[go: up one dir, main page]

 
 
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
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (561)

Search Parameters:
Keywords = SrTiO3

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 6955 KiB  
Article
Zircon U-Pb Ages of the Granitoids in Shanxi and Its Significance for Tectonic Evolution of North China Craton in Mesozoic
by Fuhui Qi, Pengpeng Li and Chao Liu
Minerals 2024, 14(9), 940; https://doi.org/10.3390/min14090940 (registering DOI) - 15 Sep 2024
Viewed by 249
Abstract
Mesozoic granitoid formations offer crucial insights into the tectonic history of the North China Craton. New zircon U-Pb ages of two Mesozoic granitoids in the Huai’an terrane yield ages of 226.4 ± 1.1 Ma for the Yihe (YH) granite and 156.3 ± 2.9 [...] Read more.
Mesozoic granitoid formations offer crucial insights into the tectonic history of the North China Craton. New zircon U-Pb ages of two Mesozoic granitoids in the Huai’an terrane yield ages of 226.4 ± 1.1 Ma for the Yihe (YH) granite and 156.3 ± 2.9 Ma for the Zhujiagou (ZJG) granodiorite. The negative Nb, Ta, and Ti anomalies; high Nb/Ta ratios (20.4 to 24.1); high (La/Yb)N (30–84); low initial 87Sr/86Sr ratios (0.707725–0.708188); and negative ƐNd(t) values of the Yihe complex suggest that it originated from the partial melting of the lower crust and lithospheric mantle. However, the geochemical and Sr-Nd isotopic results of the ZJG granodiorite are characterized by I-type granites: Na2O + K2O values of 7.27 to 7.94 wt%, negative Nb anomalies, positive Pb anomalies, higher initial 87Sr/86Sr ratios (0.710979–0.714841), and much lower ƐNd(t) values (−27.1 to −30.1). The Late Jurassic Zhujiagou complex was derived from partial melting of a thickened low crust, and during the Late Triassic, magmatic rocks were formed under a post-collisional extensional regime. Multiple upwellings of the asthenosphere facilitated the mixing of magmas derived from partial melting of the lithospheric mantle and lower crust. These mixed magmas then ascended to the upper crust after undergoing fractional crystallization, leading to the formation of the YH complex. In the Late Jurassic, the tectonic regime of the NCC shifted from compression to extension. The Late Jurassic intrusion identified in this study developed within a compressional setting linked to the subduction of the Paleo-Pacific Ocean. Full article
(This article belongs to the Section Mineral Geochemistry and Geochronology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>): Location map showing the tectonic setting of the North China Craton (modified after [<a href="#B20-minerals-14-00940" class="html-bibr">20</a>]), (<b>b</b>): Simplified geological map of the tectonic subdivision of the North China Craton and location of Shanxi Province, (<b>c</b>): The outline of Shanxi Province and the distribution of alkaline intrusive rocks, the rectangle shows the position of <a href="#minerals-14-00940-f002" class="html-fig">Figure 2</a> (DTGL: Daxinganling-Taihang gravity lineament, modified after [<a href="#B25-minerals-14-00940" class="html-bibr">25</a>]).</p>
Full article ">Figure 2
<p>Simplified geological map of the Huai’an complex and sample sites of this study. The Yihe granite–diorite complex is represented by sample YH and the Zhujiagou complex by sample ZJG. The age of Mesozoic granitoid rocks are cited by Su et al. [<a href="#B26-minerals-14-00940" class="html-bibr">26</a>] and Shao et al. [<a href="#B27-minerals-14-00940" class="html-bibr">27</a>].</p>
Full article ">Figure 3
<p>Field photograph of (<b>a</b>) ZJG granite–diorite complex; (<b>c</b>) ZJG tonalite complex; and Microphotographs (crossed nicols) of (<b>b</b>) YH granite–diorite complex; (<b>d</b>) YH tonalite complex from Huai’an terrain (Q: quartz; Pl: plagioclase; Bi: biotite; Af: alkali-feldspar).</p>
Full article ">Figure 4
<p>Representative CL images for analyzed zircons of the Mesozoic granitoids from Huai’an terrane.</p>
Full article ">Figure 5
<p>U-Pb concordia diagram of zircons of the Mesozoic granitoids from Huai’an terrane.</p>
Full article ">Figure 6
<p>Major element concentrations for the Mesozoic granitoids from Huai’an terrain: (<b>a</b>) SiO<sub>2</sub> versus total alkali (Na<sub>2</sub>O + K<sub>2</sub>O) content diagram [<a href="#B45-minerals-14-00940" class="html-bibr">45</a>]; (<b>b</b>) K<sub>2</sub>O versus SiO<sub>2</sub> diagram (after [<a href="#B46-minerals-14-00940" class="html-bibr">46</a>]); (<b>c</b>) A/CNK versus A/NK diagram (A/CNK = molar ratio of Al<sub>2</sub>O<sub>3</sub>/(CaO + Na<sub>2</sub>O + K<sub>2</sub>O), A/NK = molar ratio of Al<sub>2</sub>O<sub>3</sub>/(Na<sub>2</sub>O + K<sub>2</sub>O)); (<b>d</b>) Q-A-P (quartz-alkali feldspar-plagioclase feldspar) diagram.</p>
Full article ">Figure 7
<p>Chondrite-normalized REE patterns of the Mesozoic granitoids from Huai’an terrain. Chondrite values are from [<a href="#B47-minerals-14-00940" class="html-bibr">47</a>].</p>
Full article ">Figure 8
<p>Primitive mantle-normalized trace element patterns of the Mesozoic granitoids from the Huai’an terrain, with primitive mantle values sourced from [<a href="#B47-minerals-14-00940" class="html-bibr">47</a>].</p>
Full article ">Figure 9
<p>Variations in major element oxides for the Mesozoic granitoids from the Huai’an terrain. The shaded area indicates the compositions of partial melts derived from lower crustal rocks in experimental studies [<a href="#B22-minerals-14-00940" class="html-bibr">22</a>,<a href="#B48-minerals-14-00940" class="html-bibr">48</a>,<a href="#B49-minerals-14-00940" class="html-bibr">49</a>,<a href="#B50-minerals-14-00940" class="html-bibr">50</a>].</p>
Full article ">Figure 10
<p>Petrogenetic discrimination diagrams for the Mesozoic granitoids from the Huai’an terrain are presented as Th/Hf versus Th and Ba/Nd versus Ba. Insets depict schematic C<sup>H</sup> versus C<sup>H</sup>/C<sup>M</sup> diagrams (where C<sup>H</sup> denotes highly incompatible element concentrations and C<sup>M</sup> denotes moderately incompatible element concentrations). The curves represent calculated melt compositions resulting from partial melting, magma mixing, and fractional crystallization (revised [<a href="#B54-minerals-14-00940" class="html-bibr">54</a>]).</p>
Full article ">Figure 11
<p>Sketch map illustrating the distribution of the late Triaccic and Jurassic magmatic rocks in the NCC (modified after Zhang et al. [<a href="#B20-minerals-14-00940" class="html-bibr">20</a>]).</p>
Full article ">Figure 12
<p>Trace element discrimination diagrams for the ZJG and YH complex [<a href="#B66-minerals-14-00940" class="html-bibr">66</a>]. VAG—volcanic arc granites; ORG—ocean ridge granite; WPG—within-plate granites; syn-COLG—syn-collisional granites.</p>
Full article ">
10 pages, 2302 KiB  
Article
Study on Microwave Dielectric Materials an Adjustable Temperature Drift Coefficient and a High Dielectric Constant
by Yuan-Bin Chen, Yu Fan, Shiuan-Ho Chang and Shaobing Shen
Ceramics 2024, 7(3), 1227-1236; https://doi.org/10.3390/ceramics7030081 - 13 Sep 2024
Viewed by 215
Abstract
This paper reports the dielectric characterizations of (Ca0.95Sr0.05)(Ti1−xSnx)O3 ceramics prepared using a solid-state reaction method with various x values. X-ray diffraction spectroscopy analyses showed that the crystal structure of these pure samples was orthorhombic [...] Read more.
This paper reports the dielectric characterizations of (Ca0.95Sr0.05)(Ti1−xSnx)O3 ceramics prepared using a solid-state reaction method with various x values. X-ray diffraction spectroscopy analyses showed that the crystal structure of these pure samples was orthorhombic perovskite. With increasing Sn4+ content, the lattice constant and unit cell volume increased, while the dielectric constant decreased because of the ionic polarizability decreasing. Moreover, a maximum Q × f value of 5242 (GHz), a dielectric constant (εr) of 91.23, and a temperature coefficient (τf) of +810 ppm/°C were achieved for samples sintered at 1350 °C for 4 h. The microwave dielectric characterization was found to be strongly correlated with the sintering temperature, and the best performance was achieved for the sample sintered at 1350 °C. (Ca0.95Sr0.05)(Ti1−xSnx)O3 possesses a promising potential to be a τf compensator for a near-zero τf dielectric ceramic applied in wireless communication systems. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) X-ray diffraction patterns of (Ca<sub>0.95</sub>Sr<sub>0.05</sub>)(Ti<sub>1−x</sub>Sn<sub>x</sub>)O<sub>3</sub> (x = 0.03~0.09) sintered at 1350 °C for 4 h. (<b>b</b>) X-ray diffraction patterns of the PDF Card #01-070-8504: (Ca<sub>0.75</sub>Sr<sub>0.25</sub>)TiO<sub>3</sub>.</p>
Full article ">Figure 2
<p>(Ca<sub>0.95</sub>Sr<sub>0.05</sub>)(Ti<sub>1−x</sub>Sn<sub>x</sub>)O<sub>3</sub> (x = 0.03) sintered at various temperatures.</p>
Full article ">Figure 3
<p>(<b>a</b>) Bulk density (<b>b</b>) Relative density of (Ca<sub>0.95</sub>Sr<sub>0.05</sub>)(Ti<sub>1−x</sub>Sn<sub>x</sub>)O<sub>3</sub> (x = 0.03~0.09) sintered at various temperatures.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) Bulk density (<b>b</b>) Relative density of (Ca<sub>0.95</sub>Sr<sub>0.05</sub>)(Ti<sub>1−x</sub>Sn<sub>x</sub>)O<sub>3</sub> (x = 0.03~0.09) sintered at various temperatures.</p>
Full article ">Figure 4
<p>Dielectric constant of (Ca<sub>0.95</sub>Sr<sub>0.05</sub>)(Ti<sub>1−x</sub>Sn<sub>x</sub>)O<sub>3</sub> (x = 0.03~0.09) sintered at various temperatures.</p>
Full article ">Figure 5
<p>Value of (Ca<sub>0.95</sub>Sr<sub>0.05</sub>)(Ti<sub>1−x</sub>Sn<sub>x</sub>)O<sub>3</sub> (x = 0.03~0.09) ceramics sintered at various temperatures for 4 h.</p>
Full article ">Figure 6
<p>Temperature coefficient of the resonant frequency of (Ca<sub>0.95</sub>Sr<sub>0.05</sub>)(Ti<sub>1−x</sub>Sn<sub>x</sub>)O<sub>3</sub> (x = 0.03~0.09) ceramics sintered at various temperatures for 4 h.</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 (registering DOI) - 12 Sep 2024
Viewed by 211
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 ">
17 pages, 9225 KiB  
Article
Study of the Characteristics of Ba0.6Sr0.4Ti1-xMnxO3-Film Resistance Random Access Memory Devices
by Kai-Huang Chen, Chien-Min Cheng, Ming-Cheng Kao, Yun-Han Kao and Shen-Feng Lin
Micromachines 2024, 15(9), 1143; https://doi.org/10.3390/mi15091143 - 12 Sep 2024
Viewed by 236
Abstract
In this study, Ba0.6Sr0.4Ti1-xMnxO3 ceramics were fabricated by a novel ball milling technique followed by spin-coating to produce thin-film resistive memories. Measurements were made using field emission scanning electron microscopes, atomic force microscopes, X-ray [...] Read more.
In this study, Ba0.6Sr0.4Ti1-xMnxO3 ceramics were fabricated by a novel ball milling technique followed by spin-coating to produce thin-film resistive memories. Measurements were made using field emission scanning electron microscopes, atomic force microscopes, X-ray diffractometers, and precision power meters to observe, analyze, and calculate surface microstructures, roughness, crystalline phases, half-height widths, and memory characteristics. Firstly, the effect of different sintering methods with different substitution ratios of Mn4+ for Ti4+ was studied. The surface microstructural changes of the films prepared by the one-time sintering method were compared with those of the solid-state reaction method, and the effects of substituting a small amount of Ti4+ with Mn4+ on the physical properties were analyzed. Finally, the optimal parameters obtained in the first part of the experiment were used for the fabrication of the thin-film resistive memory devices. The voltage and current characteristics, continuous operation times, conduction mechanisms, activation energies, and hopping distances of two types of thin-film resistive memory devices, BST and BSTM, were measured and studied under different compliance currents. Full article
(This article belongs to the Special Issue Functional Ceramics: From Fundamental Research to Applications)
Show Figures

Figure 1

Figure 1
<p>Structure diagram of BST- and BSTM-film RRAM devices.</p>
Full article ">Figure 2
<p>XRD patterns of the BST material for one-time sintering method and solid-state reaction method.</p>
Full article ">Figure 3
<p>FE-SEM images of the BST material for one-time sintering method and solid-state reaction method.</p>
Full article ">Figure 4
<p>AFM diagram of BST materials for the solid-state reaction method.</p>
Full article ">Figure 5
<p>AFM diagram of BST materials for the one-time sintering method.</p>
Full article ">Figure 6
<p>XRD patterns of BSTM materials for different Mn proportions.</p>
Full article ">Figure 7
<p>FE-SEM images for different Mn proportions of BSTM materials.</p>
Full article ">Figure 8
<p>AFM surface images for different Mn proportions of BSTM materials.</p>
Full article ">Figure 9
<p>The <span class="html-italic">I-V</span> curves of BST-film RRAM devices for different compliance currents. (red symbol: oxygen ions).</p>
Full article ">Figure 10
<p>The <span class="html-italic">I-V</span> curves of BSTM-film RRAM devices for different compliance currents. (red symbol: oxygen ions).</p>
Full article ">Figure 11
<p>The retention and switching cycle properties of BST- and BSTM-film RRAM devices for different compliance currents.</p>
Full article ">Figure 12
<p>The reliability properties of BST- and BSTM-film RRAM devices for different compliance currents.</p>
Full article ">Figure 13
<p>Conduction mechanism analysis of BST- and BSTM-film RRAM devices for different compliance currents.</p>
Full article ">Figure 14
<p>The activation energy versus applied voltage of BST-film RRAM devices for different compliance currents: (<b>a</b>) temperature variation with 0.5 mA compliance current; (<b>b</b>) temperature variation with 10 mA compliance current; (<b>c</b>) electronic activation energy for 0.5 mA compliance current; (<b>d</b>) electronic activation energy for 10 mA compliance current; (<b>e</b>) voltage and activation energy for 0.5 mA compliance current; (<b>f</b>) voltage and activation energy for 10 mA compliance current. (Blue dots: activation energy versus the applied voltage).</p>
Full article ">Figure 15
<p>Electron hopping model of BST thin film RRAM devices for the compliance currents of (<b>a</b>) 0.5 mA, and (<b>b</b>) 10 mA.</p>
Full article ">Figure 16
<p>The activation energy versus applied voltage of BSTM-film RRAM devices for different compliance currents: (<b>a</b>) temperature variation with 0.5 mA compliance current; (<b>b</b>) temperature variation with 10 mA compliance current; (<b>c</b>) electronic activation energy for 0.5 mA compliance current; (<b>d</b>) electronic activation energy for 10 mA compliance current; (<b>e</b>) voltage and activation energy for 0.5 mA compliance current; (<b>f</b>) voltage and activation energy for 10 mA compliance current. (Blue dots: activation energy versus the applied voltage).</p>
Full article ">Figure 17
<p>Electron hopping model of BSTM thin film RRAM devices for the compliance currents of (<b>a</b>) 0.5 mA, and (<b>b</b>) 10 mA.</p>
Full article ">
19 pages, 7689 KiB  
Article
Development of High-Silica Adakitic Intrusions in the Northern Appalachians of New Brunswick (Canada), and Their Correlation with Slab Break-Off: Insights into the Formation of Fertile Cu-Au-Mo Porphyry Systems
by Fazilat Yousefi, David R. Lentz, James A. Walker and Kathleen G. Thorne
Geosciences 2024, 14(9), 241; https://doi.org/10.3390/geosciences14090241 - 7 Sep 2024
Viewed by 479
Abstract
High-silica adakites exhibit specific compositions, as follows: SiO2 ≥ 56 wt.%, Al2O3 ≥ 15 wt.%, Y ≤ 18 ppm, Yb ≤ 1.9 ppm, K2O/Na2O ≥ 1, MgO < 3 wt.%, high Sr/Y (≥10), and La/Yb [...] Read more.
High-silica adakites exhibit specific compositions, as follows: SiO2 ≥ 56 wt.%, Al2O3 ≥ 15 wt.%, Y ≤ 18 ppm, Yb ≤ 1.9 ppm, K2O/Na2O ≥ 1, MgO < 3 wt.%, high Sr/Y (≥10), and La/Yb (>10). Devonian I-type adakitic granitoids in the northern Appalachians of New Brunswick (NB, Canada) share geochemical signatures of adakites elsewhere, i.e., SiO2 ≥ 66.46 wt.%, Al2O3 > 15.47 wt.%, Y ≤ 22 ppm, Yb ≤ 2 ppm, K2O/Na2O > 1, MgO < 3 wt.%, Sr/Y ≥ 33 to 50, and La/Yb > 10. Remarkably, adakitic intrusions in NB, including the Blue Mountain Granodiorite Suite, Nicholas Denys, Sugar Loaf, Squaw Cap, North Dungarvan River, Magaguadavic Granite, Hampstead Granite, Tower Hill, Watson Brook Granodiorite, Rivière-Verte Porphyry, Eagle Lake Granite, Evandale Granodiorite, North Pole Stream Suite, and the McKenzie Gulch porphyry dykes all have associated Cu mineralization, similar to the Middle Devonian Cu porphyry intrusions in Mines Gaspé, Québec. Trace element data support the connection between adakite formation and slab break-off, a mechanism influencing fertility and generation of porphyry Cu systems. These adakitic rocks in NB are oxidized, and are relatively enriched in large ion lithophile elements, like Cs, Rb, Ba, and Pb, and depleted in some high field strength elements, like Y, Nb, Ta, P, and Ti; they also have Sr/Y ≥ 33 to 50, Nb/Y > 0.4, Ta/Yb > 0.3, La/Yb > 10, Ta/Yb > 0.3, Sm/Yb > 2.5, Gd/Yb > 2.0, Nb + Y < 60 ppm, and Ta + Yb < 6 ppm. These geochemical indicators point to failure of a subducting oceanic slab (slab rollback to slab break-off) in the terminal stages of subduction, as the generator of post-collisional granitoid magmatism. The break-off and separation of a dense subducted oceanic plate segment leads to upwelling asthenosphere, heat advection, and selective partial melting of the descending oceanic slab (adakite) and (or) suprasubduction zone lithospheric mantle. The resulting silica-rich adakitic magmas ascend through thickened mantle lithosphere, with minimal affect from the asthenosphere. The critical roles of transpression and transtension are highlighted in facilitating the ascent and emplacement of these fertile adakitic magmas in postsubduction zone settings. Full article
(This article belongs to the Special Issue Zircon U-Pb Geochronology Applied to Tectonics and Ore Deposits)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Major tectonic zones of the Canadian Appalachians; (<b>b</b>) Tectonic zones and cover sequences of New Brunswick (modified from [<a href="#B27-geosciences-14-00241" class="html-bibr">27</a>]).</p>
Full article ">Figure 2
<p>Regional map of the New Brunswick Appalachians, showing the location of Devonian mafic-to-felsic granitoids and major faults (modified from [<a href="#B28-geosciences-14-00241" class="html-bibr">28</a>]).</p>
Full article ">Figure 3
<p>Geochemical discrimination diagrams for adakitic samples investigated: (<b>a</b>) SiO<sub>2</sub> vs. Na<sub>2</sub>O + K<sub>2</sub>O discrimination diagram. Field boundaries from Cox et al. [<a href="#B32-geosciences-14-00241" class="html-bibr">32</a>]; (<b>b</b>) SiO<sub>2</sub> vs. K<sub>2</sub>O discrimination diagram with field boundaries from [<a href="#B33-geosciences-14-00241" class="html-bibr">33</a>]; (<b>c</b>) Al<sub>2</sub>O<sub>3</sub>/(CaO + K<sub>2</sub>O + Na<sub>2</sub>O) (A/CNK) vs. Al<sub>2</sub>O<sub>3</sub>/(Na<sub>2</sub>O + K<sub>2</sub>O) (A/NK) diagram modified from [<a href="#B34-geosciences-14-00241" class="html-bibr">34</a>]. The line with an amount of A/CNK = 1.1 is a key parameter to discriminate S- from I-type granites [<a href="#B35-geosciences-14-00241" class="html-bibr">35</a>]; (<b>d</b>) FeOt/(FeOt + MgO) vs. SiO<sub>2</sub> discrimination diagram with field boundaries from [<a href="#B36-geosciences-14-00241" class="html-bibr">36</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) (La/Yb)<sub>N</sub> vs. (Yb)<sub>N</sub> discrimination diagram with field boundaries from [<a href="#B37-geosciences-14-00241" class="html-bibr">37</a>]; (<b>b</b>) Sr/Y vs. Y discrimination diagram with field boundaries from [<a href="#B37-geosciences-14-00241" class="html-bibr">37</a>]; (<b>c</b>) SiO<sub>2</sub> vs. MgO discrimination diagram for high- and low-silica adakite; (<b>d</b>) primitive mantle-normalized extended element spider diagram. Symbols are the same as <a href="#geosciences-14-00241-f003" class="html-fig">Figure 3</a>. Normalized factors are from [<a href="#B38-geosciences-14-00241" class="html-bibr">38</a>]. TTG = tonalite–trondhjemite–granodiorite, ADR = andesite–dacite–rhyolite.</p>
Full article ">Figure 5
<p>Harker diagrams of Devonian adakitic rocks of NB. SiO<sub>2</sub> vs. (<b>a</b>) TiO<sub>2</sub>, (<b>b</b>) Al<sub>2</sub>O<sub>3</sub>, (<b>c</b>) Ni, and (<b>d</b>) Co. The same symbols as <a href="#geosciences-14-00241-f003" class="html-fig">Figure 3</a> are used. The arrows indicate a general fractionation trend towards high silica.</p>
Full article ">Figure 6
<p>Geochemical discrimination diagrams. (<b>a</b>) FeOt/MgO vs. Zr + Nb + Ce + Y (ppm) and (<b>b</b>) Zr + Nb + Ce + Y (ppm) vs. (Na<sub>2</sub>O + K<sub>2</sub>O)/CaO. Field boundaries are from [<a href="#B40-geosciences-14-00241" class="html-bibr">40</a>]. A-type: A-type granite, FG: fractionated granite rocks, OTG: unfractionated granite/other type of granite.</p>
Full article ">Figure 7
<p>Tectonomagmatic discrimination diagrams for differentiating among slab failure, arc, and A-type granites applied to the New Brunswick granites investigated. (<b>a</b>) Nb + Y vs. Ta/Yb; (<b>b</b>) Ta + Yb vs. Ta/Yb; (<b>c</b>) Nb + Y vs. La/Yb; (<b>d</b>) Ta + Yb vs. Sm/Yb; (<b>e</b>) Nb + Y vs. Gd/Yb; (<b>f</b>) Ta + Yb vs. Gd/Yb; (<b>g</b>) Nb + Y vs. Nb/Y; (<b>h</b>) Ta + Yb vs. Nb/Y. All field boundaries are from [<a href="#B48-geosciences-14-00241" class="html-bibr">48</a>,<a href="#B50-geosciences-14-00241" class="html-bibr">50</a>,<a href="#B51-geosciences-14-00241" class="html-bibr">51</a>], respectively.</p>
Full article ">Figure 8
<p>Continuation of tectonomagmatic discrimination diagrams. (<b>a</b>) Gd/Yb vs. La/Yb; (<b>b</b>) Sm/Yb vs. La/Sm; (<b>c</b>) Ta + Yb vs. Rb; (<b>d</b>) Nb + Y vs. Rb; (<b>e</b>) Y vs. Nb; (<b>f</b>) Yb vs. Ta. Symbols as in <a href="#geosciences-14-00241-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 9
<p>Tectonic discrimination diagrams for adakitic rocks investigated in this study. (<b>a</b>) Nb/Yb vs. Th/Yb, and (<b>b</b>) TiO<sub>2</sub>/Yb vs. Nb/Yb. Field boundaries are from [<a href="#B53-geosciences-14-00241" class="html-bibr">53</a>]. MORB: mid-ocean ridge basalt, OIB: ocean island basalt, Th: tholeiite, Alk: alkaline, EMORB: enriched mid-ocean ridge basalt, NMORB: normal mid-ocean ridge. Symbols as in <a href="#geosciences-14-00241-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 10
<p>Discrimination diagrams for the determination of magmatic source rocks for adakites in New Brunswick. (<b>a</b>) MgO (wt.%) vs. SiO<sub>2</sub> (wt.%), and (<b>b</b>) Mg<sup>#</sup> vs. SiO<sub>2</sub> (wt.%) diagrams for determining the effective factors in creating these adakitic magmas. Symbols as in <a href="#geosciences-14-00241-f007" class="html-fig">Figure 7</a>. Field boundaries are from [<a href="#B54-geosciences-14-00241" class="html-bibr">54</a>].</p>
Full article ">Figure 11
<p>Tectonic discrimination diagram for New Brunswick adakites. Field boundaries are from [<a href="#B55-geosciences-14-00241" class="html-bibr">55</a>]. Hb: hornblende, An: anorthite, Ab: albite, En: enstatite, Fa: fayalite, Fo: forsterite, Bt: biotite, Fs: feldspar, Sp: sphene (titanite), Hd: hedenbergite, Ha: haapalaite, and Di: diopside. Symbols as in <a href="#geosciences-14-00241-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 12
<p>Schematic model showing the Silurian–Carboniferous tectonic evolution of the northern Appalachian orogen, and the generation of slab break-off-generated magmas; (<b>a</b>) late Silurian–Early Devonian, and (<b>b</b>) Middle Devonian–Early Carboniferous. Modified from [<a href="#B66-geosciences-14-00241" class="html-bibr">66</a>].</p>
Full article ">
12 pages, 4532 KiB  
Article
Hydrothermal Synthesis of a Valence State Constant High-Entropy Perovskite Sr(TiZrHfVNb)O3 with Improved Photoresponsiveness
by Yihua Bai, Ke Gan, Xiaohu Li and Dongping Duan
Materials 2024, 17(17), 4275; https://doi.org/10.3390/ma17174275 - 29 Aug 2024
Viewed by 329
Abstract
A vanadium ion valence state constant high-entropy perovskite system was synthesized using the hydrothermal method with a trivalent vanadium ion as the vanadium source. The B-site of the perovskite crystal lattice was loaded with five atoms in equal proportions. We tried to synthesize [...] Read more.
A vanadium ion valence state constant high-entropy perovskite system was synthesized using the hydrothermal method with a trivalent vanadium ion as the vanadium source. The B-site of the perovskite crystal lattice was loaded with five atoms in equal proportions. We tried to synthesize the Sr(TiZrHfVNb)O3 high-entropy system using different methods. However, the valence state of the vanadium ion could only be kept constant using the hydrothermal process in the valence balanced high-entropy composition system. There was significant vanadium element segregation and second phase in the Sr(TiZrHfVNb)O3 system prepared using the solid-state reaction process. Also, obvious vanadium ion valence state ascending from V3+ to V5+ appeared in this high-entropy system with an increase in calcination temperature. Inconspicuous vanadium element segregation appeared at 900 °C, the significant segregation phenomenon and second phase appeared at 1200 °C, and the particle size increased with the temperature. This meant that the high-entropy value could not only stabilize the crystal phase, but also stabilize the ionic valence state. Moreover, the constant trivalent vanadium ion valence state could provide coordinated performance with a wide optical response range and a low band gap for the high-entropy system. This suggests that the system might grow a potential ceramic material for optical applications. Full article
Show Figures

Figure 1

Figure 1
<p>XRD spectra of high-entropy Sr(TiZrHfVNb)O<sub>3</sub> system: (<b>a</b>) hydrothermal and calcined at different temperatures, (<b>b</b>) solid-state reaction.</p>
Full article ">Figure 2
<p>Micromorphology of Sr(TiZrHfVNb)O<sub>3</sub> with EDS synthesized using hydrothermal and calcined at different temperatures.</p>
Full article ">Figure 3
<p>Fitted core level spectrum of the high-entropy elements in the Sr(TiZrHfVNb)O<sub>3</sub> system.</p>
Full article ">Figure 4
<p>SEM images and particle size distribution of the Sr(TiZrHfVNb)O<sub>3</sub> samples synthesized at different temperatures.</p>
Full article ">Figure 5
<p>XRD spectra of the high-entropy Sr(TiZrHfVFe)O<sub>3</sub> and fitted core level spectrum of the elements in this system.</p>
Full article ">Figure 6
<p>UV-Vis spectra (<b>a</b>) and the (αhv)<sup>2</sup>~hv curves (<b>b</b>) of Sr(TiZrHfVNb)O<sub>3</sub> system.</p>
Full article ">Figure 7
<p>Band gap of high-entropy Sr(TiZrHfVNb)O<sub>3</sub> systems synthesized using hydrothermal process and calcined different temperatures from extrapolating (αhv)<sup>2</sup>~hv curves.</p>
Full article ">
14 pages, 5319 KiB  
Article
Ultrahigh Electrostrictive Effect in Lead-Free Sodium Bismuth Titanate-Based Relaxor Ferroelectric Thick Film
by Yizhuo Li, Jinyan Zhao, Zhe Wang, Kun Zheng, Jie Zhang, Chuying Chen, Lingyan Wang, Genshui Wang, Xin Li, Yulong Zhao, Gang Niu and Wei Ren
Nanomaterials 2024, 14(17), 1411; https://doi.org/10.3390/nano14171411 - 29 Aug 2024
Viewed by 512
Abstract
In recent years, the development of environmentally friendly, lead-free ferroelectric films with prominent electrostrictive effects have been a key area of focus due to their potential applications in micro-actuators, sensors, and transducers for advanced microelectromechanical systems (MEMS). This work investigated the enhanced electrostrictive [...] Read more.
In recent years, the development of environmentally friendly, lead-free ferroelectric films with prominent electrostrictive effects have been a key area of focus due to their potential applications in micro-actuators, sensors, and transducers for advanced microelectromechanical systems (MEMS). This work investigated the enhanced electrostrictive effect in lead-free sodium bismuth titanate-based relaxor ferroelectric films. The films, composed of (Bi0.5Na0.5)0.8−xBaxSr0.2TiO3 (BNBST, x = 0.02, 0.06, and 0.11), with thickness around 1 μm, were prepared using a sol-gel method on Pt/TiO2/SiO2/Si substrates. By varying the Ba2+ content, the crystal structure, morphology, and electrical properties, including dielectric, ferroelectric, strain, and electromechanical performance, were investigated. The films exhibited a single pseudocubic structure without preferred orientation. A remarkable strain response (S > 0.24%) was obtained in the films (x = 0.02, 0.06) with the coexistence of nonergodic and ergodic relaxor phases. Further, in the x = 0.11 thick films with an ergodic relaxor state, an ultrahigh electrostrictive coefficient Q of 0.32 m4/C2 was achieved. These findings highlight the potential of BNBST films as high-performance, environmentally friendly electrostrictive films for advanced microelectromechanical systems (MEMS) and electronic devices. Full article
(This article belongs to the Section Nanofabrication and Nanomanufacturing)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the film preparation process: (<b>a</b>) BNBST precursor solution; (<b>b</b>) spin coating; (<b>c</b>) rapid thermal processing; (<b>d</b>) film cooling and forming; (<b>e</b>) electrical properties measurement.</p>
Full article ">Figure 2
<p>X-ray diffraction (XRD) patterns of (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.8−x</sub>Ba<sub>x</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> (x = 0.02, 0.06, 0.11) films: (<b>a</b>) wide scanning range 20°~60°; (<b>b</b>) slow scanning of (002) peak in the 45°~48° region and corresponding fit results; (<b>c</b>) the Williamson-Hall analysis plot, the dots are the data refined and calculated from Jade software, the lines are linear fitted results using modified Williamson-hall method.</p>
Full article ">Figure 3
<p>Morphology images of (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.8−x</sub>Ba<sub>x</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> films: (<b>a</b>–<b>c</b>) surface SEM images and corresponding grain size statistics of the film when x = 0.02, 0.06, and 0.11, respectively; (<b>d</b>–<b>f</b>) sectional SEM images of films: (<b>d</b>) x = 0.02, (<b>e</b>) 0.06, and (<b>f</b>) 0.11; AFM images of films: (<b>g</b>) x = 0.02, (<b>h</b>) 0.06, and (<b>i</b>) 0.11.</p>
Full article ">Figure 4
<p>Dielectric temperature spectrum of (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.8−x</sub>Ba<sub>x</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> films: (<b>a</b>) x = 0.02; (<b>b</b>) x = 0.06; and (<b>c</b>) x = 0.11. (<b>d</b>) The fitted curve according to the modified Curie–Weiss law.</p>
Full article ">Figure 5
<p>(<b>a</b>) Polarization loops and (<b>b</b>) corresponding switching current loops of the (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.8−x</sub>Ba<sub>x</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> (x = 0.02, 0.06, 0.11) films under an electric field of 600 kV/cm at 1 kHz. The <span class="html-italic">P</span><sub>m</sub>, <span class="html-italic">P</span><sub>r</sub>, and <span class="html-italic">P</span><sub>m</sub>-<span class="html-italic">P</span><sub>r</sub> values as a function of the Ba<sup>2+</sup> content are presented as an inset in (<b>a</b>).</p>
Full article ">Figure 6
<p>Dielectric constant: (<b>a</b>) x = 0.02; (<b>b</b>) x = 0.06; (<b>c</b>) x = 0.11,and dielectric loss: (<b>d</b>) x = 0.02; (<b>e</b>) x = 0.06; (<b>f</b>) x = 0.11 curves as a function of DC electric field of (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.8−x</sub>Ba<sub>x</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> films at room temperature and 1 kHz.</p>
Full article ">Figure 7
<p>Dielectric constant (<span class="html-italic">ε</span><sub>r</sub>) vs. AC electric field (<span class="html-italic">E</span>) measured at various frequencies for (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.8−x</sub>Ba<sub>x</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> films: (<b>a</b>) x = 0.02; (<b>b</b>) x = 0.06; and (<b>c</b>) x = 0.11. Results of the Rayleigh analysis: (<b>d</b>) initial dielectric constant (<span class="html-italic">ε</span><sub>init</sub>); (<b>e</b>) irreversible Rayleigh coefficient (<span class="html-italic">α</span>’); and (<b>f</b>) ratio of irreversible Rayleigh coefficient to the initial dielectric constant (<span class="html-italic">α</span>’/<span class="html-italic">ε</span><sub>init</sub>).</p>
Full article ">Figure 8
<p>(<b>a</b>) The bipolar strain–electric field (<span class="html-italic">S</span>–<span class="html-italic">E</span>) loops; (<b>b</b>) the unipolar strain–electric field (<span class="html-italic">S</span>–<span class="html-italic">E</span>) loops; and (<b>c</b>) the strain–polarization (<span class="html-italic">S</span>–<span class="html-italic">P</span>) loops measured at 600 kV/cm at room temperature of (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.8−x</sub>Ba<sub>x</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> (x = 0.02, 0.06, 0.11) films. The dotted symbols are experimental data, and the lines fit the data with a parabola in (<b>c</b>). (<b>d</b>) Comparison of the electrostrictive coefficient of this work with other electrostrictive films reported in the literature [<a href="#B14-nanomaterials-14-01411" class="html-bibr">14</a>,<a href="#B15-nanomaterials-14-01411" class="html-bibr">15</a>,<a href="#B17-nanomaterials-14-01411" class="html-bibr">17</a>,<a href="#B18-nanomaterials-14-01411" class="html-bibr">18</a>,<a href="#B26-nanomaterials-14-01411" class="html-bibr">26</a>,<a href="#B27-nanomaterials-14-01411" class="html-bibr">27</a>,<a href="#B28-nanomaterials-14-01411" class="html-bibr">28</a>,<a href="#B59-nanomaterials-14-01411" class="html-bibr">59</a>], all the them were prepared by a sol-gel process, BNBST* refers to (Bi<sub>0.5</sub>Na<sub>0.5</sub>)<sub>0.69</sub>Ba<sub>0.11</sub>Sr<sub>0.2</sub>TiO<sub>3</sub> films.</p>
Full article ">
20 pages, 6048 KiB  
Article
Mineralogical Characterization and Geochemical Signatures of Supergene Kaolinitic Clay Deposits: Insight of Ropp Complex Kaolins, Northcentral Nigeria
by Adamu Yunusa, Hanlie Hong, Atif Salim, Tarig Amam, Chen Liu, Yanxiao Xu, Xiaochao Zuo and Zhaohui Li
Minerals 2024, 14(9), 869; https://doi.org/10.3390/min14090869 - 27 Aug 2024
Viewed by 581
Abstract
This study presents the chemical and mineralogical composition of clay deposits and associated rock types within the Ropp Complex in order to assess the influence of parent lithology on the kaolinization, genesis, and utility of the deposit. Representative kaolin samples from E horizons [...] Read more.
This study presents the chemical and mineralogical composition of clay deposits and associated rock types within the Ropp Complex in order to assess the influence of parent lithology on the kaolinization, genesis, and utility of the deposit. Representative kaolin samples from E horizons of the weathering profiles and their bedrocks were collected from six sites in the Ropp Complex. Clay mineralogy was determined via the XRD technique, while a geochemical analysis was conducted using XRF, SEM coupled with EDS, and ICP-MS. The results showed that all kaolins dominantly contain kaolinite with a content of 77%–98% except for the AS1 kaolin with only minor kaolinite (20%) but mainly illite (65%). The notably lower crystallinity of kaolinite (HI value of 0.53–1.1) as well as its markedly small grain size is consistent with the formation of kaolinite from intensive chemical weathering of igneous rocks. The AS1 kaolin was probably formed from hydrothermal alteration in the burial stage due to the heating of groundwater by the late volcanism. Mobile trace elements (Sr, Ba, and Eu) exhibited a depletion trend, while immobile elements (Hf, Ta, Th) showed enrichment. The relatively more zirconium in kaolins implies the formation of low-temperature kaolinization. The notably high kaolinite content, accompanied by reasonable levels of Fe2O3 and TiO2, signifies a medium-grade quality. Furthermore, chondrite-normalized rare earth element (REE) patterns exhibit congruent trends in rocks and kaolin samples, indicating a relative enrichment in light rare earth elements (LREEs) alongside a discernible negative Eu anomaly. The abundant kaolinite and silicon–aluminum composition make the kaolins suitable for refractories, pharmaceutics, cosmetics, and supplementary cementitious material (SCM). Full article
(This article belongs to the Collection Clays and Other Industrial Mineral Materials)
Show Figures

Figure 1

Figure 1
<p>A geologic map of the study area (modified from [<a href="#B26-minerals-14-00869" class="html-bibr">26</a>]). (<b>a</b>,<b>b</b>) generalized maps showing the location (the purple square) of the Ropp Complex, (<b>c</b>) a geological map showing lithology of the study area and sampling sites.</p>
Full article ">Figure 2
<p>Field images showing two distinct occurrences of kaolin in the weathering profiles of the Ropp Complex. (<b>a</b>) The overview of white kaolinite, (<b>b</b>) the close-up view of white kaolinite, and (<b>c</b>) yellow kaolinite in the E horizon.</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD patterns of representative whole-rock samples of the Ropp Complex kaolinite deposits. I: illite; S: smectite; K: kaolinite; Q: quartz; Or: orthoclase; Et: ettringite; Ht: hematite. (<b>b</b>) XRD patterns of the samples showing the presence of kaolinite (K), smectite (S), and illite (I); Et: ettringite.</p>
Full article ">Figure 4
<p>Scanning electron micrographs with corresponding EDS spectra illustrating the morphology and chemical composition of kaolinite and associated minerals. (<b>a</b>) Aggregates of small elongated-plate illite in pore space showing precipitation from the solution, kaolinite occurring in relatively large plates (AS1). (<b>b</b>) Hexagonal kaolinite plates from sample AS14. (<b>c</b>) The characteristic book-like aggregates of kaolinite (AS20). (<b>d</b>–<b>f</b>) Kaolinite flakes with irregular outlines (AS8, AS9, and AS16).</p>
Full article ">Figure 5
<p>Harker diagram showing the relation between Al<sub>2</sub>O<sub>3</sub> and other oxides.</p>
Full article ">Figure 6
<p>The relation between CIW and CIA (<b>a</b>), CIA and ICV (<b>b</b>), and the ACNK diagram in molecular proportions (<b>c</b>) for the clay samples.</p>
Full article ">Figure 7
<p>Discrimination diagrams for the supergene–hypogene alteration of the Ropp Complex kaolinites (after [<a href="#B53-minerals-14-00869" class="html-bibr">53</a>]). (<b>a</b>) The log(Ce+La+Y) vs. log(Ba+Sr) plot; (<b>b</b>) The (La+Ce+Y)-(Ba+Sr)-Pb ternary plot.</p>
Full article ">Figure 8
<p>(<b>a</b>) Chondrite-normalized trace element diagram (after [<a href="#B54-minerals-14-00869" class="html-bibr">54</a>]); (<b>b</b>) Chondrite-normalized REE diagram (after [<a href="#B55-minerals-14-00869" class="html-bibr">55</a>]). Kaolin-ore samples in blue and bedrock samples in red.</p>
Full article ">Figure 9
<p>A proposed model of formation of the Ropp Complex kaolins.</p>
Full article ">
14 pages, 11847 KiB  
Article
Study on the Relationship between Electron Transfer and Electrical Properties of XLPE/Modification SR under Polarity Reversal
by Zhi-Yuan Wu, Yu-Zhi Jin, Zhe-Xu Shi, Zhi-Yuan Wang and Wei Wang
Polymers 2024, 16(16), 2356; https://doi.org/10.3390/polym16162356 - 20 Aug 2024
Viewed by 375
Abstract
The insulation of high-voltage direct-current (HVDC) cables experiences a short period of voltage polarity reversal when the power flow is adjusted, leading to sever field distortion in this situation. Consequently, improving the insulation performance of the composite insulation structure in these cables has [...] Read more.
The insulation of high-voltage direct-current (HVDC) cables experiences a short period of voltage polarity reversal when the power flow is adjusted, leading to sever field distortion in this situation. Consequently, improving the insulation performance of the composite insulation structure in these cables has become an urgent challenge. In this paper, SiC-SR (silicone rubber) and TiO2-SR nanocomposites were chosen for fabricating HVDC cable accessories. These nanocomposites were prepared using the solution blending method, and an electro-acoustic pulse (PEA) space charge test platform was established to explore the electron transfer mechanism. The space charge characteristics and field strength distribution of a double-layer dielectric composed of cross-linked polyethylene (XLPE) and nano-composite SR at different concentrations were studied during voltage polarity reversal. Additionally, a self-built breakdown platform for flake samples was established to explore the effect of the nanoparticle doping concentration on the breakdown field strength of double-layer composite media under polarity reversal. Therefore, a correlation was established between the micro electron transfer process and the macro electrical properties of polymers (XLPE/SR). The results show that optimal concentrations of nano-SiC and TiO2 particles introduce deep traps in the SR matrix, significantly inhibiting charge accumulation and electric field distortion at the interface, thereby effectively improving the dielectric strength of the double-layer polymers (XLPE/SR). Full article
(This article belongs to the Section Polymer Analysis and Characterization)
Show Figures

Figure 1

Figure 1
<p>Preparation process diagram of nano-doped composite materials.</p>
Full article ">Figure 2
<p>Flat vulcanizing machine.</p>
Full article ">Figure 3
<p>DSC curve of different SiC-SR/TiO<sub>2</sub>-SR samples.</p>
Full article ">Figure 4
<p>AFM of different SiC-SR/TiO<sub>2</sub>-SR samples.</p>
Full article ">Figure 5
<p>Schematic diagram of PEA space charge measurement platform.</p>
Full article ">Figure 6
<p>Polarity reversal voltage waveform applied during space charge testing.</p>
Full article ">Figure 7
<p>Diagram of breakdown test device for sheet-like specimens.</p>
Full article ">Figure 8
<p>Polarity reversal voltage waveform applied during breakdown test.</p>
Full article ">Figure 9
<p>Space charge distribution under polarity reversal. (<b>a</b>) Pure SR/XLPE bilayer samples; (<b>b</b>) 1 wt% SiC-doped SR/XLPE bilayer samples; (<b>c</b>) 3 wt% SiC-doped SR/XLPE bilayer samples; (<b>d</b>) 5 wt% SiC-doped SR/XLPE bilayer samples; (<b>e</b>) 2 wt% TiO<sub>2</sub>-doped SR/XLPE bilayer samples; (<b>f</b>) 4 wt% TiO<sub>2</sub>-doped SR/XLPE bilayer samples; (<b>g</b>) 8 wt% TiO<sub>2</sub>-doped SR/XLPE bilayer samples. (<b>h</b>) Space charge density of different samples at the SR electrode interface inversion during different times.</p>
Full article ">Figure 10
<p>Electric field distribution of nanocomposite SR/XLPE bilayer samples with different doping concentrations. (<b>a</b>) Electric field distribution of SiC-SR/XLPE double-layer samples with different doping concentrations; (<b>b</b>) electric field distribution of TiO<sub>2</sub>-SR/XLPE double-layer samples with different doping concentrations; (<b>c</b>) comparison of 30 s reversal end electric field distribution between 3 wt% SiC-SR/XLPE double-layer samples and 4 wt% TiO<sub>2</sub>-SR/XLPE double-layer samples.</p>
Full article ">Figure 11
<p>Weibull distribution of DC breakdown field strength in different nanocomposite materials. (<b>a</b>) Weibull distribution of DC breakdown field strength in nanocomposite SiC-SR. (<b>b</b>) Weibull distribution of DC breakdown field strength in nanocomposite TiO<sub>2</sub>-SR.</p>
Full article ">Figure 12
<p>The 30 s polarity reversal breakdown characteristics of SR/XLPE double-layer dielectric with different doping: (<b>a</b>) 30 s polarity reversal breakdown characteristics of SiC-SR/XLPE double-layer dielectric; (<b>b</b>) 30 s polarity reversal breakdown characteristics of TiO<sub>2</sub>-SR/XLPE double-layer dielectric.</p>
Full article ">
14 pages, 5631 KiB  
Article
Strengthened Removal of Tetracycline by a Bi/Ni Co-Doped SrTiO3/TiO2 Composite under Visible Light
by Weifang Chen, Na Zhao, Mingzhu Hu, Xingguo Liu and Baoqing Deng
Catalysts 2024, 14(8), 539; https://doi.org/10.3390/catal14080539 - 19 Aug 2024
Viewed by 413
Abstract
A two-step hydrothermal method was used to first obtain a SrTiO3/TiO2 composite then to dope the composite with Bi, Ni and Bi/Ni. Morphology, crystalline structures, surface valances and optical features of SrTiO3/TiO2 and Bi-, Ni-, Bi/Ni-doped SrTiO [...] Read more.
A two-step hydrothermal method was used to first obtain a SrTiO3/TiO2 composite then to dope the composite with Bi, Ni and Bi/Ni. Morphology, crystalline structures, surface valances and optical features of SrTiO3/TiO2 and Bi-, Ni-, Bi/Ni-doped SrTiO3/TiO2 were assessed. XRD and XPS analysis showed that Bi and Ni were successfully doped and existed in Bi(3+) and Ni(2+) oxidation state. UV–vis analysis further revealed that the bandgap energies of TiO2 and SrTiO3/TiO2 were calculated to be 3.14 eV and 3.04 eV. By comparison, Bi, Ni and Bi/Ni doping resulted in the narrowing of bandgaps to 2.82 eV, 2.96 eV and 2.69 eV, respectively. The removal ability of SrTiO3/TiO2 and doped SrTiO3/TiO2 were investigated with tetracycline as the representative pollutant. After 40 min of exposure to visible light, Bi/Ni co-doped SrTiO3/TiO2 photocatalyst was able to remove 90% of the tetracycline with a mineralization rate of about 70%. In addition, first-order removal rate constant was 0.0074 min−1 for SrTiO3/TiO2 and increased to 0.0278 min−1 after co-doping. The strengthened removal by co-doped photocatalyst was attributed mainly to the enhanced absorption of visible light as co-doping resulted in the decreases of bandgap energies. At the same time, the co-doped material was robust against changes in pH. Removal of tetracycline was stable as pH changed from 5 to 9. Tetracycline removal was inhibited to a certain degree by the presence of nitrate, phosphate and high concentration of humic acid. Moreover, the co-doped material exhibited strong structural stability and reusability. In addition, a photocatalysis mechanism with photogenerated holes and ·O2 radicals as main oxidative species was proposed based on entrapping experiments and EPR results. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of the photocatalysts: (<b>a</b>) TiO<sub>2</sub>; (<b>b</b>)S-TO; (<b>c</b>) Bi/S-TO; (<b>d</b>) Ni/S-TO; (<b>e</b>) Bi/Ni/S-TO.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD pattern (<b>b</b>) Locally enlarged XRD patterns of TiO<sub>2</sub>, S-TO, Bi/S-TO, Ni-S-TO and Bi/Ni/S-TO.</p>
Full article ">Figure 3
<p>XPS Spectra: (<b>a</b>) Bi/Ni/S-TO; (<b>b</b>) O 1s; (<b>c</b>) Ti 2p; (<b>d</b>) Sr 3d; (<b>e</b>) Bi 4f; (<b>f</b>) Ni 2p.</p>
Full article ">Figure 3 Cont.
<p>XPS Spectra: (<b>a</b>) Bi/Ni/S-TO; (<b>b</b>) O 1s; (<b>c</b>) Ti 2p; (<b>d</b>) Sr 3d; (<b>e</b>) Bi 4f; (<b>f</b>) Ni 2p.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–vis absorption spectra of TiO<sub>2</sub>, S-TO, Bi/S-TO, Ni-S-TO and Bi/Ni/S-TO and (<b>b</b>) bandgap energies of TiO<sub>2</sub> and Bi/Ni/S-TO.</p>
Full article ">Figure 5
<p>Tetracycline removal: (<b>a</b>) Comparison of photocatalyst; (<b>b</b>) Effects of dosage; (<b>c</b>) Effects of initial pH; (<b>d</b>) Effects of co-existing anion; (<b>e</b>) Effects of co-existing humic acid; (<b>f</b>) Mineralization; (<b>g</b>) Stability and reusability; (<b>h</b>) XRD spectra.</p>
Full article ">Figure 6
<p>(<b>a</b>) Free radical trapping (<b>b</b>) EPR signal of DMPO-·O<sub>2</sub><sup>−</sup> and (<b>c</b>) EPR signal of TEMPO-h<sup>+</sup>.</p>
Full article ">Figure 7
<p>(<b>a</b>) VB-XPS spectrum of Bi/Ni/S-TO. (<b>b</b>) Possible reaction mechanism of tetracycline removal by Bi/Ni/S-TO.</p>
Full article ">
20 pages, 7605 KiB  
Article
Geochronology and Geochemistry of the Uhelchulu Quartz Diorite-Granodiorite in Inner Mongolia of China: Implications for Evolution of the Hegenshan Ocean in the Early-Middle Devonian
by Tianshe Cheng, Wenjing Yang, Chao Teng, Xinjie Yang and Deng Xiao
Minerals 2024, 14(8), 835; https://doi.org/10.3390/min14080835 - 17 Aug 2024
Viewed by 677
Abstract
The Uhelchulu quartz diorite-granodiorite intrusions in Xiwuqi, Inner Mongolia, are exposed along the northwestern margin of the Xilinhot microcontinental block, located within the central and eastern parts of the southeastern Hegenshan suture zone. LA-ICP-MS zircon U-Pb dating yielded crystallization ages of (396 ± [...] Read more.
The Uhelchulu quartz diorite-granodiorite intrusions in Xiwuqi, Inner Mongolia, are exposed along the northwestern margin of the Xilinhot microcontinental block, located within the central and eastern parts of the southeastern Hegenshan suture zone. LA-ICP-MS zircon U-Pb dating yielded crystallization ages of (396 ± 8) Ma for the quartz diorite and (385 ± 5) Ma for the granodiorite, indicating an Early-Middle Devonian magmatic event. The quartz diorite exhibits I-type granite features, characterized by elevated Al2O3 (14.33–15.43 wt%), MgO (3.73–5.62 wt%), and Na (Na2O/K2O = 1.04–1.44), coupled with low P2O5 (0.15–0.20 wt%) and TiO2 (0.73–0.99 wt%). Trace element patterns show relative enrichments in Rb, Th, U, and Pb, while Nb, Ta, Sr and Ti are relatively depleted. Total REE contents are relatively low (123–178 ppm), with significant LREE enrichment (ΣLREE/ΣHREE = 4.75–5.20), and a non-obvious Eu anomaly (δEu = 0.75–0.84). In contrast, the granodiorite displays S-type granite characteristics, with high SiO2 (70.48–73.01 wt%), K (K2O/Na2O = 1.35–1.83), Al2O3 (A/CNK = 1.16–1.31), and a high differentiation index (DI = 76–82). Notably, MgO (1.44–2.24 wt%) contents are low, and significant depletions of Ba, Sr, Ti, and Eu are observed, while Rb, Pb, Th, U, Zr, and Hf are significantly enriched. Total REE contents are relatively low (178–314 ppm), exhibiting significant LREE enrichment (LREE/HREE = 6.17–8.36) and a pronounced negative Eu anomaly (δEu = 0.34–0.49). The overall characteristics point towards an active continental margin arc background for the Uhelchulu intrusions. Previous studies have suggested that the Hegenshan ocean continuously subducted northward from the Early Carboniferous to the Late Permian, but there is a lack of evidence for its geological evolution during the pre-Early Carboniferous. Therefore, this paper provides a certain basis for studying the geological evolution during the pre-Early Carboniferous in the Hegenshan ocean. We preliminarily believed that the Hegenshan ocean underwent a southward subduction towards the Xilinhot microcontinental block in the Xiwuqi area, at least from the Early Devonian to the Middle Devonian and the Hegenshan ocean may might have undergone a shift in subduction mechanism during the Late Devonian or Early Carboniferous. Full article
(This article belongs to the Section Mineral Geochemistry and Geochronology)
Show Figures

Figure 1

Figure 1
<p>Tectonic division of the Central Asian orogenic belt and location of study area (<b>a</b>) revised after Sengör et al., 1993 [<a href="#B1-minerals-14-00835" class="html-bibr">1</a>]; Kai-Jun Zhang, 2014 [<a href="#B8-minerals-14-00835" class="html-bibr">8</a>]. (<b>b</b>,<b>c</b>) after Wang et al., 2020 [<a href="#B28-minerals-14-00835" class="html-bibr">28</a>]). 1. Cretaceous Meiletu Formation; 2. Cretaceous Baiyin Gaolao Formation; 3. Jurassic Manketu Ebo Formation; 4. Triassic Hongqi Formation; 5. Permian Dashizhai Formation; 6. Permian Shoushangou Formation; 7. Carboniferous Amushan Formation; 8. Carboniferous Ben Batu Formation; 9. Devonian Tarbagate Formation; 10. Cretaceous granite porphyry; 11. Cretaceous monzogranite; 12. Triassic monzogranite; 13. Permian syenite granite; 14. Carboniferous monzogranite; 15. Carboniferous diorite granite; 16. Carboniferous quartz diorite; 17. Carboniferous monzogranite; 18. Devonian gabbro; 19. Devonian granodiorite; 20. Devonian quartz diorite; and 21. Fracture.</p>
Full article ">Figure 2
<p>Field and microscopic photographs of the quartz diorite in Uhelchulu and granodiorite in Guiqinkundui. (<b>a</b>,<b>b</b>) field photos of Uhelchulu quartz diorite; (<b>c</b>,<b>d</b>) field photos of Guiqinkundui granodiorite; (<b>e</b>) Uhelchulu quartz diorite with hypidiomorphic crystal (crossed polarizers); (<b>f</b>) Guiqinkundui granodiorite hypidiomorphic crystal (crossed polarizers).</p>
Full article ">Figure 3
<p>Cathodoluminescence (CL) images and analysis points of zircons from the Uhelchulu quartz diorite (<b>a</b>) and the Guiqinkundui granodiorite (<b>b</b>).</p>
Full article ">Figure 4
<p>U-Pb concordia diagram, weighted average histogram of zircons from the Uhelchulu quartz diorite (<b>a</b>,<b>b</b>) and the Guiqinkundui granodiorite (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 5
<p>TAS and AFM diagram of quartz diorite-granodiorite from Uhelchulu to Guiqinkundui in Xiwuqi ((<b>a</b>) after E.A.K.Middlemost, 1994 [<a href="#B50-minerals-14-00835" class="html-bibr">50</a>]; (<b>b</b>) after Irvine T. N., et al. 1971 [<a href="#B51-minerals-14-00835" class="html-bibr">51</a>]).</p>
Full article ">Figure 6
<p>(<b>a</b>) Chondrite-normalized REE distribution patterns (after Boynton, 1984 [<a href="#B52-minerals-14-00835" class="html-bibr">52</a>]) and (<b>b</b>) Primitive mantle-normalized trace elements distribution patterns (after Sun and McDonough, 1989 [<a href="#B53-minerals-14-00835" class="html-bibr">53</a>]) for the Uhelchulu quartz diorite and Guiqinkundui granodiorite, Xiwuqi.</p>
Full article ">Figure 7
<p>Geochemical discrimination diagrams for Uhelchulu quartz diorite and Guiqinkundui granodiorite samples in Xiwuqi. (<b>a</b>) (K<sub>2</sub>O + Na<sub>2</sub>O)/CaO versus Zr + Nb + Ce + Y (ppm) (after Whalen et al., 1987 [<a href="#B61-minerals-14-00835" class="html-bibr">61</a>]); (<b>b</b>) (K<sub>2</sub>O + Na<sub>2</sub>O)/CaO versus 10,000*Ga/Al (after Whalen et al., 1987 [<a href="#B61-minerals-14-00835" class="html-bibr">61</a>]); (<b>c</b>) Na<sub>2</sub>O versus K<sub>2</sub>O (after Collins et al., 1982 [<a href="#B64-minerals-14-00835" class="html-bibr">64</a>]); and (<b>d</b>) P<sub>2</sub>O<sub>5</sub> versus SiO<sub>2</sub> (after Collins et al., 1982 [<a href="#B64-minerals-14-00835" class="html-bibr">64</a>]).</p>
Full article ">Figure 8
<p>Discrimination diagrams for Uhelchulu quartz diorite and Guiqinkundui granodiorite in Xiwuqi, (<b>a</b>) Y-Nb and (<b>b</b>) (Y + Nb)-Rb (after Pearce et al., 1984 [<a href="#B69-minerals-14-00835" class="html-bibr">69</a>]). Syn-COLG (syn-collision granite); WPG (within plate granite); VAG (volcanic arc granite); and ORG (mid ocean ridge granite).</p>
Full article ">Figure 9
<p>Tectonic evolution model of the middle-east section of the Hegenshan ocean in the Early-Middle Devonian (<b>a</b>) and Early Carboniferous (<b>b</b>).</p>
Full article ">Figure 10
<p>Crustal thickness evolution in the research area.</p>
Full article ">
11 pages, 3201 KiB  
Article
Substrate Charge Transfer Induced Ferromagnetism in MnSe/SrTiO3 Ultrathin Films
by Chun-Hao Huang, Chandra Shekar Gantepogu, Peng-Jen Chen, Ting-Hsuan Wu, Wei-Rein Liu, Kung-Hsuan Lin, Chi-Liang Chen, Ting-Kuo Lee, Ming-Jye Wang and Maw-Kuen Wu
Nanomaterials 2024, 14(16), 1355; https://doi.org/10.3390/nano14161355 - 16 Aug 2024
Viewed by 563
Abstract
The observation of superconductivity in MnSe at 12 GPa motivated us to investigate whether superconductivity could be induced in MnSe at ambient conditions. A strain-induced structural change in the ultrathin film could be one route to the emergence of superconductivity. In this report, [...] Read more.
The observation of superconductivity in MnSe at 12 GPa motivated us to investigate whether superconductivity could be induced in MnSe at ambient conditions. A strain-induced structural change in the ultrathin film could be one route to the emergence of superconductivity. In this report, we present the physical property of MnSe ultrathin films, which become tetragonal (stretched ab-plane and shortened c-axis) on a (001) SrTiO3 (STO) substrate, prepared by the pulsed laser deposition (PLD) method. The physical properties of the tetragonal MnSe ultrathin films exhibit very different characteristics from those of the thick films and polycrystalline samples. The tetragonal MnSe films show substantial conductivity enhancement, which could be associated with the presence of superparamagnetism. The optical absorption data indicate that the electron transition through the indirect bandgap to the conduction band is significantly enhanced in tetragonal MnSe. Furthermore, the X-ray Mn L-edge absorption results also reveal an increase in unoccupied state valance bands. This theoretical study suggests that charge transfer from the substrate plays an important role in conductivity enhancement and the emergence of a ferromagnetic order that leads to superparamagnetism. Full article
(This article belongs to the Section Nanofabrication and Nanomanufacturing)
Show Figures

Figure 1

Figure 1
<p>Cross-sectional TEM image of the three studied MnSe films. The measured thicknesses are (<b>a</b>) 30 nm, (<b>b</b>) 140 nm, and (<b>c</b>) 740 nm. The ink (white layer in each image) is a carbon paste used as a layer on the top of the films to protect from damage during the TEM sample preparation process. The regions (i), (ii), (iii), and (iv) marked in the images are Pt (conduction layer for FIB process), ink (organic protection layer), MnSe film, and STO substrate.</p>
Full article ">Figure 2
<p>The temperature dependence of the resistivity of MnSe films. The resistivity decreases dramatically as the thickness of the film becomes thinner, more than one order of magnitude smaller in the 30 nm film than that of the 740 nm film.</p>
Full article ">Figure 3
<p>The magnetic susceptibility of (<b>a</b>) polycrystalline MnSe powder and the (<b>b</b>) 740 nm, (<b>c</b>) 140 nm, and (<b>d</b>) 30 nm MnSe films. The polycrystalline MnSe sample was prepared via the solid-state reaction method. Its magnetic characteristics exhibit anomalies at 180 K and 270 K, respectively. The magnetic anomaly at 270 K arises from the antiferromagnetic order of the solitary hexagonal phase, with a partial transformation into the cubic phase [<a href="#B14-nanomaterials-14-01355" class="html-bibr">14</a>]. The magnetic anomaly observed at 180 K is attributed to the antiferromagnetic order of both the collective hexagonal and cubic phases [<a href="#B14-nanomaterials-14-01355" class="html-bibr">14</a>]. The 180 K magnetic anomaly in the polycrystalline sample is only observed in thick (740 nm) films. It is noted that the magnetic susceptibility of the film increases substantially as film thickness decreases.</p>
Full article ">Figure 4
<p>An M-H curve of the MnSe films. (<b>a</b>) The M-H curve of films reveals the signatures of superparamagnetism, rapid increase at the low magnetic field region, saturation at the high magnetic region, and no hysteresis (zero coercive field). There is a small hysteresis loop in the low field region, as shown in <a href="#app1-nanomaterials-14-01355" class="html-app">Figure S2</a>. The moment is in the order of 10<sup>−4</sup> emu and higher, which is well beyond the sensitivity of SQUID, ~10<sup>−6</sup> emu. (<b>b</b>) The fitting of the M-H curve of the 20 nm film. The fitting curve (red line) agrees excellently with the experimental data (open circles). The extracted density of the magnetic domain is about 5.74 × 10<sup>18</sup> cm<sup>−3</sup>, and the effective magnetic moment of Mn is about 3.21.</p>
Full article ">Figure 5
<p>X-ray diffractions of the MnSe thin films. (<b>a</b>) A radial scan of the MnSe films with thicknesses of 30 nm, 140 nm, and 740 nm on the (001) STO substrate. The films were grown with the c-axis preferred orientation. The * mark (&gt;3 orders of magnitude smaller than the (001) peak) at q near 2.8 Å<sup>−1</sup> and 4.0 Å<sup>−1</sup> are indexed as (112) and (222) peaks of MnSe. (<b>b</b>) A radical scan near the MnSe (002) peak. The peak of the 30 nm film shifts to a higher angle, implying a shorter <span class="html-italic">c</span>-axis lattice constant. (<b>c</b>) ψ-scan profiles with respect to the (022)<sub>STO</sub> and (101)<sub>MnSe</sub> diffraction peaks of the 30 nm film. A 4-fold symmetry in the <span class="html-italic">ab</span>-plane at 45° with respect to the STO <span class="html-italic">a</span>- (<span class="html-italic">b</span>-) axis demonstrates the epitaxial growth of the MnSe film. (<b>d</b>) The (200), (020), and (002) diffraction peaks of the 140 nm film. Their positions have no significant difference. (<b>e</b>) The (200), (020), and (002) diffraction peaks of the 30 nm film. The in-plane diffraction peak is located at a lower angle compared with the (002) peak.</p>
Full article ">Figure 6
<p>Mn L-edge X-ray absorption spectra of the studied MnSe films. The XANES (X-ray absorption near edge spectroscopy) spectra of the Mn 2<span class="html-italic">p</span><sub>3/2</sub> to 3<span class="html-italic">d</span> (L<sub>3</sub>) transition. Four peaks are identified, (marked as A, B, C, and D) which originated from different Mn 3<span class="html-italic">d</span>-related final states. The inset shows the full spectrum of L-edge absorption.</p>
Full article ">Figure 7
<p>The band structure of tetragonal MnSe. (<b>a</b>) The A-AFM phase. (<b>b</b>) The FM phase. For comparison, an identical unit cell is used for both phases. A rigid shift of 0.7 eV is applied to the conduction bands, as mentioned in the text.</p>
Full article ">
23 pages, 47311 KiB  
Article
Petrogenesis and Tectonic Evolution of I- and A-Type Granites of Mount Abu Kibash and Tulayah, Egypt: Evidence for Transition from Subduction to Post-Collision Magmatism
by Amr El-Awady, Mabrouk Sami, Rainer Abart, Douaa Fathy, Esam S. Farahat, Mohamed S. Ahmed, Hassan Osman and Azza Ragab
Minerals 2024, 14(8), 806; https://doi.org/10.3390/min14080806 - 9 Aug 2024
Viewed by 478
Abstract
The Neoproterozoic granitic rocks of Mount Abu Kibash and Tulayah in the central Eastern Desert of Egypt are of geodynamic interest and provide us with important information about the evolution and growth of the northern part of the Arabian–Nubian Shield (ANS) continental crust. [...] Read more.
The Neoproterozoic granitic rocks of Mount Abu Kibash and Tulayah in the central Eastern Desert of Egypt are of geodynamic interest and provide us with important information about the evolution and growth of the northern part of the Arabian–Nubian Shield (ANS) continental crust. They are primarily composed of granodiorites and syenogranites based on new field, mineralogical, and geochemical analyses. The granodiorites are marked by an enrichment of LILEs such as Sr, K, Rb, Ba compared to HFSEs like Nb, Ta, Ti and show a higher concentration of LREEs relative to HREEs. This composition suggests a subduction-related setting and aligns with the characteristics of subducted I-type granites in the ANS. Chemistry of the analyzed primary amphiboles in the investigated granodiorites support a calc-alkaline nature, mixed source and subduction-related setting. The granodiorites represent an early magmatic phase in this setting, likely formed from a mix of mantle-derived mafic magmas and lower crust material, with subsequent fractional crystallization. On the other hand, syenogranites exhibit high SiO2 (72.02–74.02 wt%), total alkali (7.82–8.01 wt%), and Al2O3 (13.79–14.25 wt%) levels, suggesting their derivation from peraluminous (A/CNK > 1) parental magmas. Their REE-normalized patterns are flat with a pronounced negative Eu anomaly, typical of post-collisional A2-type granites worldwide. These rocks originated from the partial melting of a juvenile lower crustal source (tonalite) in a post-collisional setting, driven by lithospheric delamination that facilitated mantle upwelling and underplating to the lower crust. Interaction between the upwelled mantle and lower crust led to fertilization (enrichment with HFSE and alkalis) of the lithosphere before partial melting. Fractional crystallization coupled with less considerable crustal assimilation are the main magmatic processes during the evolution of these rocks. The transition from subduction to post-collisional setting was accompanied by crustal uplifting, thickening and extensional collapse of ANS continental crust that caused emplacement of large masses of A-type granites in the northern ANS. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geological map of the central Eastern Desert of Egypt exhibiting the distribution of granitoids and other rock units of the basement rocks adjacent to the studied area (modified after [<a href="#B10-minerals-14-00806" class="html-bibr">10</a>]). (<b>b</b>) Geological map of Mount Abu Kibash and Tulayah in the central Eastern Desert of Egypt.</p>
Full article ">Figure 2
<p>Field photographs of Mount Abu Kibash and Tulayah granitic rocks in the central Eastern Desert of Egypt. (<b>a</b>) Rounded mafic xenolith hosted in granodiorite. (<b>b</b>) High relief syenogranite at Mount Abu Kibash.</p>
Full article ">Figure 3
<p>Photomicrographs and back-scattered electron (BSE) images showing the main petrographic features of Mount Abu Kibash and Tulayah granites: (<b>a</b>) aggregations of biotite (Bt), amphibole (Amph), plagioclase (Plg), K-feldspar (Kfs), and quartz (Qtz) reflecting a typical hypidiomorphic texture of the studied granodiorite; (<b>b</b>) a close up BSE image within granodiorite showing the main mineral phases with the occurrence of ilmenite (Ilm) as an inclusion in amphiboles, (<b>c</b>) the occurrence of coarse- to medium- grained Bt, Plg, Kfs, and Qz minerals with hyidiomorphic texture in syenogranites; (<b>d</b>) BSE image showing the occurrence of magmatic biotite and the mutual relation between the mineral phases in syenogranites, (<b>e</b>) the occurrence of muscovite (Mus) in some syenogranite samples and forming with other minerals a hypidiomorphic texture (<b>f</b>) BSE image of a large ilmenite crystal sharing boundaries with muscovite and quartz in syenogranites.</p>
Full article ">Figure 4
<p>Mineral chemistry of silicate minerals in the studied granites. (<b>a</b>) Ab-An-Or ternary classification diagram indicating feldspar composition. (<b>b</b>) Classification diagram of biotite [<a href="#B13-minerals-14-00806" class="html-bibr">13</a>]. (<b>c</b>) (FeOt + MnO) − 10 * TiO<sub>2</sub> − MgO ternary diagram discriminating between primary, re-equilibrated, and secondary biotite [<a href="#B14-minerals-14-00806" class="html-bibr">14</a>]. (<b>d</b>) Mg–Ti–Na ternary diagrams discriminating between primary and secondary muscovite [<a href="#B15-minerals-14-00806" class="html-bibr">15</a>]. (<b>e</b>) Si vs. Mg/Mg + Fe + 2 binary diagram for amphiboles nomenclature [<a href="#B16-minerals-14-00806" class="html-bibr">16</a>]. (<b>f</b>) Ti vs. (Na + K) discrimination diagram of the studied amphiboles [<a href="#B17-minerals-14-00806" class="html-bibr">17</a>].</p>
Full article ">Figure 5
<p>Whole-rock chemistry of Mount Abu Kibash and Tulayah granites. (<b>a</b>) TAS classification diagram [<a href="#B19-minerals-14-00806" class="html-bibr">19</a>]. (<b>b</b>) Na<sub>2</sub>O–K<sub>2</sub>O–CaO ternary diagram of Egyptian granitic rocks [<a href="#B18-minerals-14-00806" class="html-bibr">18</a>]. Trondhjemites and calc-alkaline fields are after Barker and Arth [<a href="#B20-minerals-14-00806" class="html-bibr">20</a>].</p>
Full article ">Figure 6
<p>Harker variation diagrams of selected major oxides (TiO<sub>2</sub>, (<b>a</b>); Al<sub>2</sub>O<sub>3</sub>, (<b>b</b>); MgO, (<b>c</b>); K<sub>2</sub>O, (<b>d</b>); P<sub>2</sub>O<sub>5</sub>, (<b>e</b>); Fe<sub>2</sub>O<sub>3</sub><sup>t</sup>, (<b>f</b>) and trace elements (Rb, (<b>g</b>); Ba, (<b>h</b>); Sr, (<b>i</b>); Zr, (<b>j</b>); Y, (<b>k</b>); Sc, (<b>l</b>); V, (<b>m</b>); Cr, (<b>n</b>); Ni, (<b>o</b>) vs. SiO<sub>2</sub>. Fields in K<sub>2</sub>O vs. SiO<sub>2</sub> (<b>d</b>) after Rickwood [<a href="#B21-minerals-14-00806" class="html-bibr">21</a>]. Upper crust values are after Rudnick and Gao [<a href="#B22-minerals-14-00806" class="html-bibr">22</a>].</p>
Full article ">Figure 7
<p>Bulk-rock chemistry of Mount Abu Kibash and Tulayah granites. (<b>a</b>,<b>c</b>) Chondrite-normalized REE patterns of the investigated granodiorites and syenogranites. (<b>b</b>,<b>d</b>) Multi-element-normalized diagram of the studied granodiorites and syenogranites. Chondrite and primitive mantle normalization values and chondrite values are from Sun and McDonough [<a href="#B23-minerals-14-00806" class="html-bibr">23</a>]. Fields of Homrit Waggat A- and I-type granites [<a href="#B24-minerals-14-00806" class="html-bibr">24</a>], Gabal El-Ineigi A-type granites [<a href="#B26-minerals-14-00806" class="html-bibr">26</a>], and Qianlishan A2-type granites [<a href="#B25-minerals-14-00806" class="html-bibr">25</a>] are used for comparison.</p>
Full article ">Figure 8
<p>Crystallization conditions (P–T–<span class="html-italic">fO</span><sub>2</sub>) of the studied granites. (<b>a</b>) Ab–An–Or thermometry diagram [<a href="#B28-minerals-14-00806" class="html-bibr">28</a>]. (<b>b</b>) Ti vs. Mg/(Mg + Fe) for the analyzed biotite showing temperature isotherms [<a href="#B30-minerals-14-00806" class="html-bibr">30</a>]. (<b>c</b>) Ab–Qz–Or ternary diagram for the studied granitic rocks [<a href="#B31-minerals-14-00806" class="html-bibr">31</a>]. (<b>d</b>) Fe/(Fe + Mg) vs. AlIV + AlVI for the analyzed biotite. Ilmenite and magnetite series after Anderson et al. [<a href="#B32-minerals-14-00806" class="html-bibr">32</a>].</p>
Full article ">Figure 9
<p>Bulk-rock chemistry of the studied granites showing the role of contamination and fractional crystallization in their evolution. (<b>a</b>) Rb vs. K/Rb diagram [<a href="#B37-minerals-14-00806" class="html-bibr">37</a>]. (<b>b</b>) Zr vs. Th/Nb variation diagrams showing fractional crystallization (FC), assimilation fractional crystallization (AFC), and bulk assimilation (BA) trends [<a href="#B39-minerals-14-00806" class="html-bibr">39</a>]. (<b>c</b>) SiO<sub>2</sub> vs. CaO/Na<sub>2</sub>O diagram with AFC trend. (<b>d</b>) Zr vs. TiO<sub>2</sub> diagram for the studied granites.</p>
Full article ">Figure 10
<p>Bulk-rock chemistry showing different types and sources of the studied granites. (<b>a</b>,<b>b</b>) 10<sup>4</sup> × Ga/Al against Nb and K<sub>2</sub>O/MgO for distinguishing between I, S, M and A-type granites [<a href="#B40-minerals-14-00806" class="html-bibr">40</a>]. (<b>c</b>) Al<sub>2</sub>O<sub>3</sub>/(FeO<sup>t</sup> + MgO) − 3*CaO − 5*(K<sub>2</sub>O/Na<sub>2</sub>O) ternary diagram [<a href="#B41-minerals-14-00806" class="html-bibr">41</a>]. (<b>d</b>) Rb vs. K<sub>2</sub>O diagram [<a href="#B44-minerals-14-00806" class="html-bibr">44</a>].</p>
Full article ">Figure 11
<p>Bulk-rock chemistry of the studied granites. (<b>a</b>) SiO<sub>2</sub> vs. (Na<sub>2</sub>O + K<sub>2</sub>O) − CaO discrimination diagram [<a href="#B62-minerals-14-00806" class="html-bibr">62</a>]. (<b>b</b>) 100*(MgO + FeO<sup>t</sup> + TiO<sub>2</sub>)/SiO<sub>2</sub> vs. molar (Al<sub>2</sub>O<sub>3</sub> + CaO)/(FeO + Na<sub>2</sub>O + K<sub>2</sub>O) discrimination diagram for distinguishing between different types of granitic magma [<a href="#B61-minerals-14-00806" class="html-bibr">61</a>]. (<b>c</b>) FeO<sup>t</sup>/(FeO<sup>t</sup> + MgO) vs. SiO<sub>2</sub> [<a href="#B62-minerals-14-00806" class="html-bibr">62</a>]. (<b>d</b>) Molar Al<sub>2</sub>O<sub>3</sub>/(Na<sub>2</sub>O + K<sub>2</sub>O) vs. Al<sub>2</sub>O<sub>3</sub>/(CaO + Na<sub>2</sub>O+ K<sub>2</sub>O) for the studied granites [<a href="#B68-minerals-14-00806" class="html-bibr">68</a>]. (<b>e</b>) Y + Nb vs. Rb tectonic discrimination diagram of Pearce et al. [<a href="#B66-minerals-14-00806" class="html-bibr">66</a>]; the post-collisional granite field is from Pearce [<a href="#B65-minerals-14-00806" class="html-bibr">65</a>]. The A-type granite field in previous diagrams is after Whalen et al. [<a href="#B40-minerals-14-00806" class="html-bibr">40</a>]; Eastern Desert (ED) A2-type and I-type granites are after Azer et al. [<a href="#B24-minerals-14-00806" class="html-bibr">24</a>] and Farahat et al. [<a href="#B9-minerals-14-00806" class="html-bibr">9</a>]; Qianlishan A2-type granite field is after Chen et al. [<a href="#B25-minerals-14-00806" class="html-bibr">25</a>]. (<b>f</b>) Hf-Rb/30−3*Ta tectonic discrimination diagram after Harris et al. [<a href="#B67-minerals-14-00806" class="html-bibr">67</a>]. (<b>g</b>) SiO<sub>2</sub> vs. FeO<sup>t</sup>/(FeO<sup>t</sup> + MgO) discrimination diagram [<a href="#B68-minerals-14-00806" class="html-bibr">68</a>]. (<b>h</b>) Y-Nb−3Ga ternary plot [<a href="#B69-minerals-14-00806" class="html-bibr">69</a>]; A1 = A-type granitoids with an ocean island basalt-type source; A2 = A-type granitoids with crust-derived magma. (<b>i</b>) Na<sub>2</sub>O-K<sub>2</sub>O-CaO ternary discrimination diagram showing different types of Egyptian granitoids [<a href="#B26-minerals-14-00806" class="html-bibr">26</a>], where I = old calc-alkaline phase, II = early subphase of young calc-alkaline phase, and III = late subphase of young calc-alkaline phase.</p>
Full article ">Figure 12
<p>Mineral chemistry of amphiboles and biotite from the studied granites. (<b>a</b>) TiO<sub>2</sub> vs Al<sub>2</sub>O<sub>3</sub> for the studied primary amphiboles [<a href="#B72-minerals-14-00806" class="html-bibr">72</a>]. (<b>b</b>) Cations of Al<sup>iv</sup> vs. K binary diagram discriminating between alkaline and calc-alkaline magma [<a href="#B71-minerals-14-00806" class="html-bibr">71</a>]. (<b>c</b>) SiO<sub>2</sub> vs. Na<sub>2</sub>O binary diagram of amphiboles [<a href="#B73-minerals-14-00806" class="html-bibr">73</a>] (I-Amph = within-plate amphiboles; S-Amph = Supra-subduction amphiboles). (<b>d</b>) FeO-MgO-Al<sub>2</sub>O<sub>3</sub> ternary discrimination diagram of biotite [<a href="#B70-minerals-14-00806" class="html-bibr">70</a>].</p>
Full article ">Figure 13
<p>Sketch showing emplacement of Mount Abu Kibash and Tulayah granitic intrusions within the central Eastern Desert of Egypt in different tectonic stages during the evolution of the ANS. (<b>a</b>) Generation of granodiorites (I-type granite) during the collisional stage in a subduction-related setting (active continental margin) and (<b>b</b>) emplacement of syenogranites (A-type granites) during the post-collisional stage.</p>
Full article ">
17 pages, 5617 KiB  
Article
Impact of Thermochemical Treatments on Electrical Conductivity of Donor-Doped Strontium Titanate Sr(Ln)TiO3 Ceramics
by Aleksandr Bamburov, Ekaterina Kravchenko and Aleksey A. Yaremchenko
Materials 2024, 17(15), 3876; https://doi.org/10.3390/ma17153876 - 5 Aug 2024
Viewed by 592
Abstract
The remarkable stability, suitable thermomechanical characteristics, and acceptable electrical properties of donor-doped strontium titanates make them attractive materials for fuel electrodes, interconnects, and supports of solid oxide fuel and electrolysis cells (SOFC/SOEC). The present study addresses the impact of processing and thermochemical treatment [...] Read more.
The remarkable stability, suitable thermomechanical characteristics, and acceptable electrical properties of donor-doped strontium titanates make them attractive materials for fuel electrodes, interconnects, and supports of solid oxide fuel and electrolysis cells (SOFC/SOEC). The present study addresses the impact of processing and thermochemical treatment conditions on the electrical conductivity of SrTiO3-derived ceramics with moderate acceptor-type substitution in a strontium sublattice. A-site-deficient Sr0.85La0.10TiO3−δ and cation-stoichiometric Sr0.85Pr0.15TiO3+δ ceramics with varying microstructures and levels of reduction have been prepared and characterized by XRD, SEM, TGA, and electrical conductivity measurements under reducing conditions. The analysis of the collected data suggested that the reduction process of dense donor-doped SrTiO3 ceramics is limited by sluggish oxygen diffusion in the crystal lattice even at temperatures as high as 1300 °C. A higher degree of reduction and higher electrical conductivity can be obtained for porous structures under similar thermochemical treatment conditions. Metallic-like conductivity in dense reduced Sr0.85La0.10TiO3−δ corresponds to the state quenched from the processing temperature and is proportional to the concentration of Ti3+ in the lattice. Due to poor oxygen diffusivity in the bulk, dense Sr0.85La0.10TiO3−δ ceramics remain redox inactive and maintain a high level of conductivity under reducing conditions at temperatures below 1000 °C. While the behavior and properties of dense reduced Sr0.85Pr0.15TiO3+δ ceramics with a large grain size (10–40 µm) were found to be similar, decreasing grain size down to 1–3 µm results in an increasing role of resistive grain boundaries which, regardless of the degree of reduction, determine the semiconducting behavior and lower total electrical conductivity of fine-grained Sr0.85Pr0.15TiO3+δ ceramics. Oxidized porous Sr0.85Pr0.15TiO3+δ ceramics exhibit faster kinetics of reduction compared to the Sr0.85La0.10TiO3−δ counterpart at temperatures below 1000 °C, whereas equilibration kinetics of porous Sr0.85La0.10TiO3−δ structures can be facilitated by reductive pre-treatments at elevated temperatures. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of literature data on the electrical conductivity of Sr<sub>0.90</sub>La<sub>0.10</sub>TiO<sub>3±δ</sub> ceramics in reducing H<sub>2</sub>-based atmospheres. Sources: Moos [<a href="#B37-materials-17-03876" class="html-bibr">37</a>], Marina [<a href="#B36-materials-17-03876" class="html-bibr">36</a>], Yaremchenko [<a href="#B35-materials-17-03876" class="html-bibr">35</a>], Li [<a href="#B38-materials-17-03876" class="html-bibr">38</a>], Lv [<a href="#B39-materials-17-03876" class="html-bibr">39</a>], Wang [<a href="#B40-materials-17-03876" class="html-bibr">40</a>], Niwa [<a href="#B41-materials-17-03876" class="html-bibr">41</a>], Park [<a href="#B42-materials-17-03876" class="html-bibr">42</a>], and Hashimoto [<a href="#B43-materials-17-03876" class="html-bibr">43</a>]. See <a href="#materials-17-03876-t001" class="html-table">Table 1</a> for the experimental details.</p>
Full article ">Figure 2
<p>Examples of XRD patterns of S85L10 and S85P15 samples. Note that S85L10-1300 indicates the powder as synthesized in air at 1300 °C, and S85P15-1350 corresponds to the ceramic sample sintered in air at 1350 °C. The notations of other samples are listed in <a href="#materials-17-03876-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>SEM images of fractured cross-sections of S85L10 ceramics prepared under different conditions.</p>
Full article ">Figure 4
<p>SEM images of fractured cross-sections of S85P15 ceramics prepared under different conditions.</p>
Full article ">Figure 5
<p>Electrical conductivity of S85L10 ceramics as a function of (<b>A</b>) temperature in 10%H<sub>2</sub>-N<sub>2</sub> atmosphere and (<b>B</b>) oxygen partial pressure under reducing conditions at 900 °C.</p>
Full article ">Figure 6
<p>(<b>A</b>) Example of thermogravimetric data recorded for powdered S85L10 (Sr<sub>0.85</sub>La<sub>0.10</sub>TiO<sub>3−δ</sub>) ceramics on cooling in 10%H<sub>2</sub>-N<sub>2</sub> atmosphere and subsequent heating/cooling cycle in air. (<b>B</b>) Electrical conductivity of dense S85L10 ceramics vs. fraction of Ti<sup>3+</sup> cations in the titanium sublattice.</p>
Full article ">Figure 7
<p>Relaxation of electrical conductivity of (<b>A</b>) oxidized S85L10-1320 ceramics on reduction and (<b>B</b>) reduced S85L10-1320-R-1300 ceramics on redox cycling at 900 °C.</p>
Full article ">Figure 8
<p>Electrical conductivity of S85P15 ceramics as a function of (<b>A</b>) temperature in 10%H<sub>2</sub>-N<sub>2</sub> atmosphere and (<b>B</b>) oxygen partial pressure under reducing conditions at 900 °C.</p>
Full article ">Figure 9
<p>Changes in oxygen nonstoichiometry of S85P15 (Sr<sub>0.85</sub>Pr<sub>0.15</sub>TiO<sub>3+δ</sub>) ceramics on oxidation in air estimated from the thermogravimetric data.</p>
Full article ">Figure 10
<p>Relaxation of electrical conductivity of porous (<b>A</b>) S85P15-1350 and (<b>B</b>) S85P15-H-1350 samples on redox cycling at 850 °C. S85P15-1350 denotes the sample sintered in air at 1350 °C for 10 h.</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 486
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 ">
Back to TopTop