Next Article in Journal
Synthesis and Characterization of the Double Perovskite Y2MgRuO6
Previous Article in Journal
Research Status and Development Trend of Piezoelectric Accelerometer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the CAM18 Crystal as a Potential Reference Material for U–Pb Analysis of Zircon

1
School of Gemology, China University of Geosciences, Beijing 100083, China
2
State Key Laboratory of Geological Processes and Mineral Resources, China University of Geosciences, Beijing 100083, China
3
The Beijing SHRIMP Center, Chinese Academy of Geological Sciences, Beijing 100037, China
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(9), 1364; https://doi.org/10.3390/cryst13091364
Submission received: 2 August 2023 / Revised: 7 September 2023 / Accepted: 8 September 2023 / Published: 11 September 2023
(This article belongs to the Section Mineralogical Crystallography and Biomineralization)

Abstract

:
In the process of in situ zircon U–Pb dating, it is an effective means to overcome the matrix effect by using a matrix-matched external reference material. However, the limited number of available zircon reference materials still makes it difficult to meet the research needs. In this paper, we performed a preliminary analysis of the gemological characteristics, trace elements and U–Pb ages of natural zircon CAM18 to assess its suitability as a reference material for laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) U–Pb dating. This tawny, gem-quality zircon has no visible inclusions and weighs approximately 0.55 g. Its density, full width at half maximum (FWHM) of the Raman peak and alpha flux (Dα) indicate that the sample has suffered mild-to-moderate radiation damage without any thermal treatment. The LA-ICP-MS U–Pb dating results reveal that the trace elements content and U–Pb ages of the sample are fairly homogeneous at the 50 μm scale, and there is no obvious loss of radiogenic Pb. The 206Pb/238U age (571.0 ± 3.0 Ma, 2s) and 207Pb/235U age (573.4 ± 6.0 Ma, 2s) are consistent within the analytical uncertainty, and the calculated concordia age is 571.4 ± 1.4 Ma (2s, n = 20). The variation in the 206Pb/238U ages is small, with a measurement repeatability of 0.46% (RSD), which is within the uncertainty of the age accuracy obtained by LA-ICP-MS. The oscillatory zoning, Th/U ratio (0.2) and chondrite-normalized rare-earth element (REE) pattern imply a magmatic origin of zircon CAM18. The Ti-in-zircon temperature ranges from 714 to 742 °C, and the oxygen fugacity ranges from ΔFMQ−2.87 to ΔFMQ−3.17, suggesting that it is crystallized in a reducing environment. All the results show that zircon CAM18 may has great potential in LA-ICP-MS U–Pb dating.

1. Introduction

Gem-quality zircon has a subadamantine luster, high dispersion and high refractive index, giving it a diamond-like appearance. It is favored by consumers because it exhibits a series of rich colors, such as colorless, green, red, orange and yellow. Additionally, zircon is an accessory mineral found in magmatic, sedimentary and metamorphic rocks, possessing the following advantages: high U and Th content, low common Pb content, high closure temperature (>750 °C) and stable physico-chemical properties [1,2,3]. These properties make zircon an excellent dating mineral that can effectively record the timing of various geological events [4,5].
The chemical formula for zircon is ZrSiO4, and it may contain small amounts of Hf, U, Th, Ti, Y, P and rare-earth elements (REEs) due to the isomorphic substitutions of Zr4+ and Si4+ in the structure. The crystal structure of zircon is composed of isolated [SiO4] tetrahedra and [ZrO8] dodecahedra. In the a-axis and b-axis directions, each [ZrO8] dodecahedron is connected to the surrounding four [ZrO8] dodecahedra by a shared edge, forming a long chain in the a-axis direction. In the c-axis direction, the [SiO4] tetrahedra and the [ZrO8] dodecahedra form [001] strong chains by means of sharing edges and corners. This structure determines that zircon often develops a prismatic habit, with high birefringence and hardness [6,7].
In recent years, the spatial resolution and analysis efficiency of laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) have been rapidly developed, making it the primary method for in situ U–Pb dating and trace element analysis [8,9]. However, during laser ablation, aerosol transport and ionization in ICP, instruments often produce elemental fractionation, impacting the precision and accuracy of dating results. The utilization of reference materials offers a useful method to ameliorate these problems. First, it can help us to become acquainted with the operating status of the instrument to determine the optimal instrument parameters; second, it can well reflect the dynamic change in detector sensitivity and monitor the analytical accuracy; third, it can be used to calibrate the U–Pb data of unknown samples.
Many zircon reference materials often possess some crucial dating conditions, such as a homogeneous chemical composition and structure, high 206Pb/204Pb ratios (˃10,000, e.g., SA01), moderate content of U (tens to hundreds of ppm, e.g., 91500), few mineral inclusions and an appropriate volume (grains several mm to cm in diameter) [2,10,11]. Research on zircon reference materials has brought fruitful results, including 91500, GJ-1, M257, Plešovice, QGNG and FC-1, of which the first three zircons are gem-quality [10,11,12,13,14,15]. Nevertheless, the use of these zircon reference materials is mostly limited by small amounts of storage or difficulty in mining and can hardly meet the long-term development needs of laboratories. Therefore, it is urgent to develop novel zircon reference materials.
Compared with multi-crystal zircons, the advantages of gem-quality zircons are that (1) they have stable structures as well as uniform chemical compositions; (2) they have no obvious radiation damage; (3) they contain relatively few cracks and mineral inclusions; (4) they are more likely to be concordant at the accepted age. Thus, in this paper, a gem-quality zircon from Cambodia was studied in terms of spectroscopic analyses, trace elements and isotope measurements. On the one hand, this was accomplished to explore whether the crystal has the potential to be a reference material for LA-ICP-MS U–Pb dating; on the other hand, it was accomplished in order to discuss its possible genesis and provide an important basis for subsequent provenance studies. In addition, this study can provide useful reference information for future applications of the crystal. It is ideal to use the crystal as a reference material for unknown zircons with a similar U–Pb age and isotopic composition, which helps to alleviate the matrix effects that may result from zircon decay accumulated damage.

2. Materials and Methods

The sample is a gem-quality zircon crystal (CAM18) purchased from a gemstone dealer who claimed that it had been collected from Cambodia, though the exact provenance and petrological characteristics are unknown (Figure 1).
Routine gemological signature observation and spectroscopic testing of the sample were performed at the Gemmological Research Laboratory of the China University of Geosciences (Beijing). The color, luster, transparency, inclusions, extinction phenomena, dichroism, fluorescence weight and density were observed and recorded by using a 10× magnifying glass, gem photographic microscope, polariscope, dichroscope, ultraviolet and electronic balance.
The UV–vis spectrometer was the UV-3600 UV–vis spectrophotometer produced by Shimadzu Corporation in Japan. The sample was analyzed by absorption method, and the experimental conditions were as follows: the wavelength range was 300~800 nm, the slit width was 20 nm, the time constant was 0.1 s, and the sampling interval was 0.5 s.
The infrared spectrometer was the Tensor 27 Fourier infrared spectrometer, manufactured by Bruker in Germany. The sample was analyzed by reflection method, and the scanning range was set to 400~1200 cm−1.
The Raman spectrum was collected in the range of 100~1100 cm−1 by the Horiba HR-Evolution Raman microspectrometer. The conditions were as follows: 532 nm laser source, 50 nW laser power, 4 cm−1 spectral resolution, 600 gr/mm grating and 100 μm slit width.
The preparation of sample target and the acquisition of related pictures were carried out at Beijing Zirconia Pilot Technology Co., Ltd. First, the selected sample and zircon reference materials were pasted on a glass plate with double-sided tape, and then the prepared epoxy resin was injected into the sample target for curing. Second, the cured sample target was removed from the glass plate for cleaning, sanding and polishing. Finally, the cathodoluminescence, transmitted light and reflected light images were photographed.
LA-ICP-MS U–Pb dating and trace element analysis were carried out at the mineral laser microprobe analysis laboratory of China University of Geosciences (Beijing). The instrument was mainly composed of NewWave 193UC ArF excimer laser and Angilent 7900 ICP-MS. The main tuning parameters are shown in Table 1.
Prior to performing the analysis, the ICP-MS was tuned by ablating the NIST SRM 610 to obtain a maximum signal intensity of 238U (1 × 106 cps), while maintaining a low ThO+/Th+ ratio (<0.15%) and a U+/Th+ ratio close to 1. During the analysis, the laser ablation energy density, spot size and repetition rate were set to 4 J/cm2, 50 μm and 8 Hz, respectively. The following elements were collected by ICP-MS: 29Si, 45Sc, 49Ti, 89Y, 91Zr, 93Nb, 139La, 140Ce, 141Pr, 143Nb, 147Sm, 151Eu, 115Gd, 159Tb, 163Dy, 165Ho, 166Er, 169Tm, 173Yb, 175Lu, 179Hf, 181Ta and 202Hg, which were measured with dwell time of 6 ms; 232Th, which was measured with dwell time of 10 ms; 238U and 208Pb, which were measured with dwell time of 15 ms; 204(Pb + Hg) and 206Pb, which were measured with dwell time of 20 ms; 207Pb, which was measured with dwell time of 30 ms. Each analytical run included 2 analyses of zircon 91500 (primary reference material), 2 analyses of zircon GJ-1 (secondary reference material) and 10 analyses of unknowns. NIST SRM 610 was used as an external reference material, and 91Zr was used as an internal standard to calibrate the trace elements. Each single-spot analysis consisted of 20 s of background signal acquisition, 50 s of ablation and 30 s of washout. Isoplot R was used to calculate the weighted average age and for the generation of Concordia and Tera–Wasserburg diagrams [16].

3. Results

3.1. Gemmological Properties

Zircon CAM18 is a cut gemstone in a near-round shape, approximately 8 mm in diameter, with a density of 4.49 g/cm3 and a weight of 0.55 g. The crystal is yellow–brown and transparent and has a vitreous luster. The pleochroism is manifested as light fuchsia and light yellow. The sample is an anisotropic gemstone that shows one extinction for each 90-degree rotation under a polariscope. It is inert under long-wave UV light (365 nm) but fluoresces faintly yellow under short-wave UV light (254 nm). Under the gem microscope, we can see two of each pavilion facet of the sample, which is caused by its high birefringence (0.040~0.060), but no other inclusions are observed (Figure 2A). After crushing, the sample develops a conchoidal fracture that exhibits a greasy luster (Figure 2B).

3.2. Spectral Characteristics

3.2.1. FTIR Spectra

All five infrared spectral analyses of zircon CAM18 show four peaks at 435 cm−1, 621 cm−1, 937 cm−1 and 1078 cm−1 (Figure 3). The 621 cm−1 reflectance peak is caused by the anti-symmetric bending vibration of υ4 (SiO4), corresponding to the internal A2u infrared-active mode. The 937 cm−1 and 1078 cm−1 reflectance peaks are associated with the anti-symmetric stretching vibration of υ3 (SiO4), corresponding to the internal Eu and A2u infrared-active modes, respectively [17]. The attribution of the 435 cm−1 reflectance peak remains controversial. Nasdala et al. [17] and Dawson et al. [18] attributed it to the anti-symmetric bending vibration of υ4 (SiO4), corresponding to the internal Eu infrared-active mode, whereas Woodhead et al. [19] attributed it to the external vibration mode caused by the rotary vibration of the [SiO4]4− tetrahedron.
The width and intensity of the 435 cm−1 and 621 cm−1 peaks can reflect the degree of zircon metamictization, especially the width of the 621 cm−1 peak, which is highly sensitive to metamictization and easy to quantitatively study, so these two peaks are a useful tool for determining the degree of zircon metamictization [20,21]. From the FTIR spectral analysis results of zircon CAM18, the width and intensity of the 435 cm−1 and 621 cm−1 peaks show a broadening trend and a weakening trend, respectively. This suggests that the sample has undergone a certain degree of metamictization.

3.2.2. UV–Vis–NIR Spectra

Although colorless zircons seldom exhibit prominent absorption peaks in the UV–vis–NIR spectra, zircons of different colors tend to show various absorption peaks, mainly due to the selective absorption of light by the transition metal elements (such as Nb, Ti and Fe) and rare-earth elements (such as Nd, Pr and Er) in the structure [22]. The UV–vis–NIR spectra of zircon CAM18 consists of eight peaks (Figure 4). The peaks at 652 nm and 689 nm are the characteristic peaks of zircon, mainly attributed to the substitution of U4+ for Zr4+ in the crystal lattice [23]. The appearance of other peaks is caused by the absorption of light by some impurity ions. Among them, the light brown of zircon CAM18 is mainly affected by the absorption peaks at 351 nm, 480 nm and 511 nm, and the origin may be related to the absorption of the color centers [23,24].

3.2.3. Raman Spectra

In Figure 5, the Raman peaks at 1003 cm−1, 972 cm−1 and 438 cm−1 represent the internal structure vibration peaks of the lattice, caused by the anti-symmetric stretching vibration (B1g, υ3), symmetric stretching vibration (A1g, υ1) and symmetric bending vibration (A1g, υ2) of the [SiO4]4− tetrahedron, respectively. The Raman peaks at 223 cm−1, 210 cm−1 and 203 cm−1 belong to the external structure vibration peaks, caused by Zr4+ and [SiO4]4− tetrahedral vibrations [17]. The origin of the 352 cm−1 Raman peak (Eg) is controversial. Syme et al. [25] and Koleso et al. [26] considered it to be an external vibration mode caused by [SiO4]4− tetrahedral vibration, while Dawson et al. [18] specified it as an internal vibration mode, caused by υ4 bending vibration.
The relative intensity variation in the 352 cm−1 and 1003 cm−1 Raman peaks are related to the anisotropy of the crystal. The strongest intensity of the 352 cm−1 peak is obtained in the direction perpendicular to the c-axis of zircon, and the strongest intensity of the 1003 cm−1 peak is obtained in the direction parallel to the c-axis [27]. With the increase in metamorphism, the structure of zircon becomes more and more disordered, which is reflected in the Raman spectra by the increase in the width of the Raman peaks and the decrease in the Raman shift [28]. Although the Raman peaks of zircon CAM18 have a small shift in the direction of the low wavenumber, the peak shapes are relatively sharp and small in width, so it can be judged that the sample is less affected by metamictization and still has a high degree of crystallization.

3.3. U–Pb Ages

In the CL image, most zircon fragments are grayish white with fewer cracks, reflecting less difference in trace element content (Figure 6). In addition, the sample developed a narrow and dense oscillatory zoning, indicating that it originated from magma [29]. The results of the LA-ICP-MS U–Pb dating are shown in Table A1, and the uncertainties are reported at the 1σ level. The 206Pb/238U ages range from 565 ± 3.4 to 575 ± 3.4 Ma, and the 207Pb/235U ages range from 562 ± 5.5 to 582 ± 5.4 Ma. The data are concentrated on and near the concordia line with high concordance, yielding a concordia age of 571.4 ± 1.4 Ma (2s) (Figure 7A). The 238U/206Pb and 207Pb/206Pb ratios (not corrected for common Pb) yield a lower-intercept age of 570.5 ± 4.0 Ma (2s) (Figure 7B). The weighted average 206Pb/238U age is 571.1 ± 3.0 Ma (2s), indicating that the sample is a product of Neoproterozoic magmatic activity (Figure 7C).

3.4. Trace Elements

The trace element data are shown in Table A2. The Th and U concentrations are 136 ± 3 ppm and 673 ± 14 ppm, respectively, with a Th/U ratio of 0.20; the REE content is 62.8~66.7 ppm, and the average value is 64.8 ppm; the LREE content is 3.82~4.18 ppm, and the average value is 4.00 ppm; the HREE content is 59.0~62.6 ppm, and the average value is 60.8 ppm. As shown in Figure 8, these points exhibit fairly uniform trace element abundances, with small ranges in the mass fractions. The average values of LREE/HREE and (La/Yb)N are 0.07 and 0.00009, respectively, indicating that the enrichment degree of HREE is significantly higher than that of LREE. The average values of (La/Sm)N and (Gd/Yb)N are 0.0021 and 0.11, respectively, indicating that the degree of fractionation inside LREE is greater than that of HREE. In addition, Ce and Eu show obvious positive and negative anomalies, with average values of 80.2 and 0.31 for δCe and δEu, respectively. Overall, zircon CAM18 has REE characteristics that are similar to most magmatic zircons.

4. Discussion

4.1. Structural State and Radiation Damage

The radioactive decay of elements like U and Th can cause radiation damage to the crystal structure of zircon, resulting in a decrease in the crystallinity and physico-chemical stability. In this case, radiogenic Pb is easily lost, forming a discordant U–Pb age dataset. In order to comprehensively assess the structural state of zircon CAM18, we studied it based on the density, Raman spectroscopy parameters and α flux (Dα).
The density of the sample is 4.49 g/cm3, consistent with typical intermediate-type zircon, showing that the sample has experienced a volume expansion under the effect of α self-irradiation [30]. With the increase in radiation intensity, the υ3 peak (1008 cm−1) of crystalline zircon shows a decrease in Raman shift and an increase in full width at half maximum (FWHM). Five analyses of zircon CAM18 suggest that the υ3 peak is located at 1003 cm−1, and the FWHMs after correction using the Lorentzian–Gaussian function vary from 8.9 to 11.3 cm−1 (Figure 5). These values reconfirm the mild-to-moderate radiation damage state of zircon CAM18 [31]. The α flux (Dα), calculated using U and Th concentrations and the U–Pb age (about 571 Ma) of the sample, represents the number of α-decay events per gram, and this value can be used to further quantify the radiative intensity since the time of closure of the U–Pb system [32]. The α flux of all analysis points (except for one analysis with a α flux of 1.68 × 1018 α/g) is between 1.34 and 1.43 × 1018 α/g, which may be related to the heterogeneity of radiation damage and still needs to be experimentally verified by more fragments (Table A1).
The thermal annealing caused by heating events can eliminate partial radiation damage. The density and Raman band broadening of the sample are consistent with the calculated α flux result, suggesting that the lattice damage is at a mild-to-moderate level. Thus, the occurrence of heat treatment can be ruled out.

4.2. U–Pb Geochronology

The variation in 206Pb/238U ages between different CAM18 zircon grains is small, with a measurement repeatability of 0.46% (RSD), within the uncertainty of the age accuracy obtained by LA-ICP-MS (about 1~2%, 2s) [33]. The weighted average 206Pb/238U and 207Pb/235U ages are 571.0 ± 3.0 Ma (2s, n = 20) and 573.4 ± 6.0 Ma (2s, n = 20), respectively, and both are in agreement within analytical uncertainty. This consistency suggests that no significant loss of radiogenic Pb has occurred in the U–Pb system since the zircon formation. Since the 235U signal is calculated from 238U, potential biases may be introduced when calculating the 207Pb/235U ratio. The measured 238U/206Pb and 207Pb/206Pb ratios are plotted on a Tera–Wasserburg plot, yielding a lower-intercept age of 570.5 ± 4.0 Ma (2s). The lower-intercept age is consistent with the weighted average 206Pb/238U age within error, indicating that these U–Pb data have high reliability.
From the age results of the sample, the moderate radiation damage generated by the α-decay events of U does not cause any significant loss of radiogenic Pb. Zircon reference materials 91500, TEMORA-1 and SA01 have average U concentrations of 80, 195 and 161 ppm, respectively. In contrast, a particularly advantageous property of zircon CAM18 is the high U mass fraction of 672 ppm, which produces high count rates and good counting statistics for LA-ICP-MS analyses (Table 2). Noticeable in this dataset, however, is the large ablation spot size that could exacerbate elemental fractionation and matrix effects, reducing the accuracy of the dating results, and it is difficult to perform effective analyses on natural zircons with complicated internal structures or small particles [34,35]. As a result, considering the effect of different beam spot diameters on the accuracy of U–Pb ages by LA-ICP-MS, it is essential to carry out dating analyses with small laser beam diameters in the future. Additionally, the isotope dilution thermal ionization mass spectrometry (ID-TIMS) technique has a higher analytical accuracy (0.1%) compared to the LA-ICP-MS technique [36]. To confirm the absolute accuracy of the LA-ICP-MS age, the ID-TIMS technique is also required for this sample to further study the distributed fragments.

4.3. Magmatic Genesis

The Th/U ratio is usually used as a criterion for distinguishing magmatic zircon (>0.4) and metamorphic zircon (<0.1), but a small number of magma zircons have extremely low Th/U ratios [41,42]. For example, magmatic zircons from Concordia and Kweekfontein granites exhibit high U (2060~4650 ppm) and low Th (80~380 ppm) concentrations, with a very low Th/U ratio of 0.06 [43]. The reason for this phenomenon is that these zircons crystallized from a melt enriched with U and are not related to the decoupling migration of Th and U in a metamorphic fluid. Therefore, it is necessary to combine this information with other evidence to interpret the origin of zircon CAM18.
Zircon CAM18 has a homogeneous Th/U ratio of 0.2 (Figure 9A). In addition, based on other magmatic characteristics, such as oscillatory zoning, the positive Ce anomaly, the negative Eu anomaly and the left-dipping REE pattern, it can be determined that the sample is of magmatic origin. The Th/U ratio in zircon is controlled by multiple factors, mainly including the element content in the melt, the distribution coefficients of elements between the melt and zircon, and the crystallization of Th-rich minerals [44]. In Figure 9B, the Th/U ratio shows a negligible change as the U content increases, implying that the Th and U content in the melt is not the main factor controlling the low Th/U ratio. The crystallization of Th-rich minerals (such as monazite) can consume the Th content in the melt, resulting in later crystallized zircon with low Th content. Therefore, the Th/U ratio of zircon CAM18 may be influenced by the crystallization of Th-rich minerals as well as the distribution coefficient of elements.

4.4. Magma Oxidization State

The chondrite-normalized REE pattern of zircon depends on the valence state and ion radius of the REE [45,46]. Compared with LREEs, HREEs have similar ionic radii to Zr4+, and they are easier to replace Zr4+ in the zircon lattice, so the REE distribution shows a left-dipping pattern. The Ce and Eu in magma exist in two valence states: Ce3+ (r = 1.14 Å), Ce4+ (r = 0.97 Å), Eu2+ (r = 1.25 Å) and Eu3+ (r = 1.07 Å) [47,48]. Since Ce4+ has the same charge and similar ion radius as Zr4+, it is more compatible in zircon than Ce3+, and, likewise, Eu3+ is more compatible in zircon than Eu2+. As the oxygen fugacity increases, the Ce3+ in magma is easily oxidized to Ce4+, which leads to Ce enrichment in zircon because the partition coefficient of Ce4+ is greater than that of La3+ and Pr3+. Similarly, the oxidized Eu3+ has better compatibility in zircon, so a small Eu depletion is observed in zircon. The Ce and Eu anomalies in magma zircon can be used as an indicator to characterize the relative oxygen fugacity during magmatic differentiation [46,49].
Trail et al. [50] derived a formula for calculating Ce and Eu anomalies based on the chondrite-normalized values, which is usually restricted by elemental detection limits and mineral inclusions. When LREE-rich inclusions (such as allanite, apatite, sphene and monazite) are also detected during laser ablation, the measured La and Pr content significantly increases. Except for the CAM18-8 and CAM18-19 analysis points, the La and Pr concentrations of other analysis points are higher than the detection limit of LA-ICP-MS and less than 1~200 × 10−3 and 10~500 × 10−3 ppm, respectively, so the influence of LREE-enriched mineral inclusions can be excluded [51,52,53]. As can be seen in Figure 10A, there is no positive correlation between δCe and δEu, implying that oxygen fugacity is not the main factor affecting Ce and Eu anomalies. Two reasons may cause the negative Eu anomaly of the sample. First, the Eu depletion of zircon CAM18 is inherited from the host magma; second, there is extensive plagioclase fractional crystallization prior to or during zircon crystallization, which consumes Eu in the melt, resulting in a significant negative Eu anomaly in the zircon [51]. In order to determine the magmatic oxygen fugacity, the newly proposed zircon oxygen fugacity meter by Loucks et al. [54] is used in this paper to calculate logfO2. This method is not limited by crystallization temperature, pressure or parent magma composition, and only requires trace element concentrations (such as Ti, U and Ce) and U–Pb ages to obtain fO2 with a standard error of ±0.6 log units. The oxygen fugacity and temperature results are shown in Table A2. The temperature ranges from 714 to 742 °C, and the oxygen fugacity is below the fayalite–magnetite–quartz (FMQ) buffer, ranging from ΔFMQ−2.87 to ΔFMQ−3.17, with an average value of ΔFMQ−3.07 (Figure 10B). These results show that zircon CAM18 crystallizes in a reducing environment with low oxygen fugacity.

5. Conclusions

Integrated analyses from multiple methods reveal that zircon CAM18 is free of mineral inclusions and has a crystalline structure and uniform trace element concentrations. In addition, it has homogeneous U–Pb ages, within the uncertainty of the age accuracy obtained by LA-ICP-MS. The crystal may be a potential reference material for LA-ICP-MS U–Pb dating. However, we do not know how well the LA-ICP-MS age matches its real age. In the future, the ID-TIMS technique with high analytical accuracy will be used to examine the absolute accuracy of LA-ICP-MS age.

Author Contributions

Conceptualization, W.L.; methodology, B.X.; investigation, W.L., H.L. and Z.Z.; resources, B.X.; data curation, W.L.; writing—original draft, W.L.; writing—review and editing, W.L., Z.M. and Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (42222304 and 42073038) and the DDE Big Science Program.

Data Availability Statement

The data presented in this study are available within the article.

Acknowledgments

We thank the editor and reviewers for their constructive comments, which helped in improving our paper. This is the 19th contribution of B.X. for the National Mineral Rock and Fossil Specimens Resource Center.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. LA-ICP-MS U–Pb dating results.
Table A1. LA-ICP-MS U–Pb dating results.
204(Pb + Hg)Radiogenic Isotope RatiosIsotopic Ages (Ma)
(×1018 α/g)
207Pb/206Pb208Pb/232Th207Pb/235U206Pb/238Urho207Pb/235U206Pb/238U207Pb/206Pb208Pb/232Th
CAM18
10.530.05840.00070.02770.00040.73910.00930.09160.00060.49285625.55653.454627.85528.71.39
20.810.05880.00070.02780.00040.75170.00960.09260.00060.48955695.65713.456127.85558.51.37
30.580.05840.00070.02780.00040.74960.00920.09290.00060.51575685.35723.554625.95548.81.35
40.530.05850.00070.02780.00040.74440.00880.09210.00050.48945655.15683.255024.15557.61.40
5bd (1)0.05910.00060.02790.00040.75360.00870.09220.00050.47195705.15693.057224.15568.31.39
60.230.05800.00070.02800.00040.74110.00900.09250.00050.48075635.35703.253230.65588.51.34
70.620.05970.00070.02830.00040.76400.00930.09260.00060.49605765.45713.359425.95648.81.38
80.250.05850.00070.02820.00050.74650.00910.09230.00060.49355665.35693.355025.95629.01.40
90.440.05980.00070.02860.00040.76840.00960.09310.00060.48895795.55743.459425.95718.61.43
10bd (1)0.06000.00080.02830.00050.76150.01010.09190.00060.46065755.85673.360627.85649.51.38
110.340.06070.00070.02890.00040.77440.00930.09240.00060.51265825.45703.462821.35768.81.42
120.130.05910.00070.02870.00050.75670.00870.09280.00060.55545725.05723.556925.95739.01.37
130.200.05960.00070.02910.00040.76260.00910.09270.00050.48735765.25713.258724.15798.41.36
140.400.05960.00060.02990.00040.76820.00860.09340.00050.51615794.95753.258722.25958.71.37
150.220.06000.00070.02930.00050.77330.00920.09340.00060.51315825.35753.460620.25849.21.37
16bd (1) 0.05980.00070.02940.00040.76670.00930.09280.00050.47695785.45723.259830.55868.11.68
170.590.05980.00060.02950.00050.76910.00830.09320.00060.55635794.85743.359424.15889.21.35
180.680.05900.00070.02880.00040.75480.00890.09260.00060.52375715.25713.456924.15748.51.36
190.0200.06000.00080.02930.00050.77260.01020.09320.00060.48725815.85743.560627.85858.91.37
200.740.05940.00080.02990.00050.76240.01010.09290.00060.51335755.85723.758327.85959.31.34
91500
10.820.07630.00120.05630.00101.88850.03060.17920.00140.4876107710.810637.8110636.1110819.7
20.380.07340.00170.05170.00131.81190.03990.17910.00150.3868105014.410628.4102641.7102024.9
30.230.07520.00120.05480.00101.85770.03020.17910.00140.4735106610.710627.6107331.5107819.1
40.350.07460.00140.05330.00141.84270.03540.17930.00140.4150106112.710637.8105738.6104927.1
50.290.07630.00150.05260.00131.88870.03810.17950.00160.4419107713.410648.8110340.7103624.1
60.220.07350.00130.05550.00111.81170.03070.17880.00140.4530105011.110617.5102835.2109220.7
GJ-1
10.370.05870.00080.03000.00100.79710.01120.09830.00060.46025956.46043.856729.659819.2
20.280.06080.00080.03070.00080.82480.01140.09830.00060.47606116.36043.863528.561116.2
30.510.06040.00090.03180.00100.81520.01160.09770.00060.45566056.56013.761730.463219.3
Table A2. Trace element abundance, Ce anomalies, and Eu anomalies in zircon and Ti-in-zircon temperature.
Table A2. Trace element abundance, Ce anomalies, and Eu anomalies in zircon and Ti-in-zircon temperature.
CAM
18-1
CAM
18-2
CAM
18-3
CAM
18-4
CAM
18-5
CAM
18-6
CAM
18-7
CAM
18-8
CAM
18-9
CAM
18-10
CAM
18-11
CAM
18-12
CAM
18-13
CAM
18-14
CAM
18-15
CAM
18-16
CAM
18-17
CAM
18-18
CAM
18-19
CAM
18-20
Ti
(ppm)
3.083.283.273.323.232.613.223.613.333.443.403.353.603.393.573.473.493.643.473.48
La
(ppm)
0.00490.00280.00140.00220.00410.00210.0028bd (1)0.00350.00300.00150.00280.00630.00420.00560.00010.00020.0008bd (1)0.0029
Ce
(ppm)
2.322.292.282.262.262.232.312.352.302.342.362.272.222.322.292.362.292.252.262.26
Pr
(ppm)
0.0270.0330.0290.0440.0310.0300.0310.0340.0300.0330.0290.0240.0320.0370.0360.0500.0290.0200.0290.032
Nd
(ppm)
0.510.570.530.600.600.570.530.550.590.510.550.550.540.510.600.590.520.500.560.51
Sm
(ppm)
0.990.920.810.900.951.051.031.040.920.951.040.870.970.910.910.990.840.960.900.95
Eu
(ppm)
0.180.170.170.190.180.160.170.170.210.180.180.190.210.170.160.190.180.160.160.19
Gd
(ppm)
3.333.393.403.423.393.463.463.333.373.353.353.203.233.173.453.483.503.383.463.20
Tb
(ppm)
0.980.900.890.900.930.910.900.900.910.910.910.820.890.870.850.940.870.940.900.88
Dy
(ppm)
9.148.588.478.869.178.879.129.049.108.758.958.458.998.638.928.978.589.098.678.83
Ho
(ppm)
3.023.032.903.062.953.052.923.033.093.013.112.993.012.942.983.102.952.993.002.98
Er
(ppm)
13.1413.0912.5713.0912.8112.8612.8412.7612.9312.6312.9612.7612.8512.7812.7713.4212.7612.8613.0612.58
Tm
(ppm)
2.702.622.632.682.672.682.702.612.702.662.672.622.642.632.702.722.672.672.662.57
Yb
(ppm)
24.8623.5423.4924.2424.6624.5624.4524.9624.7823.8624.7924.5823.8924.2824.5725.2224.1423.9724.5724.21
Lu
(ppm)
4.734.574.584.634.764.814.664.794.824.654.714.574.624.664.744.744.594.654.674.61
Hf
(ppm)
10,77810,77610,73810,79310,77410,78910,78310,72710,71610,75010,68910,75810,70810,69510,75510,67210,72110,76510,72410,731
total Pb
(ppm)
67.7067.7366.6169.0668.4465.9668.5369.0071.0968.0570.6468.0367.6468.4068.5671.2867.4567.3768.4866.41
Th
(ppm)
136135130137137129137138140137140135134132135142134134136133
U
(ppm)
677670658686678653675682696674695668665668670700660665671653
Th/U0.200.200.200.200.200.200.200.200.200.200.200.200.200.200.200.200.200.200.200.20
δCe
(2)
49.4558.4187.7356.3249.1568.8860.7950.0455.0357.6687.7367.8938.3345.6339.54258.77233.13137.9143.8757.52
δEu
(3)
0.300.290.310.330.310.260.280.280.360.310.290.350.360.310.280.310.320.270.280.33
ΔFMQ
(4)
−2.98−3.04−3.03−3.10−3.06−2.87−3.02−3.10−3.08−3.05−3.06−3.08−3.17−3.05−3.12−3.08−3.09−3.16−3.12−3.10
logfO2
(5)
−19.39−19.32−19.32−19.35−19.37−19.63−19.34−19.17−19.33−19.23−19.26−19.31−19.25−19.26−19.22−19.24−19.23−19.22−19.28−19.25
T
(°C) (6)
728734733735732714732742735738737735742736741738739743738739
(1) bd = below detection. (2) δ Ce = ( Ce La × Pr ) zircon ( Ce La × Pr ) chondrite . (3) δ Eu = ( Eu Sm × Gd ) zircon ( Eu Sm × Gd ) chondrite . (4) Δ FMQ = 3.998 ×   log [ Ce ( U i   ×   Ti ) z ] + 2.284 , wherein Ui denotes age-corrected initial U content, superscript z denotes zircon. (5) logfO 2 ( sample ) = logfO 2 ( FMQ ) + 3.998   ×   log [ Ce ( U i   ×   Ti ) z ] + 2.284 , wherein Ui denotes age-corrected initial U content, FMQ represents the reference buffer fayalite + magnetite + quartz, superscript z denotes zircon. (6) T ( ° C ) = [ 4800 + 0.4748   ×   ( P     1000 ) ] / ( logTi zir   5.711     loga TiO 2 melt + loga SiO 2 melt ) , where P = 300 MPa, a TiO 2 melt = 0.8 and a SiO 2 melt = 1 .

References

  1. Wu, Y.; Zheng, Y. Genesis of zircon and its constraints on interpretation of U–Pb age. Chin. Sci. Bull. 2004, 49, 1554–1569. [Google Scholar] [CrossRef]
  2. Hu, Z.; Li, X.H.; Luo, T.; Zhang, W.; Crowley, J.; Li, Q.; Ling, X.; Yang, C.; Li, Y.; Feng, L.; et al. Tanz zircon megacrysts: A new zircon reference material for the microbeam determination of U–Pb ages and Zr–O isotopes. J. Anal. At. Spectrom. 2021, 36, 2715–2734. [Google Scholar] [CrossRef]
  3. Chen, T.; Ai, H.; Yang, M.; Zheng, S.; Liu, Y. Brownish Red Zircon from Muling, China. Gems Gemol. 2011, 47, 36–41. [Google Scholar] [CrossRef]
  4. Xu, B.; Hou, Z.Q.; Griffin, W.L.; Zheng, Y.C.; Wang, T.; Guo, Z.; Hou, J.; Santosh, M.; O’Reilly, S.Y. Cenozoic lithospheric architecture and metallogenesis in Southeastern Tibet. Earth-Sci. Rev. 2021, 214, 103472. [Google Scholar] [CrossRef]
  5. Xu, B.; Hou, Z.Q.; Griffin, W.L.; Lu, Y.; Belousova, E.; Xu, J.F.; O’Reilly, S.Y. Recycled volatiles determine fertility of porphyry deposits in collisional settings. Am. Mineral. 2021, 106, 656–661. [Google Scholar] [CrossRef]
  6. Finch, R.J.; Hanchar, J.M. Structure and chemistry of zircon and zircon-group minerals. Rev. Mineral. Geochem. 2003, 53, 1–25. [Google Scholar] [CrossRef]
  7. Robinson, K.; Gibbs, G.V.; Ribbe, P.H. The structure of zircon: A comparison with garnet. Am. Mineral. J. Earth Planet. Mater. 1971, 56, 782–790. [Google Scholar]
  8. Xu, B.; Hou, Z.Q.; Griffin, W.L.; O’Reilly, S.Y. Apatite halogens and Sr–O and zircon Hf–O isotopes: Recycled volatiles in Jurassic porphyry ore systems in southern Tibet. Chem. Geol. 2022, 605, 120924. [Google Scholar] [CrossRef]
  9. Xu, B.; Hou, Z.-Q.; Griffin, W.L.; O’Reilly, S.Y.; Zheng, Y.C.; Wang, T.; Xu, J.F. In-situ mineralogical interpretation of the mantle geophysical signature of the Gangdese Cu-porphyry mineral system. Gondwana Res. 2022, 111, 53–63. [Google Scholar] [CrossRef]
  10. Sláma, J.; Košler, J.; Condon, D.J.; Crowley, J.L.; Gerdes, A.; Hanchar, J.M.; Horstwood, M.S.; Morris, G.A.; Nasdala, L.; Norberg, N.; et al. Plešovice zircon—A new natural reference material for U–Pb and Hf isotopic microanalysis. Chem. Geol. 2008, 249, 1–35. [Google Scholar] [CrossRef]
  11. Wiedenbeck, M.; Hanchar, J.M.; Peck, W.H.; Sylvester, P.; Valley, J.; Whitehouse, M.; Kronz, A.; Morishita, Y.; Nasdala, L.; Fiebig, J.; et al. Further characterisation of the 91500 zircon crystal. Geostand. Geoanal. Res. 2004, 28, 9–39. [Google Scholar] [CrossRef]
  12. Nasdala, L.; Hofmeister, W.; Norberg, N.; Martinson, J.M.; Corfu, F.; Dörr, W.; Kamo, S.L.; Kennedy, A.K.; Kronz, A.; Reiners, P.W.; et al. Zircon M257—A homogeneous natural reference material for the ion microprobe U–Pb analysis of zircon. Geostand. Geoanal. Res. 2008, 32, 247–265. [Google Scholar] [CrossRef]
  13. 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. Et Cosmochim. Acta 2006, 70, 158. [Google Scholar] [CrossRef]
  14. Black, L.P.; Kamo, S.L.; Allen, C.M.; Davis, D.W.; Aleinikoff, J.N.; Valley, J.W.; Mundil, R.; Campbell, I.P.; Korsch, R.J.; Williams, I.S.; et al. Improved 206Pb/238U microprobe geochronology by the monitoring of a trace-element-related matrix effect; SHRIMP, ID-TIMS, ELA-ICP-MS and oxygen isotope documentation for a series of zircon standards. Chem. Geol. 2004, 205, 115–140. [Google Scholar] [CrossRef]
  15. Frei, D.; Gerdes, A. Precise and accurate in situ U–Pb dating of zircon with high sample throughput by automated LA-SF-ICP-MS. Chem. Geol. 2009, 261, 261–270. [Google Scholar] [CrossRef]
  16. Vermeesch, P. IsoplotR: A free and open toolbox for geochronology. Geosci. Front. 2018, 9, 1479–1493. [Google Scholar] [CrossRef]
  17. Nasdala, L.; Zhang, M.; Kempe, U.; Panczer, G.; Gaft, M.; Andrut, M.; Plotze, M. Spectroscopic methods applied to zircon. Rev. Mineral. Geochem. 2003, 53, 427–467. [Google Scholar] [CrossRef]
  18. Dawson, P.; Hargreave, M.M.; Wilkinson, G.R. The vibrational spectrum of zircon (ZrSiO4). J. Phys. C Solid State Phys. 1971, 4, 240. [Google Scholar] [CrossRef]
  19. Woodhead, J.A.; Rossman, G.R.; Silver, L.T. The metamictization of zircon: Radiation dose-dependent structural characteristics. Am. Mineral. 1991, 76, 74–82. [Google Scholar]
  20. Huong, L.T.T.; Vuong, B.S.; Thuyet, N.T.M.; Nguyen, N.K.; Satitkune, S.; Wanthanachaisaeng, B.; Hofmeister, W.; Hager, T.; Hauzenberger, C. Geology, gemmological properties and preliminary heat treatment of gem-quality zircon from the central highlands of Vietnam. J. Gemmol. 2016, 35, 308. [Google Scholar] [CrossRef]
  21. Deliens, M.; Delhal, J.; Tarte, P. Metamictization and U–Pb systematics—A study by infrared absorption spectrometry of Precambrian zircons. Earth Planet. Sci. Lett. 1977, 33, 331–344. [Google Scholar] [CrossRef]
  22. Klinger, M.; Kempe, U.; Pöppl, A.; Bottcher, R.; Trinkler, M. Paramagnetic hole centres in natural zircon and zircon colouration. Eur. J. Mineral. 2012, 24, 1005–1016. [Google Scholar] [CrossRef]
  23. Kempe, U.; Trinkler, M.; Poppl, A.; Himcinschi, C. Coloration of natural zircon. Can. Mineral. 2016, 54, 635–660. [Google Scholar] [CrossRef]
  24. Cao, J.; Wen, G.; Ma, Z.; Yi, W.; Wang, F. Chemical composition and some spectroscopic characteristics of gem-quality zircon. Acta Mineral. Sin. 1988, 8, 111–118. [Google Scholar]
  25. Syme, R.W.G.; Lockwood, D.J.; Kerr, H.J. Raman spectrum of synthetic zircon (ZrSiO4) and thorite (ThSiO4). J. Phys. C Solid State Phys. 1977, 10, 1335. [Google Scholar] [CrossRef]
  26. Ai, H.; Zhao, X.; Chen, Q.; Long, Q.; Qin, J. Gemmological and Geochemical Characteristic of Zircon Megacryst Related to Cenozoic Alkaline Basalt from Eastern China. J. Gems Gemol. 2018, 20, 1–13. [Google Scholar]
  27. Kolesov, B.A.; Geiger, C.A.; Armbruster, T. The dynamic properties of zircon studied by single-crystal X-ray diffraction and Raman spectroscopy. Eur. J. Mineral. 2001, 13, 939–948. [Google Scholar] [CrossRef]
  28. Nasdala, L.; Irmer, G.; Wolf, D. The degree of metamictization in zircon: A Raman spectroscopic study. Eur. J. Mineral. 1995, 7, 471–478. [Google Scholar] [CrossRef]
  29. Rubatto, D.; Gebauer, D. Use of cathodoluminescence for U-Pb zircon dating by ion microprobe: Some examples from the Western Alps. In Cathodoluminescence in Geosciences; Springer: Berlin/Heidelberg, Germany, 2000; pp. 373–400. [Google Scholar]
  30. Zhang, B.L. Systematic Gemmology, 2nd ed.; Geology Press: Beijing, China, 2006; pp. 286–293. [Google Scholar]
  31. Wang, X.L.; Coble, M.A.; Valley, J.W.; Shu, X.J.; Kitajima, K.; Spicuzza, M.J.; Sun, T. Influence of radiation damage on Late Jurassic zircon from southern China: Evidence from in situ measurements of oxygen isotopes, laser Raman, U–Pb ages, and trace elements. Chem. Geol. 2014, 389, 122–136. [Google Scholar] [CrossRef]
  32. Nasdala, L.; Corfu, F.; Valley, J.W.; Spicuzza, M.J.; Wu, F.Y.; Li, Q.L.; Yang, Y.H.; Fisher, C.; Munker, C.; Kennedy, A.K.; et al. Zircon M127–A homogeneous reference material for SIMS U–Pb geochronology combined with hafnium, oxygen and, potentially, lithium isotope analysis. Geostand. Geoanal. Res. 2016, 40, 457–475. [Google Scholar] [CrossRef]
  33. Chang, Z.; Vervoort, J.D.; McClelland, W.C.; Knaack, C. U–Pb dating of zircon by LA-ICP-MS. Geochem. Geophys. Geosyst. 2006, 7, Q05009. [Google Scholar] [CrossRef]
  34. Kroslakova, I.; Günther, D. Elemental fractionation in laser ablation-inductively coupled plasma-mass spectrometry: Evidence for mass load induced matrix effects in the ICP during ablation of a silicate glass. J. Anal. At. Spectrom. 2007, 22, 51–62. [Google Scholar] [CrossRef]
  35. Koch, J.; Wälle, M.; Pisonero, J.; Günther, D. Performance characteristics of ultra-violet femtosecond laser ablation inductively coupled plasma mass spectrometry at∼265 and∼200 nm. J. Anal. At. Spectrom. 2006, 21, 932–940. [Google Scholar] [CrossRef]
  36. Schaltegger, U.; Ovtcharova, M.; Gaynor, S.P.; Schoene, B.; Wotzlaw, J.F.; Davies, J.F.; Farina, F.; Greber, N.D.; Szymanowski, D.; Chelle-Michou, C. Long-term repeatability and interlaboratory reproducibility of high-precision ID-TIMS U–Pb geochronology. J. Anal. At. Spectrom. 2021, 36, 1466–1477. [Google Scholar] [CrossRef]
  37. Wiedenbeck, M.; Alle, P.; Corfu, F.; Griffin, W.L.; Meier, M.; Oberli, F.; Von Quadt, A.; Roddick, J.C.; Spiegel, W. Three natural zircon standards for U–Th–Pb, Lu–Hf, trace element and REE analyses. Geostand. Newsl. 1995, 19, 1–23. [Google Scholar] [CrossRef]
  38. Black, L.P.; Kamo, S.L.; Allen, C.M.; Aleinikoff, J.N.; Davis, D.W.; Korsch, R.J.; Foudoulis, C. TEMORA 1: A new zircon standard for Phanerozoic U–Pb geochronology. Chem. Geol. 2003, 200, 155–170. [Google Scholar] [CrossRef]
  39. Huang, C.; Wang, H.; Yang, J.H.; Ramezani, J.; Yang, C.; Zhang, S.B.; Yang, Y.H.; Xia, X.P.; Feng, L.J.; Lin, J.; et al. SA01—A proposed zircon reference material for microbeam U-Pb age and Hf–O isotopic determination. Geostand. Geoanalytical Res. 2020, 44, 103–123. [Google Scholar] [CrossRef]
  40. Li, X.H.; Tang, G.Q.; Gong, B.; Yang, Y.H.; Hou, K.J.; Hu, Z.C.; Li, Q.L.; Liu, Y.; Li, W.X. Qinghu zircon: A working reference for microbeam analysis of U–Pb age and Hf and O isotopes. Chin. Sci. Bull. 2013, 58, 4647–4654. [Google Scholar] [CrossRef]
  41. Maas, R.; Kinny, P.D.; Williams, I.S.; Froude, D.O.; Compston, W. The Earth’s oldest known crust: A geochronological and geochemical study of 3900–4200 Ma old detrital zircons from Mt. Narryer and Jack Hills, Western Australia. Geochim. Et Cosmochim. Acta 1992, 56, 1281–1300. [Google Scholar] [CrossRef]
  42. Cui, L.; Peng, N.; Liu, Y.; Qiao, D.; Liu, Y. Detrital Zircon Geochronology and Tectonic Evolution Implication of the Middle Jurassic Zhiluo Formation, Southern Ordos Basin, China. Minerals 2023, 13, 45. [Google Scholar] [CrossRef]
  43. Clifford, T.N.; Barton, E.S.; Stern, R.A.; Duchesne, J.C. U–Pb zircon calendar for Namaquan (Grenville) crustal events in the granulite-facies terrane of the O’okiep copper district of South Africa. J. Petrol. 2004, 45, 669–691. [Google Scholar] [CrossRef]
  44. Harley, S.L.; Kelly, N.M.; Moller, A. Zircon behaviour and the thermal histories of mountain chains. Elements 2007, 3, 25–30. [Google Scholar] [CrossRef]
  45. Hinton, R.W.; Upton, B.G. The chemistry of zircon: Variations within and between large crystals from syenite and alkali basalt xenoliths. Geochim. Et Cosmochim. Acta 1991, 55, 3287–3302. [Google Scholar] [CrossRef]
  46. 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]
  47. Guo, J.; O’Reilly, S.Y.; Griffin, W.L. Zircon inclusions in corundum megacrysts: I. Trace element geochemistry and clues to the origin of corundum megacrysts in alkali basalts. Geochim. Et Cosmochim. Acta 1996, 60, 2347–2363. [Google Scholar] [CrossRef]
  48. Belousova, E.A.; Griffin, W.L.; O’Reilly, S.Y.; Fisher, N.L. Igneous zircon: Trace element composition as an indicator of source rock type. Contrib. Mineral. Petrol. 2002, 143, 602–622. [Google Scholar] [CrossRef]
  49. Wang, F.; Liu, S.A.; Li, S.; He, Y. Contrasting zircon Hf–O isotopes and trace elements between ore-bearing and ore-barren adakitic rocks in central-eastern China: Implications for genetic relation to Cu–Au mineralization. Lithos 2013, 156, 97–111. [Google Scholar] [CrossRef]
  50. Trail, D.; Watson, E.B.; Tailby, N.D. The oxidation state of Hadean magmas and implications for early Earth’s atmosphere. Nature 2011, 480, 79–82. [Google Scholar] [CrossRef]
  51. Dilles, J.H.; Kent, A.J.; Wooden, J.L.; Tosdal, R.M.; Koleszar, A.; Lee, R.G.; Farmer, L.P. Zircon compositional evidence for sulfur-degassing from ore-forming arc magmas. Econ. Geol. 2015, 110, 241–251. [Google Scholar] [CrossRef]
  52. Claiborne, L.L.; Miller, C.F.; Wooden, J.L. Trace element composition of igneous zircon: A thermal and compositional record of the accumulation and evolution of a large silicic batholith, Spirit Mountain, Nevada. Contrib. Mineral. Petrol. 2010, 160, 511–531. [Google Scholar] [CrossRef]
  53. Shen, P.; Hattori, K.; Pan, H.; Jackson, S.; Seitmuratova, E. Oxidation condition and metal fertility of granitic magmas: Zircon trace-element data from porphyry Cu deposits in the Central Asian orogenic belt. Econ. Geol. 2015, 110, 1861–1878. [Google Scholar] [CrossRef]
  54. Loucks, R.R.; Fiorentini, M.L.; Henríquez, G.J. New magmatic oxybarometer using trace elements in zircon. J. Petrol. 2020, 61, egaa034. [Google Scholar] [CrossRef]
  55. Trail, D.; Watson, E.B.; Tailby, N.D. Ce and Eu anomalies in zircon as proxies for the oxidation state of magmas. Geochim. Et Cosmochim. Acta 2012, 97, 70–87. [Google Scholar] [CrossRef]
Figure 1. Zircon crystal CAM18.
Figure 1. Zircon crystal CAM18.
Crystals 13 01364 g001
Figure 2. The micrographs of zircon CAM18. (A) Ghosting phenomenon; (B) conchoidal fracture.
Figure 2. The micrographs of zircon CAM18. (A) Ghosting phenomenon; (B) conchoidal fracture.
Crystals 13 01364 g002
Figure 3. FTIR spectrum of zircon CAM18.
Figure 3. FTIR spectrum of zircon CAM18.
Crystals 13 01364 g003
Figure 4. UV–vis–NIR spectra of zircon CAM18.
Figure 4. UV–vis–NIR spectra of zircon CAM18.
Crystals 13 01364 g004
Figure 5. Raman spectra of zircon CAM18.
Figure 5. Raman spectra of zircon CAM18.
Crystals 13 01364 g005
Figure 6. Representative cathodoluminescence image of zircon fragments. Shown are the 206Pb/238U ages with a 1σ uncertainty. The red circles represent the analyzed spots and the value under each fragment corresponds to analytical data in Table A1.
Figure 6. Representative cathodoluminescence image of zircon fragments. Shown are the 206Pb/238U ages with a 1σ uncertainty. The red circles represent the analyzed spots and the value under each fragment corresponds to analytical data in Table A1.
Crystals 13 01364 g006
Figure 7. LA-ICP-MS U–Pb ages of zircon CAM18. (A) The concordia diagram; (B) the Tera–Wasserburg diagram; (C) the weighted average 206Pb/238U age plot. Error ellipses and error bars are 1σ.
Figure 7. LA-ICP-MS U–Pb ages of zircon CAM18. (A) The concordia diagram; (B) the Tera–Wasserburg diagram; (C) the weighted average 206Pb/238U age plot. Error ellipses and error bars are 1σ.
Crystals 13 01364 g007
Figure 8. Chondrite-normalized REE pattern of zircon CAM18.
Figure 8. Chondrite-normalized REE pattern of zircon CAM18.
Crystals 13 01364 g008
Figure 9. Th and U content and Th/U ratios of zircon CAM18. (A) Th vs. U; (B) Th/U vs. U. The black arrow indicates the trend of correlation.
Figure 9. Th and U content and Th/U ratios of zircon CAM18. (A) Th vs. U; (B) Th/U vs. U. The black arrow indicates the trend of correlation.
Crystals 13 01364 g009
Figure 10. (A) Zircon δCe vs. δEu; (B) logfO2 vs. T (modified from Trail et al., 2012) [55]. The black arrow indicates the trend of correlation.
Figure 10. (A) Zircon δCe vs. δEu; (B) logfO2 vs. T (modified from Trail et al., 2012) [55]. The black arrow indicates the trend of correlation.
Crystals 13 01364 g010
Table 1. Typical operation conditions for LA-ICP-MS analysis in Milma Lab.
Table 1. Typical operation conditions for LA-ICP-MS analysis in Milma Lab.
ICP-MS ConditionsDwell Time
RF power1350 W204(Pb + Hg)20 ms
Feedback power<5 W206Pb20 ms
RF matching1.37 V207Pb30 ms
Sampling depth7.0 mm208Pb15 ms
Plasma gas15 L/min232Th10 ms
Auxiliary gas1 L/min238U15 ms
Make-up gas (Ar)0.8–1 L/minOther elements6 ms
Laser Parameters
Wavelength193 nm
Pulse duration5 ns
Ablation styleSingle spot
Energy density4–8 J/cm2
Carrier gas (He)800–900 mL/min
Ablation spot size35 μm
Repetition rate6–8 Hz
Table 2. Summary of U–Pb age data for zircon CAM18 and other zircon reference materials.
Table 2. Summary of U–Pb age data for zircon CAM18 and other zircon reference materials.
Reference MaterialAverage U Concentration (ppm)Age (Ma, 2s 1)Reference
91500801065.4 ± 0.6[11,37]
TEMORA-1195416.8 ± 0.2[38]
SA01161535.1 ± 0.4[39]
M127896524.4 ± 0.2[32]
Qinghu1042159.5 ± 0.2[40]
CAM18672571.4 ± 3.0This study
1 2s means 2 sigma uncertainties.
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

Li, W.; Xu, B.; Miao, Z.; Zhao, Z.; Liu, H. Exploring the CAM18 Crystal as a Potential Reference Material for U–Pb Analysis of Zircon. Crystals 2023, 13, 1364. https://doi.org/10.3390/cryst13091364

AMA Style

Li W, Xu B, Miao Z, Zhao Z, Liu H. Exploring the CAM18 Crystal as a Potential Reference Material for U–Pb Analysis of Zircon. Crystals. 2023; 13(9):1364. https://doi.org/10.3390/cryst13091364

Chicago/Turabian Style

Li, Wurui, Bo Xu, Zhuang Miao, Zheyi Zhao, and Hangyu Liu. 2023. "Exploring the CAM18 Crystal as a Potential Reference Material for U–Pb Analysis of Zircon" Crystals 13, no. 9: 1364. https://doi.org/10.3390/cryst13091364

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