[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
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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (3,196)

Search Parameters:
Keywords = surface texture

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
42 pages, 113259 KiB  
Article
Hypogene Alteration of Base–Metal Mineralization at the Václav Vein (Březové Hory Deposit, Příbram, Czech Republic): The Result of Recurrent Infiltration of Oxidized Fluids
by Zdeněk Dolníček, Jiří Sejkora and Pavel Škácha
Minerals 2024, 14(10), 1038; https://doi.org/10.3390/min14101038 (registering DOI) - 17 Oct 2024
Abstract
The Václav vein (Březové Hory deposit, Příbram ore area, Czech Republic) is a base–metal vein containing minor Cu-Zn-Pb-Ag-Sb sulfidic mineralization in a usually hematitized gangue. A detailed mineralogical study using an electron microprobe revealed a complicated multistage evolution of the vein. Early siderite [...] Read more.
The Václav vein (Březové Hory deposit, Příbram ore area, Czech Republic) is a base–metal vein containing minor Cu-Zn-Pb-Ag-Sb sulfidic mineralization in a usually hematitized gangue. A detailed mineralogical study using an electron microprobe revealed a complicated multistage evolution of the vein. Early siderite and Fe-rich dolomite were strongly replaced by assemblages of hematite+rhodochrosite and hematite+kutnohorite/Mn-rich dolomite, respectively. In addition, siderite also experienced strong silicification. These changes were associated with the dissolution of associated sulfides (sphalerite, galena). The following portion of the vein contains low-Mn dolomite and calcite gangue with Zn-rich chlorite, wittichenite, tetrahedrite-group minerals, chalcopyrite, bornite, and djurleite, again showing common replacement textures in case of sulfides. The latest stage was characterized by the input of Ag and Hg, giving rise to Ag-Cu sulfides, native silver (partly Hg-rich), balkanite, and (meta)cinnabar. We explain the formation of hematite-bearing oxidized assemblages at the expense of pre-existing “normal” Příbram mineralization due to repeated episodic infiltration of oxygenated surface waters during the vein evolution. Episodic mixing of ore fluids with surface waters was suggested from previous stable isotope and fluid inclusion studies in the Příbram ore area. Our mineralogical study thus strengthens this genetic scenario, illustrates the dynamics of fluid movement during the evolution of a distinct ore vein structure, and shows that the low content of ore minerals cannot be necessarily a primary feature of a vein. Full article
(This article belongs to the Special Issue Mineralogy and Geochemistry of Polymetallic Ore Deposits)
Show Figures

Figure 1

Figure 1
<p>The youngest mineral assemblage of a drusy cavity in sample P1N 9430. (<b>a</b>) A top view across the whole drusy cavity. (<b>b</b>) Calcite crystals with acicular ore aggregates. (<b>c</b>) Smooth and finely wrinkled surface of acicular ore aggregates. (<b>d</b>) Acicular ore aggregates enclosed in a transparent crystal of calcite.</p>
Full article ">Figure 2
<p>Geological position of the Březové Hory deposit in the Příbram ore area (modified from [<a href="#B1-minerals-14-01038" class="html-bibr">1</a>]). BHD—Březové Hory base–metal district, PUD—Příbram uranium district. Positions of some other sites mentioned in the text are also indicated.</p>
Full article ">Figure 3
<p>Position of the Václav vein in geological cross-section through the Březové Hory ore district (modified from [<a href="#B9-minerals-14-01038" class="html-bibr">9</a>]). The Anna shaft is situated north of the Prokop shaft, out of the section line.</p>
Full article ">Figure 4
<p>The macroscopic appearance of the sample P1N 9430 with marked zones A–E and sites, from which samples for preparation of polished sections were cut off. The left part of the figure illustrates the distribution of selected mineral phases. Sample width is 4.5 cm.</p>
Full article ">Figure 5
<p>Mineral assemblage and textures of the Václav vein in the BSE images. (<b>a</b>) A slightly zoned relic of siderite replaced by surrounding quartz, carbonates of the dolomite-kutnohorite series, hematite, and tetrahedrite. Right part is wall rock. Zone A, sample VA-1. (<b>b</b>) Relic of siderite strongly replaced by hematite, rhodochrosite, and zoned carbonates of the dolomite series. Zone A, sample VA-2. (<b>c</b>) Relic of siderite rimmed by rhodochrosite and hematite. Zone A, sample VA-2. (<b>d</b>) Euhedral rhodochrosite crystal enclosed in hematite in the proximity of a relic of siderite, which is strongly replaced by the rhodochrosite rim and carbonates of the dolomite group. Zone A, sample VA-2. (<b>e</b>) Boundary between Zone A (hematite-rich on the left) and Zone B (hematite-poor on the right) separated by quartz crystals. Replacement of Do-I by Do-II is observed in the right part. Sample VA-5. (<b>f</b>) Detailed view on replacement of Do-I by zoned Do-II containing inclusions of hematite. Zone B, sample VA-5.</p>
Full article ">Figure 6
<p>Variations in the chemical composition of siderite, rhodochrosite, and calcite from the Václav vein in comparison with published data. (<b>a</b>) Siderite and rhodochrosite in the classification diagram by [<a href="#B27-minerals-14-01038" class="html-bibr">27</a>]. (<b>b</b>) Fe vs. Mn and Fe vs. Mg plots for calcite. Comparative data for other deposits of the Příbram uranium and base metal district are from [<a href="#B5-minerals-14-01038" class="html-bibr">5</a>,<a href="#B6-minerals-14-01038" class="html-bibr">6</a>,<a href="#B7-minerals-14-01038" class="html-bibr">7</a>,<a href="#B28-minerals-14-01038" class="html-bibr">28</a>].</p>
Full article ">Figure 7
<p>Variations in the chemical composition of carbonates of the dolomite-ankerite series from the Václav vein in comparison with published data. (<b>a</b>) All data in the classification diagram by [<a href="#B27-minerals-14-01038" class="html-bibr">27</a>]. (<b>b</b>) Data sorted according to Zones. (<b>c</b>) Data arbitrarily grouped according to compositional similarities. Comparative data for other deposits of the Příbram uranium and base metal district are from [<a href="#B5-minerals-14-01038" class="html-bibr">5</a>,<a href="#B6-minerals-14-01038" class="html-bibr">6</a>,<a href="#B7-minerals-14-01038" class="html-bibr">7</a>,<a href="#B28-minerals-14-01038" class="html-bibr">28</a>]. PUD–average dolomite from the Příbram uranium and base–metal district according to wet-chemical analyses by [<a href="#B29-minerals-14-01038" class="html-bibr">29</a>].</p>
Full article ">Figure 8
<p>Mineral assemblage and textures of the Václav vein. (<b>a</b>) Contact between Do-II and Do-III dolomites. Note the corrosion of Do-II by Do-III just along the contact. Zone C, sample VA-7. (<b>b</b>) Zoned crystals of carbonates of the dolomite group (bright Do-II is overgrown by darker Do-III) growing over lenticular hematite crystals. Residual vug was later filled up by calcite with aggregates of tetrahedrite and chalcopyrite. Zone A, black domain, sample VA-6. (<b>c</b>) The latest Do-III dolomite crystals overgrown by chalcopyrite-bornite aggregates and calcite. Zone D, sample VA-9. (<b>d</b>) Aggregates of Zn-rich chlorite filling together with calcite residual cavities in the vein composed of euhedral lenticular hematite crystals, sulfidic aggregates, and Do-II+Do-III carbonates. Zone A, black domain, sample VA-6. (<b>e</b>) Two generations of hematite strongly differing in the quality of the polished surface. Fine-grained early hemispherical aggregates are poorly polished, whereas the latest hematite preceding crystallization of Do-II carbonate followed by sulfides is well polished. Sulfide aggregate is composed of sphalerite, chalcopyrite, tetrahedrite-(Zn) (Ttd-I), and an unknown reddish AgCu<sub>6</sub>Fe<sub>2</sub>S<sub>8</sub> phase. Zone A, sample VA-1. (<b>f</b>) The central area of Figure (<b>e</b>) in BSE image. Note the zonality of hematite and Do-II carbonate. Zone A, sample VA-1. Figure (<b>e</b>) is taken in plane-polarized reflected light, whereas the other pictures are BSE images.</p>
Full article ">Figure 9
<p>Variations in the chemical composition of chlorite from the Václav vein and comparison with published data. (<b>a</b>) A Fe-Mg-Zn plot. (<b>b</b>) The Ca vs. Si plot. The comparative data from the Jerusalem deposit (Příbram uranium district) are from [<a href="#B5-minerals-14-01038" class="html-bibr">5</a>].</p>
Full article ">Figure 10
<p>Mineral assemblage and textures of the Václav vein. (<b>a</b>) Concentric zonation of hematite. Zone A, sample VA-2. (<b>b</b>) Patchy zonation of hematite caused by variable Sb contents. The youngest part of Zone A, sample VA-3. (<b>c</b>) The strongly corroded cassiterite hosted by sphalerite (partly Cu,Sn-enriched) replaced by hematite+Do-II aggregate. Zone A, sample VA-4. (<b>d</b>) Finely porous and non-porous bornite and chalcopyrite. Note a bluish tint of a part of porous bornite. Sample Dy-817. (<b>e</b>) Finely porous and non-porous bornite and chalcopyrite, with a short veinlet of tetrahedrite. Note the compositional homogeneity of bornite. Sample Dy-817. (<b>f</b>) An acicular polymineral aggregate composed of bornite, covellite, and Ag-Cu sulfides (Ag-Cu-S) with thick symmetrical rims of chalcopyrite I (Cpy-I). Zone E, sample Dy-973. Figures (<b>d</b>,<b>f</b>) are taken in plane-polarized reflected light, whereas the other pictures are BSE images.</p>
Full article ">Figure 11
<p>Variations in the chemical composition of hematite from the Václav vein. (<b>a</b>) The Si vs. Al plot. (<b>b</b>) The Si vs. Sb plot. (<b>c</b>) The Me<sup>2+</sup> vs. Sb plot. (<b>d</b>) The Pb vs. Sb plot.</p>
Full article ">Figure 12
<p>Mineral assemblage and textures of the Václav vein. (<b>a</b>) A crust formed by chalcopyrite (Cpy-I), partly filled and enclosed by bornite and overgrown by a tetrahedrite (Ttd-III) crystal. Part of the pores in bornite was filled by covellite and Ag-Cu sulfides. Bornite contains ribbons of chalcopyrite (Cpy-II). Zone E, sample Dy-973. (<b>b</b>) Three morphological forms of chalcopyrite, crust (Cpy-I), ribbon (Cpy-II), and symplectite with mckinstryite (Symplectite), hosted by bornite with late rims and fillings of covellite and Ag-Cu(-Hg)-S phases. The black rectangle shows the area of <a href="#minerals-14-01038-f012" class="html-fig">Figure 12</a>c. Zone E, sample Dy-973. (<b>c</b>) BSE detail of the central part of Figure (c) showing the nature of Ag-Cu(-Hg)-S phases: fine intergrowths of stromeyerite and mckinstryite are partly overgrown by balkanite. (<b>d</b>) Sphalerite grains are rimmed by bornite and chalcopyrite. Note the intense corrosion of both earlier sulfide phases by later ones. Zone A, sample VA-1. (<b>e</b>) Sphalerite rimmed by bornite, tetrahedrite (Ttd-I), and two generations of chalcopyrite differing in porosity. Note early porous chalcopyrite Cpy-I is replaced by Ttd-I. Zone D, sample VA-9. (<b>f</b>) Bright Ag-enriched zone in chalcopyrite. Zone D, sample VA-8. Figures (<b>c</b>,<b>f</b>) are BSE images; the other pictures are taken in plane-polarized reflected light.</p>
Full article ">Figure 13
<p>Variations in the chemical composition of some ore minerals from the Václav vein in comparison with published data. (<b>a</b>) Graph Ag versus Sb for chalcopyrite. (<b>b</b>) Graph Ag versus Cu for bornite. (<b>c</b>) Graph Sn versus Cu for sphalerite. (<b>d</b>) Graph Fe versus Cd for sphalerite. (<b>e</b>) Graph Ag versus Cu+Fe+Cd for mckinstryite. (<b>f</b>) Graph Ag versus Cu for balkanite. Comparative data for sphalerite from the Háje deposit (Příbram uranium district) are from [<a href="#B7-minerals-14-01038" class="html-bibr">7</a>], for mckinstryite from Milín from [<a href="#B31-minerals-14-01038" class="html-bibr">31</a>], for other mckinstryite data from [<a href="#B32-minerals-14-01038" class="html-bibr">32</a>], for danielsite from [<a href="#B33-minerals-14-01038" class="html-bibr">33</a>], and for published balkanite data from [<a href="#B5-minerals-14-01038" class="html-bibr">5</a>,<a href="#B34-minerals-14-01038" class="html-bibr">34</a>,<a href="#B35-minerals-14-01038" class="html-bibr">35</a>,<a href="#B36-minerals-14-01038" class="html-bibr">36</a>,<a href="#B37-minerals-14-01038" class="html-bibr">37</a>].</p>
Full article ">Figure 14
<p>Mineral assemblage and textures of the Václav vein. (<b>a</b>) Irregular aggregate of djurleite partly replaced by bornite, which is rimmed by a non-continuous zone of chalcopyrite and small grains of Ag-Cu-Hg-S phases. Note the abundant ribbons of chalcopyrite in the outer part of bornite adjacent to the chalcopyrite rim. Sample Dy-816. (<b>b</b>) Sphalerite rimmed by bornite (with brighter Ag-enriched domains) and then by chalcopyrite. Late microfractures contain Ag-Cu-S phases. Zone D, sample VA-9. (<b>c</b>) The brighter Cu,Sn-enriched domains in sphalerite in the vicinity of strongly corroded grains of cassiterite. Zone A, sample VA-4. (<b>d</b>) Sphalerite with brighter Cd-enriched domains (in the lower part of the grain), rimmed by chalcopyrite and enclosing grains of wittichenite and Bi-enriched tetrahedrite. Zone A, sample VA-4. Inset–Oscillatory zoning of a sphalerite grain due to changing Cd contents. Zone C, sample VA-7. (<b>e</b>) Mn-enriched sphalerite and hematite in the residual cavity in Mn-rich dolomite Do-II replacing Fe-rich dolomite Do-I. Zone B, sample VA-5. (<b>f</b>) Tetrahedrite Ttd-I is cut by veinlets of bornite and chalcopyrite Cpy-II. Zone D, sample VA-9. Figures (<b>a</b>,<b>f</b>) are taken in plane-polarized reflected light; the other pictures are BSE images.</p>
Full article ">Figure 15
<p>Mineral assemblage and textures of the Václav vein in BSE images. (<b>a</b>) Zonality of a tetrahedrite Ttd-I aggregate is largely caused by Fe-Zn substitution and also, exceptionally, by high Cd (arrowed). The Ttd-I is cut by veinlets of Bi-bearing Ttd-II and a narrow overgrowth of Ttd-III is observed in the lower part of the photograph. Zone D, sample VA-9. (<b>b</b>) Oscillatory zoned Bi-bearing Ttd-II rimming a grain of chalcopyrite. Zone A, sample VA-6. (<b>c</b>) Patchy zoning of Ttd-II. The brightest domain already corresponds to annivite-(Zn). Zone A, sample VA-6. (<b>d</b>) Zoned Bi-bearing tetrahedrite Ttd-II rimming and cutting tetrahedrite Ttd-I. The brightest domain corresponds to <span class="html-italic">annivite-</span>(<span class="html-italic">Cu</span>). Surrounding sphalerite encloses wittichenite. Zone A, sample VA-3. (<b>e</b>) Nature of Ag-Cu sulfides in chalcopyrite-hosted “acicular” polymineral aggregate from <a href="#minerals-14-01038-f010" class="html-fig">Figure 10</a>f: small domains of jalpaite are hosted by the mckinstryite matrix. Zone E, sample Dy-973. (<b>f</b>) A finely porous aggregate of Hg-absent native silver partly rimmed by stromeyerite. Zone E, sample Dy-973.</p>
Full article ">Figure 16
<p>Variations in the chemical composition of tetrahedrite-group minerals from the Václav vein (data points) in comparison with published data (outlined). (<b>a</b>) Graph Fe-Zn-Cd. (<b>b</b>) Graph As-Sb-Bi. Data in at. %. Comparative data for Jáchymov and Hřebečná sites are from [<a href="#B41-minerals-14-01038" class="html-bibr">41</a>]. Notably, data from the Příbram ore area are not visualized as they exhibit Fe-Zn and Sb-As substitutions only.</p>
Full article ">Figure 17
<p>Mineral assemblage and textures of the Václav vein in BSE images. (<b>a</b>) Balkanite-like phase rimming grains of probable (meta)cinnabar. Zone A, sample VA-1. (<b>b</b>) Grains of unspecified Ag-Cu-Hg-S phases (white) with variable compositions growing on a chalcopyrite finger-like aggregate. Zone E, sample Dy-973. (<b>c</b>) Individual grains of likely galena (GA) and unspecified Ag-Cu-Hg-S phases with variable compositions growing on a djurleite-bornite-chalcopyrite aggregate. Sample Dy-816. (<b>d</b>) An unknown (Cu,Ag)<sub>4</sub>FeS<sub>4</sub> phase associated with covellite in a sphalerite-chalcopyrite-hematite aggregate. Zone A, sample VA-3.</p>
Full article ">Figure 18
<p>Variations in the chemical composition of some ore minerals from the Václav vein in comparison with published data. (<b>a</b>) Graph Cu/Ag versus Hg for balkanite. (<b>b</b>) Graph Ag versus Cu or Hg for balkanite and possible intergrowths of Ag-Cu-Hg phases. (<b>c</b>) Graph Hg versus Cu for balkanite and possible intergrowths of Ag-Cu-Hg phases. (<b>d</b>) Graph Hg versus Ag+Cu+Fe for balkanite and possible intergrowths of Ag-Cu-Hg phases. Comparative published balkanite data are from [<a href="#B5-minerals-14-01038" class="html-bibr">5</a>,<a href="#B34-minerals-14-01038" class="html-bibr">34</a>,<a href="#B35-minerals-14-01038" class="html-bibr">35</a>,<a href="#B36-minerals-14-01038" class="html-bibr">36</a>,<a href="#B37-minerals-14-01038" class="html-bibr">37</a>], and data for danielsite are from [<a href="#B33-minerals-14-01038" class="html-bibr">33</a>].</p>
Full article ">Figure 19
<p>A sketch showing the interpreted textural evolution of the late ore assemblage from a drusy cavity of the sample P1N 9430. The crystallization of an acicular mineral (<b>a</b>) was followed by the deposition of a continuous layer of early chalcopyrite Cpy-I on its crystals (<b>b</b>). Then, the dissolution of acicular mineral took place (<b>c</b>), followed by the crystallization of bornite inside and outside of Cpy-I crusts (<b>d</b>). The crystallization of late chalcopyrite Cpy-II and tetrahedrite Ttd-III (<b>e</b>) was followed by minor fracturing of early ores and partial healing of the residual porosity by the latest ore minerals including covellite and Ag-Cu(-Hg) sulfides (<b>f</b>).</p>
Full article ">Figure 20
<p>The simplified paragenetic scheme of the Václav vein. Note that the position of some mineral phases is questionable (marked by ?); more problematic phases are missing.</p>
Full article ">Figure 21
<p>The interpreted scenario of the origin of the studied mineralization from the Václav vein. (<b>a</b>) Early stage characterized by the escape of “deep” fluids. (<b>b</b>) Late stage involving the circulation of basinal fluids—Scenario I. (<b>c</b>) Late stage involving the circulation of basinal fluids–Scenario II. Arrows characterize the direction of fluid movement.</p>
Full article ">
16 pages, 1844 KiB  
Article
Innovative Pathogen Reduction in Exported Sea Bass Through Atmospheric Cold Plasma Technology
by Şehnaz Yasemin Tosun, Sehban Kartal, Tamer Akan, Sühendan Mol, Serap Coşansu, Didem Üçok, Şafak Ulusoy, Hande Doğruyol and Kamil Bostan
Foods 2024, 13(20), 3290; https://doi.org/10.3390/foods13203290 (registering DOI) - 17 Oct 2024
Viewed by 136
Abstract
The safety of sea bass is critical for the global food trade. This study evaluated the effectiveness of atmospheric cold plasma in reducing food safety risks posed by Salmonella Enteritidis and Listeria monocytogenes, which can contaminate sea bass post harvest. Cold plasma [...] Read more.
The safety of sea bass is critical for the global food trade. This study evaluated the effectiveness of atmospheric cold plasma in reducing food safety risks posed by Salmonella Enteritidis and Listeria monocytogenes, which can contaminate sea bass post harvest. Cold plasma was applied to inoculated sea bass for 2 to 18 min, achieving a maximum reduction of 1.43 log CFU/g for S. Enteritidis and 0.80 log CFU/g for L. monocytogenes at 18 min. Longer treatments resulted in greater reductions; however, odor and taste quality declined to a below average quality in samples treated for 12 min or longer. Plasma treatment did not significantly alter the color, texture, or water activity (aw) of the fish. Higher levels of thiobarbituric acid reactive substances (TBARSs) were observed with increased exposure times. Cold plasma was also tested in vitro on S. Enteritidis and L. monocytogenes on agar surfaces. A 4 min treatment eliminated the initial loads of S. Enteritidis (2.71 log CFU) and L. monocytogenes (2.98 log CFU). The findings highlight the potential of cold plasma in enhancing the safety of naturally contaminated fish. Cold plasma represents a promising technology for improving food safety in the global fish trade and continues to be a significant area of research in food science. Full article
(This article belongs to the Section Food Quality and Safety)
Show Figures

Figure 1

Figure 1
<p>Original atmospheric cold plasma equipment ((A) power supply; (B) plasma generation cite; (C) glass Petri dish lid; (D) copper plate; (E) copper wire; (F) sample site; (G) cold plasma).</p>
Full article ">Figure 2
<p>In vitro reduction in <span class="html-italic">S.</span> Enteritidis and <span class="html-italic">L. monocytogenes</span> by atmospheric cold plasma (a, b: different letters show significant differences between reduction rates, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Reduction in <span class="html-italic">S.</span> Enteritidis and <span class="html-italic">L. monocytogenes</span> on sea bass by atmospheric cold plasma (a–d: different letters show significant differences in reduction rates, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Sensory analysis of sea bass treated with atmospheric cold plasma for various durations (* decreases in odor and taste after 8 min and in overall acceptability after 10 min are significant, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Color change in plasma-treated sea bass compared to untreated samples (a; no significant differences between ΔE vales, <span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 6
<p>TBARS values of sea bass treated with atmospheric cold plasma (a–i: letters indicate the significant difference between treatments, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
25 pages, 10966 KiB  
Review
Bionic Strategies for Pump Anti-Cavitation: A Comprehensive Review
by Jian Li, Xing Zhou, Hongbo Zhao, Chengqi Mou, Long Meng, Liping Sun and Peijian Zhou
Energies 2024, 17(20), 5149; https://doi.org/10.3390/en17205149 - 16 Oct 2024
Viewed by 190
Abstract
The cavitation phenomenon presents a significant challenge in pump operation since the losses incurred by cavitation adversely impact pump performance. The many constraints of conventional anti-cavitation techniques have compelled researchers to explore biological processes for innovative alternatives. Consequently, the use of bionanotechnology for [...] Read more.
The cavitation phenomenon presents a significant challenge in pump operation since the losses incurred by cavitation adversely impact pump performance. The many constraints of conventional anti-cavitation techniques have compelled researchers to explore biological processes for innovative alternatives. Consequently, the use of bionanotechnology for anti-cavitation pumping has emerged as a prominent study domain. Despite the extensive publication of publications on biomimetic technology, research concerning the use of anti-cavitation in pumps remains scarce. This review comprehensively summarizes, for the first time, the advancements and applications of bionic structures, bionic surface texture design, and bionic materials in pump anti-cavitation, addressing critical aspects such as blade leading-edge bionic structures, bionic worm shells, microscopic bionic textures, and innovative bionic coatings. Bionic technology may significantly reduce cavitation erosion and improve pump performance by emulating natural biological structures. This research elucidates the creative contributions of biomimetic designs and their anti-cavitation effects, hence boosting the anti-cavitation performance of pumps. This work integrates practical requirements and anticipates future applications of bionic technology in pump anti-cavitation, offering a significant research direction and reference for scholars in this domain. Full article
(This article belongs to the Section K: State-of-the-Art Energy Related Technologies)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Schematic diagram of centrifugal pump cavitation process. (<b>b</b>) Actual evidence of wearing on impeller blades due to cavitation [<a href="#B15-energies-17-05149" class="html-bibr">15</a>].</p>
Full article ">Figure 2
<p>Damaged centrifugal pump impellers due to cavitation [<a href="#B16-energies-17-05149" class="html-bibr">16</a>,<a href="#B17-energies-17-05149" class="html-bibr">17</a>].</p>
Full article ">Figure 3
<p>Number of relative published works obtained from Web of Science.</p>
Full article ">Figure 4
<p>The outline of this review.</p>
Full article ">Figure 5
<p>Schematic representation of the humpback whale flipper leading-edge bulge [<a href="#B43-energies-17-05149" class="html-bibr">43</a>].</p>
Full article ">Figure 6
<p>Schematic diagram of centrifugal pump bionic vane [<a href="#B43-energies-17-05149" class="html-bibr">43</a>].</p>
Full article ">Figure 7
<p>Cavity volume in impeller channel at different NPSHa in the 5th rotation cycle [<a href="#B43-energies-17-05149" class="html-bibr">43</a>]. (<b>a</b>) Variation of vacuole volume development in the impeller channel at NPSHa = 0.697; (<b>b</b>) Variation of vacuole volume development in the impeller channel at NPSHa = 0.085; (<b>c</b>) Variation of vacuole volume development in the impeller channel at NPSHa = 0.054; (<b>d</b>) Variation of vacuole volume development in the impeller channel at NPSHa = 0.044.</p>
Full article ">Figure 8
<p>Bionic humpback whale flipper blade [<a href="#B44-energies-17-05149" class="html-bibr">44</a>]. (<b>a</b>) Flipper of humpback whale; (<b>b</b>) the head shape of bionic centrifugal pump.</p>
Full article ">Figure 9
<p>Bionic impeller [<a href="#B48-energies-17-05149" class="html-bibr">48</a>]. (<b>a</b>) Leading-edge tubercle structure; (<b>b</b>) Actual object.</p>
Full article ">Figure 10
<p>Definition of bionic leading-edge parameters [<a href="#B49-energies-17-05149" class="html-bibr">49</a>].</p>
Full article ">Figure 11
<p>Simulation of gas volume fraction of bionic leading-edge blade [<a href="#B49-energies-17-05149" class="html-bibr">49</a>].</p>
Full article ">Figure 12
<p>Blade contour lines [<a href="#B50-energies-17-05149" class="html-bibr">50</a>].</p>
Full article ">Figure 13
<p>Bionic leading-edge mechanism analysis [<a href="#B50-energies-17-05149" class="html-bibr">50</a>]. (<b>a</b>) Overall distribution of pressure; (<b>b</b>) Local distribution of streamline of streamline; (<b>c</b>) Streamline diagram.</p>
Full article ">Figure 14
<p>Bionic impeller modeling [<a href="#B51-energies-17-05149" class="html-bibr">51</a>].</p>
Full article ">Figure 15
<p>Spatial modes of the DMD under cavitation conditions [<a href="#B51-energies-17-05149" class="html-bibr">51</a>]. (<b>a</b>) <span class="html-italic">f</span><sub>s</sub>, (<b>b</b>) 2<span class="html-italic">f</span><sub>s</sub>, (<b>c</b>) 3<span class="html-italic">f</span><sub>s</sub>, and (<b>d</b>) 4<span class="html-italic">f</span><sub>s</sub>, σ = 3.186.</p>
Full article ">Figure 16
<p>Differences between two types of tongues: (<b>a</b>) OVT and (<b>b</b>) STVT [<a href="#B58-energies-17-05149" class="html-bibr">58</a>].</p>
Full article ">Figure 17
<p>Distribution of pressure coefficients and streamlines at cross-sections in four pumps [<a href="#B58-energies-17-05149" class="html-bibr">58</a>].</p>
Full article ">Figure 18
<p>Nautilus cross-sectional shape [<a href="#B59-energies-17-05149" class="html-bibr">59</a>].</p>
Full article ">Figure 19
<p>Gas–liquid two-phase distribution during liquid filling [<a href="#B59-energies-17-05149" class="html-bibr">59</a>].</p>
Full article ">Figure 20
<p>Sketch of bionic impeller with bionic convex domes [<a href="#B63-energies-17-05149" class="html-bibr">63</a>].</p>
Full article ">Figure 21
<p>Representative scanning electron micrographs of cuticles and fine hairs on the mesothorax of springtails (Collembola) and sea skaters (<span class="html-italic">H. germanus</span>), respectively [<a href="#B69-energies-17-05149" class="html-bibr">69</a>]. (<b>A</b>,<b>B</b>) Springtails have primary granules (triangular) connected by ridges forming honeycomb patterns that prevent the intrusion of liquids on submersion. (<b>C</b>) Long needle-shaped hairs and tiny mushroom-shaped hairs on the dorsal and ventral mesothorax of sea skaters provide robust repellency against seawater. (<b>D</b>) Magnified micrograph of mushroom-shaped hairs.</p>
Full article ">Figure 22
<p>Principle of cavitation damage suppression by GEMS [<a href="#B69-energies-17-05149" class="html-bibr">69</a>]. (<b>A</b>) Liquid jet from a bubble collapsing close to a solid boundary affecting the substrate and causing erosion. The time scale corresponds to a cavitation bubble of Rmax ≈ 570 µm. (<b>B</b>) The gas entrapped inside the GEMS protrudes near the cavitation bubble and behaves as a free boundary. As a result, the liquid jet from the collapsing bubble is directed away from the substrate. The time scale shown is that of a cavitation bubble of Rmax ≈ 520 µm. The time in µs and maximum bubble radius depicted in (<b>A</b>,<b>B</b>) are typical values observed in the experiments. (<b>C</b>) The gas entrapped inside the GEMS expands because of the pressure field generated by the nearby cavitation bubble. Notice that the gas contained in the GEMS bulges outward and reaches an almost hemispherical shape during the expansion of the cavitation bubble as mentioned in the text. Image credit: Xavier Pita, Scientific Illustrator, KAUST.</p>
Full article ">Figure 23
<p>Impellers for centrifugal pumps with non-smooth surfaces [<a href="#B73-energies-17-05149" class="html-bibr">73</a>].</p>
Full article ">Figure 24
<p>Surface morphology of superhydrophobic materials: (<b>a</b>) optical image; (<b>b</b>) three-dimensional morphology; (<b>c</b>) typical cross-sectional image of a superhydrophobic material [<a href="#B88-energies-17-05149" class="html-bibr">88</a>].</p>
Full article ">Figure 25
<p>Bone-coupled bionic metal coatings [<a href="#B90-energies-17-05149" class="html-bibr">90</a>].</p>
Full article ">Figure 26
<p>Experimental plot comparing cavitation of AS coatings with BHS-b coating [<a href="#B90-energies-17-05149" class="html-bibr">90</a>]. (<b>a</b>,<b>b</b>) SEM images of the eroded cross-sections of the coatings. (<b>c</b>,<b>d</b>) SEM images of the eroded surfaces of the coatings. (<b>e</b>,<b>f</b>) the depth distribution on the eroded surface of the coatings according to the 2D contours given in <a href="#energies-17-05149-f026" class="html-fig">Figure 26</a>.</p>
Full article ">
24 pages, 11670 KiB  
Article
Influence of the Traverse Speed of the Stylus Tip on Changes in the Areal Texture Parameters of Machined Surfaces
by Pawel Pawlus, Rafal Reizer and Wiesław Żelasko
Materials 2024, 17(20), 5052; https://doi.org/10.3390/ma17205052 - 16 Oct 2024
Viewed by 256
Abstract
Measurements of areal (3D) surface texture using optical methods are very popular because of the short measurement time compared to the stylus tip technique. However, they are very sensitive to measurement errors. In some cases, optical measurements are not recommended. The stylus measurement [...] Read more.
Measurements of areal (3D) surface texture using optical methods are very popular because of the short measurement time compared to the stylus tip technique. However, they are very sensitive to measurement errors. In some cases, optical measurements are not recommended. The stylus measurement method is well known and can be the reference technique for surface texture measurement. The main disadvantage is the long measuring time. This time can be shortened using higher speeds of measurement. The effect of the speed of the measurement of stylus profilometer on changes in surface texture parameters was studied. Fifty surface topographies were measured using the stylus profilometer at speeds 0.5, 1, 2, 3, 4, and 5 mm/s in the same places. Surfaces after lapping, polishing, grinding, milling, laser texturing, and two-process random surfaces were measured and analyzed. Changes in parameters caused by the increase in the traverse speed depend on the characteristics and parameters of the surfaces. The random surfaces changed more than the deterministic ones. The increase in the traverse speed from 0.5 to 1 mm/s caused small changes in the parameters. Full article
Show Figures

Figure 1

Figure 1
<p>Contour plots of surface A after abrasive blasting after measurements with various speeds: (<b>a</b>) 0.5 mm/s, (<b>b</b>) 1 mm/s, (<b>c</b>) 2 mm / s, (<b>d</b>) 3 mm/s, (<b>e</b>) 4 mm/s and (<b>f</b>) 5 mm/s.</p>
Full article ">Figure 2
<p>Profile details of surface A after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 2 Cont.
<p>Profile details of surface A after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 3
<p>Contour plots of surfaces B (<b>a</b>), D (<b>b</b>), E (<b>c</b>), and F (<b>d</b>) after measurements with speeds of 0.5 mm/s.</p>
Full article ">Figure 3 Cont.
<p>Contour plots of surfaces B (<b>a</b>), D (<b>b</b>), E (<b>c</b>), and F (<b>d</b>) after measurements with speeds of 0.5 mm/s.</p>
Full article ">Figure 4
<p>Profile details of surface B after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 5
<p>Contour plots of surface C after measurements at various speeds: (<b>a</b>) 0.5 mm/s, (<b>b</b>) 1 mm/s, (<b>c</b>) 2 mm/s, (<b>d</b>) 3 mm/s, (<b>e</b>) 4 mm/s and (<b>f</b>) 5 mm/s.</p>
Full article ">Figure 5 Cont.
<p>Contour plots of surface C after measurements at various speeds: (<b>a</b>) 0.5 mm/s, (<b>b</b>) 1 mm/s, (<b>c</b>) 2 mm/s, (<b>d</b>) 3 mm/s, (<b>e</b>) 4 mm/s and (<b>f</b>) 5 mm/s.</p>
Full article ">Figure 6
<p>Profile details of surface C after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 6 Cont.
<p>Profile details of surface C after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 7
<p>Profile details of surface D after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 7 Cont.
<p>Profile details of surface D after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 8
<p>Profile details of surface E after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 8 Cont.
<p>Profile details of surface E after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 9
<p>Profile details of surface F after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 9 Cont.
<p>Profile details of surface F after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 10
<p>Contour plots of surface G after measurements with various speeds: (<b>a</b>) 0.5 mm/s, (<b>b</b>) 1 mm/s, (<b>c</b>) 2 mm/s, (<b>d</b>) 3 mm/s, (<b>e</b>) 4 mm/s and (<b>f</b>) 5 mm/s.</p>
Full article ">Figure 11
<p>Profile details of surface G after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 11 Cont.
<p>Profile details of surface G after measurements with various speeds: (<b>a</b>) 0.5 and 1 mm/s, (<b>b</b>) 0.5 and 2 mm/s, (<b>c</b>) 0.5 and 3 mm/s, (<b>d</b>) 0.5 and 4 mm/s, (<b>e</b>) 0.5 and 5 mm/s; red color—0.5 mm/s speed, blue color—higher speeds.</p>
Full article ">Figure 12
<p>Profiles after grinding (<b>a</b>) and milling (<b>b</b>) characterized by the Pdq parameter of 3.6; the PSm of the ground profile was 0.055 mm, the PSm of the milled profile was 0.13 mm.</p>
Full article ">Figure 12 Cont.
<p>Profiles after grinding (<b>a</b>) and milling (<b>b</b>) characterized by the Pdq parameter of 3.6; the PSm of the ground profile was 0.055 mm, the PSm of the milled profile was 0.13 mm.</p>
Full article ">Figure 13
<p>Dependence between the Pdq/PSm ratio and the relative change in the Sdq parameter caused by the increase in the traverse speed from: (<b>a</b>) 0.5 mm/s to 1 mm/s, (<b>b</b>) 2 mm/s, (<b>c</b>) 3 mm/s, (<b>d</b>) 4 mm/s and (<b>e</b>) 5 mm/s.</p>
Full article ">Figure 13 Cont.
<p>Dependence between the Pdq/PSm ratio and the relative change in the Sdq parameter caused by the increase in the traverse speed from: (<b>a</b>) 0.5 mm/s to 1 mm/s, (<b>b</b>) 2 mm/s, (<b>c</b>) 3 mm/s, (<b>d</b>) 4 mm/s and (<b>e</b>) 5 mm/s.</p>
Full article ">
18 pages, 2171 KiB  
Article
Changes in Texture and Collagen Properties of Pork Skin during Salt–Enzyme–Alkali Tenderization Treatment
by Qiang Zou, Yuyou Chen, Yudie Liu, Linghui Luo, Yuhan Zheng, Guilian Ran and Dayu Liu
Foods 2024, 13(20), 3264; https://doi.org/10.3390/foods13203264 - 14 Oct 2024
Viewed by 413
Abstract
The effects of salt–enzyme–alkali progressive tenderization treatments on porcine cortical conformation and collagen properties were investigated, and their effectiveness and mechanisms were analyzed. The tenderization treatment comprised three treatment stages: CaCl2 (25 °C/0–30 min), papain (35 °C/30–78 min), and Na2CO [...] Read more.
The effects of salt–enzyme–alkali progressive tenderization treatments on porcine cortical conformation and collagen properties were investigated, and their effectiveness and mechanisms were analyzed. The tenderization treatment comprised three treatment stages: CaCl2 (25 °C/0–30 min), papain (35 °C/30–78 min), and Na2CO3 (25 °C/78–120 min). The textural, microscopic, and collagenous properties (content, solubility, and structure) of pork skin were determined at the 0th, 30th, 60th, 90th, and 120th min of the treatment process. The results showed that the shear force, hardness, and chewability of the skin decreased significantly (p < 0.05), and the elasticity exhibited a gradual increase with the progression of tenderization. The content and solubility of collagen showed no significant change at the CaCl2 treatment stage. However, the soluble collagen content increased, the insoluble collagen content decreased, and the collagen solubility increased by 18.04% during the subsequent treatment with papain and Na2CO3. Meanwhile, the scanning electron microscopy results revealed that the regular, wavy structure of the pig skin collagen fibers gradually disappeared during the CaCl2 treatment stage, the overall structure revealed expansion, and the surface microscopic pores gradually increased during the papain and Na2CO3 treatment stages. The findings of the Fourier transform infrared spectroscopy analysis indicated that the hydrogen bonding interactions between the collagen molecules and the C=O, N-H and C-N bonds in the subunit structure of collagen were substantially altered during treatment and that the breakage of amino acid chains and reduction in structural ordering became more pronounced with prolonged treatment. In the tertiary structure, the maximum emission wavelength was blue-shifted and then red-shifted, and the fluorescence intensity was gradually weakened. The surface hydrophobicity was slowly increased. The salt–enzyme–alkali tenderization treatment considerably improved the physical properties and texture of edible pork skins by dissolving collagen fibers and destroying the structure of collagen and its interaction force. Full article
(This article belongs to the Section Meat)
Show Figures

Figure 1

Figure 1
<p>Effect of tenderization process on microstructure of pork skin observed via SEM with selected magnification of 100×. (<b>a</b>) Untreated stage (0 min); (<b>b</b>) 30, (<b>c</b>), 60, (<b>d</b>) 90, and (<b>e</b>) 120 min stages.</p>
Full article ">Figure 2
<p>Changes in collagen content and solubility in pig skin during salt–enzyme–alkali combined tender treatment. (<b>a</b>) Changes in the total and soluble collagen content of pig skin during tenderization treatment; (<b>b</b>) changes in the dissolution rate of pig skin collagen during tenderization treatment. Different letters (a–d) indicate significant differences between different samples (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Fluorescence visible spectrum of collagen at different tenderization treatment times.</p>
Full article ">Figure 4
<p>FTIR of collagen at different tenderization treatment times.</p>
Full article ">Figure 5
<p>Changes in surface hydrophobicity of collagen during salt–enzyme–alkali combined tenderization treatment. Different letters (a–e) indicate significant differences between different samples (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
16 pages, 4225 KiB  
Review
A Review on the Effect of Wood Surface Modification on Paint Film Adhesion Properties
by Jingyi Hang, Xiaoxing Yan and Jun Li
Coatings 2024, 14(10), 1313; https://doi.org/10.3390/coatings14101313 - 14 Oct 2024
Viewed by 542
Abstract
Wood surface treatment aims to improve or reduce the surface activity of wood by physical treatment, chemical treatment, biological activation treatment or other methods to achieve the purpose of surface modification. After wood surface modification, the paint film adhesion performance, gluing performance, surface [...] Read more.
Wood surface treatment aims to improve or reduce the surface activity of wood by physical treatment, chemical treatment, biological activation treatment or other methods to achieve the purpose of surface modification. After wood surface modification, the paint film adhesion performance, gluing performance, surface wettability, surface free energy and surface visual properties would be affected. This article aims to explore the effects of different modification methods on the adhesion of wood coating films. Modification of the wood surface significantly improves the adhesion properties of the paint film, thereby extending the service life of the coating. Research showed that physical external force modification improved the hydrophilicity and wettability of wood by changing its surface structure and texture, thus enhancing the adhesion of the coating. Additionally, high-temperature heat treatment modification reduced the risk of coating cracking and peeling by eliminating stress and moisture within the wood. Chemical impregnation modification utilized the different properties of organic and inorganic substances to improve the stability and durability of wood. Organic impregnation effectively filled the wood cell wall and increased its density, while inorganic impregnation enhanced the adhesion of the coating by forming stable chemical bonds. Composite modification methods combined the advantages of the above technologies and significantly improved the comprehensive properties of wood through multiple modification treatments, showing superior adhesion and durability. Comprehensive analysis indicated that selecting the appropriate modification method was key for different wood types and application environments. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Wettability and permeability of liquid on wood surface; (<b>B</b>) Sketch of the wood surface contact angle.</p>
Full article ">Figure 2
<p>Modification mechanism of wood impregnation.</p>
Full article ">Figure 3
<p>Cross-section schematic diagram of poplar wood before and after impregnation: (<b>A</b>) unmodified wood, (<b>B</b>) modified wood.</p>
Full article ">
11 pages, 4809 KiB  
Article
Binderless Polycrystalline Cubic Boron Nitride Sintered Compacts for Machining of Cemented Carbides
by Alexander S. Osipov, Piotr Klimczyk, Igor A. Petrusha, Yurii O. Melniichuk, Lucyna Jaworska, Kinga Momot and Yuliia Rumiantseva
Ceramics 2024, 7(4), 1477-1487; https://doi.org/10.3390/ceramics7040095 (registering DOI) - 13 Oct 2024
Viewed by 383
Abstract
High-purity, superhard, binderless polycrystalline cubic boron nitride (BL-PCBN) was obtained by direct hBN to cBN transformation in a toroid-type high-pressure apparatus at a pressure of 8.0 GPa and temperature of 2250 °C (HPHT-DCS; high-pressure, high-temperature direct conversion sintering). X-ray diffraction analysis revealed a [...] Read more.
High-purity, superhard, binderless polycrystalline cubic boron nitride (BL-PCBN) was obtained by direct hBN to cBN transformation in a toroid-type high-pressure apparatus at a pressure of 8.0 GPa and temperature of 2250 °C (HPHT-DCS; high-pressure, high-temperature direct conversion sintering). X-ray diffraction analysis revealed a prominent [111] axial texture in the sintered material when the axis was oriented perpendicular to the end surface of the sample. Vickers hardness tests conducted at a load of 49 N showed that BL-PCBN possessed an exceptional hardness value of 63.4 GPa. Finally, cutting tools made of BL-PCBN and SN-PCBN (Si3N4-doped cBN-based composite) reference materials were tested during the turning of a cemented tungsten carbide workpiece. The results of the cutting tests demonstrated that the wear resistance of the BL-PCBN material obtained with the HPHT-DCS process is 1.5–1.9 times higher compared to the conventional SN-PCBN material, suggesting its significant potential for industrial application. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>HPC view in the axial section: (<b>a</b>) central part of the assembly (half section) before compression; (<b>b</b>) HPC in compressed state (half section); (<b>c</b>) HPA-T30 appearance; (1) pyrophyllite heat-insulating ring, (2) axial heater (ZrO<sub>2</sub> + graphite), (3) pressed graphite disc, (4) CsCl + graphite, (5) graphite heater, (6) specimen, (7) outer supporting punch, (8) WC-Co anvil of HPA, (9) fastening steel ring, (10) CaCO<sub>3</sub> internal gasket; (11) external gasket (pressed CaCO<sub>3</sub> + binder).</p>
Full article ">Figure 2
<p>Justification of the choice of HPHT-DCS parameters for the present study (frame) against the background of the results of previous studies on complete hBN→cBN conversion (<b>a</b>) with the formation of an ultra-fine-grain monophase structure of the polycrystalline cBN (<b>b</b>).</p>
Full article ">Figure 3
<p>Turning of cemented carbide bushing of the WC-Co system (Co 15 wt.%) using the BL-PCBN (RNMN 09T300F) cutting insert.</p>
Full article ">Figure 4
<p>Initial hBN powder: (<b>a</b>) SEM image of the powder morphology; (<b>b</b>) XRD pattern (the Miller indices of the planes in the hBN lattice are given only for the most intense reflections).</p>
Full article ">Figure 5
<p>Translucent BL-PCBN obtained from hBN powder by the HPHT-DCS method: (<b>a</b>) sample view (sample thickness 3.6 mm); (<b>b</b>) XRD pattern of the sample.</p>
Full article ">Figure 6
<p>Change of flank wear width with cutting speed during the turning of the WC-15Co alloy (depth of cut—0.2 mm; feed—0.1 mm/rev; cutting length—75 m).</p>
Full article ">Figure 7
<p>Cutting edges of SN-PCBN (<b>a</b>,<b>c</b>) and BL-PCBN (<b>b</b>,<b>d</b>) tools after the turning of the WC-15Co material at 37 m/min (<b>a</b>,<b>b</b>) and 60 m/min (<b>c</b>,<b>d</b>). The distance between the dashed lines corresponds to the flank wear.</p>
Full article ">Figure 8
<p>The turning performance of WC-15Co alloy cutting: depth of cut—0.2 mm; feed—0.1 mm/rev; cutting length—75 m; flank wear width—0.4 mm.</p>
Full article ">
23 pages, 5488 KiB  
Article
Groundwater Recharge Response to Reduced Irrigation Pumping: Checkbook Irrigation and the Water Savings Payment Plan
by Justin Gibson, Trenton E. Franz, Troy Gilmore, Derek Heeren, John Gates, Steve Thomas and Christopher M. U. Neale
Water 2024, 16(20), 2910; https://doi.org/10.3390/w16202910 - 13 Oct 2024
Viewed by 553
Abstract
Ongoing investments in irrigation technologies highlight the need to accurately estimate the longevity and magnitude of water savings at the watershed level to avoid the paradox of irrigation efficiency. This paradox arises when irrigation pumping exceeds crop water demand, leading to excess water [...] Read more.
Ongoing investments in irrigation technologies highlight the need to accurately estimate the longevity and magnitude of water savings at the watershed level to avoid the paradox of irrigation efficiency. This paradox arises when irrigation pumping exceeds crop water demand, leading to excess water that is not recovered by the watershed. Comprehensive water accounting from farm to watershed scales is challenging due to spatial variability and inadequate socio-hydrological data. We hypothesize that water savings are short term, as prior studies show rapid recharge responses to surface changes. Precise estimation of these time scales and water savings can aid water managers making decisions. In this study, we examined water savings at three 65-hectare sites in Nebraska with diverse soil textures, management practices, and groundwater depths. Surface geophysics effectively identified in-field variability in soil water content and water flux. A one-dimensional model showed an average 80% agreement with chloride mass balance estimates of deep drainage. Our findings indicate that groundwater response times are short and water savings are modest (1–3 years; 50–900 mm over 10 years) following a 120 mm/year reduction in pumping. However, sandy soils with shallow groundwater show minimal potential for water savings, suggesting limited effectiveness of irrigation efficiency programs in such regions. Full article
(This article belongs to the Section Hydrology)
Show Figures

Figure 1

Figure 1
<p>Conceptual diagram of water savings and hypothetical case study. The lag time is defined by the amount of time that elapses following a reduction in pumping but before recharge rates begin to decrease. Lag times are a function of the depth to groundwater, soil water states and fluxes, and soil hydraulic parameters. Also note that the water savings are flat after 3 years, meaning no additional benefit, and that future management decisions can reduce water savings if pumping rates return to their initial rates or if field experiences prolonged periods of dry conditions.</p>
Full article ">Figure 2
<p>Location of the three study sites near Brule, NE (red dot on USA). Each site is ~65 ha in area and primarily under irrigated maize production. White outlines are SSURGO soil boundaries. Field sites are S1, S3 and S4 from west to east.</p>
Full article ">Figure 3
<p>Results of time-repeat ECa mapping from the Dualem 21S instrument (deep signal ~0–3.2 m) and the corresponding 1st EOF reprojected spatially for each of the three 65 ha study sites (see <a href="#water-16-02910-t001" class="html-table">Table 1</a> for sample dates). Warm EOF colors indicate drier zones/coarser soil texture and cooler colors indicate wetter zones/finer soil texture compared to the field average. White lines are SSURGO soil boundaries. White dots are locations of core extraction (20 November 2017). Red dots are the location of the groundwater observation well (closest well to S1 was ~0.4 km away and not pictured here). Geophysical data layers can be found in <a href="#app1-water-16-02910" class="html-app">Files SI1–SI3</a>.</p>
Full article ">Figure 4
<p>Volumetric water content (VWC) and chloride (Cl<sup>−</sup>) concentration profiles of soil cores extracted from the three field sites. Line colors correspond to EOF values determined at the core location (e.g., warm colors correspond to negative EOF values, green colors correspond to near-zero EOF values, and cool colors correspond to positive EOF values; see <a href="#water-16-02910-f003" class="html-fig">Figure 3</a>). Sawtooth patterns observed in VWC and Cl- profiles align with changes in soil textures. Data from this analysis can be found in SI4.</p>
Full article ">Figure 5
<p>Numerical modeling results of annual deep drainage; 2012 was an exceptionally dry year with 36% of average precipitation falling for that year. Bar colors correspond to EOF values determined at the core location (e.g., warm colors correspond to negative EOF values, green colors correspond to near-zero EOF values, and cool colors correspond to positive EOF values).</p>
Full article ">Figure 6
<p>Volumetric water content profiles from the core analysis overlain onto numerical modeling outputs. Bands are the minimum and maximum of ranges of the simulated VWC profiles and dashed lines are the corresponding simulated mean over the 10-year simulation period. Lines with circles are from the extracted volumetric analysis from core. Line and band colors correspond to the EOF values determined at the core location (e.g., warm colors correspond to negative EOF values, green colors correspond to near-zero EOF values, and cool colors correspond to positive EOF values).</p>
Full article ">Figure 7
<p>Correlation between root zone depth integrated VWC for extracted cores and the corresponding simulated root zone depth-integrated VWC (10-year average). EOF values at each core location from the repeat geophysical analysis separate the relative ranges of depth integrated VWC for both the extracted cores and simulated soil profiles. Solid line is 1:1 and dashed line is best fit to data.</p>
Full article ">Figure 8
<p>Time series of model output determined at one core (S4C) from two paired simulations that vary only in irrigation scheduling routines. In this case, the lag time is approximately 2.5 years long (determined visually when recharge reductions begin to increase). Water savings are calculated as a cumulative reduction in pumping minus the sum of the cumulative reduction in recharge and ET.</p>
Full article ">Figure 9
<p>Time series of simulated water savings calculated from the paired simulations for each core. Cores with coarser soil textures (S1A, S3E, and S4A) had the largest water savings as a result of a reduction in ET.</p>
Full article ">Figure 10
<p>Sensitivity analysis of weather year on estimated lag times and water savings. In both panels, simulations were carried out where a continuously repeated dry year is in red, a continuously repeated wet year is in blue, and the 10-year observed weather is in green. The 10th and 90th percentile weather years were selected for this analysis.</p>
Full article ">
20 pages, 13856 KiB  
Article
Clay Minerals/TiO2 Composites—Characterization and Application in Photocatalytic Degradation of Water Pollutants
by Bogna D. Napruszewska, Dorota Duraczyńska, Joanna Kryściak-Czerwenka, Paweł Nowak and Ewa M. Serwicka
Molecules 2024, 29(20), 4852; https://doi.org/10.3390/molecules29204852 - 13 Oct 2024
Viewed by 361
Abstract
TiO2 used for photocatalytic water purification is most active in the form of nanoparticles (NP), but their use is fraught with difficulties in separation from solution or/and a tendency to agglomerate. The novel materials designed in this work circumvent these problems by [...] Read more.
TiO2 used for photocatalytic water purification is most active in the form of nanoparticles (NP), but their use is fraught with difficulties in separation from solution or/and a tendency to agglomerate. The novel materials designed in this work circumvent these problems by immobilizing TiO2 NPs on the surface of exfoliated clay minerals. A series of TiO2/clay mineral composites were obtained using five different clay components: the Na-, CTA-, or H-form of montmorillonite (Mt) and Na- or CTA-form of laponite (Lap). The TiO2 component was prepared using the inverse microemulsion method. The composites were characterized with X-ray diffraction, scanning/transmission electron microscopy/energy dispersive X-ray spectroscopy, FTIR spectroscopy, thermal analysis, and N2 adsorption–desorption isotherms. It was shown that upon composite synthesis, the Mt interlayer became filled by a mixture of CTA+ and hydronium ions, regardless of the nature of the parent clay, while the structure of Lap underwent partial destruction. The composites displayed high specific surface area and uniform mesoporosity determined by the size of the TiO2 nanoparticles. The best textural parameters were shown by composites containing clay components whose structure was partially destroyed; for instance, Ti/CTA-Lap had a specific surface area of 420 m2g−1 and a pore volume of 0.653 cm3g−1. The materials were tested in the photodegradation of methyl orange and humic acid upon UV irradiation. The photocatalytic activity could be correlated with the development of textural properties. In both reactions, the performance of the most photoactive composites surpassed that of the reference commercial P25 titania. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of (<b>a</b>) clay minerals used as supports for TiO<sub>2</sub> nanoparticles; (<b>b</b>) composites of TiO<sub>2</sub> and clay minerals.</p>
Full article ">Figure 2
<p>FTIR spectra of laponite-based composites, clay supports, reference nanocrystalline TiO<sub>2,</sub> and sum spectra of individual components.</p>
Full article ">Figure 3
<p>TEM images of (<b>a</b>) nanoparticles present in hydrolyzed Ti-containing inverse microemulsion; (<b>b</b>) Ti/CTA-Mt composite; (<b>c</b>) Ti//H-Mt composite. Sample suspensions deposited on 200 mesh copper grids covered with carbon film. Magnification ×200,000.</p>
Full article ">Figure 4
<p>SEM/EDX compositional analysis of selected areas of synthesized composites: (<b>a</b>) SEM image of Ti/CTA-Mt; (<b>b</b>) EDX mapping of Ti in Ti/CTA-Mt; (<b>c</b>) EDX mapping of Si in Ti/CTA-Mt; (<b>d</b>) SEM image of Ti/Na-Mt; (<b>e</b>) EDX mapping of Ti in Ti/Na-Mt; (<b>f</b>) EDX mapping of Si in Ti/Na-Mt; (<b>g</b>) SEM image of Ti/H-Mt; (<b>h</b>) EDX mapping of Ti in Ti/H-Mt; (<b>i</b>) EDX mapping of Si in Ti/H-Mt; (<b>j</b>) SEM image of Ti/CTA-Lap; (<b>k</b>) EDX mapping of Ti in Ti/CTA-Lap; (<b>l</b>) EDX mapping of Si in Ti/CTA-Lap; (<b>m</b>) SEM image of Ti/Na-Lap; (<b>n</b>) EDX mapping of Ti in Ti/Na-Lap; (<b>o</b>) EDX mapping of Si in Ti/Na-Lap.</p>
Full article ">Figure 4 Cont.
<p>SEM/EDX compositional analysis of selected areas of synthesized composites: (<b>a</b>) SEM image of Ti/CTA-Mt; (<b>b</b>) EDX mapping of Ti in Ti/CTA-Mt; (<b>c</b>) EDX mapping of Si in Ti/CTA-Mt; (<b>d</b>) SEM image of Ti/Na-Mt; (<b>e</b>) EDX mapping of Ti in Ti/Na-Mt; (<b>f</b>) EDX mapping of Si in Ti/Na-Mt; (<b>g</b>) SEM image of Ti/H-Mt; (<b>h</b>) EDX mapping of Ti in Ti/H-Mt; (<b>i</b>) EDX mapping of Si in Ti/H-Mt; (<b>j</b>) SEM image of Ti/CTA-Lap; (<b>k</b>) EDX mapping of Ti in Ti/CTA-Lap; (<b>l</b>) EDX mapping of Si in Ti/CTA-Lap; (<b>m</b>) SEM image of Ti/Na-Lap; (<b>n</b>) EDX mapping of Ti in Ti/Na-Lap; (<b>o</b>) EDX mapping of Si in Ti/Na-Lap.</p>
Full article ">Figure 5
<p>(<b>a</b>) TG traces of clay components; (<b>b</b>) TG traces of Ti/clay composites and of reference TiO<sub>2</sub>; (<b>c</b>) DSC traces of clay components; (<b>d</b>) DSC traces of Ti/clay composites and of reference TiO<sub>2</sub>.</p>
Full article ">Figure 6
<p>N<sub>2</sub> adsorption/desorption isotherms at −196 °C of (<b>a</b>) CTA-Mt and Ti/CTA-Mt; (<b>c</b>) Na-Mt and Ti/Na-Mt; (<b>e</b>) H-Mt and Ti/H-Mt; (<b>g</b>) CTA-Lap and Ti/CTA-Lap; (<b>i</b>) Na-Lap and Ti/Na-Lap. Differential pore size distribution of (<b>b</b>) CTA-Mt and Ti/CTA-Mt; (<b>d</b>) Na-Mt and Ti/Na-Mt; (<b>f</b>) H-Mt and Ti/H-Mt; (<b>h</b>) CTA-Lap and Ti/CTA-Lap; (<b>j</b>) Na-Lap and Ti/Na-Lap.</p>
Full article ">Figure 6 Cont.
<p>N<sub>2</sub> adsorption/desorption isotherms at −196 °C of (<b>a</b>) CTA-Mt and Ti/CTA-Mt; (<b>c</b>) Na-Mt and Ti/Na-Mt; (<b>e</b>) H-Mt and Ti/H-Mt; (<b>g</b>) CTA-Lap and Ti/CTA-Lap; (<b>i</b>) Na-Lap and Ti/Na-Lap. Differential pore size distribution of (<b>b</b>) CTA-Mt and Ti/CTA-Mt; (<b>d</b>) Na-Mt and Ti/Na-Mt; (<b>f</b>) H-Mt and Ti/H-Mt; (<b>h</b>) CTA-Lap and Ti/CTA-Lap; (<b>j</b>) Na-Lap and Ti/Na-Lap.</p>
Full article ">Figure 7
<p>Results of photocatalytic experiments after 5 h irradiation: (<b>a</b>) composites ordered according to the increasing crystallinity of the titania component; (<b>b</b>) composites ordered according to the increasing pore volume.</p>
Full article ">Figure 8
<p>Scheme of setup for photocatalytic experiments.</p>
Full article ">
13 pages, 9014 KiB  
Article
Influence of Synthesis Parameters on Structure and Characteristics of the Graphene Grown Using PECVD on Sapphire Substrate
by Šarūnas Jankauskas, Šarūnas Meškinis, Nerija Žurauskienė and Asta Guobienė
Nanomaterials 2024, 14(20), 1635; https://doi.org/10.3390/nano14201635 - 12 Oct 2024
Viewed by 371
Abstract
The high surface area and transfer-less growth of graphene on dielectric materials is still a challenge in the production of novel sensing devices. We demonstrate a novel approach to graphene synthesis on a C-plane sapphire substrate, involving the microwave plasma-enhanced chemical vapor deposition [...] Read more.
The high surface area and transfer-less growth of graphene on dielectric materials is still a challenge in the production of novel sensing devices. We demonstrate a novel approach to graphene synthesis on a C-plane sapphire substrate, involving the microwave plasma-enhanced chemical vapor deposition (MW-PECVD) technique. The decomposition of methane, which is used as a precursor gas, is achieved without the need for remote plasma. Raman spectroscopy, atomic force microscopy and resistance characteristic measurements were performed to investigate the potential of graphene for use in sensing applications. We show that the thickness and quality of graphene film greatly depend on the CH4/H2 flow ratio, as well as on chamber pressure during the synthesis. By varying these parameters, the intensity ratio of Raman D and G bands of graphene varied between ~1 and ~4, while the 2D to G band intensity ratio was found to be 0.05–0.5. Boundary defects are the most prominent defect type in PECVD graphene, giving it a grainy texture. Despite this, the samples exhibited sheet resistance values as low as 1.87 kΩ/□. This reveals great potential for PECVD methods and could contribute toward efficient and straightforward graphene growth on various substrates. Full article
(This article belongs to the Section Nanofabrication and Nanomanufacturing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the PECVD chamber during graphene synthesis on C-sapphire substrate, (<b>b</b>) illustration of C-AFM analysis, (<b>c</b>) illustration of four-point probe analysis for sheet resistance measurement and (<b>d</b>) schematic representation of PECVD graphene synthesis stages. Atom spacing, electrode placement and dimensions are not to scale for clarity.</p>
Full article ">Figure 2
<p>(<b>a</b>) Comparison of Raman spectra of PECVD graphene on C−plane sapphire (violet) and SiO<sub>2</sub> (orange), (<b>b</b>) colormap of I<sub>2D</sub>/I<sub>G</sub> ratio vs. synthesis parameters, (<b>c</b>) I<sub>D</sub>/I<sub>G</sub> ratio vs. I<sub>2D</sub>/I<sub>G</sub> ratio, (<b>d</b>) colormap of I<sub>D</sub>/I<sub>G</sub> ratio vs. synthesis parameters. The samples were differentiated by shapes and color, where light green represents samples that belong to the F set and light blue represents samples that belong to the P set. Sample F3P2, which belongs to both sets, is colored red. Additional samples were produced (set S) for colormap space expansion.</p>
Full article ">Figure 3
<p>(<b>a</b>) Pos<sub>2D</sub> vs. gas-flow ratio plot, (<b>b</b>) Pos<sub>2D</sub> vs. pressure plot, (<b>c</b>) FWHM<sub>2D</sub> vs. gas-flow ratio plot, (<b>d</b>) FWHM<sub>2D</sub> vs. pressure plot of F and P samples. Conveniently, samples were given the same shapes and colors for distinction.</p>
Full article ">Figure 4
<p>Pos<sub>2D</sub> vs. Pos<sub>G</sub> plot showing vector decomposition analysis of F and P sample sets. Conveniently, samples were given the same shapes and colors for distinction.</p>
Full article ">Figure 5
<p>(<b>a</b>) Morphology of graphene synthesized using a CH<sub>4</sub>/H<sub>2</sub> flow ratio of 35/65 and a pressure of 10 (R<sub>q</sub> indicates the root mean square value of surface roughness), (<b>b</b>) R<sub>q</sub> vs. gas-flow ratio plot, (<b>c</b>) R<sub>q</sub> vs. pressure plot. The colors and shapes follow the same sample pattern.</p>
Full article ">Figure 6
<p>(<b>a</b>) Two-dimensional (2D) and G band ratio with respect to surface roughness, (<b>b</b>) plot showing correlation with surface roughness. The colors and shapes follow the same sample pattern.</p>
Full article ">Figure 7
<p>Conductive atomic force microscopy maps of four selected samples with varying synthesis conditions. (<b>a</b>) F2 sample, (<b>b</b>) F5 sample, (<b>c</b>) P1 sample, (<b>d</b>) P5 sample. I<sub>q</sub> represents the root mean square of surface conductivity values.</p>
Full article ">Figure 8
<p>(<b>a</b>) Resistance characteristics vs. I<sub>2D</sub>/I<sub>G</sub> plot showing conductance and sheet resistance variation with different thicknesses of graphene, (<b>b</b>) resistance characteristics vs. I<sub>D</sub>/I<sub>G</sub> plot showing conductance and sheet resistance variation based on defects.</p>
Full article ">
15 pages, 3373 KiB  
Article
Osteoblast Response to Widely Ranged Texturing Conditions Obtained through High Power Laser Beams on Ti Surfaces
by Federico Alessandro Ruffinatti, Tullio Genova, Ilaria Roato, Martina Perin, Giorgia Chinigò, Riccardo Pedraza, Olivio Della Bella, Francesca Motta, Elisa Aimo Boot, Domenico D’Angelo, Giorgio Gatti, Giorgia Scarpellino, Luca Munaron and Federico Mussano
J. Funct. Biomater. 2024, 15(10), 303; https://doi.org/10.3390/jfb15100303 - 12 Oct 2024
Viewed by 440
Abstract
Titanium and titanium alloys are the prevailing dental implant materials owing to their favorable mechanical properties and biocompatibility, but how roughness dictates the biological response is still a matter of debate. In this study, laser texturing was used to generate eight paradigmatic roughened [...] Read more.
Titanium and titanium alloys are the prevailing dental implant materials owing to their favorable mechanical properties and biocompatibility, but how roughness dictates the biological response is still a matter of debate. In this study, laser texturing was used to generate eight paradigmatic roughened surfaces, with the aim of studying the early biological response elicited on MC3T3-E1 pre-osteoblasts. Prior to cell tests, the samples underwent SEM analysis, optical profilometry, protein adsorption assay, and optical contact angle measurement with water and diiodomethane to determine surface free energy. While all the specimens proved to be biocompatible, supporting similar cell viability at 1, 2, and 3 days, surface roughness could impact significantly on cell adhesion. Factorial analysis and linear regression showed, in a robust and unprecedented way, that an isotropic distribution of deep and closely spaced valleys provides the best condition for cell adhesion, to which both protein adsorption and surface free energy were highly correlated. Overall, here the authors provide, for the first time, a thorough investigation of the relationship between roughness parameters and osteoblast adhesion that may be applied to design and produce new tailored interfaces for implant materials. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of the factorial experimental design followed for surface modification of titanium disks. Throughout this work, titanium samples are identified using an alphanumeric ID encoding these three features, in the following order: inter-pit distance (25 or 50 μm), pattern type (A for aligned or R for random), and pit depth (6 or 18 μm).</p>
Full article ">Figure 2
<p>SEM images showing the surface topography of the titanium samples at high magnification (1000×). Each surface is named after the alphanumeric ID defined above, which consists of the values of pit spacing (25 or 50 μm), pattern type (Aligned or Random), and pit depth (6 or 18 μm), respectively.</p>
Full article ">Figure 3
<p>Protein adsorption and cell adhesion measures. (<b>A</b>) Data from three independent protein adsorption assays are represented as mean ± SE of the amount of adsorbed protein per volume (µg/mL), for each different titanium surface. (<b>B</b>) Data from three independent cell adhesion assays are represented as mean ± SE of the number of counted cells per field of view, for each different titanium surface.</p>
Full article ">Figure 4
<p>Cell viability at 24 h and 72 h. Data from three independent experiments are presented as mean ± SE for each different titanium surface (RLU = relative light units).</p>
Full article ">Figure 5
<p>Factorial analysis. Main effects on cell adhesion of: (<b>A</b>) inter-pit distance (<span class="html-italic">p</span>-value = 2.55 × 10<sup>−2</sup>), (<b>B</b>) pattern type (<span class="html-italic">p</span>-value = 3.32 × 10<sup>−2</sup>), and (<b>C</b>) pit depth (<span class="html-italic">p</span>-value = 9.67 × 10<sup>−4</sup>). (<b>D</b>) A 3D representation of the global regression model. Light-blue and violet planes represent the linear model equation evaluated for aligned and random pattern, respectively.</p>
Full article ">Figure 6
<p>Graphical representation of the plane model. Two different views of the same plane given by Equation (1) linking cell adhesion data points (yellow dots) to the surface roughness properties of the titanium disks. Multiple linear regression analysis returned only two non-redundant and statistically significant coefficients out of the eleven initial roughness descriptors (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>β</mi> </mrow> <mrow> <mi>S</mi> <mi>t</mi> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math> coefficient: <span class="html-italic">p</span>-value = 7.10 × 10<sup>−3</sup>; <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>β</mi> </mrow> <mrow> <mi>S</mi> <mi>k</mi> <mi>u</mi> </mrow> </msub> </mrow> </semantics></math> coefficient: <span class="html-italic">p</span>-value = 3.57 × 10<sup>−2</sup>.</p>
Full article ">Figure 7
<p>Correlation analysis between cell adhesion, protein adsorption, and SFE. (<b>A</b>) Scatterplot and correlation analysis between protein adsorption and cell adhesion data. The numerical values of the correlation coefficients shown in the graph refer to the overall correlation analysis between the two datasets (Pearson correlation coefficient <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>P</mi> </mrow> </msub> <mo>=</mo> <mn>0.82</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>p</mi> <mrow> <mtext>-</mtext> <mi>value</mi> </mrow> <mo>=</mo> <mn>0.012</mn> </mrow> </semantics></math>; Spearman correlation coefficient <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>S</mi> </mrow> </msub> <mo>=</mo> <mn>0.93</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>p</mi> <mrow> <mtext>-</mtext> <mi>value</mi> </mrow> <mo>=</mo> <mn>0.002</mn> </mrow> </semantics></math>), while two independent loess curves were used to highlight the nearly deterministic relationship between protein adsorption and cell adhesion within the single subset of surfaces with 50 μm inter-pit distance (cyan curve) and 25 μm spacing (magenta curve). (<b>B</b>) Scatterplot and correlation analysis between SFE and cell adhesion data. The best fitting line (in blue) and the 95% confidence interval (shaded in gray) are superimposed on the data points (Pearson correlation coefficient <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>P</mi> </mrow> </msub> <mo>=</mo> <mn>0.88</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>p</mi> <mrow> <mtext>-</mtext> <mi>value</mi> </mrow> <mo>=</mo> <mn>0.009</mn> </mrow> </semantics></math>; Spearman correlation coefficient <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>S</mi> </mrow> </msub> <mo>=</mo> <mn>0.86</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>p</mi> <mrow> <mtext>-</mtext> <mi>value</mi> </mrow> <mo>=</mo> <mn>0.024</mn> </mrow> </semantics></math>).</p>
Full article ">
24 pages, 17951 KiB  
Article
Durability Investigation of Ultra-Thin Polyurethane Wearing Course for Asphalt Pavement
by Wenguang Wang, Baodong Liu, Dongzhao Jin, Miao Yu and Junsen Zeng
Materials 2024, 17(20), 4977; https://doi.org/10.3390/ma17204977 - 11 Oct 2024
Viewed by 303
Abstract
In this study, a wear-resistant ultra-thin wear layer was fabricated with polyurethane as an adhesive to investigate its durability for pavement applications. Its road performance was investigated based on indoor tests. First, the abrasion test was performed using a tire–pavement dynamic friction analyzer [...] Read more.
In this study, a wear-resistant ultra-thin wear layer was fabricated with polyurethane as an adhesive to investigate its durability for pavement applications. Its road performance was investigated based on indoor tests. First, the abrasion test was performed using a tire–pavement dynamic friction analyzer (TDFA), and the surface elevation information of the wear layer was obtained by laser profile scanning. The relationship between the anti-skid properties of the wear layer and the macro-texture was analyzed. Second, a Fourier infrared spectrometer and scanning electron microscope were employed to analyze the evolution of polyurethane aging properties in the pull-out test and accelerated ultraviolet (UV) aging test. The results showed that the mean profile depth (MPD), arithmetic mean wavelength of contour (λa), surface wear index (SBI), stage mass loss rate (σ), and total stage mass loss rate (ω) of the abrasive layer aggregate had significant multivariate quadratic polynomial relationships with the skidding performance of the abrasive layer. The tensile strength of the polyurethane ultra-thin abrasive layer decreased by only 2.59% after 16 days of UV aging, indicating a minimal effect of UV action on the aggregate and structural spalling of the polyurethane abrasive layer. Full article
(This article belongs to the Special Issue Innovative Materials and Technologies for Road Pavements)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Macro texture measurement area and scan path.</p>
Full article ">Figure 2
<p>Schematic diagram of surface texture parameters.</p>
Full article ">Figure 3
<p>Physical diagram of the pull-out test.</p>
Full article ">Figure 4
<p>Shear test force diagram.</p>
Full article ">Figure 5
<p>A flowchart of test program in this study.</p>
Full article ">Figure 6
<p>Variation rule of each texture parameter and DFC with abrasion time.</p>
Full article ">Figure 7
<p>Morphology of three wear layers before and after abrasion. Note: all the scanned areas are two mutually perpendicular 40 mm × 100 mm rectangular areas on the wheel track.</p>
Full article ">Figure 8
<p>Decay of pavement DFC under different loads.</p>
Full article ">Figure 9
<p>Decay of DFC at different velocities.</p>
Full article ">Figure 10
<p>DFC decay of pavement with different aggregate types.</p>
Full article ">Figure 11
<p>Decay of DFC with different aggregate particle sizes.</p>
Full article ">Figure 12
<p>Wear layer appearance after different abrasion times. Note: all the scanned areas are two mutually perpendicular 40 mm × 100 mm rectangular areas on the wheel track.</p>
Full article ">Figure 13
<p>Variation of drawing strength of two specimens with temperature.</p>
Full article ">Figure 14
<p>Changes in the apparent morphology of polyurethane specimens after UV aging.</p>
Full article ">Figure 14 Cont.
<p>Changes in the apparent morphology of polyurethane specimens after UV aging.</p>
Full article ">Figure 15
<p>Infrared spectral evolution of polyurethane material after UV aging.</p>
Full article ">Figure 16
<p>Law of the effect of UV aging on the tensile properties of polyurethane.</p>
Full article ">Figure 17
<p>The appearance of A Type wear layer before and after aging.</p>
Full article ">Figure 18
<p>Mass loss of class A abrasive layer before and after UV aging.</p>
Full article ">Figure 19
<p>Class B wear layer specimens before and after aging.</p>
Full article ">Figure 20
<p>Mass loss of class B wear layer after UV aging.</p>
Full article ">
16 pages, 9355 KiB  
Article
Enhancing Magnesium Bioactivity for Biomedical Applications: Effects of Laser Texturing and Sandblasting on Surface Properties
by Marjetka Conradi, Aleksandra Kocijan and Bojan Podgornik
Materials 2024, 17(20), 4978; https://doi.org/10.3390/ma17204978 - 11 Oct 2024
Viewed by 303
Abstract
Magnesium and its alloys, valued for their lightweight and durable characteristics, have garnered increasing attention for biomedical applications due to their exceptional biocompatibility and biodegradability. This work introduces a comparison of advanced and basic methods—laser texturing and sandblasting—on magnesium surfaces to enhance bioactivity [...] Read more.
Magnesium and its alloys, valued for their lightweight and durable characteristics, have garnered increasing attention for biomedical applications due to their exceptional biocompatibility and biodegradability. This work introduces a comparison of advanced and basic methods—laser texturing and sandblasting—on magnesium surfaces to enhance bioactivity for biomedical applications. Employing a comprehensive analysis spanning surface morphology, hardness, wettability, tribological performance, and corrosion behavior, this study elucidates the intricate relationship between varied surface treatments and magnesium’s performance. Findings reveal that both laser texturing and sandblasting induce grain refinement. Notably, sandblasting, particularly with a duration of 2 s, demonstrates superior wear resistance and reduced corrosion rates compared to untreated magnesium, thereby emerging as a promising approach for enhancing magnesium bioactivity in biomedical contexts. This investigation contributes to a deeper understanding of the nuanced interactions between diverse surface treatments and their implications for magnesium implants in chloride-rich environments, offering valuable insights for prospective biomedical applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM image of Al<sub>2</sub>O<sub>3</sub> particles used for sandblasting.</p>
Full article ">Figure 2
<p>SEM images of magnesium surfaces under investigation: diamond-polished (<b>a</b>), laser-textured (<b>b</b>), sandblasted for 2 s (<b>c</b>), sandblasted for 5 s (<b>d</b>), sandblasted for 10 s (<b>e</b>), and sandblasted for 30 s (<b>f</b>). The inset images show the details of the surfaces’ morphology at higher magnification.</p>
Full article ">Figure 3
<p>Surface profiles of diamond-polished (<b>a</b>), laser-textured (<b>b</b>), sandblasted for 2 s (<b>c</b>), sandblasted for 5 s (<b>d</b>), sandblasted for 10 s (<b>e</b>), and sandblasted for 30 s (<b>f</b>).</p>
Full article ">Figure 4
<p>EBSD band contrast image overlapping with the EBSD phase map for diamond-polished (<b>a</b>), laser-textured (<b>b</b>), and sandblasted 30 s (<b>c</b>) magnesium surface.</p>
Full article ">Figure 5
<p>EDS mapping of laser textured surface indicating increased surface oxidation of the ablated area.</p>
Full article ">Figure 6
<p>Hardness (GPa) variation with the distance from the sample’s surface into the bulk for diamond-polished (<b>a</b>), laser-textured (<b>b</b>), and sandblasted surfaces from 2 s–30 s (<b>c</b>–<b>f</b>).</p>
Full article ">Figure 7
<p>XPS spectra C 1 s, O 1 s, and Mg 1 s from polished (DP), laser-textured (LT), and sandblasted for 5 s (SB 5 s) magnesium samples.</p>
Full article ">Figure 8
<p>Comparison of friction curves for untreated (DP), laser-textured (LT), and sandblasted (SB) surfaces measured in air (black ○) and under lubrication with Hank’s solution (red □).</p>
Full article ">Figure 9
<p>Comparison of steady-state coefficient of friction (COF) for untreated (DP), laser-textured (LT), and sandblasted (SB) magnesium surfaces measured in air and under lubrication with Hank’s solution.</p>
Full article ">Figure 10
<p>Wear scar SEM micrographs in air for untreated (<b>a</b>), laser-textured (<b>b</b>), and sandblasted (<b>c</b>–<b>f</b>) magnesium surfaces. The insets show the wear scar on the counter body, Al<sub>2</sub>O<sub>3</sub> ball.</p>
Full article ">Figure 11
<p>SEM micrographs of wear scars under lubrication in Hank’s solution for untreated (<b>a</b>), laser-textured (<b>b</b>), and sandblasted (<b>c</b>–<b>f</b>) magnesium surfaces. The insets show the wear scar on the counter body, Al<sub>2</sub>O<sub>3</sub> ball.</p>
Full article ">Figure 12
<p>Wear volumes of untreated, laser-textured, and sandblasted magnesium surfaces in air and under lubrication in Hank’s solution.</p>
Full article ">Figure 13
<p>Potentiodynamic curves for polished (DP), laser-textured (LT), and sandblasted for 2–30 s (SB 2 s, SB 5 s, SB 10 s, and SB 30 s) magnesium samples measured in simulated physiological Hank’s solution at pH = 7.8 and room temperature.</p>
Full article ">
14 pages, 2649 KiB  
Article
Enhanced Photocatalytic Activity in Photocatalytic Concrete: Synthesis, Characterization, and Comprehensive Performance Assessment of Nano-TiO2-Modified Recycled Aggregates
by Xiucheng Zhang, Weizhi Chen, Wencong Lin, Jiansheng Zheng, Guohui Yan and Xuefei Chen
Catalysts 2024, 14(10), 711; https://doi.org/10.3390/catal14100711 - 11 Oct 2024
Viewed by 605
Abstract
This study presents an exhaustive exploration into the development and rigorous evaluation of nano-TiO2-modified recycled aggregates (NT@RAs) as an environmentally sustainable substitute for natural aggregates in concrete applications. A methodical framework was devised for the synthesis and thorough characterization of NT@RAs, [...] Read more.
This study presents an exhaustive exploration into the development and rigorous evaluation of nano-TiO2-modified recycled aggregates (NT@RAs) as an environmentally sustainable substitute for natural aggregates in concrete applications. A methodical framework was devised for the synthesis and thorough characterization of NT@RAs, emphasizing the optimization of nano-TiO2 loading onto the RA surface and within its intricate porous structure. The investigation encompassed three distinct types of recycled aggregates: recycled glass sands (RGSs), recycled clay brick sands (RCBSs), and recycled concrete sands (RCSs). Of particular interest, NT@RGS, with its properties of an inherently smooth surface texture and low water absorption, was found to exert a favorable influence on the rheological behavior of concrete, manifested in reduced yield stress, thereby underscoring the potential for fine-tuning mix designs to enhance workability. As the substitution levels of NT@RGS and NT@RCBS escalated, an initial decrement in compressive strength was discernible, which subsequently reversed to strength restoration at optimized substitution ratios. This phenomenon is attributed to the synergistic interplay among NT@RA components. Remarkably, NT@RA-incorporated concrete demonstrated unparalleled self-cleaning abilities, surpassing the performance of concrete with direct nano-TiO2 powder incorporation. This comprehensive research contributes significantly to the advancement in sustainable, high-performance photocatalytic construction materials within the realm of concrete technology. It underscores the potential for enhancing not only the rheological and mechanical properties but also the environmental responsiveness of concrete through the innovative utilization of NT@RAs. Full article
Show Figures

Figure 1

Figure 1
<p>Rheological properties of photocatalytic regenerated mortar: NT-RCBS (<b>left</b>); NT-RCA (<b>right</b>).</p>
Full article ">Figure 2
<p>Micrographs of the NT-RGS and NT-RCA: (<b>a</b>) micrograph of NT-RGS; (<b>b</b>,<b>c</b>) EDX map of NT-RGS; (<b>d</b>) micrograph of NT-RCA; (<b>e</b>,f) EDX map NT-RCA</p>
Full article ">Figure 3
<p>Effect of combination replacement rate of different types of photocatalytic recycled fine aggregate on 28 d compressive strength of mortar (RCBS + RGS).</p>
Full article ">Figure 4
<p>The 28 d compressive strength of mortar prepared by different photocatalytic regeneration fine aggregate replacement rates (RCBS/RCA: RGS = 1:1): (<b>a</b>) RCBS; (<b>b</b>) RCA</p>
Full article ">Figure 5
<p>The micro-morphologies of mortar prepared by different kinds of photocatalysts and different types of photocatalyzed fine aggregates were obtained: (<b>a</b>,<b>b</b>) micrographs of RCBS; (<b>c</b>,<b>d</b>) micrographs of RCA</p>
Full article ">Figure 6
<p>Degradation of methyl blue.</p>
Full article ">
11 pages, 2254 KiB  
Article
The Impact of Substrate Temperature on the Adhesion Strength of Electroplated Copper on an Al-Doped ZnO/Si System
by Jiun-Yi Tseng, Wen-Jauh Chen and Ping-Hang Chen
Materials 2024, 17(20), 4953; https://doi.org/10.3390/ma17204953 - 10 Oct 2024
Viewed by 379
Abstract
This research, which involved a comprehensive methodology, including depositing electroplated copper on a copper seed layer and Al-doped ZnO (AZO) thin films on textured silicon substrates using DC magnetron sputtering with varying substrate heating, has yielded significant findings. The study thoroughly investigated the [...] Read more.
This research, which involved a comprehensive methodology, including depositing electroplated copper on a copper seed layer and Al-doped ZnO (AZO) thin films on textured silicon substrates using DC magnetron sputtering with varying substrate heating, has yielded significant findings. The study thoroughly investigated the effects of substrate temperature (Ts) on copper adhesion strength and morphology using the peel force test and electron microscopy. The peel force test was conducted at angles of 90°, 135°, and 180°. The average adhesion strength was about 0.2 N/mm for the samples without substrate heating. For the samples with substrate heating at 100 °C, the average peeling force of the electroplated copper film was about 1 N/mm. The average peeling force increased to 1.5 N/mm as the substrate heating temperature increased to 200 °C. The surface roughness increases as the annealing temperature of the Cu/AZO/Si sample increases. These findings not only provide a reliable and robust method for applying AZO transparent conductive films onto silicon solar cells but also underscore its potential to significantly enhance the efficiency and durability of solar cells significantly, thereby instilling confidence in the field of solar cell technology. Full article
(This article belongs to the Section Energy Materials)
Show Figures

Figure 1

Figure 1
<p>The SEM images of (<b>a</b>) the as-deposited sample, (<b>b</b>) Cu/AZO/Si(100), and (<b>c</b>) Cu/AZO/Si(200).</p>
Full article ">Figure 2
<p>SEM cross-sectional image of the EP-Cu/Cu/AZO/Si(RT) sample prepared by (<b>a</b>) cleavage of the test sample and (<b>b</b>) FIB.</p>
Full article ">Figure 3
<p>Peel force diagrams for the (<b>a</b>) EP-Cu/Cu/AZO/Si(RT), (<b>b</b>) EP-Cu/Cu/AZO/Si(100), and (<b>c</b>) EP-Cu/Cu/AZO/Si(200) samples, which were tested at an angle of 135° with a constant speed of 30 mm/min.</p>
Full article ">Figure 4
<p>The relationship between peel force and substrate heating at test angles of (<b>a</b>) 135°, (<b>b</b>) 90°, and (<b>c</b>) 180°.</p>
Full article ">Figure 5
<p>The surface morphologies of the substrates after the peel test for the as-deposited EP-Cu/Cu/AZO/Si sample (<b>a</b>,<b>b</b>) and the samples annealed at 100 °C (<b>c</b>,<b>d</b>) and 200 °C (<b>e</b>,<b>f</b>). The arrow indicates the location of area 2 in <a href="#materials-17-04953-f005" class="html-fig">Figure 5</a>e.</p>
Full article ">Figure 6
<p>SEM images for the (<b>a</b>) EP-Cu/Cu/AZO/Si(RT), (<b>b</b>) EP-Cu/Cu/AZO/Si(100), and (<b>c</b>) EP-Cu/Cu/AZO/Si(200) samples after peel test. The numbers 1, 2, and 3 in (<b>a</b>–<b>c</b>) represented the zones 1, 2, and 3 of morphology. EDS spectrum for (<b>d</b>) zone 1, (<b>e</b>) zone 2, and (<b>f</b>) zone 3.</p>
Full article ">Scheme 1
<p>Schematic drawing of the peel test.</p>
Full article ">
Back to TopTop