Journal Description
Minerals
Minerals
is an international, peer-reviewed, open access journal of natural mineral systems, mineral resources, mining, and mineral processing. Minerals is published monthly online by MDPI.
- Open Access— free for readers, with article processing charges (APC) paid by authors or their institutions.
- High Visibility: indexed within Scopus, SCIE (Web of Science), GeoRef, CaPlus / SciFinder, Inspec, Astrophysics Data System, AGRIS, and other databases.
- Journal Rank: JCR - Q2 (Mineralogy) / CiteScore - Q2 (Geology)
- Rapid Publication: manuscripts are peer-reviewed and a first decision is provided to authors approximately 18 days after submission; acceptance to publication is undertaken in 2.6 days (median values for papers published in this journal in the first half of 2024).
- Recognition of Reviewers: reviewers who provide timely, thorough peer-review reports receive vouchers entitling them to a discount on the APC of their next publication in any MDPI journal, in appreciation of the work done.
- Companion journal: Mining
Impact Factor:
2.2 (2023);
5-Year Impact Factor:
2.5 (2023)
Latest Articles
Salinity Effects on the Physicochemical and Mechanical Behavior of Untreated and Lime-Treated Saline Soils
Minerals 2024, 14(12), 1217; https://doi.org/10.3390/min14121217 - 28 Nov 2024
Abstract
Improving saline soils’ properties by incorporating limes is a practical technique, generally due to cation exchange, pozzolanic reaction, and carbonation. This study explores how soil salinity, measured by electrical conductivity, affects untreated and lime-treated saline soils. An Algerian sebkha soil (from Ain M’lila)
[...] Read more.
Improving saline soils’ properties by incorporating limes is a practical technique, generally due to cation exchange, pozzolanic reaction, and carbonation. This study explores how soil salinity, measured by electrical conductivity, affects untreated and lime-treated saline soils. An Algerian sebkha soil (from Ain M’lila) with an original high salinity (ECe3 = 23.2 dS m−1) was used. The same soil was washed to create medium (ECe2 = 8.3 dS.m−1) and low (ECe1 = 2.32 dS.m−1) salinity soil samples. The results of this study indicate that salinity influenced the shape of the particle size distribution curve, particularly in the silt range. Salinity also had a significant effect on carbonate content (CaCO3) and unconfined compressive strength (UCS). For the untreated soil, when salinity decreased, the UCS and CaCO3 content increased. However, when salinity decreased for the treated soil, the UCS increased, while the CaCO3 content decreased. X-ray diffraction (XRD) analysis of untreated soils showed halite (NaCl) disappearance and gypsum (CaSO4 2H2O) reduction with decreasing salinity in ECe1. In treated soil at ECe3, these mineral phases remained constant. While XRD detected no new cementitious phases in treated ECe3 or ECe1 samples, thermogravimetric analysis confirmed the presence of portlandite in both. As Ain M’lila sebkha is a chloride–sulfate soil, the dissolution of the halite and gypsum phases released more Cl− and SO42− ions into the interstitial solution. In a low fraction of clay, these ions obstructed and slowed the pozzolanic reaction in the ECe3 soil. Identifying the season when this type of soil has lower salinity can be beneficial for treatment from a technical, economic, and environmental point of view.
Full article
Open AccessReview
Polymer/Clay Nanocomposites as Advanced Adsorbents for Textile Wastewater Treatment
by
Adel Mokhtar, Boubekeur Asli, Soumia Abdelkrim, Mohammed Hachemaoui, Bouhadjar Boukoussa, Mohammed Sassi, Gianluca Viscusi and Mohamed Abboud
Minerals 2024, 14(12), 1216; https://doi.org/10.3390/min14121216 - 28 Nov 2024
Abstract
This review explores the removal of textile dyes from wastewater using advanced polymer/clay composites. It provides an in-depth analysis of the chemical and physical properties of these composites, emphasizing how the combination of polymers and clays creates a synergistic effect that significantly improves
[...] Read more.
This review explores the removal of textile dyes from wastewater using advanced polymer/clay composites. It provides an in-depth analysis of the chemical and physical properties of these composites, emphasizing how the combination of polymers and clays creates a synergistic effect that significantly improves the efficiency of dye removal. The structural versatility of the composites, derived from the interaction between the layered clay sheets and the flexible polymer matrices, is detailed, showcasing their enhanced adsorption capacity and catalytic properties for wastewater treatment. The review outlines the key functional groups present in both polymers and clays, which are crucial for binding and degrading a wide range of dyes, including acidic, basic, and reactive dyes. The role of specific interactions, such as hydrogen bonding, ion exchange, and electrostatic attractions between the dye molecules and the composite surface, is highlighted. Moreover, the selection criteria for different types of clays such as montmorillonite, kaolinite, and bentonite and their modifications are examined to demonstrate how structural and surface modifications can further improve their performance in composite materials. Various synthesis methods for creating polymer/clay composites, including in situ polymerization, solution intercalation, and melt blending, are discussed. These fabrication techniques are evaluated for their ability to control particle dispersion, optimize interfacial bonding, and enhance the mechanical and chemical stability of the composites. Furthermore, the review introduces advanced characterization techniques, such as X-ray diffraction (XRD), scanning electron microscopy (SEM), Fourier-transform infrared spectroscopy (FTIR), and thermogravimetric analysis (TGA), to help researchers assess the morphological, structural, and thermal properties of the composites, aligning these features with their potential application in dye removal. Additionally, the review delves into the primary mechanisms involved in the dye removal process, such as adsorption, photocatalytic degradation, and catalytic reduction. It also provides an overview of the kinetic and thermodynamic models commonly used to describe the adsorption processes in polymer/clay composites. The environmental and operational factors influencing the efficiency of dye removal, such as pH, temperature, and composite dosage, are analyzed in detail, offering practical insights for optimizing performance under various wastewater conditions. In conclusion, this review not only highlights the promising potential of polymer/clay composites for textile dye removal but also identifies current challenges and future research directions. It underscores the importance of developing eco-friendly, cost-effective, and scalable solutions to address the growing concerns related to water pollution and sustainability in wastewater management.
Full article
(This article belongs to the Special Issue Environmental Pollution and Assessment in Mining Areas)
►▼
Show Figures
Figure 1
Figure 1
<p>Schematic sketch showing the preparation steps, properties, characteristics, and environmental application fields of polymer/clay composite.</p> Full article ">Figure 2
<p>Schematic representation of the structure of the clays.</p> Full article ">Figure 3
<p>Schematic diagram of the modification of the most common clays used in wastewater treatment.</p> Full article ">Figure 4
<p>The most frequently used polymers in water treatment.</p> Full article ">Figure 5
<p>Synthesis of different types of polymer/clay nanocomposites.</p> Full article ">Figure 6
<p>Main techniques for polymer/clay nanocomposite preparation.</p> Full article ">Figure 7
<p>XRD patterns of nanocomposites obtained by different organoclays. (<b>a</b>) XRD patterns of nanocomposite prepared by an in situ method (symbolized by In) or by direct dispersion of poly(glycidylmethacrylate) (Poly(GMA)) in organophilic clay galleries using ultrasound radiation (symbolized by So). CTA-Magh(2CEC) is an Algerian MMT clay modified by CTA+ surfactant using 1CEC. (<b>b</b>) XRD patterns of poly(GMA)/organophilic clay nanocomposite (containing 2CEC of CTA+). (<b>c</b>) XRD patterns of poly(GMA)/organophilic clay nanocomposite (containing 1CEC of TBA+). Reproduced with permission from Ref. [<a href="#B71-minerals-14-01216" class="html-bibr">71</a>].</p> Full article ">Figure 8
<p>TEM images of nanocomposites Nano-So4 and Nano-So7 (the abbreviation of these samples is well detailed in the XRD part). Reproduced with permission from Ref. [<a href="#B71-minerals-14-01216" class="html-bibr">71</a>].</p> Full article ">Figure 9
<p>Cellulose acetate/organophilic clay. CA: cellulose acetate; Mag: Maghnite (MMT) modified by CTA+; CA/Mag (3%–10%): nanocomposite obtained at different weights of Maghnite. Reproduced with permission from Ref. [<a href="#B72-minerals-14-01216" class="html-bibr">72</a>].</p> Full article ">Figure 10
<p>The chemical structures of synthetic dyes most frequently studied.</p> Full article ">Figure 11
<p>Dye chemical classification.</p> Full article ">Figure 12
<p>Schematic representations of mechanisms for the elimination of various dyes using polymer/clay composite via electrostatic attraction, surface hydrophobicity, π-π interaction, and hydrogen bonding.</p> Full article ">Figure 13
<p>Number of publications per year on polymer/clay nanocomposites as adsorbents.</p> Full article ">Figure 14
<p>Schematic representation of reduction/degradation mechanisms of dye using polymer/clay nanocomposite.</p> Full article ">
<p>Schematic sketch showing the preparation steps, properties, characteristics, and environmental application fields of polymer/clay composite.</p> Full article ">Figure 2
<p>Schematic representation of the structure of the clays.</p> Full article ">Figure 3
<p>Schematic diagram of the modification of the most common clays used in wastewater treatment.</p> Full article ">Figure 4
<p>The most frequently used polymers in water treatment.</p> Full article ">Figure 5
<p>Synthesis of different types of polymer/clay nanocomposites.</p> Full article ">Figure 6
<p>Main techniques for polymer/clay nanocomposite preparation.</p> Full article ">Figure 7
<p>XRD patterns of nanocomposites obtained by different organoclays. (<b>a</b>) XRD patterns of nanocomposite prepared by an in situ method (symbolized by In) or by direct dispersion of poly(glycidylmethacrylate) (Poly(GMA)) in organophilic clay galleries using ultrasound radiation (symbolized by So). CTA-Magh(2CEC) is an Algerian MMT clay modified by CTA+ surfactant using 1CEC. (<b>b</b>) XRD patterns of poly(GMA)/organophilic clay nanocomposite (containing 2CEC of CTA+). (<b>c</b>) XRD patterns of poly(GMA)/organophilic clay nanocomposite (containing 1CEC of TBA+). Reproduced with permission from Ref. [<a href="#B71-minerals-14-01216" class="html-bibr">71</a>].</p> Full article ">Figure 8
<p>TEM images of nanocomposites Nano-So4 and Nano-So7 (the abbreviation of these samples is well detailed in the XRD part). Reproduced with permission from Ref. [<a href="#B71-minerals-14-01216" class="html-bibr">71</a>].</p> Full article ">Figure 9
<p>Cellulose acetate/organophilic clay. CA: cellulose acetate; Mag: Maghnite (MMT) modified by CTA+; CA/Mag (3%–10%): nanocomposite obtained at different weights of Maghnite. Reproduced with permission from Ref. [<a href="#B72-minerals-14-01216" class="html-bibr">72</a>].</p> Full article ">Figure 10
<p>The chemical structures of synthetic dyes most frequently studied.</p> Full article ">Figure 11
<p>Dye chemical classification.</p> Full article ">Figure 12
<p>Schematic representations of mechanisms for the elimination of various dyes using polymer/clay composite via electrostatic attraction, surface hydrophobicity, π-π interaction, and hydrogen bonding.</p> Full article ">Figure 13
<p>Number of publications per year on polymer/clay nanocomposites as adsorbents.</p> Full article ">Figure 14
<p>Schematic representation of reduction/degradation mechanisms of dye using polymer/clay nanocomposite.</p> Full article ">
Open AccessArticle
Experimental Study on the Preparation of Cementitious Materials Through the Activation of Lead—Zinc Tailings
by
Xu Wu, Xiuping Xu, Shuqin Li, Xiangmei Li, Dejian Pei, Xiaojun Yang, Xiankun Yu and Xiaoman Zhu
Minerals 2024, 14(12), 1215; https://doi.org/10.3390/min14121215 - 28 Nov 2024
Abstract
The pozzolanic activity of lead–zinc tailings (LZTs) was enhanced through mechanical grinding, enabling the preparation of a lead–zinc tailing based composite cementitious material (LZTCC) by combining LZTs with ground granulated blast furnace slag (GGBS), steel slag (SS), and desulfurized gypsum (DG). The compressive
[...] Read more.
The pozzolanic activity of lead–zinc tailings (LZTs) was enhanced through mechanical grinding, enabling the preparation of a lead–zinc tailing based composite cementitious material (LZTCC) by combining LZTs with ground granulated blast furnace slag (GGBS), steel slag (SS), and desulfurized gypsum (DG). The compressive strength of LZTCC was evaluated under varying water–cement ratios (W/C) and LZTs dosages. The hydration mechanism was studied via phase composition and microstructural analyses of hydration products. The results revealed that the 28-day pozzolanic activity of LZTs improved to 76% after 2 h of mechanical grinding. LZTCC formulated with 60% LZTs, 22% GGBS, 8% SS, and 10% DG achieved compressive strengths of 13.8 MPa at 7 days and 15.7 MPa at 28 days under a W/C ratio of 0.4. XRD and SEM characterization demonstrated that AFt and amorphous C-S-H gel, along with the unreacted LZT particles, contributed to the overall microstructure, while the former two phases played a significant role in the strength development of LZTCC mortar due to their cementitious reactivity. Heavy metal pollution levels were minimized throughout the process, and the research results could provide a scientific basis for the harmless treatment and resource utilization of LZTs.
Full article
(This article belongs to the Special Issue Metallurgy Waste Used for Backfilling Materials)
►▼
Show Figures
Figure 1
Figure 1
<p>Appearance of main experimental materials.</p> Full article ">Figure 2
<p>XRD patterns of LZTs (<b>a</b>), GGBS (<b>b</b>), SS (<b>c</b>) and DG (<b>d</b>).</p> Full article ">Figure 3
<p>The fabrication process for the test piece.</p> Full article ">Figure 4
<p>Effect of mechanical activation of LZTs on their mechanical properties.</p> Full article ">Figure 5
<p>XRD analysis of LZTs grinded for 2 h and ungrinded.</p> Full article ">Figure 6
<p>Effect of LZT content on mechanical properties of samples.</p> Full article ">Figure 7
<p>Influence of the water–cement ratio on the compressive strength of test blocks.</p> Full article ">Figure 8
<p>XRD analysis of LZTCC at different maintenance ages.</p> Full article ">Figure 9
<p>SEM images of LZTCC at different maintenance ages: (<b>a</b>) 3 days, ×500 magnification, (<b>b</b>) 3 days, ×1000 magnification, (<b>c</b>) 7 days, ×500 magnification, (<b>d</b>) 7 days, ×1000 magnification, (<b>e</b>) 28 days, ×500 magnification, and (<b>f</b>) 28 days, ×1000 magnification.</p> Full article ">Figure 10
<p>EM-EDS analysis of the internal morphology of the LZTCC.</p> Full article ">
<p>Appearance of main experimental materials.</p> Full article ">Figure 2
<p>XRD patterns of LZTs (<b>a</b>), GGBS (<b>b</b>), SS (<b>c</b>) and DG (<b>d</b>).</p> Full article ">Figure 3
<p>The fabrication process for the test piece.</p> Full article ">Figure 4
<p>Effect of mechanical activation of LZTs on their mechanical properties.</p> Full article ">Figure 5
<p>XRD analysis of LZTs grinded for 2 h and ungrinded.</p> Full article ">Figure 6
<p>Effect of LZT content on mechanical properties of samples.</p> Full article ">Figure 7
<p>Influence of the water–cement ratio on the compressive strength of test blocks.</p> Full article ">Figure 8
<p>XRD analysis of LZTCC at different maintenance ages.</p> Full article ">Figure 9
<p>SEM images of LZTCC at different maintenance ages: (<b>a</b>) 3 days, ×500 magnification, (<b>b</b>) 3 days, ×1000 magnification, (<b>c</b>) 7 days, ×500 magnification, (<b>d</b>) 7 days, ×1000 magnification, (<b>e</b>) 28 days, ×500 magnification, and (<b>f</b>) 28 days, ×1000 magnification.</p> Full article ">Figure 10
<p>EM-EDS analysis of the internal morphology of the LZTCC.</p> Full article ">
Open AccessArticle
Investigation of the Testing Method of Softening–Melting Properties of Iron-Bearing Materials
by
Kai Fan, Xin Jiang, Xin Zhang, Qingyu Wang, Qiangjian Gao, Haiyan Zheng and Fengman Shen
Minerals 2024, 14(12), 1214; https://doi.org/10.3390/min14121214 - 28 Nov 2024
Abstract
The softening–melting properties of iron-bearing materials play a crucial role in the reduction process in the lumpy zone in the blast furnace (BF) and affect the height, thickness, and shape of the cohesive zone, as well as gas permeability in the BF. A
[...] Read more.
The softening–melting properties of iron-bearing materials play a crucial role in the reduction process in the lumpy zone in the blast furnace (BF) and affect the height, thickness, and shape of the cohesive zone, as well as gas permeability in the BF. A novel softening–melting method was developed based on actual BF production practices, which consistently matches the reduction index and metallization degree observed in actual BF operations compared to the conventional methods. Under the novel softening–melting testing method, the characteristic temperatures (T40 and TS) increase by about 5 °C and 49 °C, respectively, compared to the conventional method. Additionally, the permeability index (S) of the sinter in the novel method is about 707 kPa·°C lower compared to the conventional method. Clearly, the novel method results in higher softening–melting characteristic temperatures for iron-bearing materials compared to the traditional method, more closely matching actual BF conditions. This approach can provide valuable insights for improving gas permeability and enhancing the reduction process of iron-bearing materials in the BF.
Full article
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)
►▼
Show Figures
Figure 1
Figure 1
<p>Schematic diagram of softening–melting method device (RSZ-03).</p> Full article ">Figure 2
<p>Conventional softening–melting method (<b>a</b>) and novel softening–melting method (<b>b</b>) for S&M under load method.</p> Full article ">Figure 3
<p>Schematic diagram of softening–melting property curve.</p> Full article ">Figure 4
<p>Comparison of reduction index (<b>a</b>) and metallization degree (<b>b</b>) of sinter at various temperatures within the cohesive zone.</p> Full article ">Figure 5
<p>Softening–melting characteristic temperatures of the iron-bearing materials with different softening–melting methods.</p> Full article ">Figure 6
<p>Permeability of the iron-bearing materials under different softening–melting methods.</p> Full article ">
<p>Schematic diagram of softening–melting method device (RSZ-03).</p> Full article ">Figure 2
<p>Conventional softening–melting method (<b>a</b>) and novel softening–melting method (<b>b</b>) for S&M under load method.</p> Full article ">Figure 3
<p>Schematic diagram of softening–melting property curve.</p> Full article ">Figure 4
<p>Comparison of reduction index (<b>a</b>) and metallization degree (<b>b</b>) of sinter at various temperatures within the cohesive zone.</p> Full article ">Figure 5
<p>Softening–melting characteristic temperatures of the iron-bearing materials with different softening–melting methods.</p> Full article ">Figure 6
<p>Permeability of the iron-bearing materials under different softening–melting methods.</p> Full article ">
Open AccessArticle
Mineral Deposition on the Rough Walls of a Fracture
by
Nathann Teixeira Rodrigues, Ismael S. S. Carrasco, Vaughan R. Voller and Fábio D. A. Aarão Reis
Minerals 2024, 14(12), 1213; https://doi.org/10.3390/min14121213 - 28 Nov 2024
Abstract
Modeling carbonate growth in fractures and pores is important for understanding carbon sequestration in the environment or when supersaturated solutions are injected into rocks. Here, we study the simple but nontrivial problem of calcite growth on fractures with rough walls of the same
[...] Read more.
Modeling carbonate growth in fractures and pores is important for understanding carbon sequestration in the environment or when supersaturated solutions are injected into rocks. Here, we study the simple but nontrivial problem of calcite growth on fractures with rough walls of the same mineral using kinetic Monte Carlo simulations of attachment and detachment of molecules and scaling approaches. First, we consider wedge-shaped fracture walls whose upper terraces are in the same low-energy planes and show that the valleys are slowly filled by the propagation of parallel monolayer steps in the wedge sides. The growth ceases when the walls reach these low-energy configurations so that a gap between the walls may not be filled. Second, we consider fracture walls with equally separated monolayer steps (vicinal surfaces with roughness below 1 nm) and show that growth by step propagation will eventually clog the fracture gap. In both cases, scaling approaches predict the times to attain the final configurations as a function of the initial geometry and the step-propagation velocity, which is set by the saturation index. The same reasoning applied to a random wall geometry shows that step propagation leads to lateral filling of surface valleys until the wall reaches the low-energy crystalline plane that has the smallest initial density of molecules. Thus, the final configurations of the fracture walls are much more sensitive to the crystallography than to the roughness or the local curvature. The framework developed here may be used to determine those configurations, the times to reach them, and the mass of deposited mineral. Effects of transport limitations are discussed when the fracture gap is significantly narrowed.
Full article
(This article belongs to the Special Issue Mineral Dissolution and Precipitation in Geologic Porous Media)
►▼
Show Figures
Figure 1
Figure 1
<p>A region of a surface in the Kossel crystal where site colors indicate their coordinations: <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> in purple, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> in yellow, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> (kink site) in blue, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> (step site) in red, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> (terrace site) in gray, and <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> in brown. The sites surrounding this region contain molecules that affect site colors at the boundaries.</p> Full article ">Figure 2
<p>(<b>a</b>) Two-dimensional section of a fracture with wedge-shaped walls. The magnified zoom shows a three-dimensional view of a wedge with a small angle <math display="inline"><semantics> <mi>θ</mi> </semantics></math>, which is formed by wide terraces separated by monolayer steps (the bottoms and the tips of the wedges belong to two low-energy planes of the calcite crystal). (<b>b</b>) Two-dimensional <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> section of a fracture whose walls are vicinal surfaces forming angle <math display="inline"><semantics> <mi>θ</mi> </semantics></math> with the <span class="html-italic">z</span> direction.</p> Full article ">Figure 3
<p>(<b>a</b>) Cross-sections (<math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> plane) of a fracture with initially wedge-shaped walls, total length of 400 nm, and angle <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>15</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, in solution with <math display="inline"><semantics> <mrow> <msub> <mo>Ω</mo> <mrow> <mi>m</mi> <mi>o</mi> <mi>d</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mn>3.00</mn> </mrow> </semantics></math>. In all panels, orange lines indicate the projection of the initial walls on the <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> plane. (<b>b</b>) Evolution of the bottom wall of the fracture. Site colors are those defined in <a href="#minerals-14-01213-f001" class="html-fig">Figure 1</a>.</p> Full article ">Figure 4
<p>Results for growth or dissolution in wedge-shaped fracture walls: (<b>a</b>) Ratio <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math> nm and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> for the saturations indicated in the plot. (<b>b</b>) Ratio <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> for the saturations indicated in the plots. (<b>c</b>) Evolution of the growth rate for the same walls and saturations of (<b>b</b>).</p> Full article ">Figure 5
<p>Stationary value of <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced separators="" open="(" close=")"> <msub> <mi>t</mi> <mrow> <mi>s</mi> <mi>t</mi> </mrow> </msub> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> as function of <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mi>l</mi> </mrow> </semantics></math> for different angles <math display="inline"><semantics> <mi>α</mi> </semantics></math>.</p> Full article ">Figure 6
<p>Stationary times <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>s</mi> <mi>t</mi> </mrow> </msub> </semantics></math> as function of <span class="html-italic">l</span> for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> <mo>,</mo> <mn>1</mn> <mo>.</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>.</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p> Full article ">Figure 7
<p>Evolution of the cross-section of a fracture with vicinal surfaces with terrace length of 40 nm. The orange lines indicate the initial walls.</p> Full article ">Figure 8
<p>Scaled time to fill the gap between the vicinal surfaces as a function of the gap distance for two different angles and saturation ratios.</p> Full article ">Figure 9
<p>Expected time evolution of a fracture: (<b>a</b>) initial configuration with rough walls; (<b>b</b>) a configuration during calcite growth; (<b>c</b>) final configuration. Low-energy planes are indicated by parallel lines, with increasing initial density of molecules in the following order in the lower crystal: red solid line; pink dashed line; magenta dashed line; orange dashed line. Brown dashed lines in the intermediate configuration are drawn through the terraces formed around local surface peaks. Flow lines in the fracture spacing are schematically represented.</p> Full article ">Figure 10
<p>Evolution of the cross-section of a rough wall during calcite growth. The initial configuration of the wall is indicated by the orange line and the blue line indicates the low energy plane with the smallest density of molecules in the initial wall.</p> Full article ">Figure A1
<p>Snapshots of a growing surface (from left to right) with initially separated monolayer steps (<math display="inline"><semantics> <mrow> <mi>w</mi> <mo>/</mo> <mi>a</mi> <mo>=</mo> <mn>256</mn> </mrow> </semantics></math>) for <math display="inline"><semantics> <mrow> <msub> <mo>Ω</mo> <mrow> <mi>m</mi> <mi>o</mi> <mi>d</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mn>3.00</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>=</mo> <mn>0.62</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>5.2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math>. Site colors are the same defined in <a href="#minerals-14-01213-f001" class="html-fig">Figure 1</a>.</p> Full article ">Figure A2
<p>Monolayer step velocity as a function of the saturation ratio obtained in simulations (black circles) and AFM studies (blue filled squares [<a href="#B41-minerals-14-01213" class="html-bibr">41</a>] and red filled squares [<a href="#B42-minerals-14-01213" class="html-bibr">42</a>]).</p> Full article ">
<p>A region of a surface in the Kossel crystal where site colors indicate their coordinations: <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> in purple, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> in yellow, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> (kink site) in blue, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> (step site) in red, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> (terrace site) in gray, and <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>6</mn> </mrow> </semantics></math> in brown. The sites surrounding this region contain molecules that affect site colors at the boundaries.</p> Full article ">Figure 2
<p>(<b>a</b>) Two-dimensional section of a fracture with wedge-shaped walls. The magnified zoom shows a three-dimensional view of a wedge with a small angle <math display="inline"><semantics> <mi>θ</mi> </semantics></math>, which is formed by wide terraces separated by monolayer steps (the bottoms and the tips of the wedges belong to two low-energy planes of the calcite crystal). (<b>b</b>) Two-dimensional <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> section of a fracture whose walls are vicinal surfaces forming angle <math display="inline"><semantics> <mi>θ</mi> </semantics></math> with the <span class="html-italic">z</span> direction.</p> Full article ">Figure 3
<p>(<b>a</b>) Cross-sections (<math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> plane) of a fracture with initially wedge-shaped walls, total length of 400 nm, and angle <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>15</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, in solution with <math display="inline"><semantics> <mrow> <msub> <mo>Ω</mo> <mrow> <mi>m</mi> <mi>o</mi> <mi>d</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mn>3.00</mn> </mrow> </semantics></math>. In all panels, orange lines indicate the projection of the initial walls on the <math display="inline"><semantics> <mrow> <mi>x</mi> <mi>z</mi> </mrow> </semantics></math> plane. (<b>b</b>) Evolution of the bottom wall of the fracture. Site colors are those defined in <a href="#minerals-14-01213-f001" class="html-fig">Figure 1</a>.</p> Full article ">Figure 4
<p>Results for growth or dissolution in wedge-shaped fracture walls: (<b>a</b>) Ratio <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math> nm and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> for the saturations indicated in the plot. (<b>b</b>) Ratio <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced open="(" close=")"> <mi>t</mi> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> for the saturations indicated in the plots. (<b>c</b>) Evolution of the growth rate for the same walls and saturations of (<b>b</b>).</p> Full article ">Figure 5
<p>Stationary value of <math display="inline"><semantics> <mrow> <mi>N</mi> <mfenced separators="" open="(" close=")"> <msub> <mi>t</mi> <mrow> <mi>s</mi> <mi>t</mi> </mrow> </msub> </mfenced> <mo>/</mo> <msub> <mi>N</mi> <mi>I</mi> </msub> </mrow> </semantics></math> as function of <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mi>l</mi> </mrow> </semantics></math> for different angles <math display="inline"><semantics> <mi>α</mi> </semantics></math>.</p> Full article ">Figure 6
<p>Stationary times <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>s</mi> <mi>t</mi> </mrow> </msub> </semantics></math> as function of <span class="html-italic">l</span> for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0</mn> <mo>.</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> <mo>,</mo> <mn>1</mn> <mo>.</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>.</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p> Full article ">Figure 7
<p>Evolution of the cross-section of a fracture with vicinal surfaces with terrace length of 40 nm. The orange lines indicate the initial walls.</p> Full article ">Figure 8
<p>Scaled time to fill the gap between the vicinal surfaces as a function of the gap distance for two different angles and saturation ratios.</p> Full article ">Figure 9
<p>Expected time evolution of a fracture: (<b>a</b>) initial configuration with rough walls; (<b>b</b>) a configuration during calcite growth; (<b>c</b>) final configuration. Low-energy planes are indicated by parallel lines, with increasing initial density of molecules in the following order in the lower crystal: red solid line; pink dashed line; magenta dashed line; orange dashed line. Brown dashed lines in the intermediate configuration are drawn through the terraces formed around local surface peaks. Flow lines in the fracture spacing are schematically represented.</p> Full article ">Figure 10
<p>Evolution of the cross-section of a rough wall during calcite growth. The initial configuration of the wall is indicated by the orange line and the blue line indicates the low energy plane with the smallest density of molecules in the initial wall.</p> Full article ">Figure A1
<p>Snapshots of a growing surface (from left to right) with initially separated monolayer steps (<math display="inline"><semantics> <mrow> <mi>w</mi> <mo>/</mo> <mi>a</mi> <mo>=</mo> <mn>256</mn> </mrow> </semantics></math>) for <math display="inline"><semantics> <mrow> <msub> <mo>Ω</mo> <mrow> <mi>m</mi> <mi>o</mi> <mi>d</mi> <mi>e</mi> <mi>l</mi> </mrow> </msub> <mo>=</mo> <mn>3.00</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>=</mo> <mn>0.62</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>5.2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math>. Site colors are the same defined in <a href="#minerals-14-01213-f001" class="html-fig">Figure 1</a>.</p> Full article ">Figure A2
<p>Monolayer step velocity as a function of the saturation ratio obtained in simulations (black circles) and AFM studies (blue filled squares [<a href="#B41-minerals-14-01213" class="html-bibr">41</a>] and red filled squares [<a href="#B42-minerals-14-01213" class="html-bibr">42</a>]).</p> Full article ">
Open AccessArticle
Recovery of Metals from Titanium Ore Using Solvent Extraction Process: Part 1—Transition Metals
by
Nelson Kiprono Rotich, Irena Herdzik-Koniecko, Tomasz Smolinski, Paweł Kalbarczyk, Marcin Sudlitz, Marcin Rogowski, Hagen Stosnach and Andrzej G. Chmielewski
Minerals 2024, 14(12), 1212; https://doi.org/10.3390/min14121212 - 28 Nov 2024
Abstract
Solvent extraction of metals from Ti ore was investigated with a view of enhancing extraction yields by changing the concentration of the ligands, the rate of mixing, the pH, and the temperature of the solution. Norwegian Ti ore was leached with 5M HNO
[...] Read more.
Solvent extraction of metals from Ti ore was investigated with a view of enhancing extraction yields by changing the concentration of the ligands, the rate of mixing, the pH, and the temperature of the solution. Norwegian Ti ore was leached with 5M HNO3 alongside 10% ascorbic acid to obtain a pregnant solution containing transition metals and some rare earth elements (REEs). Part Two of the study will address the recovery of the REEs in the ore. The elemental analysis of solid and aqueous samples was done by two models of total reflection X-ray fluorescence spectrometers (S2 PICOFOX, Bruker Corporation, Berlin, Germany; and T-STAR, Bruker Corporation, Berlin, Germany). The same analysis was repeated using an inductively coupled plasma-mass spectrometer (Perkin Elmer Sciex ELAN DRC II, Perkin Elmer, Waltham, MA, USA). The extraction process and parameters were examined by ICP-MS. The extraction efficiencies were studied under different conditions through the use of various concentrations of ligands at different pHs, temperatures, and mixing rates of the solution. At pH 1.0, 22.5 °C, and a mixing rate of 1400 rpm, the selectivity of 150 g/L trioctyl methyl ammonium chloride (Aliquat 336) was 99% Ti4+, 94% V4⁺, and 82% Hf4+, while 99% of Co2⁺ was recovered at pH 0.8. The extraction efficiency of triethyl phosphate (10% TEP) was 58% Cu2⁺, 68% Mn2⁺, and 63% V4⁺ at 55 °C, 1400 rpm, and without a pH change. Tributyl phosphate (10% TBP) was able to retrieve 87% Cu2⁺ and 78% Zn2⁺ at pH 1.3, 1400 rpm, and 22.5 °C, and 80% Ti4+ at pH 1.2. A 10% solution of 2,4,6-tris (allyloxy)-1,3,5-triazine (TAOT) demonstrated 61% Mn2⁺ and 56% Hf4+ extraction at pH 1.3, 22.5 °C, and 1400 rpm. Under the same conditions, 10% methyl salicylate (MS) was able to recover 56% Hf4+ at pH 1.3. Using 1400 rpm, di (2-ethylhexyl) phosphoric acid (10% D2EHPA) was found to selectively extract 87% Hf4+ at 22.5 °C without a pH change, and around 99% Co2⁺, Ti4+, and Fe2⁺ at pH 1.3. This study provides valuable insights into optimizing solvent extraction conditions for transition metals’ recovery and serves as a precursor to future research on the extraction of REEs from Ti ores. This process is relevant from the environmental and economic perspectives since it provides the best approach to recycling metals to reduce the rate of raw ore mining.
Full article
(This article belongs to the Special Issue Characterization of Minerals and Raw Materials Resources Replenishment)
►▼
Show Figures
Figure 1
Figure 1
<p>The findings were obtained from analyzing the aqueous phases produced by leaching Ti. The analysis was performed before solvent extraction to assess the sample’s composition.</p> Full article ">Figure 2
<p>The results show the application of the S2 PICOFOX TXRF spectrometer to measure Ti ore samples leached by nitrate solution.</p> Full article ">Figure 3
<p>The spectrum displaying the outcome from the T-STAR Mo X-ray tube used to measure the Ti ore dissolved in nitrate solution.</p> Full article ">Figure 4
<p>The spectrum demonstrates the outcome from the W-Bremsstrahlung X-ray tube used to measure the Ti ore leached using nitrate solution.</p> Full article ">Figure 5
<p>The outcomes from TXRF ((<b>A</b>)—T-STAR and (<b>B</b>)—S2 PICOFOX) assessments on the aqueous phases produced by leaching Ti ore. The analysis was done before the solvent extraction stage.</p> Full article ">Figure 6
<p>The recovery of the metals from the nitrate solution at different concentrations of the ligands. (<b>A</b>) Extraction of Hf, Ti, and Fe using different concentrations of D2EHPA; (<b>B</b>) Recovery of Co and Hf using different concentrations of MS; (<b>C</b>) Recovery of Co, Cu, Mn, Zn, V, Ni, and Fe using different concentrations of TEP; (<b>D</b>) Recovery of Co, Hf, Cu, Mn, V and Ni using different concentrations of Aliquat 336; (<b>E</b>) Extraction of Hf, Cu, Mn, Zn and Ni using different concentrations of TBP; (<b>F</b>) Recovery of Co, Hf, Cu, Mn, and Zn utilizing different concentrations of TAOT. The metal recovery by each extractant improved with increasing ligand concentration, attributable to a higher number of binding sites.</p> Full article ">Figure 7
<p>The dependence of the extraction efficiencies on pH changes of the nitrate solutions. (<b>A</b>) Extracting Co, Hf, V, and Ti using 150 g/L of Aliquat 336 at pH values ranging from 0.8 to 1.3; (<b>B</b>) Recovering Co, Hf, Cu, Mn, and V with 10% TEP at different pH levels ranging from 0.8 to 1.3; (<b>C</b>) Recovering Co, Mn, Ti, Ni, and Fe using 10% D2EHPA at varying pH values ranging from 0.8 to 1.3; (<b>D</b>) Recovering Co and Hf with 10% MS at different pH levels ranging from 0.8 to 1.3; (<b>E</b>) Using 10% TAOT to recover Co, Hf, Cu, and Mn at different pH values ranging from 0.8 to 1.3; (<b>F</b>) Recovering Co, Cu, Zn, Ti, and Fe with 10% TBP at various pH values ranging from 0.8 to 1.3. Each ligand demonstrated distinct extraction efficiencies with pH change.</p> Full article ">Figure 8
<p>The influence of the temperature on the extraction of the metals from the nitrate solutions of Ti ore. (<b>A</b>) Extraction of Mn, V, and Ni using 2 to 10% MS at 35, 45 and 55 °C; (<b>B</b>) Recovery of Hf, Ti and Fe using 2 to 10% D2EHPA at 35, 45 and 55 °C; (<b>C</b>) Extracting Hf, Cu, Mn, and Zn utilizing 30 to 150 g/L of Aliquat 336 at 35, 45 and 55 °C; (<b>D</b>) Extraction of Cu, V, and Mn using 2 to 10% TEP at 35, 45 and 55 °C; (<b>E</b>) Extraction of Cu, Zn, Fe, and Mn using 2 to 10% TBP at 35, 45 and 55 °C; (<b>F</b>) Recovering Hf, Mn, V, Cu, Zn, Ni and Fe using 2 to 10% TAOT at 35, 45 and 55 °C. The ligands exhibited varying extraction efficiencies for different metal ions, with recovery rates generally improving at higher solution temperatures.</p> Full article ">Figure 9
<p>The recovery of the metals at different mixing rates of the solutions. (<b>A</b>) Extraction of Co, Zn, and Cu using 2 to 10% TAOT at mixing rates of 250, 600, and 1000 rpm; (<b>B</b>) Recovery of Cu and Zn with 30 to 150 g/L of Aliquat 336 at mixing rates of 250, 600, and 1000 rpm; (<b>C</b>) Extracting Cu, Zn, Mn, and Ni using 2 to 10% TBP at mixing rates of 250, 600, and 1000 rpm; (<b>D</b>) Extracting Co and Hf with 2 to 10% MS at mixing rates of 250, 600, and 1000 rpm; (<b>E</b>) Extracting Fe, Ti, and Hf using 2 to 10% D2EHPA at mixing rates of 250, 600, and 1000 rpm; (<b>F</b>) Extracting Zn, V, and Ni with 2 to 10% TEP at mixing rates of 250, 600, and 1000 rpm. The ligands showed varying extraction efficiencies for different metal ions, with higher mixing rates resulting in improved recoveries.</p> Full article ">
<p>The findings were obtained from analyzing the aqueous phases produced by leaching Ti. The analysis was performed before solvent extraction to assess the sample’s composition.</p> Full article ">Figure 2
<p>The results show the application of the S2 PICOFOX TXRF spectrometer to measure Ti ore samples leached by nitrate solution.</p> Full article ">Figure 3
<p>The spectrum displaying the outcome from the T-STAR Mo X-ray tube used to measure the Ti ore dissolved in nitrate solution.</p> Full article ">Figure 4
<p>The spectrum demonstrates the outcome from the W-Bremsstrahlung X-ray tube used to measure the Ti ore leached using nitrate solution.</p> Full article ">Figure 5
<p>The outcomes from TXRF ((<b>A</b>)—T-STAR and (<b>B</b>)—S2 PICOFOX) assessments on the aqueous phases produced by leaching Ti ore. The analysis was done before the solvent extraction stage.</p> Full article ">Figure 6
<p>The recovery of the metals from the nitrate solution at different concentrations of the ligands. (<b>A</b>) Extraction of Hf, Ti, and Fe using different concentrations of D2EHPA; (<b>B</b>) Recovery of Co and Hf using different concentrations of MS; (<b>C</b>) Recovery of Co, Cu, Mn, Zn, V, Ni, and Fe using different concentrations of TEP; (<b>D</b>) Recovery of Co, Hf, Cu, Mn, V and Ni using different concentrations of Aliquat 336; (<b>E</b>) Extraction of Hf, Cu, Mn, Zn and Ni using different concentrations of TBP; (<b>F</b>) Recovery of Co, Hf, Cu, Mn, and Zn utilizing different concentrations of TAOT. The metal recovery by each extractant improved with increasing ligand concentration, attributable to a higher number of binding sites.</p> Full article ">Figure 7
<p>The dependence of the extraction efficiencies on pH changes of the nitrate solutions. (<b>A</b>) Extracting Co, Hf, V, and Ti using 150 g/L of Aliquat 336 at pH values ranging from 0.8 to 1.3; (<b>B</b>) Recovering Co, Hf, Cu, Mn, and V with 10% TEP at different pH levels ranging from 0.8 to 1.3; (<b>C</b>) Recovering Co, Mn, Ti, Ni, and Fe using 10% D2EHPA at varying pH values ranging from 0.8 to 1.3; (<b>D</b>) Recovering Co and Hf with 10% MS at different pH levels ranging from 0.8 to 1.3; (<b>E</b>) Using 10% TAOT to recover Co, Hf, Cu, and Mn at different pH values ranging from 0.8 to 1.3; (<b>F</b>) Recovering Co, Cu, Zn, Ti, and Fe with 10% TBP at various pH values ranging from 0.8 to 1.3. Each ligand demonstrated distinct extraction efficiencies with pH change.</p> Full article ">Figure 8
<p>The influence of the temperature on the extraction of the metals from the nitrate solutions of Ti ore. (<b>A</b>) Extraction of Mn, V, and Ni using 2 to 10% MS at 35, 45 and 55 °C; (<b>B</b>) Recovery of Hf, Ti and Fe using 2 to 10% D2EHPA at 35, 45 and 55 °C; (<b>C</b>) Extracting Hf, Cu, Mn, and Zn utilizing 30 to 150 g/L of Aliquat 336 at 35, 45 and 55 °C; (<b>D</b>) Extraction of Cu, V, and Mn using 2 to 10% TEP at 35, 45 and 55 °C; (<b>E</b>) Extraction of Cu, Zn, Fe, and Mn using 2 to 10% TBP at 35, 45 and 55 °C; (<b>F</b>) Recovering Hf, Mn, V, Cu, Zn, Ni and Fe using 2 to 10% TAOT at 35, 45 and 55 °C. The ligands exhibited varying extraction efficiencies for different metal ions, with recovery rates generally improving at higher solution temperatures.</p> Full article ">Figure 9
<p>The recovery of the metals at different mixing rates of the solutions. (<b>A</b>) Extraction of Co, Zn, and Cu using 2 to 10% TAOT at mixing rates of 250, 600, and 1000 rpm; (<b>B</b>) Recovery of Cu and Zn with 30 to 150 g/L of Aliquat 336 at mixing rates of 250, 600, and 1000 rpm; (<b>C</b>) Extracting Cu, Zn, Mn, and Ni using 2 to 10% TBP at mixing rates of 250, 600, and 1000 rpm; (<b>D</b>) Extracting Co and Hf with 2 to 10% MS at mixing rates of 250, 600, and 1000 rpm; (<b>E</b>) Extracting Fe, Ti, and Hf using 2 to 10% D2EHPA at mixing rates of 250, 600, and 1000 rpm; (<b>F</b>) Extracting Zn, V, and Ni with 2 to 10% TEP at mixing rates of 250, 600, and 1000 rpm. The ligands showed varying extraction efficiencies for different metal ions, with higher mixing rates resulting in improved recoveries.</p> Full article ">
Open AccessArticle
Assessment of Mercury Uptake by Plants in Former Cinnabar Mining Areas
by
Milan Bauštein, Jiřina Száková, Luka Stefanović, Jana Najmanová, Jiřina Sysalová and Pavel Tlustoš
Minerals 2024, 14(12), 1211; https://doi.org/10.3390/min14121211 - 28 Nov 2024
Abstract
Assessment of the plant’s ability to take up mercury (Hg) from polluted soil was affected by location, plant family, and species in two former cinnabar mining areas in the Czech Republic. At each location, seven sampling points were marked out in the vicinity
[...] Read more.
Assessment of the plant’s ability to take up mercury (Hg) from polluted soil was affected by location, plant family, and species in two former cinnabar mining areas in the Czech Republic. At each location, seven sampling points were marked out in the vicinity of former shafts and dumpsites connected to the mining activity, where representative soil samples and dicotyledonous plants were collected. The individual locations were characterized by specific plant communities, where, in most cases, different plant species were found within one family at both locations. The total Hg content in the soil, as well as gaseous elemental mercury (GEMsoil-air), confirmed elevated levels of this element in the mining-affected environment, with high variability of the data. The low Hg accumulation ability of plants, especially the low root–shoot translocation in most of the plant species, indicated the predominant occurrence of excluders. Among the families, the results showed the exceptional position of the Fabaceae family regarding soil Hg pollution, as the highest Hg content in both shoots and roots was determined for Onobrychis viciifolia. Therefore, the behavior of Fabaceae plants in polluted soil, the mechanisms of their tolerance to high Hg content, and their Hg accumulation ability deserve further research.
Full article
(This article belongs to the Special Issue Geochemical Characteristics and Contamination Risk Assessment of Soil)
►▼
Show Figures
Figure 1
Figure 1
<p>The total Hg contents in the roots and shoots of plants at both locations, regardless of plant species and family; n = 54 for Jedová Hora, and n = 36 for the Horní Luby location. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers. The boxplots marked by the same letter do not significantly differ at <span class="html-italic">p</span> < 0.05.</p> Full article ">Figure 2
<p>The total Hg contents in the roots and shoots of plants at the Jedová Hora location according to the individual plant families; n = 54. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers.</p> Full article ">Figure 3
<p>The total Hg contents in roots and shoots of plants at the Horní Luby location according to the individual plant families; n = 36. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers.</p> Full article ">Figure 4
<p>The BAF values of plants at both locations, regardless of the plant species and family; n = 54 for Jedová Hora, and n = 36 for the Horní Luby location. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers. The boxplots marked by the same letter did not significantly differ at <span class="html-italic">p</span> < 0.05.</p> Full article ">Figure 5
<p>The TF values of plants at both locations, regardless of the plant species and family; n = 54 for Jedová Hora, and n = 36 for the Horní Luby location. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers. The boxplots marked by the same letter did not significantly differ at <span class="html-italic">p</span> < 0.05.</p> Full article ">
<p>The total Hg contents in the roots and shoots of plants at both locations, regardless of plant species and family; n = 54 for Jedová Hora, and n = 36 for the Horní Luby location. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers. The boxplots marked by the same letter do not significantly differ at <span class="html-italic">p</span> < 0.05.</p> Full article ">Figure 2
<p>The total Hg contents in the roots and shoots of plants at the Jedová Hora location according to the individual plant families; n = 54. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers.</p> Full article ">Figure 3
<p>The total Hg contents in roots and shoots of plants at the Horní Luby location according to the individual plant families; n = 36. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers.</p> Full article ">Figure 4
<p>The BAF values of plants at both locations, regardless of the plant species and family; n = 54 for Jedová Hora, and n = 36 for the Horní Luby location. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers. The boxplots marked by the same letter did not significantly differ at <span class="html-italic">p</span> < 0.05.</p> Full article ">Figure 5
<p>The TF values of plants at both locations, regardless of the plant species and family; n = 54 for Jedová Hora, and n = 36 for the Horní Luby location. Horizontal lines drawn inside the boxes indicate the median, x indicates the average, and the points outside the whisker boundaries are outliers. The boxplots marked by the same letter did not significantly differ at <span class="html-italic">p</span> < 0.05.</p> Full article ">
Open AccessArticle
Five Years of Leaching Experiments to Evaluate Land Spreading of a Modified Bauxite Residue Before and After Treatment of Acid Mine Drainage: Sand or Soil Capping and Revegetation
by
Patricia Merdy, Alexandre Parker, Chen Chen and Pierre Hennebert
Minerals 2024, 14(12), 1210; https://doi.org/10.3390/min14121210 - 27 Nov 2024
Abstract
The global generation of bauxite residue necessitates environmentally responsible disposal strategies. This study investigated the long-term (5-year) behavior of bauxite residue whose pH was lowered to 8.5, called modified bauxite residue (MBR), using lysimeters to test various configurations: raw MBR or used MBR
[...] Read more.
The global generation of bauxite residue necessitates environmentally responsible disposal strategies. This study investigated the long-term (5-year) behavior of bauxite residue whose pH was lowered to 8.5, called modified bauxite residue (MBR), using lysimeters to test various configurations: raw MBR or used MBR (UMBR) previously applied for acid mine drainage remediation, sand or soil capping, and revegetation. Throughout the experiment and across all configurations, the pH of the leachates stabilized between 7 and 8 and their salinity decreased. Their low sodium absorption ratio (SAR) indicated minimal risk of material clogging and suitability for salt-tolerant plant growth. Leaching of potentially toxic elements, except vanadium, decreased rapidly after the first year to low levels. Leachate concentrations consistently remained below LD50 for Hyalella azteca and were at least an order of magnitude lower by the experiment’s end, except for first-year chromium. Sand capping performed poorly, while revegetation and soil capping slightly increased leaching, though these were negligible given the low final leaching levels. Revegetated MBR shows promise as a suitable and sustainable solution for managing bauxite residues, provided the pH is maintained above 6.5. This study highlights the importance of long-term assessments and appropriate management strategies for bauxite residue disposal.
Full article
(This article belongs to the Special Issue Geochemical Characteristics and Contamination Risk Assessment of Soil)
►▼
Show Figures
Figure 1
Figure 1
<p>Experimental flow diagram.</p> Full article ">Figure 2
<p>Lysimeters experimental setup. Left: 40-L lysimeters, view and diagram of filling and revegetation; center and right: 700-L lysimeter, shown from the side and from above, respectively.</p> Full article ">Figure 3
<p>Cumulative water input (rainfall + watering) (blue columns), drainage (pink columns) and drainage/water input ratio (green columns) of lysimeters #1, #3, #5 and #16.</p> Full article ">Figure 4
<p>pH and electrical conductivity (EC) of lysimeters #1, #3, #5, #7 and #16.</p> Full article ">Figure 5
<p>Variation of element leaching over time of some selected elements (Al, Na, S, As, Cd, Co, Cr, Cu, Mo, Ni, Pb, V, Zn) in mg kg<sup>−1</sup> of MBR. Sketch of lysimeters, see <a href="#minerals-14-01210-f002" class="html-fig">Figure 2</a>.</p> Full article ">Figure 6
<p>Cumulated element quantity (mg kg<sup>−1</sup> of MBR) leached by the lysimeters during the 5-year experiment for #1, #3, #5, and #7 and the 4-year experiment for #16.</p> Full article ">Figure 7
<p>Concentrations in average annual leachate for the first year and the final year of the experiment. Stars indicate the LD50 concentrations for <span class="html-italic">Hyalella azteca</span> in soft freshwater (hardness 18, red stars) and hard freshwater (hardness 124, orange stars); green stars indicate that the LD50 concentration is over the star value [<a href="#B31-minerals-14-01210" class="html-bibr">31</a>].</p> Full article ">Figure 8
<p>SAR and electrical conductivity (EC) of lysimeters #1, #3, #5, #7 and #16. Circles and squares give first and final-year values, respectively. Areas related to clay dispersion problems were drawn after Hanson et al. [<a href="#B38-minerals-14-01210" class="html-bibr">38</a>]. Hatched area represents high and very high salinity with regard to plant growth [<a href="#B39-minerals-14-01210" class="html-bibr">39</a>].</p> Full article ">Figure 9
<p>Element leaching from vegetated lysimeter (#5 and #6) as a % of the element leaching from the corresponding non-vegetated lysimeter (#3 and #4, respectively). Cumulative leaching during the 5 years of experimentation.</p> Full article ">Figure 10
<p>Water input (rainfall + watering) (blue columns), drainage (pink columns) and drainage/water input ratio (green columns) of lysimeters #2, #4, #6 and #8.</p> Full article ">Figure 11
<p>pH and electrical conductivity (EC) of lysimeters #2, #4, #6 and #8.</p> Full article ">Figure 12
<p>Variation of leaching over time of potentially toxic elements (Al, As, Cd, Co, Cr, Cu, Mo, Ni, Pb, V, Zn) in mg kg<sup>−1</sup> of (MBR+UMBR). Sketch of lysimeters, see <a href="#minerals-14-01210-f002" class="html-fig">Figure 2</a>.</p> Full article ">Figure 13
<p>Element leaching from lysimeter with UMBR as a % of the element leaching from the corresponding lysimeter without UMBR. Cumulative leaching during the 5 years of experimentation.</p> Full article ">
<p>Experimental flow diagram.</p> Full article ">Figure 2
<p>Lysimeters experimental setup. Left: 40-L lysimeters, view and diagram of filling and revegetation; center and right: 700-L lysimeter, shown from the side and from above, respectively.</p> Full article ">Figure 3
<p>Cumulative water input (rainfall + watering) (blue columns), drainage (pink columns) and drainage/water input ratio (green columns) of lysimeters #1, #3, #5 and #16.</p> Full article ">Figure 4
<p>pH and electrical conductivity (EC) of lysimeters #1, #3, #5, #7 and #16.</p> Full article ">Figure 5
<p>Variation of element leaching over time of some selected elements (Al, Na, S, As, Cd, Co, Cr, Cu, Mo, Ni, Pb, V, Zn) in mg kg<sup>−1</sup> of MBR. Sketch of lysimeters, see <a href="#minerals-14-01210-f002" class="html-fig">Figure 2</a>.</p> Full article ">Figure 6
<p>Cumulated element quantity (mg kg<sup>−1</sup> of MBR) leached by the lysimeters during the 5-year experiment for #1, #3, #5, and #7 and the 4-year experiment for #16.</p> Full article ">Figure 7
<p>Concentrations in average annual leachate for the first year and the final year of the experiment. Stars indicate the LD50 concentrations for <span class="html-italic">Hyalella azteca</span> in soft freshwater (hardness 18, red stars) and hard freshwater (hardness 124, orange stars); green stars indicate that the LD50 concentration is over the star value [<a href="#B31-minerals-14-01210" class="html-bibr">31</a>].</p> Full article ">Figure 8
<p>SAR and electrical conductivity (EC) of lysimeters #1, #3, #5, #7 and #16. Circles and squares give first and final-year values, respectively. Areas related to clay dispersion problems were drawn after Hanson et al. [<a href="#B38-minerals-14-01210" class="html-bibr">38</a>]. Hatched area represents high and very high salinity with regard to plant growth [<a href="#B39-minerals-14-01210" class="html-bibr">39</a>].</p> Full article ">Figure 9
<p>Element leaching from vegetated lysimeter (#5 and #6) as a % of the element leaching from the corresponding non-vegetated lysimeter (#3 and #4, respectively). Cumulative leaching during the 5 years of experimentation.</p> Full article ">Figure 10
<p>Water input (rainfall + watering) (blue columns), drainage (pink columns) and drainage/water input ratio (green columns) of lysimeters #2, #4, #6 and #8.</p> Full article ">Figure 11
<p>pH and electrical conductivity (EC) of lysimeters #2, #4, #6 and #8.</p> Full article ">Figure 12
<p>Variation of leaching over time of potentially toxic elements (Al, As, Cd, Co, Cr, Cu, Mo, Ni, Pb, V, Zn) in mg kg<sup>−1</sup> of (MBR+UMBR). Sketch of lysimeters, see <a href="#minerals-14-01210-f002" class="html-fig">Figure 2</a>.</p> Full article ">Figure 13
<p>Element leaching from lysimeter with UMBR as a % of the element leaching from the corresponding lysimeter without UMBR. Cumulative leaching during the 5 years of experimentation.</p> Full article ">
Open AccessArticle
A Fuzzy Gold Mineral Prediction Model Integrating with Knowledge-Driven and Data-Driven: A Case Study of the Hatu Region in Xinjiang, China
by
Yajie Feng, Yongzhi Wang, Cheng Wang, Jiangtao Tian, Shibo Wen, Yanbin Zhou and Yigao Cheng
Minerals 2024, 14(12), 1209; https://doi.org/10.3390/min14121209 - 27 Nov 2024
Abstract
As mineral resources become increasingly scarce, the search for potential ore deposits presents a significant challenge in geological exploration. Subjective factors often constrain traditional knowledge-driven approaches, while purely data-driven methods may overlook the geological significance of data relationships, potentially compromising the accuracy and
[...] Read more.
As mineral resources become increasingly scarce, the search for potential ore deposits presents a significant challenge in geological exploration. Subjective factors often constrain traditional knowledge-driven approaches, while purely data-driven methods may overlook the geological significance of data relationships, potentially compromising the accuracy and reliability of predictions. To address these issues, this study proposes a novel mineral prediction model that integrates fuzzy comprehensive evaluation with both knowledge-driven and data-driven approaches (FCEKDD), resulting in an optimized comprehensive mineral prediction model based on multi-source geoscience data. This model establishes comprehensive exploration indicators based on expert knowledge and quantitatively assesses these indicators through fuzzy evaluation methods to evaluate mineralization potential, thereby delineating exploration targets within the study area. Ultimately, a three-layer predictive framework is constructed using the C-A fractal method. Experimental results indicate that 57.1% of known mineral points are within the primary prediction zone, thus demonstrating the model’s high predictive accuracy. Comparisons with a random forest model reveal that the FCEKDD model has advantages in addressing geoscience data uncertainty and interpreting geological phenomena. This research validates the reliability and applicability of the proposed model in mineral exploration mapping, providing an effective solution for future mineral resource exploration.
Full article
(This article belongs to the Section Mineral Exploration Methods and Applications)
►▼
Show Figures
Figure 1
Figure 1
<p>Geological map of the research area: (<b>a</b>) the geographical location of Xinjiang in China, (<b>b</b>) the geographical location of the Hatu metallogenic belt in the Xinjiang Uygur Autonomous Region, (<b>c</b>) a 1:50,000 geological map of Hatu research area. Legend: 1 = Fluvial Deposit; 2 = lake accumulation; 3 = Eluvium; 4 = Diluviumlayer (Upper Pleistocene to Holocene); 5 = Diluviumlayer (Upper Pleistocene); 6 = Dushanzi Formation; 7 = the third lithological segment of the Genghis Khan Formation; 8 = the second lithological segment of the Genghis Khan Formation; 9 = the first lithological segment of the Genghis Khan Formation; 10 = the second lithological segment of the Tailegula Formation; 11 = the first lithological segment of the Tailegula Formation; 12 = the fourth lithological segment of the Baogutu Formation; 13 = the third lithological segment of Baogutu Formation; 14 = the second lithological segment of Baogutu Formation; 15 = the first lithological segment of Baogutu Formation; 16 = the fourth lithological segment of the Xibeikulas Formation; 17 = the third lithological segment of the Xibeikulas Formation; 18 = the second lithological segment of the Xibeikulas Formation; 19 = the first lithological segment of the Xibeikulas Formation; 20 = the second lithological segment of the Balrek Formation; 21 = the first lithological segment of the Balrek Formation; 22 = Kushkuduk sequence granodiorite; 23 = Akbastao sequence potassium feldspar granite; 24 = Beilu Agaxi sequence granodiorite; 25 = Beilu Agaxi sequence quartz diorite; 26 = quartz vein; 27 = acidic rock vein; 28 = gabbro vein; 29 = gold deposit/gold mine point; 30 = geological boundary/angle unconformity boundary; 31 = phase transition boundary/ceratopsization; 32 = mylonitization zone; 33 = diagonal and numbering; 34 = anticline and number; 35 = regional fault; 36 = reverse; 37 = translational fault/general fault; 38 = inferred fault.</p> Full article ">Figure 2
<p>The framework of the fuzzy comprehensive evaluation model integrating knowledge-driven and data-driven appoaches.</p> Full article ">Figure 3
<p>Workflow for geochemical anomaly information extraction.</p> Full article ">Figure 4
<p>Result of tectonic fracture classification.</p> Full article ">Figure 5
<p>Distribution of single-element geochemical anomalies.</p> Full article ">Figure 6
<p>Results of element correlation analysis.</p> Full article ">Figure 7
<p>Spectral curves of iron-stained and hydroxylated minerals.</p> Full article ">Figure 8
<p>Results of alteration anomaly information extraction. (<b>a</b>) Iron staining alteration anomaly information; (<b>b</b>) hydroxyl alteration anomaly information; (<b>c</b>) superimposed results of iron staining and hydroxyl anomalies.</p> Full article ">Figure 9
<p>Double-logarithmic plot of the C-A method (<b>a</b>) and a grading map of mineralization prospects based on C-A method (<b>b</b>).</p> Full article ">Figure 10
<p>Prediction comparison of FCEKDD method (<b>a</b>) and RF method (<b>b</b>).</p> Full article ">Figure 10 Cont.
<p>Prediction comparison of FCEKDD method (<b>a</b>) and RF method (<b>b</b>).</p> Full article ">Figure 11
<p>Feature comparison in target areas.</p> Full article ">
<p>Geological map of the research area: (<b>a</b>) the geographical location of Xinjiang in China, (<b>b</b>) the geographical location of the Hatu metallogenic belt in the Xinjiang Uygur Autonomous Region, (<b>c</b>) a 1:50,000 geological map of Hatu research area. Legend: 1 = Fluvial Deposit; 2 = lake accumulation; 3 = Eluvium; 4 = Diluviumlayer (Upper Pleistocene to Holocene); 5 = Diluviumlayer (Upper Pleistocene); 6 = Dushanzi Formation; 7 = the third lithological segment of the Genghis Khan Formation; 8 = the second lithological segment of the Genghis Khan Formation; 9 = the first lithological segment of the Genghis Khan Formation; 10 = the second lithological segment of the Tailegula Formation; 11 = the first lithological segment of the Tailegula Formation; 12 = the fourth lithological segment of the Baogutu Formation; 13 = the third lithological segment of Baogutu Formation; 14 = the second lithological segment of Baogutu Formation; 15 = the first lithological segment of Baogutu Formation; 16 = the fourth lithological segment of the Xibeikulas Formation; 17 = the third lithological segment of the Xibeikulas Formation; 18 = the second lithological segment of the Xibeikulas Formation; 19 = the first lithological segment of the Xibeikulas Formation; 20 = the second lithological segment of the Balrek Formation; 21 = the first lithological segment of the Balrek Formation; 22 = Kushkuduk sequence granodiorite; 23 = Akbastao sequence potassium feldspar granite; 24 = Beilu Agaxi sequence granodiorite; 25 = Beilu Agaxi sequence quartz diorite; 26 = quartz vein; 27 = acidic rock vein; 28 = gabbro vein; 29 = gold deposit/gold mine point; 30 = geological boundary/angle unconformity boundary; 31 = phase transition boundary/ceratopsization; 32 = mylonitization zone; 33 = diagonal and numbering; 34 = anticline and number; 35 = regional fault; 36 = reverse; 37 = translational fault/general fault; 38 = inferred fault.</p> Full article ">Figure 2
<p>The framework of the fuzzy comprehensive evaluation model integrating knowledge-driven and data-driven appoaches.</p> Full article ">Figure 3
<p>Workflow for geochemical anomaly information extraction.</p> Full article ">Figure 4
<p>Result of tectonic fracture classification.</p> Full article ">Figure 5
<p>Distribution of single-element geochemical anomalies.</p> Full article ">Figure 6
<p>Results of element correlation analysis.</p> Full article ">Figure 7
<p>Spectral curves of iron-stained and hydroxylated minerals.</p> Full article ">Figure 8
<p>Results of alteration anomaly information extraction. (<b>a</b>) Iron staining alteration anomaly information; (<b>b</b>) hydroxyl alteration anomaly information; (<b>c</b>) superimposed results of iron staining and hydroxyl anomalies.</p> Full article ">Figure 9
<p>Double-logarithmic plot of the C-A method (<b>a</b>) and a grading map of mineralization prospects based on C-A method (<b>b</b>).</p> Full article ">Figure 10
<p>Prediction comparison of FCEKDD method (<b>a</b>) and RF method (<b>b</b>).</p> Full article ">Figure 10 Cont.
<p>Prediction comparison of FCEKDD method (<b>a</b>) and RF method (<b>b</b>).</p> Full article ">Figure 11
<p>Feature comparison in target areas.</p> Full article ">
Open AccessArticle
Experimental Study on the Influence of Rotational Speed on Grinding Efficiency for the Vertical Stirred Mill
by
Biliang Tang, Bo Cheng, Xianzhou Song, Haonan Ji, Yijiang Li and Zhaohua Wang
Minerals 2024, 14(12), 1208; https://doi.org/10.3390/min14121208 - 27 Nov 2024
Abstract
The rotational speed of the agitator is one of the important parameters that affect the grinding efficiency of the vertical stirred mill. Increasing the speed will improve the grinding effect, but it will increase energy consumption, and determining a reasonable speed setting is
[...] Read more.
The rotational speed of the agitator is one of the important parameters that affect the grinding efficiency of the vertical stirred mill. Increasing the speed will improve the grinding effect, but it will increase energy consumption, and determining a reasonable speed setting is a system issue. The effects of different speeds on energy consumption, product particle size, and grinding efficiency were analyzed in this study. An experimental vertical stirred mill was used to grind iron ore, and five different speed parameters from 175 rpm to 350 rpm were set as variables. It was found that increasing the rotational speed will increase the grinding effect, but it will trigger more energy consumption. A new evaluation index to comprehensively reflect the grinding efficiency of the mill, which was defined as the ability of a mill to grind the same product per unit of time and energy consumption, was proposed. The grinding efficiency was calculated when the particle size of iron ore powder decreased to −45, −38, and −28 μm at different speeds. It can be seen that the growth rate of energy consumption is faster than that of the percentage of particle size, which leads to a continuous decrease in grinding efficiency with the increase in rotational speed. If high processing capacity is pursued within a certain period of time, high speed can be chosen, but it will result in energy loss. On the contrary, the low speed can be chosen, if considering grinding economy.
Full article
(This article belongs to the Special Issue Comminution and Comminution Circuits Optimisation: 3rd Edition)
►▼
Show Figures
Figure 1
Figure 1
<p>Experimental prototype and internal structure of the cylinder: (<b>a</b>) experimental prototype; (<b>b</b>) schematic diagram of the internal structure of the cylinder.</p> Full article ">Figure 2
<p>Particle size distribution of feed samples.</p> Full article ">Figure 3
<p>Experiment torque of Group A during 30 min.</p> Full article ">Figure 4
<p>Relationship between the torque and speed of the agitator.</p> Full article ">Figure 5
<p>The energy consumption of the mill: (<b>a</b>) the variation law of energy consumption over time; (<b>b</b>) the variation law of energy consumption over the speed of the agitator.</p> Full article ">Figure 6
<p>The variation law of particle size over various times and speeds: (<b>a</b>) −45 μm; (<b>b</b>) −38 μm (<b>c</b>) −28 μm; (<b>d</b>) particle size of <span class="html-italic">P</span><sub>80.</sub></p> Full article ">Figure 7
<p>The variation law of energy consumption with rotational speed when the particle size reaches the same percentage.</p> Full article ">Figure 8
<p>The variation law of grinding efficiency with rotational speed: (<b>a</b>) −45 μm; (<b>b</b>) −38 μm; (<b>c</b>) −28 μm.</p> Full article ">Figure 9
<p>The variation law of energy consumption, percentage of particle size, and grinding efficiency with rotational speed.</p> Full article ">
<p>Experimental prototype and internal structure of the cylinder: (<b>a</b>) experimental prototype; (<b>b</b>) schematic diagram of the internal structure of the cylinder.</p> Full article ">Figure 2
<p>Particle size distribution of feed samples.</p> Full article ">Figure 3
<p>Experiment torque of Group A during 30 min.</p> Full article ">Figure 4
<p>Relationship between the torque and speed of the agitator.</p> Full article ">Figure 5
<p>The energy consumption of the mill: (<b>a</b>) the variation law of energy consumption over time; (<b>b</b>) the variation law of energy consumption over the speed of the agitator.</p> Full article ">Figure 6
<p>The variation law of particle size over various times and speeds: (<b>a</b>) −45 μm; (<b>b</b>) −38 μm (<b>c</b>) −28 μm; (<b>d</b>) particle size of <span class="html-italic">P</span><sub>80.</sub></p> Full article ">Figure 7
<p>The variation law of energy consumption with rotational speed when the particle size reaches the same percentage.</p> Full article ">Figure 8
<p>The variation law of grinding efficiency with rotational speed: (<b>a</b>) −45 μm; (<b>b</b>) −38 μm; (<b>c</b>) −28 μm.</p> Full article ">Figure 9
<p>The variation law of energy consumption, percentage of particle size, and grinding efficiency with rotational speed.</p> Full article ">
Open AccessArticle
Genetic Type and Formation Evolution of Mantle-Derived Olivine in Ultramafic Xenolith of Damaping Basalt, Northern North China Block
by
Cun Zhang, Fan Yang, Zengsheng Li, Leon Bagas, Lu Niu, Xinyi Zhu and Jianjun Li
Minerals 2024, 14(12), 1207; https://doi.org/10.3390/min14121207 - 27 Nov 2024
Abstract
Olivine in deep-seated ultramafic xenoliths beneath the North China Block serves as a crucial proxy for decoding the compositions, properties, and evolution of the lithospheric mantle. Here, we conduct an investigation on olivine (including gem-grade) hosted in ultramafic xenoliths from Damaping basalt in
[...] Read more.
Olivine in deep-seated ultramafic xenoliths beneath the North China Block serves as a crucial proxy for decoding the compositions, properties, and evolution of the lithospheric mantle. Here, we conduct an investigation on olivine (including gem-grade) hosted in ultramafic xenoliths from Damaping basalt in the northern part of the North China Block. This contribution presents the results from petrographic, Raman spectroscopic, and major and trace elemental studies of olivine, with the aim of characterising the formation environment and genetic type of the olivine. The analysed olivine samples are characterised by high Mg# values (close to 91%) possessing refractory to fertile features and doublet bands with unit Raman spectra beams of 822 and 853 cm−1, which are indicative of a forsterite signature. Major and trace geochemistry of olivine indicates the presence of mantle xenolith olivine. All the analytical olivine assays ≤0.1 wt % CaO, ~40 wt % SiO2, and ≤0.05 wt % Al2O3. Furthermore, olivine displays significantly different concentrations of Ti, Y, Sc, V, Co, and Ni. The Ni/Co values in olivine range from 21.21 to 22.98, indicating that the crystallisation differentiation of basic magma relates to oceanic crust recycling. The V/Sc values in mantle/xenolith olivine vary from 0.54 to 2.64, indicating a more oxidised state of the mantle. Rare earth element (REE) patterns show that the LREEs and HREEs of olivine host obviously differentiated characteristics. The HREE enrichments of olivine and the LREE depletion of clinopyroxene further assert that the mantle in the Damaping area underwent partial melting. The wide variations of Mg# values in olivine and the Cr# values in clinopyroxene, along with major element geochemistry indicate transitional characteristics of different peridotite xenoliths. This is possibly indicative of a newly accreted lithospheric mantle interaction with an old lithospheric mantle at the time of the basaltic eruption during the Paleozoic to Cenozoic.
Full article
(This article belongs to the Section Mineral Deposits)
►▼
Show Figures
Figure 1
Figure 1
<p>Location of the Damaping region showing (<b>a</b>) the tectonic framework of the northern North China Block, showing different sub-divisions and the location of the study area; and (<b>b</b>) the geology of the Damaping olivine deposits (modified after Yu et al., 2005) [<a href="#B29-minerals-14-01207" class="html-bibr">29</a>].</p> Full article ">Figure 2
<p>Representative field and sample photographs. (<b>a</b>) Damaping olivine open pit showing major volcanic rock; (<b>b</b>) basalt with deep-seated olivine; (<b>c</b>) fresh basalt without mantle xenoliths; (<b>d</b>) basalt with abundant olivine including gem-grade grains displaying different colours; (<b>e</b>) large bright green gem-grade olivine separated from xenoliths; and (<b>f</b>) green olivine with impurities (mainly magnetite) under binocular microscope. Note: the olivine (yellow circle in figure (<b>e</b>)) is magnified under a gem microscope.</p> Full article ">Figure 3
<p>Representative photomicrographs under cross-polarised light showing (<b>a</b>–<b>d</b>) mineral assemblages of olivine samples. Mineral abbreviations: Ol—olivine; Cpx—clinopyroxene; Opx—orthopyroxene.</p> Full article ">Figure 4
<p>Representative Raman spectra of olivine in the ultramafic xenolith from Damaping basalt, showing forsterite peaks.</p> Full article ">Figure 5
<p>Geochemical diagrams showing the classification of pyroxenes (modified after Morimoto (1988) [<a href="#B38-minerals-14-01207" class="html-bibr">38</a>] and Zhang et al. (2023) [<a href="#B39-minerals-14-01207" class="html-bibr">39</a>]); two main types of pyroxene are identified based on the diagram, with red and purple icons indicating clinopyroxene and orthopyroxene, respectively.</p> Full article ">Figure 6
<p>Plot of Mg# vs. SiO<sub>2</sub>, CaO, Na<sub>2</sub>O, and Al<sub>2</sub>O<sub>3</sub> for assays of the olivine samples. (<b>a</b>) Mg# vs. SiO<sub>2</sub> of olivine samples from mantle xenoliths. (<b>b</b>) Mg# vs. CaO of olivine from mantle xenoliths. (<b>c</b>) Mg# vs. Na<sub>2</sub>O of olivine from mantle xenoliths. (<b>d</b>) Mg# vs. Al<sub>2</sub>O<sub>3</sub> of olivine from mantle xenoliths.</p> Full article ">Figure 7
<p>Normalised assays for the two types of olivine showing (<b>a</b>,<b>b</b>) chondrite-normalised REE patterns and primitive mantle-normalised trace element variation diagram in olivine; (<b>c</b>,<b>d</b>) chondrite-normalised REE patterns and primitive mantle-normalised trace element variation diagram in clinopyroxene (Cpx); (<b>e</b>,<b>f</b>) chondrite-normalised REE pattern and primitive mantle-normalised trace element variation diagram of the studied orthopyroxene (Opx). The geochemical data of upper crust are from Taylor and Mclennan (1995) [<a href="#B40-minerals-14-01207" class="html-bibr">40</a>], and of chondrite are from McDonough and Sun (1995) [<a href="#B41-minerals-14-01207" class="html-bibr">41</a>].</p> Full article ">Figure 8
<p>(<b>a</b>) Mg# vs. MnO of olivine samples from mantle xenoliths. (<b>b</b>) Al<sub>2</sub>O<sub>3</sub> vs. Mg# of clinopyroxene from mantle xenoliths. (<b>c</b>) Al<sub>2</sub>O<sub>3</sub> vs. Na<sub>2</sub>O of clinopyroxene from mantle xenoliths. Previous data (olivine and clinopyroxene in websterite xenoliths from Hannuoba) are cited by Duan et al. (2022) [<a href="#B45-minerals-14-01207" class="html-bibr">45</a>]. The diagrams are plotted based on previous investigations [<a href="#B46-minerals-14-01207" class="html-bibr">46</a>,<a href="#B47-minerals-14-01207" class="html-bibr">47</a>,<a href="#B48-minerals-14-01207" class="html-bibr">48</a>,<a href="#B49-minerals-14-01207" class="html-bibr">49</a>]. (<b>d</b>) Mg# vs. Cr# values for clinopyroxene (Cpx). Published related geochemical data (Mg# and Cr#) are from Yu et al. (2006) [<a href="#B50-minerals-14-01207" class="html-bibr">50</a>].</p> Full article ">
<p>Location of the Damaping region showing (<b>a</b>) the tectonic framework of the northern North China Block, showing different sub-divisions and the location of the study area; and (<b>b</b>) the geology of the Damaping olivine deposits (modified after Yu et al., 2005) [<a href="#B29-minerals-14-01207" class="html-bibr">29</a>].</p> Full article ">Figure 2
<p>Representative field and sample photographs. (<b>a</b>) Damaping olivine open pit showing major volcanic rock; (<b>b</b>) basalt with deep-seated olivine; (<b>c</b>) fresh basalt without mantle xenoliths; (<b>d</b>) basalt with abundant olivine including gem-grade grains displaying different colours; (<b>e</b>) large bright green gem-grade olivine separated from xenoliths; and (<b>f</b>) green olivine with impurities (mainly magnetite) under binocular microscope. Note: the olivine (yellow circle in figure (<b>e</b>)) is magnified under a gem microscope.</p> Full article ">Figure 3
<p>Representative photomicrographs under cross-polarised light showing (<b>a</b>–<b>d</b>) mineral assemblages of olivine samples. Mineral abbreviations: Ol—olivine; Cpx—clinopyroxene; Opx—orthopyroxene.</p> Full article ">Figure 4
<p>Representative Raman spectra of olivine in the ultramafic xenolith from Damaping basalt, showing forsterite peaks.</p> Full article ">Figure 5
<p>Geochemical diagrams showing the classification of pyroxenes (modified after Morimoto (1988) [<a href="#B38-minerals-14-01207" class="html-bibr">38</a>] and Zhang et al. (2023) [<a href="#B39-minerals-14-01207" class="html-bibr">39</a>]); two main types of pyroxene are identified based on the diagram, with red and purple icons indicating clinopyroxene and orthopyroxene, respectively.</p> Full article ">Figure 6
<p>Plot of Mg# vs. SiO<sub>2</sub>, CaO, Na<sub>2</sub>O, and Al<sub>2</sub>O<sub>3</sub> for assays of the olivine samples. (<b>a</b>) Mg# vs. SiO<sub>2</sub> of olivine samples from mantle xenoliths. (<b>b</b>) Mg# vs. CaO of olivine from mantle xenoliths. (<b>c</b>) Mg# vs. Na<sub>2</sub>O of olivine from mantle xenoliths. (<b>d</b>) Mg# vs. Al<sub>2</sub>O<sub>3</sub> of olivine from mantle xenoliths.</p> Full article ">Figure 7
<p>Normalised assays for the two types of olivine showing (<b>a</b>,<b>b</b>) chondrite-normalised REE patterns and primitive mantle-normalised trace element variation diagram in olivine; (<b>c</b>,<b>d</b>) chondrite-normalised REE patterns and primitive mantle-normalised trace element variation diagram in clinopyroxene (Cpx); (<b>e</b>,<b>f</b>) chondrite-normalised REE pattern and primitive mantle-normalised trace element variation diagram of the studied orthopyroxene (Opx). The geochemical data of upper crust are from Taylor and Mclennan (1995) [<a href="#B40-minerals-14-01207" class="html-bibr">40</a>], and of chondrite are from McDonough and Sun (1995) [<a href="#B41-minerals-14-01207" class="html-bibr">41</a>].</p> Full article ">Figure 8
<p>(<b>a</b>) Mg# vs. MnO of olivine samples from mantle xenoliths. (<b>b</b>) Al<sub>2</sub>O<sub>3</sub> vs. Mg# of clinopyroxene from mantle xenoliths. (<b>c</b>) Al<sub>2</sub>O<sub>3</sub> vs. Na<sub>2</sub>O of clinopyroxene from mantle xenoliths. Previous data (olivine and clinopyroxene in websterite xenoliths from Hannuoba) are cited by Duan et al. (2022) [<a href="#B45-minerals-14-01207" class="html-bibr">45</a>]. The diagrams are plotted based on previous investigations [<a href="#B46-minerals-14-01207" class="html-bibr">46</a>,<a href="#B47-minerals-14-01207" class="html-bibr">47</a>,<a href="#B48-minerals-14-01207" class="html-bibr">48</a>,<a href="#B49-minerals-14-01207" class="html-bibr">49</a>]. (<b>d</b>) Mg# vs. Cr# values for clinopyroxene (Cpx). Published related geochemical data (Mg# and Cr#) are from Yu et al. (2006) [<a href="#B50-minerals-14-01207" class="html-bibr">50</a>].</p> Full article ">
Open AccessArticle
Transformation of Mackinawite to Interlayered Greigite-Pyrrhotite and Pyrite in the Gaoping Submarine Canyon Sediments off Southwestern Taiwan
by
Ko-Chun Huang and Wei-Teh Jiang
Minerals 2024, 14(12), 1206; https://doi.org/10.3390/min14121206 - 26 Nov 2024
Abstract
Iron monosulfides and neoformed pyrite below the sulfate–methane transition zone (SMTZ) of rapidly accumulating turbiditic sediments from the Gaoping submarine canyon off southwestern Taiwan were examined by SEM-EDS-EBSD, HRTEM, and HAADF STEM to investigate their microstructural characteristics and processes of formation and transformation.
[...] Read more.
Iron monosulfides and neoformed pyrite below the sulfate–methane transition zone (SMTZ) of rapidly accumulating turbiditic sediments from the Gaoping submarine canyon off southwestern Taiwan were examined by SEM-EDS-EBSD, HRTEM, and HAADF STEM to investigate their microstructural characteristics and processes of formation and transformation. Within a few meters below the SMTZ, mackinawite (Mkw) is largely replaced by interlayered greigite-pyrrhotite (Grg-Po) with {111}Grg//{001}Po and ⟨110⟩Grg//⟨110⟩Po, followed by pyrite neoformation in clusters of disseminated matrix grains consisting of coalescing pyrite microcrystals, arrays of polycrystalline interlayer pyrite grains between the cleavage planes of layer silicates, with each grain’s core having inclusions of interlayered Grg-Po locally containing relict Mkw, and amassed pyrite microcrystals on the surface of porous interlayered Grg-Po micronodules. In the deeper sediments, neoformed pyrite is absent and Mkw is largely preserved, with partial replacement by interlayered Grg-Po having an overall topotactic relationship of ⟨110⟩Grg//⟨110 ⟩Po//⟨100⟩Mkw and {111}Grg//(001)Po//~{011}Mkw and a sharp reaction front without transitional profiles. The mineral grain boundaries and structural discontinuities with Mkw resulting from extensive interlayering between Grg {111} cubic close-packed segments and Po {001} hexagonal close-packed layers could serve as conduits for fluid flow and mass transport to drive the replacement reaction. The conversion of Mkw to metastable interlayered Grg-Po is inferred to occur through interface-coupled dissolution–reprecipitation processes associated with partial oxidation while the partial replacement of interlayered Grg-Po ± minor relict Mkw by pyrite microcrystals with irregular grain boundaries and orientations probably occurred via a dissolution–precipitation mechanism. Mkw could be initially formed by sulfate reduction driven by anaerobic oxidation of methane in reactive iron-rich sediments in paleo-SMTZs and subsequently transformed into interlayered Grg-Po followed by pyrite neoformation in the sulfidization front below the SMTZ or recent SMTZs in the Gaoping submarine canyon sediments.
Full article
(This article belongs to the Section Crystallography and Physical Chemistry of Minerals & Nanominerals)
►▼
Show Figures
Figure 1
Figure 1
<p>Sampling location of the MD178-10-3291 sediment core on the west bank of the Gaoping submarine canyon in the lower-slope domain offshore SW Taiwan. DF = the deformation front; LSD = the low-slope domain; USD = the upper-slope domain; OOST = the out-of-sequence thrust; GWR = the Good Weather ridge; FLR = the Fangliao ridge.</p> Full article ">Figure 2
<p>Scanning electron micrographs of (<b>a</b>) a cluster of disseminated pyrite grains (Py) in the sediment matrix and an interlayered Grg-Po (Grg*) micronodule fringed with neoformed pyrite (1434–1439 cmbsf); (<b>b</b>) an enlargement of the small black frame area in (<b>a</b>); (<b>c</b>) arrays of interlayer pyrite grains between the cleavage planes of a chlorite stack (1434–1439 cmsbf); (<b>d</b>) a magnification of interlayered Grg-Po (+ relict Mkw) inclusions (Grg* + Mkw) in the core of a pyrite grain in the frame area in (<b>c</b>); (<b>e</b>) an iron-monosulfide micronodule consisting of dense aggregates of Mkw in association with lesser interlayered Grg-Po and relict layer silicates (2037–2042 cmbsf); and (<b>f</b>) an enlargement of the framed area in (<b>e)</b> exhibiting interweaved microtextures. The areas selected for EBSD analyses or FIB-TEM specimen preparation are indicated in (<b>a</b>,<b>d</b>,<b>f</b>).</p> Full article ">Figure 3
<p>EBSD phase and inverse pole figure (IPF) maps of (<b>a</b>,<b>b</b>) disseminated matrix pyrite grains, (<b>c</b>,<b>d</b>) arrays of interlayer pyrite grains, and (<b>e</b>,<b>f</b>) amassed pyrite microcrystals in a portion of an interlayered Grg-Po micronodule in the 1434–1439 cmbsf sediment. Layer silicate and Grg-dominant areas were unindexed (different gray-scale levels) due to poor EBSD signals. For conciseness, only pyrite was indexed and is displayed in color. Most pyrite grains consist of multiple microcrystals of different crystal orientations implied by different colors in the IPF maps. Color legends for mineral phases in the phase maps and pyrite orientations in the IPF maps are inserted in (<b>a</b>,<b>b</b>), respectively. Py = pyrite; Sd = siderite; Qz = quartz; Ab = albite. Mineral identification was assisted by EDS.</p> Full article ">Figure 4
<p>TEM micrographs of Grg-dominant inclusions in the core of an interlayer pyrite grain in the 1434–1439 cmbsf sediment, illustrating (<b>a</b>) multiple anhedral pyrite crystals (Py) enclosing Grg-dominant interlayered Grg-Po microcrystals (Grg*) and relict chlorite layers (Chl) and (<b>b</b>) infiltrative pyrite growth around a subhedral Grg-dominant crystal (the lower left corner in a) oriented in the 〈001〉<sub>Grg</sub> zone and a number of small Grg-dominant microcrystals largely elongated along 〈110〉<sub>Grg</sub> with parallel {001}<sub>Grg</sub>.</p> Full article ">Figure 5
<p>Electron micrographs of an interlayered Grg-Po microcrystal enclosed within an interlayer pyrite grain at 1434–1439 cmbsf: (<b>a</b>) the subhedral outline of a Grg-dominant composite microcrystal (the upper-right frame of <a href="#minerals-14-01206-f004" class="html-fig">Figure 4</a>a) bordered by {001}<sub>Grg</sub> faces with a slight elongation along 〈110〉<sub>Grg</sub> or equivalent directions; (<b>b</b>) the corresponding diffraction pattern having parallel Grg and Po 〈110〉 zones with the Grg <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hh</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows overlain by the Po 00<span class="html-italic">l</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows, respectively; (<b>c</b>) {111} lattice-fringe image (frame (<b>c</b>) in (<b>a</b>)) with nonperiodic contrast features; (<b>d</b>) HAADF-STEM image of part of the microcrystal in (<b>a</b>) displaying rhombic domains of Grg framed by relatively bright Po (001) layers parallel to Grg {111}, consistent with an occurrence of interlayered Grg-Po; (<b>e</b>) high-magnification Fourier filtered HAADF-STEM image of the composite microcrystal illustrating rhombic arrays of bright spots corresponding to the projected 〈110〉-parallel edge-sharing octahedral chains of Grg intervened by Po {001} octahedral layers parallel to {111} of the Grg domains; and (<b>f</b>) Fourier filtered HAADF-STEM image (frame (<b>f</b>) in (<b>a</b>)) showing an area with ~0.5 nm {001}<sub>Mkw</sub> lattice fringes consistent with an enclosure of relict Mkw by interlayered Grg-Po.</p> Full article ">Figure 6
<p>TEM micrographs of part of an iron-monosulfide micronodule in the 1434–1439 cmbsf sediment: (<b>a</b>) the micronodule is associated with microfissures and neoformed pyrite; (<b>b</b>) pyrite neoformation in a fissure of the micronodule (black arrow in <b>a</b>); (<b>c</b>) occurrence of relict phyllosilicate layers (arrowed) and 〈110〉-elongated Grg-Po composite microcrystals in an alternating manner, locally associated with infiltrative growth of neoformed pyrite; (<b>d</b>) the diffraction pattern of an elongated Grg-Po microcrystal (Grg and Po 〈110〉 zones) having <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hh</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows of Grg superimposed by the 00<span class="html-italic">l</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows of Po, respectively; (<b>e</b>,<b>f</b>) nonperiodic {111}-parallel contrast features ascribed to the interlayering of Po (001) layers within the Grg-dominant microcrystals.</p> Full article ">Figure 7
<p>Electron micrographs of part of an iron-monosulfide micronodule in the 2037–2042 cmbsf sediment: (<b>a</b>) HAADF-STEM image of an elongated iron-monosulfide aggregate consisting of mackinawite (Mkw) and interlayered Grg-Po (Grg-Po); (<b>b</b>) the elongation of the composite grain parallel to Grg 〈110〉 and mackinawite 〈010〉 as implied by the corresponding diffraction pattern in the Grg 〈110〉 and mackinawite 〈100〉 directions; (<b>c</b>,<b>d</b>) HAADF-STEM images of frames (<b>c</b>,<b>d</b>) in a illustrating zigzag boundaries between mackinawite and interlayered Grg-Po defined by Grg {111} and mackinawite {011} and in part by planar boundaries parallel to Grg and Mkw {001}; (<b>e</b>,<b>f</b>) TEM images of interlayered Grg-Po elongated along 〈110〉<sub>Grg</sub> with crossed contrast features parallel to Grg {111} and Po (001) in the absence of Mkw.</p> Full article ">Figure 8
<p>Electron micrographs of oriented Mkw and interlayered Grg-Po in an iron-monosulfide micronodule in the 2037–2042 cmbsf sediment: (<b>a</b>) TEM image of the growth microtextures of Mkw and interlayered Grg-Po (Grg*) in a composite grain of the micronodule; (<b>b</b>) the corresponding diffraction pattern with 〈100〉<sub>Mkw</sub>//〈110〉<sub>Grg</sub>//〈110〉<sub>Po</sub>, (001)<sub>Mkw</sub>//(001)<sub>Grg</sub>, and (001)<sub>Po</sub>//{111}<sub>Grg</sub>; (<b>c</b>,<b>d</b>) 〈100〉-oriented Mkw (frame c in a) having dislocations in {001} lattice fringes in company with 〈110〉-oriented Grg-dominant domains having crossed {111} fringes (frame (<b>d</b>) in (<b>a</b>)) with varied contrast due to oriented growth of intertwined Po (001) layers; (<b>e</b>,<b>f</b>) Fourier filtered HAADF-STEM images of topotactic interfaces between Mkw and interlayered Grg-Po domains.</p> Full article ">Figure 9
<p>Schematic diagram of crystallographic relationships among Mkw, Grg, and Po viewed from [110]<sub>Grg</sub>, [110]<sub>Po</sub>, and [100]<sub>Mkw</sub> with close-packed planes of sulfur atoms (<span style="color:yellow">•</span>) labeled by A, B, and C to indicate sulfur stacking sequences. Rhombic domains of Grg consist of <span class="html-italic">ccp</span> planes of sulfur atoms and alternating [110]<sub>Grg</sub>-projected columns of edge-sharing octahedral sites fully (<span style="color:#0000CC">●</span>) and half (<span style="color:#0066FF">●</span>) occupied by iron, in association with double columns of tetrahedral iron sites (each half occupied by iron) at open rhombuses. Po slabs are composed of <span class="html-italic">hcp</span> planes of sulfur atoms and (001)<sub>Po</sub> layers of [110]<sub>Po</sub>-projected edge-sharing octahedral sites fully (<span style="color:#FF0066">●</span>) and partially (<span style="color:#FF9933">●</span>) occupied by iron. Mkw (distorted ccp sulfur) is made up of (001)<sub>Mkw</sub> tetrahedral iron sheets connected by van der Waals forces. It is important to note the presence of compulsory structural discontinuities between intersecting Po layers and at intersections between Mkw and Grg and Po due to the mixed occurrence of of different sulfur stacking sequences.</p> Full article ">
<p>Sampling location of the MD178-10-3291 sediment core on the west bank of the Gaoping submarine canyon in the lower-slope domain offshore SW Taiwan. DF = the deformation front; LSD = the low-slope domain; USD = the upper-slope domain; OOST = the out-of-sequence thrust; GWR = the Good Weather ridge; FLR = the Fangliao ridge.</p> Full article ">Figure 2
<p>Scanning electron micrographs of (<b>a</b>) a cluster of disseminated pyrite grains (Py) in the sediment matrix and an interlayered Grg-Po (Grg*) micronodule fringed with neoformed pyrite (1434–1439 cmbsf); (<b>b</b>) an enlargement of the small black frame area in (<b>a</b>); (<b>c</b>) arrays of interlayer pyrite grains between the cleavage planes of a chlorite stack (1434–1439 cmsbf); (<b>d</b>) a magnification of interlayered Grg-Po (+ relict Mkw) inclusions (Grg* + Mkw) in the core of a pyrite grain in the frame area in (<b>c</b>); (<b>e</b>) an iron-monosulfide micronodule consisting of dense aggregates of Mkw in association with lesser interlayered Grg-Po and relict layer silicates (2037–2042 cmbsf); and (<b>f</b>) an enlargement of the framed area in (<b>e)</b> exhibiting interweaved microtextures. The areas selected for EBSD analyses or FIB-TEM specimen preparation are indicated in (<b>a</b>,<b>d</b>,<b>f</b>).</p> Full article ">Figure 3
<p>EBSD phase and inverse pole figure (IPF) maps of (<b>a</b>,<b>b</b>) disseminated matrix pyrite grains, (<b>c</b>,<b>d</b>) arrays of interlayer pyrite grains, and (<b>e</b>,<b>f</b>) amassed pyrite microcrystals in a portion of an interlayered Grg-Po micronodule in the 1434–1439 cmbsf sediment. Layer silicate and Grg-dominant areas were unindexed (different gray-scale levels) due to poor EBSD signals. For conciseness, only pyrite was indexed and is displayed in color. Most pyrite grains consist of multiple microcrystals of different crystal orientations implied by different colors in the IPF maps. Color legends for mineral phases in the phase maps and pyrite orientations in the IPF maps are inserted in (<b>a</b>,<b>b</b>), respectively. Py = pyrite; Sd = siderite; Qz = quartz; Ab = albite. Mineral identification was assisted by EDS.</p> Full article ">Figure 4
<p>TEM micrographs of Grg-dominant inclusions in the core of an interlayer pyrite grain in the 1434–1439 cmbsf sediment, illustrating (<b>a</b>) multiple anhedral pyrite crystals (Py) enclosing Grg-dominant interlayered Grg-Po microcrystals (Grg*) and relict chlorite layers (Chl) and (<b>b</b>) infiltrative pyrite growth around a subhedral Grg-dominant crystal (the lower left corner in a) oriented in the 〈001〉<sub>Grg</sub> zone and a number of small Grg-dominant microcrystals largely elongated along 〈110〉<sub>Grg</sub> with parallel {001}<sub>Grg</sub>.</p> Full article ">Figure 5
<p>Electron micrographs of an interlayered Grg-Po microcrystal enclosed within an interlayer pyrite grain at 1434–1439 cmbsf: (<b>a</b>) the subhedral outline of a Grg-dominant composite microcrystal (the upper-right frame of <a href="#minerals-14-01206-f004" class="html-fig">Figure 4</a>a) bordered by {001}<sub>Grg</sub> faces with a slight elongation along 〈110〉<sub>Grg</sub> or equivalent directions; (<b>b</b>) the corresponding diffraction pattern having parallel Grg and Po 〈110〉 zones with the Grg <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hh</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows overlain by the Po 00<span class="html-italic">l</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows, respectively; (<b>c</b>) {111} lattice-fringe image (frame (<b>c</b>) in (<b>a</b>)) with nonperiodic contrast features; (<b>d</b>) HAADF-STEM image of part of the microcrystal in (<b>a</b>) displaying rhombic domains of Grg framed by relatively bright Po (001) layers parallel to Grg {111}, consistent with an occurrence of interlayered Grg-Po; (<b>e</b>) high-magnification Fourier filtered HAADF-STEM image of the composite microcrystal illustrating rhombic arrays of bright spots corresponding to the projected 〈110〉-parallel edge-sharing octahedral chains of Grg intervened by Po {001} octahedral layers parallel to {111} of the Grg domains; and (<b>f</b>) Fourier filtered HAADF-STEM image (frame (<b>f</b>) in (<b>a</b>)) showing an area with ~0.5 nm {001}<sub>Mkw</sub> lattice fringes consistent with an enclosure of relict Mkw by interlayered Grg-Po.</p> Full article ">Figure 6
<p>TEM micrographs of part of an iron-monosulfide micronodule in the 1434–1439 cmbsf sediment: (<b>a</b>) the micronodule is associated with microfissures and neoformed pyrite; (<b>b</b>) pyrite neoformation in a fissure of the micronodule (black arrow in <b>a</b>); (<b>c</b>) occurrence of relict phyllosilicate layers (arrowed) and 〈110〉-elongated Grg-Po composite microcrystals in an alternating manner, locally associated with infiltrative growth of neoformed pyrite; (<b>d</b>) the diffraction pattern of an elongated Grg-Po microcrystal (Grg and Po 〈110〉 zones) having <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hh</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows of Grg superimposed by the 00<span class="html-italic">l</span> and <math display="inline"><semantics> <mrow> <mover> <mi>h</mi> <mo>¯</mo> </mover> </mrow> </semantics></math><span class="html-italic">hl</span> reflection rows of Po, respectively; (<b>e</b>,<b>f</b>) nonperiodic {111}-parallel contrast features ascribed to the interlayering of Po (001) layers within the Grg-dominant microcrystals.</p> Full article ">Figure 7
<p>Electron micrographs of part of an iron-monosulfide micronodule in the 2037–2042 cmbsf sediment: (<b>a</b>) HAADF-STEM image of an elongated iron-monosulfide aggregate consisting of mackinawite (Mkw) and interlayered Grg-Po (Grg-Po); (<b>b</b>) the elongation of the composite grain parallel to Grg 〈110〉 and mackinawite 〈010〉 as implied by the corresponding diffraction pattern in the Grg 〈110〉 and mackinawite 〈100〉 directions; (<b>c</b>,<b>d</b>) HAADF-STEM images of frames (<b>c</b>,<b>d</b>) in a illustrating zigzag boundaries between mackinawite and interlayered Grg-Po defined by Grg {111} and mackinawite {011} and in part by planar boundaries parallel to Grg and Mkw {001}; (<b>e</b>,<b>f</b>) TEM images of interlayered Grg-Po elongated along 〈110〉<sub>Grg</sub> with crossed contrast features parallel to Grg {111} and Po (001) in the absence of Mkw.</p> Full article ">Figure 8
<p>Electron micrographs of oriented Mkw and interlayered Grg-Po in an iron-monosulfide micronodule in the 2037–2042 cmbsf sediment: (<b>a</b>) TEM image of the growth microtextures of Mkw and interlayered Grg-Po (Grg*) in a composite grain of the micronodule; (<b>b</b>) the corresponding diffraction pattern with 〈100〉<sub>Mkw</sub>//〈110〉<sub>Grg</sub>//〈110〉<sub>Po</sub>, (001)<sub>Mkw</sub>//(001)<sub>Grg</sub>, and (001)<sub>Po</sub>//{111}<sub>Grg</sub>; (<b>c</b>,<b>d</b>) 〈100〉-oriented Mkw (frame c in a) having dislocations in {001} lattice fringes in company with 〈110〉-oriented Grg-dominant domains having crossed {111} fringes (frame (<b>d</b>) in (<b>a</b>)) with varied contrast due to oriented growth of intertwined Po (001) layers; (<b>e</b>,<b>f</b>) Fourier filtered HAADF-STEM images of topotactic interfaces between Mkw and interlayered Grg-Po domains.</p> Full article ">Figure 9
<p>Schematic diagram of crystallographic relationships among Mkw, Grg, and Po viewed from [110]<sub>Grg</sub>, [110]<sub>Po</sub>, and [100]<sub>Mkw</sub> with close-packed planes of sulfur atoms (<span style="color:yellow">•</span>) labeled by A, B, and C to indicate sulfur stacking sequences. Rhombic domains of Grg consist of <span class="html-italic">ccp</span> planes of sulfur atoms and alternating [110]<sub>Grg</sub>-projected columns of edge-sharing octahedral sites fully (<span style="color:#0000CC">●</span>) and half (<span style="color:#0066FF">●</span>) occupied by iron, in association with double columns of tetrahedral iron sites (each half occupied by iron) at open rhombuses. Po slabs are composed of <span class="html-italic">hcp</span> planes of sulfur atoms and (001)<sub>Po</sub> layers of [110]<sub>Po</sub>-projected edge-sharing octahedral sites fully (<span style="color:#FF0066">●</span>) and partially (<span style="color:#FF9933">●</span>) occupied by iron. Mkw (distorted ccp sulfur) is made up of (001)<sub>Mkw</sub> tetrahedral iron sheets connected by van der Waals forces. It is important to note the presence of compulsory structural discontinuities between intersecting Po layers and at intersections between Mkw and Grg and Po due to the mixed occurrence of of different sulfur stacking sequences.</p> Full article ">
Open AccessArticle
Structural Characteristics of E–W-Trending Shear Belts in the Northeastern Dabie Orogen, China: Evidence for Exhumation of High–Ultrahigh-Pressure Rocks
by
Yongsheng Wang, Xu Zhang and Qiao Bai
Minerals 2024, 14(12), 1205; https://doi.org/10.3390/min14121205 - 26 Nov 2024
Abstract
The Dabie–Sulu Orogen hosts the largest area of ultrahigh-pressure (UHP) rocks in the world. There is still significant divergence regarding the exhumation process and mechanism of UHP rocks in the Dabie Orogen, which mainly resulted from the erosion of large volumes of rocks
[...] Read more.
The Dabie–Sulu Orogen hosts the largest area of ultrahigh-pressure (UHP) rocks in the world. There is still significant divergence regarding the exhumation process and mechanism of UHP rocks in the Dabie Orogen, which mainly resulted from the erosion of large volumes of rocks in the Orogen during the post-collisional stage. Based on detailed field investigations, this study discovered the occurrence of E–W-trending sinistral shear belts that developed on the northeastern Dabie Orogen. These shear belts formed under greenschist facies conditions and are characterized by steep foliation and gentle mineral lineation. E–W-trending shear belts developed in HP rocks with metamorphic ages ranging from 227 to 219 Ma and were cut by the older phase of ductile shear belts of the Tan-Lu Fault Zone, indicating that they were formed around 219–197 Ma. Based on a comprehensive analysis of existing data, it can be concluded that E–W-trending shear belts were formed during the exhumation process of HP–UHP rocks. When HP rocks returned to the shallow crust and the lower UHP rocks continued to move, stress concentration occurred in the HP rocks and further resulted in the formation of E–W-trending shear belts. The development of E–W-trending shear belts indicates that HP–UHP rocks had essentially returned to the shallow crust by the Late Triassic, marking the near completion of the exhumation process.
Full article
(This article belongs to the Special Issue Geochemistry and Geochronology of High-Grade Metamorphic Rocks)
►▼
Show Figures
Figure 1
Figure 1
<p>Geological sketch of the northeastern Dabie Orogen. (<b>a</b>) The location of the Central China Orogenic Belt. (<b>b</b>) A simplified tectonic framework of the Dabie Orogen (modified after Xu et al. [<a href="#B25-minerals-14-01205" class="html-bibr">25</a>]). (<b>c</b>) Detailed structural map of the Tongcheng massif. (<b>d</b>) Lower-hemisphere, equal-area stereograms of poles to the mylonitic foliation and plunges of mineral elongation lineation of the E–W-trending and NE–SW-trending shear belts in the northeastern Dabie Orogen. (<b>e</b>) Cross-sections showing tectonic framework of the Dabie Orogen (Section line in (<b>b</b>)). TLF: Tan-Lu Fault Zone; SMF: Shangcheng–Macheng Fault; XMSZ: Xiaotian–Mozitan shear zone; WSF: Wuhe–Shuihou Fault; HMF: Hualiangting–Mituo Fault; TMF: Taihu–Mamiao Fault; XSF: Xishui Fault; XGF: Xiangfan–Guangji Fault.</p> Full article ">Figure 2
<p>Field photos of gneisses and mylonites in the northeastern Dabie Orogen. (<b>a</b>) Development of brittle NE–SW-trending faults in gneisses with flat-lying foliation; (<b>b</b>) steeply chlorite-bearing ultramylonite and (<b>c</b>) granitic mylonite; (<b>d</b>) S-C structures in marble mylonite indicates sinistral shear; (<b>e</b>,<b>f</b>) development of E–W-trending shear zone in gneisses with flat-lying foliation.</p> Full article ">Figure 3
<p>Micrographs of mylonites from the E–W-trending shear belts in the northeastern Dabie Orogen. (<b>a</b>) Broken feldspar and recrystallized quartz; (<b>b</b>) fine-grain recrystallized quartz with a few larger quartz grains; (<b>c</b>) crossed polarizers and (<b>d</b>) single polarizers micrographs of ultramylonite, with banded chlorite, sericite and a few epidotes, with sinistral shear; (<b>e</b>) σ-type feldspar porphyroclasts in the ultramylonite indicating sinistral shear; (<b>f</b>) e-twin of calcite in the marble mylonite. Quartz within rectangular frame for EBSD testing. Qz: quartz; Fsp: feldspar; Pl: plagioclase; Ep: epidote.</p> Full article ">Figure 4
<p>Quartz CPO pattern of mylonites from E–W-trending shear belts in the northeastern Dabie Orogen. Lower-hemisphere, equal-area projection. <span class="html-italic">n</span>: measured grain numbers. <span class="html-italic">X</span> and <span class="html-italic">Z</span> are principal axes of finite strain. Thin sections are parallel with <span class="html-italic">XZ</span> plane.</p> Full article ">Figure 5
<p>Cathodoluminescence images of a selection of the dated zircons from mylonites in the eastern Dabie Orogen. Red circles indicate analytical sites.</p> Full article ">Figure 6
<p>Concordia diagrams and weighted mean ages of zircons from mylonites in the E–W-trending shear belts in the northeastern Dabie Orogen.</p> Full article ">Figure 7
<p>Schematic diagram of tectonic evolution during syn-collisional exhumation and post-collisional uplift. (<b>a</b>) During the Late Triassic to the Early Jurassic, the subducted SCB was exhumed to the shallow crust and resulted in the formation of a fold-and-thrust belt in the northern Yangtze Block and the Hefei foreland basin in the southern NCB. (<b>b</b>) Post-collisional uplift caused rocks overlying the ND unit to be eroded and rebuilt the tectonic framework of the Dabie Orogen. The diamond represents the outcrop location of HP–UHP rocks, and the numbers in parentheses are the ages when these HP–UHP rocks returned to the shallow crust. The dashed lines show the eroded part of units in the Dabie Orogen. Age data are from Liu et al. [<a href="#B10-minerals-14-01205" class="html-bibr">10</a>], Leech et al. [<a href="#B11-minerals-14-01205" class="html-bibr">11</a>], Ayers et al. [<a href="#B31-minerals-14-01205" class="html-bibr">31</a>], and the references therein. Please refer to the manuscript for abbreviated names.</p> Full article ">Figure 8
<p>Schematic model of the formation of E–W-trending shear belts in the northeastern Dabie Orogen during the Late Triassic. To highlight the relationship between the formation of E–W-trending shear belts and the exhumation process of HP–UHP rocks, we neglected other units within the orogen in this figure. Please refer to the manuscript for abbreviated names.</p> Full article ">
<p>Geological sketch of the northeastern Dabie Orogen. (<b>a</b>) The location of the Central China Orogenic Belt. (<b>b</b>) A simplified tectonic framework of the Dabie Orogen (modified after Xu et al. [<a href="#B25-minerals-14-01205" class="html-bibr">25</a>]). (<b>c</b>) Detailed structural map of the Tongcheng massif. (<b>d</b>) Lower-hemisphere, equal-area stereograms of poles to the mylonitic foliation and plunges of mineral elongation lineation of the E–W-trending and NE–SW-trending shear belts in the northeastern Dabie Orogen. (<b>e</b>) Cross-sections showing tectonic framework of the Dabie Orogen (Section line in (<b>b</b>)). TLF: Tan-Lu Fault Zone; SMF: Shangcheng–Macheng Fault; XMSZ: Xiaotian–Mozitan shear zone; WSF: Wuhe–Shuihou Fault; HMF: Hualiangting–Mituo Fault; TMF: Taihu–Mamiao Fault; XSF: Xishui Fault; XGF: Xiangfan–Guangji Fault.</p> Full article ">Figure 2
<p>Field photos of gneisses and mylonites in the northeastern Dabie Orogen. (<b>a</b>) Development of brittle NE–SW-trending faults in gneisses with flat-lying foliation; (<b>b</b>) steeply chlorite-bearing ultramylonite and (<b>c</b>) granitic mylonite; (<b>d</b>) S-C structures in marble mylonite indicates sinistral shear; (<b>e</b>,<b>f</b>) development of E–W-trending shear zone in gneisses with flat-lying foliation.</p> Full article ">Figure 3
<p>Micrographs of mylonites from the E–W-trending shear belts in the northeastern Dabie Orogen. (<b>a</b>) Broken feldspar and recrystallized quartz; (<b>b</b>) fine-grain recrystallized quartz with a few larger quartz grains; (<b>c</b>) crossed polarizers and (<b>d</b>) single polarizers micrographs of ultramylonite, with banded chlorite, sericite and a few epidotes, with sinistral shear; (<b>e</b>) σ-type feldspar porphyroclasts in the ultramylonite indicating sinistral shear; (<b>f</b>) e-twin of calcite in the marble mylonite. Quartz within rectangular frame for EBSD testing. Qz: quartz; Fsp: feldspar; Pl: plagioclase; Ep: epidote.</p> Full article ">Figure 4
<p>Quartz CPO pattern of mylonites from E–W-trending shear belts in the northeastern Dabie Orogen. Lower-hemisphere, equal-area projection. <span class="html-italic">n</span>: measured grain numbers. <span class="html-italic">X</span> and <span class="html-italic">Z</span> are principal axes of finite strain. Thin sections are parallel with <span class="html-italic">XZ</span> plane.</p> Full article ">Figure 5
<p>Cathodoluminescence images of a selection of the dated zircons from mylonites in the eastern Dabie Orogen. Red circles indicate analytical sites.</p> Full article ">Figure 6
<p>Concordia diagrams and weighted mean ages of zircons from mylonites in the E–W-trending shear belts in the northeastern Dabie Orogen.</p> Full article ">Figure 7
<p>Schematic diagram of tectonic evolution during syn-collisional exhumation and post-collisional uplift. (<b>a</b>) During the Late Triassic to the Early Jurassic, the subducted SCB was exhumed to the shallow crust and resulted in the formation of a fold-and-thrust belt in the northern Yangtze Block and the Hefei foreland basin in the southern NCB. (<b>b</b>) Post-collisional uplift caused rocks overlying the ND unit to be eroded and rebuilt the tectonic framework of the Dabie Orogen. The diamond represents the outcrop location of HP–UHP rocks, and the numbers in parentheses are the ages when these HP–UHP rocks returned to the shallow crust. The dashed lines show the eroded part of units in the Dabie Orogen. Age data are from Liu et al. [<a href="#B10-minerals-14-01205" class="html-bibr">10</a>], Leech et al. [<a href="#B11-minerals-14-01205" class="html-bibr">11</a>], Ayers et al. [<a href="#B31-minerals-14-01205" class="html-bibr">31</a>], and the references therein. Please refer to the manuscript for abbreviated names.</p> Full article ">Figure 8
<p>Schematic model of the formation of E–W-trending shear belts in the northeastern Dabie Orogen during the Late Triassic. To highlight the relationship between the formation of E–W-trending shear belts and the exhumation process of HP–UHP rocks, we neglected other units within the orogen in this figure. Please refer to the manuscript for abbreviated names.</p> Full article ">
Open AccessReview
Potential Reuse of Ladle Furnace Slag as Cementitious Material: A Literature Review of Generation, Characterization, and Processing Methods
by
Noureddine Ouffa, Mostafa Benzaazoua, Romain Trauchessec, Tikou Belem, Yassine Taha and Cécile Diliberto
Minerals 2024, 14(12), 1204; https://doi.org/10.3390/min14121204 - 26 Nov 2024
Abstract
Ladle furnace slag (LFS), a by-product of steel refining, shows a promising reuse pathway as an alternative additive or substitute for Portland cement due to its high alkalinity and similar chemical composition to clinkers. However, LFS is often stored in large, open surface
[...] Read more.
Ladle furnace slag (LFS), a by-product of steel refining, shows a promising reuse pathway as an alternative additive or substitute for Portland cement due to its high alkalinity and similar chemical composition to clinkers. However, LFS is often stored in large, open surface areas, leading to many environmental issues. To tackle waste management challenges, LFS can be recycled as supplementary cementitious material (SCM) in many cementitious composites. However, LFS contains some mineral phases that hinder its reactivity (dicalcium silicate (γ-C2S)) and pose long-term durability issues in the cured cemented final product (free lime (f-CaO) and free magnesia (f-MgO)). Therefore, LFS needs to be adequately treated to enhance its reactivity and ensure long-term durability in the structures of the cementitious materials. This literature review assesses possible LFS treatments to enhance its suitability for valorization. Traditional reviews are often multidisciplinary and explore all types of iron and steel slags, sometimes including the recycling of LFS in the steel industry. As the reuse of industrial by-products requires a knowledge of their characteristics, this paper focuses first on LFS characterization, then on the obstacles to its use, and finally compiles an exhaustive inventory of previously investigated treatments. The main parameters for treatment evaluation are the mineralogical composition of treated LFS and the unconfined compressive strength (UCS) of the final geo-composite in the short and long term. This review indicates that the treatment of LFS using rapid air/water quenching at the end-of-refining process is most appropriate, allowing a nearly amorphous slag to be obtained, which is therefore suitable for use as a SCM. Moreover, the open-air watering treatment leads to an optimal content of treated LFS. Recycling LFS in this manner can reduce OPC consumption, solve the problem of limited availability of blast furnace slag (GGBFS) by partially replacing this material, conserve natural resources, and reduce the carbon footprint of cementitious material operations.
Full article
(This article belongs to the Special Issue Metallurgy Waste Used for Backfilling Materials)
►▼
Show Figures
Figure 1
Figure 1
<p>Summary diagram illustrating the iron slag typologies.</p> Full article ">Figure 2
<p>Principle of steel refining in a ladle furnace, modified from Yuasa et al. [<a href="#B53-minerals-14-01204" class="html-bibr">53</a>].</p> Full article ">Figure 3
<p>Secondary electron microscope (SEM) images with different scales of ladle furnace slag (LFS) (in polished section) from ArcelorMittal operation in Contrecoeur, Quebec, Canada.</p> Full article ">Figure 4
<p>Visual appearance of an as-received ladle furnace slag (LFS) from ArcelorMittal’s operations in Contrecoeur, Quebec, Canada (with an index finger as a means of comparison).</p> Full article ">Figure 5
<p>Average chemical composition of ladle furnace slag (LFS) with variation margins of main oxides; data from Table 6.</p> Full article ">Figure 6
<p>Representation of ladle furnace slag (LFS) samples in a ternary diagram.</p> Full article ">Figure 7
<p>X-ray diffractogram of a ground air-cooled ladle furnace slag (LFS) from ArcelorMittal operation at Contrecœur, Quebec, Canada (from the authors’ works).</p> Full article ">Figure 8
<p>Proposed classification of reported ladle furnace slag (LFS) treatments.</p> Full article ">Figure 9
<p>Classification of ladle furnace slag (LFS).</p> Full article ">
<p>Summary diagram illustrating the iron slag typologies.</p> Full article ">Figure 2
<p>Principle of steel refining in a ladle furnace, modified from Yuasa et al. [<a href="#B53-minerals-14-01204" class="html-bibr">53</a>].</p> Full article ">Figure 3
<p>Secondary electron microscope (SEM) images with different scales of ladle furnace slag (LFS) (in polished section) from ArcelorMittal operation in Contrecoeur, Quebec, Canada.</p> Full article ">Figure 4
<p>Visual appearance of an as-received ladle furnace slag (LFS) from ArcelorMittal’s operations in Contrecoeur, Quebec, Canada (with an index finger as a means of comparison).</p> Full article ">Figure 5
<p>Average chemical composition of ladle furnace slag (LFS) with variation margins of main oxides; data from Table 6.</p> Full article ">Figure 6
<p>Representation of ladle furnace slag (LFS) samples in a ternary diagram.</p> Full article ">Figure 7
<p>X-ray diffractogram of a ground air-cooled ladle furnace slag (LFS) from ArcelorMittal operation at Contrecœur, Quebec, Canada (from the authors’ works).</p> Full article ">Figure 8
<p>Proposed classification of reported ladle furnace slag (LFS) treatments.</p> Full article ">Figure 9
<p>Classification of ladle furnace slag (LFS).</p> Full article ">
Open AccessArticle
Influence of Chromite Ore Selection on the Pelletized Oxidative Sintering Process: A South African Case Study
by
Yolindi van Staden, Stephanus Petrus du Preez, Johan Paul Beukes, Pieter Gideon van Zyl and Jason Groenewald
Minerals 2024, 14(12), 1203; https://doi.org/10.3390/min14121203 - 26 Nov 2024
Abstract
The smelting of chromite to produce ferrochrome (FeCr) and subsequently, stainless steel, is an energy-intensive carbothermic process. Various countries apply the Outotec FeCr process, which employs oxidative sintering in air to produce mechanically strong chromite pellets. During this process, iron (Fe) is liberated
[...] Read more.
The smelting of chromite to produce ferrochrome (FeCr) and subsequently, stainless steel, is an energy-intensive carbothermic process. Various countries apply the Outotec FeCr process, which employs oxidative sintering in air to produce mechanically strong chromite pellets. During this process, iron (Fe) is liberated from the chromite spinel due to the elevated temperatures and oxidative nature of the process. It is well understood that oxidatively altered chromite requires less energy to be smelted when compared to non-oxidized chromite. This study showed that sintered pellets obtained from five South African pellet sintering plants had vastly different oxidative alteration penetrations. Additionally, sintered pellets from the same plant may also vary significantly. It was further shown that ores mined from various locations in South Africa had dissimilar sintering behaviors, suggesting that sintered pellets should be characterized before smelting to determine the extent of oxidative alteration. The benefit of a smelter consuming oxidized ore was also demonstrated by comparing the interaction between oxidized and non-oxidized chromite with a carbon (C) source.
Full article
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)
►▼
Show Figures
Figure 1
Figure 1
<p>Backscatter electron SEM micrograph of chromite grains near the edge of a cross-sectioned polished oxidative sintered pellet (<b>a</b>); a corresponding micrograph of grains toward the center of the pellet (<b>b</b>) [<a href="#B36-minerals-14-01203" class="html-bibr">36</a>] (reused under license Creative Commons Attribution (CC BY)).</p> Full article ">Figure 2
<p>Ellingham diagram (Gibbs free energy, ΔG) indicating the reduction of metal oxides with solid C and CO, constructed with HSC thermo-chemical software 10 [<a href="#B38-minerals-14-01203" class="html-bibr">38</a>] (reused with permission from Kleynhans et al., Minerals Engineering, published by Elsevier, 2016).</p> Full article ">Figure 3
<p>Comparison of TGA curves from a commercial/research TGA using a 45 mg composite pellet mixture and the large-mass TGA used in the study, loaded with 20 composite pellets with an approximate weight of 20 g. Both instruments were operated in a N<sub>2</sub> gas environment.</p> Full article ">Figure 4
<p>Backscatter electron SEM micrographs of Sinter Plant 1, Pellet 1, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 1, Pellet 4, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 5
<p>Backscatter electron SEM micrographs of Sinter Plant 2, Pellet 2, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 2, Pellet 4, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 6
<p>Backscatter electron SEM micrographs of Sinter Plant 3, Pellet 2, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 3, Pellet 3, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 7
<p>Backscatter electron SEM micrographs of Sinter Plant 3, Pellet 1, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 3, Pellet 3, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 7 Cont.
<p>Backscatter electron SEM micrographs of Sinter Plant 3, Pellet 1, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 3, Pellet 3, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 8
<p>Backscatter electron SEM micrographs of Sinter Plant 5, case study Pellet 1 near the pellet edge (<b>a</b>) and at the center (<b>b</b>).</p> Full article ">Figure 9
<p>The overall sinter process and the process used to model temperature distribution (D, drying; H, heating: S, sintering; B, balancing; C1–3, cooling) (<b>a</b>) and resulting temperature distribution (<b>b</b>) [<a href="#B49-minerals-14-01203" class="html-bibr">49</a>] (reused under license Creative Commons Attribution (CC BY)).</p> Full article ">Figure 10
<p>Average (solid lines) percentage mass gained with standard deviations (indicated by the lighter semi-transparent areas), measured with TGA, based on five measurement repeats for each case study ore (<b>a</b>). Same results as in <a href="#minerals-14-01203-f010" class="html-fig">Figure 10</a>a, but without the standard deviations indicated and over a slightly large temperature range to make differences between the case study ores more apparent (<b>b</b>).</p> Full article ">Figure 11
<p>TGA curves of base case and sintered pellets.</p> Full article ">
<p>Backscatter electron SEM micrograph of chromite grains near the edge of a cross-sectioned polished oxidative sintered pellet (<b>a</b>); a corresponding micrograph of grains toward the center of the pellet (<b>b</b>) [<a href="#B36-minerals-14-01203" class="html-bibr">36</a>] (reused under license Creative Commons Attribution (CC BY)).</p> Full article ">Figure 2
<p>Ellingham diagram (Gibbs free energy, ΔG) indicating the reduction of metal oxides with solid C and CO, constructed with HSC thermo-chemical software 10 [<a href="#B38-minerals-14-01203" class="html-bibr">38</a>] (reused with permission from Kleynhans et al., Minerals Engineering, published by Elsevier, 2016).</p> Full article ">Figure 3
<p>Comparison of TGA curves from a commercial/research TGA using a 45 mg composite pellet mixture and the large-mass TGA used in the study, loaded with 20 composite pellets with an approximate weight of 20 g. Both instruments were operated in a N<sub>2</sub> gas environment.</p> Full article ">Figure 4
<p>Backscatter electron SEM micrographs of Sinter Plant 1, Pellet 1, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 1, Pellet 4, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 5
<p>Backscatter electron SEM micrographs of Sinter Plant 2, Pellet 2, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 2, Pellet 4, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 6
<p>Backscatter electron SEM micrographs of Sinter Plant 3, Pellet 2, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 3, Pellet 3, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 7
<p>Backscatter electron SEM micrographs of Sinter Plant 3, Pellet 1, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 3, Pellet 3, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 7 Cont.
<p>Backscatter electron SEM micrographs of Sinter Plant 3, Pellet 1, near the edge (<b>a</b>) and at the pellet center (<b>b</b>); Sinter Plant 3, Pellet 3, near the edge (<b>c</b>) and at the pellet center (<b>d</b>).</p> Full article ">Figure 8
<p>Backscatter electron SEM micrographs of Sinter Plant 5, case study Pellet 1 near the pellet edge (<b>a</b>) and at the center (<b>b</b>).</p> Full article ">Figure 9
<p>The overall sinter process and the process used to model temperature distribution (D, drying; H, heating: S, sintering; B, balancing; C1–3, cooling) (<b>a</b>) and resulting temperature distribution (<b>b</b>) [<a href="#B49-minerals-14-01203" class="html-bibr">49</a>] (reused under license Creative Commons Attribution (CC BY)).</p> Full article ">Figure 10
<p>Average (solid lines) percentage mass gained with standard deviations (indicated by the lighter semi-transparent areas), measured with TGA, based on five measurement repeats for each case study ore (<b>a</b>). Same results as in <a href="#minerals-14-01203-f010" class="html-fig">Figure 10</a>a, but without the standard deviations indicated and over a slightly large temperature range to make differences between the case study ores more apparent (<b>b</b>).</p> Full article ">Figure 11
<p>TGA curves of base case and sintered pellets.</p> Full article ">
Open AccessArticle
Geochemical Anomaly Detection and Pattern Recognition: A Combined Study of the Apriori Algorithm, Principal Component Analysis, and Spectral Clustering
by
Mahsa Hajihosseinlou, Abbas Maghsoudi and Reza Ghezelbash
Minerals 2024, 14(12), 1202; https://doi.org/10.3390/min14121202 - 26 Nov 2024
Abstract
This study demonstrates the effectiveness of combining Principal Component Analysis (PCA) and the Apriori algorithm for feature selection, alongside Spectral clustering, to detect geochemical anomalies in Mississippi Valley-Type (MVT) Pb-Zn deposits in western Iran. First, PCA and Apriori enabled the identification of both
[...] Read more.
This study demonstrates the effectiveness of combining Principal Component Analysis (PCA) and the Apriori algorithm for feature selection, alongside Spectral clustering, to detect geochemical anomalies in Mississippi Valley-Type (MVT) Pb-Zn deposits in western Iran. First, PCA and Apriori enabled the identification of both syngenetic and epigenetic components, which helped in recognizing elements associated with mineralization. These elements were then modeled using Spectral clustering to detect geochemical anomalies. Unlike traditional methods like k-means, Spectral clustering does not require spherical clusters and is adept at identifying clusters of arbitrary shapes. This made it particularly suitable for analyzing the irregular shapes of geochemical anomalies in the study area. By incorporating Spectral clustering, the method effectively separated geochemical groups, revealing the underlying structure of the data. This was crucial for identifying anomalous geochemical zones and delineating areas with a high potential for Pb-Zn mineralization. The performance of the Spectral clustering algorithm was thoroughly evaluated using the Silhouette Score, the Davies–Bouldin Index, and Dunn Index. Subsampling was employed to assess the algorithm’s stability, providing a comprehensive evaluation of its effectiveness in identifying geochemical anomalies and mapping mineralization potential.
Full article
(This article belongs to the Section Mineral Exploration Methods and Applications)
►▼
Show Figures
Figure 1
Figure 1
<p>The map of Iran shows the location of the study area.</p> Full article ">Figure 2
<p>Simplified geological map.</p> Full article ">Figure 3
<p>Stream sediments samples.</p> Full article ">Figure 4
<p>A comprehensive flowchart of the methodology.</p> Full article ">Figure 5
<p>Apriori algorithm flowchart.</p> Full article ">Figure 6
<p>Geochemical layers of (<b>a</b>) Pb, (<b>b</b>) Zn, and (<b>c</b>) Ba.</p> Full article ">Figure 7
<p>Procedure of Spectral algorithm.</p> Full article ">Figure 8
<p>Elbow method analysis for identifying the optimal number of clusters.</p> Full article ">Figure 9
<p>(<b>a</b>) Two-dimensional and (<b>b</b>) three-dimensional cluster visualization.</p> Full article ">Figure 10
<p>Geochemical anomaly map derived by Spectral clustering.</p> Full article ">
<p>The map of Iran shows the location of the study area.</p> Full article ">Figure 2
<p>Simplified geological map.</p> Full article ">Figure 3
<p>Stream sediments samples.</p> Full article ">Figure 4
<p>A comprehensive flowchart of the methodology.</p> Full article ">Figure 5
<p>Apriori algorithm flowchart.</p> Full article ">Figure 6
<p>Geochemical layers of (<b>a</b>) Pb, (<b>b</b>) Zn, and (<b>c</b>) Ba.</p> Full article ">Figure 7
<p>Procedure of Spectral algorithm.</p> Full article ">Figure 8
<p>Elbow method analysis for identifying the optimal number of clusters.</p> Full article ">Figure 9
<p>(<b>a</b>) Two-dimensional and (<b>b</b>) three-dimensional cluster visualization.</p> Full article ">Figure 10
<p>Geochemical anomaly map derived by Spectral clustering.</p> Full article ">
Open AccessArticle
Characterization and Processing of Low-Grade Middle Group 2 Chromite Ore by Gravity Shaking Table and a Comparative SLon Magnetic Separation: A Case Study
by
Inga Sixhuta, Ashma Singh, Phathutshedzo Khangale, Reinout Meijboom and Mpfunzeni Raphulu
Minerals 2024, 14(12), 1201; https://doi.org/10.3390/min14121201 - 26 Nov 2024
Abstract
Chromite is considered a strategic mineral in the global economy. It is mainly used as an essential raw material in the production of stainless steel and other metal alloys due to its corrosion and heat resistance properties. High-grade chromite resources are gradually depleting;
[...] Read more.
Chromite is considered a strategic mineral in the global economy. It is mainly used as an essential raw material in the production of stainless steel and other metal alloys due to its corrosion and heat resistance properties. High-grade chromite resources are gradually depleting; with the increasing chromite demand in metallurgical applications, studies have focused on exploring low-grade and alternative chromite sources. This study proposes a cost-effective processing flowsheet for the low-grade middle group 2 (MG2) chromite layer, a poorly explored chromatite seam within the South African bushveld igneous complex (BIC). The study involved mineralogical characterization followed by gravity and magnetic separation of the low-grade MG2 ore at 18.18% Cr2O3. Characterization by XRD and Auto-SEM revealed that the ore mainly consists of pyroxene, chromite, and feldspar, with other minerals in trace quantities. The gravity separation test by shaking table upgraded the chromite (Cr2O3) to 42.0% at high chromite recoveries, whereas the laboratory Slon wet high-intensity magnetic separation method (SLon WHIMS) upgraded the chromite in the feed to 42.95% grade at lower chromite recoveries. Desliming the sample before the gravity and magnetic separation tests significantly improved the separation. The magnetic separation tests further demonstrated that chromite within the MG2 layer is sensitive to magnetic separation due to its high iron content. The adapted flowsheet is proposed as a cost-effective flowsheet for processing the low-grade MG2 layer. The flow sheet can be optimized by conducting the SLon WHIMS tests at high intensities followed by fine gravity tests by spiral circuits to maximize the chromite recovery while achieving commercial chromite grades and a Cr:Fe ratio greater than 1.5.
Full article
(This article belongs to the Special Issue Mineral Processing Technologies of Low-Grade Ores)
►▼
Show Figures
Figure 1
Figure 1
<p>Stratigraphy of Rustenburg Layered Suite of the Bushveld Complex [<a href="#B9-minerals-14-01201" class="html-bibr">9</a>].</p> Full article ">Figure 2
<p>Global chromite production over the years, 2010–2023 (USGS Reports).</p> Full article ">Figure 3
<p>Proposed flowsheet for processing a low-grade MG2 run-of-mine ore.</p> Full article ">Figure 4
<p>Schematic diagram of a Laboratory SLon-100, a wet high-intensity magnetic separator unit (Metso Outotec, Helsinki, Finland).</p> Full article ">Figure 5
<p>Particle size distribution of MG2 ore crushed to −1mm. Screen sizes used between 850 and 25 µm.</p> Full article ">Figure 6
<p>Discrete mass, chromite grade, and deportment across different sizes.</p> Full article ">Figure 7
<p>Normative bulk quantitative mineralogy results from Auto-SEM analysis.</p> Full article ">Figure 8
<p>Cumulative liberation curve for each size fraction. The red vertical line delineates 80% liberation. The +600 µm size is least liberated.</p> Full article ">Figure 9
<p>Backscatter electron (BSE) image of moderate- to fine-grained chromite grains within a larger particle from the MG2 sample. (<b>a</b>) Greyscale BSE image of the particle. (<b>b</b>) False color image of a particle with mineral classifications by the Auto-SEM.</p> Full article ">Figure 10
<p>Wide-angle XRD pattern for the MG2 chromite sample.</p> Full article ">Figure 11
<p>Cumulative chromite grades and recoveries for shaking table gravity tests on MG2 sample.</p> Full article ">Figure 12
<p>Cumulative Fe grades and recoveries from the shaking table test.</p> Full article ">Figure 13
<p>Cumulative SiO<sub>2</sub> grade and recovery curves from the shaking table test.</p> Full article ">Figure 14
<p>Summary of SLon whims results for chromite in the magnetic products at various field intensities. Primary y-axis presents both mass yield and chromite recovery, whereas the secondary y-axis shows chromite grades.</p> Full article ">Figure 15
<p>Discrete mass yield, Fe grade and recovery curves in the magnetic stream for magnetic separation tests by Slon WHIMS at various intensities.</p> Full article ">Figure 16
<p>Discrete mass yield, SiO<sub>2</sub> grade and recovery curves in the magnetic stream of Slon WHIMS tests at various intensities.</p> Full article ">
<p>Stratigraphy of Rustenburg Layered Suite of the Bushveld Complex [<a href="#B9-minerals-14-01201" class="html-bibr">9</a>].</p> Full article ">Figure 2
<p>Global chromite production over the years, 2010–2023 (USGS Reports).</p> Full article ">Figure 3
<p>Proposed flowsheet for processing a low-grade MG2 run-of-mine ore.</p> Full article ">Figure 4
<p>Schematic diagram of a Laboratory SLon-100, a wet high-intensity magnetic separator unit (Metso Outotec, Helsinki, Finland).</p> Full article ">Figure 5
<p>Particle size distribution of MG2 ore crushed to −1mm. Screen sizes used between 850 and 25 µm.</p> Full article ">Figure 6
<p>Discrete mass, chromite grade, and deportment across different sizes.</p> Full article ">Figure 7
<p>Normative bulk quantitative mineralogy results from Auto-SEM analysis.</p> Full article ">Figure 8
<p>Cumulative liberation curve for each size fraction. The red vertical line delineates 80% liberation. The +600 µm size is least liberated.</p> Full article ">Figure 9
<p>Backscatter electron (BSE) image of moderate- to fine-grained chromite grains within a larger particle from the MG2 sample. (<b>a</b>) Greyscale BSE image of the particle. (<b>b</b>) False color image of a particle with mineral classifications by the Auto-SEM.</p> Full article ">Figure 10
<p>Wide-angle XRD pattern for the MG2 chromite sample.</p> Full article ">Figure 11
<p>Cumulative chromite grades and recoveries for shaking table gravity tests on MG2 sample.</p> Full article ">Figure 12
<p>Cumulative Fe grades and recoveries from the shaking table test.</p> Full article ">Figure 13
<p>Cumulative SiO<sub>2</sub> grade and recovery curves from the shaking table test.</p> Full article ">Figure 14
<p>Summary of SLon whims results for chromite in the magnetic products at various field intensities. Primary y-axis presents both mass yield and chromite recovery, whereas the secondary y-axis shows chromite grades.</p> Full article ">Figure 15
<p>Discrete mass yield, Fe grade and recovery curves in the magnetic stream for magnetic separation tests by Slon WHIMS at various intensities.</p> Full article ">Figure 16
<p>Discrete mass yield, SiO<sub>2</sub> grade and recovery curves in the magnetic stream of Slon WHIMS tests at various intensities.</p> Full article ">
Open AccessArticle
Application of Machine Learning for Generic Mill Liner Wear Prediction in Semi-Autogenous Grinding (SAG) Mills
by
Yusuf Enes Pural, Tania Ledezma, Marko Hilden, Gordon Forbes, Feridun Boylu and Mohsen Yahyaei
Minerals 2024, 14(12), 1200; https://doi.org/10.3390/min14121200 - 25 Nov 2024
Abstract
This study explores the application of machine learning techniques for predicting generic mill liner wear in semi-autogenous grinding (SAG) mills used in mineral processing. Various models were developed and compared using data from 143 liner measurements across 36 liner cycles from ten different
[...] Read more.
This study explores the application of machine learning techniques for predicting generic mill liner wear in semi-autogenous grinding (SAG) mills used in mineral processing. Various models were developed and compared using data from 143 liner measurements across 36 liner cycles from ten different SAG mills. The research initially focused on individual mill modeling, employing simple linear regression, first-order kinetic approach, Multiple Linear Regression (MLR), tree-based methods (Decision Trees, Random Forests, XGBoost), and Multilayer Perceptron (MLP). Results showed that simple linear regression provided sufficient accuracy, with other methods only slightly improving performance. This study then developed a combined model using data from multiple mills. MLR and advanced machine learning techniques were applied for this generic model, with XGBoost emerging as the most successful. In the interpolation scenario involving a mill similar to those in the training data, the XGBoost model achieved a mean absolute percentage error (MAPE) of 5.27%. For the extrapolation scenario, with a mill larger than those in the training set, the MAPE increased slightly to 6.12%. These results demonstrate the potential of machine learning approaches in creating effective generic models for mill liner wear prediction. However, this study also highlights the potential for improving predictive models by incorporating additional key parameters such as liner and ball material properties.
Full article
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)
►▼
Show Figures
Figure 1
Figure 1
<p>3D scan result example.</p> Full article ">Figure 2
<p>PCHIP results for a high-low lifter SAG mill.</p> Full article ">Figure 3
<p>General overview of the workflow.</p> Full article ">Figure 4
<p>Train and test split for individual model building.</p> Full article ">Figure 5
<p>Train, validation, and test split for combined model building and testing.</p> Full article ">Figure 6
<p>Principle component analysis and explained variance ratio.</p> Full article ">Figure 7
<p>Multiple Linear Regression results (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">Figure 8
<p>XGBoost results on test data: (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">Figure 8 Cont.
<p>XGBoost results on test data: (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">Figure 9
<p>Shapley summary plots for XGBoost model (<b>a</b>) Lifter height (<b>b</b>) Plate thickness.</p> Full article ">Figure 10
<p>MLP test prediction results for test data (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">
<p>3D scan result example.</p> Full article ">Figure 2
<p>PCHIP results for a high-low lifter SAG mill.</p> Full article ">Figure 3
<p>General overview of the workflow.</p> Full article ">Figure 4
<p>Train and test split for individual model building.</p> Full article ">Figure 5
<p>Train, validation, and test split for combined model building and testing.</p> Full article ">Figure 6
<p>Principle component analysis and explained variance ratio.</p> Full article ">Figure 7
<p>Multiple Linear Regression results (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">Figure 8
<p>XGBoost results on test data: (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">Figure 8 Cont.
<p>XGBoost results on test data: (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">Figure 9
<p>Shapley summary plots for XGBoost model (<b>a</b>) Lifter height (<b>b</b>) Plate thickness.</p> Full article ">Figure 10
<p>MLP test prediction results for test data (<b>a</b>) SAG 7 Lifter (<b>b</b>) SAG 7 Plate (<b>c</b>) SAG 8 Lifter (<b>d</b>) SAG 8 Plate.</p> Full article ">
Open AccessArticle
Discovery and Geological Significance of Neoproterozoic Bimodal Intrusive Rocks in the Dabie Orogen, China
by
Linjing Li, Mingyi Hu, Lingyao Kong, Lin Wang and Qiqi Lyu
Minerals 2024, 14(12), 1199; https://doi.org/10.3390/min14121199 - 25 Nov 2024
Abstract
The Mingshan reservoir of the Dabie Orogen has a number of Neoproterozoic bimodal intrusive rocks. We focused on the zircon U-Pb chronology, Hf isotopes, and bulk-rock geochemistry of these rocks. The results showed the following: (1) The bimodal intrusive rocks mainly consist of
[...] Read more.
The Mingshan reservoir of the Dabie Orogen has a number of Neoproterozoic bimodal intrusive rocks. We focused on the zircon U-Pb chronology, Hf isotopes, and bulk-rock geochemistry of these rocks. The results showed the following: (1) The bimodal intrusive rocks mainly consist of monzogranitic gneiss and plagioamphibolite, with zircon U-Pb ages of 785.0 ± 7.1 Ma and 787.3 ± 6.1 Ma, respectively. These ages indicate that they were formed in the late Qingbaikou epoch of the Neoproterozoic era. (2) The monzogranitic gneiss was dominated by peraluminous features and displayed a strong right deviation of REE (rare-earth element) patterns and a negative δEu anomaly. It is enriched in the LILEs (large-ion lithophile elements) Rb, Ba, and K, but slightly depleted in Nb, Sr, P, and Ti, with low 10,000* Ga/Al values, indicating that it is similar to Al-type granite. The plagioamphibolite belongs to the metaluminous, peraluminous series. It has a Mg# (molar ratio of Mg to Mg + Fe) of 36.1~55.9 and is enriched in the LILEs Rb, Ba, and K, with a slight positive anomaly of Ba, and is depleted in Nb and Sr. (3) The monzogranite shows negative zircon εHf(t) values ranging from −13.4 to −7.2 and a Paleoproterozoic TDM2(Hf) (two-stage depleted mantle model age) of 1969–2298 Ma. The zircon εHf(t) values and TDM2(Hf) of the plagioamphibolite were concentrated around 2.9–5.7 and 1257–1410 Ma, respectively. The geochemistry and Hf isotopes show that the monzogranitic gneiss and plagioamphibolite have distinct magmatic sources. The plagioamphibolite formed from mantle and partial continental crustal materials. The monzogranitic gneiss, on the other hand, was formed as a result of the partial melting of the shallow ancient felsic crust caused by mafic rock heating or upwelling. Taking into account regional correlation, the middle Neoproterozoic bimodal intrusive rocks originated in the structural framework of an extensional setting.
Full article
(This article belongs to the Section Mineral Geochemistry and Geochronology)
►▼
Show Figures
Figure 1
Figure 1
<p>(<b>a</b>) Precambrian geological overview map of Qinling–Dabie Orogen; (<b>b</b>) geological map of the study area.</p> Full article ">Figure 2
<p>Field (<b>a</b>–<b>c</b>,<b>e</b>) and micrographs (<b>d</b>,<b>f</b>) of bimodal intrusive rocks in Mingshan reservoir in the north of Dabie Orogen. (<b>a</b>) Interbedded monzogranitic gneiss (light color) and plagioamphibolite (dark color); (<b>b</b>) contact relationship between monzogranitic gneiss (light color) and plagioamphibolite (dark color); (<b>c</b>) field photographs of monzogranitic gneiss; (<b>d</b>) orthonormal polariscope photographs of monzogranitic gneiss; (<b>e</b>) field photographs of plagioamphibolite; (<b>f</b>) orthonormal polariscope photographs of plagioamphibolite; Hbl—hornblende; Qtz—quartz; Pl—plagioclase; Kf—potassium feldspar; Bi—biotite.</p> Full article ">Figure 3
<p>Typical CL images, <sup>206</sup>Pb/<sup>238</sup>U ages, and ε<sub>Hf</sub>(t) values of zircon from samples ZJY-1-14 and ZJY-2-14 (white solid circle is for U-Pb dating, yellow dashed circle is for Lu-Hf isotope analysis).</p> Full article ">Figure 4
<p>Zircon U-Pb concordia diagrams and mean age diagrams of Neoproterozoic bimodal intrusive rocks in Dabie Orogen.</p> Full article ">Figure 5
<p>(<b>a</b>) A/NK-A/CNK geochemical discrimination diagram (after Maniar and Piccoli, 1989 [<a href="#B44-minerals-14-01199" class="html-bibr">44</a>]); (<b>b</b>) FeO<sup>t</sup>/(FeO<sup>t</sup> + MgO)-SiO<sub>2</sub> diagram (after Frost et al., 2001 [<a href="#B45-minerals-14-01199" class="html-bibr">45</a>]).</p> Full article ">Figure 6
<p>(<b>a</b>) Nb/Y-SiO<sub>2</sub> diagram (after Winchester and Floyd, 1997 [<a href="#B46-minerals-14-01199" class="html-bibr">46</a>]) and (<b>b</b>) SiO<sub>2</sub>-FeO<sup>t</sup>/MgO diagram (after Miyashiro, 1975 [<a href="#B47-minerals-14-01199" class="html-bibr">47</a>]) for the bimodal intrusive rocks in Dabie Orogen.</p> Full article ">Figure 7
<p>(<b>a</b>) Chondrite-normalized REE patterns and (<b>b</b>) N-MORB-normalized trace element patterns for the monzogranitic gneiss and plagioamphibolite in the Dabie Orogen. ((<b>a</b>) Normalizing values are from Boynton [<a href="#B48-minerals-14-01199" class="html-bibr">48</a>]; (<b>b</b>) normalizing values are from Sun and McDonough [<a href="#B49-minerals-14-01199" class="html-bibr">49</a>]).</p> Full article ">Figure 8
<p>Zircon age—<span class="html-italic">ε</span><sub>Hf</sub>(<span class="html-italic">t</span>) (<b>a</b>), Nb-Nb/U (<b>b</b>), and Ce-Ce/Pb (<b>c</b>) diagrams for plagioamphibolite ((<b>b</b>,<b>c</b>) after Hofmann et al., 1986 [<a href="#B50-minerals-14-01199" class="html-bibr">50</a>]).</p> Full article ">Figure 9
<p>(<b>a</b>) Y and (<b>b</b>) Ce vs. 10,000* Ga/Al; (c) Nb-Y-3Ga discrimination diagram of monzogranitic gneiss in the Dabie Orogen ((<b>a</b>,<b>b</b>) after Whalen et al., 1987 [<a href="#B53-minerals-14-01199" class="html-bibr">53</a>]; (<b>c</b>) after Eby G N., 1992 [<a href="#B54-minerals-14-01199" class="html-bibr">54</a>]). Abbreviations: A: A-type granites; I: I-type granites; S: S-type granites; M: M-type granites. A1 = granitoids in anorogenic settings; A2 = post-orogenic granites emplaced after a continental collision.</p> Full article ">Figure 10
<p>Zr-Zr/Y, Zr-Ti, and Ta/Hf-Th/Hf diagrams for the bimodal intrusive rocks in Dabie Orogen of Neoproterozoic era ((<b>a</b>,<b>b</b>) after Pearce et al., 1973 [<a href="#B56-minerals-14-01199" class="html-bibr">56</a>]; (<b>c</b>) after Wang Y L, 2001 [<a href="#B57-minerals-14-01199" class="html-bibr">57</a>]). I: N-MORB in diverging plate boundary; II: converging plate boundary; III: intra-oceanic plate; IV: intra-continental plate; V: basalt region in mantle plume; WPB: within plate basalt, VAB: volcanic arc basalt, MORB: mid-ocean ridge basalt.</p> Full article ">
<p>(<b>a</b>) Precambrian geological overview map of Qinling–Dabie Orogen; (<b>b</b>) geological map of the study area.</p> Full article ">Figure 2
<p>Field (<b>a</b>–<b>c</b>,<b>e</b>) and micrographs (<b>d</b>,<b>f</b>) of bimodal intrusive rocks in Mingshan reservoir in the north of Dabie Orogen. (<b>a</b>) Interbedded monzogranitic gneiss (light color) and plagioamphibolite (dark color); (<b>b</b>) contact relationship between monzogranitic gneiss (light color) and plagioamphibolite (dark color); (<b>c</b>) field photographs of monzogranitic gneiss; (<b>d</b>) orthonormal polariscope photographs of monzogranitic gneiss; (<b>e</b>) field photographs of plagioamphibolite; (<b>f</b>) orthonormal polariscope photographs of plagioamphibolite; Hbl—hornblende; Qtz—quartz; Pl—plagioclase; Kf—potassium feldspar; Bi—biotite.</p> Full article ">Figure 3
<p>Typical CL images, <sup>206</sup>Pb/<sup>238</sup>U ages, and ε<sub>Hf</sub>(t) values of zircon from samples ZJY-1-14 and ZJY-2-14 (white solid circle is for U-Pb dating, yellow dashed circle is for Lu-Hf isotope analysis).</p> Full article ">Figure 4
<p>Zircon U-Pb concordia diagrams and mean age diagrams of Neoproterozoic bimodal intrusive rocks in Dabie Orogen.</p> Full article ">Figure 5
<p>(<b>a</b>) A/NK-A/CNK geochemical discrimination diagram (after Maniar and Piccoli, 1989 [<a href="#B44-minerals-14-01199" class="html-bibr">44</a>]); (<b>b</b>) FeO<sup>t</sup>/(FeO<sup>t</sup> + MgO)-SiO<sub>2</sub> diagram (after Frost et al., 2001 [<a href="#B45-minerals-14-01199" class="html-bibr">45</a>]).</p> Full article ">Figure 6
<p>(<b>a</b>) Nb/Y-SiO<sub>2</sub> diagram (after Winchester and Floyd, 1997 [<a href="#B46-minerals-14-01199" class="html-bibr">46</a>]) and (<b>b</b>) SiO<sub>2</sub>-FeO<sup>t</sup>/MgO diagram (after Miyashiro, 1975 [<a href="#B47-minerals-14-01199" class="html-bibr">47</a>]) for the bimodal intrusive rocks in Dabie Orogen.</p> Full article ">Figure 7
<p>(<b>a</b>) Chondrite-normalized REE patterns and (<b>b</b>) N-MORB-normalized trace element patterns for the monzogranitic gneiss and plagioamphibolite in the Dabie Orogen. ((<b>a</b>) Normalizing values are from Boynton [<a href="#B48-minerals-14-01199" class="html-bibr">48</a>]; (<b>b</b>) normalizing values are from Sun and McDonough [<a href="#B49-minerals-14-01199" class="html-bibr">49</a>]).</p> Full article ">Figure 8
<p>Zircon age—<span class="html-italic">ε</span><sub>Hf</sub>(<span class="html-italic">t</span>) (<b>a</b>), Nb-Nb/U (<b>b</b>), and Ce-Ce/Pb (<b>c</b>) diagrams for plagioamphibolite ((<b>b</b>,<b>c</b>) after Hofmann et al., 1986 [<a href="#B50-minerals-14-01199" class="html-bibr">50</a>]).</p> Full article ">Figure 9
<p>(<b>a</b>) Y and (<b>b</b>) Ce vs. 10,000* Ga/Al; (c) Nb-Y-3Ga discrimination diagram of monzogranitic gneiss in the Dabie Orogen ((<b>a</b>,<b>b</b>) after Whalen et al., 1987 [<a href="#B53-minerals-14-01199" class="html-bibr">53</a>]; (<b>c</b>) after Eby G N., 1992 [<a href="#B54-minerals-14-01199" class="html-bibr">54</a>]). Abbreviations: A: A-type granites; I: I-type granites; S: S-type granites; M: M-type granites. A1 = granitoids in anorogenic settings; A2 = post-orogenic granites emplaced after a continental collision.</p> Full article ">Figure 10
<p>Zr-Zr/Y, Zr-Ti, and Ta/Hf-Th/Hf diagrams for the bimodal intrusive rocks in Dabie Orogen of Neoproterozoic era ((<b>a</b>,<b>b</b>) after Pearce et al., 1973 [<a href="#B56-minerals-14-01199" class="html-bibr">56</a>]; (<b>c</b>) after Wang Y L, 2001 [<a href="#B57-minerals-14-01199" class="html-bibr">57</a>]). I: N-MORB in diverging plate boundary; II: converging plate boundary; III: intra-oceanic plate; IV: intra-continental plate; V: basalt region in mantle plume; WPB: within plate basalt, VAB: volcanic arc basalt, MORB: mid-ocean ridge basalt.</p> Full article ">
Open AccessArticle
Increasing Structural Diversity of the Early Growth Stages in Polynesian Pearls Reveals Biological Stress Suffered by the Grafts
by
Jean-Pierre Cuif, Yannicke Dauphin, Marc Gèze, Cedrik Lo, Gergely Nemeth and Christophe Sandt
Minerals 2024, 14(12), 1198; https://doi.org/10.3390/min14121198 - 25 Nov 2024
Abstract
In Polynesian pearls produced using Pinctada margaritifera var. Cumingii, we investigated the structure of the early growth stages, from the nucleus surface up to the first deposition of the black nacre characteristic of this subspecies. Despite simultaneous grafting from the same donor
[...] Read more.
In Polynesian pearls produced using Pinctada margaritifera var. Cumingii, we investigated the structure of the early growth stages, from the nucleus surface up to the first deposition of the black nacre characteristic of this subspecies. Despite simultaneous grafting from the same donor oyster and similar cultivation conditions, we observed the deposition of various non-nacreous pre-nacre structures. These unusual microstructures, which precede the return to black nacre, varied from immediate deposition onto the nucleus surface to increasing delays, depending on the graft’s position in the grafting series. Given the similar biological conditions of grafting and cultivation, we suggest that, in line with recent data demonstrating genomic sensitivity to environmental conditions, alterations in the graft cells produced during the increasing waiting period were transmitted to the pearl sacs and the early growth stages of the grafted pearls.
Full article
(This article belongs to the Section Biomineralization and Biominerals)
►▼
Show Figures
Figure 1
Figure 1
<p>Diameter section of a Polynesian black pearl after a three-month cultivation period. (<b>a</b>–<b>d</b>) Optical view; note the perfectly black section throughout the thickness of the pearl layer. (<b>e</b>–<b>f</b>) Nacreous layer. (<b>g</b>,<b>h</b>) Details of the interface between the nucleus and the first deposits by the pearl sac: (<b>g</b>) secondary electron image; note the basal layer of the series, still rich in organic components; (<b>h</b>) BSE image of the same area.</p> Full article ">Figure 2
<p>Cross-sections of a pearl after a three-month cultivation period. (<b>a</b>–<b>c</b>) Optical images show the non-pigmented phase, characterized by a lack of a prismatic or nacreous structure. (<b>d</b>–<b>f</b>) SEM micrographs reveal the diverse layered structures formed within this region, preceding the resumption of nacre deposition. Note the variety of layered structures produced within this region.</p> Full article ">Figure 3
<p>Microstructures of the pre-nacreous layer in a pearl. (<b>a</b>) Thin pre-nacreous regions with limited prismatic areas. (<b>b</b>) Close-up of prismatic structures. (<b>c</b>,<b>d</b>) Fibrous structures perpendicular to the nucleus surface, with variations in their basal region. (<b>e</b>) Complex mineralization featuring short prism-like structures embedded in a fibrous coverage. (<b>f</b>,<b>g</b>) Details of initial secretions within the basal part of the prism-like structures, and organic envelopes surrounding the prism-like structures. (<b>h</b>) The transition from prism-like structures to black nacre is not straightforward. After the disappearance of the envelopes, prism-like material is overlaid by a granular structure. org: non-mineral nucleus coverage; yellow areas: distinct hemispherical areas; pr: prism-like structures; int: granular intermediate structures.</p> Full article ">Figure 4
<p>Pearl with an extended prism-like initial coverage. (<b>a</b>) Overall distribution of microstructures during the early mineralization period. (<b>b</b>,<b>c</b>) After the formation of a thick organic layer (org. cov), prism-like structures develop: an expansion phase (white arrow) followed by a regression phase (blue arrows). (<b>d</b>) BSE SEM view showing prismatic structures with irregular membranes and a diverse mineral sequence between them. (<b>e</b>,<b>f</b>) Oscillating mineralization: note the regular layered growth of the prism-like structures and the rapidly changing microstructure in (<b>e</b>). Clearly, instability is the main characteristic of the pre-nacreous phase.</p> Full article ">Figure 5
<p>Non-nacreous basal pearl layer after a three-month post-grafting period. (<b>a</b>) Global section showing the distribution of prismatic (blue arrows) and fibrous mineralizations. (<b>b</b>) Basal part: thick initial deposition without mineral structure. Upper part: layered deposition and irregular prism-like structures. (<b>c</b>,<b>d</b>) Optical view of a punctual deposition ((<b>a</b>): blue arrows). (<b>e</b>,<b>h</b>) Aspect of the “prism-like” structure: prisms are discontinuous with abnormal components. (<b>e</b>,<b>g</b>,<b>h</b>) SE SEM images. (<b>f</b>) BSE SEM image. (<b>i</b>) Earliest nacre deposition at the edge of the nacreous layer. (<b>j</b>,<b>k</b>) Optical view showing spots of nacre.</p> Full article ">Figure 6
<p>A three-month cultivation specimen without nacreous secretion: the entire pearl layer is made of “pre-nacreous” material. (<b>a</b>,<b>b</b>) Homogeneous structure of the pearl layer. (<b>c</b>,<b>d</b>) The mineralization mechanism shows only slight variations throughout the pearl layer. Note the consistent thickness of the layering. (<b>e</b>–<b>g</b>) Closer examination (SEM) reveals that this structure was produced by a severely disturbed graft structure. From its basal part ((<b>e</b>): blue arrow) to its outermost layers, the growth layers maintain similar structures composed of global layering (orange arrows) built by associating with short-distance radial elements (green arrows).</p> Full article ">Figure 7
<p>(<b>a</b>–<b>c</b>) Polynesian pearls with an unusually extended pre-nacreous period. (<b>a</b>) Only a thin nacreous layer was formed at the lower part (black arrows). In the pre-nacreous mineralization, red arrows indicate calcite, and green arrows indicate aragonite. (<b>d</b>) An enlarged view of pearl c showing the aragonite and calcite regions from the nucleus to the first layers of black nacre. (<b>e</b>,<b>f</b>) A thick basal part in contact with the nucleus, composed of repeated non-mineralized layers (yellow arrows) and covered by simultaneously produced early stages of the calcite area (red arrows) and aragonite area (green arrows). Note the overall expansion of calcite against aragonite (blue arrows in (<b>d</b>–<b>f</b>)). (<b>g</b>,<b>h</b>) Mixed calcite–aragonite layers with varying thicknesses depending on mineralogy. (<b>i</b>) Progressive regression of calcite and expansion of aragonite layers, leading to the formation of Polynesian black nacre.</p> Full article ">Figure 8
<p>Maps of the closely separating membrane between calcite and aragonite in the <span class="html-italic">Pinctada</span> shell contrasting with the absence of structural separation between calcite and aragonite in pearls. (<b>a</b>) Separating membrane between the calcite and aragonite compartments in a <span class="html-italic">Pinctada</span> shell. SEM image. (<b>b</b>) Raman map illustrating the organic membrane separating aragonite and calcite at the termination of the prismatic region; 2800–3100 cm<sup>−1</sup> indicative of lipids; red color for high organic content , blue dark for low organic content. (<b>c</b>) Optical view of the contact between calcite and aragonite areas in the pre-nacreous domain of a Polynesian <span class="html-italic">Pinctada</span> pearl. (<b>d</b>,<b>e</b>) Raman maps of the calcite–aragonite interface in this region reveal no distinct chemical boundary separating the two mineral components; d: 280 cm<sup>−1</sup> for calcite, e: 1085 cm<sup>−1</sup> for aragonite. Green color for high contents. (<b>f</b>–<b>h</b>) SEM images of a mixed calcite–aragonite region reveal no discernible boundary between the two minerals (see also <a href="#minerals-14-01198-f007" class="html-fig">Figure 7</a>g,h). Dark blue: low contents; yellow and red: high contents.</p> Full article ">
<p>Diameter section of a Polynesian black pearl after a three-month cultivation period. (<b>a</b>–<b>d</b>) Optical view; note the perfectly black section throughout the thickness of the pearl layer. (<b>e</b>–<b>f</b>) Nacreous layer. (<b>g</b>,<b>h</b>) Details of the interface between the nucleus and the first deposits by the pearl sac: (<b>g</b>) secondary electron image; note the basal layer of the series, still rich in organic components; (<b>h</b>) BSE image of the same area.</p> Full article ">Figure 2
<p>Cross-sections of a pearl after a three-month cultivation period. (<b>a</b>–<b>c</b>) Optical images show the non-pigmented phase, characterized by a lack of a prismatic or nacreous structure. (<b>d</b>–<b>f</b>) SEM micrographs reveal the diverse layered structures formed within this region, preceding the resumption of nacre deposition. Note the variety of layered structures produced within this region.</p> Full article ">Figure 3
<p>Microstructures of the pre-nacreous layer in a pearl. (<b>a</b>) Thin pre-nacreous regions with limited prismatic areas. (<b>b</b>) Close-up of prismatic structures. (<b>c</b>,<b>d</b>) Fibrous structures perpendicular to the nucleus surface, with variations in their basal region. (<b>e</b>) Complex mineralization featuring short prism-like structures embedded in a fibrous coverage. (<b>f</b>,<b>g</b>) Details of initial secretions within the basal part of the prism-like structures, and organic envelopes surrounding the prism-like structures. (<b>h</b>) The transition from prism-like structures to black nacre is not straightforward. After the disappearance of the envelopes, prism-like material is overlaid by a granular structure. org: non-mineral nucleus coverage; yellow areas: distinct hemispherical areas; pr: prism-like structures; int: granular intermediate structures.</p> Full article ">Figure 4
<p>Pearl with an extended prism-like initial coverage. (<b>a</b>) Overall distribution of microstructures during the early mineralization period. (<b>b</b>,<b>c</b>) After the formation of a thick organic layer (org. cov), prism-like structures develop: an expansion phase (white arrow) followed by a regression phase (blue arrows). (<b>d</b>) BSE SEM view showing prismatic structures with irregular membranes and a diverse mineral sequence between them. (<b>e</b>,<b>f</b>) Oscillating mineralization: note the regular layered growth of the prism-like structures and the rapidly changing microstructure in (<b>e</b>). Clearly, instability is the main characteristic of the pre-nacreous phase.</p> Full article ">Figure 5
<p>Non-nacreous basal pearl layer after a three-month post-grafting period. (<b>a</b>) Global section showing the distribution of prismatic (blue arrows) and fibrous mineralizations. (<b>b</b>) Basal part: thick initial deposition without mineral structure. Upper part: layered deposition and irregular prism-like structures. (<b>c</b>,<b>d</b>) Optical view of a punctual deposition ((<b>a</b>): blue arrows). (<b>e</b>,<b>h</b>) Aspect of the “prism-like” structure: prisms are discontinuous with abnormal components. (<b>e</b>,<b>g</b>,<b>h</b>) SE SEM images. (<b>f</b>) BSE SEM image. (<b>i</b>) Earliest nacre deposition at the edge of the nacreous layer. (<b>j</b>,<b>k</b>) Optical view showing spots of nacre.</p> Full article ">Figure 6
<p>A three-month cultivation specimen without nacreous secretion: the entire pearl layer is made of “pre-nacreous” material. (<b>a</b>,<b>b</b>) Homogeneous structure of the pearl layer. (<b>c</b>,<b>d</b>) The mineralization mechanism shows only slight variations throughout the pearl layer. Note the consistent thickness of the layering. (<b>e</b>–<b>g</b>) Closer examination (SEM) reveals that this structure was produced by a severely disturbed graft structure. From its basal part ((<b>e</b>): blue arrow) to its outermost layers, the growth layers maintain similar structures composed of global layering (orange arrows) built by associating with short-distance radial elements (green arrows).</p> Full article ">Figure 7
<p>(<b>a</b>–<b>c</b>) Polynesian pearls with an unusually extended pre-nacreous period. (<b>a</b>) Only a thin nacreous layer was formed at the lower part (black arrows). In the pre-nacreous mineralization, red arrows indicate calcite, and green arrows indicate aragonite. (<b>d</b>) An enlarged view of pearl c showing the aragonite and calcite regions from the nucleus to the first layers of black nacre. (<b>e</b>,<b>f</b>) A thick basal part in contact with the nucleus, composed of repeated non-mineralized layers (yellow arrows) and covered by simultaneously produced early stages of the calcite area (red arrows) and aragonite area (green arrows). Note the overall expansion of calcite against aragonite (blue arrows in (<b>d</b>–<b>f</b>)). (<b>g</b>,<b>h</b>) Mixed calcite–aragonite layers with varying thicknesses depending on mineralogy. (<b>i</b>) Progressive regression of calcite and expansion of aragonite layers, leading to the formation of Polynesian black nacre.</p> Full article ">Figure 8
<p>Maps of the closely separating membrane between calcite and aragonite in the <span class="html-italic">Pinctada</span> shell contrasting with the absence of structural separation between calcite and aragonite in pearls. (<b>a</b>) Separating membrane between the calcite and aragonite compartments in a <span class="html-italic">Pinctada</span> shell. SEM image. (<b>b</b>) Raman map illustrating the organic membrane separating aragonite and calcite at the termination of the prismatic region; 2800–3100 cm<sup>−1</sup> indicative of lipids; red color for high organic content , blue dark for low organic content. (<b>c</b>) Optical view of the contact between calcite and aragonite areas in the pre-nacreous domain of a Polynesian <span class="html-italic">Pinctada</span> pearl. (<b>d</b>,<b>e</b>) Raman maps of the calcite–aragonite interface in this region reveal no distinct chemical boundary separating the two mineral components; d: 280 cm<sup>−1</sup> for calcite, e: 1085 cm<sup>−1</sup> for aragonite. Green color for high contents. (<b>f</b>–<b>h</b>) SEM images of a mixed calcite–aragonite region reveal no discernible boundary between the two minerals (see also <a href="#minerals-14-01198-f007" class="html-fig">Figure 7</a>g,h). Dark blue: low contents; yellow and red: high contents.</p> Full article ">
Journal Menu
► ▼ Journal Menu-
- Minerals Home
- Aims & Scope
- Editorial Board
- Reviewer Board
- Topical Advisory Panel
- Instructions for Authors
- Special Issues
- Topics
- Sections & Collections
- Article Processing Charge
- Indexing & Archiving
- Editor’s Choice Articles
- Most Cited & Viewed
- Journal Statistics
- Journal History
- Journal Awards
- Editorial Office
Journal Browser
► ▼ Journal BrowserHighly Accessed Articles
Latest Books
E-Mail Alert
News
7 November 2024
Prof. Dr. Urs Klötzli Appointed Section Editor-in-Chief of Section “Mineral Geochemistry and Geochronology” in Minerals
Prof. Dr. Urs Klötzli Appointed Section Editor-in-Chief of Section “Mineral Geochemistry and Geochronology” in Minerals
5 November 2024
MDPI INSIGHTS: The CEO's Letter #17 - OA Week, Basel Open Day, Beijing Graphene Forum
MDPI INSIGHTS: The CEO's Letter #17 - OA Week, Basel Open Day, Beijing Graphene Forum
Topics
Topic in
Energies, Geosciences, Geotechnics, Minerals, Applied Sciences
Advances in Oil and Gas Wellbore Integrity
Topic Editors: Kai Wang, Jie Wu, Yongjun Deng, Lin ChenDeadline: 31 December 2024
Topic in
Applied Sciences, Energies, Geosciences, Minerals, Remote Sensing
Geomechanics and Engineering Evaluation of Fractured Oil and Gas Reservoirs
Topic Editors: Hu Li, Ahmed E. Radwan, Shuai YinDeadline: 31 March 2025
Topic in
Energies, Geosciences, Minerals, Resources
Drilling, Completion and Well Engineering for the Natural Energy Resources Extraction, Storage and Sustainable Management, 2nd Volume
Topic Editors: Mofazzal Hossain, Hisham Khaled Ben Mahmud, Md Motiur RahmanDeadline: 20 May 2025
Topic in
Energies, Minerals, Safety, Sensors, Sustainability
Mining Safety and Sustainability, 2nd Volume
Topic Editors: Longjun Dong, Ming Xia, Yanlin Zhao, Wenxue ChenDeadline: 30 May 2025
Conferences
Special Issues
Special Issue in
Minerals
Advances in Pyrometallurgy of Minerals and Ores
Guest Editors: Lei Gao, Guo Chen, Bangfu Huang, Fan ZhangDeadline: 29 November 2024
Special Issue in
Minerals
Chemical Weathering Studies
Guest Editors: Diego De Souza Sardinha, Vania Rosolen, Leticia Hirata GodoyDeadline: 29 November 2024
Special Issue in
Minerals
Characterization and Provenance of Archaeological Materials Using Multi-Method Analysis
Guest Editors: Luminita Ghervase, Monica Dinu, Ioana Maria CorteaDeadline: 29 November 2024
Special Issue in
Minerals
New Methods and Technologies for Mineral Geological and Geophysical Exploration in China, 2nd Edition
Guest Editors: Guoqiang Xue, Nannan Zhou, Weiying Chen, Xin WuDeadline: 29 November 2024
Topical Collections
Topical Collection in
Minerals
New Minerals
Collection Editors: Irina O. Galuskina, Igor V. Pekov, Zhenyu Chen
Topical Collection in
Minerals
Clays and Other Industrial Mineral Materials
Collection Editors: Manuel Pozo Rodríguez, Francisco Franco, Michael Stamatakis