[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline

Search Results (196)

Search Parameters:
Keywords = monazite

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
28 pages, 25856 KiB  
Article
Geophysical Methods Applied to the Mineralization Discovery of Rare-Earth Elements at the Fazenda Buriti Alkaline Complex, Goiás Alkaline Province, Brazil
by Fabrício Pereira dos Santos, Marcelo Henrique Leão-Santos, Welitom Rodrigues Borges and Patrícia Caixeta Borges
Minerals 2024, 14(11), 1163; https://doi.org/10.3390/min14111163 - 17 Nov 2024
Viewed by 594
Abstract
In this case study, exploratory techniques were applied for the selection of potential targets for rare-earth elements (REEs) in the Fazenda Buriti Mafic–Ultramafic Complex, part of the Goiás Alkaline Province. The results of the processing and interpretation of aeromagnetic and radiometric data associated [...] Read more.
In this case study, exploratory techniques were applied for the selection of potential targets for rare-earth elements (REEs) in the Fazenda Buriti Mafic–Ultramafic Complex, part of the Goiás Alkaline Province. The results of the processing and interpretation of aeromagnetic and radiometric data associated with the direct measurements of magnetic susceptibility and radiometry in rock samples collected in the study area allowed for the characterization and delimitation of the geological units. The application of Boolean logic in the radiometric data of uranium (U), thorium (Th), and the U/Th ratio allowed for the generation of a prospective map with the delimitation of two exploration targets. A 100 m deep exploratory drill hole was drilled at the main target, intercepting REE mineralization and validating the developed prospective technique. The results contributed to the detailing of a 1:25,000 scale geological map and the interpretation of shallow and deep magnetic structures. Petrophysical data allowed for the estimation of the magnetite content in the main units of the study area. The delimitation of targets with the applied method proved to be efficient after positive geochemical results for REE from the drilled rocks. The total sum of ∑REEs reached 19,629 ppm in the superficial part of the hole and 3,560 ppm in the fresh rock. Mineralogical results in two follow-up drill core samples indicated that monazite was the main REE mineral. Total REE ranged from 34,746 ppm in HG1 to 30,017 ppm in HG2, with LREEs in its majority. The bulk and clay XRD analyses indicated that monazite consisted of 5.7% (HG1) and 5.1% (HG2). The mineral abundance from the TIMA-X analysis indicated 4.2% (HG1) and 4.4% (HG2) in monazite content. Full article
(This article belongs to the Section Mineral Exploration Methods and Applications)
Show Figures

Figure 1

Figure 1
<p>Map of Southern Brazil showing the main occurrences of alkaline rocks and the location of the study area (red dot). Detail for the metamorphic belts represented by BB (Brasília) and RB (Ribeira) (modified from [<a href="#B3-minerals-14-01163" class="html-bibr">3</a>]).</p>
Full article ">Figure 2
<p>Geology of the central-eastern sector of the Tocantins Province, with an emphasis on the Brasília Belt and the study area marked with a yellow dot (modified from [<a href="#B31-minerals-14-01163" class="html-bibr">31</a>]).</p>
Full article ">Figure 3
<p>Goiás Alkaline Province (GAP) with the study area marked by the green polygon. Location of GAP in the Tocantins Province (modified from [<a href="#B34-minerals-14-01163" class="html-bibr">34</a>]).</p>
Full article ">Figure 4
<p>Normal distribution (Gaussian) and anomalous thresholds defined based on the average (μ) and standard deviation (σ). The area under the curve shows the percentage that each threshold represents (modified from [<a href="#B54-minerals-14-01163" class="html-bibr">54</a>]).</p>
Full article ">Figure 5
<p>Map of the geophysical and geological units of the Fazenda Buriti Alkaline Complex containing the location of rock samples collected, interpreted magnetic lineaments and integration of the data used to produce the map. Previous works in [<a href="#B40-minerals-14-01163" class="html-bibr">40</a>].</p>
Full article ">Figure 6
<p>(<b>A</b>): In situ outcrops on a rock ledge at the margins of drainage with granitic gneiss. (<b>B</b>): Sigmoidal porphyroblasts of plagioclase rotated in granitic gneiss. (<b>C</b>): Metagranite outcrop intruded by a centimetric mafic alkaline dike with NW-SE direction. (<b>D</b>): Mylonite sample.</p>
Full article ">Figure 7
<p>(<b>A</b>,<b>B</b>): Decametric and metric boulders and blocks of granitic rock. (<b>C</b>): Granite outcrop with the intrusion of an alkaline mafic porphyritic dike. The red dashed line marks the contact between the dike and the granite. (<b>D</b>): Hand sample of porphyritic granite composed of very coarse phenocrysts of potassium feldspar inserted in a medium grain equigranular matrix with quartz, plagioclase, biotite, and magnetite.</p>
Full article ">Figure 8
<p>(<b>A</b>–<b>C</b>): Decametric blocks and boulders of sandstones outcropping from the Furnas Formation showing preserved sedimentary structures. (<b>D</b>): well-selected fine-grained cream sandstone block—hand specimen.</p>
Full article ">Figure 9
<p>(<b>A</b>): Outcrop with metric alkaline gabbro blocks in the meadow. (<b>B</b>): Medium-grained pyroxenite hand sample composed essentially of clinopyroxene, magnetite, olivine, and plagioclase. (<b>C</b>): Outcrop with alkaline breccia formed by fragments of pyroxenite/gabbro and syenite matrix. (<b>D</b>): Syenite sample found in a decametric “pocket” in the middle of alkaline gabbro.</p>
Full article ">Figure 10
<p>(<b>A</b>): Thin section—Pyroxenite formed by fractured and zoned phenocrysts of augite, biotite, and opaque minerals in the interstices and as inclusions. A 25× magnification with parallel polarizers. (<b>B</b>): Same thin section as A but with crossed polarizers. Op = opaque mineral, Aug = augite, Bt = biotite.</p>
Full article ">Figure 11
<p>(<b>A</b>): Blocks with microsyenite. (<b>B</b>): Sample of porphyritic microsyenite with a grayish color.</p>
Full article ">Figure 12
<p>Box plot for the petrophysical data of magnetic susceptibility and radioelements Potassium, Thorium, and Uranium recorded in rock samples.</p>
Full article ">Figure 13
<p>Magnetic maps, interpreted magnetic lineaments, and drainage. (<b>A</b>): Inclination of Total Gradient Amplitude (ITGA). (<b>B</b>): Block diagram with 3.5 times vertical exaggerations, with the Total Gradient (TG) associated with elevations. (<b>C</b>): First horizontal derivative in the X direction (dX). (<b>D</b>): First horizontal derivative in the Y direction (dY). (<b>E</b>): First Vertical derivative (dZ). Elevation [<a href="#B39-minerals-14-01163" class="html-bibr">39</a>].</p>
Full article ">Figure 14
<p>Radiometric and calculated ratios maps, lithological contacts, and drainage. (<b>A</b>): Potassium (K) concentrations; (<b>B</b>): Equivalent concentrations of Thorium (eTh); (<b>C</b>): Equivalent concentrations of Uranium (eU); (<b>D</b>): Block diagram with 3.5 vertical exaggerations, with RGB ternary image (K-eTh-eU) associated with elevations. (<b>E</b>): Estimated <span class="html-italic">F</span>-parameter [<a href="#B50-minerals-14-01163" class="html-bibr">50</a>,<a href="#B51-minerals-14-01163" class="html-bibr">51</a>], and; (<b>F</b>): U/Th ratio. Altitude data [<a href="#B39-minerals-14-01163" class="html-bibr">39</a>].</p>
Full article ">Figure 15
<p>Flowchart for integrating eU, eTh, and eU/eTh anomalies using the arithmetic average (μ) and standard deviation (σ) for the first threshold (μ + 1 × σ), second threshold (μ + 2 × σ), and third threshold (μ + 3 × σ). Arrows indicate the direction of overlay on the prospective map.</p>
Full article ">Figure 16
<p>(<b>A</b>): Radiometric anomalies were selected based on the average (µ) plus one standard deviation (σ), indicating targets and the location of the drill hole. (<b>B</b>): Block diagram with the selected anomalies over the relief. (<b>C</b>): Target 1. (<b>D</b>): Target 2. Altitude data [<a href="#B39-minerals-14-01163" class="html-bibr">39</a>].</p>
Full article ">Figure 17
<p>Drill hole log, total sum of rare-earth elements (∑REEs) in ppm (parts per million), light rare-earth elements (LREEs) sum in ppm, and heavy rare-earth elements (HREEs) sum in ppm.</p>
Full article ">Figure 18
<p>Summary of modal mineralogy of the samples HG1 and HG2.</p>
Full article ">Figure 19
<p>Summary of REE and Nb minerals of the samples HG1 and HG2.</p>
Full article ">
19 pages, 4996 KiB  
Article
Characterization of Heavy Minerals and Their Possible Sources in Quaternary Alluvial and Beach Sediments by an Integration of Microanalytical Data and Spectroscopy (FTIR, Raman and UV-Vis)
by Adel A. Surour and Amira M. El-Tohamy
Quaternary 2024, 7(4), 46; https://doi.org/10.3390/quat7040046 - 22 Oct 2024
Viewed by 791
Abstract
Quaternary stream sediments and beach black sand in north-western Saudi Arabia (namely Wadi Thalbah, Wadi Haramil and Wadi Al Miyah) are characterized by the enrichment of heavy minerals. Concentrates of the heavy minerals in two size fractions (63–125 μm and 125–250 μm) are [...] Read more.
Quaternary stream sediments and beach black sand in north-western Saudi Arabia (namely Wadi Thalbah, Wadi Haramil and Wadi Al Miyah) are characterized by the enrichment of heavy minerals. Concentrates of the heavy minerals in two size fractions (63–125 μm and 125–250 μm) are considered as potential sources of “strategic” accessory minerals. A combination of mineralogical, geochemical and spectroscopic data of opaque and non-opaque minerals is utilized as clues for provenance. ThO2 (up to 17.46 wt%) is correlated with UO2 (up to 7.18 wt%), indicating a possible uranothorite solid solution in zircon. Hafnoan zircon (3.6–5.75 wt% HfO2) is a provenance indicator that indicates a granitic source, mostly highly fractionated granite. In addition, monazite characterizes the same felsic provenance with rare-earth element oxides (La, Ce, Nd and Sm amounting) up to 67.88 wt%. These contents of radionuclides and rare-earth elements assigned the investigated zircon and monazite as “strategic” minerals. In the bulk black sand, V2O5 (up to 0.36 wt%) and ZrO2 (0.57 wt%) are correlated with percentages of magnetite and zircon. Skeletal or star-shaped Ti-magnetite is derived from the basaltic flows. Mn-bearing ilmenite, with up to 5.5 wt% MnO, is derived from the metasediments. The Fourier-transform infrared transmittance (FTIR) spectra indicate lattice vibrational modes of non-opaque silicate heavy minerals, e.g., amphiboles. In addition, the FTIR spectra show O-H vibrational stretching that is related to magnetite and Fe-oxyhydroxides, particularly in the magnetic fraction. Raman data indicate a Verwey transition in the spectrum of magnetite, which is partially replaced by possible ferrite/wüstite during the measurements. The Raman shifts at 223 cm−1 and 460 cm−1 indicate O-Ti-O symmetric stretching vibration and asymmetric stretching vibration of Fe-O bonding in the FeO6 octahedra, respectively. The ultraviolet-visible-near infrared (UV-Vis-NIR) spectra confirm the dominance of ferric iron (Fe3+) as well as some Si4+ transitions of magnetite (226 and 280 nm) in the opaque-rich fractions. Non-opaque heavy silicates such as hornblende and ferrohornblende are responsible for the 192 nm intensity band. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location map of the study area in north-western Saudi Arabia; (<b>b</b>) Sample locations in three investigated wadis shown on a Google-based satellite image; (<b>c</b>) Modified geological map of Al Wajh quadrangle based on lithological boundaries from [<a href="#B32-quaternary-07-00046" class="html-bibr">32</a>] and updated nomenclature from [<a href="#B30-quaternary-07-00046" class="html-bibr">30</a>].</p>
Full article ">Figure 2
<p>Opaque minerals in the heavy concentrates. (<b>a</b>) Skeletal Ti-bearing magnetite (Ti-Mag) and titanite (Ttn), coarse black sand, Wadi Thalbah; (<b>b</b>) Homogeneous ilmenite (Ilm) partly replaced by sub-graphic intergrowth of rutile and hematite (RH), coarse black sand, Wadi Thalbah; (<b>c</b>) Homogeneous ilmenite (Ilm) with continuous titanite reaction rim (Tnt), fine black sand, Wadi Thalbah. (<b>d</b>) Freshness and euhedrality of homogeneous ilmenite (Ilm), fine black sand, Wadi Thalbah; (<b>e</b>) Coarse-trellis ilmenite (Ilm)-magnetite (Mag) intergrowth, fine wadi alluvium, Wadi Al Miyah. (<b>f</b>) Banded ilmenite (Ilm)-magnetite (Mag) intergrowth, fine wadi alluvium, Wadi Al Miyah; (<b>g</b>) Fine exsolved ilmenite (Ilm) confined to the (111) octahedral planes in host magnetite (Mag) forming fine network intergrowth, coarse wadi alluvium, Wadi Haramil; (<b>h</b>) Alteration of pyrite (Py) to goethite (Gth) along fractures and peripheries, coarse wadi alluvium, Wadi Haramil.</p>
Full article ">Figure 3
<p>Binary relationships of some major oxides in the heavy concentrates whole-fractions (magnetic and non-magnetic).</p>
Full article ">Figure 4
<p>BSE images and spectral analyses of non-opaque accessory minerals from black sand from Wadi Thalbah. (<b>a</b>) Secondary titanite (Tnt) forms at the expense of Ti-bearing magnetite (Ti-Mag); (<b>b</b>) Fractured euhedral zircon (Zrn); (<b>c</b>) Fractured anhedral zircon (Zrn); (<b>d</b>) Interlocked silicate (grey) and magnetite (Mag) with minute inclusions of monazite (Mnz). A red Astrix is added to indicate exact position of analysis.</p>
Full article ">Figure 5
<p>BSE images and spectral analyses of non-opaque accessory minerals from the wadi alluvium. (<b>a</b>) Extremely fine zircon (Zrn) at the peripheral zone of coarse hornblende (Hbl), fine fraction, Wadi Thalbah; (<b>b</b>) Subhedral U- and Th-bearing monazite (Mnz), fine fraction, Wadi Thalbah; (<b>c</b>) Anhedral zoned zircon (Zrn), fine fraction Wadi Al Miyah; (<b>d</b>) Fractured anhedral zircon (Zrn), coarse fraction, Wadi Al Miyah. A red Astrix is added to indicate exact position of analysis.</p>
Full article ">Figure 6
<p>FTIR spectra of the heavy concentrates. (<b>a</b>) Bulk/whole-fraction (magnetic and non-magnetic); (<b>b</b>) Three representative magnetic fractions. (1) Fine black sand, Wadi Thalbah, (2) Coarse black sand, Wadi Thalbah, (3) Fine wadi alluvium, Wadi Al Miyah, (4) Fine wadi alluvium, Wadi Thalbah, (5) Coarse black sand, Wadi Thalbah, and (6) Coarse wadi alluvium, Wadi Haramil.</p>
Full article ">Figure 7
<p>Raman and UV spectra of the whole-fraction heavy concentrates. (<b>a</b>) Raman spectra; (<b>b</b>) UV spectra up to 400 nm; (<b>c</b>) UV spectra up to 2500 nm. (1) Fine black sand, Wadi Thalbah, (2) Coarse black sand, Wadi Thalbah, (3) Fine wadi alluvium, Wadi Al Miyah, (4) Fine wadi alluvium, Wadi Thalbah, (5) Coarse black sand, Wadi Thalbah, and (6) Coarse wadi alluvium, Wadi Haramil.</p>
Full article ">
17 pages, 5039 KiB  
Article
Occurrences of the Rare, REE Minerals Daqingshanite, Törnebohmite, Biraite, Sahamalite, and Ferriperbøeite from the Sheep Creek Area, Montana, USA
by Christopher H. Gammons, Sarah Risedorf, Gary Wyss and Heather Lowers
Minerals 2024, 14(10), 1047; https://doi.org/10.3390/min14101047 - 18 Oct 2024
Viewed by 694
Abstract
Over 30 small, discontinuous, tabular carbonatite bodies are located in the Sheep Creek area, Ravalli County, southwest Montana. The age and origin of these REE-Nb-rich deposits are currently being investigated. The purpose of this paper is to document the occurrence of several rare [...] Read more.
Over 30 small, discontinuous, tabular carbonatite bodies are located in the Sheep Creek area, Ravalli County, southwest Montana. The age and origin of these REE-Nb-rich deposits are currently being investigated. The purpose of this paper is to document the occurrence of several rare minerals, including daqingshanite, törnebohmite, biraite, sahamalite, and ferriperbøeite, in two of the carbonatite bodies. These minerals are found in association with monazite, hydroxylbastnäsite, ferriallanite, calcite, dolomite, baryte, quartz, actinolite, apatite, celsian, and Sr-rich aragonite. Automated SEM-EDS was used to target the areas of interest in polished specimens for more detailed spot SEM-EDS and electron probe microanalysis. Raman spectra were also acquired for each of the rare minerals. The complex mineralogy of the Sheep Creek carbonatites is most likely due to several overlapping thermal events, including primary magmatic, overprinting hydrothermal, and supergene weathering stages. The rare minerals described in this study are believed to be hydrothermal and/or carbothermal in origin, although no estimates of temperature are available at this time. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Simplified geology of the igneous rocks of the Idaho batholith (modified from [<a href="#B15-minerals-14-01047" class="html-bibr">15</a>]); (<b>B</b>) approximate extent of the 1.37 Ga meta-igneous complex showing REE deposits and districts mentioned in the text; (1) Sheep Creek; (2) Mineral Hill; (3) Diamond Creek; (4) Lemhi Pass; (5) Idaho Cobalt Belt; NFT = North Fork Thrust; Ysu = Mesoproterozoic Belt Supergroup and cover rocks; (<b>C</b>) map of the Sheep Creek area showing known outcrops of carbonatite (after [<a href="#B2-minerals-14-01047" class="html-bibr">2</a>,<a href="#B16-minerals-14-01047" class="html-bibr">16</a>]).</p>
Full article ">Figure 2
<p>Photographs of the samples. (<b>A</b>) Thin section billet (2.7 cm × 4.6 cm) of sample 722-3A. (<b>B</b>) Sawn slab (8 cm wide) of sample 720-3A. In both photos, the black mineral is allanite/ferriallanite, the milky white mineral is calcite, and the pinkish zones are complex intergrowths of REE-carbonates, phosphates, and silicates. The yellowish areas contain dolomite and/or baryte.</p>
Full article ">Figure 3
<p>(<b>A</b>) Automated TIMA-SEM mineral map of sample 720-3A. The most abundant minerals indexed are allanite/ferriallanite (dark gray), calcite (pale blue), baryte (yellow), hydroxylbastnäsite (red), quartz (tan), and törnebohmite (dark blue). The white vertical band is a crack in the sample. (<b>B</b>) Backscattered electron (BSE) image of a törnebohmite grain (tör) intergrown with hydroxylbastnäsite (bst). (<b>C</b>) BSE closeup image of a porous mass of hydroxylbastnäsite (bst) with abundant inclusions of hematite (hem).</p>
Full article ">Figure 4
<p>(<b>A</b>) Automated TIMA-SEM mineral map of sample 722-3A. The yellow dashed line shows a possible precursor grain with a hexagonal outline. (<b>B</b>,<b>C</b>) Closeups showing the intergrowths of calcite (cal), dolomite (dol), baryte (bar), apatite (ap), monazite (mz), celsian (cls), magnetite (mt), actinolite (act), daqingshanite (dq), hydroxylbastnäsite (bst), allanite (aln), quartz (qtz), and Sr-rich aragonite (arg). The panels labeled 5A, 5B, and 5C in A are enlarged in <a href="#minerals-14-01047-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 5
<p>(<b>A</b>–<b>C</b>) BSE images of the areas of interest in sample 722-3A (see <a href="#minerals-14-01047-f004" class="html-fig">Figure 4</a> for locations). (<b>D</b>) REE-bearing silicates from a sample collected at the stockpile at Crowley Adit 3. Abbreviations: cal = calcite; fal = ferriallanite; dq = daqingshanite; fpb = ferriperbøeite; mz = monazite; qtz = quartz; shm = sahamalite; bir = biraite; epd = epidote; bst = hydroxylbastnäsite.</p>
Full article ">Figure 6
<p>Representative Raman spectra (514 nm) for the REE minerals in this study (black traces). (<b>A</b>) Biraite-(Ce) from this study is compared to biraite-(La) from [<a href="#B21-minerals-14-01047" class="html-bibr">21</a>]; (<b>B</b>) sahamalite-(Ce) from this study is compared to RRUFF ID:R080043 (red dashed line); (<b>C</b>) spectrum for daqingshanite-(Ce) from this study (top pattern, no background subtraction) compared to RRUFF ID:R070389 (red dashed line). (<b>D</b>), (<b>E</b>), and (<b>F</b>) show Raman spectra for törnebohmite, ferriallanite, and ferriperbøeite, respectively.</p>
Full article ">Figure 7
<p>Fluid inclusions at room temperature. (<b>A</b>) Carbonic (3-phase, CO<sub>2</sub>-H<sub>2</sub>O) fluid inclusions in coarse (&gt;1 cm) monazite; (<b>B</b>) hypersaline inclusions in allanite.</p>
Full article ">
27 pages, 11358 KiB  
Article
Geochemistry and Mineralogy of Upper Paleozoic Coal in the Renjiazhuang Mining District, Northwest Ordos Basin, China: Evidence for Sediment Sources, Depositional Environment, and Elemental Occurrence
by Meng Wu, Yong Qin, Guchun Zhang, Jian Shen, Jianxin Yu, Xiaoyan Ji, Shifei Zhu, Wenqiang Wang, Yali Wan, Ying Liu and Yunhu Qin
Minerals 2024, 14(10), 1045; https://doi.org/10.3390/min14101045 - 18 Oct 2024
Viewed by 635
Abstract
This study aims to investigate the depositional environment, sediment sources, and elemental occurrence of Upper Paleozoic coal in the Renjiazhuang Mining District, Western Ordos Basin. Furthermore, SEM-EDX, optical microscope (OM), ICP-AES, ICP-MS, and AAS were used. Compared with hard coal of the world, [...] Read more.
This study aims to investigate the depositional environment, sediment sources, and elemental occurrence of Upper Paleozoic coal in the Renjiazhuang Mining District, Western Ordos Basin. Furthermore, SEM-EDX, optical microscope (OM), ICP-AES, ICP-MS, and AAS were used. Compared with hard coal of the world, M3 coals were enriched in Ga, Li, Zr, Be, Ta, Hf, Nb, Pb, and Th, M5 coals were enriched in Li (CC = 10.21), Ta (CC = 6.96), Nb (CC = 6.95), Be, Sc, Ga, Hf, Th, Pb, Zr, In, and REY, while M9 coals were enriched in Li (CC = 14.79), Ta (CC = 5.41), Ga, W, Hf, Nb, Zr, Pb, and Th. In addition, minerals were mainly composed of kaolinite, dolomite, pyrite, feldspar, calcite, and quartz, locally visible minor amounts of monazite, zircon, clausthalite, chalcopyrite, iron dolomite, albite, fluorite, siderite, galena, barite, boehmite, and rutile. In addition, maceral compositions of M3 coals and M9 coals were dominated by vitrinite (up to 78.50%), while M5 coals were the main inertite (up to 76.26%), and minor amounts of liptinite. REY distribution patterns of all samples exhibited light REY enrichment and negative Eu anomalies. The geochemistry of samples (TiO2 and Al2O3, Nb/Y and Zr × 0.0001/TiO2 ratios, and REY enrichment types) indicates that the sediment sources of samples originated from felsic igneous rocks. Indicator parameters (TPI, GI, VI, GWI, V/I, Sr/Ba, Th/U, and CeN/CeN*) suggest that these coals were formed in different paleopeat swamp environments: M3 coal was formed in a lower delta plain and terrestrial (lacustrine) facies with weak oxidation and reduction, and M5 coal was formed in a terrestrial and dry forest swamp environment with weak oxidation–oxidation, while M9 coal was formed in a seawater environment of humid forest swamps and the transition from the lower delta plain to continental sedimentation with weak oxidation and reduction. Statistical methods were used to study the elemental occurrence. Moreover, Li, Ta, Hf, Nb, Zr, Pb, and Th elements were associated with aluminosilicates, and Ga occurred as silicate. Full article
(This article belongs to the Section Mineral Geochemistry and Geochronology)
Show Figures

Figure 1

Figure 1
<p>Geological map of the northwest Ordos Basin and structural sketch-map of Renjiazhuang Mining District, modified after Wu et al. [<a href="#B32-minerals-14-01045" class="html-bibr">32</a>], Zhao [<a href="#B38-minerals-14-01045" class="html-bibr">38</a>], and Zhang [<a href="#B39-minerals-14-01045" class="html-bibr">39</a>]. (<b>a</b>) locations of the Ordos Basin in China, (<b>b</b>). geological map of the northwest Ordos Basin (<b>c</b>) structural sketch map of the Renjiazhuang Mining District.</p>
Full article ">Figure 2
<p>Lithological sequence and sample sections of the Renjiazhuang Mining District. The sample numbers from D-R-1 to D-R-15 are from Wu et al. [<a href="#B32-minerals-14-01045" class="html-bibr">32</a>].</p>
Full article ">Figure 3
<p>Liptinite, inertinite, and vitrinite in the samples. UV -light and reflected light reflectance, oil immersion. (<b>a</b>) Collodetrinite with distribution of clay minerals and fusinite, sample N-H-1; (<b>b</b>) collotelinite, sample T-H-1; (<b>c</b>) clay filling the telinite, sample T-H-2; (<b>d</b>) semifusinite cells filled with clay minerals, sample F-H-M; (<b>e</b>) vitrodetrinite and inertodetrinite distributing in clay, sample F-H-2; (<b>f</b>) fusinite, sample T-H-2; (<b>g</b>) clay minerals embedded with vitrodetrinite and macrinite, sample F-H-2; (<b>h</b>) micrinite, semifusinite, and clay minerals, sample T-H-M; (<b>i</b>) sporinite, sample T-H-2; (<b>j</b>) resinite and sporinite, sample T-H-M; (<b>k</b>) collotelinite embedded with banded cutinite, sample N-H-2; (<b>l</b>) barkinite, sample T-H-2.</p>
Full article ">Figure 4
<p>The concentration coefficients of trace elements and REY in the samples. (<b>a</b>) the enrichment coefficient of trace elements in the M3 coals. (<b>b</b>) the enrichment coefficient of trace elements in the M5 coals. (<b>c</b>) the enrichment coefficient of trace elements in the M9 coals. (<b>d</b>) the enrichment coefficient of trace elements in the non-coals.</p>
Full article ">Figure 5
<p>The REY distribution in coal (<b>a</b>) and non-coal (<b>b</b>) samples.</p>
Full article ">Figure 6
<p>Mineral composition and distribution in the samples, reflected light, oil immersion. (<b>a</b>) pyrite-filled fractures, sample T-H-1; (<b>b</b>) pyrite in cell lumens, sample T-H-1; (<b>c</b>) massive pyrite, sample N-H-1; (<b>d</b>) well-developed spheroid pyrite occurring in calcite, sample F-H-2; (<b>e</b>) framboidal pyrite and clay minerals, sample N-H-1; (<b>f</b>) granular and disseminated pyrite, sample N-H-2; (<b>g</b>) clay minerals filling in cell lumens, sample T-H-2; (<b>h</b>) irregular massive calcite-filled fractures, sample T-H-2; (<b>i</b>) granular quartz and kaolinite, sample T-H-2.</p>
Full article ">Figure 7
<p>Images of kaolinite in the samples under scanning electron microscopy. (<b>a</b>) Fissured-filling kaolinite, sample T-H-2. (<b>b</b>) Cell-filling kaolinite, sample N-H-1. (<b>c</b>) Flaky and aggregated kaolinite, sample F-H-2. (<b>d</b>) Dispersed kaolinite, sample F-H-2. (<b>e</b>) Lens-like kaolinite, sample T-H-1. (<b>f</b>) Irregular massive kaolinite, sample N-H-2.</p>
Full article ">Figure 8
<p>SEM-EDX images of minerals in the samples. (<b>a</b>) Agglomerate diaspore, sample N-H-1; (<b>b</b>) EDX spectrum of spot 1; (<b>c</b>) boehmite, kaolinite, and brannerite, sample N-H-2; EDX spectrum corresponding to spot 2 (<b>d</b>), spot 3 (<b>e</b>), and spot 4 (<b>f</b>).</p>
Full article ">Figure 9
<p>SEM of minerals in the samples. (<b>a</b>) Calcite and dolomite with elongated columnar and irregular blocky forms, sample T-H-M; (<b>b</b>) massive dolomite, sample N-H-2; (<b>c</b>) cell-filling calcite and flaky kaolinite, sample T-H-1; (<b>d</b>) quartz and calcite, sample T-H-M.</p>
Full article ">Figure 10
<p>SEM images of minerals in the samples. (<b>a</b>,<b>b</b>) Agglomerated ankerite, sample N-H-1; (<b>c</b>,<b>d</b>) granular zircon and flocculent kaolinite, sample T-H-1; (<b>e</b>,<b>f</b>) agglomerated fluorite, sample N-H-2; (<b>g</b>,<b>h</b>) albite with elongated columnar and irregular massive forms, sample N-H-1; (<b>i</b>,<b>j</b>) agglomerated monazite and flocculent kaolinite, sample T-H-1; (<b>k</b>,<b>l</b>) chalcopyrite and kaolinite, sample F-H-2; (<b>m</b>,<b>n</b>) irregular blocky barite, sample N-H-2; (<b>o</b>,<b>p</b>) fine-grained crystalline aggregated siderite, sample T-H-M; (<b>q</b>,<b>r</b>) irregular massive galena, sample T-H-M; (<b>s</b>,<b>t</b>) agglomerated clausthalite, sample T-H-2.</p>
Full article ">Figure 11
<p>Relationship diagrams for Al<sub>2</sub>O<sub>3</sub> vs. TiO<sub>2</sub> (<b>a</b>) and Nb/Y vs. Zr × 10<sup>−3</sup>/TiO<sub>2</sub> (<b>b</b>) for identifying the source rock [<a href="#B83-minerals-14-01045" class="html-bibr">83</a>]. M5 coal* and M9 coal* samples are from Ji et al. [<a href="#B27-minerals-14-01045" class="html-bibr">27</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) The intersection map of TPI and GI for indicating the sedimentary environment of coal seams [<a href="#B50-minerals-14-01045" class="html-bibr">50</a>]. (<b>b</b>) The intersection map of GWI and VI for indicating the coal-forming environment [<a href="#B49-minerals-14-01045" class="html-bibr">49</a>].</p>
Full article ">Figure 13
<p>Variations in total sulfur, ash yield, Sr/Ba, Th/U, and Ce<span class="html-italic"><sub>N</sub></span>/Ce<span class="html-italic"><sub>N</sub></span>* in the samples.</p>
Full article ">
16 pages, 4792 KiB  
Article
Leaching Efficacy of Ethylenediaminetetraacetic Acid (EDTA) to Extract Rare-Earth Elements from Monazite Concentrate
by Ammar S. A. Al Sheidi, Laurence G. Dyer and Bogale Tadesse
Crystals 2024, 14(10), 829; https://doi.org/10.3390/cryst14100829 - 24 Sep 2024
Viewed by 1077
Abstract
Alkaline EDTA solution has been previously identified as an effective leaching agent for solubilising rare-earth oxalates. These oxalates are the product of an oxalic acid conversion leach dissolving monazite and redepositing the salt. Pervious work suggested a significant increase in recovery was observed [...] Read more.
Alkaline EDTA solution has been previously identified as an effective leaching agent for solubilising rare-earth oxalates. These oxalates are the product of an oxalic acid conversion leach dissolving monazite and redepositing the salt. Pervious work suggested a significant increase in recovery was observed between pH 8 and 10; we have demonstrated that, in an excess of EDTA, this is not the case, and the dissolution is similar. While demonstrating that, at a nominal solid loading of 100 g/L, 0.2 M EDTA solution produced the highest dissolution, elevated solids require an equivalent increase in lixiviant concentration driven by consumption. Very-high-solution concentrations (>50 g/L dissolved TREEs) were achieved at a high solid loading, indicating both that a solution equilibrium is yet to be reached and that a build-up of oxalate in the system (estimated at ~1 M) does not impact the leach efficiency. We have also demonstrated the recycling of EDTA to use in multiple stages as well as the ability to recover oxalate from this solution. Full article
(This article belongs to the Special Issue Crystallization and Purification)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Stability constant of organic solvents with rare-earth elements.</p>
Full article ">Figure 2
<p>Molecular structure of oxalic acid.</p>
Full article ">Figure 3
<p>Molecular structure of EDTA.</p>
Full article ">Figure 4
<p>Distribution of the particle size. Red line is the particle size of monazite concentrates. Green line is the particle size of rare-earth oxalate.</p>
Full article ">Figure 5
<p>X-ray diffractogram of oxalic acid leach residue demonstrating characteristic rare-earth oxalate (O) peaks along with the two other predominant primary minerals monazite (M) and Goethite (G).</p>
Full article ">Figure 6
<p>Dissolution of elements without an oxalic acid step: concentrations (ppm) of various elements recovered from the monazite concentrate directly leached with EDTA without the oxalic acid conversion step. The data show significantly lower recovery efficiencies for key rare-earth elements (Ce, La, Pr, and Nd), highlighting the importance of the oxalic acid step in enhancing the solubility and reactivity of these elements for efficient extraction.</p>
Full article ">Figure 7
<p>The effect of temperature on the leaching of rare-earth elements.</p>
Full article ">Figure 8
<p>The impact of EDTA molarity on the leaching of rare-earth elements.</p>
Full article ">Figure 9
<p>Presents the pH dependency of REE dissolution.</p>
Full article ">Figure 10
<p>X-ray diffractograms for the residue samples from the varied pH leach tests and a reference pattern for rare-earth hydroxide. The marked peaks are either labelled M (monazite) or H (hydroxide).</p>
Full article ">Figure 11
<p>Leaching of rare-earth elements over time.</p>
Full article ">Figure 12
<p>The relationship between the solid-to-liquid ratio and the dissolution of rare-earth elements.</p>
Full article ">Figure 13
<p>The recovery of rare-earth elements at different S/L ratios.</p>
Full article ">Figure 14
<p>Concentration of rare-earth elements during leaching and precipitation cycles.</p>
Full article ">Figure 15
<p>X-ray diffractogram of precipitate formed upon the addition of calcium to the EDTA leach residue following REE removal.</p>
Full article ">
21 pages, 9535 KiB  
Article
Petrogenesis of Eocene A-Type Granite Associated with the Yingpanshan–Damanbie Regolith-Hosted Ion-Adsorption Rare Earth Element Deposit in the Tengchong Block, Southwest China
by Zhong Tang, Zewei Pan, Tianxue Ming, Rong Li, Xiaohu He, Hanjie Wen and Wenxiu Yu
Minerals 2024, 14(9), 933; https://doi.org/10.3390/min14090933 - 12 Sep 2024
Viewed by 634
Abstract
The ion-adsorption-type rare earth element (iREE) deposits dominantly supply global resources of the heavy rare earth elements (HREEs), which have a critical role in a variety of advanced technological applications. The initial enrichment of REEs in the parent granites controls the formation of [...] Read more.
The ion-adsorption-type rare earth element (iREE) deposits dominantly supply global resources of the heavy rare earth elements (HREEs), which have a critical role in a variety of advanced technological applications. The initial enrichment of REEs in the parent granites controls the formation of iREE deposits. Many Mesozoic and Cenozoic granites are associated with iREE mineralization in the Tengchong block, Southwest China. However, it is unclear how vital the mineralogical and geochemical characteristics of these granites are to the formation of iREE mineralization. We conducted geochronology, geochemistry, and Hf isotope analyses of the Yingpanshan–Damanbie granitoids associated with the iREE deposit in the Tengchong block with the aims to discuss their petrogenesis and illustrate the process of the initial REE enrichment in the granites. The results showed that the Yingpanshan–Damanbie pluton consists of syenogranite and monzogranite, containing REE-bearing accessory minerals such as monazite, xenotime, apatite, zircon, allanite, and titanite, with a high REE concentration (210–626 ppm, mean value is 402 ppm). The parent granites have Zr + Nb + Ce + Y (333–747 ppm) contents and a high FeOT/MgO ratio (5.89–11.4), and are enriched in Th (mean value of 43.6 ppm), U (mean value of 4.57 ppm), Zr (mean value of 305 ppm), Hf (mean value of 7.94 ppm), Rb (mean value of 198 ppm), K (mean value of 48,902 ppm), and have depletions of Sr (mean value of 188 ppm), Ba (mean value of 699 ppm), P (mean value of 586 ppm), Ti (mean value of 2757 ppm). The granites plot in the A-type area in FeOT/MgO vs. Zr + Nb + Ce + Y and Zr vs. 10,000 Ga/Al diagrams, suggesting that they are A2-type granites. These granites are believed to have formed through the partial melting of amphibolites at a post-collisional extension setting when the Tethys Ocean closed. REE-bearing minerals (e.g., apatite, titanite, allanite, and fluorite) and rock-forming minerals (e.g., potassium feldspar, plagioclase, biotite, muscovite) supply rare earth elements in weathering regolith for the Yingpanshan–Damanbie iREE deposit. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Tectonic map of the eastern Tethys domain (modified after Wang et al. [<a href="#B47-minerals-14-00933" class="html-bibr">47</a>]); (<b>b</b>) Geological map of the Tengchong block with iREE deposits (modified after Deng et al. [<a href="#B48-minerals-14-00933" class="html-bibr">48</a>]); (<b>c</b>) U–Pb ages histogram of zircons from magmatic rocks in the Tengchong block (date from He et al. [<a href="#B15-minerals-14-00933" class="html-bibr">15</a>], Dong et al. [<a href="#B16-minerals-14-00933" class="html-bibr">16</a>], Xu et al. [<a href="#B18-minerals-14-00933" class="html-bibr">18</a>], Yang et al. [<a href="#B19-minerals-14-00933" class="html-bibr">19</a>], Li et al. [<a href="#B26-minerals-14-00933" class="html-bibr">26</a>], Zou et al. [<a href="#B30-minerals-14-00933" class="html-bibr">30</a>], Cong et al. [<a href="#B35-minerals-14-00933" class="html-bibr">35</a>], Cao et al. [<a href="#B39-minerals-14-00933" class="html-bibr">39</a>], Xie et al. [<a href="#B42-minerals-14-00933" class="html-bibr">42</a>], Chen et al. [<a href="#B44-minerals-14-00933" class="html-bibr">44</a>], Cong et al. [<a href="#B49-minerals-14-00933" class="html-bibr">49</a>], Li et al. [<a href="#B50-minerals-14-00933" class="html-bibr">50</a>], Chen et al. [<a href="#B51-minerals-14-00933" class="html-bibr">51</a>], Zhu et al. [<a href="#B52-minerals-14-00933" class="html-bibr">52</a>]).</p>
Full article ">Figure 2
<p>(<b>a</b>) Geological map of the Yingpanshan–Damanbie pluton with the Yingpanshan–Damanbie iREE deposit. (<b>b</b>) A profile of the regolith with iREE mineralization from the Yingpanshan–Damanbie iREE deposit.</p>
Full article ">Figure 3
<p>Characteristics of petrography and REE-bearing accessory minerals of syenogranite from the Yingpanshan–Damanbie iREE deposit in western Yunnan. (<b>a</b>) Photograph of a sample specimen; (<b>b</b>) photomicrograph; and (<b>c</b>) TIMA images of thin section; abbreviations: Kfs = K–feldspar, Qtz = quartz, Pl = plagioclase, Bt = biotite, Ep = Epidote.</p>
Full article ">Figure 4
<p>Characteristics of petrography and REE accessory minerals of monzogranite from the Yingpanshan–Damanbie iREE deposit in western Yunnan. (<b>a</b>) Photograph of a sample specimen; (<b>b</b>) photomicrograph; and (<b>c</b>) TIMA images of representative thin sections; abbreviations: Kfs = K–feldspar, Qtz = quartz, Pl = plagioclase, Bt = biotite, Ep = Epidote, Hb = hornblende.</p>
Full article ">Figure 5
<p>U–Pb concordia diagrams for (<b>a</b>) monzogranite (L–1–B6) and (<b>b</b>) syenogranite (L–1–B5) from the Yingpanshan–Damanbie iREE deposit and CL images of representative zircon grains.</p>
Full article ">Figure 6
<p>Plots of (<b>a</b>) (K<sub>2</sub>O + Na<sub>2</sub>O) versus SiO<sub>2</sub>, (<b>b</b>) A/NK versus A/CNK, (<b>c</b>) K<sub>2</sub>O versus SiO<sub>2</sub>, (<b>d</b>) K<sub>2</sub>O/Na<sub>2</sub>O versus SiO<sub>2</sub> of monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit.</p>
Full article ">Figure 7
<p>Plots of chondrite-normalized REE patterns (<b>a</b>,<b>c</b>) and primitive mantle (PM)-normalized spider diagrams (<b>b</b>,<b>d</b>) for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. Values for normalization are from Sun and McDonough [<a href="#B68-minerals-14-00933" class="html-bibr">68</a>], respectively. UCC = upper continental crust; LCC = lower continental crust; UCC and LCC data from Jahn et al. [<a href="#B69-minerals-14-00933" class="html-bibr">69</a>].</p>
Full article ">Figure 8
<p>Plots of (<b>a</b>) La/Sm versus La, (<b>b</b>) Zr/Hf versus SiO<sub>2</sub>, (<b>c</b>) FeO<sup>T</sup>/MgO versus (Zr + Y + Nb + Ce), (<b>d</b>) Zr versus 10,000 Ga/Al) [<a href="#B83-minerals-14-00933" class="html-bibr">83</a>], (<b>e</b>) Nb-Y-3Ga [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>]; and (<b>f</b>) Nb-Y-Ce [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>] for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. A = A-type granite; A<sub>1</sub> = A<sub>1</sub>-type granite; A<sub>2</sub> = A<sub>2</sub>-type granite; I = I-type granite; S = S-type granite; FG = Fractionated felsic granite; OGT = Unfractionated M-, I- and S-type granite.</p>
Full article ">Figure 8 Cont.
<p>Plots of (<b>a</b>) La/Sm versus La, (<b>b</b>) Zr/Hf versus SiO<sub>2</sub>, (<b>c</b>) FeO<sup>T</sup>/MgO versus (Zr + Y + Nb + Ce), (<b>d</b>) Zr versus 10,000 Ga/Al) [<a href="#B83-minerals-14-00933" class="html-bibr">83</a>], (<b>e</b>) Nb-Y-3Ga [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>]; and (<b>f</b>) Nb-Y-Ce [<a href="#B73-minerals-14-00933" class="html-bibr">73</a>] for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. A = A-type granite; A<sub>1</sub> = A<sub>1</sub>-type granite; A<sub>2</sub> = A<sub>2</sub>-type granite; I = I-type granite; S = S-type granite; FG = Fractionated felsic granite; OGT = Unfractionated M-, I- and S-type granite.</p>
Full article ">Figure 9
<p>Plots of (<b>a</b>) Al<sub>2</sub>O<sub>3</sub>/(Fe<sub>2</sub>O<sub>3</sub><sup>T</sup> + MgO + TiO<sub>2</sub>) versus (Al<sub>2</sub>O<sub>3</sub> + Fe<sub>2</sub>O<sub>3</sub><sup>T</sup> + MgO + TiO<sub>2</sub>) (after Patiňo Douce. (1999) [<a href="#B84-minerals-14-00933" class="html-bibr">84</a>]); (<b>b</b>) (Na<sub>2</sub>O + K<sub>2</sub>O) versus (FeO<sup>T</sup> + MgO + TiO<sub>2</sub>); (<b>c</b>) Mg<sup>#</sup> versus SiO<sub>2</sub>; and (<b>d</b>) ε<sub>Hf</sub>(t) versus U–Pb ages for monzogranite and syenogranite from the Yingpanshan–Damanbie iREE deposit. (<b>b</b>) Compositional fields of experimental melts are from Patiño Douce [<a href="#B84-minerals-14-00933" class="html-bibr">84</a>], Sylvester [<a href="#B85-minerals-14-00933" class="html-bibr">85</a>], Patiño Douce [<a href="#B84-minerals-14-00933" class="html-bibr">84</a>], and Altherr et al. [<a href="#B86-minerals-14-00933" class="html-bibr">86</a>], respectively; (<b>c</b>) fields shown are as follows: pure crustal partial melts obtained in experimental studies by the dehydration melting of low-K basaltic rocks at 8–16 kbar and 1000–1050 °C [<a href="#B87-minerals-14-00933" class="html-bibr">87</a>]; pure crustal melts obtained in experimental studies by the moderately hydrous (1.7–2.3 wt.% H<sub>2</sub>O) melting of medium- to high-K basaltic rocks at 7 kbar and 825–950 °C [<a href="#B88-minerals-14-00933" class="html-bibr">88</a>]; mantle melts (basalts) and Quaternary volcanic rocks from the Andean southern volcanic zone [<a href="#B89-minerals-14-00933" class="html-bibr">89</a>]; melts from meta-igneous sources under crustal pressure and temperature conditions of 0.5–1.5 GPa and 800–1000 °C, respectively, which are based on the work completed by Wolf and Wyllie [<a href="#B90-minerals-14-00933" class="html-bibr">90</a>]; (<b>d</b>) data for the Gangdese belt from Ji et al. [<a href="#B91-minerals-14-00933" class="html-bibr">91</a>]; data for the southern Lhasa block from Jiang et al. [<a href="#B92-minerals-14-00933" class="html-bibr">92</a>], Ji et al. [<a href="#B93-minerals-14-00933" class="html-bibr">93</a>], Hou et al. [<a href="#B94-minerals-14-00933" class="html-bibr">94</a>], Zheng et al. [<a href="#B95-minerals-14-00933" class="html-bibr">95</a>], Zhu et al. [<a href="#B96-minerals-14-00933" class="html-bibr">96</a>], and Huang et al. [<a href="#B97-minerals-14-00933" class="html-bibr">97</a>]; data for the central Lhasa block from Hou et al. [<a href="#B94-minerals-14-00933" class="html-bibr">94</a>], Gao et al. [<a href="#B98-minerals-14-00933" class="html-bibr">98</a>], Zheng et al. [<a href="#B99-minerals-14-00933" class="html-bibr">99</a>], and Wang et al. [<a href="#B100-minerals-14-00933" class="html-bibr">100</a>]; data for the eastern Himalayan syntaxis from Chui et al. [<a href="#B101-minerals-14-00933" class="html-bibr">101</a>], Gou et al. [<a href="#B102-minerals-14-00933" class="html-bibr">102</a>], and Pan et al. [<a href="#B103-minerals-14-00933" class="html-bibr">103</a>]; and data for the Tengchong block including the Guyong area from Xu et al. [<a href="#B18-minerals-14-00933" class="html-bibr">18</a>], Xie et al. [<a href="#B42-minerals-14-00933" class="html-bibr">42</a>], Chen et al. [<a href="#B51-minerals-14-00933" class="html-bibr">51</a>], and Qi et al. [<a href="#B104-minerals-14-00933" class="html-bibr">104</a>].</p>
Full article ">
14 pages, 2765 KiB  
Article
Evaluation of Reservoir Damage by Hydrothermal Fluid from Clay Metamorphism, Particle Migration, and Heavy-Component Deposition in Offshore Heavy Oilfields
by Zuhao Zheng, Lu Zhang, Hongchao Yin, Dong Liu, Wei He, Leilei Shui, Ning Wang, Hao Chen, Shenglai Yang and Yiqi Zhang
Processes 2024, 12(9), 1959; https://doi.org/10.3390/pr12091959 - 12 Sep 2024
Viewed by 578
Abstract
Marine heavy-oil reserves are enormous, and thermal recovery technology is one of the most effective ways to utilize them. However, steam as a high-energy external fluid will affect the geological characteristics of the reservoir. In this paper, the sensitivity of the reservoir was [...] Read more.
Marine heavy-oil reserves are enormous, and thermal recovery technology is one of the most effective ways to utilize them. However, steam as a high-energy external fluid will affect the geological characteristics of the reservoir. In this paper, the sensitivity of the reservoir was analyzed in terms of the high-temperature metamorphic characteristics of clay minerals and the coupling damage of particle migration and heavy component deposition. Firstly, long-core cyclic steam stimulation experiments were conducted using supersaturated steam, saturated steam, and superheated steam to quantify the differences in oil recovery capabilities. Subsequently, the variation characteristics of clay components in the core under different steam temperatures were analyzed by X-ray diffraction spectroscopy. Finally, the influence of particle migration and heavy-component deposition on reservoir permeability was clarified through displacement experiments. The results show that the recovery of superheated steam is more than 12% higher than that of supersaturated steam, and the throughput cycle is effectively shortened. In the laboratory, only the clay metamorphism due to superheated steam was more effective, and the metamorphism was mainly concentrated in kaolinite and monazite. Particle migration causes little damage to the reservoir, but the formation of particle migration coupled with heavy-component deposition can lead to more than 30% of the reservoir becoming damaged. Full article
(This article belongs to the Section Energy Systems)
Show Figures

Figure 1

Figure 1
<p>The experimental process of long-core steam-cycle stimulation.</p>
Full article ">Figure 2
<p>Comparison of cumulative recovery situation.</p>
Full article ">Figure 3
<p>Comparison of recovery conditions in different rounds.</p>
Full article ">Figure 4
<p>Determination results of clay components: (<b>a</b>) changes in clay components after different steam-cycle stimulation experiments; (<b>b</b>) the change rate of clay components after different steam-cycle stimulation experiments.</p>
Full article ">Figure 5
<p>Speed sensitivity experiment results.</p>
Full article ">Figure 6
<p>The impact of sedimentation and migration coupling on permeability.</p>
Full article ">Figure 7
<p>The decrease in permeability caused by particles.</p>
Full article ">Figure 8
<p>The decrease in permeability caused by particles and heavy components, respectively.</p>
Full article ">
13 pages, 1700 KiB  
Article
Investigation of the Flotation of an Ore Containing Bastnaesite and Monazite: Kinetic Study and Process Flowsheet Simulation
by Claude Bazin and Jean-François Boulanger
Minerals 2024, 14(9), 906; https://doi.org/10.3390/min14090906 - 4 Sep 2024
Cited by 1 | Viewed by 773
Abstract
Laboratory flotation tests carried out using an ore sample containing Rare Earth Elements (REEs) present as monazite and bastnaesite show that the flotation of monazite is slower and yielded lower recovery than that of bastnaesite. Results show that when studying the performances of [...] Read more.
Laboratory flotation tests carried out using an ore sample containing Rare Earth Elements (REEs) present as monazite and bastnaesite show that the flotation of monazite is slower and yielded lower recovery than that of bastnaesite. Results show that when studying the performances of a concentration process for an REE ore, it is essential to not look only at the behavior of the individual REEs but to convert elemental assays into mineral assays to obtain the mineral’s actual response to the concentration process. The results of the laboratory flotation tests are used to calibrate a flotation simulator applied to study different circuit configurations for the concentration of the REE minerals. Indeed, it is shown that for the studied ore, two cleaning stages of a rougher concentrate are sufficient to produce a concentrate with a Total Rare Earth Oxide (TREO) grade above 40%, which is acceptable for the subsequent hydrometallurgical process. The simulation also shows that it may be feasible, if required for the hydrometallurgy step, to separate bastnaesite and monazite by taking advantage of the different flotation kinetics of the two minerals. Full article
Show Figures

Figure 1

Figure 1
<p>Variation of the ground ore D<sub>80</sub> as a function of grinding time in the laboratory rod mill (dashed line shows the 12 min grinding time established).</p>
Full article ">Figure 2
<p>Ore and element size distributions in the ground product.</p>
Full article ">Figure 3
<p>Flotation test procedure.</p>
Full article ">Figure 4
<p>Elements cumulated recovery as a function of time.</p>
Full article ">Figure 5
<p>Mineral cumulated recoveries as a function of time.</p>
Full article ">Figure 6
<p>Recovery of gangue minerals as a function of time.</p>
Full article ">Figure 7
<p>Simulation results of open and closed flotation circuit configurations for processing the studied REE ore.</p>
Full article ">Figure 8
<p>Simulated intermediate products in a bastnaesite/monazite separation circuit.</p>
Full article ">
20 pages, 5613 KiB  
Article
Alteration and Non-Formula Elements Uptake of Zircon from Um Ara Granite, South Eastern Desert, Egypt
by Hamdy H. Abd El-Naby
Minerals 2024, 14(8), 834; https://doi.org/10.3390/min14080834 - 17 Aug 2024
Viewed by 745
Abstract
The Um Ara granites are a suite of granitoid rocks located in the southern part of the Eastern Desert of Egypt. The integration of various electron probe micro analyzer (EPMA) techniques, such as backscattered electron (BSE) imaging, X-ray compositional mapping, and wavelength dispersive [...] Read more.
The Um Ara granites are a suite of granitoid rocks located in the southern part of the Eastern Desert of Egypt. The integration of various electron probe micro analyzer (EPMA) techniques, such as backscattered electron (BSE) imaging, X-ray compositional mapping, and wavelength dispersive spectrometry (WDS), has provided valuable insights into the alteration process of zircon in the Um Ara granite. The zircon exhibits high concentrations of non-formula elements such as P, Al, Ca, Fe, Ti, and REEs, suggesting that the alteration involved coupled dissolution-reprecipitation processes influenced by aqueous fluids. The negative correlations between Zr and the non-formula elements indicate that these elements were incorporated into zircon at the expense of Zr and Si, significantly affecting the distribution and fractionation of REEs in the original zircon. Based on the presented data and literature knowledge, the sequence of alteration events is proposed as follows: (1) initial zircon crystallization around 603 Ma accompanied by the formation of other U- and Th-bearing minerals like xenotime, thorite, monazite, and apatite; (2) long-term metamictization leading to fractures and cracks that facilitated fluid circulation and chemical changes; (3) a major hydrothermal event around 20 Ma that released a suite of non-formula elements from the metamicted zircon and associated minerals, with the enriched hydrothermal fluids subsequently incorporating these elements into the modified zircon structure; and (4) further low-temperature alteration during subsequent pluvial periods (around 50,000–159,000 years ago), facilitated by the shear zones in the Um Ara granites, may have allowed further uptake of non-formula elements. The interplay between hydrothermal fluids, meteoric water, and the shear zone environments appears to have been a key driver for the uptake of non-formula elements into the altered zircon. Full article
(This article belongs to the Section Mineral Geochemistry and Geochronology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geological map of Um Ara area (modified from [<a href="#B14-minerals-14-00834" class="html-bibr">14</a>]).</p>
Full article ">Figure 2
<p>BSE imaging showing morphological characteristics of zircon grains separated from Um Ara granites, south Eastern Desert of Egypt. (<b>a</b>,<b>b</b>) show unzoned zircon grains that exhibit a porous texture. (<b>c</b>) Zoned zircon grains containing apatite inclusions.</p>
Full article ">Figure 3
<p>(<b>a</b>) BSE showing zoning in zircon. It contains inclusions of apatite, uranothorite, and many micropores (dark areas). (<b>b</b>–<b>g</b>) peak intensity mappings showing the distribution of Zr, Th, U, Hf, Al, and Ca in the grain of image (<b>a</b>). Outer rims show a higher concentration of Th, U, Hf, and Al and a lower concentration of Zr.</p>
Full article ">Figure 4
<p>(<b>a</b>) BSE showing inclusions of apatite, uranothorite, uranophane, and many micropores (dark areas). (<b>b</b>–<b>g</b>) peak intensity mappings showing the distribution of Zr, U, Th, Hf, Ca, and Ali in the grain of image (<b>a</b>).</p>
Full article ">Figure 5
<p>Analyses of the different sectors along the lines X–Y in <a href="#minerals-14-00834-f002" class="html-fig">Figure 2</a>c reveal variations in SiO<sub>2</sub>, ZrO<sub>2</sub>, HfO<sub>2</sub>, UO<sub>2</sub>+ThO<sub>2</sub>, Al<sub>2</sub>O<sub>3</sub>, FeO, CaO, and REE<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 6
<p>(<b>a</b>) Binary plot of total major elements (wt%) versus SiO<sub>2</sub> and ZrO<sub>2</sub> for the Um Ara zircon. Theoretical endmember pure zircon indicated by the orange asterisk (where total major elements = 100 wt%, SiO<sub>2</sub> = 32.8 wt%, and ZrO<sub>2</sub> = 67.2 wt%, [<a href="#B24-minerals-14-00834" class="html-bibr">24</a>]). (<b>b</b>) ThO<sub>2</sub> vs. UO<sub>2</sub> diagram that reveals a wide range of Th/U ratios for the Um Ara zircon.</p>
Full article ">Figure 7
<p>Rare earth element patterns of Um Ara-altered zircon, normalized to the C1 chondrite values of [<a href="#B37-minerals-14-00834" class="html-bibr">37</a>]. The dashed black line represents the mean chondrite pattern of Um Ara-altered zircon. The chondrite pattern of unaltered magmatic zircon is shown for comparison.</p>
Full article ">Figure 8
<p>Chondrite-normalized Sm/La ratio vs. La (ppm) discrimination diagram (after [<a href="#B32-minerals-14-00834" class="html-bibr">32</a>,<a href="#B38-minerals-14-00834" class="html-bibr">38</a>,<a href="#B39-minerals-14-00834" class="html-bibr">39</a>]) with zircons from granitoids of the Um Ara area (red circles). The positioning of the studied zircon grains, along with their elevated La contents and lower (Sm/La)N ratios, indicates that the primary zircon has been subjected to fluid-driven alteration processes.</p>
Full article ">Figure 9
<p>Dominant simple and coupled substitution in zircon from the Um Ara granite: (<b>a</b>) Zr<sup>4+</sup> vs. Th<sup>4+</sup>; (<b>b</b>) Zr<sup>4+</sup><sub>2</sub> vs. U<sup>4+</sup>; (<b>c</b>) Zr<sup>4+</sup> vs. Hf<sup>4+</sup>; (<b>d</b>) (Y, REE)<sup>3+</sup> + P<sup>5+</sup> vs. Zr<sup>4+</sup>+Si<sup>4+</sup>; (<b>e</b>) (Al, Fe)<sup>3+</sup> + 4(Y, REE)<sup>3+</sup> + P<sup>5+</sup> vs. 4Zr<sup>4+</sup>+Si<sup>4+</sup>; (<b>f</b>) (Mg, Fe)<sup>2+</sup> + 3(Y, REE)<sup>3+</sup> + P<sup>5+</sup> vs. 3Zr<sup>4+</sup>+Si<sup>4+</sup>. Negative correlations in these diagrams indicate that Th, U, Hf, Al, Fe, Mg, P, Y, and REE were incorporated in the parental zircon at the expense of Zr and Si, leading to the non-formula elements uptake of the Um Ara-altered zircon.</p>
Full article ">
22 pages, 5061 KiB  
Article
A Protocol for Electron Probe Microanalysis (EPMA) of Monazite for Chemical Th-U-Pb Age Dating
by Bernhard Schulz, Joachim Krause and Wolfgang Dörr
Minerals 2024, 14(8), 817; https://doi.org/10.3390/min14080817 - 12 Aug 2024
Cited by 1 | Viewed by 1132
Abstract
A protocol for the monazite (LREE,Y,Th,U,Si,Ca)PO4 in situ Th-U-Pb dating by electron probe microanalyser (EPMA) involves a suitable reference monazite. Ages of several potential reference monazites were determined by TIMS-U-Pb isotope analysis. The EPMA protocol is based on calibration with REE-orthophosphates and [...] Read more.
A protocol for the monazite (LREE,Y,Th,U,Si,Ca)PO4 in situ Th-U-Pb dating by electron probe microanalyser (EPMA) involves a suitable reference monazite. Ages of several potential reference monazites were determined by TIMS-U-Pb isotope analysis. The EPMA protocol is based on calibration with REE-orthophosphates and a homogeneous Th-rich reference monazite at beam conditions of 20 kV, 50 nA, and 5 µm for best possible matrix matches and avoidance of dead time bias. EPMA measurement of samples and repeated analysis of the reference monazite are performed at beam conditions of 20 kV, 100 nA, and 5 µm. Analysis of Pb and U on a PETL crystal requires YLg-on-PbMa and ThMz-on-UMb interference corrections. Offline re-calibration of the Th calibration on the Th-rich reference monazite, to match its nominal age, is an essential part of the protocol. EPMA-Th-U-Pb data are checked in ThO2*-PbO coordinates for matching isochrones along regressions forced through zero. Error calculations of monazite age populations are performed by weighted average routines. Depending on the number of analyses and spread in ThO2*-PbO coordinates, minimum errors <10 Ma are possible and realistic for Paleozoic monazite ages. A test of the protocol was performed on two garnet metapelite samples from the Paleozoic metamorphic Zone of Erbendorf-Vohenstrauß (NE-Bavaria, western Bohemian Massif). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Th-U-Pb isochrones and chemical model ages of various reference monazites. (<b>a</b>) Total ThO<sub>2</sub>* vs. PbO (wt%) isochrons. ThO<sub>2</sub>* is ThO<sub>2</sub> + UO<sub>2</sub> equivalents expressed as ThO<sub>2</sub>. (<b>b</b>) Isochrones and weighted average ages in Th<sub>SUZ</sub> vs. Pb [<a href="#B20-minerals-14-00817" class="html-bibr">20</a>]. Regression lines with the coefficient of determination R<sup>2</sup> are forced through zero [<a href="#B15-minerals-14-00817" class="html-bibr">15</a>,<a href="#B20-minerals-14-00817" class="html-bibr">20</a>]. Weighted average ages in Ma and minimal error of 2σ are calculated from the single analyses belonging to an isochrone. Reference monazites with homogeneous compositions and with data along isochrons can be distinguished.</p>
Full article ">Figure 2
<p>Monazite U-Pb concordia diagrams, with insets showing the concordia ages at 95% confidence with MSWD and probability of concordance. Error ellipses of 2σ of single analyses. Madmon (<b>a</b>), Steenkampskraal (<b>b</b>), and Tsaratana (<b>c</b>) are reference monazites in <a href="#minerals-14-00817-f001" class="html-fig">Figure 1</a> and <a href="#minerals-14-00817-t002" class="html-table">Table 2</a> and <a href="#minerals-14-00817-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 3
<p>Spectrometer linescans on Madmon reference monazites at 20 kV acceleration voltage, 100 nA beam current, 5 µm beam diameter, and 300 ms dwell time with step size 25 with a PETL crystal. Spectrometer position (in mm) vs. intensity (in counts per second). (<b>a</b>) Peaks of PbMa and PbMb and potential background (BKG) positions. (<b>b</b>) Peaks of UMa and UMb and potential background (BKG) positions.</p>
Full article ">Figure 4
<p>Combined spectrometer linescans on Madmon reference monazite at 20 kV acceleration voltage, 100 nA beam current, 5 µm beam diameter, and 300 ms dwell time with step size 25 with a PETL crystal. Spectrometer position (in mm) vs. intensity (in counts per second). (<b>a</b>) Peak PbMa displays interference with the positive slope of YLg of Y-garnet and no interference with ThMz metal. (<b>b</b>) Peak of UMb shows interference with the positive slope of ThMg metal.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) ThO<sub>2</sub>* vs. PbO (wt%) isochrone diagrams of monazite analyses from garnet micaschists in the Saxothuringian Zone of Erbendorf-Vohenstrauß (ZEV). ThO<sub>2</sub>* is ThO<sub>2</sub> + UO<sub>2</sub> equivalents expressed as ThO<sub>2</sub>. Regression lines (isochrones) with the coefficient of determination R<sup>2</sup> are forced through zero [<a href="#B15-minerals-14-00817" class="html-bibr">15</a>,<a href="#B20-minerals-14-00817" class="html-bibr">20</a>]. Weighted average ages in Ma with MSWD and minimal error of 2σ are calculated from the single analyses belonging to an isochrone according to [<a href="#B69-minerals-14-00817" class="html-bibr">69</a>]. Analyses from reference monazite Madmon are added. (<b>c</b>,<b>d</b>) Monazite ages for the main age population calculated by the zircon age extractor routine according to [<a href="#B69-minerals-14-00817" class="html-bibr">69</a>]. Note that there is an age difference between both methods applied to sample ZEV-10PUL.</p>
Full article ">Figure 6
<p>(<b>a</b>) Mineral chemistry of monazite in garnet micaschists from the Zone of Erbendorf-Vohenstrauß ZEV in age vs. Y<sub>2</sub>O<sub>3</sub> coordinates. Trend toward higher Y<sub>2</sub>O<sub>3</sub> in the younger monazites. (<b>b</b>) Similar Th+U vs. Ca (per formula unit p.f.u.) along the cheralite substitution trend in the monazites. (<b>c</b>,<b>d</b>) Garnet core (c) to rim (r) zonations in almandine (Alm-50%, due to scale), pyrope (Prp), grossular (Grs), and spessartine (Sps) components (in Mol %, calculated from mole fraction × 100). (<b>e</b>) Different <span class="html-italic">X</span>Mg vs. <span class="html-italic">X</span>Ca core (c) to rim (r) zonation trends of the garnets in assemblages with biotite, muscovite, plagioclase, quartz, kyanite, and sillimanite. Numbers mark garnet analyses used for thermobarometry. Sample ZEV-1349 displays first prograde then retrograde zonation trend and sample ZEV-10PUL a retrograde zonation trend.</p>
Full article ">Figure 7
<p>P-T estimates and core-to-rim P-T trends for garnet mineral assemblages in coloured marks in reference to the monazite stability field. Numbers refer to garnet analyses in <a href="#minerals-14-00817-f006" class="html-fig">Figure 6</a>c–e and <a href="#minerals-14-00817-t006" class="html-table">Table 6</a>; error of ± 50 °C/1.0 kbar on P-T estimates by cation exchange and net transfer GBMP geothermobarometers [<a href="#B92-minerals-14-00817" class="html-bibr">92</a>]. The aluminosilicates (Ky, Sill), muscovite-out (Ms−), and cordierite-in (Cd+) univariant lines are after [<a href="#B93-minerals-14-00817" class="html-bibr">93</a>]. Stability fields of monazite (Mnz) and allanite (Aln) at different bulk rock contents as a function of Ca wt% and with xenotime (Xtm) stability field [<a href="#B94-minerals-14-00817" class="html-bibr">94</a>,<a href="#B95-minerals-14-00817" class="html-bibr">95</a>]. Interpretations of monazite ages in reference to P-T path: (<b>a</b>) Early 461 Ma monazite crystallisation after Ordovician metamorphism, followed by Devonian ~390–370 Ma garnet crystallisation. A Carboniferous 352–343 Ma monazite crystallisation appears as a separate event. (<b>b</b>) Alternative interpretation with Carboniferous monazite crystallisation at higher temperatures, following the Devonian garnet crystallisation.</p>
Full article ">
22 pages, 8347 KiB  
Article
Geochronology, Geochemistry, and In Situ Sr-Nd-Hf Isotopic Compositions of a Tourmaline-Bearing Leucogranite in Eastern Tethyan Himalaya: Implications for Tectonic Setting and Rare Metal Mineralization
by Yangchen Drolma, Kaijun Li, Yubin Li, Jinshu Zhang, Chengye Yang, Gen Zhang, Ruoming Li and Duo Liu
Minerals 2024, 14(8), 755; https://doi.org/10.3390/min14080755 - 26 Jul 2024
Viewed by 854
Abstract
Himalayan leucogranite is an excellent target for understanding the orogenic process of the India–Asia collision, but its origin and tectonic significance are still under debate. An integrated study of geochronology, geochemistry, and in situ Sr-Nd-Hf isotopes was conducted for a tourmaline-bearing leucogranite in [...] Read more.
Himalayan leucogranite is an excellent target for understanding the orogenic process of the India–Asia collision, but its origin and tectonic significance are still under debate. An integrated study of geochronology, geochemistry, and in situ Sr-Nd-Hf isotopes was conducted for a tourmaline-bearing leucogranite in the eastern Tethyan Himalaya using LA-ICP-MS, X-ray fluorescence spectroscopy, and ICP-MS and LA-MC-ICP-MS, respectively. LA-ICP-MS U-Pb dating of zircon and monazite showed that it was emplaced at ~19 Ma. The leucogranite had high SiO2 and Al2O3 contents ranging from 73.16 to 73.99 wt.% and 15.05 to 15.24 wt.%, respectively. It was characterized by a high aluminum saturation index (1.14–1.19) and Rb/Sr ratio (3.58–6.35), which is characteristic of S-type granite. The leucogranite was enriched in light rare-earth elements (LREEs; e.g., La and Ce) and large ion lithophile elements (LILEs; e.g., Rb, K, and Pb) and depleted in heavy rare-earth elements (e.g., Tm, Yb, and Lu) and high field strength elements (HFSEs; e.g., Nb, Zr, and Ti). It was characterized by high I Sr (t) (0.7268–0.7281) and low ε Nd (t) (−14.6 to −13.2) and ε Hf (t) (−12.6 to −9.47), which was consistent with the isotopic characteristics of the Higher Himalayan Sequence. Petrogenetically, the origin of the leucogranite is best explained by the decompression-induced muscovite dehydration melting of an ancient metapelitic source within the Higher Himalayan Sequence during regional extension due to the movement of the South Tibetan Detachment System (STDS). The significantly high lithium and beryllium contents of the leucogranite and associated pegmatite suggest that Himalayan leucogranites possess huge potential for lithium and beryllium exploration. Full article
Show Figures

Figure 1

Figure 1
<p>Geological sketch map of the Himalayas showing the distribution of Himalayan leucogranites (after [<a href="#B5-minerals-14-00755" class="html-bibr">5</a>]).</p>
Full article ">Figure 2
<p>Simplified geological map of the Luozha tourmaline-bearing leucogranite (after [<a href="#B25-minerals-14-00755" class="html-bibr">25</a>]). Mineral abbreviations [<a href="#B26-minerals-14-00755" class="html-bibr">26</a>]: And, andalusite; Grt, garnet; St, staurolite.</p>
Full article ">Figure 3
<p>Representative field photographs and photomicrographs of the LTLG and spodumene-bearing pegmatites. (<b>a</b>) Field photograph showing oriented tourmalines of the LTLG; (<b>b</b>) Photomicrograph of the LTLG; (<b>c</b>) Field photograph of the spodumene-bearing pegmatite and (<b>d</b>) Photomicrograph of the spodumene-bearing pegmatite. Mineral abbreviations [<a href="#B26-minerals-14-00755" class="html-bibr">26</a>]: Bt, biotite; Kfs, K-feldspar; Ms, muscovite; Pl, plagioclase; Qz, quartz; Spd, Spodumene; Tur, tourmaline.</p>
Full article ">Figure 4
<p>U-Pb dating results of the LTLG. (<b>a</b>) Cathodoluminescence images for representative zircons from the LTLG; (<b>b</b>) U-Pb zircon concordia diagram of the LTLG; Tera–Wasserburg concordia diagram for zircons (<b>c</b>) and monazites (<b>d</b>) of the LTLG. The red circle indicate the location of U-Pb dating analysis.</p>
Full article ">Figure 5
<p>Plots of (<b>a</b>) SiO<sub>2</sub> vs. (Na<sub>2</sub>O+K<sub>2</sub>O) (after [<a href="#B39-minerals-14-00755" class="html-bibr">39</a>]), (<b>b</b>) SiO<sub>2</sub> vs. (Na<sub>2</sub>O+K<sub>2</sub>O-CaO) (after [<a href="#B40-minerals-14-00755" class="html-bibr">40</a>]), (<b>c</b>) SiO<sub>2</sub> vs. K<sub>2</sub>O (after [<a href="#B41-minerals-14-00755" class="html-bibr">41</a>]); and (<b>d</b>) A/CNK vs. A/NK (after [<a href="#B42-minerals-14-00755" class="html-bibr">42</a>]) for the LTLG.</p>
Full article ">Figure 6
<p>(<b>a</b>) REE patterns and (<b>b</b>) Spidergrams of the LTLG. The values of chondrite and primitive mantle are from McDonough and Sun [<a href="#B43-minerals-14-00755" class="html-bibr">43</a>]. The data of S-type (blue field) and highly fractional (green field) leucogranites from the Ramba area are from Liu et al. [<a href="#B12-minerals-14-00755" class="html-bibr">12</a>].</p>
Full article ">Figure 7
<p>Plots of (<b>a</b>) in situ and whole rock Sr-Nd isotopic data and (<b>b</b>) Zircon Hf isotopic data of the LTLG.</p>
Full article ">Figure 8
<p>Diagrams of (<b>a</b>) (Zr + Nb + Ce + Y) vs. FeO*/MgO (after [<a href="#B48-minerals-14-00755" class="html-bibr">48</a>]); (<b>b</b>) (Zr + Nb + Ce + Y) vs. (Na<sub>2</sub>O + K<sub>2</sub>O)/CaO (after [<a href="#B48-minerals-14-00755" class="html-bibr">48</a>]); (<b>c</b>) Rb vs. Th and (<b>d</b>) Rb vs. Y (after [<a href="#B50-minerals-14-00755" class="html-bibr">50</a>]) for the LTLG. The data of S-type granites from the Interview River Suite are from Chappell [<a href="#B45-minerals-14-00755" class="html-bibr">45</a>]. The data of S-type leucogranites from the Ramba area are from Liu et al. [<a href="#B12-minerals-14-00755" class="html-bibr">12</a>].</p>
Full article ">Figure 9
<p>Diagrams of (<b>a</b>) Nb/Ta vs. Zr/Hf and (<b>b</b>) Rb/Sr vs. (La/Yb)<sub>N</sub> for the LTLG. The data of S-type and highly fractional leucogranites from the Ramba area are from Liu et al. [<a href="#B12-minerals-14-00755" class="html-bibr">12</a>].</p>
Full article ">Figure 10
<p>Plots of (<b>a</b>) (Na<sub>2</sub>O + K<sub>2</sub>O + TiO<sub>2</sub> + TFeO + MgO) vs. (Na<sub>2</sub>O + K<sub>2</sub>O)/(TiO<sub>2</sub> + TFeO + MgO) (after [<a href="#B60-minerals-14-00755" class="html-bibr">60</a>]), (<b>b</b>) (CaO + TiO<sub>2</sub> + TFeO + MgO) vs. CaO/(TiO<sub>2</sub> + TFeO + MgO) (after [<a href="#B60-minerals-14-00755" class="html-bibr">60</a>]), (<b>c</b>) Al<sub>2</sub>O<sub>3</sub>/TiO<sub>2</sub> vs. CaO/TiO<sub>2</sub> (after [<a href="#B61-minerals-14-00755" class="html-bibr">61</a>]); and (<b>d</b>) Rb/Sr vs. Rb/Ba (after [<a href="#B61-minerals-14-00755" class="html-bibr">61</a>]) for the LTLG. MP, metapelites; MGW, metagreywackes; AMP, amphibolites.</p>
Full article ">Figure 11
<p>Plots of (<b>a</b>) Ba vs. Rb/Sr and (<b>b</b>) Sr vs. Rb/Sr (after [<a href="#B62-minerals-14-00755" class="html-bibr">62</a>]).</p>
Full article ">Figure 12
<p>Plots of <b>ε<sub>Nd</sub></b>(t) vs. <b>I<sub>Sr</sub></b>(t) for the LTLG. Fields of Gangdese batholith, Higher Himalayan Sequence, and Lesser Himalayan Sequence are from Wu et al. [<a href="#B5-minerals-14-00755" class="html-bibr">5</a>].</p>
Full article ">Figure 13
<p>Plots of (<b>a</b>) Rb/Sr vs. Li; (<b>b</b>) Rb/Sr vs. Be; (<b>c</b>) Zr/Hf vs. Li; (<b>d</b>) Zr/Hf vs. Be; (<b>e</b>) Nb/Ta vs. Li; and (<b>f</b>) Nb/Ta vs. Be for the LTLG.</p>
Full article ">
14 pages, 794 KiB  
Article
Removal and Recovery of Europium with a New Functionalized Mesoporous Silica-Based Adsorbent
by Charith Fonseka, Seongchul Ryu, Jaya Kandasamy, Harsha Ratnaweera and Saravanamuthu Vigneswaran
Sustainability 2024, 16(13), 5636; https://doi.org/10.3390/su16135636 - 30 Jun 2024
Viewed by 1266
Abstract
The discharge of acid mine drainage (AMD), characterized by a high concentration of rare earth elements (REEs), poses a significant threat to the health of ecosystems surrounding water sources. The global market demand for REEs has experienced a notable surge in the past [...] Read more.
The discharge of acid mine drainage (AMD), characterized by a high concentration of rare earth elements (REEs), poses a significant threat to the health of ecosystems surrounding water sources. The global market demand for REEs has experienced a notable surge in the past decade. Consequently, recovering REEs from waste streams like AMD not only benefits the environment but also offers financial advantages. Europium (Eu), the rarest among REEs, constitutes only 0.1% w/w in monazite and bastnaesite ores. Eu is extensively used in the production of phosphors, alloys, and additives, and is a critical raw material for developing smart devices, ranging from high-resolution color screens to circuitry. Traditional adsorbents typically exhibit limited selectivity towards REE recovery. Mesoporous silica materials, such as SBA15 (Santa Barbara Amorphous-15), provide excellent tunability and modification capabilities, making them an attractive and cost-effective alternative. This research focused on two key aspects: (i) evaluating the dynamic adsorption column performance of granulated SBA15–NH–PMIDA to preferentially recover Eu, and (ii) employing mathematical modeling to optimize the dynamic adsorption column’s operating conditions for real-world applications with a minimal number of experimental runs. Granulated SBA15–NH–PMIDA was chosen as the adsorbent due to its high adsorptive capacity and selectivity in capturing Eu. The study revealed that granulated SBA15–NH–PMIDA exhibited 57.47 mg/g adsorption capacity and an 81% selectivity towards Eu. Furthermore, SBA15–NH–PMIDA demonstrated preferential adsorption toward Eu in complex multi-component solutions, such as AMD. The linear driven force approximation model (LDFAM) provided an acceptable simulation (R2 > 0.91) under varying operational conditions. This validates the use of the model as a tool to effectively simulate and optimize column experiments that used granulated SBA15–NH–PMIDA to recover Eu. Full article
(This article belongs to the Section Environmental Sustainability and Applications)
Show Figures

Figure 1

Figure 1
<p>SEM images of (<b>a</b>) SBA15–NH–PMIDA powder and (<b>b</b>) SBA15–NH–PMIDA granule.</p>
Full article ">Figure 2
<p>Validation of single solute Eu breakthrough curve for granular SBA15–NH–PMIDA with LDFAM. (Bed height = 0.1 m, linear flow rate = 0.637 m/h, R<sup>2</sup> = 0.95).</p>
Full article ">Figure 3
<p>Validation of simulation LDFAM for inlet concentration of 3 mg/L (Bed Height = 0.1 m, velocity = 0.637 m/h, R<sup>2</sup> = 0.92).</p>
Full article ">Figure 4
<p>Validation of simulation LDFAM for linear filtration of 0.955 m/h (Inlet concentration = 3 mg/L, bed height = 0.1 m, R<sup>2</sup> = 0.94).</p>
Full article ">Figure 5
<p>Validation of simulation LDFAM for adsorbent medium depth of 0.125 m (inlet Eu concentration = 3 mg/L, linear filtration rate = 0.637 m/h, R<sup>2</sup> = 0.91).</p>
Full article ">
20 pages, 23085 KiB  
Article
Origin of the Kunduleng Granite and Its Associated Uranium Anomaly in the Southern Great Xing’an Range, NE China
by Jiaxing Sun, Deyou Sun, Jun Gou, Dongguang Yang, Changdong Wang, Li Tian and Duo Zhang
Minerals 2024, 14(7), 666; https://doi.org/10.3390/min14070666 - 27 Jun 2024
Viewed by 730
Abstract
The Kunduleng granite hosts one of several significant uranium anomalies within the southern Great Xing’an Range, NE China. Whole-rock geochemistry and mineral chemistry data, along with the zircon U-Pb-Hf isotope have been used to constrain the petrogenesis of this granitic intrusion and the [...] Read more.
The Kunduleng granite hosts one of several significant uranium anomalies within the southern Great Xing’an Range, NE China. Whole-rock geochemistry and mineral chemistry data, along with the zircon U-Pb-Hf isotope have been used to constrain the petrogenesis of this granitic intrusion and the origin of the uranium anomaly. Microscopically, quartz, alkali-feldspar, and plagioclase are the essential mineral constituents of the granite, with minor biotite, while monazite, apatite, xenotime, and zircon are accessory minerals. Geochemically, the silica- and alkali-rich granites show a highly fractionated character with “seagull-shaped” REE patterns and significant negative anomalies of Ba and Sr, along with low Zr/Hf and Nb/Ta ratios. The granite has positive zircon εHf(t) values ranging from +12.7 to +14.5 and crustal model ages (TDM2) of 259–376 Ma, indicating a Paleozoic juvenile crustal source. Uraninite and brannerite are the main radioactive minerals responsible for the uranium anomaly within the Kunduleng granite. Uraninite presents well-developed cubic crystals and occurs as tiny inclusions in quartz and K-feldspar with magmatic characteristics (e.g., elevated ThO2, Y2O3, and REE2O3 contents and low CaO, FeO, and SiO2 concentrations). The calculated U-Th-Pb chemical ages (135.4 Ma) are contemporaneous with the U-Pb zircon age (135.4–135.6 Ma) of the granite, indicating a magmatic genesis for uraninite. The granites are highly differentiated, and extreme magmatic fractionation might be the main mechanism for the initial uranium enrichment. Brannerite is relatively less abundant and typically forms crusts on ilmenite and rutile or it cements them, representing the local redistribution and accumulation of uranium. Full article
(This article belongs to the Special Issue Mineralization in Subduction Zone)
Show Figures

Figure 1

Figure 1
<p>The tectonic position of the studied granite (after reference [<a href="#B24-minerals-14-00666" class="html-bibr">24</a>]).</p>
Full article ">Figure 2
<p>Distribution of the granite in the Kunduleng area [<a href="#B25-minerals-14-00666" class="html-bibr">25</a>,<a href="#B26-minerals-14-00666" class="html-bibr">26</a>,<a href="#B27-minerals-14-00666" class="html-bibr">27</a>,<a href="#B28-minerals-14-00666" class="html-bibr">28</a>].</p>
Full article ">Figure 3
<p>Field and petrographic characteristics of the Kunduleng granites. (<b>a</b>) Well-developed joints are present in the granite. (<b>b</b>) A vein composed of fine-grained granite within a host rock of porphyritic granite. Hand specimen (<b>c</b>) and thin section (<b>d</b>) images of the porphyritic granite. Hand specimen (<b>e</b>) and thin section (<b>f</b>) images illustrate the characteristics of the fine-grained granite. Abbreviations: Qtz = quartz; Pl = plagioclase; Mic = microcline; and Pth = perthite.</p>
Full article ">Figure 4
<p>Photomicrographs illustrating typical petrographic characteristics of altered granites. (<b>a</b>) Sericitization of plagioclase. (<b>b</b>) Locally, K-feldspar is muscovitized. (<b>c</b>) Muscovitization of biotite. (<b>d</b>) Fluorite in altered biotite. Abbreviations: Qtz = quartz; Pl = plagioclase; Kfs = K-feldspar; Bt = biotite; Ms = muscovite; and Fl = fluorite.</p>
Full article ">Figure 5
<p>BSE images of uraninite and brannerite. Minute uraninite is present in quartz (<b>a</b>–<b>c</b>) and K-feldspar (<b>d</b>). Brannerite is closely associated with ilmenite and rutile (<b>e</b>,<b>f</b>). Abbreviations: Urn = uraninite; Brn = brannerite; Rt = rutile; Ilm = ilmenite; Qtz = quartz; and Kfs = K-feldspar.</p>
Full article ">Figure 6
<p><sup>206</sup>Pb/<sup>238</sup>U versus <sup>207</sup>Pb/<sup>235</sup>U concordia diagram for U–Pb data obtained on zircons from porphyritic granite (<b>a</b>) and fine-grained granite (<b>b</b>).</p>
Full article ">Figure 7
<p>(<b>a</b>) K/Rb vs. SiO<sub>2</sub> plot (after [<a href="#B49-minerals-14-00666" class="html-bibr">49</a>]); (<b>b</b>) Nb/Ta vs. Zr/Hf diagram (after [<a href="#B50-minerals-14-00666" class="html-bibr">50</a>]); (<b>c</b>) the Ba-Rb-Sr ternary diagram (after [<a href="#B51-minerals-14-00666" class="html-bibr">51</a>]); (<b>d</b>) 10000Ga/Al vs. Zr diagram (after [<a href="#B45-minerals-14-00666" class="html-bibr">45</a>,<a href="#B52-minerals-14-00666" class="html-bibr">52</a>]); (<b>e</b>) A-B diagram (after [<a href="#B23-minerals-14-00666" class="html-bibr">23</a>,<a href="#B46-minerals-14-00666" class="html-bibr">46</a>,<a href="#B47-minerals-14-00666" class="html-bibr">47</a>]); and (<b>f</b>) U vs. Th diagram (after [<a href="#B23-minerals-14-00666" class="html-bibr">23</a>,<a href="#B48-minerals-14-00666" class="html-bibr">48</a>]).</p>
Full article ">Figure 8
<p>REE patterns (<b>a</b>) and trace element spider diagrams (<b>b</b>) for the Kunduleng granites. Values for normalization are sourced from references [<a href="#B53-minerals-14-00666" class="html-bibr">53</a>,<a href="#B54-minerals-14-00666" class="html-bibr">54</a>]. The uraniferous leucogranites are sourced from reference [<a href="#B55-minerals-14-00666" class="html-bibr">55</a>].</p>
Full article ">Figure 9
<p>Zircon ε<sub>Hf</sub>(t) versus age diagram for the Kunduleng granite. The U–Pb ages and Hf isotopic composition of zircon in magmatic rocks from the Great Xing’an Range are sourced from references [<a href="#B56-minerals-14-00666" class="html-bibr">56</a>,<a href="#B57-minerals-14-00666" class="html-bibr">57</a>,<a href="#B58-minerals-14-00666" class="html-bibr">58</a>,<a href="#B59-minerals-14-00666" class="html-bibr">59</a>,<a href="#B60-minerals-14-00666" class="html-bibr">60</a>,<a href="#B61-minerals-14-00666" class="html-bibr">61</a>,<a href="#B62-minerals-14-00666" class="html-bibr">62</a>,<a href="#B63-minerals-14-00666" class="html-bibr">63</a>,<a href="#B64-minerals-14-00666" class="html-bibr">64</a>,<a href="#B65-minerals-14-00666" class="html-bibr">65</a>,<a href="#B66-minerals-14-00666" class="html-bibr">66</a>,<a href="#B67-minerals-14-00666" class="html-bibr">67</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) Ba vs. Sr and (<b>b</b>) K/Rb vs. Sr diagrams of the Kunduleng granites. For modeling Rayleigh fractionation, the FC–AFC–FCA and mixing modeler developed by Ersoy and Helvacı (2010) [<a href="#B68-minerals-14-00666" class="html-bibr">68</a>] is employed.</p>
Full article ">Figure 11
<p>(<b>a</b>) Total REE content (wt%) versus U/Th diagram of uraninite (after [<a href="#B73-minerals-14-00666" class="html-bibr">73</a>,<a href="#B74-minerals-14-00666" class="html-bibr">74</a>]); (<b>b</b>) ThO<sub>2</sub> + Y<sub>2</sub>O<sub>3</sub> versus UO<sub>2</sub> (wt%) diagram of uraninite (after [<a href="#B70-minerals-14-00666" class="html-bibr">70</a>]); (<b>c</b>) chondrite-normalized rare earth element patterns for uraninite from the Kunduleng granite. The data for uraninite in intrusive, vein-type, and volcanic-related deposits are from reference [<a href="#B75-minerals-14-00666" class="html-bibr">75</a>]. The values utilized for normalization come from the reference [<a href="#B76-minerals-14-00666" class="html-bibr">76</a>]. (<b>d</b>) Weighted mean U–Th–Pb EPMA chemical ages of uraninite.</p>
Full article ">
20 pages, 9384 KiB  
Article
Petrogenetic Implications of the Lithium-Rich Tongtianmiao Granite Pluton, South China: Evidence from Geochemistry and Geochronology
by Xinhui Yu, Yongzhang Zhou, Wei Cao, Hanyu Wang, Can Zhang, Lifeng Zhong, Wu Wei, Zhiqiang Wang, Jianying Yao, Zhiqiang Chen and Qinghe Xu
Minerals 2024, 14(7), 637; https://doi.org/10.3390/min14070637 - 21 Jun 2024
Viewed by 1065
Abstract
The South China Block, a region renowned for its extensive granite distribution and rich metal deposits, serves as a natural laboratory for the study of granite-related mineralization. This research focuses on the Tongtianmiao granite pluton, which is located at the intersection of the [...] Read more.
The South China Block, a region renowned for its extensive granite distribution and rich metal deposits, serves as a natural laboratory for the study of granite-related mineralization. This research focuses on the Tongtianmiao granite pluton, which is located at the intersection of the Qin-Hang and Nanling metallogenic belts and has been confirmed as a significant lithium mineral resource. Despite its discovery and ongoing development, the lithium-rich Tongtianmiao pluton has been understudied, particularly concerning its petrogenesis, which has only recently come to the forefront of scientific inquiry. By integrating an array of petrogeochemical data with geochronological studies derived from zircon and monazite dating, this study provides insights into the magmatic processes related to lithium enrichment in the Tongtianmiao granites. The Tongtianmiao granites are classified as A-type granites characterized by high SiO2 contents (69.18–78.20 wt.%, average = 74.08 wt.%), K2O + Na2O contents (4.59–8.34 wt.%, average = 6.86 wt.%), A/CNK > 1.2, and low concentrations of Ca, Mg, and Fe. These granites are enriched in alkali metals such as Li, Rb, and Cs but are significantly depleted in Ba, Sr, and Eu. They show no significant fractionation of light or heavy rare-earth elements but present characteristic tetrad effects. A finding of this study is the identification of multiple ages from in situ zircon U–Pb dating, which implies a prolonged history of magmatic activity. However, given the high uranium content in zircons, which could render U–Pb ages unreliable, emphasis is placed on the monazite U–Pb ages. These ages cluster at approximately 172.1 ± 1.1 Ma and 167.9 ± 1.6 Ma, indicating a Middle Jurassic period of granite formation. This timing correlates with the retreat of the Pacific subduction plate and the associated NE-trending extensional fault activity, which likely provided favorable conditions for lithium enrichment. The study concluded that the Tongtianmiao granites were formed through partial melting of crustal materials and subsequent underplating by mantle-derived materials, and were contaminated by strata materials. This process resulted in the formation of highly differentiated granite through magmatic differentiation and external forces. These findings have significant implications for understanding the petrogenesis of lithium-rich granites and are expected to inform future exploration endeavors in the Tongtianmiao pluton. Full article
Show Figures

Figure 1

Figure 1
<p>Geological sketch map of the South China Block (modified from Chen et al., 1989 [<a href="#B28-minerals-14-00637" class="html-bibr">28</a>], Zhou et al., 2012 [<a href="#B21-minerals-14-00637" class="html-bibr">21</a>]), showing the distribution of Yanshanian granitic plutons. The map is presented in the WGS84 coordinate system.</p>
Full article ">Figure 2
<p>(<b>a</b>) Geological sketch showing the strata and intrusion distribution in the Xianghualing ore field (modified from Yuan et al., 2008 [<a href="#B14-minerals-14-00637" class="html-bibr">14</a>]); (<b>b</b>) geological sketch showing the horizontal distribution of sampling boreholes ZK6404, ZK7412 and ZK8203, CRS: WGS84.</p>
Full article ">Figure 3
<p>Photographs and micrographs of representative rocks: (<b>a</b>) lithium-rich mica quartz vein; (<b>b</b>) zinnwaldite specimen; (<b>c</b>) fine-grained zinnwaldite granite veins intersecting medium-grained zinnwaldite granite; (<b>d</b>) grain size variation and greisen vein in zinnwaldite granite; (<b>e</b>) greisen; (<b>f</b>) albitized zinnwaldite granite; (<b>g</b>) zinnwaldite granite; (<b>h</b>) biotite monzonitic granite; (<b>i</b>) sericitized granite thin section; (<b>j</b>) zinnwaldite granite thin section; (<b>k</b>) greisen thin section; (<b>l</b>) altered quartz-enriched granite thin section. Pl—plagioclase, Kfs—potassium feldspar, Qz—quartz, Znw—zinnwaldite, Ms—muscovite, Bt—biotite, Ser—sericite.</p>
Full article ">Figure 4
<p>Geologic cross sections of drill holes ZK6404, ZK7412 and ZK8203 showing the spatial relationships among different rocks and sampling locations.</p>
Full article ">Figure 5
<p>Plots of SiO<sub>2</sub> vs. (<b>a</b>) Al<sub>2</sub>O<sub>3</sub>; (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub><sup>T</sup>; (<b>c</b>) MnO; (<b>d</b>) MgO; (<b>e</b>) CaO; (<b>f</b>) Na<sub>2</sub>O; (<b>g</b>) Rb; (<b>h</b>) Sr; and (<b>i</b>) Ba for the Tongtianmiao granitic rocks and other intrusions in the Xianghualing ore field (data obtained from previous research [<a href="#B7-minerals-14-00637" class="html-bibr">7</a>,<a href="#B8-minerals-14-00637" class="html-bibr">8</a>,<a href="#B11-minerals-14-00637" class="html-bibr">11</a>,<a href="#B18-minerals-14-00637" class="html-bibr">18</a>]).</p>
Full article ">Figure 6
<p>Chemical classification of Tongtianmiao plutonic rocks using the total alkali versus silica (TAS) diagram, modified from Middlemost (1994) [<a href="#B38-minerals-14-00637" class="html-bibr">38</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) SiO<sub>2</sub> vs. K<sub>2</sub>O diagram, modified from Peccerillo and Taylor (1976) [<a href="#B39-minerals-14-00637" class="html-bibr">39</a>]; (<b>b</b>) A/CNK versus A/NK plots of granites from the Tongtianmiao pluton, based on the diagram of Maniar and Piccoli (1989) [<a href="#B40-minerals-14-00637" class="html-bibr">40</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) The primitive mantle normalized diagram of trace elements and (<b>b</b>) the chondrite-normalized REE patterns, with normalization values from Sun and McDonough (1989) [<a href="#B41-minerals-14-00637" class="html-bibr">41</a>].</p>
Full article ">Figure 9
<p>Zircon CL images of samples ZK7412-1 and ZK7412-5.</p>
Full article ">Figure 10
<p>Monazite CL images of samples ZK7412-5 and ZK7412-6.</p>
Full article ">Figure 11
<p>Monazite U–Pb ages of sample ZK7412-5.</p>
Full article ">Figure 12
<p>Monazite U–Pb ages of sample ZK7412-6.</p>
Full article ">Figure 13
<p>Granite discrimination diagram of granite genetic type (modified from [<a href="#B55-minerals-14-00637" class="html-bibr">55</a>]): (<b>a</b>) 10,000 Ga/Al vs. Zr; (<b>b</b>) Zr + Nb + Ce + Y vs. FeO<sup>T</sup>/MgO; (<b>c</b>) 10,000 Ga/Al vs. Ce; (<b>d</b>) Zr + Nb + Ce + Y vs. (Na<sub>2</sub>O + K<sub>2</sub>O)/CaO. FG—fractionated felsic granite; OTG—other I-, S- and M-type granite.</p>
Full article ">Figure 14
<p>Granite discrimination diagram of the tectonic setting (modified from [<a href="#B62-minerals-14-00637" class="html-bibr">62</a>]): (<b>a</b>) Y vs. Nb; (<b>b</b>) Yb vs. Ta. Syn-COLG—syn-collisional granite; VAG—volcanic arc granite; ORG—ocean ridge granite; WPG—within plate granite; MORG—mantle-derived ocean ridge granite.</p>
Full article ">
24 pages, 12991 KiB  
Article
Petrogenesis and Geodynamic Evolution of A-Type Granite Bearing Rare Metals Mineralization in Egypt: Insights from Geochemistry and Mineral Chemistry
by Mohamed M. Ghoneim, Ahmed E. Abdel Gawad, Hanaa A. El-Dokouny, Maher Dawoud, Elena G. Panova, Mai A. El-Lithy and Abdelhalim S. Mahmoud
Minerals 2024, 14(6), 583; https://doi.org/10.3390/min14060583 - 31 May 2024
Viewed by 1359
Abstract
During the Late Precambrian, the North Eastern Desert of Egypt underwent significant crustal evolution in a tectonic environment characterized by strong extension. The Neoproterozoic alkali feldspar granite found in the Homret El Gergab area is a part of the Arabian Nubian Shield and [...] Read more.
During the Late Precambrian, the North Eastern Desert of Egypt underwent significant crustal evolution in a tectonic environment characterized by strong extension. The Neoproterozoic alkali feldspar granite found in the Homret El Gergab area is a part of the Arabian Nubian Shield and hosts significant rare metal mineralization, including thorite, uranothorite, columbite, zircon, monazite, and xenotime, as well as pyrite, rutile, and ilmenite. The geochemical characteristics of the investigated granite reveal highly fractionated peraluminous, calc–alkaline affinity, A-type granite, and post-collision geochemical signatures, which are emplaced under an extensional regime of within-plate environments. It has elevated concentrations of Rb, Zr, Ba, Y, Nb, Th, and U. The zircon saturation temperature ranges from 753 °C to 766 °C. The formation of alkali feldspar rare metal granite was affected by extreme fractionation and fluid interactions at shallow crustal levels. The continental crust underwent extension, causing the mantle and crust to rise, stretch, and become thinner. This process allows basaltic magma from the mantle to be injected into the continental crust. Heat and volatiles were transferred from these basaltic bodies to the lower continental crust. This process enriched and partially melted the materials in the lower crust. The intrusion of basaltic magma from the mantle into the lower crust led to the formation of A-type granite. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geologic map showing the Arabian Nubian Sheild (ANS). (<b>b</b>) Geological map showing the distribution of the Neoproterozoic basement rocks in the Eastern Desert, Egypt [<a href="#B2-minerals-14-00583" class="html-bibr">2</a>].</p>
Full article ">Figure 2
<p>Geologic map of Homret El Gergab, North Eastern Desert, Egypt, modified after Abd El-Hadi [<a href="#B22-minerals-14-00583" class="html-bibr">22</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Quartz vein (Qz) cutting the Dokhan Volcanics (DV). (<b>b</b>) Sharp intrusive contact between the alkali feldspar granite (Gr) and Dokhan Volcanics (DV). (<b>c</b>) Exfoliation in alkali feldspar granite, (<b>d</b>,<b>e</b>). Open cut in granite for prospecting feldspars. (<b>f</b>) Distribution of alkali feldspar granite at Wadi Abu Masananah.</p>
Full article ">Figure 4
<p>Photomicrographs of the studied alkali feldspar granite at Homret El Gergab, North Eastern Desert of Egypt, clarifying that (<b>a</b>) flame and patchy perthites are associated with antiperthite; (<b>b</b>) perthite encloses plagioclase; (<b>c</b>) biotite is highly altered to ferrichlorite and is associated with quartz; (<b>d</b>) fine muscovite flacks are associated with quartz; (<b>e</b>) quartz encloses euhedral zircon crystal; and (<b>f</b>) zircon is enclosed in iron oxides. Abbreviations: Per, perthite; Ant, antiperthite; Plg, plagioclase; Bt, Biotite; Qz, quartz; Mus, muscovite; Kfs, K-feldspar; Zrn, zircon; Irx, iron oxide.</p>
Full article ">Figure 5
<p>Harker variation diagrams illustrate the distributions of major oxides and trace elements in relation to silica in the studied granite.</p>
Full article ">Figure 6
<p>Geochemical discrimination diagrams present the nomenclature of the studied granite. (<b>a</b>) Ternary An–Ab–Or normative diagram by Barker [<a href="#B23-minerals-14-00583" class="html-bibr">23</a>]. (<b>b</b>) Binary diagram shows the total alkalis versus silica, according to Middlemost [<a href="#B24-minerals-14-00583" class="html-bibr">24</a>].</p>
Full article ">Figure 7
<p>Normalized multi-element pattern according to Sun and McDonough [<a href="#B25-minerals-14-00583" class="html-bibr">25</a>] of the studied granite.</p>
Full article ">Figure 8
<p>Histograms showing the concentrations of U and Th in the studied granitic samples from Homret El Gergab, North Eastern Desert, Egypt.</p>
Full article ">Figure 9
<p>Back-scattered images (BSE) of radioactive minerals associated with other accessory minerals from alkali feldspar granite, Homret El Gergab, North Eastern Desert, Egypt. (<b>a</b>) Fine-grained uranothorite is associated with monazite and is enclosed in feldspar. (<b>b</b>) Fine-grained xenotime and thorite are associated with rutile. (<b>c</b>) Fine-grained xenotime and thorite occur along the rims of zircon. (<b>d</b>) Thin films of thorite occur along the periphery of zircon. (<b>e</b>) Thorite and zircon are enclosed in biotite. (<b>f</b>) Columbite crystals are enclosed in quartz. (<b>g</b>) Highly deformed columbite crystals are enclosed in K-feldspar. (<b>h</b>) Large zircon crystals enclose microinclusions of columbite. Abbreviations: uthr, uranothorite; thr, thorite; Zrn, zircon; Mnz, monazite; Xtm, xenotime; Col, columbite; Rt, rutile; Qz, quartz; Ab, Albite; Kfs, K-feldspar; Bt, Biotite.</p>
Full article ">Figure 10
<p>Back-scattered images (BSE) of Zr and REE minerals associated with other accessories from alkali feldspar granite, Homret El Gergab, North Eastern Desert, Egypt. (<b>a</b>) Zoned zircon crystals are enclosed in feldspar. (<b>b</b>) Fine-grained monazite and zircon are enclosed in feldspar. (<b>c</b>) Xenotime is overgrown along the rims of zircon. (<b>d</b>) Fine-grained xenotime is adjacent to zircon. (<b>e</b>) Hematite is along the periphery of monazite. (<b>f</b>) Quartz encloses monazite. (<b>g</b>) Pyrite encloses rutile. (<b>h</b>) Quartz encloses rutile. Abbreviations: Zrn, zircon; Mnz, monazite; Xtm, xenotime; Rt, rutile; Hem, hematite; Qz, quartz; Ab, Albite; Kfs, K-feldspar.</p>
Full article ">Figure 11
<p>Magma-type discrimination diagrams of the studied granite. (<b>a</b>) Binary molar Al<sub>2</sub>O<sub>3</sub>/(Na<sub>2</sub>O + K<sub>2</sub>O) versus Al<sub>2</sub>O<sub>3</sub>/(CaO + Na<sub>2</sub>O + K<sub>2</sub>O) diagram of Maniar and Piccoli [<a href="#B30-minerals-14-00583" class="html-bibr">30</a>]. (<b>b</b>) Binary K<sub>2</sub>O versus SiO<sub>2</sub> diagram of Rickwood [<a href="#B31-minerals-14-00583" class="html-bibr">31</a>].</p>
Full article ">Figure 12
<p>Tectonic setting discrimination diagrams of the studied granite. (<b>a</b>) Binary Rb versus (Y + Nb) diagram of Pearce et al. [<a href="#B32-minerals-14-00583" class="html-bibr">32</a>]. (<b>b</b>) Binary Nb versus Y diagram of Pearce et al. [<a href="#B32-minerals-14-00583" class="html-bibr">32</a>]. (<b>c</b>) Binary discrimination K<sub>2</sub>O/MgO versus 10,000*Ga/Al diagram of Whalen et al. [<a href="#B33-minerals-14-00583" class="html-bibr">33</a>]. (<b>d</b>) Binary discrimination Nb versus 10,000*Ga/Al diagram of Whalen et al. [<a href="#B33-minerals-14-00583" class="html-bibr">33</a>].</p>
Full article ">Figure 13
<p>Petrogenesis discrimination diagrams show the distribution of the analyzed alkali feldspar granite samples. (<b>a</b>) Binary K<sub>2</sub>O versus Rb diagram, where MT refers to the magmatic trend and PH refers to the pegmatitic hydrothermal trend, according to Shaw [<a href="#B40-minerals-14-00583" class="html-bibr">40</a>]. The shaded area illustrates the field of the Ras ed Dome ring complex, Sudan, according to O’Halloran [<a href="#B37-minerals-14-00583" class="html-bibr">37</a>]. (<b>b</b>) Binary Rb versus Sr diagram. (<b>c</b>) Binary Ba versus Rb diagram of Mason [<a href="#B42-minerals-14-00583" class="html-bibr">42</a>].</p>
Full article ">Figure 14
<p>Simplified geodynamic model for the origin of the Homret El Gergab alkali feldspar granite modified after Tavakoli et al. [<a href="#B59-minerals-14-00583" class="html-bibr">59</a>].</p>
Full article ">
Back to TopTop