.
Table 3. Results of LA-ICPMS U–Pb dating for the single-grain zircon from the monzogranite.
Spot No.Content (ppm)Isotopic RatiosAges (Ma)
ThUTh/U207Pb/206Pb207Pb/235U206Pb/238U207Pb/235U206Pb/238U
CAPN1-01116619090.610.05051680.0010.1883120.00520.0272760.000417541733
CAPN1-0393411570.810.04994820.0020.1885020.00570.0277340.000417551762
CAPN1-0588510740.820.05280130.0050.1961950.01930.0277040.0007182161764
CAPN1-075658140.690.04945490.0020.1828350.00620.0270330.000417051723
CAPN1-08125621660.580.04999370.0010.1888350.00540.0275140.000417651752
CAPN1-09129512741.020.05074140.0020.1914760.00880.0278710.000917881776
CAPN1-10179442650.420.04944650.0010.1876140.00510.0277090.000517541763
CAPN1-1159713300.450.04981940.0010.1906120.00550.0279370.000417751782
CAPN1-1291514600.630.05045640.0040.1853560.01380.0274010.0006173121744
CAPN1-1390613090.690.05000060.0020.1859350.00620.0271830.000417351732
CAPN1-14159331570.50.04777140.0010.1757820.00510.027130.000516441733
CAPN1-15131919120.690.0499470.0010.1905640.00520.027890.000517741773
CAPN1-17151422420.680.05154320.0030.1865140.01370.0274350.0006174121744
CAPN1-18122110301.190.04873570.0020.1797840.00530.0270080.000416851722
CAPN1-19172431410.550.04972130.0020.1943860.00730.0275770.000618061754
CAPN1-2010259491.080.05006140.0020.1892310.00750.0276190.000417661763
CAPN2-024166590.630.05201460.0030.1970990.01040.0275320.000518391753
CAPN2-035286340.830.05102650.0020.1959210.01040.027570.000518291753
CAPN2-055568330.670.05127140.0020.1920380.01010.0277930.000717891774
CAPN2-0669918630.380.04957160.0020.188080.00640.0274960.000417551753
CAPN2-0743810620.410.05023620.0020.1888450.00620.0273590.000317651742
CAPN2-087018610.810.05206150.0030.1878270.00860.0273550.000517571743
CAPN2-098208820.930.05085210.0020.1925260.00710.0275820.000317961752
CAPN2-124067680.530.05120030.0020.1955540.00760.0278210.000318161772
CAPN2-144908220.60.05122110.0060.1796940.01150.0273620.0004168101742
CAPN2-155566410.870.04818570.0030.1884620.01370.0274440.0005175121753
CAPN2-2069713910.50.04611230.0020.1757790.01040.0274030.000416491742
CAPN3-013093670.840.050530.0040.189990.0140.027280.0005177121743
CAPN3-022895490.530.051090.0030.194540.01120.027620.0005180101763
CAPN3-03102911100.930.050640.0030.188930.01170.027060.0005176101723
CAPN3-043964160.950.049310.0090.181550.03160.026710.0011169271707
CAPN3-051582630.60.049780.0080.192390.02820.028040.0009179241786
CAPN3-062382910.820.050780.0060.188880.020.026990.0007176171724
CAPN3-072473240.760.050190.0030.190990.01270.027610.0005177111763
CAPN3-105196890.750.049340.0040.186790.01580.027470.0006174131754
CAPN3-1281910360.790.050260.0030.189910.0110.027420.000517791743
CAPN3-131633280.50.048920.0080.186850.02840.027730.0009174241766
CAPN3-141141940.590.050550.010.180680.03610.025950.0011169311657
CAPN3-1523823810.050080.0090.196440.03570.028480.0012182301817
CAPN3-163213021.070.052290.010.190210.03420.026420.0009177291686
CAPN3-174243861.10.051250.0040.198320.01550.02810.0006184131794
CAPN3-183785350.710.052050.0070.189610.02590.026460.0008176221685
CAPN3-1910158091.250.050180.0030.193730.00950.028050.000418081783
CAPN3-2011114582.420.051440.0030.191290.01240.027020.0005178111723
Table 4. Zircon Hf isotopic compositions for monzogranite in the Chang’anpu Mo deposit.
Table 4. Zircon Hf isotopic compositions for monzogranite in the Chang’anpu Mo deposit.
Spot No.t (Ma)176Yb/177Hf176Lu/177Hf176Hf/177Hf±2σεHf(0)εHf(t)TDM1(Hf)TDM2(Hf)fLu/Hf
CAP1-11680.025927660.00098401580.28288610.00002194.047.62518618−0.97
CAP1-21790.017299840.00069200990.28285320.00001872.876.72561673−0.98
CAP1-31750.019013830.00077039590.28286790.00001553.397.15541648−0.98
CAP1-41650.013632660.00051998550.28290330.00001644.648.21488585−0.98
CAP1-51760.015514190.00059057860.28286480.00002103.287.08543652−0.98
CAP1-61630.018581290.00069989500.28292310.00001925.348.85462551−0.98
Note: ɛHf(t) = {[(176Hf/177Hf) − (176Lu/177Hf) × (eλt−1)]/[(176Hf/177Hf)CHUR(0) − (176Lu/177Hf)CHUR×(eλt−1)] − 1} × 10,000. (176Hf/177Hf)CHUR(0) = 0.282772 and (176Lu/177Hf)CHUR = 0.0332 (Blichert-Toft and Albarède, 1997); λLu = 1.867 × 10−11 year−1 [77].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, J.; Yang, Y.; Han, S.; Wutiepu, W. Geochronology, Geochemistry, and Lu-Hf Isotopic Compositions of Monzogranite Intrusion from the Chang’anpu Mo Deposit, NE China: Implications for Tectonic Setting and Mineralization. Minerals 2022, 12, 967. https://doi.org/10.3390/min12080967

AMA Style

Zhang J, Yang Y, Han S, Wutiepu W. Geochronology, Geochemistry, and Lu-Hf Isotopic Compositions of Monzogranite Intrusion from the Chang’anpu Mo Deposit, NE China: Implications for Tectonic Setting and Mineralization. Minerals. 2022; 12(8):967. https://doi.org/10.3390/min12080967

Chicago/Turabian Style

Zhang, Jian, Yanchen Yang, Shijiong Han, and Wukeyila Wutiepu. 2022. "Geochronology, Geochemistry, and Lu-Hf Isotopic Compositions of Monzogranite Intrusion from the Chang’anpu Mo Deposit, NE China: Implications for Tectonic Setting and Mineralization" Minerals 12, no. 8: 967. https://doi.org/10.3390/min12080967

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