[go: up one dir, main page]

Next Issue
Volume 14, August
Previous Issue
Volume 14, June
 
 

Minerals, Volume 14, Issue 7 (July 2024) – 103 articles

Cover Story (view full-size image): Volcanic millstones are common in archaeological sites. Their characterisation is performed petrographically, but their provenance is often studied by comparing geochemical data from reference outcrops using biplots or multivariate analysis. In this paper, volcanic stones from two sites (Iulia Libica, Spain, and Sidi Zahruni, Tunisia) are studied. New strategies to quantify petrographic elements and to determine the provenance of the tools are presented. In particular, supervised machine learning has been tested using the GEOROC database as a source of labelled data. The Tunisian materials have been identified as basalts from Pantelleria (Italy). In contrast, the provenance of the Spanish materials remained unknown. The results illustrate the advantages and limitations of all the examined methods. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
21 pages, 15719 KiB  
Article
Lithium-Rich Deposits in the Liangshan Formation during the Permian in the Upper Yangtze Plate, China
by Yin Gong, Chun-Yao Liu, Yin Liu, Li Lei, Meng Xiang, Bo Yang, Zhou Zhou, Yang Zhang, Xiang-Rong Yang, Lei Yan and Yuan Xiong
Minerals 2024, 14(7), 735; https://doi.org/10.3390/min14070735 - 22 Jul 2024
Viewed by 403
Abstract
With the increasing demand for lithium (Li) resources in industry, there has been new attention on clay-type lithium-rich deposits recently. In this study, a Li-rich clay deposit with a Li2O content up to 0.3% in the Liangshan Formation in the upper [...] Read more.
With the increasing demand for lithium (Li) resources in industry, there has been new attention on clay-type lithium-rich deposits recently. In this study, a Li-rich clay deposit with a Li2O content up to 0.3% in the Liangshan Formation in the upper Yangtze, South China Block was demonstrated. We analysed the mineralogy and element geochemistry of the samples from the Liangshan Formation and its underlying and overlying layers. Kaolinite (average 53%, up to 93%) was the major mineral in the samples from the Liangshan Formation. The Li concentrations increased with increasing kaolinite compositions and Al2O3 concentrations. Furthermore, based on the geochemical indicators, it was suggested that the clay formation and Li enrichment were related to the weathering processes of the bottom impure limestone under the hot and wet climate, and the sedimentary processes in the anoxic, still, and flat land–sea interaction area in the Upper Yangtze. The Li was probably sourced from the bottom impure limestone during the weathering stage. The samples from the Liangshan Formation also showed REE enrichment from 117 to 729 μg/g. Full article
Show Figures

Figure 1

Figure 1
<p>The reconstructed palaeographic maps of the South China Block during the late Carboniferous and early Permian. The subfigure (<b>A</b>) is modified from [<a href="#B15-minerals-14-00735" class="html-bibr">15</a>], and (<b>B</b>–<b>E</b>) are modified from [<a href="#B16-minerals-14-00735" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>The sampling maps and field photos. Subfigure (<b>B</b>) is the Li-rich deposit area, and the details of formation symbols (e.g., P<sub>2</sub>l) are described in the supplementary material (<a href="#app1-minerals-14-00735" class="html-app">Table S1</a>). In Subfigure B, the black lines are stratigraphic boundaries and the red lines represent structural lines. The subfigures (<b>A</b>,<b>C</b>,<b>D</b>) are the maps and field photos of sampling points trail trench TC08, drill core ZK1402, and trail trench BT06, respectively. The legends for subfigure (<b>A</b>,<b>C</b>,<b>D</b>) are at the left bottom. The yellow circles in the field photos are sample IDs and the red lines in the field photos are the boundary between the layers.</p>
Full article ">Figure 3
<p>The image of the thin section under the polarizing microscope and XRD patterns of sample BT06-13 from the Liangshan Formation. The image of the thin section was observed through cross-polarized light. (<b>A</b>) The image of the thin section of sample BT06-13 under the polarizing microscope. (<b>B</b>) The XRD patterns of sample BT06-13.</p>
Full article ">Figure 4
<p>The geochemistry of drill core ZK1402. The red dots are the sample ZK1402-01 to 07 from up to bottom.</p>
Full article ">Figure 5
<p>The geochemistry of trail trench BT06. The red dots are the samples BT06-01 to 17 from bottom to up. The samples from the bottom (BT06-01 from the Huanglong Formation) and top (BT06-17 from the Qixia Formation) were not shown (see the data in <a href="#minerals-14-00735-t002" class="html-table">Table 2</a>, <a href="#minerals-14-00735-t003" class="html-table">Table 3</a> and <a href="#minerals-14-00735-t004" class="html-table">Table 4</a>).</p>
Full article ">Figure 6
<p>The geochemistry of trail trench TC08.</p>
Full article ">Figure 7
<p>The plots of Zr versus Hf (<b>A</b>) and Ta versus Nb (<b>B</b>). The black dash lines are the best-fit lines, and the grey dash lines represent the boundaries of the fits with <span class="html-italic">p</span> less than 0.05.</p>
Full article ">Figure 8
<p>The relationship between Hinckley index (H.I.) and Fe<sub>2</sub>O<sub>3</sub> content (<b>A</b>) and Li concentration (<b>B</b>). The H.I. calculation method is shown in <a href="#sec3dot1-minerals-14-00735" class="html-sec">Section 3.1</a> and <a href="#app1-minerals-14-00735" class="html-app">Figure S1</a>. Samples BT06-09 to 13, TC08-01 to 04, and ZK1402-02 to 05 are shown.</p>
Full article ">Figure 9
<p>The relationships between Li and Al<sub>2</sub>O<sub>3</sub> composition (<b>A</b>), CIA (<b>B</b>), and kaolinite composition (<b>C</b>).</p>
Full article ">Figure 10
<p>The influences of the first two components on variables based on PCA. The details on the PCA method are in the <a href="#app1-minerals-14-00735" class="html-app">Supplementary Material</a>.</p>
Full article ">Figure 11
<p>The normalized REE patterns of trail trenches BT06 (<b>A</b>) and TC08 (<b>B</b>) and drill core ZK1402 (<b>C</b>).</p>
Full article ">
27 pages, 380 KiB  
Review
Recent Uses of Ionic Liquids in the Recovery and Utilization of Rare Earth Elements
by Francisco Jose Alguacil, Jose Ignacio Robla and Olga Rodriguez Largo
Minerals 2024, 14(7), 734; https://doi.org/10.3390/min14070734 - 22 Jul 2024
Viewed by 487
Abstract
The importance of rare earth elements as a basis for the development of new technologies or the improvement of existing ones makes their recovery from raw and waste materials necessary. In this recovery, hydrometallurgy and its derivative solvometallurgy play key roles due to [...] Read more.
The importance of rare earth elements as a basis for the development of new technologies or the improvement of existing ones makes their recovery from raw and waste materials necessary. In this recovery, hydrometallurgy and its derivative solvometallurgy play key roles due to their operational characteristics, which are emphasized with the use of ionic liquids. This manuscript reviews the most recent advances (2023 and 2024) in the use of ionic liquids in unit operations (leaching and separation technologies) aimed at the recovery of these valuable and strategic metals. Moreover, a comprehensive review is presented of the use of these chemicals in the development of advanced materials containing some of these rare earth elements. Full article
21 pages, 6293 KiB  
Article
The Formation Age and Magma Source of the Xiaonanshan–Tunaobao Cu-Ni-PGE Deposit in the Northern Margin of the North China Craton
by Guanlin Bai, Jiangang Jiao, Xiaotong Zheng, Yunfei Ma and Chao Gao
Minerals 2024, 14(7), 733; https://doi.org/10.3390/min14070733 - 22 Jul 2024
Viewed by 410
Abstract
The Xiaonanshan–Tunaobao Cu-Ni-PGE deposit is located in the northern margin of the North China Craton (N-NCC) in central Inner Mongolia. However, the age, magma source, petrogenesis, and sulfide mineralization mechanism of the ore-related Xiaonanshan-Tunaobao pluton remain unclear. Zircon U-Pb dating indicates the Tunaobao [...] Read more.
The Xiaonanshan–Tunaobao Cu-Ni-PGE deposit is located in the northern margin of the North China Craton (N-NCC) in central Inner Mongolia. However, the age, magma source, petrogenesis, and sulfide mineralization mechanism of the ore-related Xiaonanshan-Tunaobao pluton remain unclear. Zircon U-Pb dating indicates the Tunaobao pluton formed at 275.9 ± 2.8 Ma (Early Permian), similar to the Xiaonanshan pluton (272.7 ± 2.9 Ma). The ore-related gabbro is enriched in LREE and LILE (e.g., Rb) and depleted in HREE and HFSE (e.g., Nb and Ti). It likely originated from enriched mantle metasomatized by subduction fluids, supported by enriched Hf-Nd isotopes (–34.34 to –6.16 for zircon εHf(t) and –7.24 to –5.92 for whole-rock εNd(t) values) and high Ba/La but low Rb/Y ratios. The δ34S values of the Xiaonanshan sulfides range from 4.5‰ to 11.4‰, indicating a mantle origin with contribution from surrounding rocks. Combining previous recognition with this study, we propose that the Xiaonanshan–Tunaobao pluton formed in a post-collision extensional setting. Full article
(This article belongs to the Special Issue Mineral Resources in North China Craton)
Show Figures

Figure 1

Figure 1
<p>Division of geological tectonic units in the central part of Inner Mongolia [<a href="#B10-minerals-14-00733" class="html-bibr">10</a>,<a href="#B11-minerals-14-00733" class="html-bibr">11</a>,<a href="#B13-minerals-14-00733" class="html-bibr">13</a>,<a href="#B15-minerals-14-00733" class="html-bibr">15</a>,<a href="#B17-minerals-14-00733" class="html-bibr">17</a>] (<b>a</b>) [<a href="#B13-minerals-14-00733" class="html-bibr">13</a>] and geological sketch map for the XNS-TNB area (<b>b</b>) [<a href="#B18-minerals-14-00733" class="html-bibr">18</a>].</p>
Full article ">Figure 2
<p>Geological exploration line for the Xiaonanshan–Tunaobao deposit [<a href="#B18-minerals-14-00733" class="html-bibr">18</a>]. (<b>a</b>) Geological sketch map of the Xiaonanshan deposit; (<b>b</b>) geological sketch map of the Tunaobao deposit; (<b>c</b>) exploration line for the Xiaonanshan deposit; (<b>d</b>) exploration line for the Tunaobao deposit.</p>
Full article ">Figure 3
<p>Representative photographs and photomicrographs showing major mineral assemblages and sulfide mineralization in the Xiaonanshan (<b>a</b>–<b>c</b>) and Tunaobao gabbro (<b>d</b>–<b>f</b>). Abbreviations: Ccp—chalcopyrite; Cpx—clinopyroxene; Ilm—ilmenite; Opx—orthopyroxene; Pn—pentlandite; Po—pyrrhotite; Py—pyrite; Srp—serpentine.</p>
Full article ">Figure 4
<p>Cathode luminescence photos of zircon grains from the Tunaobao pluton.</p>
Full article ">Figure 5
<p>Concordia (<b>a</b>) and weighted mean (<b>b</b>) diagrams of zircon U-Pb dating of the Tunaobao pluton.</p>
Full article ">Figure 6
<p>Chondrite-normalized REE patterns (<b>a</b>) and primitive mantle-normalized trace elements spider diagrams (<b>b</b>) of XNS-TNB gabbro. The data of chondrite and primitive mantle are from [<a href="#B44-minerals-14-00733" class="html-bibr">44</a>].</p>
Full article ">Figure 7
<p>Na<sub>2</sub>O + K<sub>2</sub>O vs. SiO<sub>2</sub> diagram (<b>a</b>) [<a href="#B46-minerals-14-00733" class="html-bibr">46</a>] and SiO<sub>2</sub> vs. K<sub>2</sub>O diagram (<b>b</b>) [<a href="#B47-minerals-14-00733" class="html-bibr">47</a>] of the XNS-TNB gabbro.</p>
Full article ">Figure 8
<p>Heatmap of immobile elements and geochemical data.</p>
Full article ">Figure 9
<p>Harker diagrams for the Xiaonanshan–Tunaobao plutons.</p>
Full article ">Figure 10
<p>(Th/Ta)<sub>PM</sub> vs. (La/Nb)<sub>PM</sub> diagram (<b>a</b>) and (Ta/Th)<sub>PM</sub> vs. (Th/Yb)<sub>PM</sub> diagram (<b>b</b>) of the Xiaonanshan–Tunaobao pluton [<a href="#B15-minerals-14-00733" class="html-bibr">15</a>,<a href="#B54-minerals-14-00733" class="html-bibr">54</a>].</p>
Full article ">Figure 11
<p>ε<sub>Hf</sub>(t) vs. <span class="html-italic">t</span> diagram of Tunaobao pluton (<b>a</b>) and ε<sub>Nd</sub>(t) vs. (<sup>87</sup>Sr/<sup>86</sup>Sr)<sub>i</sub> diagram of the Xiaonanshan–Tunaobao pluton (<b>b</b>) [<a href="#B57-minerals-14-00733" class="html-bibr">57</a>].</p>
Full article ">Figure 12
<p>Nb/Y vs. La/Yb diagram (<b>a</b>) and Ba/La vs. Th/Yb diagram (<b>b</b>) of the Xiaonanshan–Tunaobao pluton [<a href="#B60-minerals-14-00733" class="html-bibr">60</a>,<a href="#B62-minerals-14-00733" class="html-bibr">62</a>].</p>
Full article ">Figure 13
<p>δ<sup>34</sup>S of sulfides from the Xiaonanshan Cu-Ni-PGE deposit. δ<sup>34</sup>S of sulfides of other deposits are from the data published in previous studies [<a href="#B35-minerals-14-00733" class="html-bibr">35</a>,<a href="#B72-minerals-14-00733" class="html-bibr">72</a>,<a href="#B73-minerals-14-00733" class="html-bibr">73</a>,<a href="#B74-minerals-14-00733" class="html-bibr">74</a>,<a href="#B75-minerals-14-00733" class="html-bibr">75</a>,<a href="#B76-minerals-14-00733" class="html-bibr">76</a>,<a href="#B77-minerals-14-00733" class="html-bibr">77</a>,<a href="#B78-minerals-14-00733" class="html-bibr">78</a>].</p>
Full article ">
16 pages, 1241 KiB  
Article
Açaí Seed Biochar-Based Phosphate Fertilizers for Improving Soil Fertility and Mitigating Arsenic-Related Impacts from Gold Mining Tailings: Synthesis, Characterization, and Lettuce Growth Assessment
by Yan Nunes Dias, Wendel Valter da Silveira Pereira, Cecílio Frois Caldeira, Sílvio Junio Ramos, Edna Santos de Souza, Paula Godinho Ribeiro and Antonio Rodrigues Fernandes
Minerals 2024, 14(7), 732; https://doi.org/10.3390/min14070732 - 22 Jul 2024
Viewed by 357
Abstract
Biochar represents a promising alternative for enhancing substrates and remediating contaminants in mining areas. Given that arsenic (As) and phosphorus (P) share similar chemical forms, the combination of biochar and P fertilizers may reduce As uptake, thereby mitigating As-related impacts. This study aimed [...] Read more.
Biochar represents a promising alternative for enhancing substrates and remediating contaminants in mining areas. Given that arsenic (As) and phosphorus (P) share similar chemical forms, the combination of biochar and P fertilizers may reduce As uptake, thereby mitigating As-related impacts. This study aimed to evaluate the potential of biochar-based P fertilizers in improving soil fertility and mitigating human health risks from gold mining tailings in the eastern Brazilian Amazon. Biochar from açaí palm (Euterpe oleracea Mart.) seeds was produced through enrichment with single and triple superphosphate at a ratio of 1:4, at 400 °C, and applied to mining tailings at 0.5%, 1%, and 2%. After one year of incubation, lettuce plants were grown for 70 days. Biochar reduced As absorption by lettuce and improved biomass and nutrient accumulation, resulting in improved vegetation indices. Biochar was effective in reducing non-carcinogenic As risks via ingestion of soil and plants to acceptable levels. Regression equations explained the As absorption behavior as affected by the biochar and the importance of biochar-related nutrients in reducing As stress. This study demonstrates the potential of P-enriched biochar as an amendment for As-contaminated soils, reducing As absorption, increasing P availability, and improving plant growth. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Arsenic fractionation after biochar application.</p>
Full article ">Figure 2
<p>Carcinogenic and non-carcinogenic risk index for soil and plant ingestion. Different letters indicate a significant difference between treatments using Tukey’s test (<span class="html-italic">p</span> &lt; 0.05). Red line represents the levels of non-carcinogenic risk (1) and carcinogenic risk (0.0001).</p>
Full article ">
41 pages, 35191 KiB  
Review
Laser-Induced Breakdown Spectroscopy in Mineral Exploration and Ore Processing
by Russell S. Harmon
Minerals 2024, 14(7), 731; https://doi.org/10.3390/min14070731 - 22 Jul 2024
Viewed by 599
Abstract
Laser-induced breakdown spectroscopy (LIBS) is a type of optical emission spectroscopy capable of rapid, simultaneous multi-element analysis. LIBS is effective for the analysis of atmospheric gases, geological fluids, and a broad spectrum of minerals, rocks, sediments, and soils both in and outside the [...] Read more.
Laser-induced breakdown spectroscopy (LIBS) is a type of optical emission spectroscopy capable of rapid, simultaneous multi-element analysis. LIBS is effective for the analysis of atmospheric gases, geological fluids, and a broad spectrum of minerals, rocks, sediments, and soils both in and outside the traditional laboratory setting. With the recent introduction of commercial laboratory systems and handheld analyzers for use outside the laboratory for real-time in situ analysis in the field, LIBS is finding increasing application across the geosciences. This article first overviews the LIBS technique and then reviews its application in the domain of mineral exploration and ore processing, where LIBS offers some unique capabilities. Full article
Show Figures

Figure 1

Figure 1
<p>Example broadband LIBS spectrum between 200 and 800 nm for the Limica lepidolite [KLi<sub>2</sub>Al(Si,Al)<sub>4</sub>O<sub>10</sub>(F,OH)<sub>2</sub>] from the Grosmont Pegmatite District (Australia). Sample chemical composition can be determined from such LIBS spectra because the position of the emission lines is specific to each element present in a sample and its intensity is related to its concentration. Ar lines are present in the spectrum since this is the purge gas used by the handheld LIBS analyzer.</p>
Full article ">Figure 2
<p>Successive screen shots for handheld LIBS analysis of a spodumene [LiAl(SiO<sub>3</sub>)<sub>2</sub>] crystal from the Carolina Tin-Spodumene Belt (CTSB), North Carolina (USA) showing the elements identified in the real-time analysis. Values in green text are arbitrary relative abundances determined by the instrument software and the black text indicates the percentage of emission lines for each element contained in the on-board spectral library that are present in the acquired LIBS spectrum.</p>
Full article ">Figure 3
<p>Handheld LIBS analysis of vivianite [Fe<sub>3</sub>(PO<sub>4</sub>)<sub>2</sub>·8H<sub>2</sub>O] from the CTSB, North Carolina (USA), modified from [<a href="#B19-minerals-14-00731" class="html-bibr">19</a>]. (<b>a</b>) Photograph of secondary vivianite on a drill core fracture surface. (<b>b</b>) Broadband LIBS spectrum acquired using a handheld analyzer showing the prominent P peaks at 213.6, 214.9, and 255.3 nm and the suites of Fe lines between 240 and 280 nm and between 438 and 441 nm. The unlabeled peaks in the infrared portion of the of the LIBS spectrum between 700 and 870 nm result from the Ar purge gas used for the LIBS analysis. The very strong H peak at 653.6 nm indicates that this is a hydrated mineral, further supporting its identification as vivianite.</p>
Full article ">Figure 4
<p>Approximate LIBS detection limits (LODs) for solid materials, modified from [<a href="#B47-minerals-14-00731" class="html-bibr">47</a>]. These LODs are conservative estimates that depend on the material analyzed, the LIBS instrument used, and the specific conditions of the analysis.</p>
Full article ">Figure 5
<p>LIBS spectral pre-processing, modified from [<a href="#B63-minerals-14-00731" class="html-bibr">63</a>]. (<b>a</b>) Flowchart for the five pre-processing steps of LIBS spectra. (<b>b</b>) An example of smoothing, normalization, and baseline correction of a broadband LIBS spectrum.</p>
Full article ">Figure 6
<p>Examples of LIBS compositional imaging, modified from [<a href="#B76-minerals-14-00731" class="html-bibr">76</a>,<a href="#B77-minerals-14-00731" class="html-bibr">77</a>]. (<b>a</b>) LIBS spectrum between 285 and 335 nm (left), surface spatial distributions of Ca and Si over a 0.785 × 0.785 mm<sup>2</sup> area (center), and Si distributions on the surface and at 3 μm depth intervals in a sample acquired by LIBS analysis of the REE-rich mineral bastnäsite [(Ce, La)CO<sub>3</sub>F] by Chirinos et al. [<a href="#B76-minerals-14-00731" class="html-bibr">76</a>] as determined from the emission lines for Ca at 315.9 nm and Si at 288.2 nm. In this and subsequent figures, unless otherwise stated, warm colors (e.g., yellow and red) in the concentration distribution maps denote elevated elemental abundances whereas cool colors (e.g., blue and purple) indicate diminished elemental abundances. (<b>b</b>) Composite image of Fe distribution across the polished surface of a specimen obtained by Moncayo et al. [<a href="#B77-minerals-14-00731" class="html-bibr">77</a>] from merging individual concentration maps for its three constituent minerals—turquoise [CuAl<sub>6</sub>(PO<sub>4</sub>)<sub>4</sub>(OH)<sub>8</sub>·4H<sub>2</sub>O], pyrite [FeS<sub>2</sub>], and silica (mainly quartz). The intensity of the LIBS Fe emission line at 259.9 nm was recorded across the 468 mm<sup>2</sup> surface area of the sample at a step resolution of 15 μm. The image comprises more than 2 × 10<sup>6</sup> pixels (i.e., 1300 by 1600 pixels along the x- and y-axes, respectively).</p>
Full article ">Figure 7
<p>Optical image of the cut and polished surface a hydrothermal ore sample from the Tighza mine (Morocco), modified from [<a href="#B78-minerals-14-00731" class="html-bibr">78</a>]. (<b>a</b>) Optical image of the sample. (<b>b</b>) Typical single-pulse LIBS emission spectra covering the spectral range between 270 and 340 nm recorded for the five different minerals indicated in (<b>d</b>). (<b>c</b>) Multi-elemental LIBS image for the major elements of the sample, Pb (red), Cu (yellow), Zn (magenta), Ca (blue), and Si (gray), with the same scale as (<b>a</b>). (<b>d</b>) A 10× magnification of the region shown by the white rectangle in (<b>c</b>) showing the five different minerals analyzed (1 = sphalerite, 2 = chalcopyrite, 3 = galena, 4 = calcite, and 5 = silica).</p>
Full article ">Figure 8
<p>False-color image of mineralogy in a pegmatite from the Fregeneda–Almendra pegmatite field between Spain and Portugal in the Central Iberian Zone of western Europe acquired from kHz-µLIBS scanning, modified from [<a href="#B73-minerals-14-00731" class="html-bibr">73</a>]. Lpd = lepidolite mica, Qtz = quartz, Ab = albite [NaAlSi<sub>3</sub>O<sub>8</sub>], Kfs + Ms = K-feldspar [KAlSi<sub>3</sub>O<sub>8</sub>] plus muscovite mica [KAl<sub>2</sub>(Si<sub>3</sub>Al)O<sub>10</sub>(OH)<sub>2</sub>], Brl = beryl [Be<sub>3</sub>Al<sub>2</sub>Si<sub>6</sub>O<sub>18</sub>], AGM = amblygonite-group minerals, Cst = cassiterite [SnO<sub>2</sub>], CGM = Nb–Ta oxides/columbite-group minerals, Phos = phosphates, and Zrn = zircon [ZrSiO<sub>4</sub>].</p>
Full article ">Figure 9
<p>Handheld LIBS imaging of a mantle xenolith in a drill core from the Muskox kimberlite pipe near Nunavut (Canada), modified from [<a href="#B71-minerals-14-00731" class="html-bibr">71</a>]. (<b>a</b>) Core photograph with the black rectangle indicating the area of handheld LIBS analysis. (<b>b</b>–<b>d</b>) LIBS imaging results undertaken directly on the sawed core surface for the major element Ca using the 527.0 nm emission line, the minor element Na using the 589.0 nm emission line, and the trace element Li using the 670.8 nm emission line. The heat maps to the right of each set of elemental distribution maps indicate relative elemental concentrations as percent values for sum-normalized emission intensities. The elemental distribution maps for Ca, Na, and Li stitch together 20 (10 × 2) individual 16 × 16 shot grids (i.e., 5120 total laser shots). Qualitative LIBS concentrations suggest that Ca and, to a lesser extent, Li are concentrated along veinlets that are only a few pixels wide (i.e., ~100 s of μm) that comprise a mixture of calcite and clinopyroxene [Ca(Mg,Fe)Si<sub>2</sub>O<sub>6</sub>]. It would not have been possible to document the location and distribution of the trace element Li without such geochemical micro-imaging. The feature-of-interest maps, which highlight the distribution of veinlets, effectively separate clinopyroxene (Cpx), garnet (Gt), and olivine (Olv) based mostly on Ca emissions.</p>
Full article ">Figure 10
<p>Schema illustrating a post-processing approach to LIBS quantitative analysis using the calibration curve approach, modified from [<a href="#B88-minerals-14-00731" class="html-bibr">88</a>]. (<b>a</b>) LIBS spectrum, (<b>b</b>) magnitudes of the Ag emission line at 546.6 nm for different silver contents in the standards, and (<b>c</b>) regression calibration model based on principal components regression (PCR) and partial least squares regression (PLSR) model developed for quantitative silver analysis. Analytical figures of merit for assessing the performance of a calibration include the coefficient of determination (R<sup>2</sup>), the root mean square error (RSME), mean absolute error (MAE), and limit of quantification (LOQ).</p>
Full article ">Figure 11
<p>In-line measurement of drill dust in an active quarry by LIBS in real time, as modified from [<a href="#B100-minerals-14-00731" class="html-bibr">100</a>]. Unlike the downhole deployable LIBS instrument for geochemical analysis described in [<a href="#B99-minerals-14-00731" class="html-bibr">99</a>], this approach continuously analyzes the dust that is generated from rock drilling using the dust extraction hose connected to the LIBS analyzer. (<b>a</b>) LIBS system mounted on mobile drill rig in quarry. (<b>b</b>) Measured spatial variation of Al concentration (wt. %) as a function of drill depth and horizontal location of the drill rig. A depth resolution of a few centimeters is possible because the measurements are directly related to the depth in the deposit using the position monitor of the drill rig.</p>
Full article ">Figure 12
<p>LIBS publications related to mineral exploration and ore processing from 1990–2023.</p>
Full article ">Figure 13
<p>LIBS muscovite analysis in Li-pegmatites from southwestern North Carolina (USA) by handheld LIBS, modified from [<a href="#B129-minerals-14-00731" class="html-bibr">129</a>]. K/Rb versus Li plots for (<b>a</b>) the Carolina Tin-Spodumene Belt (CTSB), (<b>b</b>) the Carolina Lithium Prospect (CLP) within the CTSB, and (<b>c</b>) four CLP drill cores. Symbols containing an ‘x’ indicate spodumene-bearing pegmatites. According to Wise et al. [<a href="#B131-minerals-14-00731" class="html-bibr">131</a>], common pegmatites characterized by low degrees of fractionation typically exhibit K/Rb ratios &gt; 40, moderately fractionated pegmatites have K/Rb ratios between 40 and 10, and highly fractionated pegmatites have K/Rb ratios &lt; 10. The spodumene-bearing granite pegmatites of the CTSB and CLP are Li-poor but moderately to highly fractionated in this schema.</p>
Full article ">Figure 14
<p>LIBS analysis of a Li-bearing pegmatite in drill core through the metagreywacke of the Rapasaari complex (Finland), modified from figures in [<a href="#B61-minerals-14-00731" class="html-bibr">61</a>]. (<b>top</b>) Example LIBS spectra for the accessory minerals beryl, apatite [<sub>Ca5</sub>(PO<sub>4</sub>)<sub>3</sub>(OH,F,Cl)] and a Zn-bearing phase with specific element lines indicated. (<b>middle</b>) Optical image of muscovite pegmatite drill core BK204 with the area of analysis indicated by the red rectangle. (<b>bottom</b>) False-color LIBS-derived images for element distributions (left) and mineral classification results (right).</p>
Full article ">Figure 15
<p>Garnet analysis by handheld LIBS, modified from [<a href="#B81-minerals-14-00731" class="html-bibr">81</a>]. (<b>a</b>) PCA score plot for igneous garnets from granitic pegmatites and rhyolites, chromite pods and kimberlites by handheld LIBS. (<b>b</b>) PLSDA classification matrix for pyrope-type garnets from five South African kimberlite locations: KZNP = KwaZulu-Natal Province; NCP-BW = Barkly West, in Northern Cape Province; NCP-K = Kimberley in Northern Cape Province; NK = North of Kimberley in Northern Cape Province; and OFS = Orange Free State. This bivariate plot portrays the PLSDA reclassification of the LIBS observations, where each entry in the matrix indicates the percentage of samples that were identified as belonging to the column class when they are actually members of the row class.</p>
Full article ">Figure 16
<p>Automated LIBS system for real-time production line analysis, modified from [<a href="#B158-minerals-14-00731" class="html-bibr">158</a>]. (<b>left</b>) LIBS system installed and operating on a moving belt conveyer at an open-pit phosphate mine in Florida (USA). (<b>right</b>) LIBS measurements of MgO (■), Fe<sub>2</sub>O<sub>3</sub> (▲), percentage bone phosphate lime, BPL (▼), and metal equivalent ratio, MER (●) as functions of time during a continuous operation over five days. Two different rock compositions were observed as documented by the changing values of MgO, Fe<sub>2</sub>O<sub>3</sub>, BPL, and MER. Such real-time recognition of unsuitable material by LIBS enables its removal from the conveyer before ore processing.</p>
Full article ">Figure 17
<p>Sulfide and gangue mineral distribution in three drill core segments from LIBS analysis, modified from [<a href="#B91-minerals-14-00731" class="html-bibr">91</a>].</p>
Full article ">Figure 18
<p>Example LIBS spectra for the Cu-bearing mineral species pyrite, chalcopyrite, bornite, molybdenite, enargite, covellite, and chalcocite (modified from [<a href="#B59-minerals-14-00731" class="html-bibr">59</a>]).</p>
Full article ">Figure 19
<p>LIBS processing schema for primary (<b>top row</b>) and secondary (<b>bottom row</b>) Cu-sulfide ores from northern Chile, modified from [<a href="#B96-minerals-14-00731" class="html-bibr">96</a>]. (<b>left top and bottom</b>) Single-element images for Cu, S, Fe, Si, O, and Ca. (<b>center top and bottom</b>) Merged images for element combinations. (<b>right top and bottom</b>) Mineral images generated from element image combinations and sample optical images. anh = anhydrite, bn = bornite, cc = calcite, ccp = chalcopyrite, cv = covelite, mag = magnetite, qtz = quartz, pl = plagioclase feldspar.</p>
Full article ">Figure 20
<p>Principal component score plot for coltan samples from mines in California (USA)—El Molino, Ingram, Katerina and Olla, and from pegmatites in Connecticut (USA)—Branchville, Glastonbury, Haddam, Middletown, Portland and Strickland Quarry, modified from [<a href="#B87-minerals-14-00731" class="html-bibr">87</a>].</p>
Full article ">Figure 21
<p>Elemental and mineral mapping for a segment of drill core from Stillwater, Montana (USA), modified from [<a href="#B200-minerals-14-00731" class="html-bibr">200</a>]. (<b>a</b>) Spatial distribution of Si (gray), Mg (blue), Cu (orange), Fe (green), Ni (cyan), Pd (magenta), and Pt (yellow) with the minerals olivine, calcic plagioclase feldspar, chalcopyrite, pentlandite [(Fe,Ni)<sub>9</sub>S<sub>8</sub>], pyrrhotite, and braggite [(Pt, Pd, Ni)S] identified. (<b>b</b>) Spatial distribution of K (cyan), Ca (magenta), and Na (yellow) with the minerals olivine, bytownite plagioclase feldspar, and actinolite [Ca<sub>2</sub>(MgFe)Si<sub>8</sub>O<sub>22</sub>(OH)<sub>2</sub>] identified.</p>
Full article ">Figure 22
<p>Au distributions in cores from gold mines in Québec (Canada), modified from [<a href="#B58-minerals-14-00731" class="html-bibr">58</a>,<a href="#B200-minerals-14-00731" class="html-bibr">200</a>]. (<b>a</b>) Concentration versus area mapping over of 5 × 20 cm<sup>2</sup> domain of a drill core from the Lapa Mine [<a href="#B58-minerals-14-00731" class="html-bibr">58</a>]. The averaged Au concentration over the entire drill core is 46 ppm, although its actual distribution is highly spatially heterogeneous on a millimeter scale, with concentrations overall quite low but very high at a few discrete locations. (<b>b</b>) Volume segmentation of a Au-bearing sample from the Cadillac Mine reconstructed by LIBS tomography [<a href="#B200-minerals-14-00731" class="html-bibr">200</a>] showing the 3D spatial distribution of gold (yellow), chalcopyrite (red), and Al-rich mineral aggregates of biotite–chlorite [K(Mg,Fe)<sub>3</sub>AlSi<sub>3</sub>O<sub>10</sub>(F,OH)<sub>2</sub> -(Mg,Fe)<sub>3</sub>(Si,Al)<sub>4</sub>O<sub>10</sub>(OH)<sub>2</sub>·(Mg,Fe)<sub>3</sub>(OH)<sub>6</sub>] (blue). The volume without color mostly comprises tremolite [Ca<sub>2</sub>Mg<sub>5</sub>Si<sub>8</sub>O<sub>22</sub>(OH)<sub>2</sub>].</p>
Full article ">Figure 23
<p>LIBS provenance discrimination of cassiterite, modified from [<a href="#B80-minerals-14-00731" class="html-bibr">80</a>]. The PLSDA classification matrix was generated using 80 components and 10-fold cross-validation for 38 cassiterite samples from six different geographic regions. Numbers in parentheses indicate the numbers of samples analyzed from each cassiterite source and shaded boxes denote a classification success of &gt;90%.</p>
Full article ">Figure 24
<p>Elemental and molecular LIBS imaging of a cerite sample, modified from [<a href="#B219-minerals-14-00731" class="html-bibr">219</a>]. (<b>a</b>) Example single-shot elemental emission spectrum between 280 and 340 nm. (<b>b</b>) Example single-shot molecular emission spectrum between 460 and 650 nm; (<b>c</b>) Optical, elemental (Ca, Fe, Si, Cu, La and Y), molecular (LiO and YO), and combined false-color images.</p>
Full article ">Figure 25
<p>Optical image of a partially oxidized and hydrothermally altered ore sample from the Zálesí uranium deposit in the Rychlebské hory Mountains (Czech Republic) modified from [<a href="#B240-minerals-14-00731" class="html-bibr">240</a>]. The optical image (upper left) shows the surface texture of the sample and the areas examined by EDX imaging (red rectangle) to distinguish uraninite and LIBS imaging (green rectangle) used for elemental analysis, and the boundary (black line) between aggregates of primary uraninite—(Urn) and secondary uranophane—Urp [Ca(UO<sub>2</sub>)<sub>2</sub>(SiO<sub>3</sub>OH)<sub>2</sub>·5H<sub>2</sub>O]. The distributions of U, O, H, Si, and Ca in the sample are displayed in the five LIBS scans. Areas of secondary fluid alteration are indicated by elevated O, H, Si, and Ca abundances that generated secondary uranophane across the majority of the ore sample.</p>
Full article ">Figure 26
<p>LIBS-derived concentrations for Li, Si, and K of minerals in a drill core from the Rapasaari lithium deposit (Finland), modified from [<a href="#B105-minerals-14-00731" class="html-bibr">105</a>]. Calibration models were developed on a pixel-by-pixel basis for the core segments scanned by both LIBS and LA-ICP-MS.</p>
Full article ">Figure 27
<p>Selected elemental, molecular, and radiogenic isotope distributions in a uraninite sample from the Rožná-Olší ore field (Czech Republic), modified from [<a href="#B239-minerals-14-00731" class="html-bibr">239</a>]. (<b>top left</b>) Photographic images showing bands of uraninite (1), hydrothermal calcite (2), and metasomatite (3) along with the areas of the sample imaged by LIBS (gray rectangle) and LA-ICP-MS (white rectangles) analysis. (<b>lower left</b>) Distribution maps for Ca, Fe, Pb, U, and molecular UO in the three lithologies obtained by LIBS. (<b>top right</b>) Photographic image of spherulitic uraninite aggregates overgrown and/or crosscut by hydrothermal calcite (top left) and distributions of selected isotopes (<sup>43</sup>Ca, <sup>51</sup>V, <sup>235</sup>U, <sup>204</sup>Pb, <sup>206</sup>Pb, <sup>207</sup>Pb, <sup>208</sup>Pb) obtained by LA-ICP-MS. (<b>lower right</b>) Distribution of U in the sample domain shown in top left: (a) U by LIBS and (b–d) <sup>235</sup>U by LA-ICP-MS.</p>
Full article ">
14 pages, 6797 KiB  
Article
Telluride Mineralogy of the Kochbulak Epithermal Gold Deposit, Tien Shan, Eastern Uzbekistan
by Yongwei Lu, Xiaobo Zhao, Chunji Xue, Bakhtiar Nurtaev, Yiwei Shi, Yangtao Liu and Shukhrat Shukurov
Minerals 2024, 14(7), 730; https://doi.org/10.3390/min14070730 - 22 Jul 2024
Viewed by 332
Abstract
The Kochbulak gold deposit is situated on the northern slope of the Kurama range of eastern Uzbekistan and is one of the largest Tellurium-rich epithermal gold deposits in the world. Based on a detailed field and petrological investigation, three stages of mineralization can [...] Read more.
The Kochbulak gold deposit is situated on the northern slope of the Kurama range of eastern Uzbekistan and is one of the largest Tellurium-rich epithermal gold deposits in the world. Based on a detailed field and petrological investigation, three stages of mineralization can be classified, including, from early to late, quartz–pyrite vein stage, quartz–telluride–sulfide–sulphosalt–native gold stage, and pyrite–chalcopyrite vein stage. Abundant tellurides, including tellurobismuthite, rucklidgeite, tetradymite, altaite, volynskite, and hessite, have been well recognized in the second (main) mineralization stage. Based on the mineral assemblages and petrogenetic occurrence, the sequence of tellurides in the second mineralization stage can be approximately identified as altaite+calaverite+native tellurium, calaverite+native gold, Bi-telluride (e.g., tellurobismuthite and rucklidgeite)+petzite+native gold, Ag-Bi telluride (e.g., volynskite), and Ag-telluride (e.g., hessite)+native gold. By depicting the Log ƒ(Te2)-Log ƒ(S2) relationship diagram of the Kochbulak gold deposit under 250 °C and 200 °C, the Log ƒ(S2) value ranges from −14.7 to −8.6 and from −16.7 to −10.9, respectively, with Log ƒ(Te2) value varies from −12.3 to −7.8 under 250 °C and ranges from −13.8 to −11.2 under 200 °C. Full article
(This article belongs to the Special Issue Selenium, Tellurium and Precious Metal Mineralogy)
Show Figures

Figure 1

Figure 1
<p>Sketch tectonic map of the western Tien Shan showing locations of major gold deposits (modified from [<a href="#B6-minerals-14-00730" class="html-bibr">6</a>,<a href="#B11-minerals-14-00730" class="html-bibr">11</a>]).</p>
Full article ">Figure 2
<p>Geological map of the Kurama district (modified from [<a href="#B2-minerals-14-00730" class="html-bibr">2</a>]).</p>
Full article ">Figure 3
<p>(<b>a</b>) Geological map of the Kochbulak gold deposit. (<b>b</b>) Cross-section of the Kochbulak gold deposit (modified after [<a href="#B13-minerals-14-00730" class="html-bibr">13</a>]).</p>
Full article ">Figure 4
<p>(<b>a</b>) Typical hydrothermal gold ore of the Kochbulak deposit showing the crosscutting relationship of the three mineralization stages. (<b>b</b>) Miarolitic texture of quartz. (<b>c</b>) Crusty texture developed in the second (main) mineralization stage. (<b>d</b>) Breccias in the second (main) mineralization stage.</p>
Full article ">Figure 5
<p>BSE images of typical tellurides of the Kochbulak gold deposit. (<b>a</b>) Tellurobismuthite, pyrite, tetrahedrite, and chalcopyrite. (<b>b</b>) Rucklidgeite, pyrite, and chalcopyrite. (<b>c</b>) Tetradymite, pyrite, tetrahedrite, and galena. (<b>d</b>) Altaite, tetrahedrite, and chalcopyrite. (<b>e</b>) Altaite, tellurobismuthite, and volynskite. (<b>f</b>) Altaite, tellurobismuthite, hessite, and tetrahedrite. Py = pyrite, Ccp = chalcopyrite, Teb = tellurobismuthite, Td = tetrahedrite, Rkl = rucklidgeite, Gn = galena, Ttr = tetradymite, Alt = altaite, Vol = volynskite, Hes = hessite.</p>
Full article ">Figure 6
<p>Photomicrographs and paragenetic relationship of typical tellurides of the Kochbulak gold deposits. (<b>a</b>) Tellurobismuthite, altaite, tetrahedrite, and chalcopyrite. (<b>b</b>) Rucklidgeite, pyrite, and chalcopyrite. (<b>c</b>) Tellurobismuthite and volynskite in altaite. (<b>d</b>) Altaite, tetrahedrite, and chalcopyrite. (<b>e</b>) Altaite, tellurobismuthite, and hessite in tetrahedrite. (<b>f</b>) Altaite, tellurobismuthite, and volynskite in tetrahedrite. Py = pyrite, Ccp = chalcopyrite, Td = tetrahedrite, Rkl = rucklidgeite, Alt = altaite, Teb = tellurobismuthite, Vol = volynskite, Hes = hessite.</p>
Full article ">Figure 7
<p>Paragenetic sequence of the Te-rich Kochbulak gold deposit.</p>
Full article ">Figure 8
<p>The Log ƒ(Te<sub>2</sub>)-Log ƒ(S<sub>2</sub>) relationship diagram of the Kochbulak gold deposit calculated for 250 °C (modified after [<a href="#B4-minerals-14-00730" class="html-bibr">4</a>]). Abbreviations: Bn-bornite; Cp-chalcopyrite; Po-pyrrhotite; Py-pyrite.</p>
Full article ">Figure 9
<p>The Log ƒ(Te<sub>2</sub>)-Log ƒ(S<sub>2</sub>) relationship diagram of the Kochbulak gold deposit calculated for 200 °C (modified after [<a href="#B17-minerals-14-00730" class="html-bibr">17</a>]). Abbreviations: Bn-bornite; Cp-chalcopyrite; Fo-frohbergite; Po-pyrrhotite; Py-pyrite.</p>
Full article ">
12 pages, 5585 KiB  
Article
FIB-SEM Study of Archaeological Human Petrous Bones: 3D Structures and Diagenesis
by Jamal Ibrahim, Eugenia Mintz, Lior Regev, Dalit Regev, Ilan Gronau, Steve Weiner and Elisabetta Boaretto
Minerals 2024, 14(7), 729; https://doi.org/10.3390/min14070729 - 21 Jul 2024
Viewed by 470
Abstract
The petrous bone generally preserves ancient DNA better than other fossil bones. One reason for this is that the inner layer of the petrous bone of pigs and humans contains about three times as many osteocytes as other bones, and hence more DNA. [...] Read more.
The petrous bone generally preserves ancient DNA better than other fossil bones. One reason for this is that the inner layer of the petrous bone of pigs and humans contains about three times as many osteocytes as other bones, and hence more DNA. A FIB-SEM study of modern pig petrous bones showed that the 3D structure of the thin inner layer is typical of woven bone that forms in the fetus, whereas the thicker outer layer has a lamellar structure. The lamellar structure is common in mammalian bones. Here we study human petrous bones that are about 2500 years old, obtained from three Phoenician sites in Sicily, Italy. A detailed FIB-SEM study of two of these bones, one well preserved and the other poorly preserved, shows that the 3D bone type structure of the human petrous inner layer is woven bone, and the outer layer is lamellar bone. These are the same bone type structures found in pig petrous bones. Furthermore, by comparing nine differently preserved petrous bones from the same archaeological region and age, we show that their collagen contents vary widely, implying that organic material can be significantly altered during diagenesis. The mineral crystals are better preserved and hence less crystalline in the inner layers compared to the outer layers. We therefore infer that the best-preserved DNA in fossil petrous bones should be found in the thin inner layers immediately adjacent to the otic cavity where much more DNA is initially present and the mineral phase tends to be better preserved. Full article
(This article belongs to the Section Biomineralization and Biominerals)
Show Figures

Figure 1

Figure 1
<p>Plot of the SFs versus the full width at half maximum (FWMH) of the main peak of the mineral infrared spectrum of the inner (triangles) and outer (squares) layers of each of the 9 bones listed in <a href="#minerals-14-00729-t001" class="html-table">Table 1</a>. Dotted lines connect the different layers from the same petrous bone specimen. Samples selected for FIB-SEM imaging are indicated by sample number.</p>
Full article ">Figure 2
<p>FIB-SEM image of the outer layer obtained from one surface of the image stack from Sample 10610. (<b>A</b>) A 2D ESB image showing aligned lineations attributed mainly to the less-mineralized collagen fibrils. (<b>B</b>) A 3D image of the Eigenvector projections viewed in the same orientation as the image in (<b>A</b>). (<b>C</b>) A 3D image of the Eigenvector projections viewed top-down, i.e., orthogonal to the other views. The presence of only two colors shows that the collagen fibril bundles are oriented in two different directions within one plane. Vector directions parallel to the X-axis direction are blue, vectors parallel to the Y-axis are red, and vectors parallel to Z-axis are represented in green. Scale bar and grid size = 1 µm.</p>
Full article ">Figure 3
<p>A 3D representation of the FIB-SEM analysis using Eigenvectors for the outer layer of sample 10619. (<b>A</b>) A 2D ESB image of one of the surfaces. (<b>B</b>) Eigenvectors in 3D as viewed in the milling plane. <a href="#minerals-14-00729-f002" class="html-fig">Figure 2</a>B shows a 1µm thick layer where the collagen bundle follow the direction of the <span class="html-italic">X</span>-axis (blue), whereas the rest of the bundles follow the direction of the <span class="html-italic">Z</span>-axis (green). (<b>C</b>). A 3D image of the Eigenvector projections viewed top-down i.e. orthogonal to the other views. The color code for the EV directions is shown along each axis. Scale bar and grid size = 1 µm.</p>
Full article ">Figure 4
<p>The 2D and 3D images obtained from sample 10610 inner layer. (<b>A</b>) A 2D ESB image from the original milling plane. (<b>B</b>) A 3D image showing the Eigenvector from the same plane as in (<b>A</b>). The color code of the Eigenvectors shows the directions of the vectors which follows the axis colors, namely the x-axis direction is indicated by the blue color, the y-axis is indicated by red, and the z-axis is indicated by green. (<b>C</b>). A 3D image of the Eigenvector projections viewed top-down, i.e., orthogonal to the other views. Note the color mixtures of the EVs and the lack of layer organization. Note the random distribution of the colors across the image. Scale bar = 1 µm.</p>
Full article ">Figure 5
<p>The 2D and 3D images of the FIB-SEM analysis using Eigenvectors for sample 10619 inner layer. (<b>A</b>) A 2D image obtained from ESB detector. (<b>B</b>) A 3D image showing the Eigenvectors analysis at the same position as in (<b>A</b>). The color code for the Eigenvectors shows collagen fibril bundle directions as shown in the axis indicator (<span class="html-italic">x</span>-axis blue, <span class="html-italic">y</span>-axis red, and <span class="html-italic">z</span>-axis green). Note that at least 3 different directions can be discerned by eye. Scale bar = 1 µm.</p>
Full article ">
16 pages, 5131 KiB  
Article
Exploring the Effect of Particle Loading Density on Respirable Dust Classification by SEM-EDX
by Daniel Sweeney, Cigdem Keles and Emily Sarver
Minerals 2024, 14(7), 728; https://doi.org/10.3390/min14070728 - 20 Jul 2024
Viewed by 529
Abstract
Exposure to respirable coal mine dust (RCMD) still poses health risks to miners. Scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM-EDX) is a powerful tool for RCMD characterization because it provides particle-level data, including elemental ratios (via the EDX signals) that can enable [...] Read more.
Exposure to respirable coal mine dust (RCMD) still poses health risks to miners. Scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM-EDX) is a powerful tool for RCMD characterization because it provides particle-level data, including elemental ratios (via the EDX signals) that can enable classification by inferred mineralogy. However, if the particle loading density (PLD) is high on the analyzed substrate (filter sample), interference between neighboring particles could cause misclassification. To investigate this possibility, a two-part study was conducted. First, the effect of PLD on RCMD classification was isolated by comparing dust particles recovered from the same parent filters under both low- and high-PLD conditions, and a set of modified classification criteria were established to correct for high PLD. Second, the modified criteria were applied to RCMD particles on pairs of filters, with each pair having one filter that was analyzed directly (frequently high PLD) and another filter from which particles were recovered and redeposited prior to analysis (frequently lower PLD). It was expected that application of the modified criteria would improve the agreement between mineralogy distributions for paired filters; however, relatively little change was observed for most pairs. These results suggest that factors other than PLD, including particle agglomeration, can have a substantial effect on the particle EDX data collected during direct-on-filter analysis. Full article
(This article belongs to the Special Issue Size Distribution, Chemical Composition and Morphology of Mine Dust)
Show Figures

Figure 1

Figure 1
<p>Conceptual illustration of low and high particle loading density (PLD) and particle agglomeration on RCMD sample filters. (Note that colors in this figure are arbitrary and are simply meant to illustrate, in general, different particle types.)</p>
Full article ">Figure 2
<p>Sample preparation process for Set 1.</p>
Full article ">Figure 3
<p>Sample preparation process for Set 2.</p>
Full article ">Figure 4
<p>Example images of SEM-EDX analysis fields (at 1000× magnification) on the recovered RCMD filters with varying PLDs. The top row of images are SE micrographs; the bottom row shows BSE images, with the first 50 particles color-coded according to their EDX spectra. (The legend shows relative descriptions of the primary elemental abundance since multiple particle classification criteria were applied in this study.)</p>
Full article ">Figure 5
<p>Difference between the number percentage of particles in each major mineralogy class observed on the high PLD filter versus the low PLD filter in each Set 1 pair (excluding Pair 5), plotted as a function of the difference in the PLD observed on the high PLD filter versus the low PLD filter. Results show when particles were classified using (<b>a</b>) STD criteria (closed points) or (<b>b</b>) MOD criteria (open points).</p>
Full article ">Figure 6
<p>Difference between the number percentage of particles in each major mineralogy class observed on the high PLD filter versus the low PLD filter in each Set 1 pair, plotted as a function of the increase in the normalized elemental percent relative to the STD classification criteria percent. The difference in PLD observed on the high PLD filter versus the low PLD filter.</p>
Full article ">Figure 7
<p>Difference between the number percentage of particles in each major mineralogy class observed on the high PLD fields versus the low PLD fields for each filter with both high PLD and low PLD fields in Set 2, plotted as a function of the difference in the PLD observed on the high PLD fields versus the low PLD fields. Results show when particles were classified using (<b>a</b>) STD criteria (closed points) or (<b>b</b>) MOD criteria (open points).</p>
Full article ">Figure 8
<p>Difference between the number percentage of particles in each major mineralogy class observed on the D filter versus the R filter in each Set 2 pair, plotted as a function of the difference in the PLD observed on the D filter versus the R filter. Results show when particles were classified using (<b>a</b>) STD criteria (closed points) or (<b>b</b>) MOD criteria (open points).</p>
Full article ">Figure 9
<p>Difference between the number percentage of particles in each major mineralogy class observed on the D filter versus the R filter in each Set 2 pair, plotted against filter pair number. Again, results with STD criteria are shown with closed points, and those with MOD criteria are shown with open points.</p>
Full article ">
19 pages, 12914 KiB  
Article
Preparation of High-Purity Quartz Sand by Vein Quartz Purification and Characteristics: A Case Study of Pakistan Vein Quartz
by Mei Xia, Xiaoyong Yang and Zhenhui Hou
Minerals 2024, 14(7), 727; https://doi.org/10.3390/min14070727 - 19 Jul 2024
Viewed by 570
Abstract
This study focuses on the purification and evaluation of the high-purity quartz (HPQ) potential of vein quartz ore from Pakistan. Vein quartz is grayish-white and translucent, with its mineral composition mainly comprising quartz crystal. Processed quartz sand is obtained from quartz raw ore [...] Read more.
This study focuses on the purification and evaluation of the high-purity quartz (HPQ) potential of vein quartz ore from Pakistan. Vein quartz is grayish-white and translucent, with its mineral composition mainly comprising quartz crystal. Processed quartz sand is obtained from quartz raw ore through purifying technologies, including crushing, ultrasonic desliming, flotation, high-temperature calcination, water quenching, hot pressure acid leaching, and chlorination roasting. The microscopic characteristics show that the vein quartz raw ore has a medium-coarse granular metacrystalline structure, high quartz content, with only a small quantity of fine-grained K-feldspar. The inclusions primarily consist of large-sized primary inclusions and secondary fluid inclusions developed along the micro-fractures, and the content of inclusions in most areas of the crystal is very low or even nonexistent. The quartz ore with such inclusion characteristics is considered a relatively good raw material for quartz. Component analysis shows that the main impurity elements in the quartz ore are Al, K, Ca, Na, Ti, Fe, and Li, with a total impurity element content of 128.86 µg·g−1. After purification, only lattice impurity elements Al, Ti, and Li remain in the processed quartz sand, resulting in a total impurity element content of 24.23 µg·g−1, an impurity removal rate of 81.20%, and the purity of SiO2 reaching 99.998 wt.%. It is suggested that when the quartz raw ore contains high content of lattice impurity elements, such as Al, Li, and Ti, it is difficult to remove them by the current purification method. In industrial production, considering the economic cost, if quartz sand still contains high content of lattice impurity elements Al, Ti, and Li after flotation, it cannot be used as a raw material for high-end HPQ. Full article
(This article belongs to the Special Issue Physicochemical Properties and Purification of Quartz Minerals)
Show Figures

Figure 1

Figure 1
<p>Hand specimen of B6-1 vein quartz. (<b>a</b>) Translucent grayish-white. (<b>b</b>) Light yellowish-brown rust.</p>
Full article ">Figure 2
<p>Flow chart of purification experiment.</p>
Full article ">Figure 3
<p>The experimental setup used for chlorination roasting. Arrow represents rotation.</p>
Full article ">Figure 4
<p>Microphotographs (TPM and BSE) of the B6-1 raw vein quartz. (<b>a</b>) Single-polarized photos. (<b>b</b>,<b>c</b>) Orthographic polarized photos. (<b>d</b>–<b>f</b>) SEM microphotographs: K-feldspar.</p>
Full article ">Figure 5
<p>Microphotographs (TPM) of fluid inclusions in the B6-1 raw vein quartz. (<b>a</b>–<b>f</b>) TPM of 2.5×, 10× and 20× of primary inclusions. A certain number of isolated or clustered primary fluid inclusions, mainly H<sub>2</sub>O and CO<sub>2</sub>. (<b>g</b>–<b>l</b>) TPM of 2.5×, 10×, and 20× of secondary inclusions. A large number of small secondary inclusions distributed in bands or clusters along the fractures.</p>
Full article ">Figure 6
<p>The content of SiO<sub>2</sub> and impurity elements in in raw ore and different processed quartz sand. Acid leaching sand: flotation + acid leaching. (<b>a</b>) The content of SiO<sub>2</sub>. (<b>b</b>) The content of impurity elements. Chlorinated sand: flotation + acid leaching + chlorination roasting.</p>
Full article ">Figure 7
<p>Microphotographs (BSE) of B6-1 quartz sand. (<b>a</b>) Ultrasonic scrubbing quartz sand. (<b>b</b>) Flotation quartz sand. (<b>c</b>,<b>d</b>) Calcination quartz sand. (<b>e</b>,<b>f</b>) Acid leaching quartz sand.</p>
Full article ">Figure 8
<p>Microphotographs (TPM) of B6-1 quartz sand. (<b>a</b>–<b>c</b>) Flotation quartz sand. (<b>d</b>–<b>f</b>) Acid leaching quartz sand. (<b>g</b>–<b>i</b>) Chlorination quartz sand. Acid leaching quartz sand: flotation + acid leaching. Chlorinated quartz sand: flotation + acid leaching + chlorination roasting.</p>
Full article ">Figure 9
<p>Typical isomorphic substitution in quartz crystal.</p>
Full article ">
13 pages, 4358 KiB  
Article
Synthesis of Geopolymers Incorporating Mechanically Activated Fly Ash Blended with Alkaline Earth Carbonates: A Comparative Analysis
by Alexander M. Kalinkin, Elena V. Kalinkina, Ekaterina A. Kruglyak and Alla G. Ivanova
Minerals 2024, 14(7), 726; https://doi.org/10.3390/min14070726 - 19 Jul 2024
Viewed by 478
Abstract
The objective of this study is to perform a comparative analysis of the impact of incorporating alkaline earth metal carbonates (MCO3, where M–Mg, Ca, Sr, Ba) into low-calcium fly ash (FA) on the geopolymerization processes and the resultant properties of composite [...] Read more.
The objective of this study is to perform a comparative analysis of the impact of incorporating alkaline earth metal carbonates (MCO3, where M–Mg, Ca, Sr, Ba) into low-calcium fly ash (FA) on the geopolymerization processes and the resultant properties of composite geopolymers. Mechanical activation was employed to enhance the reactivity of the mixtures. The reactivity of the mechanically activated (FA + alkaline earth carbonate) blends towards NaOH solution was experimentally studied using XRD analysis and FTIR spectroscopy. In agreement with thermodynamic calculations, MgCO3 demonstrated the most active interaction with the alkaline solution, whereas strontium and barium carbonates exhibited little to no chemical interaction, and calcite was situated in the transition region. As the calcite content in the mixture with FA increased, the compressive strength of the geopolymers continuously improved. The addition of Mg, Sr, and Ba carbonates to the FA did not enhance the strength of geopolymers. However, the strength of geopolymers based on these blends was comparable with that of geopolymers based on 100% FA. The strength of geopolymers synthesized from the 100% FA and from the (90% FA + 10% MCO3) blends, mechanically activated for 180 s, at the age of 180 days was 11.0 MPa (0% carbonate), 11.1 MPa (10% MgCO3), 36.5 MPa (10% CaCO3), 13.6 MPa (10% SrCO3), and 12.4 MPa (10% BaCO3) MPa, respectively. The influence of carbonate additives on the properties of the composite geopolymers was examined, highlighting filler, dilution, and chemical effects. The latter determined the unique position of calcite among the carbonates of alkaline earth metals. Full article
Show Figures

Figure 1

Figure 1
<p>The XRD patterns: (<b>a</b>) SrCO<sub>3</sub>; (<b>b</b>) BaCO<sub>3</sub>. The phases marked are as follows: S—strontianite (SrCO<sub>3</sub>) (ICDD 00-005-0418), W—witherite (BaCO<sub>3</sub>) (ICDD 00-045-1471).</p>
Full article ">Figure 2
<p>Effect of carbonate content in the (FA + MCO<sub>3</sub>) (M–Mg, Ca, Sr, Ba) blends, milled for 180 s, on the compressive strength of geopolymers cured for 7 d (<b>a</b>), 28 d (<b>b</b>), and 180 d (<b>c</b>).</p>
Full article ">Figure 2 Cont.
<p>Effect of carbonate content in the (FA + MCO<sub>3</sub>) (M–Mg, Ca, Sr, Ba) blends, milled for 180 s, on the compressive strength of geopolymers cured for 7 d (<b>a</b>), 28 d (<b>b</b>), and 180 d (<b>c</b>).</p>
Full article ">Figure 3
<p>The compressive strength of geopolymers synthesized using the (90% FA + 10% MCO<sub>3</sub>) (M–Mg, Ca, Sr, Ba) blends, milled for 180 s, at the ages of 7 d, 28 d, and 180 d depending on the radius of the alkaline earth metal cation.</p>
Full article ">Figure 4
<p>The standard Gibbs energy (Δ<sub>r</sub>G°(298)) of reactions between alkaline earth metal carbonates and sodium hydroxide solution.</p>
Full article ">Figure 5
<p>The XRD patterns of the (80% FA + 20% magnesite) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test [<a href="#B28-minerals-14-00726" class="html-bibr">28</a>]. The phases marked are as follows: Q—quartz SiO<sub>2</sub> (ICDD 00-046-1045), M—mullite 3Al<sub>2</sub>O<sub>3</sub>∙2SiO<sub>2</sub> (ICDD 00-015-0776), G—magnesite MgCO<sub>3</sub> (ICDD 01-086-2348), B—brucite Mg(OH)<sub>2</sub> (ICCD 01-071-5972), and H—hydrotalcite Mg<sub>6</sub>Al<sub>2</sub>CO<sub>3</sub>(OH)<sub>16</sub>⋅4H<sub>2</sub>O (ICCD 00-041-1428).</p>
Full article ">Figure 6
<p>The XRD patterns of the (80% FA + 20% calcite) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test. The phases marked are as follows: Q—quartz SiO<sub>2</sub> (ICDD 00-046-1045), M—mullite 3Al<sub>2</sub>O<sub>3</sub>∙2SiO<sub>2</sub> (ICDD 00-015-0776), C—calcite CaCO<sub>3</sub> (ICDD 00-005-0586), and P—portlandite Ca(OH)<sub>2</sub> (ICDD 00-001-1079).</p>
Full article ">Figure 7
<p>The XRD patterns of the (80% FA + 20% SrCO<sub>3</sub>) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test. The phases marked are as follows: Q—quartz (ICDD 00-046-1045), M—mullite (ICDD 00-015-0776), and S—strontianite SrCO<sub>3</sub> (ICDD 00-005-0418).</p>
Full article ">Figure 8
<p>The XRD patterns of the (80% FA + 20% BaCO<sub>3</sub>) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test. The phases marked are as follows: Q—quartz (ICDD 00-046-1045), M—mullite (ICDD 00-015-0776), and W—witherite BaCO<sub>3</sub> (ICDD 00-045-1471).</p>
Full article ">Figure 9
<p>FTIR spectra of the (80% FA + 20% magnesite) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test.</p>
Full article ">Figure 10
<p>FTIR spectra of the (80% FA + 20% calcite) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test.</p>
Full article ">Figure 11
<p>FTIR spectra of the (80% FA + 20% SrCO<sub>3</sub>) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test.</p>
Full article ">Figure 12
<p>FTIR spectra of the (80% FA + 20% BaCO<sub>3</sub>) blend mechanically activated for 180 s: 1—before reactivity test; 2—residue after reactivity test.</p>
Full article ">
26 pages, 24038 KiB  
Article
Petrogenesis of the Early Jurassic–Early Cretaceous Adakite-like Rocks in the Erguna Block, NE China: Implications for the Tectonic Evolution of the Mongol–Okhotsk Ocean
by Yuanchao Wang, Yuanyi Zhao, Xinfang Shui and Zaili Tao
Minerals 2024, 14(7), 725; https://doi.org/10.3390/min14070725 - 19 Jul 2024
Viewed by 440
Abstract
The petrogenesis and geodynamic setting of the Mesozoic magmatic rocks in the Erguna Block, NE China remains controversial, especially the relationship between magmatism and the subduction history of the Mongol–Okhotsk oceanic plate. Here we present data for the Early Jurassic–Early Cretaceous adakite-like magmatic [...] Read more.
The petrogenesis and geodynamic setting of the Mesozoic magmatic rocks in the Erguna Block, NE China remains controversial, especially the relationship between magmatism and the subduction history of the Mongol–Okhotsk oceanic plate. Here we present data for the Early Jurassic–Early Cretaceous adakite-like magmatic rocks from Chaoman Farm in the northeastern part of the Erguna Block. Zircon U-Pb dating reveals that the syenogranites crystallized at around 190–180 Ma, while the monzonites, quartz diorite porphyries, and quartz monzonite porphyries were emplaced at around 147–143 Ma. The syenogranites, monzonites, quartz diorite porphyries, and quartz monzonite porphyries are adakite-like rocks. The syenogranites and quartz monzonite porphyries were produced by the partial melting of a thickened ancient mafic lower continental crust and a thickened juvenile lower crust, respectively. Meanwhile, the monzonites and quartz diorite porphyries were formed as a result of partial melting of the oceanic crust. In conclusion, the occurrence of these Early Jurassic magmatic rocks was closely linked to the process of southward subduction of the Mongol–Okhotsk oceanic plate. On the contrary, the Late Jurassic to early Early Cretaceous magmatism (147–143 Ma) occurred in an extensional environment, and was probably triggered by upwelling of the asthenosphere. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Simplified geological sketch map of the CAOB (the Central Asian Orogenic Belt) showing the main tectonic subdivisions [<a href="#B28-minerals-14-00725" class="html-bibr">28</a>]. (<b>b</b>) Tectonic sketch map of NE China [<a href="#B29-minerals-14-00725" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Geological map of the Chaoman Forest Farm polymetallic exploration area in the northeastern part of the Erguna Block.</p>
Full article ">Figure 3
<p>Representative photomicrographs of the Chaoman Forest Farm Mesozoic magmatic rocks in the Erguna Block; (<b>a</b>) ZK21001-55 Syenogranite; (<b>b</b>) ZK21001-59 Syenogranite; (<b>c</b>) ZK0401-9 Monzonite; (<b>d</b>) ZK0401-1 Quartz diorite porphyry; (<b>e</b>) ZK0401-3 Quartz diorite porphyry; (<b>f</b>) ZK21001-43 Quartz monzonite porphyry; Bi = biotite; Amp = amphibole; Kfs = K-feldspar; Pl = pla-gioclase; Qtz = quartz.</p>
Full article ">Figure 4
<p>CL images of zircon grains of the Chaoman Forest Farm Mesozoic magmatic rocks. Solid and dashed circles indicate the locations of U–Pb dating and Hf isotope analyses, respectively (<b>a</b>–<b>h</b>).</p>
Full article ">Figure 5
<p>U–Pb concordia diagrams of the Chaoman Forest Farm Mesozoic magmatic rocks. 190 ± 1.4 Ma (MSWD = 0.03) (<b>a</b>); 190 ± 1.2 Ma (MSWD = 0.28) (<b>b</b>); 189 ± 0.94 Ma (MSWD = 0.2) (<b>c</b>); 146 ± 1.5 Ma (MSWD = 0.08) (<b>d</b>); 144 ± 1.2 Ma (MSWD = 0.08) (<b>e</b>); 150 ± 1.5 Ma (MSWD = 0.02) (<b>f</b>); 143 ± 0.9 Ma (MSWD = 0.1) (<b>g</b>); 190 ± 3.7 Ma (MSWD = 1.3) (<b>h</b>).</p>
Full article ">Figure 6
<p>(<b>a</b>) Total alkalis vs. silica diagram [<a href="#B63-minerals-14-00725" class="html-bibr">63</a>]; (<b>b</b>) K<sub>2</sub>O vs. SiO<sub>2</sub> diagram [<a href="#B64-minerals-14-00725" class="html-bibr">64</a>]; (<b>c</b>) A/NK vs. A/CNK diagram [<a href="#B65-minerals-14-00725" class="html-bibr">65</a>]; (<b>d</b>) Na<sub>2</sub>O + K<sub>2</sub>O vs. 10,000 Ga/Al discrimination diagram [<a href="#B68-minerals-14-00725" class="html-bibr">68</a>]. Data sources: Lower crust-derived adakites in the Lhasa terrane and Subducted oceanic crust-derived adakites in the Lhasa terrane [<a href="#B69-minerals-14-00725" class="html-bibr">69</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Sr/Y vs. Y diagrams [<a href="#B66-minerals-14-00725" class="html-bibr">66</a>] and (<b>b</b>) diagram of batch-melting modeling of chondrite-normalized (La/Yb)<sub>N</sub> ratios vs. (Yb)<sub>N</sub> [<a href="#B70-minerals-14-00725" class="html-bibr">70</a>], where N means normalized to chondrite [<a href="#B71-minerals-14-00725" class="html-bibr">71</a>]. An Eastern Pontides gabbro (G518) [<a href="#B72-minerals-14-00725" class="html-bibr">72</a>] is used as the source rock for the REE modeling under amphibolite and eclogite conditions, with varying garnet contents and respective partition coefficients (I–VI).</p>
Full article ">Figure 8
<p>Chondrite-normalized REE pattern (<b>a</b>,<b>c</b>) [<a href="#B73-minerals-14-00725" class="html-bibr">73</a>] and spider diagrams (<b>b</b>,<b>d</b>) [<a href="#B71-minerals-14-00725" class="html-bibr">71</a>] for the Chaoman Forest Farm Mesozoic magmatic rocks.</p>
Full article ">Figure 9
<p>(<b>a</b>) Plot of zircon ε<sub>Hf</sub>(<span class="html-italic">t</span>) values vs. U–Pb ages, (<b>b</b>) diagrams of ε<sub>Nd</sub>(<span class="html-italic">t</span>) vs. (<sup>87</sup>Sr/<sup>86</sup>Sr)<sub>i</sub> for the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: Taipingchuan igneous rocks [<a href="#B74-minerals-14-00725" class="html-bibr">74</a>]; Fukeshan igneous rocks [<a href="#B8-minerals-14-00725" class="html-bibr">8</a>]; Wunugetushan igneous rocks [<a href="#B19-minerals-14-00725" class="html-bibr">19</a>,<a href="#B75-minerals-14-00725" class="html-bibr">75</a>]; Badaguan igneous rocks [<a href="#B15-minerals-14-00725" class="html-bibr">15</a>]. The fields for the Erguna Block are from Deng et al., (2019a) [<a href="#B8-minerals-14-00725" class="html-bibr">8</a>]. The fields for MORB (Mid-Oceanic Ridge Basalt), OIB (Ocean Island Basalt) and IAB (Island Arc Basalt) are from Vervoort et al., (1999) [<a href="#B76-minerals-14-00725" class="html-bibr">76</a>]. EMI and EMII represent two types of mantle end-members [<a href="#B77-minerals-14-00725" class="html-bibr">77</a>]. The new continental crust (island arc) evolutionary line is defined by isotopic growth from <sup>176</sup>Hf/<sup>177</sup>Hf = 0.279703 at 4.55 Ga to 0.283145 at present, with <sup>176</sup>Lu/<sup>177</sup>Hf = 0.0375 [<a href="#B78-minerals-14-00725" class="html-bibr">78</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) (<sup>207</sup>Pb/<sup>204</sup>Pb)<sub>i</sub> vs. (<sup>206</sup>Pb/<sup>204</sup>Pb)<sub>i</sub> and (<b>b</b>) (<sup>208</sup>Pb/<sup>204</sup>Pb)<sub>i</sub> vs. (<sup>206</sup>Pb/<sup>204</sup>Pb)<sub>i</sub> for the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: subducted oceanic slab-derived adakites and MORB [<a href="#B80-minerals-14-00725" class="html-bibr">80</a>]; Northern Hemisphere Reference Line (NHRL) [<a href="#B79-minerals-14-00725" class="html-bibr">79</a>]; mantle source reservoirs BSE, DMM, EM I and EM II [<a href="#B81-minerals-14-00725" class="html-bibr">81</a>].</p>
Full article ">Figure 11
<p>Discrimination diagrams for the Chaoman Forest Farm Mesozoic magmatic rocks. (<b>a</b>) MgO vs. SiO<sub>2</sub> diagram [<a href="#B98-minerals-14-00725" class="html-bibr">98</a>]. Data for metabasaltic and eclogite experimental melts (1–4 GPa) are from Rapp et al., (1999) and references therein [<a href="#B96-minerals-14-00725" class="html-bibr">96</a>]; (<b>b</b>) Mg<sup>#</sup> vs. SiO<sub>2</sub> diagram [<a href="#B84-minerals-14-00725" class="html-bibr">84</a>]. Mantle AFC curves are after Rapp et al., (1999) (Curve 1); the proportion of assimilated peridotite is also shown. The crustal AFC curve is after Stern and Kilian (1996) (Curve 2) [<a href="#B95-minerals-14-00725" class="html-bibr">95</a>]; (<b>c</b>) Ni vs. SiO<sub>2</sub> diagram [<a href="#B98-minerals-14-00725" class="html-bibr">98</a>]; (<b>d</b>) Cr vs. SiO<sub>2</sub> diagram [<a href="#B98-minerals-14-00725" class="html-bibr">98</a>]; (<b>e</b>) Ni versus Cr diagram [<a href="#B69-minerals-14-00725" class="html-bibr">69</a>]; (<b>f</b>) Th/La versus Th diagram. The data for upper continental crust are from Plank (2005) and references therein [<a href="#B99-minerals-14-00725" class="html-bibr">99</a>]. The data for marine sediments are from Plank and Langmuir (1998) and for MORB are from Niu and Batiza (1997) [<a href="#B100-minerals-14-00725" class="html-bibr">100</a>,<a href="#B101-minerals-14-00725" class="html-bibr">101</a>].</p>
Full article ">Figure 12
<p>Diagrams of whole-rock ε<sub>Nd</sub>(t) (<b>a</b>) and zircon ε<sub>Hf</sub>(t) (<b>b</b>) against U–Pb ages of the Chaoman Forest Farm Mesozoic magmatic rocks. Data sources: Mesozoic mafic rocks in the Erguna Block [<a href="#B5-minerals-14-00725" class="html-bibr">5</a>,<a href="#B19-minerals-14-00725" class="html-bibr">19</a>,<a href="#B111-minerals-14-00725" class="html-bibr">111</a>]; Mesozoic intermediate–felsic rocks in the Erguna Block [<a href="#B2-minerals-14-00725" class="html-bibr">2</a>,<a href="#B4-minerals-14-00725" class="html-bibr">4</a>,<a href="#B5-minerals-14-00725" class="html-bibr">5</a>,<a href="#B8-minerals-14-00725" class="html-bibr">8</a>,<a href="#B10-minerals-14-00725" class="html-bibr">10</a>,<a href="#B13-minerals-14-00725" class="html-bibr">13</a>,<a href="#B15-minerals-14-00725" class="html-bibr">15</a>,<a href="#B16-minerals-14-00725" class="html-bibr">16</a>,<a href="#B17-minerals-14-00725" class="html-bibr">17</a>,<a href="#B74-minerals-14-00725" class="html-bibr">74</a>,<a href="#B75-minerals-14-00725" class="html-bibr">75</a>,<a href="#B117-minerals-14-00725" class="html-bibr">117</a>,<a href="#B118-minerals-14-00725" class="html-bibr">118</a>,<a href="#B119-minerals-14-00725" class="html-bibr">119</a>,<a href="#B120-minerals-14-00725" class="html-bibr">120</a>,<a href="#B121-minerals-14-00725" class="html-bibr">121</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) Nb vs. Y and (<b>b</b>) Rb vs. (Y + Nb) diagrams [<a href="#B127-minerals-14-00725" class="html-bibr">127</a>]. Abbreviations: WPG: Within-plate granitoid; VAG: volcanic arc granitoid; Syn-COLG: syn-collision granitoid; ORG: ocean ridge granitoid.</p>
Full article ">Figure 14
<p>Conceptual diagram illustrating the proposed tectonic model and magma genesis of the Εarly Jurassic–late Early Cretaceous Mesozoic magmatic rocks in the Erguna Block. (<b>a</b>) Εarly Jurassic, (<b>b</b>) Middle Jurassic–early Cretaceous.</p>
Full article ">
26 pages, 14806 KiB  
Article
Genesis of Fe-Ti-(V) Oxide-Rich Rocks by Open-System Evolution of Mafic Alkaline Magmas: The Case of the Ponte Nova Massif, SE Brazil
by Amanda Andrade de Souza, Rogério Guitarrari Azzone, Luanna Chmyz, Lina Maria Cetina Tarazona, Fábio Ramos Dias de Andrade, José Vinicius Martins, Excelso Ruberti and Celso de Barros Gomes
Minerals 2024, 14(7), 724; https://doi.org/10.3390/min14070724 - 19 Jul 2024
Viewed by 404
Abstract
The formation of Fe-Ti oxides-rich layers is commonly associated with open-system magma chamber dynamics. These processes are widely discussed due to the economic importance of Fe-Ti-(V) deposits, although an alkaline-system approach to the matter is still scarce. In this study, we use petrography, [...] Read more.
The formation of Fe-Ti oxides-rich layers is commonly associated with open-system magma chamber dynamics. These processes are widely discussed due to the economic importance of Fe-Ti-(V) deposits, although an alkaline-system approach to the matter is still scarce. In this study, we use petrography, mineral chemistry, X-ray diffraction and elemental geochemical analyses (whole-rock and Sr isotopes) to discuss the process associated with the formation of Fe-Ti-(V) oxide-rich clinopyroxenite (OCP, 7–15 vol.%) and magnetitite (MTT, 85 vol.%) from the Ponte Nova alkaline mafic–ultramafic massif (PN, K-Ar 87.6 Ma). Ilmenite and Ti-magnetite from both OCP and MTT exhibit higher MgO contents (MgO > 5.0 wt%) than other PN rocks. OCP shows high 87Sr/86Sri ratios, equivalent to crustal-contaminated lithotypes of the PN Central Intrusion, while MTTs are less radiogenic. The oxide supersaturation in silicate mafic magmas is typically associated with the dislocation of the liquid cotectic evolution line, shifting to Fe-Ti-(V) oxide minerals stability field, mainly Ti-magnetite. Different magmatic processes can lead to these changes such as crustal contamination and magma recharge. For the PN massif, the OCP was formed by the assimilation of crustal contaminants in a mush region, near the magma chamber upper walls, which was associated with the evolution of the main pulse. Differently, the MTT would have its origin related to the interaction between magma chamber evolved liquids and more primitive liquids during a new episode of magma recharge. Lastly, post-magmatic events were superimposed on these rocks, generating sulfides. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geological setting of alkaline intrusions of Cretaceous magmatism in the southeastern part of the South American Platform showing the distribution of major alkaline occurrences of Alto Paranaíba, Serra do Mar and Ponta Grossa Arch alkaline provinces (modified after [<a href="#B20-minerals-14-00724" class="html-bibr">20</a>]). (<b>b</b>) Detail of (<b>a</b>), with the location of the main occurrences. Legend: I—Sediments or sedimentary rocks (Cenozoic); II—Alkaline complexes (Meso-Cenozoic); III—Sedimentary rocks of the Paraná Basin (Ordovician-Silurian); IV—Crystalline Basement, Brasília and Ribeira orogenic belts (Precambrian). (<b>c</b>) Simplified geological map of the Ponte Nova alkaline mafic–ultramafic massif (modified after [<a href="#B20-minerals-14-00724" class="html-bibr">20</a>]). (<b>d</b>) Schematic model of the lithostratigraphy of the studied rocks, after [<a href="#B36-minerals-14-00724" class="html-bibr">36</a>].</p>
Full article ">Figure 2
<p>Petrographic aspects of the OCP facies. (<b>a</b>) OCP scanned thin section. Transmitted and plane-polarized light: (<b>b</b>) Photomicrograph of OCP cumulatic texture and mineral assemblage constituted of clinopyroxene, ilmenite and apatite (granular and acicular) disseminated through the matrix. (<b>c</b>) Orientation and compaction structures. (<b>d</b>) Apatite-rich domain with subordinate kaersutite, ilmenite and biotite. Reflected and plane-polarized light: (<b>e</b>,<b>f</b>) Anhedral and acicular ilmenite associated with cumulus clinopyroxene and weathered sulfides along fractured zones. Backscattered electron image (BSE): (<b>g</b>) Ilmenite (in triple-junction contacts) and sulfides; and (<b>h</b>) Ilmenite corroded rims and substitution by hematite along fractures. Abbreviations: bt—biotite, cpx—clinopyroxene, cpy—chalcopyrite, hem—hematite, ilm—ilmenite, py—pyrite, sul—weathered sulfide, opq—opaque minerals.</p>
Full article ">Figure 3
<p>Petrographic aspects of the MTT facies. (<b>a</b>) MTT scanned thin section. Transmitted light: (<b>b</b>) Photomicrograph of MTT cumulatic texture. Reflected and plane-polarized light: (<b>c</b>) Euhedral magnetite crystals with submillimetric ilmenite lamellae and fine oxi-exsolutions. (<b>d</b>) Substitution textures to spinel predominantly at crystal rims. Backscattered electron image (BSE): (<b>e</b>) Cumulus Ti-magnetite, clinopyroxene and ilmenite. (<b>f</b>) Zoned clinopyroxene macrocrystal. Reflected and plane-polarized light: (<b>g</b>) Clinopyroxene grains with magnetite inclusions, cumulus magnetite and ilmenite (reflected and plane-polarized light). (<b>h</b>) Intercumulus ilmenite, cumulus magnetite and clinopyroxene (reflected and plane-polarized light). Abbreviations: cpx—clinopyroxene, ilm—ilmenite, Ti-mgt—Ti-magnetite, ulv—ulvöspinel.</p>
Full article ">Figure 4
<p>OCP and MTT clinopyroxene compositions compared to compiled clinopyroxene data (after [<a href="#B38-minerals-14-00724" class="html-bibr">38</a>]) of CI (green line) and other PN intrusions (gray line). (<b>a</b>) Mg#<sub>cpx</sub> vs. Cr<sub>2</sub>O<sub>3</sub> (<b>b</b>) Mg#<sub>cpx</sub> vs. TiO<sub>2</sub> (<b>c</b>) Mg#<sub>cpx</sub> vs. Na<sub>2</sub>O (<b>d</b>) Mg#<sub>cpx</sub> vs. Al<sub>2</sub>O<sub>3</sub>. Pink dotted lines refer to core–intermediate–rim composition of one selected crystal from OCP facies. Orange dotted lines refer to core–rim composition of one selected crystal from MTT facies.</p>
Full article ">Figure 5
<p>Ti-R<sub>2</sub>-R<sub>3</sub> ternary diagram of Fe-Ti oxides of PN with the magnetite–ulvöspinel and ilmenite–hematite solid solution lines, in atoms per formula unit (apfu).</p>
Full article ">Figure 6
<p>Mg# index of Ti-magnetite and ilmenite from OCP, MTT, CI and other PN intrusions. The grey field represents analyses with equivalent Mg# to other PN intrusions. (<b>a</b>) Mg# in Ti-magnetite. (<b>b</b>) Mg# in ilmenite. Abbreviations: OCP—oxide-rich clinopyroxenites, MTT—magnetitites, CI—Central Intrusion.</p>
Full article ">Figure 7
<p>Mg# Fe-Ti oxides compositions vs.TiO<sub>2</sub> and Cr<sub>2</sub>O<sub>3</sub> from OCP, MTT, CI and other PN intrusions. (<b>a</b>) Mg# vs. TiO<sub>2</sub> for Ti-magnetite populations. (<b>b</b>) Mg# vs. TiO<sub>2</sub> for ilmenite populations. (<b>c</b>) Mg# vs. Cr<sub>2</sub>O<sub>3</sub> for Ti-magnetite populations. (<b>d</b>) Mg# vs. Cr<sub>2</sub>O<sub>3</sub> for ilmenite populations.</p>
Full article ">Figure 8
<p>(<b>a</b>) SiO<sub>2</sub> vs. FeO (wt%) analyses of Ti-magnetite populations from MTT facies. (<b>b</b>) SiO<sub>2</sub> vs. Fe<sub>2</sub>O<sub>3</sub> (wt%) analyses of Ti-magnetite populations from MTT facies. Results obtained in this study via laser ablation ICP-MS and electron microprobe (EPMA).</p>
Full article ">Figure 9
<p>(<b>a</b>) Discriminant diagram of log(Al) + log(Ti) + log(V) vs. log(Mn)/[log(Co) + log(Mg)] for Ti-magnetite from MTT facies and compiled data of Fe-Ti-V deposits, adapted from [<a href="#B51-minerals-14-00724" class="html-bibr">51</a>]. (<b>b</b>) Trace elements (ppm) diagrams for Ti-magnetite from MTT facies. Ti-magnetite compiled data from worldwide Fe-Ti-V deposits are presented for comparison [<a href="#B51-minerals-14-00724" class="html-bibr">51</a>].</p>
Full article ">Figure 10
<p>TiO<sub>2</sub> (wt%) vs. Mg, Mn, V, Co, Cu, Zr, Zn and Cr (ppm) contents of ilmenites from the OCP and MTT facies.</p>
Full article ">Figure 11
<p>(<b>a</b>) Mg/Mn equilibrium test after [<a href="#B55-minerals-14-00724" class="html-bibr">55</a>] in magnetite–ilmenite pairs from PN. 2σ represents the standard deviation of the equilibrium line (in orange). (<b>b</b>) Kernel density estimation (KDE) and box plot diagrams for the obtained temperatures. Temperature estimation after [<a href="#B53-minerals-14-00724" class="html-bibr">53</a>]. (<b>c</b>) Temperature vs. oxygen fugacity of magnetite-ilmenite pairs from PN massif. (<b>d</b>) Temperature vs. ΔFMQ of magnetite-ilmenite pairs from PN massif. Oxygen fugacity estimation after [<a href="#B54-minerals-14-00724" class="html-bibr">54</a>]. Buffers were obtained in [<a href="#B56-minerals-14-00724" class="html-bibr">56</a>]. Abbreviations: MH—magnetite-hematite, NNO—nickel-nickel-oxide; FMQ—fayalite-magnetite-quartz; WM—wüstite-magnetite.</p>
Full article ">Figure 12
<p>(<b>a</b>) R<sub>1</sub>-R<sub>2</sub> diagram [<a href="#B57-minerals-14-00724" class="html-bibr">57</a>] for rocks from the PN massif and host rocks. Average mineral poles were compiled from [<a href="#B36-minerals-14-00724" class="html-bibr">36</a>] and are representative of the main assemblage found in PN (modified after [<a href="#B20-minerals-14-00724" class="html-bibr">20</a>]). (<b>b</b>) Al-Si-(Ti + Fe + Mg + Mn) ternary diagram for rocks from the PN massif and host rocks in molar proportions. Mineral chemistry data (in gray) for clinopyroxene and Fe-Ti oxides are from this work. Nepheline, feldspar and olivine compositions in the ternary are from their ideal formula. Abbreviations: cpx—clinopyroxene, Ti-mgt—Ti-magnetite, ol—olivine, ilm—ilmenite, nph—nepheline, fsp—alkali-feldspar, pl—plagioclase, ap—apatite, bt—biotite, krs—kaersutite.</p>
Full article ">Figure 13
<p>Bivariant diagrams of TiO<sub>2</sub> (wt%) vs. V, Zn, Co and Cu (ppm) from the PN massif rocks, and MgO (wt%) vs. Cr and Ni (ppm) from the PN massif rocks.</p>
Full article ">Figure 14
<p>(<b>a</b>) <sup>87</sup>Sr/<sup>86</sup>Sr<sub>i</sub> of PN massif intrusions. Sr isotopes obtained in whole-rock, apatite and plagioclase analyses. (<b>b</b>) <sup>87</sup>Sr/<sup>86</sup>Sr<sub>i</sub> vs. Sr (ppm) from the PN massif rocks and host rocks. Compiled data from [<a href="#B20-minerals-14-00724" class="html-bibr">20</a>]. Abbreviations: AFC—assimilation + fractional crystallization.</p>
Full article ">Figure 15
<p>XRD diffractogram of the OCP facies. Abbreviations: rmb—rhomboclase, mlnt—melantherite, aug—augite, ilm—ilmenite.</p>
Full article ">Figure 16
<p>Schematic model for the proposed origin to OCP and MTT facies. (<b>a</b>) Assimilation of partially melted crustal components into the Central Intrusion upper walls, leading to the displacement of the crystallization cotecticinto the oxide phase and formation of OCP facies. (<b>b</b>) In the final stages of crystallization, and after the formation of OCP facies, a new episode of magma recharge in the PN massif. (<b>c</b>) Displacement of crystallization cotectic into the oxide phase due to the interaction of primitive and evolved liquids during a magma recharge episode, forming the MTT facies. (<b>d</b>) Eventual erosion of the uppermost sequences of PN, exposing the lithologies approached in this study.</p>
Full article ">
10 pages, 6264 KiB  
Article
Bismuth White (Bismuth Oxychloride) and Its Use in Portrait Miniatures Painted by George Engleheart
by Lucia Burgio
Minerals 2024, 14(7), 723; https://doi.org/10.3390/min14070723 - 19 Jul 2024
Viewed by 475
Abstract
This article documents the discovery of ‘bismuth white’ on three late eighteenth-century portrait miniatures in the Victoria and Albert Museum collections, painted by renowned English artist George Engleheart. Metallic bismuth and bismuth-containing minerals have been known for centuries and were used on various [...] Read more.
This article documents the discovery of ‘bismuth white’ on three late eighteenth-century portrait miniatures in the Victoria and Albert Museum collections, painted by renowned English artist George Engleheart. Metallic bismuth and bismuth-containing minerals have been known for centuries and were used on various types of artistic production, from German Wismutmalerei to medieval manuscripts and Renaissance paintings. However, until now they had never been documented on portrait miniatures, despite documentary evidence that suggests their use. The Raman analysis of the three miniatures shows that bismuth oxychloride (BiOCl, corresponding to the mineral bismoclite) is present, and XRF data prove that this material was used as a white pigment in its own right. This work is a pilot study: it represents the first step in the rediscovery of bismuth white as an artist’s pigment, and hopes to provide encouragement to other institutions to look deeper in their collections and map out the use of a relatively rare white material which until now had not been detected or documented in fine art objects. Full article
(This article belongs to the Special Issue Geomaterials and Cultural Heritage)
Show Figures

Figure 1

Figure 1
<p>The three miniatures analysed in this study: (<b>a</b>) P.26-1922, (<b>b</b>) P.20-1929 and (<b>c</b>) P.22-1937.</p>
Full article ">Figure 2
<p>XRF spectrum from a spot (see inset) in the pale blue dress in miniature P.22-1937, showing that bismuth is present.</p>
Full article ">Figure 3
<p>Representative Raman spectra of bismuth oxychloride obtained from the three miniatures, compared to the spectrum of a reference sample of bismuth oxychloride from the Natural History Museum collection.</p>
Full article ">Figure 4
<p>(<b>a</b>) Detail of ruff near the neck in P.20-1929 and (<b>b</b>) detail of dress near the neck in P.26-1922.</p>
Full article ">Figure 5
<p>(<b>a</b>) XRF distribution map of bismuth (La1) of (<b>b</b>) P.26-1922.</p>
Full article ">Figure 6
<p>Occurrences of ‘bizmuth white’ in William Wood’s handwritten ledgers (highlighted in red): (<b>a</b>) first page of the memorandum; (<b>b</b>) page marked as no. 39; (<b>c</b>) entry for miniature 5005; (<b>d</b>) entry for miniature 5006; (<b>e</b>) entry for miniature 5011; (<b>f</b>) entry for miniature 5025.</p>
Full article ">
21 pages, 33803 KiB  
Article
Clarification of Distinguishing Natural Super-Reduced Phase from Synthetics Based on Inclusions
by Yutong Ma, Mengqi Miao, Ming Chen and Shan Qin
Minerals 2024, 14(7), 722; https://doi.org/10.3390/min14070722 - 18 Jul 2024
Viewed by 364
Abstract
Super-reduced phases (SRPs), such as silicon carbide (SiC) and metal silicides, have increasingly been reported in various geological environments. However, their origin remains controversial. SRP inclusions (e.g., metal silicides and metallic silicon (Si0)) within SiC are commonly believed to indicate a [...] Read more.
Super-reduced phases (SRPs), such as silicon carbide (SiC) and metal silicides, have increasingly been reported in various geological environments. However, their origin remains controversial. SRP inclusions (e.g., metal silicides and metallic silicon (Si0)) within SiC are commonly believed to indicate a natural origin. Here, we identified an unusual SRP assemblage (SiC, (Fe,Ni)Si2, and Si0) in situ in an H5-type Jingshan ordinary chondrite. Simultaneously, our analysis showed that the SiC abrasives contain (Fe,Ni)Si2 and Si0 inclusions. Other inclusions in the artificial SiC were similar to those in natural SiC (moissanite) reported in reference data, including diverse metal silicides (e.g., FeSi, FeSi2, Fe3Si7, and Fe5Si3), as well as a light rare earth element-enriched SiO phase and Fe-Mn-Cr alloys. These inclusions were produced by the in situ reduction of silica and the interaction between Si-containing coke and hot metals during the synthesis of the SiC abrasives. The results demonstrate that the SRP assemblage in the Jingshan chondrite originates from abrasive contamination and that the SRP inclusions (with a low content of Ca, Al, Ti, and Zr) cannot be used as a conclusive indicator for natural SiC. Additionally, the morphologies, biaxiality, and polytypes (determined by Raman spectroscopy) of SiC abrasives bear resemblance to those reported for natural SiC, and caution must be exercised when identifying the origin of SRP in samples processed by conventional methods using SiC abrasives. At the end of this paper, we propose more direct and reliable methods for distinguishing between natural and synthetic SiC. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) BSE image of the region containing the Jingshan-SRP assemblage in the thin section J-1. The (Fe,Ni)Si<sub>2</sub> grain is located in a crack between the orthopyroxene (Opx) matrix and an impact melt vein (IMV); chromite (Chr), kamacite (Kam) blebs, olivine (Ol), and pyroxene (Px) are distributed within the vein; (<b>b</b>) optical microscope image under reflected light; (Fe,Ni)Si<sub>2</sub> shows a bright white, metallic lustre; (<b>c</b>) the magnified BSE image shows the intergrowth of (Fe,Ni)Si<sub>2</sub> and Si<sup>0</sup>, surrounded by the fragments of SiC and Opx; (<b>d</b>,<b>e</b>) the BSE images of the crack area between the Opx matrix and an IMV in the thin section J-2; no SRP assemblage was identified; (<b>f</b>) the BSE image and (<b>g</b>) SE image of the alumina contamination mixed into J-2. Tro—troilite; Pl—plagioclase; EP—epoxy resin.</p>
Full article ">Figure 2
<p>BSE image and X-ray elemental maps for Si, Fe, Ni, Ti, Cr, Mg, and O in the region of the Jingshan-SRP assemblage.</p>
Full article ">Figure 3
<p>(<b>a</b>) Raman spectra of SiC-<span class="html-italic">6H</span>, (Fe,Ni)Si<sub>2</sub>, and Si<sup>0</sup> in the Jingshan chondrite; (<b>b</b>) the comparison of the Raman spectra of (Fe,Ni)Si<sub>2</sub> with similar synthetic silicides β-FeSi<sub>2</sub> [<a href="#B51-minerals-14-00722" class="html-bibr">51</a>] and NiSi<sub>2</sub> [<a href="#B52-minerals-14-00722" class="html-bibr">52</a>]. The red arrows in (<b>b</b>) represent the shift of the corresponding peaks of (Fe,Ni)Si<sub>2</sub> relative to β-FeSi<sub>2</sub> and NiSi<sub>2</sub>, respectively.</p>
Full article ">Figure 4
<p>EBSD patterns of (Fe,Ni)Si<sub>2</sub> were recorded from two orientations (<b>left</b>) and their patterns indexed with the <span class="html-italic">Fm</span><math display="inline"><semantics> <mrow> <mover> <mn>3</mn> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">m</span> NiSi<sub>2</sub> structure (<b>right</b>). Six diffraction bands were set.</p>
Full article ">Figure 5
<p>Morphology of the artificial green SiC abrasives (120#) under the microscope: (<b>a</b>,<b>b</b>) colour varies from bluish green, green and light green to colourless, and the conchoidal fractures of the artificial SiC are visible; (<b>c</b>) an artificial SiC grain with a regular shape; (<b>d</b>) residual coke is visible between the SiC abrasives.</p>
Full article ">Figure 6
<p>BSE images of different inclusions in the SiC abrasives (the serial numbers I–XV and #1 of the inclusions in this figure correspond to those in <a href="#minerals-14-00722-t001" class="html-table">Table 1</a>). (<b>a,b</b>) elongated triangular (Fe,Ni)Si<sub>2</sub> and Fe<sub>2</sub>Si<sub>7</sub>; (<b>c</b>) granular FeSi<sub>2</sub>; (<b>d</b>) granular Fe<sub>2</sub>Si<sub>5</sub>; (<b>e</b>) irregular planar (Fe,Mn)Si<sub>3</sub>; (<b>f</b>) irregular planar Fe<sub>2</sub>Si<sub>9</sub>; (<b>g</b>) ‘oil splash’ droplet-shaped FeSi<sub>3</sub>; (<b>h</b>) lath-shaped Fe<sub>3</sub>Si<sub>7</sub>; (<b>i</b>) SiO<sub>2</sub> was detected in SiC, which contains an irregular planar inclusion; (<b>j</b>) shows a magnified image of (<b>i</b>)—the distribution of Fe-silicides with different compositions in an irregular plane; (<b>k</b>) ‘oil splash’ droplet-shaped LREE-enriched SiO were dispersed in the planar FeSi<sub>2</sub>; (<b>l</b>) irregular planar Si<sup>0</sup>; (<b>m</b>–<b>o</b>) lamellar Fe-Mn-Cr.</p>
Full article ">Figure 7
<p>(<b>a</b>) Several representative Raman spectra of the SiC abrasives (three grey lines) and Si<sup>0</sup> inclusions (red line). The upper right illustration shows the test location of the Si<sup>0</sup> inclusion; (<b>b</b>) the high-magnification detail of the SiC-FTO mode from the 750 to 820 cm<sup>−1</sup> band; (<b>c</b>) the high-magnification detail of the SiC-FLO mode from the 850 to 1200 cm<sup>−1</sup> band. FTA = folded transverse acoustic phonon; FLA = folded longitudinal acoustic phonon; FTO = folded transverse optic phonon; FLO = folded longitudinal optic phonon.</p>
Full article ">
19 pages, 7350 KiB  
Article
Inhibitory Effects of Polysaccharides on the Dolomitization Reaction of Calcite at 200 °C
by Yang Wei and Hiromi Konishi
Minerals 2024, 14(7), 721; https://doi.org/10.3390/min14070721 - 18 Jul 2024
Viewed by 335
Abstract
This study investigates the impact of dissolved carboxymethyl cellulose (CMC) and agar on the dolomitization reaction of calcite at 200 °C. Previous studies have suggested that CMC and agar promote dolomite precipitation at room temperature. However, this study found that their decomposition products [...] Read more.
This study investigates the impact of dissolved carboxymethyl cellulose (CMC) and agar on the dolomitization reaction of calcite at 200 °C. Previous studies have suggested that CMC and agar promote dolomite precipitation at room temperature. However, this study found that their decomposition products hinder the reaction at 200 °C, with uncertainty about their role at other temperatures. The inhibitory effect of the decomposition products could be attributed to their adsorption onto calcite surfaces, which hinders their dissolution. This results in a longer reaction induction period and replacement period. Regression analysis demonstrates that the 0.1 g/L agar and 0.2 g/L CMC series decrease the cation ordering rate of dolomite produced from synthetic calcite when compared with series without polysaccharides. In contrast, the 0.1 g/L CMC series shows a slight increase in the cation ordering rate compared with series without polysaccharides. The findings of this study suggest a notable potential impact of the decomposition products of polysaccharides on the ordering of dolomite, although it is uncertain whether they inhibit this ordering process. The inhibitory effect observed in the decomposition products of CMC and agar could also exist in the decomposition products of the extracellular polymeric substances (EPS) and bacteria cell walls found in sedimentary rocks during burial diagenesis. Therefore, further research is necessary to understand the role of EPS and bacteria cell walls in dolomitization, since their impact is not always predictable. Full article
(This article belongs to the Special Issue Microbial Biomineralization and Organimineralization)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The graphs show the progress of the dolomitization reaction of synthetic calcite in various experiments: (<b>A</b>) parallel control experiment (referred to as “without polysaccharide”), (<b>B</b>) agar 0.1 g/L series experiment, (<b>C</b>) CMC 0.1 g/L series experiment, and (<b>D</b>) CMC 0.2 g/L series experiment. Dashed vertical lines mark the boundaries of the three reaction stages, with numbers indicating the total overlapping points at each position. Each graph has three cross-plots: (<b>a</b>) ordering degree vs. reaction time (h), (<b>b</b>) stoichiometry (mol% MgCO<sub>3</sub>) vs. reaction time (h), and (<b>c</b>) percentage of product vs. reaction time (h). In the figures, blue circles denote dolomite abundances below 90%, while orange triangles indicate abundances exceeding 90%. The data points represent multiple experiment batches, so the boundaries of the three stages were established at the point where the characteristics specific to each stage initially appeared. The numbers above or below the points indicate the total number of overlapping points at each position.</p>
Full article ">Figure 1 Cont.
<p>The graphs show the progress of the dolomitization reaction of synthetic calcite in various experiments: (<b>A</b>) parallel control experiment (referred to as “without polysaccharide”), (<b>B</b>) agar 0.1 g/L series experiment, (<b>C</b>) CMC 0.1 g/L series experiment, and (<b>D</b>) CMC 0.2 g/L series experiment. Dashed vertical lines mark the boundaries of the three reaction stages, with numbers indicating the total overlapping points at each position. Each graph has three cross-plots: (<b>a</b>) ordering degree vs. reaction time (h), (<b>b</b>) stoichiometry (mol% MgCO<sub>3</sub>) vs. reaction time (h), and (<b>c</b>) percentage of product vs. reaction time (h). In the figures, blue circles denote dolomite abundances below 90%, while orange triangles indicate abundances exceeding 90%. The data points represent multiple experiment batches, so the boundaries of the three stages were established at the point where the characteristics specific to each stage initially appeared. The numbers above or below the points indicate the total number of overlapping points at each position.</p>
Full article ">Figure 1 Cont.
<p>The graphs show the progress of the dolomitization reaction of synthetic calcite in various experiments: (<b>A</b>) parallel control experiment (referred to as “without polysaccharide”), (<b>B</b>) agar 0.1 g/L series experiment, (<b>C</b>) CMC 0.1 g/L series experiment, and (<b>D</b>) CMC 0.2 g/L series experiment. Dashed vertical lines mark the boundaries of the three reaction stages, with numbers indicating the total overlapping points at each position. Each graph has three cross-plots: (<b>a</b>) ordering degree vs. reaction time (h), (<b>b</b>) stoichiometry (mol% MgCO<sub>3</sub>) vs. reaction time (h), and (<b>c</b>) percentage of product vs. reaction time (h). In the figures, blue circles denote dolomite abundances below 90%, while orange triangles indicate abundances exceeding 90%. The data points represent multiple experiment batches, so the boundaries of the three stages were established at the point where the characteristics specific to each stage initially appeared. The numbers above or below the points indicate the total number of overlapping points at each position.</p>
Full article ">Figure 1 Cont.
<p>The graphs show the progress of the dolomitization reaction of synthetic calcite in various experiments: (<b>A</b>) parallel control experiment (referred to as “without polysaccharide”), (<b>B</b>) agar 0.1 g/L series experiment, (<b>C</b>) CMC 0.1 g/L series experiment, and (<b>D</b>) CMC 0.2 g/L series experiment. Dashed vertical lines mark the boundaries of the three reaction stages, with numbers indicating the total overlapping points at each position. Each graph has three cross-plots: (<b>a</b>) ordering degree vs. reaction time (h), (<b>b</b>) stoichiometry (mol% MgCO<sub>3</sub>) vs. reaction time (h), and (<b>c</b>) percentage of product vs. reaction time (h). In the figures, blue circles denote dolomite abundances below 90%, while orange triangles indicate abundances exceeding 90%. The data points represent multiple experiment batches, so the boundaries of the three stages were established at the point where the characteristics specific to each stage initially appeared. The numbers above or below the points indicate the total number of overlapping points at each position.</p>
Full article ">Figure 2
<p>Percentage of dolomite product plotted against reaction time. Data points from experiments in the without-polysaccharide series are plotted in blue, while data points from experiments in the with-polysaccharide series are plotted in red. Colored dashed vertical lines demarcate the boundaries of the three reaction stages in the respective experiments. In all three plots (<b>a</b>–<b>c</b>), the blue data points are shifted to the left on the reaction time axis compared to the red data points. This indicates that the dolomitization reaction was delayed in the experiments with polysaccharides compared with the parallel control experiments in the without-polysaccharide series. The numbers above or below the points indicate the total number of overlapping points at each position.</p>
Full article ">Figure 3
<p>Cross-plot of the ordering degree against reaction time (h) for the synthetic calcite series.</p>
Full article ">Figure 4
<p>SEM images of reaction products from control experiments without polysaccharides using cleaved calcite. Etch pits on the calcite (104) plane are shown after 120 h (<b>a</b>). Rhombohedral dolomite crystallites at 150 h (<b>b</b>) form on the calcite (104) plane and continue to grow, covering the surface after 250 h (<b>c</b>). EDS spectra were inserted in images (<b>a</b>,<b>b</b>), with the analyzing points denoted by yellow pluses (+).</p>
Full article ">Figure 5
<p>SEM images of reaction products from the 0.1 g/L CMC series (<b>a</b>–<b>d</b>) and 0.2 g/L agar series (<b>e</b>–<b>j</b>) are presented, with each series separately heated for varying durations. In the CMC series, microneedles were observed on the calcite (104) plane after 120 h (<b>a</b>,<b>b</b>), followed by pyramidal hillocks after 150 h (<b>c</b>) and the formation of quadrangle typical micropyramids after 250 h (<b>d</b>). The agar series exhibited tiny typical quadrangle micropyramids, microcone-like protrusions, and large-sized irregular micropyramids with tips after 130 h ((<b>e</b>–<b>g</b>), respectively), and hexagon typical micropyramids after 200 h (<b>h</b>), followed by parallel platy microstructures after 250 h (<b>i</b>), and densely packed hillocks after 350 h (<b>j</b>). The names of the microstructures are based on previous literature [<a href="#B80-minerals-14-00721" class="html-bibr">80</a>,<a href="#B81-minerals-14-00721" class="html-bibr">81</a>,<a href="#B82-minerals-14-00721" class="html-bibr">82</a>,<a href="#B83-minerals-14-00721" class="html-bibr">83</a>].</p>
Full article ">Figure 6
<p>The [110] HRTEM image of an APB in dolomite is shown after being heated for 61 h at 200 °C (<b>a</b>), along with the corresponding electron diffraction (ED) pattern (<b>b</b>). The Fast Fourier Transform (FFT)-filtered image (<b>a</b>) of (<b>c</b>) and its FFT (<b>d</b>) are also presented. Arrows in images (<b>a</b>,<b>c</b>) point to the location of the APB.</p>
Full article ">
15 pages, 5079 KiB  
Article
Effect of Solid Concentration on Particle Size Distribution and Grinding Kinetics in Stirred Mills
by Wang Guo and Keqi Guo
Minerals 2024, 14(7), 720; https://doi.org/10.3390/min14070720 - 17 Jul 2024
Viewed by 432
Abstract
In this study, the evolution behavior of the particle size distribution during the grinding process was examined with fractal theory. According to the distribution index k of the Rosin–Rammler–Benne model, the relationship between the fractal dimension D of the fractal theory and the [...] Read more.
In this study, the evolution behavior of the particle size distribution during the grinding process was examined with fractal theory. According to the distribution index k of the Rosin–Rammler–Benne model, the relationship between the fractal dimension D of the fractal theory and the distribution index k is discussed. The fractal dimension D was used to evaluate the uniformity of the particle size distribution of the grinding product. In addition, the population balance model was used to simulate the breakage behavior of each size interval. The result indicates that the non-first-order model presented a better fitting performance in the breakage behavior of the coarse size and the desired size when compared with the other type of model. It can be found that the breakage rate increased with the solid concentration. However, the breakage distribution function is independent of the solid concentration in this study. These results suggest that the effect of the solid concentration on the fraction of the coarse size broken into the desired size was not significant. Furthermore, the simulated data are discussed and analyzed with the attainable region method as well as the difference in the change rate of the desired size and the overgrinding size. It can be found that to produce a higher fraction of the desired size in the grinding products, the residence time of the material in the mill needs be shortened with a higher solid concentration. Full article
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)
Show Figures

Figure 1

Figure 1
<p>Plot of the coarse size versus the desired size for silica sand with a particle size of 0.60 mm during the batch milling process [<a href="#B16-minerals-14-00720" class="html-bibr">16</a>]. Reproduced with permission from Katubilwa, F.O.M., Powder Technology; published by Elsevier, 2024.</p>
Full article ">Figure 2
<p>Particle size distribution of the feed.</p>
Full article ">Figure 3
<p>The evolution of the cumulative undersize distribution for different grinding times with a solid concentration of 40% (markers = experimental; data lines = regression).</p>
Full article ">Figure 4
<p>The measured and calculated <span class="html-italic">d<sub>e</sub></span> for different solid concentrations.</p>
Full article ">Figure 5
<p>The relationship between the parameter <span class="html-italic">k</span> and the particle size <span class="html-italic">d</span><sub>50.</sub></p>
Full article ">Figure 6
<p>The <span class="html-italic">lg</span>(<span class="html-italic">F</span>)<span class="html-italic">-lg</span>(<span class="html-italic">d</span>) curves for a solid concentration of 40% (lines = regression; markers = experimental data).</p>
Full article ">Figure 7
<p>The fractal dimension versus the particle size <span class="html-italic">d</span><sub>50</sub> under different solid concentrations.</p>
Full article ">Figure 8
<p>Temporal variation in the fraction of the coarse size under different solid concentrations (markers = experimental data; lines = regression). (<b>a</b>) First-order model. (<b>b</b>) Two-component model. (<b>c</b>) Non-first-order model.</p>
Full article ">Figure 9
<p>The values of the Adj. R<sup>2</sup>, RMSE, and SSE obtained by fitting different kinetic equations (means and error bars).</p>
Full article ">Figure 10
<p>Temporal variation in the fraction of the desired size under different solid concentrations (markers = experimental data; lines = regression). (<b>a</b>) First-order model. (<b>b</b>) Non-first-order model.</p>
Full article ">Figure 11
<p>Temporal variation in the fraction of the overgrinding size under different solid concentrations (markers = experimental data; lines = regression). (<b>a</b>) First-order model. (<b>b</b>) Non-first-order model.</p>
Full article ">Figure 12
<p>Construction of the attainable Region. (<b>a</b>) The fraction of the desired size versus the fraction of the coarse size. (<b>b</b>) The fraction of the overgrinding size versus the fraction of the coarse size.</p>
Full article ">Figure 13
<p>The difference in the change rates between the desired size and overgrinding size vs. the grinding times for different solid concentrations.</p>
Full article ">
30 pages, 5946 KiB  
Article
Geochronology, Geochemical Characterization and Tectonic Background of Volcanic Rocks of the Longjiang Formation in the Lengjimanda Plate Area, Middle Da Hinggan Mountains
by Shi-Chang Wang, Yu-Jie Hao, Lu Shi, Zhen Tang and Shuang Zhu
Minerals 2024, 14(7), 719; https://doi.org/10.3390/min14070719 - 16 Jul 2024
Viewed by 381
Abstract
The Lengjimanda plate is situated in the middle section of the Da Hinggan mountains, in the eastern section of the Tianshan Xingmeng orogenic belt. To determine the formation age of the volcanic rocks in the Longjiang formation in this area, to explore their [...] Read more.
The Lengjimanda plate is situated in the middle section of the Da Hinggan mountains, in the eastern section of the Tianshan Xingmeng orogenic belt. To determine the formation age of the volcanic rocks in the Longjiang formation in this area, to explore their origin and tectonic background, and to reconstruct the geodynamic evolution of the region, this study conducted petrological, zircon U–Pb geochronological, geochemical, and isotopic analyses of the volcanic rocks in the Longjiang formation. The Longjiang formation’s volcanic rocks are primarily composed of trachyandesite, trachyte trachydacite, and andesite, which are intermediate basic volcanic rocks. They are enriched in large-ion lithophile elements, are depleted in high-field-strength elements, are significantly fractionated between light and heavy rare earth elements, and exhibit a moderate negative Eu anomaly in most samples. The results of the LA–ICP–MS zircon U–Pb dating indicate that the volcanic rocks in this group were formed in the Early Cretaceous period at 129.1 ± 0.82 Ma. The zircon εHf(t) ranges from +1.13 to +43.77, the tDM2 ranges from +655 to +1427 Ma, the initial Sr ratio (87Sr/86Sr)i ranges from 0.7030 to 0.7036, and the εNd(t) ranges from +2.1 to +6.6. Based on the geochemical compositions and isotopic characteristics of the rocks, the initial magma of the volcanic rocks in the Longjiang formation originated from the partial melting of basaltic crustal materials, with a source material inferred to be depleted mantle-derived young crustal. These rocks were formed in a superimposed post-collisional and continental arc environment, possibly associated with the Mongol-Okhotsk Ocean closure and the oblique subduction of the Pacific plate. This study addresses a research gap regarding the volcanic rocks of the Longjiang formation in this area. Its findings can be applied to exploration and prospecting in the region. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The sample structural map in the Northeast region [<a href="#B40-minerals-14-00719" class="html-bibr">40</a>]; (<b>b</b>) plate geological map of LingjiManda plate, Da Hinggan Mountains [<a href="#B41-minerals-14-00719" class="html-bibr">41</a>]; and (<b>c</b>) geological cross-section view of Lengjimanda plate.</p>
Full article ">Figure 2
<p>Hand specimen photo (<b>a</b>) and microscope photos (<b>b</b>–<b>f</b>) of the Lengjimanda plate. (<b>a</b>) TWG05; (<b>b</b>) TWG01; (<b>c</b>) TWG02; (<b>d</b>) TWG04; (<b>e</b>) TWG05; (<b>f</b>) TWG06, Sa: sanidine; Pl: plagioclase; Mag: magnetite; Afs: alkalifeldspar; Bt: biotite; Qtz: quartz.</p>
Full article ">Figure 3
<p>U–Pb zircons cathodoluminescence images from the volcanic rocks of the Longjiang formation at Lengjimanda plate. Circles indicate the locations of U–Pb analysis on zircons, along with their respective zircon numbers, and age information listed below. (<b>a</b>) TWG01, (<b>b</b>) TWG04, (<b>c</b>) TWG05, (<b>d</b>) TWG06.</p>
Full article ">Figure 4
<p>Diagram of (N<sub>2</sub>O+K<sub>2</sub>O) vs. SiO<sub>2</sub> of volcanic rocks of Longjiang formation [<a href="#B49-minerals-14-00719" class="html-bibr">49</a>].</p>
Full article ">Figure 5
<p>Major elements classification diagrams of Longjiang formation volcanic rocks, Lengjimanda plate. (<b>a</b>) K<sub>2</sub>O vs. SiO<sub>2</sub> diagram [<a href="#B50-minerals-14-00719" class="html-bibr">50</a>]; (<b>b</b>) A/NK vs. A/CNK diagram [<a href="#B50-minerals-14-00719" class="html-bibr">50</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Chondrite-normalized REE distribution patterns [<a href="#B51-minerals-14-00719" class="html-bibr">51</a>] and (<b>b</b>) primitive mantle-normalized spider diagrams [<a href="#B52-minerals-14-00719" class="html-bibr">52</a>] of the Longjiang formation volcanic rocks, Lengjimanda plate.</p>
Full article ">Figure 7
<p>Concordant and weighted diagrams of the zircon U–Pb age of the Longjiang formation volcanic rocks, Lengjimanda plate. Error ellipses are shown for 1-sigma level of uncertainty. (<b>a</b>) TWG01, (<b>b</b>) TWG04, (<b>c</b>) TWG05, (<b>d</b>) TWG06, (<b>e</b>) Longjiang formation volcanic rocks, (<b>f</b>) weighted diagrams of the Longjiang formation volcanic rocks.</p>
Full article ">Figure 8
<p>Illustration of magmatic evolution of volcanic rocks in the Longjiang formation of Lengjimanda plate [<a href="#B65-minerals-14-00719" class="html-bibr">65</a>]. (<b>a</b>) La/Yb vs. La and (<b>b</b>) La/Sm vs. La.</p>
Full article ">Figure 9
<p>Adakite discrimination diagram of Longjiang formation volcanic rocks, Lengjimanda plate. (<b>a</b>) Sr/Y vs. Y diagram; (<b>b</b>) (La/Yb) <sub>N</sub> vs. (Yb) n diagram [<a href="#B71-minerals-14-00719" class="html-bibr">71</a>].</p>
Full article ">Figure 10
<p><span class="html-italic">ε</span><sub>Hf</sub> (<span class="html-italic">t</span>) vs. age diagram of Longjiang formation volcanic rocks, Lengjimanda plate. East CAOB: East Asian Orogenic Belt Eastern part [<a href="#B21-minerals-14-00719" class="html-bibr">21</a>].</p>
Full article ">Figure 11
<p><span class="html-italic">ε</span><sub>Nd</sub>(t) vs. <span class="html-italic">ε</span><sub>Sr</sub>(t) diagram of volcanic rocks in the Longjiang formation of Lengjimanda plate.</p>
Full article ">Figure 12
<p>Pb isotope composition diagram of Longjiang formation volcanic rocks, Lengjimanda plate. Orogene: Orogenic belt. ((<b>a</b>), [<a href="#B83-minerals-14-00719" class="html-bibr">83</a>]; (<b>b</b>,<b>c</b>), [<a href="#B84-minerals-14-00719" class="html-bibr">84</a>]; (<b>d</b>), [<a href="#B80-minerals-14-00719" class="html-bibr">80</a>]).</p>
Full article ">Figure 13
<p>Structural identification diagram of volcanic rocks in Longjiang formation, Lengjimanda plate [<a href="#B107-minerals-14-00719" class="html-bibr">107</a>]. (<b>a</b>) Intraplate environment (WIP), (<b>b</b>) continental arc environment (CAP), post-collision environment (PAP), initial ocean arc environment (IOP), late ocean arc environment (LOP).</p>
Full article ">
30 pages, 15303 KiB  
Article
Discovery and Exploration of the Luming Porphyry Mo Deposit, Northeastern China: Implications for Regional Prospecting
by Bangfei Gao, Minghua Dong, Hui Xie, Zhiliang Liu, Yihang Li and Tong Zhou
Minerals 2024, 14(7), 718; https://doi.org/10.3390/min14070718 - 16 Jul 2024
Viewed by 548
Abstract
Over the past two decades, significant deposit discoveries were made in Northeastern China, including the super-large Chalukou, Daheishan, and Luming porphyry Mo deposits. The discovery of the Luming deposit was accomplished through verification of stream sediment anomalies, with mineralization closely associated with early [...] Read more.
Over the past two decades, significant deposit discoveries were made in Northeastern China, including the super-large Chalukou, Daheishan, and Luming porphyry Mo deposits. The discovery of the Luming deposit was accomplished through verification of stream sediment anomalies, with mineralization closely associated with early Jurassic monzogranite and granite porphyry. Previous studies primarily focused on the mineralization mechanisms of these deposits without adequately addressing the exploration methods and prospecting criteria. This study involved a comprehensive re-evaluation of geological observations, analysis of rock primary halo, gravity and magnetic surveys, and induced polarization surveys conducted during exploration campaigns at the Luming porphyry Mo deposit. The results suggest that hydrothermal breccias play a critical role in controlling the mineralization by forming a central low-grade core within the deposit while the Mo mineralization and hydrothermal alteration exhibit a donut-shaped distribution around it. The primary halo shows a distinct metal zonation moving from a central W-Bi-Mo-(Sb) to a peripheral Cu-Co-Ni and a distal Pb-Zn-Ag-In. The mineralization zone exhibits a low Bouguer gravity anomaly, negative magnetic anomaly, medium to low resistivity, and moderate to high chargeability, indicating the effectiveness of geophysical methods in defining the extent of the ore body. The Luming porphyry Mo deposit and distal skarn-epithermal Pb-Zn mineralization are parts of a porphyry-related magmatic-hydrothermal system. The results of this study offer valuable insights into the genesis of porphyry Mo deposits and their implications for prospecting in the forested region of Northeastern China. Full article
Show Figures

Figure 1

Figure 1
<p>Geological map of the Lesser Xing’an Range–Zhangguangcai Range Metallogenic Belt (LXRZRMB) in Northeastern China and the distribution of Jurassic Mo deposits; the Luming porphyry Mo deposit is located in the northern section of the LXRZRMB [<a href="#B5-minerals-14-00718" class="html-bibr">5</a>,<a href="#B16-minerals-14-00718" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>(<b>A</b>) Regional aeromagnetic anomaly and distribution of mineral deposits in the Yichun-Tieli area. (<b>B</b>) Regional aerogravity anomaly and mineral deposit distribution in the Yichun-Tieli area. The 1:1,000,000 aeromagnetic anomaly data are derived from [<a href="#B60-minerals-14-00718" class="html-bibr">60</a>], while the mineral deposit data are sourced from [<a href="#B41-minerals-14-00718" class="html-bibr">41</a>,<a href="#B61-minerals-14-00718" class="html-bibr">61</a>,<a href="#B62-minerals-14-00718" class="html-bibr">62</a>]. The 1:1,000,000 aerogravity anomaly data are based on [<a href="#B63-minerals-14-00718" class="html-bibr">63</a>]. 1 = Luming Mo(Cu); 2 = Xulaojiugou Pb-Zn; 3 = Cuiling Mo-Au; 4 = Kunlunqi Pb-Zn; 5 = Qianjin Dongshan Pb-Zn; 6 = Xiling Nanshan Pb-Zn; 7 = Xiaoxilin Pb-Zn; 8 = Fengmao Fe; 9 = Mantoushan Pb-Zn; 10 = Milin Cu-Fe; 11 = Wuxing Pb-Zn; 12 = Pingdingshan Au; 13 = Lianzhushan Au; 14 = Shouhushan Cu-Pb-Zn; 15 = Kubing Pb-Zn; 16 = Cuihongshan Mo-Fe-W-Pb-Zn; 17 = Huojihe Mo; 18 = Qianjin Nanshan Mo; 19 = Xilinshi’er Fe; 20 = Daxilin Fe; 21 = Hongqi Mo; 22 = Kuyuan Fe; 23 = Hongtieshan Fe-Mo; 24 = Yongxu Mo; 25 = Hongqishan Fe; 26 = Cuibei Fe; 27 = 712 Highland (712 HL).</p>
Full article ">Figure 3
<p>Regional stream sediment anomalies at scale of 1:50,000 show a zoning with W-Mo-Cu and Pb-Zn-Ag-As-Sb, modified after [<a href="#B41-minerals-14-00718" class="html-bibr">41</a>,<a href="#B44-minerals-14-00718" class="html-bibr">44</a>]. Two Mo ore bodies were discovered at the NW corner of the prospecting area during 2003–2005 (gray rectangle), offering valuable guidance for the following exploration campaigns during 2005–2011(red rectangle).</p>
Full article ">Figure 4
<p>The layout depicts diamond drilling and ground geophysical surveys, with the delineation of ore body outline following drilling exploration, modified after [<a href="#B68-minerals-14-00718" class="html-bibr">68</a>]. 1 = Quaternary; 2 = Indosinian-Yanshanian intrusive rocks; 3 = low-grade ores (0.03% ≤ Mo &lt; 0.06%); 4 = economic ores (Mo ≥ 0.06%); 5 = drillings during general exploration; 6 = verification drillings after general exploration; 7 = drillings of resource verification stage; 8 = drillings of complementary exploration stage; 9 = exploration lines and numbers; 10 = locations of geophysical surveys. Gravity and magnetic surveys were conducted at A–A’, B–B’, C–C’, and D–D’, while induced polarization surveys were carried out at A–A’, E–E’, F–F’, and G–G’.</p>
Full article ">Figure 5
<p>Geological map of the Luming PMD, modified after [<a href="#B68-minerals-14-00718" class="html-bibr">68</a>]. 1 = Quaternary; 2 = F3 fault; 3 = K-feldspar granite; 4 = hydrothermal breccias; 5 = granite porphyry; 6 = fine-grained granite; 7 = biotite monzogranite; 8 = monzogranite; 9 = economic Mo ores; 10 = low-grade Mo ores; 11 = alteration zones; ② = silicification-illitization-K-feldsparization; ③ = silicification-illitization-chloritization; ④ = silicification-pyritization-chloritization.</p>
Full article ">Figure 6
<p>Geological cross section of the 04N exploration line at Luming PMD, modified after [<a href="#B68-minerals-14-00718" class="html-bibr">68</a>]. 1 = hydrothermal breccias; 2 = granite porphyry; 3 = monzogranite; 4 = economic Mo ores; 5 = low-grade Mo ores; 6 = alteration zones; ① = silicification-K-feldsparization-biotitization; ② = silicification-illitization-K-feldsparization; ③ = silicification-illitization-chloritization; ④ = silicification-pyritization-chloritization.</p>
Full article ">Figure 7
<p>Intrusive rocks and their alteration characteristics of the Luming PMD. (<b>A</b>) Monzogranite exhibits the silicification-K-feldsparization-biotitization overlain by silicification-illitization-chloritization. (<b>B</b>) Fine-grained granite intruded into the monzogranite, which underwent silicification-kaolinization while enveloping the earlier silicification-K-feldsparization. (<b>C</b>) Fine-grained granite encloses breccias of monzogranite. The fine-grained granite experienced silicification-pyrite-chloritization while monzogranite underwent silicification-K-feldsparization-biotitization. Both were superposed by late-stage veinlet silicification. (<b>D</b>) Granite porphyry was intruded into monzogranite; the former experienced strong silicification with residual plagioclase phenocrysts, while the latter underwent early silicification-K-feldsparization-biotitization and late veinlet silicification. (<b>E</b>) Granite porphyry, characterized by plagioclase phenocrysts, exhibited veinlet Mo mineralization. (<b>F</b>) K-feldspar granite underneath F3 fault without any mineralization. (<b>G</b>) F3 fault developed at the contact zone between the silicification-illitization monzogranite (grey) and the K-feldspar granite (red). (<b>H</b>) Mineralized hydrothermal breccias are composed of clasts of altered monzogranite and a cement matrix comprising silica and molybdenite. (<b>I</b>) Barren hydrothermal breccias comprise clasts made up of altered monzogranite within a silica-rich cement matrix. Qz = quartz; Kfs = K-feldspar; Pl = plagioclase; Bi = biotite; ill = illite; Kaol = kaolinite; Chl = chlorite; Py = pyrite; Mol = molybdenite; ηγ = monzogranite; γ = fine-grained granite; γπ = granite porphyry; κγ = K-feldspar granite; HB = hydrothermal breccias.</p>
Full article ">Figure 8
<p>The 3D geological model of the Luming PMD indicates a low strip ratio with a bowl-shaped structure, where intercalated rock can be separately delineated within the low-grade core, which is mainly composed of hydrothermal breccias.</p>
Full article ">Figure 9
<p>The occurrence of molybdenite at Luming PMD exhibits a multi-phase mineralization. (<b>A</b>) Molybdenite is found in clusters within quartz-molybdenite veins. (<b>B</b>) Molybdenite fills fractures in the monzogranite (D-type veinlet). (<b>C</b>) Molybdenite occurs as quartz-molybdenite veins (early phase, A-type veinlet) and pyrite-molybdenite veins (late phase, D-type veinlet). (<b>D</b>) Molybdenite in a radial aggregate. (<b>E</b>) Tabular crystals of molybdenite. (<b>F</b>) Molybdenite intergrowths with pyrite. (<b>G</b>) Molybdenite fills the cracks between mineral grains. (<b>H</b>) The paragenesis of molybdenite with pyrite and chalcopyrite. (<b>I</b>) Intergrowth of molybdenite and chalcopyrite. (<b>J</b>) Late-phase bornite and galena fill the spaces between early-phase molybdenite particles. (<b>K</b>) Molybdenite encases early-phase chalcopyrite. (<b>L</b>) The late-phase bornite-chalcopyrite solid solution fills the spaces between the early-phase molybdenite particles. (<b>D</b>–<b>L</b>) were observed under reflected light; Qz = quartz; Mol = molybdenite; Py = pyrite; Cpy = chalcopyrite; Ga = galena; Sph = sphalerite.</p>
Full article ">Figure 10
<p>Characteristics of polymetallic mineralization under reflected light at Luming PMD. (<b>A</b>) Chalcopyrite (late phase) fills fractures of early pyrite. (<b>B</b>) Fractured pyrite was replaced by late-phase chalcopyrite, with oxidized edges forming blue chalcocite. (<b>C</b>) Late-phase pyrite replaces rutile encasing early ilmenite. (<b>D</b>) Chalcopyrite (late phase) replaces early pyrrhotite, forming a bay contact. (<b>E</b>) Chalcopyrite encases early pyrrhotite. (<b>F</b>) Chalcopyrite is dispersed in sphalerite. (<b>G</b>) Sphalerite is dispersed in chalcopyrite. (<b>H</b>) Chalcopyrite fills fractures of early pyrite cutting veinlets formed by rutile and ilmenite. (<b>I</b>) Chalcopyrite replaces rutile, both encasing early euhedral pyrite. (<b>J</b>) Late galena envelops early chalcopyrite. (<b>K</b>) Late galena replaces early sphalerites-chalcopyrite. (<b>L</b>) Titanite replaces early ilmenite. Ilm = ilmenite; Rt = rutile; Ttn = titanite; Po = pyrrhotite; Mol = molybdenite; Py = pyrite; Cpy = chalcopyrite; Ga = galena; Sph = sphalerite; Bcc = blue chalcocite.</p>
Full article ">Figure 11
<p>Alteration characteristics of the Luming PMD, showing a general evolution sequence of K-feldsparization-biotitization→illitization→chloritization→carbonatization, with silicification and pyritization throughout the hydrothermal processes. (<b>A</b>) Late calcite veins enclose the early K-feldsparization-biotitization monzogranite. (<b>B</b>) Potassic alteration characterized by pervasive K-feldsparization and veinlet biotitization. (<b>C</b>) Early silicification-K-feldsparization-biotitization overlain by late silicification-illitization-chloritization. (<b>D</b>) Quartz-molybdenite fills the early silicification-K-feldsparization monzogranite, which is cut by late silicification-molybdenite veinlets. (<b>E</b>) Early silicification-K-feldsparization ± biotitization overlain by late silicification-molybdenite veinlets shows brecciated structure. (<b>F</b>) Late silicification-pyritization intersects early silicification-K-feldsparization. (<b>G</b>) Early pyrite-calcite veins enclose early silicification-K-feldsparization. (<b>H</b>) Polysynthetic twinning plagioclase is replaced by perthite and quartz. (<b>I</b>) Veinlet biotitization. (<b>J</b>) Late quartz-calcite vein cut-through altered rocks of silicification-illitization-chloritization. Qz = quartz; Kfs = K-fedlspar; Pl = plagioclase; Bi = biotite; ill = illite; Chl = chlorite; Cc = calcite; Py = pyrite; Mol = molybdenite.</p>
Full article ">Figure 12
<p>Tentative timeframe for intrusion, brecciation, alteration, and mineralization of the Luming PMD, indicating multi-phase hydrothermal processes. Qz = quartz; Cc = calcite; Py = pyrite; Mol = molybdenite; Cpy = chalcopyrite; Ga = galena; Sph = sphalerite.</p>
Full article ">Figure 13
<p>Classification of the Luming monzogranite and granite porphyry samples. (<b>A</b>) (Na<sub>2</sub>O + K<sub>2</sub>O) vs. SiO<sub>2</sub> [<a href="#B69-minerals-14-00718" class="html-bibr">69</a>]; (<b>B</b>) K<sub>2</sub>O vs. SiO<sub>2</sub> [<a href="#B70-minerals-14-00718" class="html-bibr">70</a>]; (<b>C</b>) A/NK vs. A/CNK [<a href="#B71-minerals-14-00718" class="html-bibr">71</a>]; (<b>D</b>) Fe<sub>2</sub>O<sub>3</sub>/FeO vs. SiO<sub>2</sub> diagrams [<a href="#B2-minerals-14-00718" class="html-bibr">2</a>]. Major element data are derived from [<a href="#B22-minerals-14-00718" class="html-bibr">22</a>,<a href="#B46-minerals-14-00718" class="html-bibr">46</a>,<a href="#B47-minerals-14-00718" class="html-bibr">47</a>,<a href="#B49-minerals-14-00718" class="html-bibr">49</a>].</p>
Full article ">Figure 14
<p>Primitive mantle-normalized trace element (<b>A</b>) and chondrite-normalized REE (<b>B</b>) spider diagrams for the Luming monzogranite and granite porphyry samples. Data for normalization are from [<a href="#B72-minerals-14-00718" class="html-bibr">72</a>]. Trace element data are derived from [<a href="#B22-minerals-14-00718" class="html-bibr">22</a>,<a href="#B46-minerals-14-00718" class="html-bibr">46</a>,<a href="#B47-minerals-14-00718" class="html-bibr">47</a>,<a href="#B49-minerals-14-00718" class="html-bibr">49</a>].</p>
Full article ">Figure 15
<p>The zoning characteristics of the primary halo at Luming PMD. (<b>A</b>) W. (<b>B</b>) Bi. (<b>C</b>) Mo (<b>D</b>) Cu. (<b>E</b>) Co. (<b>F</b>) Ni. (<b>G</b>) Zn. (<b>H</b>) Ag. (<b>I</b>) Pb. (<b>J</b>) Sb. (<b>K</b>) In. (<b>L</b>) Section 00E. Qz-Kfs-Bi = silicification-K-feldsparization-biotitization; Qz-Ill-Kfs = silicification-illitization-K-feldsparization; Qz-Ill-Chl = silicification-illitization-chloritization; Qz-Py-Chl = silicification-pyritization-chloritization; ηγ = monzogranite; γπ = granite porphyry; HB=hydrothermal breccias; F3 = F3 fault.</p>
Full article ">Figure 16
<p>Gravity and magnetic anomalies of the Luming PMD. (<b>A</b>) Contour map showing Bouguer gravity anomalies. (<b>B</b>) Contour map showing magnetic anomalies. (<b>C</b>) Cross section showing gravity and magnetic anomalies along C–C’. Qz-Kfs-Bi = silicification-K-feldsparization-biotitization; Qz-Ill-Kfs = silicification-illitization-K-feldsparization; Qz-Ill-Chl = silicification-illitization-chloritization; Qz-Py-Chl = silicification-pyritization-chloritization; Q = Quaternary; ηγ = monzogranite; βγ = biotite granite; γ = fine-grained granite; γπ = granite porphyry; κγ = K-feldspar granite; HB = hydrothermal breccias; F3 = F3 fault.</p>
Full article ">Figure 17
<p>Induced polarization anomalies of the Luming PMD. (<b>A</b>) Contour map shows resistivity anomalies. (<b>B</b>) Contour map shows chargeability anomalies. (<b>C</b>) Profile shows induced polarization anomalies along A–A’. Qz-Kfs-Bi =silicification-K-feldspathization-biotitization; Qz-Ill-Kfs = silicification-illitization-K-feldspathization; Qz-Ill-Chl = silicification-illitization-chloritization; Qz-Py-Chl = silicification-pyritization-chloritization; Q = Quaternary; ηγ = monzogranite; βγ = biotite granite; γ = fine-grained granite; γπ = granite porphyry; κγ = K-feldspar granite; HB = hydrothermal breccias; F3 = F3 fault.</p>
Full article ">Figure 18
<p>3D geochemical model of the Luming PMD showing the influence of hydrothermal breccias on the distribution of mineralization. (<b>A</b>) Mineralization at 400 m elevation. (<b>B</b>) Mineralization at 100 m elevation. (<b>C</b>) Mineralization at 03N, 04N, and 12N cross sections. (<b>D</b>) Mineralization at 04E, 00E, and 03E cross sections. γπ = granite porphyry; HB = hydrothermal breccias.</p>
Full article ">
30 pages, 15297 KiB  
Article
Geochronology and Geochemistry of Paleoproterozoic Mafic Rocks in Northern Liaoning and Their Geological Significance
by Jingsheng Chen, Yi Tian, Zhonghui Gao, Bin Li, Chen Zhao, Weiwei Li, Chao Zhang and Yan Wang
Minerals 2024, 14(7), 717; https://doi.org/10.3390/min14070717 - 16 Jul 2024
Viewed by 454
Abstract
Petrological, geochronological, and geochemical analyses of mafic rocks in northern Liaoning were conducted to constrain the formation age of the Proterozoic strata, and to further study the source characteristics, genesis, and tectonic setting. The mafic rocks in northern Liaoning primarily consist of basalt, [...] Read more.
Petrological, geochronological, and geochemical analyses of mafic rocks in northern Liaoning were conducted to constrain the formation age of the Proterozoic strata, and to further study the source characteristics, genesis, and tectonic setting. The mafic rocks in northern Liaoning primarily consist of basalt, diabase, gabbro, and amphibolite. Results of zircon U-Pb chronology reveal four stages of mafic magma activities in northern Liaoning: the first stage of basalt (2209 ± 12 Ma), the second stage of diabase (2154 ± 15 Ma), the third stage of gabbro (2063 ± 7 Ma), and the fourth stage of magmatic protolith of amphibolite (2018 ± 13 Ma). Combined with the unconformity overlying Neoproterozoic granite, the formation age of the Proterozoic strata in northern Liaoning was found to be Paleoproterozoic rather than Middle Neoproterozoic by the geochronology of these mafic rocks. A chronological framework of mafic magmatic activities in the eastern segment of the North China Craton (NCC) is proposed. The mafic rocks in northern Liaoning exhibit compositional ranges of 46.39–50.33 wt% for SiO2, 2.95–5.08 wt% for total alkalis (K2O + Na2O), 6.17–7.50 wt% for MgO, and 43.32–52.02 for the Mg number. TiO2 contents lie between 1.61 and 2.39 wt%, and those of MnO between 0.17 and 0.21 wt%. The first basalt and the fourth amphibolite show low total rare earth element contents. Normalized against primitive mantle, they are enriched in large ion lithophile elements (Rb, Ba, K), depleted in high field strength elements (Th, U, Nb, Ta, Zr, Ti), and exhibit negative anomalies in Sr and P, as well as slight positive anomalies in Zr and Hf. The second diabase and the third gabbro have similar average total rare earth element contents. The diabase shows slight negative Eu anomalies (Eu/Eu* = 0.72–0.88), enrichment in large ion lithophile elements (Ba), depletion in Rb, and slight positive anomalies in high field strength elements (Th, U, Nb, Ta, Zr, Hf, Ti), with negative anomalies in K, Sr, and P. The gabbro is enriched in large ion lithophile elements (Rb, Ba, K), depleted in high field strength elements (Th, U, Nb, Ta, Zr, Hf), and exhibits positive anomalies in Eu (Eu/Eu* = 1.31–1.37). The contents of Cr, Co, and Ni of these four stages of mafic rocks are higher than those of N-MORB. The characteristics of trace element ratios indicate that the mafic rocks belong to the calc-alkaline series and originate from the transitional mantle. During the process of magma ascent and emplacement, it is contaminated by continental crustal materials. There are residual hornblende and spinel in the magma source of the first basalt. The other three magma sources contain residual garnet and spinel. The third gabbro was formed in an island arc environment, and the other three stages of mafic rocks originated from the Dupal OIB and were formed in an oceanic island environment. The discovery of mafic rocks in northern Liaoning suggests that the Longgang Block underwent oceanic subduction and extinction in both the north and south in the Paleoproterozoic, indicating the possibility of being in two different tectonic domains. Full article
Show Figures

Figure 1

Figure 1
<p>The tectonic location (<b>a</b>) (modified after Zhao, 2005 [<a href="#B1-minerals-14-00717" class="html-bibr">1</a>]); simplified geological map and sample location (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 2
<p>Occurrence and micro-pictures of mafic rocks from northern Liaoning. (<b>a</b>,<b>b</b>,<b>c</b>): D1917 basalt; (<b>a</b>): pillow-shaped basalt; (<b>b</b>): edge and center phases of pillow-shaped basalt; (<b>c</b>): microscopic characteristics of basalt; (<b>d</b>–<b>f</b>): D1918 diabase; (<b>d</b>): field occurrence of diabase dyke; (<b>e</b>): specimen of diabase; (<b>f</b>): microscopic characteristics of diabase; (<b>g</b>–<b>i</b>): D1919 gabbro; (<b>g</b>): gabbro intruded into marble; (<b>h</b>): spherical weathering of gabbro; (<b>i</b>): microscopic characteristics of gabbro; (<b>j</b>–<b>l</b>): D2012 amphibolite; (<b>j</b>): field occurrence of amphibolite; (<b>k</b>): amphibolite wrapped in marble in a pillow shape; (<b>l</b>): microscopic characteristics of amphibolite; Pl: plagioclase; Px: pyroxene; Ol: olivine; Hbl: hornblende; Spn: sphene; Ep: epidote.</p>
Full article ">Figure 3
<p>Micro-pictures of zircons in the mafic rocks from northern Liaoning. (<b>a</b>,<b>b</b>): D1917 zircon in pillow-shaped basalt; (<b>c</b>,<b>d</b>): D1918 zircon in diabase; (<b>e</b>,<b>f</b>): D1919 zircon in gabbro; (<b>g</b>,<b>h</b>): D2012 zircon in amphibolite; Zr: zircon.</p>
Full article ">Figure 4
<p>Cathodoluminescence (CL) images of the selected zircons from the mafic rocks in northern Liaoning. The circles on zircons represent analyzed spots.</p>
Full article ">Figure 5
<p>Chondrite-normalized REE distribution diagrams for different zircons from mafic rocks in northern Liaoning.</p>
Full article ">Figure 6
<p>Concordia diagrams for zircons analysed from mafic rocks in the northern Liaoning.</p>
Full article ">Figure 7
<p>SiO<sub>2</sub> vs. total alkali (Na<sub>2</sub>O + K<sub>2</sub>O) ((<b>a</b>), after [<a href="#B51-minerals-14-00717" class="html-bibr">51</a>]), R1 vs. R2 ((<b>b</b>), after [<a href="#B52-minerals-14-00717" class="html-bibr">52</a>]), SiO<sub>2</sub>-K<sub>2</sub>O ((<b>c</b>), after [<a href="#B53-minerals-14-00717" class="html-bibr">53</a>]) and Ta/Yb vs. Ce/Yb ((<b>d</b>), after [<a href="#B54-minerals-14-00717" class="html-bibr">54</a>]) diagrams for mafic rocks from northern Liaoning. (<b>b</b>): 1—alkaline gabbro (alkaline basalt); 2—olivine gabbro (olivine basalt); 3—gabbro norite (tholeiite); 4—syenite gabbro (trachyte basalt); 5—monzonite gabbro (andesite coarse basalt); 6—gabbro (basalt); 7—trachyandesite (syenite); 8—monzonite (andesite); 9—monzodiorite (trachyte); 10—diorite (andesite); 11—nepheline syenite (trachyte phonolite); 12—syenite (trachyte); 13—quartz syenite (quartz trachyte); 14—quartz monzonite (quartz andesite); 15—tonalite (dacite); 16—alkaline granite (alkaline rhyolite); 17—syenogranite (rhyolite); 18—monzogranite (dacite rhyolite); 19—granodiorite (rhyolite dacite); 20—essenite aegirine gabbro; 21—peridotite (picrite); 22—nepheline (picrite nepheline); 23—qilieyan (basanite); 24—neonite (nepheline); 25—essenite; 26—nepheline syenite (phonolite).</p>
Full article ">Figure 8
<p>Chondrite-normalized rare earth element patterns (<b>a</b>) and primitive mantle-normalized trace element spider diagram (<b>b</b>) for the mafic rocks from northern Liaoning. (The values of chondrite and primitive mantle are from [<a href="#B55-minerals-14-00717" class="html-bibr">55</a>]).</p>
Full article ">Figure 9
<p>Age histogram of mafic rocks from the eastern segment of the NCC.</p>
Full article ">Figure 10
<p>Source characteristics of the Triassic gabbro from the Kaiyuan Area. (<b>a</b>) After [<a href="#B69-minerals-14-00717" class="html-bibr">69</a>], DEP—depleted mantle, EN—enriched mantle, N-MORB—normal mid-ocean ridge basalt, PM—primitive mantle, REC—recycled plate, UC—upper crust. (<b>b</b>,<b>c</b>) After [<a href="#B70-minerals-14-00717" class="html-bibr">70</a>], Grt—garnet, SP—spinel. (<b>c</b>) Cpx—clinopyroxene. (<b>d</b>) After [<a href="#B71-minerals-14-00717" class="html-bibr">71</a>]. (<b>e</b>,<b>f</b>) After [<a href="#B72-minerals-14-00717" class="html-bibr">72</a>].</p>
Full article ">Figure 11
<p>Identification diagram of tectonic setting for the mafic rocks from northern Liaoning. (<b>a</b>) Hf/3 versus Th versus Nb/16 (after [<a href="#B83-minerals-14-00717" class="html-bibr">83</a>]); (<b>b</b>) Nb/Yb versus Th/Yb diagram (after [<a href="#B83-minerals-14-00717" class="html-bibr">83</a>]).</p>
Full article ">
17 pages, 10314 KiB  
Article
Investigating Dynamic Behavior in SAG Mill Pebble Recycling Circuits: A Simulation Approach
by Haijie Li, Gauti Asbjörnsson, Kanishk Bhadani and Magnus Evertsson
Minerals 2024, 14(7), 716; https://doi.org/10.3390/min14070716 - 16 Jul 2024
Viewed by 420
Abstract
The dynamics of milling circuits, particularly those involving Semi-Autogenous Grinding (SAG) mills, are not adequately studied, despite their critical importance in mineral processing. This paper aims to investigate the dynamic behavior of an SAG mill pebble recycling circuit under varying feed ore conditions, [...] Read more.
The dynamics of milling circuits, particularly those involving Semi-Autogenous Grinding (SAG) mills, are not adequately studied, despite their critical importance in mineral processing. This paper aims to investigate the dynamic behavior of an SAG mill pebble recycling circuit under varying feed ore conditions, focusing on both uncontrollable parameters (such as ore hardness) and controllable parameters (including circuit layout and pebble crusher configurations). The study is carried out with Simulink dynamic simulations. Our findings reveal several key insights. Firstly, plant designs based solely on static simulations may not be adequate for large or complex circuits, as they fail to account for the dynamic nature of milling processes. Second, incorporating stockpiles after pebble crushing can effectively mitigate the impact of dynamic fluctuations, leading to more stable circuit performance. Third, different circuit layouts can facilitate easier maintenance and operational flexibility. Notably, finer pebble crushing can enhance circuit throughput by 5% to 10%. Full article
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)
Show Figures

Figure 1

Figure 1
<p>Schematic of data structure in the SAG mill simulation.</p>
Full article ">Figure 2
<p>Product PSD of different cone crusher settings. Data from Evertsson [<a href="#B8-minerals-14-00716" class="html-bibr">8</a>], tested on a Svedala Hydrocone cone crusher.</p>
Full article ">Figure 3
<p>The capacity of cone crushers with different settings. Data from Lindqvist [<a href="#B9-minerals-14-00716" class="html-bibr">9</a>], tested on an FLS Raptor 900 cone crusher.</p>
Full article ">Figure 4
<p>Flowsheet of a typical SAG mill pebble crusher circuit with options adding pre-crushers, stockpiles and other circuit configurations. These different layouts will be simulated using a DoE approach.</p>
Full article ">Figure 5
<p>Simulation setup in Simulink involves two identical SAG mills paired with two pebble crushers. The highlighted signals are shown in the simulation results.</p>
Full article ">Figure 6
<p>The simulation results from Run1 to Run4 include the circuit <b>with stockpile and fine crusher settings</b>. Figure (<b>a</b>) shows the throughput, fresh feed rate, and recycle rate of each SAG mill. Figure (<b>b</b>) shows the pebble crusher utilization rate. Figure (<b>c</b>) shows the pebble rate and crusher bin level.</p>
Full article ">Figure 7
<p>Normal plots to assess the significance of effects in the DoE results. (<b>a</b>) is the fresh feed rate, (<b>b</b>) is the pebble rate, (<b>c</b>,<b>d</b>) are the crusher capacity utilization. The blue line has a slope equal to Lenth’s PSE, and the red line slope is 1.</p>
Full article ">Figure A1
<p>The SAG mill model used in this study was validated and calibrated using data from a copper mine. This plot shows the simulation data versus the plant survey data. It should be noted that the results shown in this plot do not represent a specific simulation in the DoE, but the selection function parameters and hard ore A × b values are used.</p>
Full article ">Figure A2
<p>Fresh feed size distribution used in the simulation. The fine feed has a F80 = 50 mm and coarse feed F80 = 84 mm.</p>
Full article ">Figure A3
<p>Simulation results of the SAG mill product size P80, transfer size T80 (undersize) and pebble size P80 (oversize).</p>
Full article ">Figure A4
<p>The simulation results from Run 5 to Run 8 include the circuit with stockpile and coarse crusher settings.</p>
Full article ">Figure A5
<p>The simulation results from Run 9 to Run 12 include the circuit with NO stockpile and coarse crusher settings.</p>
Full article ">Figure A6
<p>The simulation results from Run 13 to Run 16 include the circuit with NO stockpile and with fine crusher settings.</p>
Full article ">
11 pages, 2013 KiB  
Article
In Situ XRD Measurement for High-Pressure Iron in Laser-Driven Off-Hugoniot State
by Liang Sun, Hao Liu, Xiaoxi Duan, Huan Zhang, Zanyang Guan, Weimin Yang, Xiaokang Feng, Youjun Zhang, Yulong Li, Sanwei Li, Dong Yang, Zhebin Wang, Jiamin Yang, Jin Liu, Wenge Yang, Toshimori Sekine and Zongqing Zhao
Minerals 2024, 14(7), 715; https://doi.org/10.3390/min14070715 - 15 Jul 2024
Viewed by 476
Abstract
The investigation of iron under high pressure and temperatures is crucial to understand the Earth’s core structure and composition and the generation of magnetic fields. Here, we present new in situ XRD measurements for iron in an off-Hugoniot state by laser-driven ramp compression [...] Read more.
The investigation of iron under high pressure and temperatures is crucial to understand the Earth’s core structure and composition and the generation of magnetic fields. Here, we present new in situ XRD measurements for iron in an off-Hugoniot state by laser-driven ramp compression at pressure of 200–238 GPa. The lattice parameters for the hexagonal (hcp)-Fe phase and the c/a ratios were obtained to compare them with previous static and dynamical data, which provides the direct confirmation of such parameters via the different compression paths and strain rates. This work indicates that laser ramp compression can be utilized to provide crystal structure information and direct key information on the crystal structure of Fe at the ultrahigh pressure–temperature conditions relevant for planetology. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup, target configuration and laser profiles in laser ramp compression experiments. (<b>a</b>) Dynamical compression with in situ X-ray diffraction measurement was created by nanosecond lasers. Several beams irradiated the Fe foil to generate the monochromatic X-ray flash for XRD measurement. The XRD snapshot was documented by image plates in the steel diagnostic box when a series of laser pulses drove the sample in the center of box to achieve an extremely high-pressure state. (<b>b</b>) Dynamical compression for the diamond–Fe–diamond layer target with a specific thickness was generated by drive laser beams. A line-imaging velocimetry (VISAR) device monitored the diamond window to reconstruct the off-Hugoniot pressure history of iron under dynamical compression. (<b>c</b>) Laser profile in shot 092. The main driven laser for the sample was the ~12 ns pulse shape (purple curve) with a gradually increasing intensity as ramp compression. A 500 ps square pulse with a 1-nanosecond initial pre-pulse (red curve) at the proper relative time was adopted to create X-ray source. (<b>d</b>) The detected Fe X-ray source emission in shot 092.</p>
Full article ">Figure 2
<p>The X-ray diffraction patterns of iron observed in shot 092 (<b>a</b>) and shot 093 (<b>b</b>). The diffraction patterns were projected into the 2<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>−</mo> <mi mathvariant="normal">∅</mi> </mrow> </semantics></math> plane after subtracting the background. The blue dashed lines show the peak positions of the reference Pt at ambient conditions for calibration. The red dashed lines show the peaks from compressed Fe as indexed (<span class="html-italic">h l k</span>) for hcp-Fe. The peaks marked by stars (☆) are ghosts from IP faults and background noise.</p>
Full article ">Figure 3
<p>The constructed compression state for iron at high pressure in shot 092. (<b>a</b>) The laser pulses indicate that the X-ray diffraction recording time at ~14 ns was near the time for peak compression. (<b>b</b>) The VISAR data recorded the extracted free surfaces of diamond windows with experimental uncertainties (red, black, green curves). The velocity history shows that the first velocity jump at ~1.5 km/s is followed by ramp loading in ~5 ns to peak compression velocity ~7.5 km/s, and the single-crystal diamond becomes opaque under ramp compression. (<b>c</b>) The constructed stress map in the C-Fe-C target as a function of time by characteristic calculation. The horizontal dashed lines show the boundaries of the diamond ablator, iron sample, and diamond window in the target. The vertical dashed lines show the ~0.5 ns time duration in X-ray diffraction measurement. (<b>d</b>) The averaged stress for the iron sample as a function of time. In this shot, the deduced pressure for iron at the XRD time is 238 GPa.</p>
Full article ">Figure 4
<p>The c/a ratio of hcp-iron at high pressure by static and dynamical compression. The ideal c/a ratio for the hcp structure (grey dashed line) and the typical measurement data (black and red squares) with fitting data in Ref. [<a href="#B18-minerals-14-00715" class="html-bibr">18</a>] (black dotted, blue dashed, red, blue lines) are plotted for comparison. The ratios from the previous dynamical compression (black, grey, blue circles) from Refs. [<a href="#B24-minerals-14-00715" class="html-bibr">24</a>,<a href="#B27-minerals-14-00715" class="html-bibr">27</a>,<a href="#B49-minerals-14-00715" class="html-bibr">49</a>] are illustrated. Our ramp compression data are between those of previous shock experiments and the static compression data.</p>
Full article ">Figure 5
<p>The volume–c/a ratio relationship for iron and iron alloy at high pressure. The c/a ratios in static compression (squares) from Ref. [<a href="#B18-minerals-14-00715" class="html-bibr">18</a>] and in the previous dynamical compression (grey, blue circles) from Refs. [<a href="#B27-minerals-14-00715" class="html-bibr">27</a>,<a href="#B49-minerals-14-00715" class="html-bibr">49</a>] are illustrated for comparison. Our data obtained in ramp-compressed iron are illustrated in red circles. The initial volume used for V/V<sub>0</sub> calculation in static compression from Ref. [<a href="#B18-minerals-14-00715" class="html-bibr">18</a>] is 22.468 Å<sup>3</sup>.</p>
Full article ">
16 pages, 3199 KiB  
Article
Study of a Copper Oxide Leaching in Alkaline Monosodium Glutamate Solution
by Carlos G. Perea, Christian Ihle, Laurence Dyer, Simón Díaz Quezada and Humberto Estay
Minerals 2024, 14(7), 714; https://doi.org/10.3390/min14070714 - 15 Jul 2024
Viewed by 419
Abstract
Oxide copper minerals are commonly extracted via acidic leaching, using acids such as H2SO4, HCl, or HNO3. These strong acids are the most widely used because of their high dissolution kinetics. However, their main concern is the [...] Read more.
Oxide copper minerals are commonly extracted via acidic leaching, using acids such as H2SO4, HCl, or HNO3. These strong acids are the most widely used because of their high dissolution kinetics. However, their main concern is the high acid consumption because copper oxide deposits contain large amounts of acid-consuming gangue. This paper proposes using an alternative aqueous alkaline monosodium glutamate (MSG) system to leach copper oxide minerals. Tenorite (CuO) was used as the copper oxide mineral under study. The influence of process variables (such as temperature and glutamate concentration) and kinetics of this system on copper leaching from tenorite were studied. The results showed that temperature has a significant effect on copper dissolution rates. Increased temperature from 15 °C to 60 °C enhanced the copper extraction from 9.1% to 97.7% after 2 h. Leaching kinetics were analyzed using the shrinking core model (SCM) under various conditions, indicating that the leaching rate presented a mixed control. This method, however, fails to describe leaching for broad particle sizes due to its requirement for single-sized solid grains. This study demonstrated that a large particle size distribution in tenorite supported a successful extension of the SCM for leaching it from mixed glutamate solutions. The activation energy for the 15–60 °C temperature range was calculated to be 102.6 kJ/mol for the chemical control. Full article
Show Figures

Figure 1

Figure 1
<p>Eh–pH stability diagram for the CuO–glut–H<sub>2</sub>O system at 25 °C; [Cu] = 3.14 × 10<sup>−2</sup> M and [glut] = 0.5 M.</p>
Full article ">Figure 2
<p>Particle size distribution in tenorite samples via laser diffraction using Malvern Mastersizer 2000.</p>
Full article ">Figure 3
<p>X-ray diffraction pattern of tenorite sample.</p>
Full article ">Figure 4
<p>Effect of monosodium glutamate concentration on the dissolution of tenorite. Working conditions: pH = 9.4, temperature = 30 °C, and solids content = 0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>.</p>
Full article ">Figure 5
<p>Effect of temperature on dissolution of tenorite. Working conditions: pH = 9.4, glutamate = 0.5 M, and solids content = 0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>.</p>
Full article ">Figure 6
<p>Plots of <math display="inline"><semantics> <mrow> <mfrac bevelled="true"> <mrow> <mn>1</mn> </mrow> <mrow> <mn>3</mn> </mrow> </mfrac> <mi>l</mi> <mi>n</mi> <mfenced separators="|"> <mrow> <mn>1</mn> <mo>−</mo> <mi>x</mi> </mrow> </mfenced> <mo>+</mo> <mo>(</mo> <mn>1</mn> <mo>−</mo> <mi>x</mi> <msup> <mrow> <mo>)</mo> </mrow> <mrow> <mo>−</mo> <mfrac> <mrow> <mn>1</mn> </mrow> <mrow> <mn>3</mn> </mrow> </mfrac> </mrow> </msup> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> versus time at different glutamate concentrations and temperatures.</p>
Full article ">Figure 7
<p>Arrhenius plot for copper dissolution in the mixed control stage.</p>
Full article ">Figure 8
<p>Arrhenius plot for copper dissolution in the chemical control stage considering the particle size distribution.</p>
Full article ">Figure 9
<p>Validation plot of the chemical control with PSD kinetic model and experimental values of the copper dissolution.</p>
Full article ">
14 pages, 3344 KiB  
Article
Characterization of Kazakhstan’s Clays by Mössbauer Spectroscopy and X-ray Diffraction
by Adilkhan Shokanov, Irina Manakova, Mikhail Vereshchak and Anastassiya Migunova
Minerals 2024, 14(7), 713; https://doi.org/10.3390/min14070713 - 13 Jul 2024
Viewed by 519
Abstract
Studies of the mineralogical composition were carried out, and the features of the clays from the deposits of Kazakhstan were established using Mössbauer spectroscopy (MS) and X-ray diffraction analysis (XRD). According to the XRD results, all the samples were mixed-layer clays of the [...] Read more.
Studies of the mineralogical composition were carried out, and the features of the clays from the deposits of Kazakhstan were established using Mössbauer spectroscopy (MS) and X-ray diffraction analysis (XRD). According to the XRD results, all the samples were mixed-layer clays of the kaolinite–illite type. The lattice parameters of the kaolinite were determined, and it was shown that its structure was disordered and contained a certain amount of impurity in some of the clay samples. A special feature of two of the samples was the additionally identified muscovite polytype 2M1. The spectra of the iron-containing clays were amenable to being resolved into separate components, with similar Mössbauer parameters of the kaolinite, muscovite, illite, and glauconite. The oxidation state of the iron was determined using MS. The predominant part of paramagnetic iron in most samples was in the trivalent state. The primary minerals contributing to Fe2+ were illite and muscovite. The results obtained during the study of the clay samples with complex mineralogical compositions using MS and XRD methods both complemented one another and were found to be in good agreement. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of the No. 1 (<b>a</b>), No. 2 (<b>b</b>), No. 3 (<b>c</b>), No. 4 (<b>d</b>), No. 5 (<b>e</b>), No. 4 (<b>f</b>), and No. 7 (<b>g</b>) samples: <span class="html-fig-inline" id="minerals-14-00713-i001"><img alt="Minerals 14 00713 i001" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i001.png"/></span>—hematite, <span class="html-fig-inline" id="minerals-14-00713-i002"><img alt="Minerals 14 00713 i002" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i002.png"/></span>—kaolinite, <span class="html-fig-inline" id="minerals-14-00713-i003"><img alt="Minerals 14 00713 i003" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i003.png"/></span>—muscovite, <span class="html-fig-inline" id="minerals-14-00713-i004"><img alt="Minerals 14 00713 i004" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i004.png"/></span>—illite, <span class="html-fig-inline" id="minerals-14-00713-i005"><img alt="Minerals 14 00713 i005" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i005.png"/></span>—quartz, <span class="html-fig-inline" id="minerals-14-00713-i006"><img alt="Minerals 14 00713 i006" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i006.png"/></span>—albite, <span class="html-fig-inline" id="minerals-14-00713-i007"><img alt="Minerals 14 00713 i007" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i007.png"/></span>—calcite, <span class="html-fig-inline" id="minerals-14-00713-i008"><img alt="Minerals 14 00713 i008" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i008.png"/></span>—dolomite, <span class="html-fig-inline" id="minerals-14-00713-i009"><img alt="Minerals 14 00713 i009" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i009.png"/></span>—microcline, and <span class="html-fig-inline" id="minerals-14-00713-i010"><img alt="Minerals 14 00713 i010" src="/minerals/minerals-14-00713/article_deploy/html/images/minerals-14-00713-i010.png"/></span>—dellaite.</p>
Full article ">Figure 2
<p>MS spectra of the samples: No. 1 (<b>a</b>), No. 3 (<b>b</b>), No. 4 (<b>c</b>), No. 5 (<b>d</b>), and No. 7 (<b>e</b>). The left panel exhibits the Mössbauer spectra in the velocity range of ±10 mm/s, and the right panel shows the enlarged central parts of the corresponding spectra in the velocity range of −2–+3 mm/s.</p>
Full article ">
14 pages, 6880 KiB  
Article
Characteristics of Nanoparticles in Late Pliocene Paleo-Mountain Fire Relics in Jinsuo Basin, Yunnan Province and Their Implications for Paleoclimate Evolution
by Peng Zhang, Bangjun Liu, Yaqin Wang, Lei Zuo, Rui Liu, Jialong Wang and Ru Wang
Minerals 2024, 14(7), 712; https://doi.org/10.3390/min14070712 - 13 Jul 2024
Viewed by 411
Abstract
Wildfires significantly affect climate and environmental changes, closely tied to extreme weather responses. Vegetation combustion emits greenhouse gases (CO2, CH4, CO), warming the climate. Climate shifts, in turn, impact vegetation growth, altering combustible material types and quantities, thus affecting [...] Read more.
Wildfires significantly affect climate and environmental changes, closely tied to extreme weather responses. Vegetation combustion emits greenhouse gases (CO2, CH4, CO), warming the climate. Climate shifts, in turn, impact vegetation growth, altering combustible material types and quantities, thus affecting wildfire intensity, duration, and frequency. Wildfires profoundly affect ecosystems, influenced by factors like atmospheric oxygen and climate. Their combustion gases impact climate and vegetation growth. Recent advancements in studying ancient wildfires include analyzing nanoparticles as key indicators. This study discovered six types of nanoparticles in ancient wildfire remains, with sizes ranging from 50 nm to 500 nm and diverse compositions including elements such as C, O, Mg, Al, Ti, Fe, S, Ca, and P. These findings indicate that wildfires generate a variety of nanoparticles, offering new insights into ancient fire events. Elemental analysis revealed low magnesium but high calcium and aluminum levels, suggesting a warm, humid paleoclimate during these fires. The presence of high Ti-O ratios and carbon-rich nanoparticles points to ground fires with incomplete combustion. This research underscores the significance of nanoparticles in understanding the history and characteristics of ancient wildfires. Full article
(This article belongs to the Section Environmental Mineralogy and Biogeochemistry)
Show Figures

Figure 1

Figure 1
<p>A regional geological map of paleo-wildfire remains in Yunnan (<b>a</b>) and a profile of the lignite seam, showing the distribution of charcoal and wood remains (<b>b</b>).</p>
Full article ">Figure 2
<p>Examples of macroscopic charcoal particles in samples from the lignite seam. Samples from pale layers: (<b>a</b>,<b>b</b>); samples from dark layers: (<b>c</b>–<b>d</b>) (modified from [<a href="#B23-minerals-14-00712" class="html-bibr">23</a>]).</p>
Full article ">Figure 3
<p>Topography of nanoparticles in paleo-wildfire remains. Among them, (<b>a</b>–<b>d</b>) are different types of nanoparticles.</p>
Full article ">Figure 4
<p>Diagrams of morphology (<b>a</b>), diffraction (<b>b</b>), HRTEM (<b>c</b>), and lattice spacing (<b>d</b>) of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 5
<p>EDS diagram of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 6
<p>Diagrams of morphology (<b>a</b>), diffraction (<b>b</b>), HRTEM (<b>c</b>), and lattice spacing (<b>d</b>) of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 7
<p>EDS diagram of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 8
<p>Diagrams of topography (<b>a</b>), diffraction (<b>b</b>), HRTEM (<b>c</b>), and lattice spacing (<b>d</b>) of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 9
<p>EDS diagram of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 10
<p>Topography (<b>a</b>) and selection diffraction (<b>b</b>) of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 11
<p>EDS diagram of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 12
<p>Diagrams of topography (<b>a</b>) and diffraction (<b>b</b>) of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 13
<p>EDS diagram of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 14
<p>Diagrams of topography (<b>a</b>) and diffraction (<b>b</b>) of nanoparticles in paleo-wildfire remains.</p>
Full article ">Figure 15
<p>EDS diagram of nanoparticles in paleo-wildfire remains.</p>
Full article ">
17 pages, 48563 KiB  
Article
Molecular Dynamic Simulation of the Interaction of a Deep Eutectic Solvent Based on Tetraethylammonium Bromide with La3+ in Acidic Media
by Luver Echeverry-Vargas, Luz M. Ocampo-Carmona and Leopoldo Gutiérrez
Minerals 2024, 14(7), 711; https://doi.org/10.3390/min14070711 - 13 Jul 2024
Viewed by 429
Abstract
In recent years, noticeable progress has been made in the development of alternative extraction systems characterized by greater sustainability. In this context, deep eutectic solvents (DESs) have emerged as a promising alternative to the conventional solvents commonly used in metal extraction. This work [...] Read more.
In recent years, noticeable progress has been made in the development of alternative extraction systems characterized by greater sustainability. In this context, deep eutectic solvents (DESs) have emerged as a promising alternative to the conventional solvents commonly used in metal extraction. This work focuses on investigating the extraction of lanthanum in an aqueous solution of sulfuric acid using a deep eutectic solvent, employing molecular dynamics simulations (MD). The structural characteristics of the solvent and its interactions with the components of the aqueous solution are explored. In this study, tetraethylammonium bromide (TEABr) is combined with ethylene glycol (EG) to form a DES, in which sodium cyanide (NaCN) is subsequently solubilized. According to the results obtained from the MD simulation, the primary interactions in the DESs are established through hydrogen bonds between the bromine and the hydrogens of the methyl group of tetraethylammonium at 3.5 Å, as well as between the bromine and the hydrogens of the methylene group of ethylene glycol at 3.5 Å. Similarly, the main interactions between the binary DES and sodium cyanide occur through the hydrogens of the hydroxyl group of EG and the carbon of cyanide at 1.7 Å, and between the oxygen of the hydroxyl group of EG and the sodium at 2.5 Å. In the acidic solution, the primary interaction is highlighted between the lanthanum ion and the oxygen of the bisulfate at 2.8 Å. Additionally, it is observed that the interaction between the DES and the aqueous solution occurs between the lanthanum and the oxygen of the hydroxyl group of EG, as well as between the lanthanum and the carbon of cyanide at 4.4 Å. It is important to note that, when increasing the temperature from 25 to 80 °C, the interaction distance between the lanthanum and the carbon of cyanide decreases to 2.4 Å, suggesting a possible correlation with the increase in lanthanum extraction, as experimentally observed. Overall, this study underscores the importance of considering the fundamental structural interactions of the DES with the lanthanum acid solution, providing an essential theoretical basis for future experimental investigations. Full article
Show Figures

Figure 1

Figure 1
<p>Structure and atomic definition of the DES components.</p>
Full article ">Figure 2
<p>Structure and atomic definition of the components in the aqueous solution.</p>
Full article ">Figure 3
<p>Simulation box of the system composed of 150 ETG molecules, 40 TEABr molecules, 10 NaCN molecules, 15,000 water molecules, 1500 HSO<sub>4</sub><sup>−</sup> molecules, and 500 La<sup>3+</sup> ions.</p>
Full article ">Figure 4
<p>Molecular dynamics simulation methodology employed.</p>
Full article ">Figure 5
<p>Temperature effect on lanthanum extraction with the DES synthesized from an H<sub>2</sub>SO<sub>4</sub> leaching liquor.</p>
Full article ">Figure 6
<p>Radial distribution functions calculated for a central Br<sup>−</sup> ion with respect to the H<sub>S</sub> and H<sub>T</sub> hydrogen atoms of tetraethylammonium (<b>a</b>), and the H<sub>G</sub> and H<sub>O</sub> hydrogen atoms of ethylene glycol (<b>b</b>).</p>
Full article ">Figure 7
<p>Representative snapshot from a MD simulation of the local environment of a Br<sup>−</sup> ion in the TEA and ETG system (color code: magenta—Br; red—O; navy green—C; white—H; blue—N).</p>
Full article ">Figure 8
<p>Radial distribution functions calculated at 25 °C and 80 °C for a central Br<sup>−</sup> ion with respect to: the H<sub>S</sub> atoms (<b>a</b>), and H<sub>G</sub> hydrogen atoms (<b>b</b>).</p>
Full article ">Figure 9
<p>Radial distribution functions calculated at 25 °C for a central Br<sup>−</sup> ion with respect to the H<sub>S</sub> hydrogen atoms (<b>a</b>), and H<sub>G</sub> hydrogen atoms (<b>b</b>), in the system composed of TEABr–ETG–NaCN.</p>
Full article ">Figure 10
<p>Radial distribution functions calculated at 25 °C for the C<sub>M</sub>–H<sub>O</sub> pairs (<b>a</b>), and Na–O<sub>G</sub> pairs (<b>b</b>), in the system composed of TEABr–ETG–NaCN.</p>
Full article ">Figure 11
<p>Representative snapshot from a MD simulation of the local environment of the system composed of TEABr–ETG–NaCN (color code: magenta—Br; red—O; navy green—C; white—H; blue—N; apple green—Na).</p>
Full article ">Figure 12
<p>Radial distribution functions calculated at 25 °C and 80 °C for a central Br<sup>−</sup> ion with respect to the Hs hydrogens (<b>a</b>), and H<sub>G</sub> hydrogens (<b>b</b>), in the system composed of TEABr–ETG–NaCN.</p>
Full article ">Figure 13
<p>Radial distribution functions calculated at 25 °C and 80 °C for the C<sub>M</sub>–H<sub>O</sub> pairs (<b>a</b>), and Na–O<sub>G</sub> pairs (<b>b</b>), in the system composed of TEABr–ETG–NaCN.</p>
Full article ">Figure 14
<p>Radial distribution functions calculated at 25 °C for the La<sup>3+</sup>–O<sub>T</sub> pairs (<b>a</b>), La<sup>3+</sup>–O pairs (<b>b</b>), and H–O<sub>T</sub> pairs (<b>c</b>), in the aqueous system composed of H<sub>2</sub>O–La<sup>3+</sup>–HSO<sub>4</sub><sup>−</sup>.</p>
Full article ">Figure 15
<p>Radial distribution functions calculated at 25 °C for the La<sup>3+</sup>–O<sub>G</sub> pairs (<b>a</b>), and La<sup>3+</sup>–C<sub>M</sub> pairs (<b>b</b>), in the interaction between the components of the DES and the aqueous solution of lanthanum.</p>
Full article ">Figure 16
<p>Radial distribution functions calculated at 25 °C and 80 °C for the La<sup>3+</sup>–O<sub>G</sub> pairs (<b>a</b>), and La<sup>3+</sup>–C<sub>M</sub> pairs (<b>b</b>), in the interaction between the components of the DES and the aqueous solution of lanthanum.</p>
Full article ">Figure 16 Cont.
<p>Radial distribution functions calculated at 25 °C and 80 °C for the La<sup>3+</sup>–O<sub>G</sub> pairs (<b>a</b>), and La<sup>3+</sup>–C<sub>M</sub> pairs (<b>b</b>), in the interaction between the components of the DES and the aqueous solution of lanthanum.</p>
Full article ">Figure 17
<p>Schematic representation of the main interactions between (TEA–ETG–NaCN) and (H<sub>2</sub>O–La<sup>3+</sup>–HSO<sub>4</sub><sup>−</sup>). * As the temperature increases, the interaction distance decreases to 2.4 Å.</p>
Full article ">
18 pages, 84655 KiB  
Article
Petrogenesis and Tectonic Implications of the Granite Porphyry in the Sinongduo Ag-Pb-Zn Deposit, Central Tibet: Constraints from Geochronology, Geochemistry, and Sr-Nd Isotopes
by Peng Zhang, Zhuang Li, Feng Zhao and Xinkai Liu
Minerals 2024, 14(7), 710; https://doi.org/10.3390/min14070710 - 12 Jul 2024
Viewed by 461
Abstract
The Paleocene ore deposits related to the India–Asia continental collision are widely distributed in the Gangdese metallogenic belt. Among these, Sinongduo is the first discovered epithermal Ag-Pb-Zn deposit in the Lhasa terrane. However, there is still controversy over the ore-forming magma in this [...] Read more.
The Paleocene ore deposits related to the India–Asia continental collision are widely distributed in the Gangdese metallogenic belt. Among these, Sinongduo is the first discovered epithermal Ag-Pb-Zn deposit in the Lhasa terrane. However, there is still controversy over the ore-forming magma in this deposit. This study mainly reports new zircon U-Pb isotopic ages, whole-rock geochemistry, and Sr-Nd isotopic data for the granite porphyry from the Sinongduo deposit, aiming to discuss the petrogenesis and tectonic setting of the granite porphyry and its genetic link between the Ag-Pb-Zn mineralization. The results show that zircon U-Pb analyses yield ages of 62.9 ± 0.5 Ma and 59.0 ± 0.7 Ma for the granite porphyry, indicating that it formed during the Paleocene period. The timing of the granite porphyry intrusion is contemporaneous with the mineralization, suggesting that it is most likely the ore-forming magma in the Sinongduo deposit. The granite porphyry has high SiO2 and K2O, moderate Al2O3, and low Na2O, CaO, and FeOT contents, and it displays significant enrichments in LREEs and LILEs and depletions in HREEs and HFSEs, with negative Eu anomaly. The granite porphyry is a peraluminous series and can be classified as S-type granite. Moreover, the granite porphyry shows relatively high ratios of (87Sr/86Sr)i and low values of εNd(t). The geochemical and isotopic compositions of the granite porphyry from the Sinongduo area are similar to those of the upper continental crust, which suggests that the granite porphyry was most likely derived from the melting of the upper continental crust in the Lhasa terrane during the India–Asia collisional tectonic setting. Full article
(This article belongs to the Special Issue Genesis and Evolution of Pb-Zn-Ag Polymetallic Deposits: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Simplified map showing the location of the Himalayan–Tibetan orogeny; (<b>b</b>) tectonic framework of the Lhasa terrane (modified after [<a href="#B7-minerals-14-00710" class="html-bibr">7</a>]); (<b>c</b>) diagram showing the distribution of the magmatic rocks and the associated deposits in the Lhasa terrane (modified after [<a href="#B52-minerals-14-00710" class="html-bibr">52</a>]).</p>
Full article ">Figure 2
<p>The simplified geological map (<b>a</b>) and lithostratigraphy of borehole BZK1502 (<b>b</b>) of the Sinongduo deposit (modified after [<a href="#B20-minerals-14-00710" class="html-bibr">20</a>]).</p>
Full article ">Figure 3
<p>The hand specimen photographs and photomicrographs showing the main ore structure and textures in the mineral assemblages from the Sinongduo deposit. (<b>a</b>) The rhyolite porphyry and crystal tuff with sphalerite, galena, and pyrite sulfide minerals; (<b>b</b>) the rhyolite porphyry cut by the sphalerite–galena vein; (<b>c</b>) the chalcopyrite in the sphalerite; (<b>d</b>) the euhedral pyrite in the quartz; (<b>e</b>) the galena and sphalerite; (<b>f</b>) the pearceite and argentite; (<b>g</b>) the hematite and pearceite developed in the pyrite; (<b>h</b>) the acanthite in the jasper; (<b>i</b>) the pyrargyrite developed in pyrite fractures. Abbreviations: Sp, sphalerite; Gn, galena; Py, pyrite; Ccp, chalcopyrite; Arn, argentite; Pea, pearceite; Hem, hematite; Aca, acanthite; Pyr, pyrargyrite; III, illite; Jas, jasper; Ser, sericite; Chl, chlorite; Q, quartz.</p>
Full article ">Figure 4
<p>(<b>a</b>) The field relationship of rocks, (<b>b</b>) hand specimen photograph, and (<b>c</b>–<b>e</b>) cross-polarized photomicrographs of the granite porphyry in the Sinongduo deposit. Abbreviations: Kfs—potassium feldspar; Q—quartz; Ser—sericite; Bt—biotite.</p>
Full article ">Figure 5
<p>Representative cathodoluminescence images of zircon grains for the (<b>a</b>) SND-G1 and (<b>b</b>) 1502-98 granite porphyry samples from the Sinongduo deposit. The yellow circles are 32 μm in diameter and show the location of the U-Pb analytical sites.</p>
Full article ">Figure 6
<p>LA-ICP-MS zircon U-Pb concordia and weighted mean age diagrams of samples (<b>a</b>,<b>b</b>) SND-G1 and (<b>c</b>,<b>d</b>) 1502-98 for the Sinongduo granite porphyry.</p>
Full article ">Figure 7
<p>Geochemical diagrams for the granite porphyry from the Sinongduo deposit. (<b>a</b>) SiO<sub>2</sub> versus Na<sub>2</sub>O + K<sub>2</sub>O diagram after [<a href="#B69-minerals-14-00710" class="html-bibr">69</a>]; (<b>b</b>) A/NK versus A/CNK diagram after [<a href="#B70-minerals-14-00710" class="html-bibr">70</a>]. Data for the Sinongduo volcanic rocks are from [<a href="#B71-minerals-14-00710" class="html-bibr">71</a>]; data for the Paleocene granites are from [<a href="#B72-minerals-14-00710" class="html-bibr">72</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) The chondrite-normalized REE patterns and (<b>b</b>) primitive mantle normalized trace element patterns for the Sinongduo granite porphyry. Data for the chondrite and primitive mantle normalization are from [<a href="#B73-minerals-14-00710" class="html-bibr">73</a>], data for the Indian Ocean sediments, UCC, and LCC are from [<a href="#B74-minerals-14-00710" class="html-bibr">74</a>], Sinongduo volcanic rock data are from [<a href="#B71-minerals-14-00710" class="html-bibr">71</a>], and Paleocene granite data are from [<a href="#B72-minerals-14-00710" class="html-bibr">72</a>]. Abbreviations: UCC, upper continental crust; LCC, lower continental crust.</p>
Full article ">Figure 9
<p>The Sr-Nd isotopic compositions for the granite porphyry from the Sinongduo deposit. Data for the Indian Ocean MORB, UCC, and LCC are from [<a href="#B75-minerals-14-00710" class="html-bibr">75</a>,<a href="#B76-minerals-14-00710" class="html-bibr">76</a>,<a href="#B77-minerals-14-00710" class="html-bibr">77</a>]. Data for global subducting sediment (GLOSS) are from [<a href="#B78-minerals-14-00710" class="html-bibr">78</a>]. Data for the Sinongduo volcanic rocks are from [<a href="#B71-minerals-14-00710" class="html-bibr">71</a>]; data for the Paleocene granites are from [<a href="#B72-minerals-14-00710" class="html-bibr">72</a>].</p>
Full article ">Figure 10
<p>Al<sub>2</sub>O<sub>3</sub>-(Na<sub>2</sub>O + K<sub>2</sub>O) versus CaO versus FeO<sup>T</sup> + MgO diagram (after [<a href="#B81-minerals-14-00710" class="html-bibr">81</a>]). Data for the Paleocene I-type volcanic rocks are from [<a href="#B32-minerals-14-00710" class="html-bibr">32</a>,<a href="#B33-minerals-14-00710" class="html-bibr">33</a>]; other data from the literature are from [<a href="#B10-minerals-14-00710" class="html-bibr">10</a>,<a href="#B71-minerals-14-00710" class="html-bibr">71</a>,<a href="#B72-minerals-14-00710" class="html-bibr">72</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) SiO<sub>2</sub> versus ε<sub>Nd</sub>(<span class="html-italic">t</span>), (<b>b</b>) SiO<sub>2</sub> versus (<sup>87</sup>Sr/<sup>86</sup>Sr)<sub>i</sub>, (<b>c</b>) Nb/Ta versus Zr, and (<b>d</b>) Nb/Ta versus Nb diagrams (after [<a href="#B88-minerals-14-00710" class="html-bibr">88</a>]. Data for the BCC are from [<a href="#B74-minerals-14-00710" class="html-bibr">74</a>]; other data sources are the same as that in the <a href="#minerals-14-00710-f009" class="html-fig">Figure 9</a>. Abbreviations: UCC, upper continental crust; LCC, lower continental crust; BCC, basin continental crust.</p>
Full article ">
22 pages, 23827 KiB  
Article
The Role of Hydrocarbons in the Formation of Uranium Mineralization, Louzhuangzi District, Southern Junggar Basin (China)
by Zhong-Bo He, Bao-Qun Hu, Lin-Fei Qiu, Yun Wang, Hong Chen, Wei-Wei Jia, Yi-Fei Li, Hua-Li Ji and Man-Huai Zhu
Minerals 2024, 14(7), 709; https://doi.org/10.3390/min14070709 - 12 Jul 2024
Viewed by 443
Abstract
In recent years, there have been important breakthroughs in the exploration for sandstone-hosted uranium (U) deposits in the Louzhuangzi district of the southern Junggar Basin. Between 2020 and 2023, a medium-sized sandstone-hosted uranium deposit production area was identified in the region. Only a [...] Read more.
In recent years, there have been important breakthroughs in the exploration for sandstone-hosted uranium (U) deposits in the Louzhuangzi district of the southern Junggar Basin. Between 2020 and 2023, a medium-sized sandstone-hosted uranium deposit production area was identified in the region. Only a few investigations have been conducted at the Louzhuangzi U deposit, including those analyzing its geological–tectonic evolution, basic geological features, hydrogeology, and ore-controlling factors. It is generally believed that uranium mineralization at the Louzhuangzi U deposit is controlled by a redox zone. Organic matter (referred to as OM hereafter) consisting of bitumen and carbonaceous debris is very common in the uranium ores (especially in high-grade ores) at the Louzhuangzi U deposit. However, the characteristics of the OM and its contribution to uranium’s mineralization have not been studied in detail. In this study, OM-rich U-ores, altered sandstone, and barren sandstone samples were collected for petrography, mineralogical, micro-spectroscopy, carbon, and sulfur isotope studies. The results of this study show that the distribution of U minerals and metal sulfides (pyrite, sphalerite, etc.) was strictly controlled by bitumen at the Louzhuangzi U deposit. The bitumen may have been formed by hydrocarbon-rich and U-rich ore-forming fluids, which were formed after hydrocarbon generation and expulsion in the underlying Jurassic coal-bearing source rocks. The fluids contained U, Zn, Fe, and other metal elements, which migrated together and then precipitated into the oxidized Toutunhe Formation sandstone through cracking and differentiation processes. Therefore, the results indicate that migrated hydrocarbons were involved in U mineralization, in addition to oxidation–reduction processes, in the Louzhuangzi district, south of the Junggar Basin (China). Full article
(This article belongs to the Special Issue Uranium: Geochemistry and Mineralogy)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the study area; (<b>b</b>) map showing tectonic units of the study area, modified from [<a href="#B19-minerals-14-00709" class="html-bibr">19</a>,<a href="#B22-minerals-14-00709" class="html-bibr">22</a>]; (<b>c</b>) geological map of the Louzhuangzi U deposit, modified from [<a href="#B19-minerals-14-00709" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>Map showing the comprehensive stratigraphic column of the study area.</p>
Full article ">Figure 3
<p>A schematic cross-section (L6) showing the spatial positions of the U-ore bodies considered in this study, modified from [<a href="#B20-minerals-14-00709" class="html-bibr">20</a>]. 1—Qigu Fm; 2—Toutunhe Fm; 3—Xishanyao Fm; 4—sand–conglomerate; 5—sandstone; 6—mudstone; 7—gray-blackish carbonaceous mudstone; 8—gray, gray-green color; 9—variegated color; 10—brown color; 11—coal; 12—lithological boundary and formation contact boundary; 13—U-ore body (U% &gt; 0.01%); 14—U-mineralized sandstone; 15—gamma log curve and grade (%) of ore body (m); 16—drill holes with elevation and depth annotations; 17—grayish white alteration zone; 18—sampling location.</p>
Full article ">Figure 4
<p>Images showing the macroscopic features of the OM in the U-ores and the primary zone of unaltered sandstone in the Louzhuangzi U deposit. (<b>a</b>–<b>d</b>) The different color specimens of OM-bearing U-ores; (<b>e</b>) carbonaceous debris in grayish white unaltered (barren) sandstone; (<b>f</b>) a mosaic of images showing the microscopic features of fine-veined disseminated OM in a U-ore; (<b>g</b>) an image showing the clastic mineral structure of the ore-bearing sandstone and the cementation in its pores; (<b>h</b>) an image showing the carbonaceous debris in the unaltered sandstones of the primary zone. OM in U-ore. OM = organic matter; Qz = quartz; Cal = calcite; Dtr = rock detritus; Ab = albite.</p>
Full article ">Figure 5
<p>Images showing the distributions of the U minerals in the Louzhuangzi U deposit. Images (<b>b</b>,<b>e</b>) were captured via EDS; (<b>c</b>) is a reflected light image captured by a microscope; all others are BSE images captured via SEM. (<b>a</b>–<b>e</b>) Coffinite in close symbiosis with pyrite in OM vein; (<b>f</b>) Coffinite and two stages of pyrite (Py1 and Py2) in bitumen; (<b>g</b>) Sphalerite symbiotic with pyrite containing cloudy-like U; (<b>h</b>) Tiny coffinite and pyrite particles symbiotic with silicon in bitumen; (<b>i</b>) Blocky coffinite in micropores and symbiotic with pyrite in sandstone. U = uranium; Pit = pitchblende; Cof = coffinite; Cc = calcite; Py = pyrite; Dol = dolomite; Sph = Sphalerite; Si = silicon.</p>
Full article ">Figure 6
<p>EPMA mapping images showing the distribution of different elements (Y, Si, Th, U, As, Ti, Ca, Pb, Zr, Nd, P) in the U-rich bitumen and a multi-element weight composition image (Wt).</p>
Full article ">Figure 7
<p>Raman (<b>a</b>,<b>b</b>) and infrared spectra (<b>c</b>) of the OM in U-ores at the Louzhuangzi U deposit, Junggar Basin, and the Honghaigou U deposit, Yili Basin.</p>
Full article ">Figure 8
<p>Characteristic trace element change curves of U-ores and primary barren gray sandstone of Toutunhe Fm in the Louzhuangzi U deposit.</p>
Full article ">Figure 9
<p>Diagrams representing the relationship between typical trace elements (Mo, Pb, Cu, Y, Ni, Zn) and the U content.</p>
Full article ">Figure 10
<p>Sulfur isotope distribution of U-symbiotic pyrite from the Louzhuangzi U deposit and typical rock and energy resources (modified from references [<a href="#B35-minerals-14-00709" class="html-bibr">35</a>,<a href="#B38-minerals-14-00709" class="html-bibr">38</a>,<a href="#B39-minerals-14-00709" class="html-bibr">39</a>]).</p>
Full article ">Figure 11
<p>A conceptual model of the U mineralization and metallogenic process of the Louzhuangzi U deposit. (<b>a</b>) The gray sandstone containing carbonaceous debris was formed via sedimentation, and the first stage of U mineralization at the Louzhuangzi U deposit was formed via the interlayer oxidation of supergene fluids, and then the formation was oxidized to a brown or light yellow color. (<b>b</b>) The second stage of U mineralization at the Louzhuangzi U deposit occurred during the cracking of hydrocarbon-containing fluids in oxidized formations, and the hydrocarbon-containing fluids evolved from the coal-bearing strata in the lower part of the Toutunhe Formation. The strata were also reduced to gray or grayish-white compounds.</p>
Full article ">
24 pages, 6179 KiB  
Article
The Devonian Kalarskoe Epithermal Occurrence of the Kaburchak Au-Ag Cluster in the Altai-Sayan Folded Area, Russia: Geological Setting; Mineralogical, Geochemical, and Geochronological Features
by Alexander I. Chernykh, Polina N. Leibham, Lidia A. Sokolova, Olga V. Yakubovich, Maria O. Anosova and Evgeny A. Naumov
Minerals 2024, 14(7), 708; https://doi.org/10.3390/min14070708 - 12 Jul 2024
Viewed by 438
Abstract
Prospecting efforts to located Au mineralization within the Altai-Sayan fold area (ASFA) over previous decades have revealed that Devonian epithermal Au-Ag mineralization is more widespread than previously recognized. The preservation of this type of mineralization in Paleozoic rocks offers new prospects for the [...] Read more.
Prospecting efforts to located Au mineralization within the Altai-Sayan fold area (ASFA) over previous decades have revealed that Devonian epithermal Au-Ag mineralization is more widespread than previously recognized. The preservation of this type of mineralization in Paleozoic rocks offers new prospects for the exploration of Au-Ag deposits in the underexplored region of Gornaya Shoria. The Kalarskoe epithermal Au-Ag occurrence represents Devonian epithermal mineralization within the Kaburchak cluster, Gornaya Shoria, Russia. This occurrence is confined to zones of argillic alteration that were superimposed on previously formed propylites. The argillic-altered rocks host quartz-sulfide veinlet zones. The mineralization of the Kalarskoe site is characterized by a high abundance of sulfide minerals: commonly, 5%–10%; often, up to 20%; and in some cases, up to 60%–70%. The ore minerals are represented by pyrite, arsenopyrite, sphalerite, galena, chalcopyrite, fahlores, native Au, and electrum, as well as by the sulfosalts Pb, Bi, Ag, Cu, and the tellurides of Au, Ag, and Pb. Based on mineralogical observations, at least four generations of sulfide mineral formations are distinguished within the ore occurrence. The mineralization of the Kalarskoe ore occurrence may be assigned to the intermediate sulfidation (IS) type. The results of the (U,Th)-He dating of pyrite from the pyrite-arsenopyrite massive body (pyr-3 and 4) revealed the protracted history of the mineralization in the intervals from ~399 to ~371 Ma. The obtained results substantially enhance the prospecting models for the exploration of epithermal Au-Ag deposits in the western part of the Altai-Sayan fold area (ASFA). Full article
Show Figures

Figure 1

Figure 1
<p>Schematic map showing the distribution of epithermal Au-Ag mineralization in the western segment of the ASFA (Figure 2).</p>
Full article ">Figure 2
<p>Schematic map showing geological features and Au-bearing potential of the Kaburchak cluster (Figure 3).</p>
Full article ">Figure 3
<p>Schematic geological prospecting map showing the Au-Ag Kalarskoe occurrence.</p>
Full article ">Figure 4
<p>(<b>a</b>) The mineral composition and zoning of altered wall rocks of the Kalarskoe occurrence, according to Ref. [<a href="#B9-minerals-14-00708" class="html-bibr">9</a>] and new data. Samples from drill holes C1, C2, 7k, and S-17; (A, B, G—massive pyrite-arsenopyrite body; C—superimposed argillic alteration on propylite with disseminated pyrite and its veinlets; D, E, H, I—argillic metasomatites, with sulfide and later, carbonate (Ca, Mn) veinlets; F—relict porphyry texture of argillic-altered rock; J—argillic metasomatite with disseminated pyrite; K—argillic-altered rock with schistose structure. (<b>b</b>) Simplified geological cross-section of the Kalarskoe occurrence (all volcanic rocks were propylitically altered) (modified after Ref. [<a href="#B10-minerals-14-00708" class="html-bibr">10</a>]) (1) propylitic alteration; (2) hypergene zone; (3) argillic alteration with disseminated pyrite and its veinlets; (4) pyrite-arsenopyrite veins; (5) quaternary sediment; (6) number of drill holes. Letters corresponds to the sample numbers from (<b>a</b>).</p>
Full article ">Figure 5
<p>Photomicrograph of a massive pyrite-arsonopyrite assemblage at a 33.3 m depth from drill hole 7k. Some of the pyrite grains in this sample were used for (U-Th)-He dating. Py—pyrite; Apy—arsenopyrite.</p>
Full article ">Figure 6
<p>Changes in the content of elements with depth in the mineralized rocks and metasomatites from drill hole C2, based on the results of ICP-MS analysis. The vertical axis is the content of elements (ppm) on the logarithmic scale. The horizontal axis is the sampling depth.</p>
Full article ">Figure 7
<p>Images of polished sections (updated based on the work of Leibham, 2022 [<a href="#B9-minerals-14-00708" class="html-bibr">9</a>]). (<b>A</b>)—disseminated pyrite and arsenopyrite in the argillic-altered rock; (<b>B</b>,<b>C</b>)—massive pyrite-arsenopyrite vein; (<b>D</b>)—pyrrhotite and chalcopyrite in the propylite; (<b>E</b>)—pyrite vein crossed by galena-carbonate veinlet; (<b>F</b>)—pyrite and arsenopyrite fragments in sphalerite with inclusions of chalcopyrite; (<b>G</b>)—sphalerite, galena, and chalcopyrite in the carbonate veinlet; (<b>H</b>)—later chalcopyrite with fahlore; (<b>I</b>,<b>J</b>)—fahlore in the carbonate veinlet; (<b>K</b>)—sulfosalt with relicts of fahlore; (<b>L</b>)—the intergrowth of fahlore, chalcopyrite, and sulfosalt. Apy—arsenopyrite; Cb—carbonates; Ccp—chalcopyrite; Gn—galena; Py—pyrite; Pyh—pyrrhotite; Qz—quartz; Rds—rhodochrosite; Rt—rutile; Slf—sulfosalts; Sp—sphalerite; Ttr—tetrahedrite.</p>
Full article ">Figure 8
<p>BSE images of jamesonite (Ja), bournonite (Bnn), tetrahedrite-(Fe) (Ttr-Fe), and tetrahedrite-(Zn) (Ttr-Zn). The numbers of the analyses are shown in parentheses.</p>
Full article ">Figure 9
<p>BSE images. The compositions of electrum and native Au at the analysis points are indicated (wt.%).</p>
Full article ">Figure 10
<p>BSE image. A fractured pyrite-arsenopyrite aggregate with a marginal destruction zone, within which Au occurrences are noted (orange pins).</p>
Full article ">Figure 11
<p>Morphology of Au from eluvial sediments above the central Au-bearing zone of the Kalarskoe occurrence, with geochemical composition data.</p>
Full article ">Figure 12
<p>Paragenetic sequence of alteration and mineralization at the Kalarskoe occurrence. The numbers on the scheme indicate the different morphological forms of the minerals (see text). * calcite, kutnohorite, and rhodochrosite were identified and grouped.</p>
Full article ">Figure 13
<p>Distribution of (U,Th)-He ages of pyrite from the massive pyrite-arsenopyrite body, probability density plot (grey), and previous age estimations of the various rocks within the Kalarskoe field. Blue dots corresponds to the age of pyrite from the sample 7k-33.3; green dots—sample 7k-40.1. The plot is constructed using DensityPlotter software ver. 8.5 [<a href="#B21-minerals-14-00708" class="html-bibr">21</a>]. Error bars represents 2σ.</p>
Full article ">Figure 14
<p>Images of the pyrite from the sample 7k-33.3 m, which shows at least two coexisting generations of pyrite: grains with zonation (<b>A</b>,<b>B</b>) and grains without signs of the zonation. Dotted lines are used to emphasizes the internal zoning.</p>
Full article ">
21 pages, 3649 KiB  
Article
Comparing the Performance of Hydrocyclones and High-Frequency Screens in an Industrial Grinding Circuit: Part I—Size Separation Assessments
by Bruna Silveira Costa, Maurício Guimarães Bergerman and Homero Delboni Júnior
Minerals 2024, 14(7), 707; https://doi.org/10.3390/min14070707 - 12 Jul 2024
Viewed by 386
Abstract
Industrial ball milling circuits usually include hydrocyclones in a closed configuration to achieve a specified grinding size. Although hydrocyclones are relatively simple to operate, their classification performance is generally low, leading to significant fines recirculation within the circuit, consequently overgrinding the product. Conversely, [...] Read more.
Industrial ball milling circuits usually include hydrocyclones in a closed configuration to achieve a specified grinding size. Although hydrocyclones are relatively simple to operate, their classification performance is generally low, leading to significant fines recirculation within the circuit, consequently overgrinding the product. Conversely, high-frequency screening potentially shows a relatively higher separation efficiency, as the entrainment of fines to the coarse product is significantly reduced. The present work compares the performance of hydrocyclones—HC and high-frequency screens—HFS based on four surveys conducted in Nexa’s Vazante Zinc ore industrial grinding circuit in Vazante, Brazil, which processes zinc silicate ore. The comparisons included the partition of solids, water split, and particle size distributions. Whiten’s partition curve model was adopted to obtain the selected performance parameters through mass balancing the experimental data. The industrial surveys comprised three different size separation configurations, i.e., HC-Only, HFS-Only, and a combined HC-HFS setup. In all cases, the assessments consistently indicated higher separation performances with HFS compared to the HC operation. The final product associated with the HC+HFS configuration showed a narrower size distribution around the grinding size. Full article
Show Figures

Figure 1

Figure 1
<p>Partition curves: (<b>a</b>) experimental partition curve; (<b>b</b>) reduced partition curve.</p>
Full article ">Figure 2
<p>Grinding circuit C sampling points—HFS+HC configuration.</p>
Full article ">Figure 3
<p>Grinding circuit C sampling points—HC-Only configuration.</p>
Full article ">Figure 4
<p>Industrial installation of C grinding circuit.</p>
Full article ">Figure 5
<p>Experimental and mass-balanced size distributions for HFS+HC—Campaign 1.</p>
Full article ">Figure 6
<p>Reduced partition curves for HFS+HC and the separate HFS and HC conditions—Campaign 1.</p>
Full article ">Figure 7
<p>Experimental and mass-balanced size distributions, HC-Only—Campaign 1.</p>
Full article ">Figure 8
<p>Experimental and mass-balanced size distributions, HFS+HC—Campaign 2.</p>
Full article ">Figure 9
<p>Reduced partition curves for HFS+HC and separate HFS and HC conditions—Campaign 2.</p>
Full article ">Figure 10
<p>Experimental and mass-balanced size distributions, HC-Only—Campaign 2.</p>
Full article ">Figure 11
<p>Reduced partition curves for the HFS+HC and HC-Only configurations—Campaigns 1 and 2.</p>
Full article ">Figure 12
<p>Reduced partition curves for HFS+HC and separate HFS and HC conditions—Campaigns 1 and 2.</p>
Full article ">Figure 13
<p>Size distributions of the grinding circuit product in (<b>a</b>) Campaign 1, (<b>b</b>) Campaign 2, and (<b>c</b>) both campaigns.</p>
Full article ">
13 pages, 4836 KiB  
Article
Effect of the Interaction between Clays and Cations on Froth Rheology in Flotation
by Chao Li, Zhongren Wu, Zhihang Wu, Xianggen Chen and Yijun Cao
Minerals 2024, 14(7), 706; https://doi.org/10.3390/min14070706 - 12 Jul 2024
Viewed by 365
Abstract
The significance of froth rheology in affecting flotation performance is widely acknowledged. Clays could deteriorate flotation performance by altering froth rheology. The presence of cations further complicates the flotation system. Thus far, the interaction between clay minerals and cations and their impact on [...] Read more.
The significance of froth rheology in affecting flotation performance is widely acknowledged. Clays could deteriorate flotation performance by altering froth rheology. The presence of cations further complicates the flotation system. Thus far, the interaction between clay minerals and cations and their impact on froth rheology remains unclear. The present work selected three typical clays and cations with two valences (Na+ and Ca2+) to investigate their interacting influences on froth rheology. The results indicate that clays exhibit diverse froth rheological behaviors, with increasing cation strength from 0 to 0.1 mol/L. For montmorillonite, the froth viscosity initially decreased and subsequently increased. For kaolinite, upon the addition of cations, there was a significant decrease in froth viscosity; nevertheless, froth viscosity barely changed as the valency and concentration of the cations increased. Talc produced a considerably more viscous froth, and froth viscosity continued to rise with increasing concentrations of cations. The underlying mechanisms of the different responses in froth rheology were also investigated. The findings of this work have the potential to advance the optimization of flotation for complex ores containing clay minerals in high-salt processing water. Full article
(This article belongs to the Special Issue Interfacial Chemistry of Critical Mineral Flotation)
Show Figures

Figure 1

Figure 1
<p>Schematic of the flotation rig and the rheometer.</p>
Full article ">Figure 2
<p>Froth rheology of three clay minerals under different Na<sup>+</sup> concentrations (<b>a</b>) Montmorillonite, (<b>b</b>) Kaolinite, (<b>c</b>) Talc, (<b>d</b>) apparent viscosity at 2 s<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Froth rheology of three clay minerals under different Ca<sup>2+</sup> concentrations (<b>a</b>) Montmorillonite, (<b>b</b>) Kaolinite, (<b>c</b>) Talc, (<b>d</b>) apparent viscosity at 2 s<sup>−1</sup>.</p>
Full article ">Figure 4
<p>Correlation between froth properties and cation type, as well as concentration (<b>a</b>) water holdup vs. Na<sup>+</sup> concentration, (<b>b</b>) Solids concentration vs. Na<sup>+</sup> concentration, (<b>c</b>) water holdup vs. Ca<sup>2+</sup> concentration, (<b>d</b>) Solids concentration vs. Ca<sup>2+</sup> concentration.</p>
Full article ">Figure 5
<p>The correlation between water holdup and solids volumetric concentration in froth: (<b>a</b>) hydrophilic minerals and (<b>b</b>) hydrophobic talc.</p>
Full article ">Figure 6
<p>Froth apparent viscosity in relation to water holdup (<b>a</b>) and solids concentration (<b>b</b>) in the froth.</p>
Full article ">Figure 7
<p>Influence of cations on the zeta potential of montmorillonite (<b>a</b>) and setting tests (<b>b</b>) Na<sup>+</sup>, (<b>c</b>) Ca<sup>2+</sup>.</p>
Full article ">Figure 8
<p>Influence of cations on the zeta potential of kaolinite (<b>a</b>) and setting tests ((<b>b</b>) Na<sup>+</sup>, (<b>c</b>) Ca<sup>2+</sup>).</p>
Full article ">Figure 9
<p>Influence of cations on the zeta potential (<b>a</b>) and hydrophobic interaction between talc particles (<b>b</b>).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop