Next Article in Journal
α-Fe2O3 Nanoparticles/Iron-Containing Vermiculite Composites: Structural, Textural, Optical and Photocatalytic Properties
Next Article in Special Issue
Re Variation Triggered from the Paleo-Pacific Plate Evolution: Constrains from Mo Polymetallic Deposits in Zhejiang Province, South China Mo Province
Previous Article in Journal
Sulfate Sources Required for Thermochemical Sulfate Reduction in Dolostone Reservoirs in the Upper Permian Changxing Formation, Yuanba Gas Field, Sichuan Basin, China: Insights from the Origin of Celestite
Previous Article in Special Issue
Ranges of Physical Parameters and Geochemical Features of Mineralizing Fluids at Porphyry Deposits of Various Types of the Cu−Mo−Au System: Evidence from Fluid Inclusions Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genesis and Metallogenic Characteristics of the Zhunsujihua Granitic Intrusions in Sonid Left Banner, Inner Mongolia, China

Laboratory of Orogenic Belts and Crustal Evolution, School of Earth and Space Sciences, Peking University, Beijing 100871, China
*
Author to whom correspondence should be addressed.
Submission received: 22 March 2022 / Revised: 8 May 2022 / Accepted: 9 May 2022 / Published: 11 May 2022

Abstract

:
The Zhunsujihua porphyry Mo-Cu deposit is located in the Erenhot–East Ujimqin metallogenic belt in northeastern China. Granodioritic intrusions in the mining area are dominated by granodiorite and granodiorite porphyry, but the Mo mineralization is limited within the granodiorite. Zircon LA-ICP-MS U-Pb dating yields crystallization ages of 301.5 ± 3.0 Ma for granodiorite and 296.0 ± 3.0 Ma for granodiorite porphyry. These ages constrain the magmatic activity at the Zhunsujihua deposit that took place during the subduction of the Paleo-Asian oceanic plate. Whole-rock geochemical data suggest that the granodioritic intrusions belong to calc-alkaline and high-K calc-alkaline series, and are characterized by enrichment in K, Rb, U, Th, and Pb, and depletion in Nb, Ta, Ti, and P. The negative Eu, Ba, and Sr anomalies suggest that they have experienced extensive fractionation of plagioclase. Trace element compositions of zircons from the Zhunsujihua deposit provide constraints on the oxygen fugacity (ƒO2) of the magma, which is shown to high values with ∆FMQ = +0.5 to +5.6. The wide range of zircon εHf (t) (+1.3~+9.4) values, positive whole-rock εNd (t) (+2.5~+3.9) values, and relatively low initial (87Sr/86Sr)i (0.70367~0.70561) ratios indicate that the magmas mainly originated from a juvenile lower crust source derived from depleted mantle, but mixed with pre-existing crustal components. Moreover, the juvenile lower crust represents the main source of Mo for the Zhunsujihua deposit. A high magmatic oxygen fugacity and fractional crystallization played key roles in forming the Zhunsujihua deposit.

1. Introduction

The Xing’an–Mongolia Orogenic Belt (XMOB) is the eastern part of the Central Asian Orogenic Belt (CAOB), which lies between the Siberian and North China Cratons (Figure 1a) and is characterized by widespread arc–basin systems and continental margin accretions [1,2,3]. In the Paleozoic, the XMOB was dominated by the subduction and termination of the Paleo-Asian oceanic plate with the amalgamation of micro-continent massifs, including Erguna, Xing’an, Songliao, and Liaoyuan blocks (Figure 1b). During the Mesozoic, the XMOB was successively overprinted by the Mongol–Okhotsk tectonic domain in the west and the Circum–Pacific tectonic domain in the east [2,4]. Meanwhile, the region produced numerous magmatic hydrothermal deposits with large reserves of Mo, Cu, Au, Ag, Pb, Sn, and W. It is worth noting that the porphyry deposit represents the most important Mo and Cu deposit, contributing 95% of the Cu and 99% of the Mo reserves [5].
The Erenhot–East Ujimqin metallogenic belt in eastern CAOB stretches 184 km from east to west, and 120–160 km from north to south (Figure 1b) [9]. Over the last decade, an increasing number of middle to large-scale Mo deposits have been successively discovered in the belt, such as Wurinitu, Wulandele, Zhunsujihua, and Wuhua’aobao (Figure 1c) [3,10]. In the Paleozoic, the Erenhot–East Ujimqin district was the site of the accretion zone on the southeastern continental margin of the Siberian craton. After the closure of the Paleo-Asian Ocean, this region changed from a compression to an extensional setting, resulting in a series of magmatic activities, accompanied by the mineralization of Mo, Cu, Au, Pb, Zn, and other metals [11]. Most metallic ore deposits in this belt are porphyry-type and hosted in high-K calc-alkaline rocks, and the mineralization mainly occurred in the Yanshanian Period (Jurassic and Cretaceous) (Figure 1c), except for the Zhunsujihua porphyry Mo-Cu deposit, which is a Paleozoic deposit [11]. Therefore, the study of the Zhunsujihua deposit is of special significance to our understanding of the Late Paleozoic tectonic evolution processes and the associated metallogenesis in the eastern part of CAOB.
Previous studies have revealed that the granitoids in the Zhunsujihua deposit are high-K calc-alkaline series with I-type granite characteristics, and have reported zircon U-Pb dating ages of 299–300 Ma [12], consistent with the molybdenite Re-Os isochron age of 298.1 ± 3.6 Ma [9]. He et al. (2017) [13] investigated the nature and evolution of ore-forming fluids. However, the relationship between magmatic evolution and mineralization has not been well constrained. In this paper, we report zircon U-Pb ages and trace elements of the Zhunsujihua granitoids, along with whole-rock geochemical and Sr-Nd-Hf isotope data, to better constrain the origin and evolution of the intrusions, and their implications for the formation of the deposit.

2. Geological Setting

The Erenhot–East Ujimqin metallogenic belt is located in southwest Xing’an Terrane bounded by the Tayuan–Xiguitu fault in the north, and by the Hegenshan–Heihe fault in the south (Figure 1b) [7]. Several main NE-striking faults developed throughout this region, such as the Chagan Obo–Arongqi fault and Hegenshan–Heihe fault, providing ideal channels for magma intrusion [12]. The secondary faults are mainly NW-NWW trending, and generally filled by later dikes [12].
Outcropping strata in this region include the Ordovician Bayanhushu Formation of marine sedimentary rocks, Carboniferous–Permian Baoligaomiao Formation of volcano-sedimentary rocks, Upper Jurassic volcanic rocks, Tertiary sandstone, and Quaternary sediments (Figure 1c). Phanerozoic granitic intrusions were mainly emplaced during the Variscan and Yanshanian (Figure 1c). The Variscan intrusions are mainly composed of diorite, granodiorite, diabase, diorite porphyry, and granite porphyry; the Yanshanian intrusions are dominated by granite, diorite porphyry, granite porphyry, quartz monzonite porphyry, and aplite.

3. Deposit Geology

The study area is situated in the northwest of Sonid Left Banner, Xilin Gol League, Inner Mongolia Autonomous Region. Its geographic coordinates are from 112°41′00″ E to 112°45′00″ E, and from 44°30′00″ N to 44°32′00″ N, with a mining area of 16.02 km2. The total explored resource of the Zhunsujihua deposit is 15.24 Mt Mo and 0.43 Mt Cu with an average grade of 0.127% Mo and 0.793% Cu (reference from the internal data of the Ninth Exploration Institute of Inner Mongolia). The stratigraphic sequences exposed in the Zhunsujihua porphyry Mo-Cu deposit are mainly the Carboniferous–Permian Baoligaomiao Formation of metamorphic siltstone and Ordovician Bayanhushu Formation of metamorphic clastic rock, partially covered by Quaternary sediments (Figure 2) [9]. The strata generally occur in a NE trending monoclinic structure controlled by the regional main fractures; dike rocks and orebodies are mainly controlled by the NW-trending faults (Figure 2) [9]. The Zhunsujihua granite intrusive complex consists of a large body of granodiorite and some granodiorite porphyry dikes and leucogranite dikes, which intruded into the Baoligaomiao Formation strata (Figure 2 and Figure 3) [14]. In addition, a barren Yanshanian granodioritic stock outcrops in the study area (Figure 2).
The granodiorite is medium-grained and equigranular in texture (Figure 4a,d). Plagioclase is the predominant mineral and occupies approximately 40–45% of the rock. Other major minerals include K-feldspar (15–20%), quartz (20–25%), amphibole (5%), and biotite (10%), with accessory minerals of zircon, titanite, apatite, and magnetite (Figure 4d). The granodiorite porphyry has a typical porphyritic structure (Figure 4b,e). The phenocrysts of granodiorite porphyry are plagioclase (~45%), quartz (~50%), and biotite (~5%), while the groundmass comprises quartz, plagioclase, K-feldspar, and biotite (Figure 4e). The granodiorite shows a direct genetic linkage to the Mo mineralization (Figure 4c). In the contact zone between granodiorite and wall rocks, hornification and skarnization are locally developed. No obvious mineralization was observed in the granodiorite porphyry.
Most of the orebodies of the Zhunsujihua porphyry Mo-Cu deposit occurred in a lenticular shape, with local pinching out in both the strike and dip (Figure 3). The length is generally about 300 m, with a few reaching up to 800 m; the extension depth varies from 10 to 400 m; the thickness of the ore body is generally between 1 and 10 m, with an average of 3 m. Alteration in the mining area shows characteristics typical of those associated with porphyry deposits, including potassic alteration, silicification, sericitization, and carbonation, which are distributed in the interior of the granodiorite body or in the contact zone between granodiorite and country rocks. Weak sericitization and epidotization are observed in the granodiorite porphyry body (Figure 4e). Sericitization and silicification are very closely related to the molybdenum mineralization and are the main precipitation stages of molybdenite (Figure 4f). The ore occurs as dominant sulfide-bearing quartz veins (Figure 4c). Pyrite, molybdenite, and chalcopyrite constitute the main metal minerals. In addition, there are also small amounts of pyrrhotite, azurite, sphalerite, magnetite, etc. Gangue minerals are dominated by quartz and carbonate, with a small amount of sericite, biotite, chlorite, epidote, anhydrite, etc. In the early stage, molybdenite usually occurs as fine scales or stockwork veinlets on the outer edges of quartz veins, and a little occurs inside the veins. In the late stage, molybdenite veins often coexist with chalcopyrite and pyrite veins, which show disseminated or massive occurrences and crosscut the early molybdenite veins [13].

4. Sampling and Analytical Methods

The thirteen least altered samples selected for analyses were all from drilling cores or ore heaps in the Zhunsujihua mining area. Four samples (two of granodiorite and two of granodiorite porphyry) were used for zircon U-Pb dating, in situ Hf isotope, and trace element determination. Whole-rock and trace elements analyses were examined on eight samples (including four granodiorite porphyry and four granodiorite). Three granodiorite porphyry samples and three granodiorite samples were selected for Sr-Nd isotope analyses.

4.1. In Situ Zircon U-Pb and Trace Element Analyses

Cathodoluminescence (CL) images of zircons were taken at the SEM-CL-TIMA Laboratory in Peking University, China. U-Pb isotopes and trace element analyses were conducted simultaneously using LA-ICP-MS at the Key Laboratory of Orogenic Belt and Crustal Evolution, Peking University, China. An Agilent 7500 ce/cs ICP-MS coupled with a 193 nm ArF laser-ablation system (ComPex 102) was used. The analyses were carried out using a 32 μm diameter spot with a flux of 10 J/cm2 at rate of 5 Hz. The ablated aerosol was transported into the mass spectrometer by high-purity helium at a rate of 0.65 L/min. The NIST610 and zircon 91,500 were employed as the primary reference materials for instrument optimization and elemental fractionation calibration. The Plešovice zircon was used for data quality control; sixteen zircon spots yielded a Concordia age of 337.6 ± 1.7 Ma (95% confidence, MSWD = 0.18), which agrees well with the recommended value of 337.13 ± 0.37 Ma [15,16]. The data reduction was performed using the Glitter 4.4.2 software and the weighted mean 206Pb/238U was calculated using the Isoplot/Ex rev. 2.49 Program [17].

4.2. Whole-Rock Geochemistry

Major and trace elements were analyzed at the Analytical Chemistry & Testing Services (ALS) Laboratory, Guangzhou, China. Samples were first cleaned and then crushed in a steel jaw crusher and powered to 200 mesh in an agate mill. The major elements analyses were performed using an X-ray fluorescence spectrometer (XRF) (PANalytical PW2424). Before XRF analysis, the mixtures of 0.4 g of rock powder, 4 g of Li2B4O7 (flux), and 4 or 5 drops of NH4Br (release agent) were fused in a furnace at 1100 °C to form a glass disk. The loss on ignition (LOI) was determined by placing a 1 g sample in a crucible at 980 °C for 20 min before cooling in the desiccator and reweighing. The standards NCSDC73303 and SARM-4 were used to control the analytical precisions. The analytical precisions were generally better than 0.5%. For detailed experimental procedures, refer to [18].
For whole-rock trace elements analyses, 25 mg of crushed sample powder was dissolved in HNO3 (1.5 mL) and HF (1.5 mL) in a closed Teflon bomb at 150 °C for 24 h. Subsequently, 10 mL of HClO4, 1.5 mL of HF, and 1.5 mL of HNO3 were added to the Teflon bomb and heated in an oven at 150 °C for 10 h to dissolve the sample completely. The cooled solution was placed on a hot plate and evaporated to dryness. After the addition of 1 mL of HNO3, the bomb was placed on a hot plate until the sample dissolved. Then, the final solution was fully transferred to a 50 mL colorimetric tube and the volume was metered with 1% HNO3. The final solution was subsequently analyzed by ICP-MS (Agilent 7900, USA). The procedures are described in detail by [19]. The MRGeo08 and OREAS-100a standards were employed to monitor the analytical accuracy. The precisions of the ICP-MS analyses were generally better than 5%.

4.3. Whole-Rock Sr-Nd Isotope Analyses

The separation and purification of Rb, Sr, and light rare earth elements (LREEs) were performed in the ultraclean room of the Laboratory of Orogenic Belt and Crust Evolution, Peking University, China. The whole-rock powder samples were dissolved using acids (HClO4 + HNO3 + HF) in Teflon capsules, and separated by the conventional cation-exchange technique. Isotopic ratios were measured in Shangpu Analysis Technology Co., Ltd. (Wuhan, China) by thermal ionization mass spectrometry (TIMS) using a Triton instrument (ThermoFisher, Waltham, MA, USA), operating in the positive ionization mode with a 10 kV acceleration voltage and 1011 Ω for the Faraday cups. Rb, Sr, Sm, and Nd concentrations were measured by the isotopic dilution method. All water used in the experiments was purified by a high-purity water machine (Millipore Element, Millipore Corporation, Burlington, MA, USA) to a resistivity of 18.2 MΩ·cm−1. Procedural blanks were 200 pg for Rb and Sr, and 50 pg for Sm and Nd. The mass fractionation correction for 87Sr/86Sr and 143Nd/144Nd ratios was normalized against 86Sr/88Sr = 0.1194 and 146Nd/144Nd = 0.7219, respectively. 87Sr/86Sr ratios were adjusted relative to the NBS-987 Sr standard = 0.710256 ± 0.000016 (2σ; n = 5), and 143Nd/144Nd ratios were adjusted to the JNdi-1 Nd standard = 0.512113 ± 0.000012 (2σ; n = 5). The reproducibility of 87Sr/86Sr and 143Nd/144Nd from JR-2 and BIR-1 is approximately 0.002–0.003% (2σ). The detailed analytical procedures for Sr and Nd isotopic measurements are outlined in [20,21].

4.4. In Situ Hf Isotope Analyses of Zircon

In situ zircon Lu-Hf isotope analyses were obtained on the same zircon grains that were previously analyzed for U-Pb isotopes, and performed on a Neptune Plus laser-ablation multi-collector inductively coupled plasma mass spectrometer (LA-MC-ICP-MS) equipped with a GeolasPro HD laser ablation system. A spot size of 44 μm, an ablation time of 26 s, a repetition rate of 8 Hz, and a pulse energy of 8 mJ/cm2 were used for analyses. The isobaric interference of 176Lu and 176Yb on 176Hf was corrected by values of 176Lu/175Lu (0. 02669) and 176Yb/172Yb (0. 5886). The 91,500 zircon was used as the internal standard with recommended 176Hf/177Hf ratios (0.282307) [22]. GJ-1 standard was analyzed to monitor the data quality, which yielded weighted mean 176Hf/177Hf ratios of 0.282010 ± 0.000024 (2σ, n = 4). This value agrees well with the recommended value of 0.282015 ± 0.000019 (2σ) [23]. εHf (t) was calculated based on the decay constant of 176Lu of 1.867 × 10−11 a, and the 176Hf/177Hf and 176Lu/177Hf ratios of chondrite of 0.282785 and 0.0336, respectively. One-stage model ages (TDM1) of the Hf isotope were calculated using the current depleted mantle value of 176Hf/177Hf = 0.28325, and 176Lu/177Hf = 0.0384 [24]. Two-stage model ages (TDM2) were calculated using the average crust value of 176Lu/177Hf = 0.015 [24]. The detailed analytical procedures are described in [22]. All analyses were carried out in Shangpu Analysis Technology Limited Liability Company, Wuhan, China.

5. Results

5.1. Zircon U-Pb Dating and Trace Element Results

Zircon CL images are shown in Figure 5 and the U-Pb dating results are summarized in Table 1. Most zircons have euhedral crystals that exhibit typical magmatic oscillatory zoning without distinctive older cores or younger overgrowths (Figure 5). Their size varies between 40 and 200 μm in length, and these crystals have variable U (162.4–1086.9 ppm) and Th (74.6–591.6 ppm) contents with a Th/U between 0.12 and 1.51. All these features indicate that the zircons are of magmatic origin [25,26]. The data for sixteen measuring points from granodiorite porphyry are closely distributed, constituting a relevant age group, with a calculated 206Pb/238U weighted mean age of 296.0 ± 3.0 Ma (95% confidence, MSWD = 3.0) (Figure 6). Sixteen zircon spots analyses from the granodiorite plotted on concordia yield a well-sorted population with a mean 206Pb/238U age of 301.5 ± 3.0 Ma (95% confidence, MSWD = 2.7) (Figure 6). Data for the trace elements in zircon are listed in Table 2. The chondrite-normalized REE diagram for zircon shows a strong left-leaning trend, with an obvious positive Ce anomaly and negative Eu anomaly (Figure 7).

5.2. Major and Trace Element Compositions

The results of major and trace elements for granodiorite and granodiorite porphyry are summarized in Table 3. The petrographic photos of representative rocks show that the studied samples are virtually unaltered (Figure 4), which is consistent with their low LOI values, ranging from 0.92 to 1.45 wt.% (Table 3) [28]. Granodiorite and granodiorite porphyry have similar geochemical characteristics in terms of both major and trace elements. All the samples are characterized by high silicon contents (SiO2: 69–71 wt.%) and are alkali-rich (K2O + Na2O: 7.02–7.76 wt.%) with relatively high K2O/Na2O ratios (0.62–0.99), and have low TiO2, MgO, and P2O5 contents (Table 3). They plot into the granodiorite and granite fields in the SiO2 vs. (Na2O + K2O) diagram (Figure 8), and fall into the calc-alkaline and high K calc-alkaline series in the SiO2 vs. K2O diagram (Figure 9). The majority of calculated A/CNK values are over 1.0, with only one exception of 0.99 (Table 3), indicating that these rocks are peraluminous or only slightly metaluminous (Figure 10). In the chondrite-normalized rare earth element (REE) diagram (Figure 11), they are enriched in LREEs at different degrees, relatively depleted in heavy rare earth elements (HREEs), and have a significant fractionation between LREEs and HREEs. The (La/Yb)N values of granodiorite and granodiorite porphyry are 10.44–11.93 and 9.37–10.84, respectively. The granodiorite shows an obvious negative Eu anomaly (δEu = 0.51–0.58), while the granodiorite porphyry shows a weak negative Eu anomaly (δEu = 0.65–0.68). In the primitive mantle-normalized trace element spider diagram, both granodiorite and granodiorite porphyry show enrichment in certain large-ion lithophile elements (LILEs) (K, Rb, U, Th, Pb), but depletion in Ba and high-field-strength elements (HFSEs) (Nb, Ta, Ti, P) (Figure 11).

5.3. Whole-Rock Sr-Nd Isotopic Compositions and Zircon Hf Isotopes

The results of the whole-rock Sr-Nd isotopic compositions and in situ zircon Lu-Hf isotopic analyses are presented in Table 4 and Table 5, respectively. The initial 87Sr/86Sr ratios and εNd (t) values are calculated at 302 Ma for granodiorite and at 296 Ma for granodiorite porphyry. The granodiorite has (87Sr/86Sr)i = 0.70412–0.70561, and εNd (t) = +2.55 to +3.39, yielding two-stage depleted mantle Nd model ages (TDM2) of 787 to 807 Ma. The granodiorite porphyry has (87Sr/86Sr)i = 0.70367–0.70378, and εNd (t) = +3.08 to +3.86, yielding TDM2 of 754 to 787 Ma. The (87Sr/86Sr)i values of granodiorite are slightly higher than those of granodiorite porphyry, but the εNd (t) values are slightly lower.
The 176Yb/177Hf and 176Lu/177Hf ratios for zircons from granodiorite are relatively low, ranging from 0.045504 to 0.075724 and from 0.001378 to 0.002395, respectively. The calculated εHf (t) values vary from +1.3 to + 8.9 and the obtained two-stage depleted mantle Hf model ages (TDM2) are 693–1118 Ma. The 176Yb/177Hf and 176Lu/177Hf ratios for zircons from granodiorite porphyry are relatively high, varying in a large range of 0.049679–0.137264 and 0.001301–0.004089, respectively. The calculated εHf (t) values vary from +1.9 to +9.4 and the obtained TDM2 are 666–1151 Ma (Table 5). Both two-stage model ages obtained from Hf isotopes are in good agreement with those obtained from Nd isotopes. The range of εHf (t) variation is significantly larger than the error caused by the analytical error. Therefore, zircons from the Zhunsujihua granitic intrusions have heterogeneous Hf isotopic compositions.

6. Discussion

6.1. Timing of Magmatism and Mineralization

The zircon U-Pb ages of the granodiorite and granodiorite porphyry in the Zhunsujihua deposit are 301.5 ± 3.0 Ma and 296.0 ± 3.0 Ma, respectively. These age data are similar to the previously reported U-Pb age of granodiorite (300 ± 2 Ma) and granodiorite porphyry (299.0 ± 2.6 Ma) [12], and are consistent with the reported molybdenite Re-Os isochron age of 297.2 ± 4.3 Ma [31], suggesting that the Zhunsujihua granodioritic intrusions and mineralization are related to Late Carboniferous–Early Permian magmatic–hydrothermal events [31]. Given that the granodiorite is the major host to the orebodies, and no mineralization and alteration were observed in the granodiorite porphyry, it is reasonable to propose that granodiorite was the causative intrusion.

6.2. Petrogenesis of the Magmatic Rocks

All of the granodiorite and granodiorite porphyry samples contain magmatic biotite and hornblende, suggesting that they have a typical I-type affinity [32]. This is further supported by the Sr isotope data (ISr = 0.70367–0.70561) and the geochemical features of the granodioritic intrusions, which are metaluminous, high-K calc-alkaline, with A/CNK < 1.1.
Similar Nd and Hf isotope compositions of granodiorite and granodiorite porphyry, together with the same zircon U-Pb ages, indicate that they have a unified magma source region [33]. The positive whole-rock εNd (t) values (+2.55~+3.86) and relatively low (87Sr/86Sr)i values (0.70367~0.70561) suggest that the source region might be from the simple juvenile lower crust derived from the depleted mantle and mixed with pre-existing crustal materials. According to the mixing calculations using different end-members, which are shown in the εNd (t)-(87Sr/86Sr)i diagram (Figure 12a) [34], the source of Zhunsujihua intrusions contains 80–90% juvenile basalt components, with 10–20% ancient crustal materials. The average Ce/Pb and Nb/Ta ratios of granodiorite and granodiorite porphyry are 5.3 and 11.1, respectively, which are close to those of the crust (Ce/Pb = 3.9 [35]; Nb/Ta = 12–13 [36]), but quite different from those of the mantle-derived magma (Ce/Pb = 27 [37]; Nb/Ta = 17.5 ± 2.0 [38]). Therefore, the magma source region should be dominated by the juvenile lower crustal materials.
The zircon Lu-Hf isotopic system has a higher closure temperature compared with the whole-rock Rb-Sr or Sm-Nd isotopic system [40], so it can be applied to track the source region and identify the magma mingling process more effectively [41]. In the εHf (t)-Age diagram (Figure 12b), all samples plot into the field between the chondrite uniform reservoir reference line and depleted mantle evolution line. The two-stage Hf model ages from 564 to 1150 Ma suggest that the crust-mantle differentiation could be dated back to the Mesoproterozoic. The εHf (t) values of zircons from the Zhunsujihua granodiorite and granodiorite porphyry vary widely (+1.3~+9.4) (Table 5) and even show negative values [12], indicating that the source magma was mainly derived from the juvenile crust, but incorporated ancient crustal components [42]. Mao et al. (1999) [43] concluded that the rhenium contents in molybdenite decrease when going from a mantle source (≈10−4), to mixtures between mantle and crust (≈10−5), and then to crust (≈10−6). The rhenium content of molybdenite in the Zhunsujihua deposit ranges from 9 × 10−6 to 2.8 × 10−5 [31], which also indicates that the ore materials originated from a mixed source between the crust and mantle. Finally, we can conclude that the Zhunsujihua granitic intrusions most likely originated from partial melting of the juvenile lower crust derived from the depleted mantle, but mixed with pre-existing crustal components.

6.3. Tectonic Implications

Both granodiorite and granodiorite porphyry are characterized by enrichments in LILEs and LREEs, and a relative depletion in HFSEs, similar to those of arc volcanic rocks [44]. In the Sr/Y vs. Y and Rb vs. (Y + Nb) diagrams, all rock samples fall into the island arc or the continental margin arc field (Figure 13) [45,46]. It seems that Zhunsujihua granodioritic intrusions were formed in the arc tectonic setting. This is supported by other evidence. From the late Carboniferous to the early Permian, there was a successive volcanic-sedimentary sequence in the region, i.e., the Baoligaomiao Formation [47]. The volcanic rocks of the Baoligaomiao Formation are primarily rhyolite, dacite, and andesite [48], which show analogous geochemical characteristics to the Zhunsujihua granitic intrusions and igneous rocks in an active continental margin setting [6,47,49]. Furthermore, Chai et al. (2018) [50] reported an arc-like adakitic granodiorite in the Baiyinwula, which yielded a zircon U-Pb age of 318 ± 1.8 Ma; Chen et al. (2000) [51] proposed that Baolidao rocks in Sonidzuoqi was emplaced at ca. 310 Ma under the subduction setting.

6.4. Oxygen Fugacity of Magmas

The oxygen fugacity of the magma is an important parameter of the redox state during fractional crystallization, which affects the mineralization of many metals, especially chalcophile and siderophile elements, such as Cu and Au [52]. Under relatively low oxygen fugacity conditions, the sulfur in magma mainly exists in the form of S2−. The sulphophile elements such as Cu preferentially combine with S2− to form sulfide and remain in the magma chamber. Under high-oxygen fugacity conditions, S2− is easily oxidized to SO42− or SO2 and dissolved in the silicate melt. A higher concentration of Cu and other sulphophile elements can be further transported into magmatic hydrothermal fluids and enriched for mineralization. In this study, we calculated the C e 4 + / C e 3 + ratio of zircons to represent the relative oxygen futility of the magma during fractional crystallization based on the following formula [53]:
[ C e 4 + / C e 3 + ] z i r c o n = ( C e m e l t Ce z i r c o n / D C e 3 + z i r c o n / m e l t ) / ( Ce z i r c o n / D C e 4 + z i r c o n / m e l t C e m e l t )
where D C e 3 + z i r c o n / m e l t and D C e 4 + z i r c o n / m e l t are calculated from the linear fit of trivalent cations (REE3+) and tetravalent cations, including U4+, Th4+, and Hf4+, respectively, and C e z i r c o n and C e m e l t are the Ce contents in zircon and whole rock, respectively. The detailed calculation process is described in [53]. Trail et al. (2012) [54] converted the relative oxygen fugacity to the absolute oxygen fugacity (logƒ(O2)) by calibrating the Ce anomaly of zircon as a function of temperature and oxygen fugacity:
ln (δCe)D = (0.1156 ± 0.0050) × ln (ƒ(O2)) + (13860 ± 708)/T(K) − 6.125 ± 0.484
where the values of (δCe)D can be replaced by [ C e 4 + / C e 3 + ] z i r c o n , and T is the zircon crystallization temperature in K, which can be obtained using the Ti thermometer of zircon [55]:
T(K) = (5080 ± 30)/[(6.01 ± 0.03) − lg (Ti)]
The oxygen fugacity data of granodiorite and granodiorite porphyry calculated with the above empirical formula are listed in Table 2 and plotted in the logƒ(O2)-1000/T diagram. As shown in Figure 14, most of the oxygen fugacity (logƒ(O2)) data for granodiorite porphyry and granodiorite are above the FMQ buffer line, with an average value of ∆FMQ + 1.1 and ∆FMQ + 3.5, respectively (Table 2). It is notable that the clean zircon is generally considered as a La, Pr-free mineral with La, Pr ≤ 0.1 ppm [56]. Those zircons with La or Pr > 0.1 ppm may contain undetected inclusions, which would result in unreliable oxygen fugacity calculations based on the REE composition of zircons [57]. In other words, the oxygen fugacity values (∆FMQ = +0.5 to +5.6; Table 2) calculated using REE contents of clean zircons are robust (Figure 14), and it can be concluded that both granodioritic intrusions had a high oxygen fugacity, which was favorable for the migration of ore-forming elements.

6.5. Controls on the Formation of the Zhusujihua Mo-Cu Deposit

A previous study has revealed that source regions [58], the magmatic process (e.g., fractionation) [59], and high-oxygen fugacities [60] were all critical for the formation of porphyry Mo deposits. Although many scholars have agreed that the Mo is mainly derived from the ancient lower crust [61,62], our study suggests that partial melting of the juvenile lower crust played an important role in the genesis of granodioritic magma related to the Mo mineralization in the Zhunsujihua deposit, which is similar to other porphyry deposits in the XMOB [63] (e.g., the Daheishan Mo deposit [64]; the Badaguan Cu-Mo deposit [65], the Luming Mo deposit [66]). In addition, our data indicate a high oxygen fugacity for the granodioritic magma in Zhunsujihua. Such oxidized magmas are common in subduction zones [67]. A high magmatic oxygen fugacity can enable Mo to migrate in the form of Mo6+ in the residual melt until fluid exsolution, rather than in the form of Mo4+ partitioning into Ti-bearing magmatic minerals or magmatic sulfides [68]. Furthermore, the negative Eu, Ba, and Sr anomalies indicate that the parental magma that formed granodiorite and granodiorite porphyry experienced extensive fractionation of plagioclase or K-feldspar. The differentiation index (DI) of magma was used to quantify the degree of fractional crystallization. The results show that both the granodiorite and granodiorite porphyry have high DI values of 92.3–93.9 and 92.7–94.3, respectively. These values are similar to those (93–97) of the Chalukou porphyry Mo deposit [69]. Li et al. (2014) [69] argued that the felsic magma with high DI values could maintain long-term fractional crystallization and ultimately contribute to Mo enrichment in the residual magma.
In summary, the Mo (probably also Cu) of the Zhunsujihua deposit was mainly sourced from the juvenile lower crust. A high magmatic oxygen fugacity and fractional crystallization played key roles in forming the Zhunsujihua deposit.

7. Conclusions

The granodioritic intrusions exposed in the Zhunsujihua porphyry Mo-Cu deposit mainly consisted of granodiorite and granodiorite porphyry, and both were emplaced at ca. 296.0–301.5 Ma in a continental margin arc setting.
Granodiorite and granodiorite porphyry have a unified magma source region, both of which were derived from partial melting of the juvenile lower crust, but mixed with ancient crustal components.
The Mo in the Zhunsujihua deposit was mainly sourced from the juvenile lower crust. A high magmatic oxygen fugacity and preliminary enrichment of Mo through fractional crystallization were critically important in the formation of the Zhunsujihua deposit.

Author Contributions

Conceptualization, Q.S. and Y.L.; methodology, Y.L.; software, Q.S.; validation, Q.S., Y.L. and H.G.; formal analysis, C.L.; investigation, H.G.; resources, H.G.; data curation, C.L.; writing—original draft preparation, Q.S.; writing—review and editing, Q.S.; visualization, Q.S.; supervision, Y.L.; project administration, Y.L.; funding acquisition, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key R&D Program, China [grant number 2021YFA0719002].

Data Availability Statement

All data appear in the submitted article.

Acknowledgments

The authors are thankful to Pingping Liu, Fanghua Zhang and Fa Ma for valuable discussion and help with zircon data processing, respectively. Wasiq Lutfi and Huijuan Li are thanked for language polishing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, Y.J.; Zhang, C.; Li, N.; Yang, Y.F.; Deng, K. Geology of molybdenum deposits in Northeast China. J. Jilin Univ. Sci. Ed. 2012, 42, 1223–1268, (In Chinese with English Abstract). [Google Scholar]
  2. Lv, B.; Wang, T.; Tong, Y. Spatial and Temporal Distribution and Tectonic Settings of Magmatic-Hydrothermal Ore Deposits in the Eastern Central Asian Orogenic Belt. J. Jilin Univ. Sci. Ed. 2017, 47, 305–343, (In Chinese with English Abstract). [Google Scholar]
  3. 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 Daxing’anling mineralization belt, China. Geol. Bull. China 2009, 28, 955–962, (In Chinese with English Abstract). [Google Scholar]
  4. Zeng, Q.D.; Qin, K.Z.; Liu, J.M.; Li, G.M.; Zhai, M.G.; Chu, S.X. Porphyry molybdenum deposits in the Tianshan-Xingmeng orogenic belt, northern China. Int. J. Earth Sci. 2015, 104, 991–1023. [Google Scholar] [CrossRef]
  5. Zeng, Q.D.; Liu, J.M.; Chu, S.X.; Fu, G.L.; Yu, W.B.; Li, Z.M.; Gao, Y.Y.; Li, Y.J.; Sun, Y.; Zhou, L.L.; et al. Mesozoic granitic magmatism and molybdenum ore–forming process in the Xilamulun metallogenic belt. J. Jilin Univ. (Earth Sci. Ed.) 2011, 46, 1705–1714, (In Chinese with English Abstract). [Google Scholar]
  6. 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, 8–20. [Google Scholar] [CrossRef] [Green Version]
  7. Wu, F.Y.; Sun, D.Y.; Ge, W.C.; Zhang, Y.B.; Grant, M.L.; Wilde, S.A. Geochronology of the Phanerozoic granitoids in northeastern China. J. Asian Earth Sci. 2011, 41, 1–30. [Google Scholar] [CrossRef] [Green Version]
  8. Yang, Z.H.; Wang, J.P.; Liu, J.J.; Wang, S.G.; Wang, Q.Y.; Kang, S.G.; Zhang, J.X.; Zhao, Y. Isotope geochemistry of the Wurinitu W-Mo deposit in Sunid Zuoqi, Inner Mongolia, China. Geoscience 2013, 27, 13–23, (In Chinese with English Abstract). [Google Scholar]
  9. Liu, Y.F.; Nie, F.J.; Jiang, S.H.; Hou, W.R.; Liang, Q.L.; Zhang, K. Geochronology of Zhunsujihua molybdenum deposit in Sonid Left Banner, Inner Mongolia, and its geological significance. Miner. Depos. 2012, 31, 119–128, (In Chinese with English Abstract). [Google Scholar]
  10. Huang, D.H.; Du, A.D.; Wu, C.Y.; Liu, L.S.; Sun, Y.L.; Zou, X.Q. Metallochronology of molybdenum (copper) deposits in the north China platform: Re-Os age of molybdenite and its geological significance. Miner. Depos. 1996, 15, 289–297, (In Chinese with English Abstract). [Google Scholar]
  11. Yu, Q.A.; Sima, X.Z.; Qu, K. Prospecting progress of copper and molybdenum polymetallic metallogenic belt in the middle and western of Erenhot-East Ujimqin, Inner Mongolia. Geol. Rev. 2016, 62, 139–140, (In Chinese with English Abstract). [Google Scholar]
  12. Zhang, X.J.; Lentz, D.R.; Yao, C.L.; Liu, R.; Yang, Z.; Mei, Y.X. Geochronology, geochemistry, and Sr-Nd-Pb-Hf isotopes of the Zhunsujihua granitoid intrusions associated with the molybdenum deposit, northern Inner Mongolia, China: Implications for petrogenesis and tectonic setting. Int. J. Earth Sci. 2018, 107, 687–710. [Google Scholar] [CrossRef]
  13. He, L.Y.; Wang, Y.H.; Liu, J.J.; He, M.Y.; Yao, Y.; Xian, X.C. Genesis of the Zhunsujihua porphyry Mo deposit in the Erenhot-East Ujimqin metallogenic belt, Inner Mongolia, China: Evidence from geology, fluid inclusions, and isotope. Geol. J. 2017, 55, 223–238. [Google Scholar] [CrossRef]
  14. Lv, J.K.; Zhang, J.X.; Feng, R.; Zhao, J.; Wu, Y.A. The Census Report for the Zhunsujihua Mo-Cu Deposit, Sonid Left Banner, Inner Mongolia; The 9th Geology and Mineral Resources Exploration Department of Inner Mongolia Autonomous Region: Inner Mongolia, China, 2008; pp. 1–48, (In Chinese with English Abstract).
  15. Sláma, J.; Košler, J.; Condon, D.J.; Crowley, J.L.; Gerdes, A.; Hanchar, J.M.; Norberg, N. Plešovice zircon-a new natural reference material for U-Pb and Hf isotopic microanalysis. Chem. Geol. 2008, 249, 1–35. [Google Scholar] [CrossRef]
  16. Wiedenbeck, M. Three natural zircon standards for U-Th-Pb, Lu-Hf, trace element and REE analyses. Geostand. Geoanal. Res. 1995, 19, 1–23. [Google Scholar] [CrossRef]
  17. Andersen, T. Correction of common lead in U-Pb analyses that do not report 204Pb. Chem. Geol. 2002, 192, 59–79. [Google Scholar] [CrossRef]
  18. Liu, S.W.; Pan, Y.M.; Xie, Q.L.; Zhang, J.; Li, Q.G. Archean geodynamics in the Central Zone, North China Craton: Constraints from geochemistry of two contrasting series of granitoids in the Fuping and Wutai complexes. Precambrian Res. 2004, 130, 229–249. [Google Scholar] [CrossRef]
  19. Han, Y.G.; Zhang, S.H.; Pirajno, F.; Zhang, Y.H. Evolution of the Mesozoic granites in the Xiong’ershan-Waifangshan region, western Henan province, China, and its tectonic implications. J. Geol. Engl. Ed. 2007, 81, 253–265. [Google Scholar]
  20. Thirlwall, M.F. Long-term reproducibility of multicollector Sr and Nd isotope ratio analysis. Chem. Geol. 1991, 94, 85–104. [Google Scholar] [CrossRef]
  21. Li, C.F.; Li, X.H.; Li, Q.L.; Guo, J.H.; Yang, Y.H. Rapid and precise determination of Sr and Nd isotopic ratios in geological samples from the same filament loading by thermal ionization mass spectrometry employing a single-step separation scheme. Anal. Chem. Acta 2012, 727, 54–60. [Google Scholar] [CrossRef]
  22. Yuan, H.L.; Gao, S.; Dai, M.N.; Zong, C.L.; Detlef, G.; Fontaine, G.H. Simultaneous determinations of U-Pb age, Hf isotopes and trace element compositions of zircon by excimer laser-ablation quadrupole and multiple-collector ICP-MS. Chem. Geol. 2008, 247, 100–118. [Google Scholar] [CrossRef]
  23. Elhlou, S.; Belousova, E.; Griffin, W.L.; Pearson, N.J.; O’reilly, S.Y. Trace element and isotopic composition of GJ red zircon standard by laser ablation. Geochim. Cosmochim. Acta 2006, 70, A158. [Google Scholar] [CrossRef]
  24. Griffin, W.L.; Pearson, N.J.; Belousova, E.; Jackson, S.E.; van Achterbergh, E.; O’Reilly, S.Y.; Shee, S.R. The Hf isotope composition of -msonic mantle: LAM-MC-ICPMS analysis of zircon megacrysts in kimberlites. Geochim. Cosmochim. Acta 2000, 64, 133–147. [Google Scholar] [CrossRef]
  25. Hanchar, J.M.; Hoskin, P.W.O. Zircon, 1st ed.; Gruyter: Berlin, Germany, 2003; pp. 1–50. [Google Scholar]
  26. Pidgeon, R.T.; Nemchin, A.A.; Hitchen, G.J. Internal structures of zircons from Archaean granites from the darling range batholith: Implications for zircon stability and the interpretation of zircon U-Pb ages. Contrib. Mineral. Petrol. 1998, 132, 288–299. [Google Scholar] [CrossRef]
  27. Sun, S.S.; McDonough, W.F. Chemical and isotopic systematics of oceanic basalts: Implications for mantle composition and processes. Geol. Soc. 1989, 42, 313–345. [Google Scholar] [CrossRef]
  28. Lu, Y.J.; Kerrich, R.; Kemp, A.I.; McCuaig, T.C.; Hou, Z.Q.; Hart, C.J.; Yang, Z.M. Intracontinental Eocene-Oligocene porphyry Cu mineral systems of Yunnan, western Yangtze Craton, China: Compositional characteristics, sources, and implications for continental collision metallogeny. Econ. Geol. 2013, 108, 1541–1576. [Google Scholar] [CrossRef]
  29. Middlemost, E.A.K. Naming materials in the magma/igneous rock system. Earth-Sci. Rev. 1994, 37, 215–224. [Google Scholar] [CrossRef]
  30. Jacobsen, S.B.; Wasserburg, G.J. Sm-Nd isotopic systematics of Chondrites and Achondrites. Meteoritics 1980, 15, 307. [Google Scholar]
  31. Liu, C.; Guo, H.; Lai, Y. Study on grannitoid intrusions characteristics and metallogenetic mechanism of Zhunsujihua porphyry molybdenum deposit. J. Peking Univ. Sci. Ed. 2020, 56, 679–691, (In Chinese with English Abstract). [Google Scholar]
  32. Chappell, B.W.; White, A.J.R. I- and S-type granites in the Lachlan fold belt. Trans. R. Soc. Edinb. Earth Sci. 1992, 83, 1–26. [Google Scholar]
  33. Zhu, D.C.; Pan, G.T.; Mo, X.X.; Duan, L.P.; Liao, Z.L.; Wang, L.Q. Sr-Nd-Pb isotopic variations of the Cenozoic volcanic rocks from the Qinghai-Xizang plateau and its adjacent areas. Sediment. Geol. Tethyan Geol. 2003, 23, 1–11, (In Chinese with English Abstract). [Google Scholar]
  34. 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]
  35. Rudnick, R.L.; Gao, S. Composition of the continental crust. In Treatise on Geochemistry; Holland, H.D., Turekian, K.K., Eds.; Elsevier-Pergamon: Oxford, UK, 2003; pp. 1–64. [Google Scholar]
  36. Barth, M.G.; Mcdonough, W.F.; Rudnick, R.L. Tracking the budget of Nb and Ta in the continental crust. Chem. Geol. 2000, 165, 197–213. [Google Scholar] [CrossRef]
  37. Hofmann, A.W.; Jochum, K.P.; Seufert, M.; White, W.M. Nb and Pb in oceanic basalts: New constraints on mantle evolution. Earth Planet. Sci. Lett. 1986, 79, 33–45. [Google Scholar] [CrossRef]
  38. Green, T.H. Significance of Nb/Ta as an indicator of geochemical processes in the crust-mantle system. Chem. Geol. 1995, 120, 347–359. [Google Scholar] [CrossRef]
  39. 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]
  40. Patchett, P.J. Importance of the Lu-Hf isotopic system in studies of planetary chronology and chemical evolution. Geochim. Cosmochim. Acta 1983, 47, 81–91. [Google Scholar] [CrossRef]
  41. Wu, F.Y.; Li, X.H.; Zheng, Y.F.; Gao, S. Lu-Hf isotopic systematics and their applications in petrology. Acta Petrol. Sin. 2007, 23, 185–220, (In Chinese with English Abstract). [Google Scholar]
  42. Kemp, A.I.S.; Hawkesworth, C.J.; Foster, G.L.; Paterson, B.A.; Woodhead, J.D.; Hergt, J.M. Magmatic and crustal differentiation history of granitic rocks from Hf-O isotopes in zircon. Science 2007, 315, 980–983. [Google Scholar] [CrossRef]
  43. Mao, J.W.; Zhang, Z.C.; Zhang, Z.H.; Du, A.D. Re-Os isotopic dating of molybdenites in the Xiaoliugou W (Mo) deposit in the northern Qilian mountains and its geological significance. Geochim. Cosmochim. Acta 1999, 63, 1815–1818. [Google Scholar]
  44. Mcculloch, M.T.; Gamble, J.A. Geochemical and geodynamical constraints on subduction zone magmatism. Earth Planet. Sci. Lett. 1991, 102, 358–374. [Google Scholar] [CrossRef]
  45. Castillo, P.R. An overview of adakite petrogenesis. Chin. Sci. Bull. 2006, 51, 617–627. [Google Scholar] [CrossRef]
  46. 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] [Green Version]
  47. Chai, H.; Ma, Y.F.; Santosh, M.; Hao, S.L.; Luo, T.W.; Fan, D.Q.; Gao, B.; Zong, L.B.; Mao, H.; Wang, Q.F. Late Carboniferous to Early Permian oceanic subduction in central Inner Mongolia and its correlation with the tectonic evolution of the southeastern Central Asian Orogenic Belt. Gondwana Res. 2020, 84, 245–259. [Google Scholar] [CrossRef]
  48. He, S.S.; Li, Q.G.; Wang, Z.Q.; Xu, X.Y.; Liu, S.W.; Chen, J.L. Zircon U-Pb-Hf Isotopic Characteristics from Felsic Volcanic Rocks of Baoligaomiao Formation, the Middle Segment of Inner Mongolia: Implications for Geological Evolution. Acta Sci. Nat. Univ. Pekin. 2015, 51, 50–64, (In Chinese with English Abstract). [Google Scholar]
  49. Zhang, X.H.; Wilde, S.A.; Zhang, H.F.; Zhai, M.G. Early Permian high-K calc-alkaline volcanic rocks from NW Inner Mongolia, North China: Geochemistry, origin and tectonic implications. J. Geol. Soc. 2011, 168, 525–543. [Google Scholar] [CrossRef]
  50. Chai, H.; Wang, Q.; Tao, J.; Santosh, M.; Ma, T.; Zhao, R. Late Carboniferous to Early Permian magmatic pulses in the Uliastai continental margin linked to slab rollback: Implications for evolution of the Central Asian Orogenic Belt. Lithos 2018, 308–309, 134–158. [Google Scholar] [CrossRef]
  51. Chen, B.; Jahn, B.; 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]
  52. Streck, M.J.; Dilles, J.H. Sulfur evolution of oxidized arc magmas as recorded in apatite from a porphyry copper batholith. Geology 1998, 26, 523–526. [Google Scholar] [CrossRef]
  53. Ballard, J.R.; Palin, M.J.; Campbell, I.H. Relative oxidation states of magmas inferred from Ce(IV)/Ce(III) in zircon: Application to porphyry copper deposits of Northern Chile. Contrib. Mineral. Petrol. 2002, 144, 347–364. [Google Scholar] [CrossRef]
  54. Trail, D.; Watson, E.B.; Tailby, N.D. Ce and Eu anomalies in zircon as proxies for the oxidation state of magmas. Geochim. Cosmochim. Acta 2012, 97, 70–87. [Google Scholar] [CrossRef]
  55. Watson, E.B.; Wark, D.A.; Thomas, J.B. Crystallization thermometers for zircon and rutile. Contrib. Mineral. Petrol. 2006, 151, 413–433. [Google Scholar] [CrossRef]
  56. 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]
  57. Zou, X.; Qin, K.; Han, X.; Li, G.; Evans, N.J.; Li, Z.; Yang, W. Insight into zircon REE oxy-barometers: A lattice strain model perspective. Earth Planet. Sci. Lett. 2019, 506, 87–96. [Google Scholar] [CrossRef]
  58. Xu, K.Q.; Zhu, J.C. Time-Space Distribution of Tin/Tungsten Deposits in South China and Controlling Factors of Mineralization. In Geology of Tin Deposits in Asia and the Pacific: Selected Papers from the International Symposium on the Geology of Tin Deposits Held in Nanning, China, October 26–30; Hutchison, C.S., Ed.; Springer: Berlin/Heidelberg, Germany, 1984; pp. 265–277. [Google Scholar]
  59. Lehmann, B. Metallogeny of tin magmatic differentiation versus geochemical heritage. Econ. Geol. 1982, 77, 50–59. [Google Scholar] [CrossRef]
  60. Kelley, K.A.; Cottrell, E. Water and the oxidation state of subduction zone magmas. Science 2009, 325, 605–607. [Google Scholar] [CrossRef] [Green Version]
  61. Chen, Y.-J.; Wang, P.; Li, N.; Yang, Y.-F.; Pirajno, F. The collision-type porphyry Mo deposits in Dabie Shan, China. Ore Geol. Rev. 2017, 81, 405–430. [Google Scholar] [CrossRef]
  62. Lu, X.X.; Yu, P.; Feng, Y.L.; Wang, Y.T.; Ma, W.F.; Cui, H.F. Mineralization and Tectonic Setting of Deep hypabyssal Granites in East Qinling Mountain. Depos. Geol. 2002, 21, 168–178, (In Chinese with English Abstract). [Google Scholar]
  63. Gao, J.; Klemd, R.; Zhu, M.; Wang, X.; Li, J.; Wan, B.; Xiao, W.; Zeng, Q.; Shen, P.; Sun, J.; 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]
  64. Zheng, W.; Yu, X.-F. Geochronological and Geochemical Constraints on the Petrogenesis and Geodynamic Setting of the Daheishan Porphyry Mo Deposit, Northeast China. Resour. Geol. 2017, 68, 1–21. [Google Scholar] [CrossRef] [Green Version]
  65. Kang, Y.; She, H.; Lai, Y.; Wang, Z.; Li, J.; Zhang, Z.; Xiang, A.; Jiang, Z. Evolution of Middle-Late Triassic granitic intrusions from the Badaguan Cu-Mo deposit, Inner Mongolia: Constraints from zircon U-Pb dating, geochemistry and Hf isotopes. Ore Geol. Rev. 2018, 95, 195–215. [Google Scholar] [CrossRef]
  66. Chen, L.; Zhang, Y. In situ major-, trace-elements and Sr-Nd isotopic compositions of apatite from the Luming porphyry Mo deposit, NE China: Constraints on the petrogenetic-metallogenic features. Ore Geol. Rev. 2018, 94, 93–103. [Google Scholar] [CrossRef]
  67. Richards, J.P. Magmatic to hydrothermal metal fluxes in convergent and collided margins. Ore Geol. Rev. 2011, 40, 1–26. [Google Scholar] [CrossRef]
  68. Xing, K.; Shu, Q.; Lentz, D.R. Constraints on the Formation of the Giant Daheishan Porphyry Mo Deposit (NE China) from Whole-Rock and Accessory Mineral Geochemistry. J. Petrol. 2021, 62, egab018. [Google Scholar] [CrossRef]
  69. 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 juvenile lower crust in intra-plate extensional environment. Lithos 2014, 202–203, 138–156. [Google Scholar] [CrossRef]
Figure 1. (a) Tectonic frameworks of the CAOB; modified after [6]. (b) Tectonic location map of north-east China; modified after [7]. (c) Geological map of the southwest Erenhot-East Ujimqi district, showing distribution of major Mo-bearing deposits, modified from [8].
Figure 1. (a) Tectonic frameworks of the CAOB; modified after [6]. (b) Tectonic location map of north-east China; modified after [7]. (c) Geological map of the southwest Erenhot-East Ujimqi district, showing distribution of major Mo-bearing deposits, modified from [8].
Minerals 12 00606 g001
Figure 2. Simplified geological map of intrusions, wall rocks, and structures in the Zhunsujihua porphyry Mo-Cu mining area; modified after [9,13].
Figure 2. Simplified geological map of intrusions, wall rocks, and structures in the Zhunsujihua porphyry Mo-Cu mining area; modified after [9,13].
Minerals 12 00606 g002
Figure 3. Geological section along No. 8 exploration line of the Zhunsujihua Mo-Cu deposit; modified after [9,13].
Figure 3. Geological section along No. 8 exploration line of the Zhunsujihua Mo-Cu deposit; modified after [9,13].
Minerals 12 00606 g003
Figure 4. Hand specimens and photomicrographs of rocks from the Zhunsujihua porphyry Mo-Cu deposit and mineralization features. (a) Granodiorite. (b) Granodiorite porphyry. (c) Quartz-molybdenite vein in the granodiorite. (d,e) Granodiorite and granodiorite porphyry under orthogonal polarizing microscope, respectively. (f) Flaky molybdenite in close association with sericite. Abbreviations: Bi = biotite, Kfs = K-feldspar, Mo = molybdenite, Pl = plagioclase, Ser = sericite, Qtz = quartz.
Figure 4. Hand specimens and photomicrographs of rocks from the Zhunsujihua porphyry Mo-Cu deposit and mineralization features. (a) Granodiorite. (b) Granodiorite porphyry. (c) Quartz-molybdenite vein in the granodiorite. (d,e) Granodiorite and granodiorite porphyry under orthogonal polarizing microscope, respectively. (f) Flaky molybdenite in close association with sericite. Abbreviations: Bi = biotite, Kfs = K-feldspar, Mo = molybdenite, Pl = plagioclase, Ser = sericite, Qtz = quartz.
Minerals 12 00606 g004
Figure 5. Representative cathodoluminescence (CL) images of dated zircon crystals from granodiorite (Sample ZK310-277) and granodiorite porphyry (Sample ZK1105-135). The yellow numbers represent 206Pb/238U ages ± 1σ in Ma. LA-ICP-MS pits are marked with circles. The white numbers represent the spot number.
Figure 5. Representative cathodoluminescence (CL) images of dated zircon crystals from granodiorite (Sample ZK310-277) and granodiorite porphyry (Sample ZK1105-135). The yellow numbers represent 206Pb/238U ages ± 1σ in Ma. LA-ICP-MS pits are marked with circles. The white numbers represent the spot number.
Minerals 12 00606 g005
Figure 6. Zircon U-Pb concordia diagrams of the Zhunsujihua granodiorite (Sample ZK310-277) and granodiorite porphyry (Sample ZK1105-135).
Figure 6. Zircon U-Pb concordia diagrams of the Zhunsujihua granodiorite (Sample ZK310-277) and granodiorite porphyry (Sample ZK1105-135).
Minerals 12 00606 g006
Figure 7. Zircon chondrite-normalized REE distribution patterns of granodiorite porphyry and granodiorite in the Zhunsujihua Mo-Cu deposit (data normalized after [27]).
Figure 7. Zircon chondrite-normalized REE distribution patterns of granodiorite porphyry and granodiorite in the Zhunsujihua Mo-Cu deposit (data normalized after [27]).
Minerals 12 00606 g007
Figure 8. Classification of the Zhunsujihua granite samples on SiO2 vs. (Na2O + K2O) diagram, modified from [29].
Figure 8. Classification of the Zhunsujihua granite samples on SiO2 vs. (Na2O + K2O) diagram, modified from [29].
Minerals 12 00606 g008
Figure 9. SiO2-K2O diagram for granodiorite porphyry and granodiorite in the Zhunsujihua porphyry Mo-Cu deposit.
Figure 9. SiO2-K2O diagram for granodiorite porphyry and granodiorite in the Zhunsujihua porphyry Mo-Cu deposit.
Minerals 12 00606 g009
Figure 10. A/NK-A/CNK diagram for granodiorite porphyry and granodiorite in Zhunsujihua porphyry Mo-Cu deposit.
Figure 10. A/NK-A/CNK diagram for granodiorite porphyry and granodiorite in Zhunsujihua porphyry Mo-Cu deposit.
Minerals 12 00606 g010
Figure 11. (a) Whole-rock chondrite-normalized REE diagram and (b) whole-rock primitive mantle-normalized trace element spider diagram for the Zhunsujihua granodiorite porphyry and granodiorite. Chondrite and primitive mantle data are both from [27].
Figure 11. (a) Whole-rock chondrite-normalized REE diagram and (b) whole-rock primitive mantle-normalized trace element spider diagram for the Zhunsujihua granodiorite porphyry and granodiorite. Chondrite and primitive mantle data are both from [27].
Minerals 12 00606 g011
Figure 12. εNd (t)-(87Sr/86Sr)i and εHf (t)-Age diagrams. (a) The field for NE China Granite is from [34]. Isotope modeling for a simple mixing in the source region of Zhunsujihua granitoid plutons is also shown (only the samples with (87Rb/86Sr)i ≤ 10 are plotted in this diagram). The mantle-derived component is represented by basaltic rock (B) with the following calculated parameters: εNd (t) = +8, Nd = 15 ppm, (87Sr/86Sr)i = 0.704, Sr = 200 ppm. The upper continental crust (UC) component (the red star) is represented by Proterozoic Mashan Complex metamorphic rocks with εNd (t) = −12, Nd = +30 ppm, (87Sr/86Sr)i = 0.740, Sr = 250 ppm; for lower continental crust (LC) (the yellow square): εNd (t) = −15, Nd = 20 ppm, (87Sr/86Sr)i = 0.708, Sr = 20 ppm. All end-number data are derived from [34]. (b) Plot of εNd (t) vs. crystallization ages of zircons for the Zhunsujihua granitic intrusions. The Hf isotopic evolution line of the Archean average crust with ƒLu/Hf = 0.015 is after [39].
Figure 12. εNd (t)-(87Sr/86Sr)i and εHf (t)-Age diagrams. (a) The field for NE China Granite is from [34]. Isotope modeling for a simple mixing in the source region of Zhunsujihua granitoid plutons is also shown (only the samples with (87Rb/86Sr)i ≤ 10 are plotted in this diagram). The mantle-derived component is represented by basaltic rock (B) with the following calculated parameters: εNd (t) = +8, Nd = 15 ppm, (87Sr/86Sr)i = 0.704, Sr = 200 ppm. The upper continental crust (UC) component (the red star) is represented by Proterozoic Mashan Complex metamorphic rocks with εNd (t) = −12, Nd = +30 ppm, (87Sr/86Sr)i = 0.740, Sr = 250 ppm; for lower continental crust (LC) (the yellow square): εNd (t) = −15, Nd = 20 ppm, (87Sr/86Sr)i = 0.708, Sr = 20 ppm. All end-number data are derived from [34]. (b) Plot of εNd (t) vs. crystallization ages of zircons for the Zhunsujihua granitic intrusions. The Hf isotopic evolution line of the Archean average crust with ƒLu/Hf = 0.015 is after [39].
Minerals 12 00606 g012
Figure 13. Non-active element diagram of granite tectonic environment discrimination. (a) Sr/Y vs. Y diagram, according to [45]. (b) Rb vs. (Y + Nb) diagram, according to [46]. VAG = volcanic arc granite, WPG = within plate granite, Syn-COLG = syn-collision granite, ORG = ocean ridge granite. Legends are the same as in Figure 9.
Figure 13. Non-active element diagram of granite tectonic environment discrimination. (a) Sr/Y vs. Y diagram, according to [45]. (b) Rb vs. (Y + Nb) diagram, according to [46]. VAG = volcanic arc granite, WPG = within plate granite, Syn-COLG = syn-collision granite, ORG = ocean ridge granite. Legends are the same as in Figure 9.
Minerals 12 00606 g013
Figure 14. The oxygen fugacity logƒ(O2)-1000/T diagram of granodiorite porphyry and granodiorite. Abbreviations: F-M-Q = fayalite-magnetite-quartz buffer line, H-M = hematite-magnetite buffer line. SO2-H2S buffer lines are also shown in the diagram.
Figure 14. The oxygen fugacity logƒ(O2)-1000/T diagram of granodiorite porphyry and granodiorite. Abbreviations: F-M-Q = fayalite-magnetite-quartz buffer line, H-M = hematite-magnetite buffer line. SO2-H2S buffer lines are also shown in the diagram.
Minerals 12 00606 g014
Table 1. Zircon U-Pb LA-ICP-MS analytical results of granodiorite and granodiorite porphyry from Zhunsujihua Mo-Cu deposit.
Table 1. Zircon U-Pb LA-ICP-MS analytical results of granodiorite and granodiorite porphyry from Zhunsujihua Mo-Cu deposit.
UTh Isotopic RatiosDates (Ma)
Spot No.ppmppmTh/U207Pb/206Pb±2σ207Pb/235U±2σ206Pb/238U±2σ207Pb/235U±2σ206Pb/238U±2σ
Granodiorite
Z310-03306.5461.71.510.052830.0020.358040.025920.049150.00104311103096
Z310-04543.9218.60.400.052650.001760.342870.021840.047220.0009829982976
Z310-10270.3114.80.420.05280.002160.349880.027520.048050.00104305103036
Z310-12409.3160.20.390.052040.001840.338950.023040.047230.0009829682976
Z310-15185.2149.70.810.051890.002460.340390.031160.047570.00108297123006
Z310-16322.1161.30.500.05230.002020.34860.025960.048330.00104304103046
Z310-17481.0225.80.470.052620.00180.35470.023360.048880.0010230883086
Z310-18231.4151.50.650.052340.002340.352170.03040.048790.00108306123076
Z310-21235.7145.80.620.053320.002340.351360.02960.047780.00106306123016
Z310-22603.3238.30.400.051750.001720.332260.02120.046560.0009629182936
Z310-23630.6342.40.540.051230.001680.340830.021480.048240.00129883046
Z310-25813.3361.80.440.053080.001660.355560.021280.048580.00130983066
Z310-26866.8350.80.400.052230.001640.336660.020160.046740.0009629582946
Z310-31368.3109.10.300.05290.002040.351460.026120.048180.00104306103036
Z310-40865.3363.60.420.05290.001720.349010.021720.047840.00130483016
Z310-41659.6478.00.720.052080.00180.338150.022360.047090.0009829682976
Granodiorite porphyry
Z1105-01827.1284.90.340.054080.00160.341170.009760.045770.0009429882886
Z1105-04291.9271.30.930.054910.002420.352180.015060.046520.00106306122936
Z1105-05268.8282.81.050.053640.002160.349720.013660.04730.00104305102986
Z1105-08332.2156.80.470.057970.002180.377670.013760.047260.00102325102986
Z1105-11429.4133.40.310.060940.002640.397650.01660.047330.00108340122986
Z1105-14562.1581.61.030.057990.001860.367840.011380.046010.0009631882906
Z1105-16283.5262.30.930.054610.002160.345930.01320.045950.001302102906
Z1105-17202.899.80.490.063610.002760.423810.017640.048330.00110329123046
Z1105-19329.9163.60.50.056870.002420.380050.013960.048470.00104327103056
Z1105-20252.374.60.30.052840.002160.346750.01370.04760.00104302103006
Z1105-21634.0250.40.390.055990.00190.362590.01180.046970.0009831482966
Z1105-29461.5340.80.740.054770.001980.349260.012120.046250.00098304102916
Z1105-31166.483.60.50.052260.002740.34090.017260.047310.00112298142986
Z1105-341086.987.40.080.051940.001540.342710.00970.047860.0009829983016
Z1105-351040.1122.30.120.052650.001560.343810.009760.047360.0009630082986
Z1105-37162.4130.20.80.053320.002720.339920.016740.046240.00108297122916
Note: All analyses in the table are used in the mean age calculation.
Table 2. Trace (ppm) element data, Ce4+/Ce3+ values, and logƒ(O2) calculated for zircons from granodiorite, granodiorite porphyry in the Zhunsujihua Mo-Cu deposit.
Table 2. Trace (ppm) element data, Ce4+/Ce3+ values, and logƒ(O2) calculated for zircons from granodiorite, granodiorite porphyry in the Zhunsujihua Mo-Cu deposit.
Spot No.LaCePrNdSmEuGdTbDyHoErTmYbLuUThTiT (℃)Ce4+/Ce3+δEulgƒ (O2)∆FMQ
ZK1105-010.035.090.070.872.110.6214.96.2383.934.616335.034567.9230484.18713390.50−15.90+0.5
ZK1105-020.0326.80.132.187.073.7453.218.421177.033166.862312129227111.8813500.88−10.12+3.9
ZK1105-030.0526.30.457.7116.46.7875.023.524786.136773.87051392692838.25777110.88−17.27−2.4
ZK1105-040.074.870.080.992.280.3916.17.2410343.822150.452410456210516.3849420.29−9.27+4.0
ZK1105-050.0713.50.151.763.911.0822.48.9611245.220944.844186.53321573.46697450.52−16.32+0.5
ZK1105-060.0523.80.070.892.610.6219.57.9510340.718840.238775.86573122.346641600.39−13.44+4.2
ZK1105-070.0913.50.181.764.030.6421.38.8311243.720644.143785.54291333.33693440.31−16.55+0.3
ZK1105-080.0726.50.152.546.962.5744.616.318970.329959.55531062832628.07774440.66−12.43+2.4
ZK1105-090.0612.70.121.513.050.3717.37.4398.240.119141.640881.02031003.58699570.23−15.30+1.4
ZK1105-100.0824.20.141.673.711.0422.79.1811847.522046.245187.53301645.29734870.51−11.86+4.0
ZK1105-110.0753.50.213.217.192.5541.015.418771.431965.66261184623416.01746820.67−11.49+4.1
ZK1105-120.024.770.122.525.090.6327.59.8411642.918637.735367.1166847.9777380.24−18.71−3.8
ZK1105-130.098.560.111.963.640.2218.76.4474.027.612224.323346.0156573.17689190.12−19.91−2.9
ZK1105-140.0211.30.183.078.382.1243.914.214752.222143.841682.81621306.15748120.50−18.44−3.0
ZK1105-150.0921.70.101.473.450.9824.59.4412046.721645.744889.34842595.29734870.48−11.86+4.0
Average-----------------740520.48−14.59+1.1
ZK310-010.0239.70.070.992.811.0914.04.7254.320.696.721.622446.17804712.816791900.97−11.93+5.3
ZK310-020.0320.20.040.762.700.6719.88.3411045.921947.847394.25442192.996841710.51−12.05+5.1
ZK310-030.0317.60.071.383.481.0021.38.2810441.218840.039077.83362244.77725700.65−13.15+2.9
ZK310-040.0517.60.040.672.220.5716.16.5987.535.916836.836372.24091603.867061650.53−10.98+5.6
ZK310-050.0621.80.061.092.770.8320.38.2010643.020243.843186.34812264.207131280.62−11.52+4.8
ZK310-060.0922.20.050.702.720.5718.47.7510140.919041.140178.66032383.106871820.45−11.63+5.4
ZK310-070.025.820.111.865.140.7124.47.4376.026.411022.121042.03681098.7078290.36−17.79−3.1
ZK310-080.0718.10.152.926.761.7136.013.415860.827056.15301043882155.59739280.61−15.72+0.0
ZK310-090.1036.20.101.705.271.3132.112.415056.925351.849494.46725347.23764990.57−9.94+5.2
ZK310-100.0327.70.081.373.870.6723.49.6112751.324454.05431108653642.866811270.40−13.35+3.9
Average-----------------7161170.57−12.81+3.5
Table 3. Major (wt.%) and trace (ppm) elements of granodiorite and granodiorite porphyry from the Zhunsujihua Mo-Cu deposit.
Table 3. Major (wt.%) and trace (ppm) elements of granodiorite and granodiorite porphyry from the Zhunsujihua Mo-Cu deposit.
RocksGranodioriteGranodiorite Porphyry
SamplesZK310-277ZK1105-81ZK1105-206ZK1105-264Z1105-155ZK701-03ZK310-52ZK1105-135
SiO269.4370.7669.9469.7271.4571.0170.5170.62
TiO20.340.340.360.350.280.280.280.30
Al2O314.8214.6214.8314.7515.2915.0614.8815.02
TFe2O33.012.862.892.862.312.542.492.80
MnO0.030.050.040.030.040.050.070.05
MgO0.850.820.880.850.750.750.760.84
CaO2.602.612.562.052.192.172.241.90
Na2O3.853.933.843.904.454.404.374.59
K2O3.413.283.183.862.982.782.702.96
P2O50.110.120.120.120.100.100.100.10
LOI1.411.131.451.391.000.921.091.09
Total98.5998.87 98.55 98.61 99.00 99.08 98.91 98.91
NK7.267.21 7.02 7.76 7.43 7.18 7.07 7.55
A/CNK1.00 0.99 1.03 1.03 1.05 1.06 1.05 1.06
A/NK1.48 1.46 1.52 1.39 1.45 1.47 1.47 1.40
Mg#36.09 36.44 37.85 37.28 39.37 37.13 37.91 37.50
A.I.0.68 0.68 0.66 0.72 0.69 0.68 0.68 0.72
Cs2.931.381.583.441.461.192.031.90
Rb107.591.497.2135.086.878.181.690.4
Ba503510477463397349374400
Sr309386344372482431445406
Th7.5111.607.468.186.687.467.247.13
U2.932.431.153.052.512.392.002.72
Nb6.46.35.96.35.75.35.45.5
Ta0.60.40.50.60.60.50.50.6
Zr163158160166129143141140
Hf4.24.94.04.23.33.63.53.6
La25.427.120.922.721.219.817.520.7
Ce51.158.245.148.441.544.639.345.0
Pr5.866.845.325.674.655.194.575.34
Nd22.625.920.522.017.819.217.320.5
Sm4.975.474.384.643.813.993.584.41
Eu0.780.840.760.820.790.830.740.86
Gd3.854.313.583.813.283.312.973.46
Tb0.590.640.540.590.530.480.470.51
Dy3.203.392.903.092.822.772.522.81
Ho0.620.670.560.610.550.530.510.54
Er1.691.791.521.681.521.441.421.48
Tm0.260.270.220.250.250.230.220.23
Yb1.531.631.411.561.451.471.341.37
Lu0.260.270.250.260.250.210.230.24
Y18.519.616.718.417.015.215.116.1
Ga20.320.419.8519.8019.3519.2519.2518.75
Tl0.620.430.550.820.520.480.490.59
V4339405333172931
W22121112
Sn1.71.71.81.91.51.61.51.9
Mo2.121.981.341.232.291.982.182.56
Bi0.160.250.180.310.450.190.310.08
Cu29.833.953.680.30.90.70.70.7
Pb9.38.510.18.59.58.58.97.6
Zn2442302527284332
Ag0.040.050.080.11<0.010.01<0.01<0.01
As3.93.73.54.53.33.24.13.0
Be4.714.882.822.913.422.862.532.86
Li17.315.916.923.913.114.514.917.9
∑REE122.71 137.32 107.94 116.08 100.40 104.05 92.67 107.45
(La/Yb)N11.91 11.93 10.63 10.44 10.49 9.66 9.37 10.84
Sr/Y16.70 19.69 20.60 20.22 28.35 28.36 29.47 25.22
δEu0.53 0.51 0.57 0.58 0.67 0.68 0.67 0.65
Ce/Pb5.49 6.85 4.47 5.69 4.37 5.25 4.42 5.92
Nb/Ta10.67 15.75 11.80 10.50 9.50 10.60 10.80 9.17
Table 4. Sr-Nd isotopic data of the Zhunsujihua granitic intrusions.
Table 4. Sr-Nd isotopic data of the Zhunsujihua granitic intrusions.
Rock RbSr SmNd t TDM2
TypeSampleppmppm87Rb/86Sr87Sr/86SrISrppmppm147Sm/144Nd143Nd/144NdMaεNd(t)Ma
GDZS-013168.52911.670.71279090.705614.9624.60.12200.5126464302+3.04807
PG-01100.03400.850.70826790.704623.7915.20.15090.5126785302+2.55785
ZK310-01151.52871.520.71066690.704123.5518.20.11800.5126565302+3.39787
GPZK701-0379.44570.500.705796110.703684.2420.00.12830.5126635296+3.08787
ZK1105-16583.84560.530.70601280.703783.7918.30.12530.5126644296+3.21782
ZK707-0274.8449.30.480.705695100.703672.5418.90.11320.5126749296+3.86754
Note: 87Rb decay λ = 1.42 × 10−11 year−1; 147Sm decay k = 6.54 × 10−12 year−1; εNd (t) were calculated with modern (143Nd/144Nd)CHUR = 0.512638 and (147Sm/144Nd)CHUR = 0.1967, and TDM were calculated using present-day (147Sm/144Nd)DM = 0.2137 and (143Nd/144Nd)DM = 0.51315 [30]. GD = granodiorite; DP = granodiorite porphyry; ISr = (87Sr/86Sr)i.
Table 5. Hf isotope data of zircons from granodiorite, granodiorite porphyry of the Zhunsujihua Mo-Cu deposit.
Table 5. Hf isotope data of zircons from granodiorite, granodiorite porphyry of the Zhunsujihua Mo-Cu deposit.
t TDM1TDM2
Spots176Hf/177Hf176Yb/177Hf176Lu/177Hf(176Hf/177Hf)iMaεHf(t)MaMa
Granodiorite
ZS-013-010.2827710.0000090.0569920.0005170.0015240.0000070.2827623026.31.2691837
ZS-013-020.2826930.0000130.0712570.0002420.0023270.0000090.282683023.41.4820999
ZS-013-030.2828190.0000110.0512560.0009250.0017170.0000150.282813028.01.3625744
ZS-013-040.2828440.000010.0510470.0001160.0014270.0000060.2828363028.91.3585693
ZS-013-050.2828080.0000130.0622930.0004360.002010.0000150.2827973027.51.4647770
ZS-013-060.2828090.0000110.0455040.0004890.0015670.0000220.28283027.61.3638763
ZS-013-070.282840.0000090.072630.0003420.0020590.0000180.2828283028.61.2601708
ZS-013-080.2826440.0000120.0757240.0007990.0023950.0000210.2826313021.61.48941097
ZS-013-090.2828320.000010.0644370.0004540.0017620.0000130.2828223028.41.3608721
ZS-013-100.2826310.0000140.0544150.0006390.0018860.0000280.282623021.31.59011118
ZS-013-110.2828260.0000120.0543080.0005510.0015450.0000160.2828173028.21.4613730
ZS-013-120.2828420.0000110.0501720.0010440.0013780.0000270.2828353028.91.3586695
Granodiorite porphyry
ZK-701-010.2828480.0000090.0759130.0010620.0020440.0000410.2828362968.81.3590694
ZK-701-020.282850.0000090.0776240.0006570.0021820.0000130.2828382968.81.2588691
ZK-701-030.2828320.0000120.0577930.0009220.0018110.0000460.2828212968.31.4609723
ZK-701-040.2827330.0000160.0770750.0011790.0025750.0000480.2827192964.61.6767925
ZK-701-050.2826640.0000140.1372640.001370.0040890.0000560.2826412961.91.59071151
ZK-701-060.282760.0000120.1277530.0034590.0037830.0001410.2827392965.31.4753941
ZK-701-070.2828650.0000090.0606110.0005760.0017770.000030.2828552969.41.3561666
ZK-701-080.282840.000010.0545380.0010710.0014250.0000210.2828322968.61.3590712
ZK-701-090.2828460.0000110.0708340.0013590.0019460.0000440.2828352968.81.3590713
ZK-701-100.2828190.0000130.0968860.0011030.0026230.0000490.2828042967.71.4641785
ZK-701-110.2828570.000010.0496790.0006550.0013010.000010.282852969.31.3564670
ZK-701-120.2828480.0000110.074640.0004610.0019680.000020.2828372968.81.3588708
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shi, Q.; Guo, H.; Liu, C.; Lai, Y. Genesis and Metallogenic Characteristics of the Zhunsujihua Granitic Intrusions in Sonid Left Banner, Inner Mongolia, China. Minerals 2022, 12, 606. https://0-doi-org.brum.beds.ac.uk/10.3390/min12050606

AMA Style

Shi Q, Guo H, Liu C, Lai Y. Genesis and Metallogenic Characteristics of the Zhunsujihua Granitic Intrusions in Sonid Left Banner, Inner Mongolia, China. Minerals. 2022; 12(5):606. https://0-doi-org.brum.beds.ac.uk/10.3390/min12050606

Chicago/Turabian Style

Shi, Qianxiong, Hu Guo, Cong Liu, and Yong Lai. 2022. "Genesis and Metallogenic Characteristics of the Zhunsujihua Granitic Intrusions in Sonid Left Banner, Inner Mongolia, China" Minerals 12, no. 5: 606. https://0-doi-org.brum.beds.ac.uk/10.3390/min12050606

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