Next Article in Journal
Chemical and Physical Denudation Rates in the Poços de Caldas Alkaline Massif, Minas Gerais State, Brazil
Previous Article in Journal
Study on Chalcopyrite Dissolution Mechanism and Bioleaching Community Behavior Based on Pulp Concentration Gradient at 6 °C
Previous Article in Special Issue
Genesis of the Mahour Base Metal Deposit, Iran: Constraints from Fluid Inclusions and Sulfur Isotopes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geochronological, Geochemical and Pb Isotope Inferences for Genesis of Wulandele Porphyry Molybdenum Deposit, Inner Mongolia, Northeast China

1
School of Earth Sciences and Resources, China University of Geosciences, Beijing 100083, China
2
Natural Resources Bureau of Maoxian, Aba Prefecture 623200, China
*
Author to whom correspondence should be addressed.
Minerals 2024, 14(7), 699; https://doi.org/10.3390/min14070699
Submission received: 22 May 2024 / Revised: 4 July 2024 / Accepted: 4 July 2024 / Published: 9 July 2024
(This article belongs to the Special Issue Role of Granitic Magmas in Porphyry, Epithermal, and Skarn Deposits)

Abstract

:
Integrated geochemical, U-Pb zircon, and Pb isotopic data from granitoids of the Wulandele porphyry molybdenum deposit, northeastern Inner Mongolia, are reported to disclose the possible magmatic process and Mo ore-forming process. LA-ICP-MS zircon U-Pb dating constrains the timing of the quartz diorite and monzonitic granite to 282 ± 2.4 Ma and 135.4 ± 2.1 Ma, respectively. The ages are accordant with geological facts which state that the shallow Permian granitoids are only the ore-hosting rock while the concealed Cretaceous fine-grained granite is the causative intrusion. Whole-rock geochemical data show that the granitoids belong to the high-K calc-alkaline series, and are enriched in LILEs, but depleted in HSFEs. Permian granitoids exhibit I-type characteristics, while Cretaceous granite is akin to A-type granite. Pb isotopic ratios are consistent between Permian granitoids and Cretaceous granite with ratios of 206Pb/204Pb = 18.048–18.892, 207Pb/204Pb = 15.488–15.571, and 208Pb/204Pb = 37.066–38.441. Considering geological and geochemical features together, Permian granitoids are interpreted as subduction-related continental margin high-K calc-alkaline rocks, while Cretaceous granite may be the result of the remelting of the relic Permian arc in an extensional environment induced by the rollback of the Paleo-Pacific plate. Different from classical porphyry-type deposits, the Wulandele Mo deposit, which formed in an intraplate tectonic setting, indicates that the intraplate porphyry Mo deposit should be one of the important exploration targets in the Central Asian Orogenic Belt, especially its eastern segment.

1. Introduction

The Central Asian Orogenic Belt (CAOB), comprised of island arcs, seamounts, accretionary wedges, oceanic plateaus, and microcontinents accreted in the closure process of the Paleo-Asian Ocean (PAO), is a world-famous accretionary orogen [1,2,3] and evolved for some 720 Ma [4,5,6]. It is also a giant metallogenic belt, hosting huge amounts of Cu, Au, Pb, and Zn deposits from the Neoproterozoic to Cretaceous [7,8,9,10,11,12,13,14], with particularly significant porphyry-type and epithermal-type mineralization [15,16].
Situated in the eastern section of the CAOB, the Xingmeng Orogenic Belt (XMOB) has long been an important Au, Ag, Cu, Mo, Pb-Zn ore belt in China [17,18,19]. More than 40 molybdenum deposits have been confirmed in this area in the past 20 years [18,19,20]. The great discoveries include the superlarge Luming (890Mt @ 0.084% Mo), Chalukou (1340 Mt @ 0.080% Mo), Diyanqijamu (800 Mt @ 0.081% Mo), and Daheishan (1090 Mt @ 0.081% Mo) Mo deposits [21,22,23,24,25,26]. Also, two porphyry Mo-(Cu-W) belts, the Xilamulun tectononic–magmatic–metallogenic belt (Chehugou, Jiguanshan, Nianzigou, etc. [27,28]) and the Chagan Obo-Chaobulen tectononic–magmatic–metallogenic belt (Wulandele, Wurinitu, Zhunsujihua, Bogda Uul, etc. [29,30]), have been delineated through years of exploration (Figure 1).
The Wulandele Mo deposit is a medium-sized porphyry-type deposit that occurred in the Chagan Obo-Chaobulen belt (Figure 1). Approximately 56 Mt of ore (0.093% Mo) was proven during a general exploration in 2008 [31]. Geological characteristics, fluid inclusions, and ore-forming age have been studied by former researchers [32,33]. Mo mineralization occurred both in the shallow quartz diorite and granodiorite and in the deep monzonitic granite. Tao et al. (2009) determined a zircon SHRIMP U-Pb age of 131.3 ± 1.6 Ma (2σ) for fine-grained monzonitic granite and a molybdenite Re-Os age of 134.1 ± 3.3 Ma [31]. However, the role of shallow granitoids and deep granite in the ore-forming processes, and tectonometallogenic setting of the Wulandele molybdenum deposit are not well constrained. In this contribution, we provide new zircon U-Pb ages, whole-rock geochemistry, and Pb isotopic compositions for the granitoids from the Wulandele Mo deposit. The results give us new insight into the ore-forming process and allow us to reveal the crystallization age, magmatism, and tectonic environment.

2. Geological Framework

2.1. Regional Geology

The Wulandele porphyry Mo deposit, approximately 150 km northeast of the Erenhot, is located in the Wulandele area of the Baiyinwula town, Sonid Left Banner (Figure 1). It is extremely close to the Chinese–Mongolian border and is not far from the superlarge Oyu Tolgoi and Tsagaan Suvarga deposits. Some researchers indicated that they developed in a similar tectonic environment (Silurian to Carboniferous arcs [10,34,35]).
Tectonically, the Wulandele Mo deposit is situated within the Chagan Obo-Chaobulen belt in the western part of the Great Xing’an Range (GXR) [29], which is considered as the easternmost extension of the CAOB. The GXR underwent complex tectonic evolution, including the Palaeozoic PAO stage, the Mesozoic Mongol–Okhotsk Ocean (MOO) stage, and the Mesozoic Paleo-Pacific Ocean (PPO) stage. This geological complexity has rendered the GXR renowned for its diverse tectonic–magmatic–metallogenetic signatures, showcasing both Paleozoic deposits like the Bainaimiao Cu deposit, Duobanshan and Tongshan Cu-Mo deposit, as well as Mesozoic deposits such as the Luming Mo deposit, Chalukou Mo deposit, Diyanqinamu Mo deposit, and Wunugetushang Cu deposit.
In the Wulandele region, a sequence of Paleozoic strata is exposed, consisting of the Lower Ordovician Wubinaobao Formation, the Middle Ordovician Bayanhusu Formation, and the Upper Carboniferous Baoligaomiao Formation [31]. The Wubinaobao Formation is characterized by silty slate, neritic sericite slate, metamorphic feldspar sandstone, and interbedded limestone and meta-siltstone layers, with local occurrences of andesite, dacite, sedimentary tuff, and andesitic volcanic breccia. The Bayanhusu Formation primarily consists of neritic sandy slate and metamorphic siltstone, with limestone intrusions. The Baoligaomiao Formation is composed of grayish green metamorphic feldspar-quartz sandstone and siltstone, representing typical continental clastic sedimentary rocks. In the southern and southwestern regions, Mesozoic Upper Jurassic continental volcanic rocks can be found sporadically.
Magmatic activity is intense and frequent in the study area. A substantial amount of intrusions dominated by acidic granitoids comprise an NE-trend magmatic rock belt [31,32,33]. Two epochs of magmatism are obvious: the Hercynian period is marked by widespread intrusive events, while the Yanshanian epoch is characterized by a combination of volcanic eruptions and intrusive activity. Regional Mo-polymetallic mineralization is mainly related to Yanshanian intrusive magmatism. Notable porphyry deposits include the large-scale Wurinitu Mo-W deposit, the medium-sized Dalaiaobao Mo deposit, and the medium-sized Wulandele Mo deposit.

2.2. Ore Deposit Geology

2.2.1. Host Rocks

The Wulandele deposit is surrounded by large areas of Permian middle-coarse biotite granite (Figure 2), but no mineralization occurred in the biotite granite. Most ore-bearing quartz veins (Figure 3a) are found in Permian quartz diorite and granodiorite (Zircon U-Pb age of 299.3 ± 2.4 Ma [31]), with little amount of biotite monzonitic granite at the surface. In addition, disseminated ores are disclosed by drill cores at the boundary between concealed Cretaceous monzonitic granite and the Permian granitoids mentioned above. The current manuscript focuses on Cretaceous monzonitic granite and Permian quartz diorite and granodiorite.
Quartz diorite (Figure 3b), one of the important ore-hosting rocks, covered an area of about 2 square kilometers at the center of the ore district. It is dark gray, with a fine–middle-grained unequal grain texture and mass structure. Mineral compositions are plagioclase (50%–75%), hornblende (15%–35%), and quartz (5%–10%) with K-feldspar content lower than 5%. Euhedral plagioclase is 0.1–1.0 mm long, euhedral reddish brown-to-brown hornblende is 0.4–2.0 mm long, and xenomorphic granular quartz is 0.4–2.8 mm in diameter. This is due to the fact that intrusion of the concealed Cretaceous monzonitic granite in the deep, epidotization, chloritization, and sericitization are common in quartz diorite. West of the Quartz diorite, there occurred some middle–fine-grained biotite monzonitic granite, which is thought of as the contemporaneous marginal face of the quartz diorite [31].
Granodiorite (Figure 3c) is outcropped east of quartz diorite and is also one of the ore-hosting rocks. It gradually changed to quartz diorite and no obvious boundary has been observed between them. Alteration is rather strong and sericitization is common. Granodiorite is grayish red with fine–middle-grained unequal grain texture and mass structure. The mineral compositions are plagioclase (35%–55%), K-feldspar (15%–25%), quartz (15%–20%), hornblende (5%–5%), and biotite (about 5%). Grains of most minerals range from 0.2 to 0.8 mm except for a small amount of quartz that can reach 2.4 mm. Carlsbad twins and polysynthetic twins are often seen in K-feldspar and plagioclase. Most K-feldspars look very dirty under the microscope due to sericitization.
Monzonitic granite (Figure 3d) is concealed in the deep and was revealed by the drill holes. It intruded into the granodiorite and quartz diorite, resulting in extensive Mo mineralization near the contact zone. Being the main ore-hosting rock, it was interpreted as the mother rock of the Wulandele Mo deposit [31]. Monzonitic granite is flesh-red in color with fine-grain granitic texture and mass structure. Its constituents include quartz (30%–40%), plagioclase (30%–45%), K-feldspar (30%–40%), and trace amounts of hornblende and biotite. Quartz is clear, with a xenomorphic granular structure and diameters ranging from 0.4 to 2.8 mm. Both plagioclase and K-feldspar are euhedral tabular in shape, with diameters of 0.4–2.4 mm. Carlsbad twins and polysynthetic twins are common under the microscope.

2.2.2. Ore Bodies

The Wulandele Mo deposit developed two distinct types of ore bodies (Figure 2): quartz vein type (Figure 3e) in the shallow Permian granitoids and disseminated type (Figure 3f) close to the boundary between Cretaceous monzonitic granite and Permian granitoids.
A series of ore-bearing quartz veins occur in fractures of quartz diorites and granodiorite. The occurrence of these ore veins is affected by the occurrence of the contact zone between Cretaceous monzonitic granite and Permian quartz diorite and granodiorite, trending in NW and dipping to SW with a dip angle of about 60°. Three types of felsic ore veins have been observed in quartz diorites, including pyrite-bearing felsic veins, pyrite–molybdenite quartz veins, and pyrite–molybdenite–chalcopyrite quartz veins. These veins constitute the upper ore bodies, with a maximum length of 900 m and an average thickness of 38 m. The average Mo grade of the ores of the upper vein-type ore bodies is 0.092%, with copper as an important accompanying element. The ores of the vein-type ore bodies constitute about 70% of the reserve of the whole deposit. Weak alteration (usually thinner than 2 cm) of chloritization, epidotization, silicification, sericitization, and kaolinization can usually be seen along these ore veins.

2.2.3. Ore Characteristics

According to different mineral associations, two types of ores occurred in the upper quartz vein-type ore bodies, namely molybdenite ore and chalcopyrite–molybdenite ore. Ore minerals are irregular flakes, xenomorphic granular, or disseminated in quartz veins. The mineral composition is simple in the molybdenite ore, and molybdenite predominates in ore minerals, with some pyrite and bismuthine. Molybdenite and chalcopyrite are major ore minerals in chalcopyrite–molybdenite ore, together with minor pyrite, pyrrhotite, and bismuthine.
Scale-like molybdenite in the disseminated ores from the lower ore body is usually euhedral or sub-euhedral, scattered in the monzonitic granite, showing typical features of a porphyry-type ore deposit. The main ore minerals are molybdenite and minor chalcopyrite and pyrite. Copper is not economical in the disseminated ore due to its very low content.

2.3. Sampling

Twenty-two representative samples were selected from drill cores of the Wulandele Mo deposit (there are almost no fresh outcrops in this area due to strong weathering). Among them are seven quartz diorites (WL-6, WL-7, WL-9, WL-20, WL-29, WL-32, WL-33), seven granodiorites (WL-14, WL-15, WL-23, WL-24, WL-26, WL-27, WL-45), and eight biotite monzonitic granites (WL-10, WL-17, WL-22, WL-25, WL-30, WL-31, WL-34, WL-42). The samples were chosen with great care to have the fewest alterations, cracks, and inclusions.

3. Analytical Methods

3.1. Zircon U-Pb Dating

Zircon grains were selected from WL-17 (biotite monzonitic granite) and WL-20 (quartz diorite) samples via a conventional density and magnetic separation method at the Institute of Regional Geology and Resource Survey, Langfang, Hebei, China. To inspect the zircon crystal morphology and their internal textures, cathodoluminescence (CL) images and transmitted and reflected light microphotos were performed. LA-ICP-MS U–Pb dating was completed in the State Key Laboratory of Geological Processes and Mineral Resources, Beijing, China. The ICP-MS is an Agilent 7500a coupled with a UP 193 system. The analytical laser was 10 Hz in frequency with an approximately 36 μm ablation spot size. To improve transport efficiency, helium was employed as the carrier gas at a flow rate of 0.7 L/min [36,37]. The comprehensive experimental protocol was identical to that documented by Wang et al. [38]. U, Th, and Pb concentrations were calibrated using 29Si as an internal standard and NIST 610 glass as the reference standard. 207Pb/206Pb and 206Pb/238U ratios, calculated using GLITTER 4.4.1 of Macquarie University (http://www.glitter-gemoc.com, accessed on 1 June 2021), were corrected for both instrumental mass bias and depth-dependent elemental and isotopic fractionation using Harvard zircon 91500 as the external standard. The filtering of the U-Pb ages was mainly based on a 10% cut off of the U-Pb discordance [39,40]. Age calculations and Concordia plots were performed using Isoplot (ver. 3.0).

3.2. Whole-Rock Major and Trace Elements

Major, trace, and rare earth element (REE) compositions were measured at the Langfang Institute of Geophysical and Geochemical Exploration, Chinese Academy of Geological Sciences. Major element concentrations were assessed through X-ray fluorescence (XRF) spectrometry, while trace element compositions were quantified using an Agilent 7500a Inductively Coupled Plasma Mass Spectrometer (ICP-MS). To prepare the samples, approximately 60 milligrams of crushed whole-rock powder underwent a multi-step dissolution process. First, the powder was mixed with a 10:1 ratio of hydrofluoric acid (HF) and nitric acid (HNO3) in screw-top Teflon beakers, which was then heated to 100 °C for six days. The mixture was subsequently evaporated to dryness and further processed by refluxing in 7N HNO3 until nearly dry. Finally, the residue was dissolved in a 2% HNO3 solution, adjusting the sample-to-solution ratio to 1:1000 for analysis. The analysis precision was better than 3% for major oxides and 5% for trace and REE elements.

3.3. Pb Isotopes

Pb isotope compositions were tested at the Beijing Research Institute of Uranium Geology (BRIUG) under the guidance of GB/T17672-1999 (https://www.chinesestandard.net/PDF/English.aspx/GBT17672-1999, accessed on 21 May 2024). For isotope dilution, powder samples were combined and dissolved in sealed Teflon capsules on a heated plate for 24 h using HF + HNO3 + HClO4. After dilution with NHCl, Pb was purified and extracted through the traditional cation-exchange process, producing procedure blanks with Pb contents of less than 100 pg. An Isoprobe-T thermal ionization mass spectrometry was applied with the standard sample NBS 981 in the experiment. The testing precision for the Pb isotope is better than ±0.09‰.

4. Results

4.1. LA-ICP-MS Zircon U-Pb Age

Zircons from quartz diorite (WL20) and biotite monzonitic granite (WL17) are euhedral, with a length range of 50–200 lm and a length/width ratio of 1.3:1 and 4:1, respectively. In the CL pictures, most of the zircons exhibit obvious zoning structures, showing a sign of magmatic origin [41,42,43]. The zircon LA-ICP-MS data are presented in Table A1 in Appendix.
Measured Th and U contents of zircons from WL20 range from 109–437 ppm to 101–300 ppm, with relative constant Th/U ratios of 1.03–1.46. A total of 23 spots give a weighted mean age of 282.5 ± 2.4 Ma (MSWD = 0.14) with 206Pb/238U ages ranging from 278 ± 6 Ma to 287 ± 6 Ma (Figure 4).
The examined zircon grains from WL17 exhibit wide variability in their Th (46–4158 ppm) and U (54–20,760 ppm) contents. Their Th/U ratios generally vary from 0.43 to 1.44 (except for the three below 0.4). Excluding seven obvious old ages, the weighted mean age fitting the 18 remaining data is computed to be 135.4 ± 2.1 Ma (MSWD = 1.7, 95% confidence) (Figure 4).

4.2. Major and Trace Elements

The major and trace element contents of different rock types from the Wulandele Mo deposit are listed in Table A2. Most samples fit into the fields of gabbro, diorite, and granite in the total alkali (K2O + Na2O) vs. silica (SiO2) classification diagram (Figure 5a). No pyroxenes have been observed in samples of the gabbro area; thus, they are named after their mineral composition as quartz diorite. Most of the samples are classified as shoshonitic series and high-K calc-alkaline series according to the SiO2 vs. K2O diagram (Figure 5b). Major elements are also displayed in the Harker diagrams (Figure 6). In most of the Harker diagrams, the Permian samples (quartz diorite and granodiorite) exhibit a continuous variation trend. MgO, Fe2O3, CaO, P2O5, Sr, and V linearly decrease with increasing SiO2, while Al2O3, K2O, Rb, Ba, Zr, and U present a clear increasing trend. The Cretaceous samples (biotite monzonitic granite) are similar to the Permian samples except that Al2O3, Ba, and Zr reveal an opposite trend compared to the Permian samples. In addition, contents of MgO, P2O5, and V are constant for Cretaceous samples, but K and U, especially U are much more concentrated in Cretaceous samples than in Permian samples. Robust associations between selected main oxides and SiO2 indicate notable fractionations that occurred during the evolution of magma.
REE patterns (chondrite-normalized) are displayed in Figure 7a. The REE distribution patterns of all rock types from the Wulandele Mo deposit are distinguished by a fractionation between LREE and HREE and a negative Eu anomaly. The Permian quartz diorite and granodiorite exhibit enrichment in REE than that of biotite monzonitic granite (total REE, 188–464 ppm). The biotite monzonitic granite samples exhibit a much more fractionated feature ((La/Yb)N = 2.9–32.3) with strong negative Eu-anomalies, low total REE from 72 to 179 ppm, and extremely low Eu/Eu* of 0.11 to 0.50. The negative Eu anomalies may be a suggestion that plagioclase has been removed by crystal fractionation or that the plagioclase is a residual mineral during the process of partial melting. Increases in Rb with increasing SiO2, together with decreases in Sr and Ba, may also corroborate the plagioclase fractionation.
Most of the samples exhibit positive Rb, Th, U, La, Ce, and Hf anomalies and negative Sr, Ba, P (except the positive P anomaly in quartz diorite), and Ti anomalies in the primitive-mantle-normalized spidergram (Figure 7b). These anomalies might imply the fractional crystallization of feldspar, apatite, and ilmenite. Cretaceous biotite monzonitic granite samples stand out from Permian quartz diorite and granodiorite samples by a much higher concentration of U and depletion of Sr, P, and Ti, indicating a more crustal component involvement during the magma evolution.

4.3. Pb Isotope

Each of the three samples for Permian quartz diorite, Permian granodiorite, and Cretaceous monzonitic granite are tested for Pb isotopes and the results are listed in Table A3. According to the U–Pb ages of this study, the initial isotopic compositions were back-calculated to 283 Ma and 135 Ma.
The range of Pb isotopes is small with a notable variation in 208Pb/204Pb (38.124–39.589) and 206Pb/204Pb (18.510–19.388) at relatively constant 207Pb/204Pb ratios (15.507–15.597). The initial values of 206Pb/204Pb, 207Pb/204Pb, and 208Pb/204Pb for the measured ages vary from 18.048 to 18.892, 15.488 to 15.571, and 37.066 to 38.441. There is no clear difference between Permian samples and Cretaceous samples.

5. Discussion

5.1. Timing of Granitoids from WLDL

Voluminous granitoids characterize the CAOB and cover a long geohistory from the Ordovician to Cretaceous [46,47]. Our new U-Pb zircon data show that granitoids in the Wulandele Mo deposit are the products of Permian magmatism and Cretaceous magmatism.
Permian granitic plutons are widely distributed in the East Ujumqin area [48,49]. Hong et al. pointed out four ca. 286 Ma (Rb-Sr isochron) alkaline granitic plutons (Baiyinwula, Zuhengdeleng, Zuhengdeleng, and Zhanawula granites) to the east of Baiyinwula [48]. Zhang et al. reported a 284 ± 5 Ma age (zircon U-Pb) of the Jingesitai alkaline granitic pluton in the northern area of East Ujumqin [50]. Zhang et al. acquired four zircon U-Pb ages of ca. 290 Ma for peralkaline granitoids from the Baiyinwula district [51]. Tong et al. carried out a detailed study on four alkaline granitic plutons in this area (Hongol, Saiyinwusu, Baolag, and Baiyinwula). U-Pb zircon data also point to an Early Permian (ca. 280 Ma) magmatism [52]. Our zircon U-Pb dating on the quartz diorite from the Wulandele Mo deposit yielded a U-Pb age of 282.5 ± 2.4 Ma. It is identical to the ages mentioned above, placing the intrusive age of the quartz diorite in Early Permian.
Abundant granites were intruded in the GXR and its neighboring areas [53,54,55] during the Late Jurassic–Early Cretaceous. Especially, the Early Cretaceous witnessed a period of A-type granitic magmatism, including the Baerzhe, Baishileizi, and Nianzishan plutons [56]. This period agrees with the magmatic climax in the XMOB [53,57,58,59]. Our new U-Pb dating on the monzonitic granite gave a Concordia age of 135.4 ± 2.1 Ma, which is identical to the Re-Os age (134.1 ± 3.3 Ma) obtained by Tao et al. (2010) [31], indicating that the Wulandele Mo deposit formed in the Early Cretaceous.

5.2. Petrogenesis and Magma Source

Both the Permian and Cretaceous samples have high K, Na, Rb, U, and Th content and bear characteristics of the calc-alkaline series. Permian quartz diorites are metaluminous, while Permian granodiorites and Cretaceous granites are peraluminous (normative corundum < 1%, A/CNK < 1.1, and A/NK > 1), indicating that they are not S-type granotoids (Figure 8a). While the anti-correlation between P2O5 concentration and SiO2 content (not clear in the cretaceous samples) is indicative of I-type igneous rocks for the Permian sample (Figure 8b). In the classification diagram of Nb-SiO2, all the Permian data fit into the I-type area while most Cretaceous granites belong to the A-type field. Also, Cretaceous granites exhibit features of the A-type granites with high SiO2, alkali, Ga, Zn, and Y contents and very low CaO, MgO, and Sr contents. In the diagrams of Ce, Zr, and K2O + Na2O vs. 10,000 Ga/Al [60] (Figure 9), all Cretaceous samples fall within the category of A-type granites due to their high Ga/Al ratios [60]. The zircon saturation temperatures (often lower than the initial melt temperature) of the Cretaceous granites calculated with equation proposed by Watson [61] cover a range of 723–795 °C, with an average of 768 °C (excluding one exception low data). The crystallization pressures of the Cretaceous granites obtained through equation with normative quartz (Qtz) content [62] vary from 114.2 Mpa to 370.1 Mpa (ca. 3.8–12.3 km). Thus, the Cretaceous monzonitic granites formed in a low-pressure relative high-temperature condition, which is in accordance with the formation of A-type granites, implying its A-type affinity. In addition, Y/Nb ratios of the Cretaceous granite (0.67–0.86 with an average of 0.76) suggest that it belongs to A1-type granite as suggested by Eby [63]. In summary, the Permian samples bear I-type affinity and the Cretaceous samples belong to A-type granite.
Cretaceous granites are featured with high silica content (72.82–79.78 wt.%) and high K2O content (3.95–4.81 wt.%), which is similar to post-orogenic granite [64], suggesting a major crustal origin. The Rb/Sr (2.28–13.24), Ti/Y (16–98, except WL31 with the value of 119), and Ti/Zr (7–11) ratios of Cretaceous samples are within the limits of the crust source (>0.5, <100, and <20) [65,66], showing that crustal partial melting is the source of the magma. The crustal source for the Cretaceous granites is slightly different from its A1 classification, which is mainly derived from the mantle. Permian samples, particularly quartz diorites, show higher Ti/Y (236–430) and Ti/Zr (38–160) and lower Rb/Sr (0.11–0.48) ratios, indicating a deeper source than the Cretaceous monzonitic granite and an increase in the mantle component. Furthermore, the Nb/Ta ratios of Cretaceous samples (4.94–9.54, with an average value of 7.01) point to a crustal source (<11.4) [67]. The Nb/Ta ratios of the Permian samples (11.06–24.44, average at 15.34) are higher than the crustal value and close to that of the primitive mantle (17.8) [68], implying a significant add-in of mantle material in the magma evolution process.
Similarly to the Pb model diagram (Figure 10a,b), most plots fall between the orogenic and upper mantle evolution lines. Additionally, these plots parallel to the Northern Hemisphere Reference Line (NHRL), which is frequently taken to mean that there is a mixed source for lead [69,70,71]. Although the plots are scattered in the initial 208Pb/204Pb-versus-initial 206Pb/204Pb diagram (Figure 10b), most samples fit into the MORB field in the initial 207Pb/204Pb-versus-initial 206Pb/204Pb diagram (Figure 10a), indicating that the magma source might contain mantle components.
As mentioned above, Harker diagrams (Figure 6) exhibit strong linear patterns in most oxides and elements, which are thought to be related to the fractional crystallization process. Diagrams showing a strong relation between the major and trace elements and SiO2 are in accord with the fractionation of feldspar, plagioclase, apatite, and ilmenite that occurred throughout the evolution of magma. Fractional crystallization was important to the formation of Permian granitoids while the partial melting process was crucial to the genesis of Cretaceous granites; this is also shown in the La/Yb-vs.-La diagram (Figure 11). However, the low Nb/Ta (17.31–34.69, with an average of 25.10) and Zr/Hf (4.92–9.54, with an average of 7.01) ratios show that the Cretaceous granites also experienced a certain degree of differentiation.

5.3. Implication for Tectonic Setting and Porphyry Mo Mineralization

It is important to know the regional geotectonic evolution to discuss the intrusion of the granitoids as well as the formation of the Wulandele Mo deposit. However, it has long been a contentious subject when the PAO closed in XMOB. Different final closure times of the PAO have been put forward, including the Early Permian [9,48,72,73], the Permian–Triassic [74,75,76], the Triassic [77], and the Early Cretaceous [78]. A diachronous closure of the PAO was suggested by Xiao et al. [9] and Wilde [79], firstly shutting down near the Tarim Craton in the west and finally near Changchun in the east.
As to the Erenhot area, Wu et al. [54] put forward a five-stage modal. The Early Permian–Middle Permian (290–260 Ma) was controlled by a subduction regime, while the Triassic (<250 Ma) was a post-collisional tectonic environment. Li et al. [80] indicate that both sides of the PAO have experienced intense Permian igneous activity resulting from a subduction scenario. Considering the geochemical features and chronology data mentioned above, we prefer that Permian high-K calc-alkaline granitoids to be generated in a continental arc setting, while cretaceous samples belong to a within-plate origin.
In the classical tectonic discrimination diagram of Ta vs. Yb and Nb vs. Y (Figure 12) [81,82], most plots of the Permian quartz diorite fall into the volcanic arc granite (VAG) field. While Permian granodiorite samples are plotted near the tri-point area among VAG, Syn-COLG, and WPG fields. The Cretaceous monzonitic granite samples stride across the collision granite (COLG) field and the within-plate granite (WPG) field, although some data are plotted into the Syn-COLG fields. The monzonitic granite samples align with the field of “slab failure” proposed by Whalen and Hildebrand [82]. Considering the regional geological evolution, we prefer that the Permian samples originated from a volcanic arc tectonic setting, while the Cretaceous samples formed in a transitional tectonic setting from post-collisional to within-plate.
As mentioned above, strong Cretaceous magmatic activity has been confirmed in the GXRs [53,54,55]. At the same time, large-scale mineralization also developed in the study area, including the large Chaobuleng Fe-polymetallic skarn-type deposit [83], the large Wurinitu Mo-W deposit (two fine-grained granite zircon U-Pb ages of 139.3 ± 1.4 Ma and 139.2 ± 1.7 Ma [84]), the Huanggang skarn Sn-Fe deposit (zircon U-Pb age of the K-feldspar granite and porphyritic granite at 136.7 ± 1.1 Ma and 136.8 ± 0.57 Ma [73]), and the Xiaodonggou porphyry Mo deposit (porphyritic granite SHRIMP zircon age of 142 ± 2 Ma [85]).
The late Triassic to early Jurassic marked the end of the orogenic phase of the XOB. Consequently, it is necessary to link Cretaceous A-type granites to a different tectonic regime rather than the tectonic development of this massive Paleozoic orogenic belt [52]. From the Early Jurassic, the Paleo-Pacific plate started westward subduction, which was responsible for the extensive formation of I-type granites that extended far inland in the GXR. From ~140 Ma, the Paleo-Pacific plate began to retreat eastward, creating an extensional environment in the Early Cretaceous, which was exacerbated by localized lithosphere thinning that is typically linked to delamination [75]. The Cretaceous granites of the Wulandele Mo deposit are most likely the result of highly fractionated slab-breakoff magmas evolving into the crustal A-type granite. Thus, the Wulandele Mo deposit, which is different from classical subduction-related porphyry-type deposits (continental-arc-related, e.g., Endako [86] and MAX [87], or back-arc rift-related, e.g., Climax [88,89]), is the result of a special A-type magmatism in this tectonic setting.
A-type-related porphyry molybdenum deposits have long been recognized [88,89,90,91], with the Climax-type Mo deposits being the most prominent and well-documented examples. The Climax-type Mo deposits are linked to siliceous, incompatible element-bearing and F-rich oxidized magmas, resembling certain aspects of A-type granites [88]. According to Keith et al., the high grade of molybdenum is likely inherited from magma with concentrations as high as 4 to 5 ppm (about 3 times more than the Clark value of 1.5 ppm). This high molybdenum content is associated with elevated magmatic oxygen fugacity levels [89]. The ore-hosting rocks at Climax exhibit enrichments in elements like fluorine, rubidium, niobium, tantalum, and rare metals, indicating that extensive magmatic differentiation played a significant role in the formation of these deposits [89].
The ore-forming related Cretaceous monzonitic granite of the Wulandele Mo deposit exhibits low molybdenum content, around 1 ppm. Thus, there must be other factors controlling the enrichment of molybdenum than source control. The high contents of SiO2 and K2O of the monzonitic granite signify a highly evolved magma. This is also evident from the elevated K/Rb and Rb/Sr ratios and decreased Zr/Hf and Nb/Ta ratios. The Fe2O3/FeO ratio of seven monzonitic samples varies between 0.61 and 1.18, averaging at 0.83, suggesting a moderately oxidized magma [92]. The presence of oxidized, highly evolved magma appears crucial to the formation of the large Wulandele Mo deposit, similar to the Climax-type deposits. However, the lower molybdenum content in the monzonitic granite contrasts with the Climax-type deposits. Further investigation is needed to discern subtle differences between A-type porphyry molybdenum deposits formed in rift settings and those in intraplate settings.

6. Conclusions

Based on the geological, geochronological, and Pb isotope data presented above, the following conclusions can be drawn:
  • U-Pb zircon data indicate that the quartz diorite and monzonitic granite crystallized at 282.5 ± 2.4 Ma (N = 23, MSWD = 0.14) and 135.4 ± 2.1 Ma (N = 18, MSWD = 1.7), respectively. These ages agree with geological observations that Permian granitoids (quartz diorite and granodiorite) are the host rocks of the Wulandele Mo deposit, while Cretaceous granite is the causative intrusion.
  • Geochemical and isotopic data reveal that the high-K calc-alkaline Permian granitoids, classified as I-type intrusions, underwent extensive fractional crystallization in a continental arc environment. In contrast, the Cretaceous monzonitic granite, characterized as A-type, resulted from partial melting of the crust within an intraplate tectonic setting.
  • Distinct from the majority of classical porphyry deposits, which are primarily associated with I-type granitoids, the Wulandele porphyry Mo deposit stands out as a product of an intraplate Cretaceous A-type magmatic event. The widespread presence of A-type intrusions in the GXR region implies that A-type granite-related porphyry Mo deposits should be considered as important targets for future exploration.

Author Contributions

Conceptualization, J.W.; data curation, J.Z. and Z.L.; funding acquisition, J.W.; investigation, J.W., J.Z., Z.L., Y.Z. and F.Z.; methodology, Z.L.; software, J.Z.; supervision, J.W.; visualization, F.Z.; writing—original draft, J.W.; writing—review and editing, Z.L. and F.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundations of China (No. 92062219, 42003032, 41272106) and the Inner Mongolia Academician Project (2022-TZH03).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

We are grateful to Shouguang Wang for his help in accessing the Wulandele exploration program.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

Table A1. Major (%) and trace element (ppm) compositions of the granitoids from the Wulandele Mo deposit.
Table A1. Major (%) and trace element (ppm) compositions of the granitoids from the Wulandele Mo deposit.
SampleWL-6WL-7WL-9WL-20WL-29WL-32WL-33WL-14WL-15WL-23WL-24
LithologyQuartz DioriteGranodiorite
SiO246.045053.9749.8746.6547.4748.7764.6166.1264.2162.14
Al2O313.6113.8815.7314.5712.8414.3214.716.2515.6515.8615.03
Fe2O34.595.13.383.254.444.33.552.041.681.941.36
FeO9.568.326.588.799.948.918.793.243.214.023.32
MgO5.934.644.365.035.855.354.51.561.461.781.38
CaO9.397.656.77.989.019.156.472.942.813.262.57
Na2O2.532.883.233.482.722.943.023.253.423.062.8
K2O1.211.852.161.31.31.341.544.113.733.584.36
MnO0.210.20.160.20.230.240.240.090.120.110.1
P2O52.251.590.841.691.931.991.930.320.340.340.27
TiO23.993.171.853.244.493.523.370.780.730.90.7
LOI0.40.450.780.410.370.212.840.620.590.775.89
Total99.7299.7399.7599.7999.899.7399.7299.8199.8599.8399.94
σ4.223.082.583.234.23.943.022.492.22.062.52
DI30.9340.9947.1840.0333.3734.8142.367374.4269.5174.23
A/CNK0.60.670.790.670.580.630.81.071.061.071.07
A/NK3.642.942.923.053.193.353.232.212.192.392.1
Rb898192507973189169262149193
Ba1923574042992622661499889288961025
Th30.816.512.613.81825.311.99.41419.817.2
U1.11.92.21.71.61.934.34.12.73.5
Nb12.219.118.612.715.815.326.421.421.622.921.9
Ta0.921.361.180.961.111.071.081.441.341.411.49
K10,04515,35817,93110,79210,79211,12412,78434,11930,96529,71936,195
Pb8.110.113.89.17.47.98.7252221.926
Sr446366415444359494391244256269257
P98196939366673758423868584231397148414841178
Zr145203295214206144169383458487430
Hf8.212.613.712.811.489.917.21818.919.8
Sc33.634.426.926.332.435.231.114.7151715.4
Ti23,92019,00411,09119,42426,91821,10320,2034676437653964197
Y57.2766.646.9854.262.576070.1749.3644.8754.9650.93
V49934122134847336229765647360
Cr2.52.624.33.83.42.52.518.217.82215.2
Co52.837.530.639.551.943.230.598.59.88.6
Ni15.312.120.711.417.810.37.211.210.812.811.1
La5057434651516033545147
Ce119138981091271241466711210794
Pr17.519.21315.518.217.719.98.813.813.412.2
Nd8386577184839138555649
Sm17.217.611.714.717.817.318.48.510.311.710.3
Eu3.53.582.433.253.483.73.522.392.322.522.36
Gd14.515.4210.3312.7615.2915.0415.838.159.0510.439.21
Tb2.182.341.621.922.262.242.371.441.431.721.54
Dy11.2512.668.9410.3311.9211.4912.558.387.929.878.93
Ho2.042.331.671.892.192.12.341.641.521.911.71
Er5.386.374.515.085.775.526.224.744.285.254.79
Tm0.760.940.70.720.810.780.910.780.670.810.77
Yb4.465.594.314.354.914.695.464.794.095.034.74
Lu0.640.830.640.620.740.690.80.710.610.760.71
ΣREE331.28368.65257.44297.02344.66339.73384.91187.27277.8277.55247.02
LREE/HREE7.046.936.876.896.856.997.285.118.46.766.62
LaN/YbN8.087.337.157.547.447.87.914.889.547.227.09
δEu0.660.650.660.710.630.690.610.860.720.680.73
SiO265.6766.1768.2576.0275.1373.9372.9274.2672.8579.7876.65
Al2O315.7415.5915.0112.9913.7313.914.5514.114.3811.813.13
Fe2O31.561.411.40.430.460.630.520.510.650.110.29
FeO3.493.442.730.740.390.650.850.720.820.160.27
MgO1.451.431.260.320.190.330.370.20.390.120.1
CaO2.642.562.30.930.73111.011.180.220.59
Na2O3.142.923.043.183.773.583.654.13.612.313.8
K2O4.354.64.284.544.714.484.744.074.813.954.44
MnO0.10.090.090.040.030.040.040.030.050.010.02
P2O50.260.260.220.120.050.120.10.130.120.110.04
TiO20.690.720.590.190.110.190.20.190.20.070.08
LOI0.740.70.690.470.511.060.880.520.841.270.53
Total99.8299.8899.8699.9799.8399.999.8399.8599.8899.9399.94
σ2.462.432.111.82.242.092.352.132.371.062.01
DI74.5575.1178.1291.8293.6691.1590.591.7890.079595.08
A/CNK1.071.081.091.11.091.111.121.091.081.391.08
A/NK2.12.072.051.681.621.731.731.721.711.891.59
Rb163172211336408422400303396297331
Ba10279888242791744435391985445235
Th31.320.821.210.812.718.520.316.5265.311
U2.34.35.111.618.99.817.87.37.719.814.9
Nb20.319.918.815.624.515.717.826.614.635.419.7
Ta1.251.221.72.564.381.692.174.031.537.193.36
K36,11238,18735,53037,68939,10037,19139,34933,78739,93032,79136,859
Pb25.726.524.425.927.129.731.125.72729.332.2
Sr2432352149762134122931742125
P11351135960524218524436567524480175
Zr439378363119841461471361705045
Hf18.215.113.54.74.14.64.55.64.93.52.6
Sc16.211.212.133.62.63.72.12.33.62.5
Ti41374316353711396591139119911391199420480
Y52.8339.8635.0312.4519.5411.6213.2920.210.126.7616.17
V67675515101717171797
Cr15.115.811.72.13.32.92.52.92.52.54.5
Co8.98.67.61.50.71.81.31.11.80.20.5
Ni11.311.210.62.62.32.732.532.12.3
La9051632221313338371112
Ce2101101134441565876642729
Pr22.813.713.65.356.56.897.33.43.8
Nd8754532018242434261315
Sm159.99.93.84.24.24.26.84.33.74
Eu2.52.181.910.480.310.560.60.490.630.130.22
Gd12.228.348.172.883.393.183.194.983.173.183.07
Tb1.871.31.240.430.60.440.460.730.430.630.52
Dy9.647.056.682.053.252.012.253.691.883.82.85
Ho1.781.331.220.360.580.350.380.650.320.710.5
Er5.133.733.281.021.740.951.061.860.862.141.46
Tm0.780.590.450.170.30.150.170.320.130.390.25
Yb4.853.782.621.172.080.991.072.250.822.731.64
Lu0.730.580.370.170.310.150.160.330.110.410.25
ΣREE464.58267.48277.21103.65102.13130.3134.53179.44147.2772.2974.52
LREE/HREE11.569.0210.5311.597.3414.8514.3911.1218.14.176.07
LaN/YbN13.379.6817.1813.647.1622.5521.7612.232.72.855.25
δEu0.550.710.630.420.250.450.480.250.490.110.18
Table A2. Zircon U-Pb isotopic data for quartz diorite (WL20) and biotite monzonitic granite (WL17).
Table A2. Zircon U-Pb isotopic data for quartz diorite (WL20) and biotite monzonitic granite (WL17).
SpotsPbUThU/PbTh /U207Pb/206Pb207Pb/235U206Pb/238U206Pb/238U%U-Pb disc
WL20-019.415120116.06 1.340.05210.001560.32480.009790.045210.0006928570.35
WL20-025.910110917.12 1.080.051960.001930.315530.011680.044040.000727880.00
WL20-0317.430031417.24 1.050.051960.001230.323110.007850.04510.0006628470.00
WL20-0510.116122515.94 1.400.052120.001520.323640.009540.045040.0006828470.35
WL20-067.411616915.68 1.460.051920.001760.323070.010950.045130.000712857−0.35
WL20-076.210213016.45 1.270.052010.001860.32170.01150.044860.000728370.00
WL20-0818.429739816.14 1.340.052040.001250.320350.007870.044640.0006628270.00
WL20-0917.727043715.25 1.620.051930.001260.319510.007890.044620.0006628170.36
WL20-1011.918926115.88 1.390.052020.001390.322250.008750.044930.0006728370.35
WL20-118.112517915.43 1.440.053360.001830.3310.011370.044990.000728482.11
WL20-128.713819815.86 1.430.051860.001780.3170.010860.044330.000728070.00
WL20-1311.317925215.84 1.410.052040.001470.319630.009140.044540.0006728170.36
WL20-148.313816416.63 1.190.051920.001640.320950.010160.044830.0006928370.00
WL20-157.111316015.92 1.420.052040.001940.321930.012010.044870.0007128370.00
WL20-1611.518026015.65 1.440.051970.001380.321320.008640.044840.0006828370.00
WL20-1712.020421017.00 1.030.052040.001420.323320.008890.045060.0006828480.00
WL20-189.114719016.15 1.300.052080.001560.320960.009660.04470.0006828270.35
WL20-209.615720016.35 1.270.051980.001540.321390.009610.044850.0006828370.00
WL20-2115.023334015.53 1.460.051870.001270.320230.008030.044770.0006628270.00
WL20-2211.518025015.65 1.390.051980.001410.321340.00880.044830.0006728380.00
WL20-2311.317725415.66 1.430.051910.001440.32040.008980.044770.0006828280.00
WL20-2418.229634916.26 1.180.052040.001220.326260.007850.045470.0006628770.00
WL20-2515.322837814.90 1.660.051940.001290.32110.008120.044840.0006628370.00
WL17-0119.780539240.86 0.490.048820.001280.146180.003860.021710.0003213840.72
WL17-032.4878736.25 1.000.051060.004170.149840.012140.021280.000413661.45
WL17-048.331629138.07 0.920.048340.001640.139890.004730.020990.000331344−2.17
WL17-0521.081453438.76 0.660.048360.001270.146190.003870.021930.000331404−2.16
WL17-077.417416723.51 0.960.046050.006830.129790.01910.020440.000371308−4.58
WL17-089.629228630.42 0.980.046050.002020.125970.005210.019840.000312735.36
WL17-1110.441825940.19 0.620.04880.001660.142840.004820.021230.0003313550.74
WL17-1438.1141794937.19 0.670.046830.00270.140130.007780.02170.000331383−2.90
WL17-1588.53242306036.63 0.940.054720.001410.16130.004160.021380.0003213648.82
WL17-1663.82497126639.14 0.510.052470.001370.15980.004170.022090.0003314156.43
WL17-1733.3124385537.33 0.690.051580.001380.158280.004240.022260.0003314256.43
WL17-18164.86708286940.70 0.430.05470.002950.15730.008110.020860.0003313322.59
WL17-1972.22966110441.08 0.370.052920.002550.156220.007160.021410.0003213727.30
WL17-218.225536931.10 1.440.05350.006940.149560.01920.020280.0003812922.98
WL17-22459.020670415845.03 0.200.048480.001280.141550.003750.021180.000321354−0.74
WL17-23229.410159156344.29 0.150.050360.001890.146960.005070.021160.0003213522.92
WL17-241.5594639.33 0.790.051530.004470.152830.013080.021510.00045137100.00
WL17-251.5545236.00 0.960.05180.006360.152660.018580.021380.0004713692.19
Table A3. Whole-rock Pb isotopic compositions for the granitoids from the WLDL Mo deposit.
Table A3. Whole-rock Pb isotopic compositions for the granitoids from the WLDL Mo deposit.
Sample no.WL-7WL-9WL-11WL-14WL-26WL-27WL-30WL-34WL-42
LithologyQuartz DioriteGranodioriteMonzonitic Granite
U(ppm)1.92.21.74.32.34.37.319.814.9
Th(ppm)16.512.613.89.431.320.816.55.311
Pb(ppm)10.113.89.12525.726.525.729.332.2
Age/Ma283283283283283283135135135
206Pb/204Pb18.75118.6818.58119.38818.72518.58218.5119.34419.024
207Pb/204Pb15.5515.56315.55215.59715.5715.5515.50715.57215.56
208Pb/204Pb38.71538.50238.47138.43339.58938.75338.23938.12438.152
(206Pb/204Pb)i18.21118.22418.04818.89218.46518.11718.12918.42918.400
(207Pb/204Pb)i15.52215.53915.52415.57115.55615.52615.48815.52715.53
(208Pb/204Pb)i37.19337.65437.06638.08138.44138.02337.95838.04438.001

References

  1. Sengör, A.M.C.; Natal’in, B.A.; Burtman, V.S. Evolution of the Altaid tectonic collage and Paleozoic crustal growth in Eurasia. Nature 1993, 364, 299–307. [Google Scholar] [CrossRef]
  2. Badarch, G.; Dickson, C.W.; Windley, B.F. A new terrane subdivision for Mongolia: Implications for the Phanerozoic crustal growth of Central Asia. J. Asian Earth Sci. 2002, 21, 87–110. [Google Scholar] [CrossRef]
  3. Windley, B.F.; Alexeiev, D.V.; Xiao, W.; Kröner, A.; Badarch, G. Tectonic models for accretion of the Central Asian Orogenic Belt. J. Geol. Soc. Lond. 2007, 164, 31–47. [Google Scholar] [CrossRef]
  4. Khain, E.V.; Bibikova, E.V.; Salnikova, E.B.; Kröner, A.; Gibsher, A.S.; Didenko, A.N.; Degtyarev, K.E.; Fedotova, A.A. The Palaeo-Asian Ocean in the Neoproterozoic and early Paleozoic: New geochronologic data and palaeotectonic reconstructions. Precambrian Res. 2003, 122, 329–358. [Google Scholar] [CrossRef]
  5. Kröner, A.; Windley, B.F.; Badarch, G.; Tomurtogoo, O.; Hegner, E.; Jahn, B.M.; Gruschka, S.; Khain, E.V.; Demoux, A.; Wingate, M.T.D. Accretionary growth and crust-formation in the Central Asian orogenic belt and comparison with the Arabian-Nubian shield. In 4-D Frame Work of Continental Crust; Memoirs Hatcher, R.D., Carlson, M.P., McBride, J.H., Martínez Catalán, J.R., Eds.; Geological Society of America: Boulder, CO, USA, 2007; Volume 200, pp. 181–209. [Google Scholar]
  6. Rytsk, E.Y.; Kovach, V.P.; Yarmolyuk, V.V.; Kovalenko, V.I. Structure and evolution of the continental crust in the Baikal Fold Region. Geotectonics 2007, 41, 440–464. [Google Scholar] [CrossRef]
  7. Tu, G.Z. A preliminary discussion on the Central Asian metallogenic belt. Sci. Geol. Sin. 1999, 34, 397–404. [Google Scholar]
  8. Xiao, W.J.; Windley, B.F.; Hao, J.; Zhai, M.G. Accretion leading to collision and the Permian Solonker suture; Inner Mongolia; China: Termination of the central Asian orogenic belt. Tectonics 2003, 22, 1069. [Google Scholar] [CrossRef]
  9. Xiao, W.J.; Kroner, A.; Windley, B. Geodynamic evolution of Central Asia in the Paleozoic and Mesozoic. Int. J. Earth Sci. 2009, 98, 1185–1188. [Google Scholar] [CrossRef]
  10. Wainwright, A.J.; Tosdal, R.M.; Wooden, J.L.; Mazdab, F.K.; Friedman, R.M. U–Pb (zircon) and geochemical constraints on the age; origin; and evolution of Paleozoic arc magmas in the Oyu Tolgoi porphyry Cu–Au district, southern Mongolia. Gondwana Res. 2011, 19, 764–787. [Google Scholar] [CrossRef]
  11. Shen, P.; Shen, Y.C.; Pan, H.D.; Li, X.H.; Dong, L.H.; Wang, J.B.; Zhu, H.P.; Dai, H.W.; Guan, W.N. Geochronology and isotope geochemistry of the Baogutu porphyry copper deposit in the west Junggar region, Xinjiang, China. J. Asian Earth Sci. 2012, 49, 99–115. [Google Scholar] [CrossRef]
  12. Zeng, Q.D.; Qin, K.Z.; Liu, J.M.; Li, G.M.; Zhai, M.G.; Chu, S.X.; Guo, Y.P. Porphyry molybdenum deposits in the Tianshan–Xingmeng orogenic belt; northern China. Int. J. Earth Sci. 2014, 104, 991–1023. [Google Scholar] [CrossRef]
  13. 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. 2013, 25, 48–102. [Google Scholar] [CrossRef]
  14. Gao, J.; Klemd, R.; Zhu, M.T.; Wang, X.S.; Li, J.L.; Wan, B.; Xiao, W.J.; Zen, Q.D.; Shen, P.; Sun, J.G.; et al. Large-scale porphyry-type mineralization in the Central Asian metallogenic domain: A review. J. Asian Earth Sci. 2018, 165, 7–36. [Google Scholar] [CrossRef]
  15. Mao, J.W.; Pirajno, F.; Lehmann, B.; Luo, M.C.; Berzina, A. Distribution of porphyry deposits in the Eurasian continent and their corresponding tectonic settings. J. Asian Earth Sci. 2014, 79, 576–584. [Google Scholar] [CrossRef]
  16. Seltmann, R.; Mike Porter, T.; 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]
  17. Liu, J.M.; Zhang, R.; Zhang, Q.Z. The regional metallogeny of Da Hinggan Ling, China. Earth Sci. Front. 2004, 11, 269–277, (In Chinese with English abstract). [Google Scholar]
  18. Han, C.M.; Xiao, W.J.; Zhao, G.C.; Su, B.X.; Ao, S.J.; Zhang, J.; Zhang, J.Y.; Wan, B.; Wang, Z.M. Tectonic implications of Re-Os dating of molybdenum deposits in the Tianshan–Xingmeng Orogenic Belt, Central Asia. Int. Geol. Rev. 2014, 56, 985–1006. [Google Scholar] [CrossRef]
  19. Zeng, Q.D.; Liu, J.M.; Chu, S.X.; Wang, Y.B.; Sun, Y.; Duan, X.X.; Zhou, L.L. Mesozoic molybdenum deposits in the East Xingmeng orogenic belt, northeast China: Characteristics and tectonic setting. Int. Geol. Rev. 2012, 54, 1843–1869. [Google Scholar] [CrossRef]
  20. 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]
  21. Hu, X.L.; Ding, Z.J.; He, M.C.; Yao, S.Z.; Zhu, B.P.; Shen, J.; Chen, B. A porphyry-skarn metallogenic system in the Lesser Xing’an Range, NE China: Implications from U–Pb and Re–Os geochronology and Sr–Nd–Hf isotopes of the Luming Mo and Xulaojiugou Pb–Zn deposit. J. Asian Earth Sci. 2014, 90, 88–100. [Google Scholar] [CrossRef]
  22. 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 GXR, NE China: Partial melting of the juvenile lower crust in intra-plate extensional environment. Lithos 2014, 202–203, 138–156. [Google Scholar] [CrossRef]
  23. Nie, F.J.; Sun, Z.J.; Li, C.; Liu, Y.F.; Lu, K.P.; Zhang, K.; Liu, Y. Re–Os isotopic dating of molybdenite separates from Chalukou porphyry Mo polymetallic deposit in Heilongjiang Province. Min. Deposits 2011, 30, 828–836. [Google Scholar]
  24. Shao, J.; Yang, H.Z.; Jia, B.; Peng, M.S. Geological characteristics and ore-forming age of Luming Mo deposit in Heilongjiang Province. Min. Deposits 2012, 31, 1301–1310. [Google Scholar]
  25. Ju, N.; Zhang, D.; Zhang, G.B.; Zhang, S.; Ren, C.T.; Ren, Y.S.; Wang, H.; Wu, Y.; Liu, X.; Shi, L.; et al. Geology and mineralization of the Daheishan supergiant porphyry molybdenum deposit (1.65 Bt), Jilin, China: A review. China Geol. 2023, 6, 494–530. [Google Scholar] [CrossRef]
  26. Leng, C.B.; Zhang, X.C.; Huang, Z.L.; Huang, Q.Y.; Wang, S.X.; Ma, D.Y.; Luo, T.Y.; Li, C.; Li, W.B. Geology, Re-Os ages, sulfur and lead isotopes of the Diyanqinamu porphyry Mo deposit, Inner Mongolia, NE China. Econ. Geol. 2015, 110, 557–574. [Google Scholar] [CrossRef]
  27. Zeng, Q.D.; Liu, J.M.; Zhang, Z.L.; Chen, W.J.; Zhang, W.Q. Geology and geochronology of the Xilamulun molybdenum metallogenic belt in eastern Inner Mongolia, China. Int. J. Earth Sci. 2011, 100, 1791–1809. [Google Scholar] [CrossRef]
  28. Zhang, L.C.; Wu, H.Y.; Wan, B.; Chen, Z.G. Ages and geodynamic settings of Xilamulun Mo-Cu metallogenic belt in the northern part of the North China craton. Gondwana Res. 2009, 16, 243–254. [Google Scholar] [CrossRef]
  29. Nie, F.J.; Jiang, S.H. Geological setting and origin of Mo-W-Cu deposits in the Honggor-Shamai district, Inner Mongolia, North China. Resour. Geol. 2011, 61, 344–355. [Google Scholar] [CrossRef]
  30. Shao, J.D.; Tao, J.X.; Li, S.W.; Shang, H.S.; Wu, L.W.; Gong, Z.Z. The new progress in ore prospecting within the Daxing’anling mineralization belt, China. Geol. Bull. China 2009, 28, 955–962, (In Chinese with English abstract). [Google Scholar]
  31. Tao, J.X.; Wang, T.; Cheng, Z.H.; Luo, Z.Z.; Xu, L.Q.; Hao, X.Y.; Cui, L.W. The Re–Os isotopic dating of molybdenite from the Wulandele molybdenum–copper polymetallic deposit in Sonid Zuoqi of Inner Mongolia and its geological significance. Rock Miner. Anal. 2009, 28, 249–253. (In Chinese) [Google Scholar]
  32. Zhang, F.F.; Wang, Y.H.; Liu, J.J.; Wang, J.C. Ore genesis and hydrothermal evolution of the Wulandele Mo deposit, Inner Mongolia, Northeast China: Evidence from geology, fluid inclusions and H–O–S–Pb isotopes. Ore Geol. Rev. 2018, 93, 181–199. [Google Scholar] [CrossRef]
  33. Zhang, X.H.; Yao, C.L.; Lentz, D.R.; Qin, Y.; Wei, Y.W.; Zhao, F.S.; Yang, Z.J.; Liu, R.; Zhang, Z.F. Re-Os Geochronology, Whole-Rock and Radiogenic Isotope Geochemistry of the Wulandele Porphyry Molybdenum Deposit in Inner Mongolia, China, and Their Geological Significance. Minerals 2020, 10, 374. [Google Scholar] [CrossRef]
  34. Nie, F.J.; Jiang, S.H.; Zhang, Y.; Liu, Y.; Hu, P. Geological features and origin of porphyry copper deposits in China-Mongolia border region and its neighboring areas. Min. Deposits 2004, 23, 176–189. [Google Scholar]
  35. Porter, T.M. The geology; structure and mineralisation of the Oyu Tolgoi porphyry copper-gold-molybdenum deposits, Mongolia: A review. Geosci. Front. 2015, 6, 1–60. [Google Scholar] [CrossRef]
  36. Eggins, S.M.; Kinsley, L.P.J.; Shelley, J.M.M. Deposition elemental fractionation processes during atmospheric pressure laser sampling for analysis by ICP–MS. Appl. Surf. Sci. 1998, 127, 278–286. [Google Scholar] [CrossRef]
  37. Jackson, S.E.; Pearson, N.J.; Griffin, W.L.; Belousova, E.A. The application of laser ablation-inductively coupled plasma–mass spectrometry to in situ U–Pb zircon geochronology. Chem. Geol. 2004, 211, 47–69. [Google Scholar] [CrossRef]
  38. Wang, W.; Liu, S.W.; Bai, X.; Li, Q.G.; Yang, P.T.; Zhao, Y.; Zhang, S.H. and Guo, R.R. Geochemistry and zircon U–Pb–Hf isotopes of the late Paleoproterozoic Jianping diorite–monzonite–syenite suite of the North China Craton: Implications for petrogenesis and geodynamic setting. Lithos 2013, 162, 175–194. [Google Scholar] [CrossRef]
  39. Schaltegger, U.; Davies, J.H.F.L. Petrochronology of Zircon and Baddeleyite in Igneous Rocks: Reconstructing Magmatic Processes at High Temporal Resolution. Rev. Mineral. Geochem. 2017, 83, 297–328. [Google Scholar] [CrossRef]
  40. Fornelli, A.; Festa, V.; Micheletti, F.; Spiess, R.; Tursi, F. Building an Orogen: Review of U-Pb Zircon Ages from the Calabria–Peloritani Terrane to Constrain the Timing of the Southern Variscan Belt. Minerals 2020, 10, 944. [Google Scholar] [CrossRef]
  41. Corfu, F.; Hanchar, J.M.; Hoskin, P.W.; Kinny, P. Atlas of zircon textures. Rev. Mineral. Geochem. 2003, 53, 469–500. [Google Scholar] [CrossRef]
  42. Fornelli, A.; Piccarreta, G.; Micheletti, F. In situ U-Pb dating combined with SEM imaging on zircon–an analytical bond for effective geological reconstructions. In Geochronology–Methods and Case Studies; Mörner, N.A., Ed.; IntechOpen Limited: London, UK, 2014; pp. 109–139. [Google Scholar]
  43. 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]
  44. Middlemost, E.A.K. Naming materials in the magma/igneous rock system. Earth-Sci. Rev. 1994, 37, 215–224. [Google Scholar] [CrossRef]
  45. Peccerillo, A.; Taylor, S.R. Geochemistry of eocene calc-alkaline volcanic rocks from the Kastamonu area, Northern Turkey. Contrib. Mineral. Petrol. 1976, 58, 63–81. [Google Scholar] [CrossRef]
  46. Jahn, B.M.; Wu, F.Y.; 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] [PubMed]
  47. Hong, D.W.; Wang, S.G.; Xie, X.L.; Zhang, J.S. Genesis of positive ε(Nd,t) granitoids in the Da Hinggan Mts.-Mongolia orogenic belt and growth continental crust. Earth Sci. Front. 2000, 7, 440–456, (In Chinese with English abstract). [Google Scholar]
  48. Hong, D.W.; Wang, S.G.; Han, B.F.; Jin, M.Y. The Permian alkaline granites in Central Inner Mongolia and their geodynamic significance. J. Southeast Asian Earth Sci. 1994, 10, 169–176. [Google Scholar]
  49. Hong, D.W.; Wang, S.G.; Han, B.F.; Jin, M.Y. Post-orogenic alkaline granites from China and comparisons with anorogenic alkaline granites elsewhere. J. Southeast Asian Earth Sci. 1996, 30, 13–27. [Google Scholar]
  50. Zhang, Y.Q.; Xu, L.Q.; Kang, X.L.; Bao, Y. Age dating of alkali granite in the Jingesitai area of Dongujimqin Banner; Inner Mongolia, and its significance. Geol. China 2009, 36, 988–992, (In Chinese with English abstract). [Google Scholar]
  51. Zhang, X.H.; Yuan, L.L.; Xue, F.H.; Yan, X.; Mao, Q. Early Permian A-type granites from central Inner Mongolia, North China: Magmatic tracer of post-collisional tectonics and oceanic crustal recycling. Gondwana Res. 2015, 28, 311–327. [Google Scholar] [CrossRef]
  52. Tong, Y.; Jahn, B.M.; Wang, T.; Hong, D.W.; Smith, E.I.; Sun, M.; Gao, J.F.; Yang, Q.D.; Huang, W. Permian alkaline granites in the Erenhot–Hegenshan belt, northern Inner Mongolia; China: Model of generation, time of emplacement and regional tectonic significance. J. Asian Earth Sci. 2015, 97, 320–336. [Google Scholar] [CrossRef]
  53. Zhang, J.H.; Gao, S.; Ge, W.C.; Wu, F.Y.; Yang, J.H.; Wilde, S.A.; Li, M. Geochronology of the Mesozoic volcanic rocks in the GXR, northeast China: Implications for subduction-induced delamination. Chem. Geol. 2010, 276, 144–165. [Google Scholar] [CrossRef]
  54. 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]
  55. 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 systems. J. Asian Earth Sci. 2013, 74, 167–193. [Google Scholar] [CrossRef]
  56. Wu, F.Y.; Sun, D.Y.; Li, H.M.; Jahn, B.M.; Wilde, S.A. A-type granites in NE China: Age and geochemical constraints on their petrogenesis. Chem. Geol. 2002, 187, 143–173. [Google Scholar] [CrossRef]
  57. 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]
  58. Wang, F.; Zhou, X.H.; Zhang, L.C.; Ying, J.F.; Zhang, Y.T.; Wu, F.Y.; Zhu, R.X. Late Mesozoic volcanism in the GXR (NE China): Timing and implications for the dynamic setting of NE Asia. Earth Planet. Sci. Lett. 2006, 251, 179–198. [Google Scholar] [CrossRef]
  59. Yang, W.B.; Niu, H.C.; Cheng, L.R.; Shan, Q.; Li, N.B. Geochronology, geochemistry and geodynamic implications of the Late Mesozoic volcanic rocks in the southern Great Xing’an Mountains, NE China. J. Asian Earth Sci. 2015, 113, 454–470. [Google Scholar] [CrossRef]
  60. 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]
  61. Watson, E.B.; Harrison, T.M. Zircon saturation revisited: Temperature and composition effects in a variety of crustal magma types. Earth Planet Sc. Lett. 1983, 64, 295–304. [Google Scholar] [CrossRef]
  62. Yang, X.M. Estimation of crystallization pressure of granite intrusions. Lithos 2017, 286, 324–329. [Google Scholar] [CrossRef]
  63. Eby, G.N. The A-type granitoids: A review of their occurrence and chemical characteristics and speculations on their petrogenesis. Lithos 1990, 26, 115–134. [Google Scholar] [CrossRef]
  64. Bernard, B.; Abla, A.S.; François, B.; Sandrine, F. Alkali-calcic and alkaline postorogenic (PO) granite magmatism: Petrologic constraints and geodynamic settings. Lithos 1998, 45, 45–70. [Google Scholar]
  65. Pearce, J.A. Role of sub-continental lithosphere in magma genesis at destructive plate margins. In Continental Basalts and Mantle Xenoliths; Hawkesworth, C.J., Norry, M.J., Eds.; Shiva Publishing Ltd.: Nantwich, Cheshire, 1983; pp. 230–249. [Google Scholar]
  66. Wilson, M. Igneous Petrogenesis; Kluwer Academic Publishers: Dordrecht, The Netherland, 1989; p. 466. [Google Scholar]
  67. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Oxford Press: New York, NY, USA, 1985; p. 312. [Google Scholar]
  68. McDonough, W.F.; Sun, S.S. The composition of the Earth. Chem. Geol. 1995, 120, 223–253. [Google Scholar] [CrossRef]
  69. Zhou, Z.H.; Mao, J.W.; Lyckberg, P. Geochronology and isotopic geochemistry of the A-type granites from the Huanggang Sn–Fe deposit, southern Great Hinggan Range, NE China: Implication for their origin and tectonic setting. J. Asian Earth Sci. 2012, 49, 272–286. [Google Scholar] [CrossRef]
  70. Canals, A.; Cardellach, E. Ore lead and sulphur isotope pattern from the low temperature veins of the Catalonian Coastal Ranges (NE Spain). Miner. Deposita 1997, 32, 2343–2349. [Google Scholar] [CrossRef]
  71. Li, W.B.; Huang, Z.L.; Zhang, G. Sources of the ore metals of the Huize ore field in Yunnan province constraints from Pb; S; C; H; O and Sr isotope geochemistry. Acta Petrol. Sin. 2006, 22, 2567–2580, (In Chinese with English abstract). [Google Scholar]
  72. Shi, Y.R.; Liu, D.Y.; Zhang, Q.; Jian, P.; Zhang, F.Q.; Miao, L.C.; Shi, G.H.; Zhang, L.Q.; Tao, H. SHRIMP geochronology of dioritic-granitic intrusions in Sunidzuoqi area, Inner Mongolia. Acta Geol. Sin. 2004, 78, 789–799, (In Chinese with English abstract). [Google Scholar]
  73. Cui, F.H.; Zheng, C.Q.; Xu, X.C.; Yao, W.G.; Shi, L.; Xu, J.L. Late Carboniferous magmatic activities in the Quanshenglinchang Area, GXR: Constraints on the timing of amalgamation between Xing’an and Songnen massifs. Acta Geol. Sin. 2013, 87, 1247–1263, (In Chinese with English abstract). [Google Scholar]
  74. Chen, B.; Jahn, B.M.; Wilde, S.; Xu, B. Two contrasting Paleozoic magmatic belts in northern Inner Mongolia, China: Petrogenesis and tectonic implications. Tectonophysics 2000, 328, 157–182. [Google Scholar] [CrossRef]
  75. Shi, G.H.; Miao, L.C.; Zhang, F.Q.; Jian, P.; Fan, W.M.; Liu, D.Y. The age and its district tectonic implications on the Xilinhot A-type granites, Inner Mongolia. Chin. Sci. Bull. 2004, 4, 384–389, (In Chinese with English abstract). [Google Scholar]
  76. Han, J.; Zhou, J.B.; Wang, B.; Cao, J.L. The final collision of the CAOB: Constraint from the zircon U–Pb dating of the Linxi Formation, Inner Mongolia. Geosci. Front. 2015, 6, 211–225. [Google Scholar] [CrossRef]
  77. Miao, L.C.; Fan, W.M.; Liu, D.Y.; Zhang, F.Q.; Shi, Y.R.; Guo, F. Geochronology and geochemistry of the Hegenshan ophiolitic complex: Implications for late stage tectonic evolution of the Inner Mongolia-Daxinganling Orogenic Belt. J. Asian Earth Sci. 2008, 32, 348–370. [Google Scholar] [CrossRef]
  78. Nozaka, T.; Liu, Y. Petrology of the Hegenshan ophiolite and its implication for the tectonic evolution of northern China. Earth Planet. Sci. Lett. 2002, 202, 89–104. [Google Scholar] [CrossRef]
  79. Wilde, S.A. Final amalgamation of the Central Asian Orogenic Belt in NE China: Paleo–Asian Ocean closure versus Paleo–Pacific plate subduction—A review of the evidence. Tectonophysics 2015, 662, 345–362. [Google Scholar] [CrossRef]
  80. Li, D.P.; Jin, Y.; Hou, K.J.; Chen, Y.L.; Lu, Z. Late Paleozoic final closure of the Paleo-Asian Ocean in the eastern part of the Xing-Meng Orogenic Belt: Constrains from Carboniferous-Permian (meta-) sedimentary strata and (meta-) igneous rocks. Tectonophysics 2015, 665, 251–262. [Google Scholar] [CrossRef]
  81. Pearce, J.A.; Harris, N.B.W.; Tindle, A.G. Trace element discrimination diagrams for the tectonic interpretation of granitic rocks. J. Petrol. 1984, 25, 956–983. [Google Scholar] [CrossRef]
  82. Whalen, J.B.; Hildebrand, R.S. Trace element discrimination of arc, slab failure, and A-type granitic rocks. Lithos 2019, 348, 105179. [Google Scholar] [CrossRef]
  83. Xu, L.Q.; Chen, Z.Y.; Chen, Z.H.; Zhang, T.; Zhang, Y.Q. SHRIMP dating of medium-coarse-grained granite in Chaobuleng iron deposit, Dong Ujimqin Banner, Inner Mongolia. Miner. Deposits 2010, 29, 317–320, (In Chinese with English abstract). [Google Scholar]
  84. Yang, Z.H. Metallogenic Characteristics and Metallogenic Mechanism in the Wurinitu W-Mo deposit in Inner Mongolia, China. Master’s Thesis, China University of Geosciences, Beijing, China, 2013. [Google Scholar]
  85. Zeng, Q.D.; Liu, J.M.; Qin, F.; Zhang, Z.L. Geochronology of the Xiaodonggou porphyry Mo deposit in northern margin of North China Craton. Resour. Geol. 2010, 60, 192–202. [Google Scholar] [CrossRef]
  86. Selby, D.; Nesbitt, B.E.; Muehlenbachs, K.; Prochaska, W. Hydrothermal alteration and fluid chemistry of the Endako porphyry molybdenum deposit, British Columbia. Econ. Geol. 2000, 95, 183–202. [Google Scholar] [CrossRef]
  87. Lawley, C.J.M.; Richards, J.P.; Anderson, R.G.; Creaser, R.A.; Heaman, L.M. Geochronology and geochemistry of the MAX porphyry Mo deposit and its relationship to Pb-Zn-Ag mineralization, Kootenay arc, southeastern British Columbia, Canada. Econ. Geol. 2010, 105, 1113–1142. [Google Scholar] [CrossRef]
  88. White, W.H.; Bookstrom, A.A.; Kamilli, R.J.; Ganster, M.W.; Smith, R.; Ranta, D.E.; Steininger, R.C. Character and origin of Climax-type molybdenum deposits. Econ. Geol. 1981, 75, 270–316. [Google Scholar]
  89. Ludington, S.; Plumlee, G.S. Climax-type porphyry molybdenum deposits. In US Geological Survey Open-File Report; U.S. Geological Survey: Reston, VA, USA, 2009; Volume 1215. [Google Scholar]
  90. Keith, J.D.; Christiansen, E.H.; Carten, R.B. The genesis of giant porphyry molybdenum deposits. In Giant Ore Deposits; Special Publications of the Society of Economic Geologists; Whiting, B.H., Hodgson, C.J., Mason, R., Eds.; 1993; pp. 285–317. Available online: https://pubs.geoscienceworld.org/segweb/books/book/1226/chapter-abstract/107024848/The-Genesis-of-Giant-Porphyry-Molybdenum-Deposits?redirectedFrom=fulltext (accessed on 21 May 2024).
  91. Ye, H.S.; Mao, J.W.; Li, Y.F.; Guo, B.J.; Zhang, C.Q.; Liu, J.; Yan, Q.R.; Liu, G.Y. SHRIMP Zircon U-Pb and Molybdenite Re-Os Datings of the Superlarge Donggou Porphyry Molybdenum Deposit in the East Qinling, China, and Its Geological Implications. Acta Geol. Sin. 2008, 82, 134–145. [Google Scholar]
  92. Blevin, P.L. Redox and compositional parameters for interpreting the granitoid metallogeny of eastern Australia: Implications for gold-rich ore systems. Resour. Geol. 2004, 54, 241–252. [Google Scholar] [CrossRef]
Figure 1. Simplified geologic map of the Great Xing’an region and its adjacent areas, showing the location of the Chagan Obo-Chaobulen tectono-magmatic belt (COC), which hosts the Wulandele porphyry Mo deposit (following [26]).
Figure 1. Simplified geologic map of the Great Xing’an region and its adjacent areas, showing the location of the Chagan Obo-Chaobulen tectono-magmatic belt (COC), which hosts the Wulandele porphyry Mo deposit (following [26]).
Minerals 14 00699 g001
Figure 2. Geologic map and cross section of the Wulandele porphyry Mo deposit.
Figure 2. Geologic map and cross section of the Wulandele porphyry Mo deposit.
Minerals 14 00699 g002
Figure 3. Petrographs of outcrop and textural features. (a) Quartz vein on the field; (b) quartz diorite; (c) granodiorite; (d) contact between monzonitic granite and quartz diorite; (e) vein-type mineralization; (f) disseminated mineralization.
Figure 3. Petrographs of outcrop and textural features. (a) Quartz vein on the field; (b) quartz diorite; (c) granodiorite; (d) contact between monzonitic granite and quartz diorite; (e) vein-type mineralization; (f) disseminated mineralization.
Minerals 14 00699 g003
Figure 4. Zircon U-Pb Concordia plots for quartz diorite (WL-20) and biotite monzonitic granite (WL-17) from the Wulandele porphyry Mo deposit.
Figure 4. Zircon U-Pb Concordia plots for quartz diorite (WL-20) and biotite monzonitic granite (WL-17) from the Wulandele porphyry Mo deposit.
Minerals 14 00699 g004
Figure 5. Total alkali (K2O + Na2O) vs. silica (SiO2) (a) [44] and K2O vs. SiO2 (b) [45] classification diagrams for granitoids from the Wulandele porphyry Mo deposit. The lithology symbols will be the same as Figure 5 unless it is noted.
Figure 5. Total alkali (K2O + Na2O) vs. silica (SiO2) (a) [44] and K2O vs. SiO2 (b) [45] classification diagrams for granitoids from the Wulandele porphyry Mo deposit. The lithology symbols will be the same as Figure 5 unless it is noted.
Minerals 14 00699 g005
Figure 6. Harker diagram for granitoids from the Wulandele porphyry Mo deposit; see symbols in Figure 5.
Figure 6. Harker diagram for granitoids from the Wulandele porphyry Mo deposit; see symbols in Figure 5.
Minerals 14 00699 g006
Figure 7. Chondrite-normalized REE patterns. (a) Primitive-mantle (PM)-normalized spidergrams (b) for granitoids from the Wulandele porphyry Mo deposit; see symbols in Figure 5.
Figure 7. Chondrite-normalized REE patterns. (a) Primitive-mantle (PM)-normalized spidergrams (b) for granitoids from the Wulandele porphyry Mo deposit; see symbols in Figure 5.
Minerals 14 00699 g007
Figure 8. A/NK vs. A/CNK plot (a) and Nb vs. SiO2 plot (b) for granitoids from the Wulandele porphyry Mo deposit. I = I-type granite, A = A-type granite.
Figure 8. A/NK vs. A/CNK plot (a) and Nb vs. SiO2 plot (b) for granitoids from the Wulandele porphyry Mo deposit. I = I-type granite, A = A-type granite.
Minerals 14 00699 g008
Figure 9. K2O + Na2O, Ce, and Zr vs. 10,000 Ga/Al discrimination diagrams [60], showing the A-type nature of the Cretaceous granite. I = I-type granite, S = S-type granite, A = A-type granite.
Figure 9. K2O + Na2O, Ce, and Zr vs. 10,000 Ga/Al discrimination diagrams [60], showing the A-type nature of the Cretaceous granite. I = I-type granite, S = S-type granite, A = A-type granite.
Minerals 14 00699 g009
Figure 10. Initial 207Pb/204Pb-versus-initial 206Pb/204Pb (a) and initial 208Pb/204Pb-versus-initial 206Pb/204Pb (b) for granitoids from the Wulandele porphyry Mo deposit. NHRL = North Hemisphere Reference Line. The fields for DMM, EMI, EMII, and HIMU end members are from [69].
Figure 10. Initial 207Pb/204Pb-versus-initial 206Pb/204Pb (a) and initial 208Pb/204Pb-versus-initial 206Pb/204Pb (b) for granitoids from the Wulandele porphyry Mo deposit. NHRL = North Hemisphere Reference Line. The fields for DMM, EMI, EMII, and HIMU end members are from [69].
Minerals 14 00699 g010
Figure 11. Plot of La/Yb vs. La for granitoids from the Wulandele porphyry Mo deposit. Partial melting (P.M) and fractional crystallization (F.C) trends are marked.
Figure 11. Plot of La/Yb vs. La for granitoids from the Wulandele porphyry Mo deposit. Partial melting (P.M) and fractional crystallization (F.C) trends are marked.
Minerals 14 00699 g011
Figure 12. Ta vs. Yb (a) [81] and Nb vs. Y (b) [82] discrimination diagrams for granitoids from the Wulandele porphyry Mo deposit.
Figure 12. Ta vs. Yb (a) [81] and Nb vs. Y (b) [82] discrimination diagrams for granitoids from the Wulandele porphyry Mo deposit.
Minerals 14 00699 g012
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, J.; Zhang, J.; Liu, Z.; Zhao, Y.; Zhang, F. Geochronological, Geochemical and Pb Isotope Inferences for Genesis of Wulandele Porphyry Molybdenum Deposit, Inner Mongolia, Northeast China. Minerals 2024, 14, 699. https://doi.org/10.3390/min14070699

AMA Style

Wang J, Zhang J, Liu Z, Zhao Y, Zhang F. Geochronological, Geochemical and Pb Isotope Inferences for Genesis of Wulandele Porphyry Molybdenum Deposit, Inner Mongolia, Northeast China. Minerals. 2024; 14(7):699. https://doi.org/10.3390/min14070699

Chicago/Turabian Style

Wang, Jianping, Jiexian Zhang, Zhenjiang Liu, Yun Zhao, and Fangfang Zhang. 2024. "Geochronological, Geochemical and Pb Isotope Inferences for Genesis of Wulandele Porphyry Molybdenum Deposit, Inner Mongolia, Northeast China" Minerals 14, no. 7: 699. https://doi.org/10.3390/min14070699

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