[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
remove_circle_outline
remove_circle_outline

Search Results (3,015)

Search Parameters:
Keywords = crystal growth

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 5555 KiB  
Article
The Introduction of a BaTiO3 Polarized Coating as an Interface Modification Strategy for Zinc-Ion Batteries: A Theoretical Study
by Diantao Chen, Jiawei Zhang, Qian Liu, Fan Wang, Xin Liu and Minghua Chen
Int. J. Mol. Sci. 2024, 25(20), 11172; https://doi.org/10.3390/ijms252011172 (registering DOI) - 17 Oct 2024
Abstract
Aqueous zinc-ion batteries (AZIBs) have become a promising and cost-effective alternative to lithium-ion batteries due to their low cost, high energy, and high safety. However, dendrite growth, hydrogen evolution reactions (HERs), and corrosion significantly restrict the performance and scalability of AZIBs. We propose [...] Read more.
Aqueous zinc-ion batteries (AZIBs) have become a promising and cost-effective alternative to lithium-ion batteries due to their low cost, high energy, and high safety. However, dendrite growth, hydrogen evolution reactions (HERs), and corrosion significantly restrict the performance and scalability of AZIBs. We propose the introduction of a BaTiO3 (BTO) piezoelectric polarized coating as an interface modification strategy for ZIBs. The low surface energy of the BTO (110) crystal plane ensures its thermodynamic preference during crystal growth in experimental processes and exhibits very low reactivity toward oxidation and corrosion. Calculations of interlayer coupling mechanisms reveal a stable junction between BTO (110) and Zn (002), ensuring system stability. Furthermore, the BTO (110) coating also effectively inhibits HERs. Diffusion kinetics studies of Zn ions demonstrate that BTO effectively suppresses the dendrite growth of Zn due to its piezoelectric effect, ensuring uniform zinc deposition. Our work proposes the introduction of a piezoelectric material coating into AZIBs for interface modification, which provides an important theoretical perspective for the mechanism of inhibiting dendrite growth and side reactions in AZIBs. Full article
(This article belongs to the Section Physical Chemistry and Chemical Physics)
Show Figures

Figure 1

Figure 1
<p>A schematic illustration of the action mechanism of an artificially introduced solid electrolyte interphase (SEI) in BTO (110). (<b>a</b>) The unpolarized phase and (<b>b</b>) the polarized phase, highlighting how dendrite formation induces interface pressure on BTO. This pressure triggers a piezoelectric effect, resulting in substantial polarization which effectively restricts Zn-ion diffusion and further suppresses dendrite growth. (<b>c</b>) A schematic representation of the HER inhibition mechanism.</p>
Full article ">Figure 2
<p>The surface energies of different terminations for the (110) surface of BTO: (<b>a</b>) the surface energy of the Ba-terminated (110) surface; (<b>b</b>) the surface energy of the TiO-terminated (110) surface; (<b>c</b>) the surface energy of the O-terminated (110) surface. The deep blue represents the barium atom, the light blue represents the titanium atom, and the red represents the oxygen atom.</p>
Full article ">Figure 3
<p>The interaction mechanism between BTO (110) and Zn (002). (<b>a</b>) A schematic of the BTO (110) and Zn (002) heterojunction, including the corresponding charge density difference (CDD) map and the interlayer binding energy between BTO (110) and Zn (002). (<b>b</b>) The plane-averaged CDD along the z direction. The observed extensive charge transfer between BTO (110) and Zn (002) layers indicates a strong Coulombic interaction. In the figure, the purple sphere is zinc atom, the dark blue is barium atom, the light blue is titanium atom, and the red is oxygen atom.</p>
Full article ">Figure 4
<p>Calculation of the electron tunneling barrier (ΔE) by aligning the Fermi level (E<sub>f</sub>), work function (Φ), and band gap (E<sub>g</sub>) of the Zn anode and BTO–solid electrolyte interphase (SEI). The work function of Zn (ΦZn) is 0.76 eV, the work function of BTO (ΦBTO) is 3.49 eV, resulting in a calculated tunneling barrier of ΔE = 2.73 eV, and the band gap of BTO–bulk is 3.2 eV.</p>
Full article ">Figure 5
<p>Analysis of the diffusion mechanism of Zn ions in BTO (110). (<b>a</b>) The diffusion barrier for Zn ions along the (110) direction in the non-polarized phase. (<b>b</b>) The diffusion barrier for Zn ions along the (110) direction in the polarized phase, with 5% applied strain to simulate the piezoelectric effect. It is observed that the piezoelectric effect induces significant polarization in the material, leading to a higher diffusion barrier that suppresses further growth of Zn dendrites compared to the non-polarized phase.</p>
Full article ">Figure 6
<p>The hydrogen adsorption free energy for the HER on the BTO (110) surface and Zn (002) surface. H* represents the intermediate state of hydrogen ions adsorbed on the surface.</p>
Full article ">Figure 7
<p>The interaction mechanism between the BTO (110) O1a vacancy and Zn (002). (<b>a</b>) The schematic diagram of the BTO (110) O1a vacancy and the Zn (002) heterojunction, including the corresponding charge density difference (CDD) diagram and the interlayer binding energy between the BTO (110) O1a vacancy and Zn (002). (<b>b</b>) The average CDD along the z direction. The observed extensive charge transfer between the BTO (110) O1a vacancy and the Zn (002) layer indicates a strong Coulomb interaction. In the figure, the purple sphere is zinc atom, the dark blue is barium atom, the light blue is titanium atom, and the red is oxygen atom.</p>
Full article ">Figure 8
<p>The interaction mechanism between the BTO (110) O2d vacancy and Zn (002). (<b>a</b>) The schematic diagram of the BTO (110) O2d vacancy and the Zn (002) heterojunction, including the corre-sponding charge density difference (CDD) diagram and the interlayer binding energy between the BTO (110) O2d vacancy and Zn (002). (<b>b</b>) The average CDD along the z direction. The observed extensive charge transfer between the BTO (110) O2d vacancy and the Zn (002) layer indicates a strong Coulomb interaction. In the figure, the purple sphere is zinc atom, the dark blue is barium atom, the light blue is titanium atom, and the red is oxygen atom.</p>
Full article ">Figure 9
<p>Analysis of the diffusion mechanism of Zn ions in BTO (110). (<b>a</b>) The diffusion barrier for Zn ions along the (110) O1a vacancy direction in the non-polarized phase. (<b>b</b>) The diffusion barrier for Zn ions along the (110) O2d vacancy direction in the non-polarized phase.</p>
Full article ">
18 pages, 12177 KiB  
Article
Isotopically Enriched Lithium Fluoride Crystals for Detection of Neutrons with the Fluorescent Track Technique
by Małgorzata Sankowska, Paweł Bilski, Mariusz Kłosowski, Anna Kilian, Wojciech Gieszczyk and Barbara Marczewska
Materials 2024, 17(20), 5029; https://doi.org/10.3390/ma17205029 (registering DOI) - 14 Oct 2024
Viewed by 325
Abstract
In this work, the properties of LiF crystals grown using Li of different isotopic compositions are described from the standpoint of their application as fluorescent nuclear track detectors used in measurements in the neutron radiation fields. The crystals were grown using two techniques: [...] Read more.
In this work, the properties of LiF crystals grown using Li of different isotopic compositions are described from the standpoint of their application as fluorescent nuclear track detectors used in measurements in the neutron radiation fields. The crystals were grown using two techniques: the Czochralski method and the micro-pulling-down method. Three isotopic compositions of Li were studied: natural, highly enriched in 6Li, and highly enriched in 7Li. It was found that 6LiF detectors are about six times more sensitive to thermal (low energy) neutrons than natural LiF, which significantly decreases the lower detection limit. 7LiF detectors are insensitive to thermal neutrons, which makes it easier to detect tracks due to other radiation modalities, such as energetic ions or nuclei recoiled in collisions with high-energy neutrons. Besides the response to neutron radiation, no other significant differences in the crystal properties were identified, irrespective of the isotopic composition and crystal growth method employed. Full article
Show Figures

Figure 1

Figure 1
<p>Photoluminescence excitation (PLE) and emission (PL) spectra for LiF crystal after exposure to Sr-90/Y-90 beta source. The emission spectrum was measured using a 505 nm long-pass filter to cut off the excitation light.</p>
Full article ">Figure 2
<p>Examples of the fluorescent tracks created by the products of (n,α) reaction at <sup>6</sup>Li: alpha particle and tritium nucleus in LiF crystal enriched with <sup>6</sup>Li. Thermal neutron fluence around 8.6 × 10<sup>4</sup> n.cm<sup>−2</sup>. Maximum intensity projection from images taken at depths from 13 µm below surface to 18 µm below the crystal surface (see <a href="#sec2dot3-materials-17-05029" class="html-sec">Section 2.3</a> for details).</p>
Full article ">Figure 3
<p>Examples of LiF crystals grown with (<b>a</b>) Czochralski method; (<b>b</b>) µPD method. Panel (<b>c</b>) shows a comparison between FNTD detectors cut from crystals produced with both methods.</p>
Full article ">Figure 4
<p>Examples of images registered after irradiations with thermal neutrons. Maximum intensity projection of 21 images taken at depths ranging from 10 µm to 30 µm below the crystal surface for (<b>a</b>) natural LiF, neutron fluence c.a. 8.6 × 10<sup>4</sup> n.cm<sup>−2</sup>; (<b>b</b>) LiF enriched with <sup>6</sup>Li, neutron fluence c.a. 8.6 × 10<sup>4</sup> n.m<sup>−2</sup>; (<b>c</b>) natural LiF, neutron fluence c.a. 4.0 × 10<sup>5</sup> n.cm<sup>−2</sup>; (<b>d</b>) LiF enriched with <sup>6</sup>Li, neutron fluence c.a. 4.0 × 10<sup>5</sup> n.cm<sup>−2</sup>.</p>
Full article ">Figure 5
<p>Relationship between a number of the registered tracks per field of view and the thermal neutron fluence for detectors made of natural and <sup>6</sup>Li-enriched LiF. One track per field of view corresponds to about 7000 tracks per mm<sup>3</sup>.</p>
Full article ">Figure 6
<p>Total fluorescent intensity in the field of view versus thermal neutron fluence for detectors made of LiF enriched with <sup>6</sup>Li. Calculations were made for the maximum intensity projection of 21 images taken at depths ranging from 10 µm to 30 µm below the crystal surface.</p>
Full article ">Figure 7
<p>Examples of images registered after irradiations with different individual equivalent doses of neutrons for natural (<b>a</b>,<b>c</b>) and <sup>6</sup>Li-enriched (<b>b</b>,<b>d</b>) detectors. Maximum intensity projection of 21 images taken at depths ranging from 6 to 26 µm under the crystal surface.</p>
Full article ">Figure 8
<p>Number of tracks per field of view versus neutron dose equivalent Hp(10) for detectors made of natural and <sup>6</sup>Li-enriched LiF. The inset graph shows in detail the results in the range up to 2 mSv.</p>
Full article ">Figure 9
<p>Examples of images registered after irradiations with high energy neutrons at CERF for (<b>a</b>) natural LiF crystals and (<b>b</b>) <sup>7</sup>Li-enriched LiF crystals. Maximum intensity projection of 21 images taken at depths ranging from 6 to 26 µm under the crystal surface.</p>
Full article ">Figure 10
<p>Example of a very long track (about 220 µm) registered with <sup>7</sup>Li-enriched LiF FNTD detector at CERF. The image consists of a superposition of three maximum intensity projections of image stacks (depth from 11 µm to 15 µm under crystal surface) acquired in the adjacent areas. The image background was numerically subtracted.</p>
Full article ">Figure 11
<p>Comparison of images acquired after irradiation with fast neutrons from Pu-Be source for (<b>a</b>) natural LiF, (<b>b</b>) LiF enriched with <sup>6</sup>Li, and (<b>c</b>) LiF enriched with <sup>7</sup>Li. Maximum intensity projection of 21 images taken at depths ranging from 10 to 30 µm under the crystal surface.</p>
Full article ">Figure 12
<p>Comparison of images obtained after 6-month exposure of FNTDs on the International Space Station in Earth orbit: (<b>a</b>) natural LiF and (<b>b</b>) LiF enriched with <sup>7</sup>Li. The pictures show maximum intensity projection of 21 images taken at depths ranging from the crystal surface to 21 µm under the surface. Acquisition time for a single image was 20 s.</p>
Full article ">Figure 13
<p>Example of an image for <sup>7</sup>Li-enriched LiF crystal after exposure during the flight to the lunar orbit in the frame of the MARE experiment (Artemis-1 mission). Maximum intensity projection of 21 images taken at depths ranging from 6 to 26 µm under the crystal surface. The acquisition time for a single image was 30 s.</p>
Full article ">Figure 14
<p>Influence of post-irradiation annealing at 290 °C on track intensity. Maximum intensity projection for stacks of images acquired at depths ranging from the surface to 20 µm in 1 µm steps for the same LiF crystal before and after heating at 290 °C. The acquisition time for a single image 5 s. The sample was irradiated with thermal neutrons (moderated Pu-Be source). The brightness, contrast, and other graphic parameters of both images are the same.</p>
Full article ">
15 pages, 4011 KiB  
Article
Performance Enhancement of Hole Transport Layer-Free Carbon-Based CsPbIBr2 Solar Cells through the Application of Perovskite Quantum Dots
by Qi Yu, Wentian Sun and Shu Tang
Nanomaterials 2024, 14(20), 1651; https://doi.org/10.3390/nano14201651 - 14 Oct 2024
Viewed by 419
Abstract
CsPbIBr2, with its suitable bandgap, shows great potential as the top cell in tandem solar cells. Nonetheless, its further development is hindered by a high defect density, severe carrier recombination, and poor stability. In this study, CsPbI1.5Br1.5 quantum [...] Read more.
CsPbIBr2, with its suitable bandgap, shows great potential as the top cell in tandem solar cells. Nonetheless, its further development is hindered by a high defect density, severe carrier recombination, and poor stability. In this study, CsPbI1.5Br1.5 quantum dots were utilized as an additive in the ethyl acetate anti-solvent, while a layer of CsPbBr3 QDs was introduced between the ETL and the CsPbIBr2 light-harvester film. The combined effect of these two QDs enhanced the nucleation, crystallization, and growth of CsPbIBr2 perovskite, yielding high-quality films characterized by an enlarged crystal size, reduced grain boundaries, and smooth surfaces. And a wider absorption range and better energy band alignment were achieved owing to the nano-size effect of QDs. These improvements led to a decreased defect density and the suppression of non-radiative recombination. Additionally, the presence of long-chain organic molecules in the QD solution promoted the formation of a hydrophobic surface, significantly enhancing the long-term stability of CsPbIBr2 PSCs. Consequently, the devices achieved a PCE of 9.20% and maintained an initial efficiency of 85% after 60 days of storage in air. Thus, this strategy proves to be an effective approach for the preparation of efficient and stable CsPbIBr2 PSCs. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of QDs prepared by the solvothermal method (<b>a</b>); images of CsPbI<sub>1.5</sub>Br<sub>1.5</sub> QDs and CsPbBr<sub>3</sub> QDs in sunlight (<b>b</b>,<b>c</b>), and UV light (<b>d</b>,<b>e</b>); XRD patterns of CsPbI<sub>1.5</sub>Br<sub>1.5</sub> (green) and CsPbBr<sub>3</sub> (orange) QDs (<b>f</b>); UV-vis absorption (upper) and PL spectra (lower) of CsPbI<sub>1.5</sub>Br<sub>1.5</sub> QDs (<b>g</b>) and UV-vis absorption (green) and PL spectra (orange) of CsPbBr<sub>3</sub> QDs (<b>h</b>).</p>
Full article ">Figure 2
<p>SEM image (<b>a</b>) and particle size statistics (<b>b</b>) of CsPbI<sub>1.5</sub>Br<sub>1.5</sub> QDs; SEM images of 0% (<b>c</b>,<b>d</b>), 20% (<b>e</b>,<b>f</b>), 25% (<b>g</b>,<b>h</b>), and 30% (<b>i</b>,<b>j</b>) CsPbI<sub>1.5</sub>Br<sub>1.5</sub> QD additive in EA.</p>
Full article ">Figure 3
<p>XRD patterns (<b>a</b>) and the best-performing devices’ J–V curves (<b>b</b>) after varying CsPbI<sub>1.5</sub>Br<sub>1.5</sub> QD additive amount modification; the EQE and integrated Jsc curves (<b>c</b>) and stability tests (<b>d</b>) of the unoptimized and the 25%-treated devices.</p>
Full article ">Figure 4
<p>SEM image (<b>a</b>) and particle size distribution (<b>b</b>) of CsPbBr<sub>3</sub> QDs; SEM images of 0 (<b>c</b>,<b>d</b>), 0.2 (<b>e</b>,<b>f</b>), 0.4 (<b>g</b>,<b>h</b>), and 0.6 mg/mL (<b>i</b>,<b>j</b>) CsPBr<sub>3</sub> QD layers on the TiO<sub>2</sub> films.</p>
Full article ">Figure 5
<p>CsPbIBr<sub>2</sub> PSC configuration (<b>a</b>); J–V curves of devices modified by different concentrations of CsPbBr<sub>3</sub> QDs (<b>b</b>); the EQE curves (<b>c</b>), Nyquist plots, inset: equivalent circuit diagram (<b>d</b>), stability tests (<b>e</b>) of devices optimized with 0 and 0.4 mg/mL CsPbBr<sub>3</sub> QD layers and the reproducibility of the devices (<b>f</b>).</p>
Full article ">Figure 6
<p>UV-vis absorption spectra of CsPbIBr<sub>2</sub> films treated by CsPbI<sub>1.5</sub>Br<sub>1.5</sub> QDs (<b>a</b>) and CsPbBr<sub>3</sub> QDs (<b>b</b>).</p>
Full article ">
25 pages, 4840 KiB  
Article
Mechanism of Intracellular Elemental Sulfur Oxidation in Beggiatoa leptomitoformis, Where Persulfide Dioxygenase Plays a Key Role
by Tatyana S. Rudenko, Liubov I. Trubitsina, Vasily V. Terentyev, Ivan V. Trubitsin, Valentin I. Borshchevskiy, Svetlana V. Tishchenko, Azat G. Gabdulkhakov, Alexey A. Leontievsky and Margarita Yu. Grabovich
Int. J. Mol. Sci. 2024, 25(20), 10962; https://doi.org/10.3390/ijms252010962 (registering DOI) - 11 Oct 2024
Viewed by 301
Abstract
Representatives of the colorless sulfur bacteria of the genus Beggiatoa use reduced sulfur compounds in the processes of lithotrophic growth, which is accompanied by the storage of intracellular sulfur. However, it is still unknown how the transformation of intracellular sulfur occurs in Beggiatoa [...] Read more.
Representatives of the colorless sulfur bacteria of the genus Beggiatoa use reduced sulfur compounds in the processes of lithotrophic growth, which is accompanied by the storage of intracellular sulfur. However, it is still unknown how the transformation of intracellular sulfur occurs in Beggiatoa representatives. Annotation of the genome of Beggiatoa leptomitoformis D-402 did not identify any genes for the oxidation or reduction of elemental sulfur. By searching BLASTP, two putative persulfide dioxygenase (PDO) homologs were found in the genome of B. leptomitoformis. In some heterotrophic prokaryotes, PDO is involved in the oxidation of sulfane sulfur. According to HPLC-MS/MS, the revealed protein was reliably detected in a culture sample grown only in the presence of endogenous sulfur and CO2. The recombinant protein from B. leptomitoformis was active in the presence of glutathione persulfide. The crystal structure of recombinant PDO exhibited consistency with known structures of type I PDO. Thus, it was shown that B. leptomitoformis uses PDO to oxidize endogenous sulfur. Additionally, on the basis of HPLC-MS/MS, RT-qPCR, and the study of PDO reaction products, we predicted the interrelation of PDO and Sox-system function in the oxidation of endogenous sulfur in B. leptomitoformis and the connection of this process with energy metabolism. Full article
(This article belongs to the Special Issue Current Research on Omics of Microorganisms)
Show Figures

Figure 1

Figure 1
<p>Organization of genes encoding enzymes for dissimilatory sulfide oxidation in the genome of <span class="html-italic">B. leptomitoformis</span>. <span class="html-italic">fccA</span>—cytochrome c subunit of flavocytochrome c sulfide dehydrogenase (EC 1.8.2.3; AUI67863.1); <span class="html-italic">fccB</span>—sulfide dehydrogenase (flavocytochrome c) flavoprotein chain (EC 1.8.2.3; AUI67862.2); <span class="html-italic">tauE</span>—sulfite exporter TauE/SafE family protein (AUI67490.1); <span class="html-italic">soxB</span>—thiosulfohydrolase SoxB (EC 3.1.6.20; AUI67367.1); <span class="html-italic">rhd</span>—rhodanese (sulfur transferase) (EC 2.8.1.1; AUI70268.1); <span class="html-italic">tusA</span>—sulfur carrier protein TusA (AUI70013.1); <span class="html-italic">dsrE</span>—DsrE/DsrF/DrsH-like family protein (AUI70012.1); <span class="html-italic">soxZ</span>—thiosulfate oxidation carrier complex protein SoxZ (EC 1.8.2.6; AUI69817.1); <span class="html-italic">soxY</span>—thiosulfate oxidation carrier protein SoxY (EC 1.8.2.6; AUI69816.1); <span class="html-italic">soeABC</span>—membrane-bound cytoplasmic sulfite:quinone oxidoreductase SoeABC (EC 1.8.5.6; AUI70655.1, AUI69788.1, AUI69789.1); <span class="html-italic">soxAX</span>—sulfur oxidation c-type cytochrome (EC 1.8.2.3; AUI69545.1); <span class="html-italic">pdo</span>—persulfide dioxygenase (EC 1.13.11.18; AUI69048.1); <span class="html-italic">sqrF</span>—sulfide:quinone oxidoreductase type VI SqrF (EC 1.8.5.4; AUI68992.1); <span class="html-italic">sqrA</span>—sulfide:quinone oxidoreductase type I SqrA (EC 1.8.5.4; AUI68548.1).</p>
Full article ">Figure 2
<p>Phylogenetic tree based on the predicted amino acid sequences of PDOs. Clustering sequences were deduced using the neighbor-joining method. The amino acid sequences of related proteins from the superfamily of metallo-β-lactamase, glyoxalase II proteins were used as an outgroup. Phylogenetic analysis was performed based on a representative selection of amino acids sequences of PDOs from the study of Xia et al. [<a href="#B14-ijms-25-10962" class="html-bibr">14</a>]. The phylogenetic analysis included 110 amino acid sequences, including sequences for members of the family <span class="html-italic">Beggiatoaceae</span> identified by BLASTP searches in this study. Two amino acid sequences, presumably PDOs, found in this study in <span class="html-italic">B. leptomitoformis</span> are highlighted in red font. The type I PDO cluster is shown in a green semicircle, and the glyoxalase II proteins cluster in dark purple; type II and III PDOs are condensed within two separate branches. The protein accession numbers are indicated on the tree.</p>
Full article ">Figure 3
<p>SDS-PAGE of recombinant PDO from <span class="html-italic">B. leptomitoformis</span>. The concentrating gel is 5% acrylamide, the separating gel is 12% acrylamide; colored with Coomassie brilliant blue R250. Mw, molecular weight marker (kDa) (Thermo Fisher Scientific, Waltham, MA, USA). The arrow shows the band with the target protein PDO.</p>
Full article ">Figure 4
<p>The effects of pH (<b>a</b>) and temperature (<b>b</b>) on the activity of recombinant PDO of <span class="html-italic">B. leptomitoformis</span> in the presence of GSSH. The bars on the curves show the standard deviations of the triplicate measurements.</p>
Full article ">Figure 5
<p>Stability assay of recombinant PDO of <span class="html-italic">B. leptomitoformis</span> at different temperatures in the presence of GSSH. Residual activity assay was determined in 50 mM KPi buffer (pH 7.0) at 35 °C. The bars on the curves show the standard deviations of the triplicate measurements.</p>
Full article ">Figure 6
<p>Stability assay of recombinant PDO of <span class="html-italic">B. leptomitoformis</span> at different pH values in the presence of GSSH. Residual activity assay was determined in 50 mM KPi buffer (pH 7.0) at 35 °C. The bars on the curves show the standard deviations of the triplicate measurements.</p>
Full article ">Figure 7
<p>General view of the crystal structure of PDO from <span class="html-italic">B. leptomitoformis</span> (pdb id 8ZBD) and superimposed structures of PDO from <span class="html-italic">B. leptomitoformis</span> and hETHE1 (pdb id 4CHL). (<b>a</b>) <span class="html-italic">B. leptomitoformis</span> PDO dimer; secondary structure elements in one of the monomers are shown in different colors (β-sheets in red, α-helixes in blue, and loops in purple). (<b>b</b>) From left to right: Cα-atom trajectories of PDO monomers from <span class="html-italic">B. leptomitoformis</span> in dimer (orange) and Cα-atom trajectories of hETHE1 monomers (blue).</p>
Full article ">Figure 8
<p>Superposition of <span class="html-italic">B. leptomitoformis</span> PDO active site structures (orange) and hETHE1 (blue). The iron ions and water molecules (W1, W2, W3) which coordinate (dashed lines) to the iron are shown as spheres.</p>
Full article ">Figure 9
<p>Expression levels of the <span class="html-italic">pdo</span> gene (<b>A</b>) and Sox-system genes (<b>B</b>) in <span class="html-italic">B. leptomitoformis</span> D-402 during chemolithoautotrophic growth in the presence of Na<sub>2</sub>S, during the exponential growth phase (3 days) (a), organoheterotrophic growth, during the exponential growth phase (3 days) (b), chemolithoautotrophic growth in the presence of intracellular sulfur, during 4 days of growth (c), during 7 days of growth (d), and during 24 days of growth (e), and chemolithoautotrophic growth in the presence of intracellular sulfur, on day 18 of growth (f).</p>
Full article ">Figure 10
<p>Intermediates detected during the chemolithoautotrophic growth of <span class="html-italic">B. leptomitoformis</span> D-402 in the presence of intracellular sulfur. Thiosulfate concentration is shown by gray curves and sulfate concentration by black curves. The bars on the curves show the standard deviations of triplicate chemical measurements. The diagram shows nine points representing 600 h (25 days) of incubation of the strain D-402 in the presence of intracellular elemental sulfur and CO<sub>2</sub> alone.</p>
Full article ">Figure 11
<p>Superposition of the active site structures of PDO from <span class="html-italic">B. leptomitoformis</span> (orange), hETHE1 (cyan), and <span class="html-italic">Bp</span>PRF (lilac). Panels (<b>a</b>,<b>b</b>) show different views of the active sites of the three type I PDOs.</p>
Full article ">Figure 12
<p>Superposition of the active site structures of PDO <span class="html-italic">B. leptomitoformis</span> (orange) and <span class="html-italic">Mx</span>PDOI (green). The iron ion is shown as a sphere.</p>
Full article ">Figure 13
<p>A hypothetical of sulfide oxidation scheme for <span class="html-italic">B. leptomitoformis</span> D-402. The black dashed line shows non-enzymatic reactions, the red dashed line shows reactions whose mechanism is still unknown, and the purple dashed line is not shown for <span class="html-italic">B. leptomitoformis</span>. FCSD, flavocytochrome c sulfide dehydrogenase; PDO, persulfide dioxygenase; SQR, sulfide:quinone oxidoreductase; GSH, glutathione; GSSH, glutathione persulfide. The figure was adapted from Xin et al., 2020 [<a href="#B15-ijms-25-10962" class="html-bibr">15</a>], and Dahl, 2020 [<a href="#B26-ijms-25-10962" class="html-bibr">26</a>], under the terms of the Creative Commons Attribution 4.0 International License.</p>
Full article ">
14 pages, 1867 KiB  
Article
Spectroscopic Properties of TmF3-Doped CaF2 Crystals
by Carla Schornig, Marius Stef, Gabriel Buse, Maria Poienar, Philippe Veber and Daniel Vizman
Materials 2024, 17(20), 4965; https://doi.org/10.3390/ma17204965 - 11 Oct 2024
Viewed by 274
Abstract
In this study, we report the growth and comprehensive spectroscopic analysis of TmF3-doped CaF2 crystals, grown using the vertical Bridgman method. The optical absorption and photoluminescence properties of both trivalent (Tm3+) and divalent (Tm2+) thulium ions [...] Read more.
In this study, we report the growth and comprehensive spectroscopic analysis of TmF3-doped CaF2 crystals, grown using the vertical Bridgman method. The optical absorption and photoluminescence properties of both trivalent (Tm3+) and divalent (Tm2+) thulium ions were investigated. Optical absorption spectra in the UV-VIS-NIR range reveal characteristic transitions of Tm3+ ions, as well as weaker absorption bands corresponding to Tm2+ ions. The Judd–Ofelt (JO) formalism was applied to determine the intensity parameters Ω2, Ω4, and Ω6, which were used to calculate radiative transition probabilities, branching ratios, and radiative lifetimes for the Tm3+ ions. The emission spectra showed concentration-dependent quenching effects, with significant emissions observed for the concentration of 0.1 mol% TmF3 under excitation at 260 nm and 353 nm for Tm3+ ions and at 305 nm for Tm2+ ions. A new UV emission associated with divalent Thulium is reported. The results indicate that higher TmF3 concentrations lead to increased non-radiative energy transfer, which reduces luminescence efficiency. These findings contribute to the understanding of the optical behavior of Tm-doped fluoride crystals, with implications for their application in laser technologies and radiation dosimetry. Full article
(This article belongs to the Section Optical and Photonic Materials)
Show Figures

Figure 1

Figure 1
<p>Preparation stage of the Bridgman setup to reach the melting temperature in the graphite heater. The inset shows CaF<sub>2</sub>:x mol% TmF<sub>3</sub>, (<b>a</b>) x = 0.1; (<b>b</b>) x = 1; (<b>c</b>) x = 5.</p>
Full article ">Figure 2
<p>(<b>a</b>) Absorption spectra of CaF<sub>2</sub> doped with three concentrations of TmF<sub>3</sub> (x = 0.1, 1, and 5 mol%) measured at room temperature. The spectral features correspond to transitions of Tm<sup>3+</sup> and Tm<sup>2+</sup> ions, with absorption peaks labelled according to their electronic transitions. The absorption spectrum corresponding to the concentration of 0.1 mol% TmF<sub>3</sub> is multiplied by 10 (black line); (<b>b</b>) Linear fitting of the experimental data. The values of the correlation coefficients, r<sup>2</sup>, are indicated on the graph.</p>
Full article ">Figure 3
<p>Room temperature emission spectra of CaF<sub>2</sub> crystals doped with different concentrations of Tm<sup>3</sup><sup>+</sup> ions (0.1, 1, and 5 mol%) under excitation at (<b>a</b>) 260 nm, corresponding to the <sup>3</sup>H<sub>6</sub>→<sup>3</sup>P<sub>2</sub> transition, and (<b>b</b>) 353 nm, corresponding to the <sup>3</sup>H<sub>6</sub> → <sup>1</sup>D<sub>2</sub> transition.</p>
Full article ">Figure 4
<p>Energy level diagram of Tm<sup>3+</sup> ions in CaF<sub>2</sub> crystal illustrating the observed electronic transitions corresponding to the emission peaks in the spectra. The diagram shows excitation processes (solid upward arrows) at 260 nm, leading to the population of the excited states. The subsequent radiative transitions (downward arrows) result in emission at various wavelengths.</p>
Full article ">Figure 5
<p>CIE chromaticity diagram of CaF<sub>2</sub>:Tm<sup>3+</sup> crystal under excitation (<b>a</b>) at 260 and (<b>b</b>) 356 nm.</p>
Full article ">Figure 6
<p>(<b>a</b>) Emission spectra of CaF<sub>2</sub> crystals doped with various concentrations of TmF<sub>3</sub> (0.1, 1, and 5 mol%) under 305 nm excitation, corresponding to the Tm<sup>2+</sup> transition. The prominent emission peak at 353 nm is observed, with the intensity significantly decreasing as the concentration of TmF<sub>3</sub> increases; (<b>b</b>) Emission (black-line) and excitation spectrum (blue-line) of CaF<sub>2</sub>: 0.1 mol% TmF<sub>3</sub>.</p>
Full article ">
16 pages, 5818 KiB  
Article
Biomineralization Process Inspired In Situ Growth of Calcium Carbonate Nanocrystals in Chitosan Hydrogels
by Xinyue Zeng, Zheng Zhu, Wei Chang, Bin Wu and Wei Huang
Appl. Sci. 2024, 14(20), 9193; https://doi.org/10.3390/app14209193 - 10 Oct 2024
Viewed by 757
Abstract
Biological composites such as bone, nacre, and teeth show excellent mechanical efficiency because of the incorporation of biominerals into the organic matrix at the nanoscale, leading to hierarchical composite structures. Adding a large volume of ceramic nanoparticles into an organic molecular network uniformly [...] Read more.
Biological composites such as bone, nacre, and teeth show excellent mechanical efficiency because of the incorporation of biominerals into the organic matrix at the nanoscale, leading to hierarchical composite structures. Adding a large volume of ceramic nanoparticles into an organic molecular network uniformly has been a challenge in engineering applications. However, in natural organisms, biominerals grow inside organic fibers, such as chitin and collagen, forming perfect ceramic/polymer composites spontaneously via biomineralization processes. Inspired from these processes, the in situ growth of calcium carbonate nanoparticles inside the chitosan network to form ceramic composites was proposed in the current work. The crystal growth of CaCO3 nanoparticles in the chitosan matrix as a function of time was investigated. A weight percentage of ~35 wt% CaCO3 composite was realized, resembling the high weight percentage of mineral phase in bones. Scanning and transmission electron microscopy indicated the integration of CaCO3 nanocrystals with chitosan macromolecules. By growing CaCO3 minerals inside the chitosan matrix, the elastic modulus and tensile strength increases by ~110% and ~90%, respectively. The in situ crystal growth strategy was also demonstrated in organic frameworks prepared via 3D printing, indicating the potential of fabricating ceramic/polymer composites with complicated structures, and further applications in tissue engineering. Full article
(This article belongs to the Section Chemical and Molecular Sciences)
Show Figures

Figure 1

Figure 1
<p>The design strategy, preparation process, and microstructural characteristics of chitosan/CaCO<sub>3</sub> composites. (<b>a</b>) Schematic diagram illustrating the hierarchical structure of the natural skeleton. (<b>b</b>) Schematic representation of the compositional components of the mantis shrimp’s dactyl club. (<b>c</b>) Preparation method of the chitosan/CaCO<sub>3</sub> composite hydrogel and a schematic diagram of its structure.</p>
Full article ">Figure 2
<p>Microstructures of chitosan hydrogel before and after mineralization. (<b>a</b>) Images of 3D printed pure chitosan hydrogel samples at their unmineralized state. (<b>b</b>) Images of chitosan/CaCO<sub>3</sub> composite hydrogel samples after 12 h of mineralization. (<b>c</b>) SEM images of the surface of pure chitosan hydrogel at its unmineralized state. (<b>d</b>) SEM image of the chitosan/CaCO<sub>3</sub> composite hydrogel surface after 12 h of mineralization. (<b>e</b>) SEM images of the cross-section of unmineralized chitosan hydrogel. (<b>f</b>) SEM image of the cross-section of the chitosan/CaCO<sub>3</sub> composite hydrogel after 12 h of mineralization. Nanoparticles are noticed.</p>
Full article ">Figure 3
<p>Testing and analysis of chitosan/CaCO<sub>3</sub> composite hydrogels. (<b>a</b>) XRD test curve of the chitosan/CaCO<sub>3</sub> composite hydrogel. (<b>b</b>) TGA of chitosan/CaCO<sub>3</sub> composite hydrogel. (<b>c</b>) Raman spectral profile of chitosan/CaCO<sub>3</sub> composite hydrogels. (<b>d</b>) FTIR analysis of hydrogel before and after mineralization. (<b>e</b>) WAXD diffractogram of unmineralized pure chitosan hydrogel. (<b>f</b>) WAXD diffractogram of chitosan/CaCO<sub>3</sub> composite hydrogel.</p>
Full article ">Figure 4
<p>SEM images of chitosan/CaCO<sub>3</sub> composite hydrogels after 5 min and 1 h mineralization. (<b>a</b>–<b>c</b>) show SEM images of the chitosan/CaCO<sub>3</sub> composite hydrogel cross-sections after 5 min of mineralization, with (<b>b</b>) representing the area near the surface and (<b>c</b>) representing the area near the center. (<b>d</b>–<b>f</b>) show SEM images of the chitosan/CaCO<sub>3</sub> composite hydrogel cross-sections after 1 h of mineralization, with (<b>e</b>) representing the area near the surface and (<b>f</b>) representing the area near the center.</p>
Full article ">Figure 5
<p>SEM images of chitosan/CaCO<sub>3</sub> composite hydrogel after 3 and 6 h mineralization. (<b>a</b>–<b>c</b>) depict the cross-sections after 3 h of mineralization, where (<b>b</b>) is near the surface and (<b>c</b>) is near the center. (<b>d</b>–<b>f</b>) present the cross-sections after 6 h of mineralization, with (<b>e</b>) representing the area near the surface and (<b>f</b>) representing the area near the center.</p>
Full article ">Figure 6
<p>TEM images of chitosan/CaCO<sub>3</sub> composite hydrogel. (<b>a</b>,<b>b</b>) TEM image of nanoparticles formed via in situ growth. (<b>c</b>–<b>f</b>) TEM images and diffraction patterns of chitosan/CaCO<sub>3</sub> composite hydrogel. Nanoparticles shows single crystal-like diffraction pattern. Amorphous regions are noticed inside and on the edge of the calcite nanoparticles. (<b>g</b>,<b>h</b>) TEM images of the nanocomposite coating in mantis shrimp dactyl clubs, showing the combination of chitin macromolecules and hydroxyapatite nanocrystals. Images are taken from [<a href="#B16-applsci-14-09193" class="html-bibr">16</a>] with permission.</p>
Full article ">Figure 7
<p>Schematic of in situ growth of CaCO<sub>3</sub> nanocrystals inside chitosan hydrogels via the double diffusion method. Calcium ions are diffused into the 3D-printed chitosan hydrogel first, until an equilibrium is reached. Samples are then immersed in NaHCO<sub>3</sub> solution. As the CO<sub>3</sub><sup>2−</sup> is introduced, calcite crystals start growing at the surface first. Microcrystals are formed on the surface, while crystal sizes are limited to ~100 nm inside the hydrogel due to the confinement of the chitosan molecular network.</p>
Full article ">Figure 8
<p>Stress–strain curve of chitosan/CaCO<sub>3</sub> composite hydrogel at different mineralization times. Compared to the unmineralized chitosan films, mineralized films show an increase in both stiffness and tensile strength.</p>
Full article ">
13 pages, 6854 KiB  
Article
Enhancing Ice Nucleation: The Role of Surface Roughness in Electrofreezing Using Laser Shock Processed Al6061 T6 Electrodes
by E. G. Espinosa-Yañez, G. C. Mondragón-Rodríguez, E. José-Trujillo and D. P. Luis
Appl. Sci. 2024, 14(19), 9145; https://doi.org/10.3390/app14199145 - 9 Oct 2024
Viewed by 462
Abstract
The present study investigates the impact of the electrode surface roughness on the electrofreezing of water. This research focuses on how the electrode microstructure induced by a laser treatment affects the nucleation and growth of ice crystals under controlled electric fields. For this, [...] Read more.
The present study investigates the impact of the electrode surface roughness on the electrofreezing of water. This research focuses on how the electrode microstructure induced by a laser treatment affects the nucleation and growth of ice crystals under controlled electric fields. For this, electrofreezing experiments of deionized water over electrodes with varying surface roughnesses and crystalline textures were conducted. The electrodes of the Al6061 T6 alloy were microstructured via the Laser Shock Processing (LSP) method. For this purpose, the pulse densities during the LSP process were varied (900, 1600, and 2500 pulses/cm2). The increase in pulse density was correlated to the microstructural features and average roughness of the LSP-treated Al6061 alloy. A wave-like microstructure was induced upon the LSP treatment, with roughnesses between 3.5 and 6 µm at the selected pulse densities. The results indicate that electrode roughness significantly influences the electrofreezing process. Rougher electrodes were found to increase the nucleation temperature, suggesting enhanced ice nucleation activity. These findings are attributed to the increased electric field concentration at the asperities of the rough surfaces and the (111) planes of the Al6061 alloy, which may facilitate the alignment of water molecules and the formation of critical ice nuclei. Full article
Show Figures

Figure 1

Figure 1
<p>Diagram of the experimental design applied in the electrofreezing investigation. Deionized water was placed in the sample holder and the lower electrode was cooled using a cooling plate through the Peltier effect.</p>
Full article ">Figure 2
<p>LSP treatment application process on Al6061 T6 electrodes: (<b>a</b>) motion generation by robot, (<b>b</b>) laser generation module.</p>
Full article ">Figure 3
<p>The effect of the pulse density on the surface microstructure of the Al6061 T6 alloy after the LSP treatment.</p>
Full article ">Figure 4
<p>Profile mapping of the initial Al6061 alloy surface (<b>a</b>) and after LSP, (<b>b</b>) 900 pulses/cm<sup>2</sup>, (<b>c</b>) 1600 pulses/cm<sup>2</sup>, and (<b>d</b>) 2500 pulses/cm<sup>2</sup>.</p>
Full article ">Figure 5
<p>Average roughness (R<sub>a</sub>) of the Al6061 T6 alloy electrodes after LSP treatment as a function of the pulse density.</p>
Full article ">Figure 6
<p>The effect of the pulse density during the LSP treatment on the crystalline state of the Al6061 T6 alloy.</p>
Full article ">Figure 7
<p>Cooling curves of the electrofreezing experiments using electrodes with different pulse densities without high voltage.</p>
Full article ">Figure 8
<p>Cooling curves of the electrofreezing experiments using electrodes without treatment with increasing high voltage.</p>
Full article ">Figure 9
<p>Cooling curves of the electrofreezing experiments using electrodes with 900 pulses/cm<sup>2</sup> with increasing high voltage.</p>
Full article ">Figure 10
<p>Cooling curves of the electrofreezing experiments using electrodes with 1600 pulses/cm<sup>2</sup> with increasing high voltage.</p>
Full article ">Figure 11
<p>Cooling curves of the electrofreezing experiments using electrodes with 2500 pulses/cm<sup>2</sup> with increasing high voltage.</p>
Full article ">Figure 12
<p>Nucleation temperature of deionized water vs. voltage applied to the Al6061 T6 electrodes with different pulse densities during the LSP treatment.</p>
Full article ">
27 pages, 9632 KiB  
Article
In Situ Raman Spectroscopy as a Valuable Tool for Monitoring Crystallization Kinetics in Molecular Glasses
by Roman Svoboda, Nicola Koutná, Magdalena Hynková and Marek Pakosta
Molecules 2024, 29(19), 4769; https://doi.org/10.3390/molecules29194769 - 9 Oct 2024
Viewed by 557
Abstract
The performance of in situ Raman microscopy (IRM) in monitoring the crystallization kinetics of amorphous drugs (griseofulvin and indomethacin) was evaluated using a comparison with the data obtained via differential scanning calorimetry (DSC). IRM was found to accurately and sensitively detect the initial [...] Read more.
The performance of in situ Raman microscopy (IRM) in monitoring the crystallization kinetics of amorphous drugs (griseofulvin and indomethacin) was evaluated using a comparison with the data obtained via differential scanning calorimetry (DSC). IRM was found to accurately and sensitively detect the initial stages of the crystal growth processes, including the rapid glass–crystal surface growth or recrystallization between polymorphic phases, with the reliable localized identification of the particular polymorphs being the main advantage of IRM over DSC. However, from the quantitative point of view, the reproducibility of the IRM measurements was found to be potentially significantly hindered due to inaccurate temperature recording and calibration, variability in the Raman spectra corresponding to the fully amorphous and crystalline phases, and an overly limited number of spectra possible to collect during acceptable experimental timescales because of the applied heating rates. Since theoretical simulations showed that, from the kinetics point of view, the constant density of collected data points per kinetic effect results in the smallest distortions, only the employment of the fast Raman mapping functions could advance the performance of IRM above that of calorimetric measurements. Full article
(This article belongs to the Special Issue Exclusive Feature Papers in Physical Chemistry, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) DSC crystallization peaks obtained at a low q<sup>+</sup> for the 20–50 µm GRIS powder. (<b>B</b>) DSC crystallization peaks obtained at a high <span class="html-italic">q</span><sup>+</sup> for the 20–50 µm GRIS powder. (<b>C</b>) DSC melting peaks obtained for the 20–50 µm GRIS powder; the inset shows a zooming-in on the curves measured at a low <span class="html-italic">q</span><sup>+</sup>. (<b>D</b>) Selected DSC curves zoomed in on the glass transition region; an evaluation of the half-height midpoint <span class="html-italic">T<sub>g</sub></span> is indicated. The solid black lines represent extrapolated temperature dependences of heat flow in the glassy and undercooled liquid regions; the dashed black line indicates a half distance between the black solid lines.</p>
Full article ">Figure 2
<p>(<b>A</b>) Example Raman spectra of fully amorphous and fully crystalline GRIS. (<b>B</b>) Example series of Raman spectra monitoring the crystallization of the 20–50 µm GRIS powder at <span class="html-italic">q</span><sup>+</sup> = 4.55 °C·min<sup>−1</sup>. The spectra are zoomed in on the spectral region of interest; roughly every tenth spectrum is displayed for clarity. (<b>C</b>) Raman spectra similar to those from (<b>B</b>); fine selection of the spectra (roughly every second measured spectrum) collected at the time of the amorphous-to-crystalline transition is shown. (<b>D</b>) Temperature evolution of α determined by means of the multicomponent analysis from the data depicted in (<b>B</b>). The horizontal dashed line indicates α = 0.63, which was used to evaluate the characteristic temperature, <span class="html-italic">T<sub>IRM</sub></span>.</p>
Full article ">Figure 3
<p>(<b>A</b>) Kissinger plot obtained for the DSC crystallization of the amorphous 20–50 µm GRIS powder—data shown in <a href="#molecules-29-04769-f001" class="html-fig">Figure 1</a>A,B. The quantities <span class="html-italic">T<sub>ini</sub></span>, <span class="html-italic">T<sub>p</sub></span><sub>1</sub>, and <span class="html-italic">T<sub>p</sub></span><sub>2</sub> correspond to the temperatures of the initial crystallization onset, first crystallization peak, and second crystallization peak, respectively. The dashed line indicates the estimated development of <span class="html-italic">T<sub>g</sub></span> with q<sup>+</sup>. (<b>B</b>) Kissinger plot comparing the GRIS dependencies obtained for <span class="html-italic">T<sub>ini</sub></span>, <span class="html-italic">T<sub>IRM</sub></span>, and <span class="html-italic">T<sub>g</sub></span>.</p>
Full article ">Figure 4
<p>(<b>A</b>) DSC crystallization peaks obtained at a low <span class="html-italic">q</span><sup>+</sup> for the 20–50 µm IMC powder. (<b>B</b>) DSC crystallization peaks obtained at a high <span class="html-italic">q</span><sup>+</sup> for the 20–50 µm IMC powder. (<b>C</b>) DSC melting peaks obtained at a low <span class="html-italic">q</span><sup>+</sup> for the 20–50 µm IMC powder. (<b>D</b>) DSC melting peaks obtained at a high <span class="html-italic">q</span><sup>+</sup> for the 20–50 µm IMC powder.</p>
Full article ">Figure 5
<p>(<b>A</b>) Example Raman spectra of fully amorphous and fully crystalline IMC. (<b>B</b>) Example series of Raman spectra monitoring the crystallization of the 20–50 µm IMC powder at <span class="html-italic">q</span><sup>+</sup> = 4.55 °C·min<sup>−1</sup>. The spectra are zoomed in on the spectral region of interest; roughly every fifth spectrum is displayed for clarity. (<b>C</b>) Raman spectra similar to those from (<b>B</b>); fine selection of the spectra (every measured spectrum in that time frame) collected at the time of the amorphous-to-crystalline transition is shown. (<b>D</b>) Temperature evolution of α determined by means of the multicomponent analysis from the data depicted in (<b>B</b>). The horizontal dashed line indicates α = 0.63, which was used to evaluate the characteristic temperature <span class="html-italic">T<sub>IRM</sub></span>.</p>
Full article ">Figure 6
<p>(<b>A</b>) Comparison of the DSC curves obtained for the GRIS and IMC powders at <span class="html-italic">q</span><sup>+</sup> = 6.50 °C·min<sup>−1</sup>. (<b>B</b>) Kissinger plot comparing the IMC dependencies obtained for <span class="html-italic">T<sub>ini</sub></span>, <span class="html-italic">T<sub>p</sub></span>, <span class="html-italic">T<sub>IRM</sub></span>, and <span class="html-italic">T<sub>g</sub></span>.</p>
Full article ">Figure 7
<p>(<b>A</b>,<b>B</b>) Examples of the experimental GRIS DSC curves (points) fit with the sc-MKA method (using the reaction mechanism consisting of three to four independent AC processes; red line). (<b>C</b>) Relative crystallization enthalpy of the GRIS crystallization pre-peak (normalized with respect to the overall Δ<span class="html-italic">H</span>; filled circles and left Y axis) and the overall correlation coefficients (in the form of log(1-<span class="html-italic">r</span>); half-filled triangles and right Y axis), as obtained during the sc-MKA processing of the DSC crystallization data.</p>
Full article ">Figure 8
<p>(<b>A</b>) Kissinger plot constructed for the IMC <span class="html-italic">T<sub>p</sub></span> data (points) fit with the linear dependence (the solid line). The dashed line paired with the right-hand Y axis shows the corresponding temperature dependence of the activation energy, <span class="html-italic">E</span>. (<b>B</b>) Kissinger plot constructed for the IMC <span class="html-italic">T<sub>p</sub></span> data (points) fit with the second-order polynomial dependence (the solid line). The dashed line paired with the right-hand Y axis shows the corresponding temperature dependence of the activation energy <span class="html-italic">E</span>. (<b>C</b>) Values of the pre-exponential factor <span class="html-italic">A</span> determined by the sc-MKA method for the <span class="html-italic">E</span>–<span class="html-italic">T</span> dependence shown in (<b>B</b>). The A values were determined at temperatures corresponding to the experimental points from (<b>B</b>). The solid line shows the almost perfect fit with a linear dependence, demonstrating the <span class="html-italic">E</span>–<span class="html-italic">A</span> compensation effect.</p>
Full article ">Figure 9
<p>(<b>A</b>) Temperature dependence of the rate constant (right-hand axis) calculated using the constant <span class="html-italic">E</span> and <span class="html-italic">A</span> values determined for the crystallization of IMC (see <a href="#molecules-29-04769-f008" class="html-fig">Figure 8</a>A); the graph depicts two selected DSC crystallization peaks for comparison. (<b>B</b>) Temperature dependence of the rate constant (right-hand axis) calculated using the variable <span class="html-italic">E</span> and <span class="html-italic">A</span> values determined for the crystallization of IMC (see <a href="#molecules-29-04769-f008" class="html-fig">Figure 8</a>B,C); the graph depicts two selected DSC crystallization peaks for comparison. (<b>C</b>) Example of the IMC crystallization peak processed by means of the sc-MKA method.</p>
Full article ">Figure 10
<p>(<b>A</b>) Filled-in points: α-T dependence simulated for <span class="html-italic">E</span> = 200,000 kJ·mol<sup>−1</sup>, ln(<span class="html-italic">A</span>/s<sup>−1</sup>) = 62.3, <span class="html-italic">m</span> = 3, <span class="html-italic">q</span><sup>+</sup> = 1 °C·min<sup>−1</sup>, and <span class="html-italic">t<sub>IRM</sub></span> = 83 s. Half-filled-in points: the α-<span class="html-italic">T</span> dependence delayed by subtracting <span class="html-italic">t<sub>IRM</sub></span>/2 (41.5 s). The inset shows the construction of the delayed data point from the original α-<span class="html-italic">T</span> dependence according to Equation (8). (<b>B</b>) The derivative dα·d<span class="html-italic">t</span><sup>−1</sup>-<span class="html-italic">T</span> dependencies corresponding to the similar types of points from <a href="#molecules-29-04769-f010" class="html-fig">Figure 10</a>A. The lines indicate the JMA fits of these dependencies. (<b>C</b>) Points: Kissinger dependencies constructed for the series of theoretically simulated kinetic peaks with <span class="html-italic">E</span> = 200,000 kJ·mol<sup>−1</sup>, ln(<span class="html-italic">A</span>/s<sup>−1</sup>) = 62.3, and <span class="html-italic">q</span><sup>+</sup> = 0.5, 1, 2, 5, 10, and 20 °C·min<sup>−1</sup>; the <span class="html-italic">m</span> and <span class="html-italic">t<sub>IRM</sub></span> values are listed in the legend. Solid lines: linear or polynomial fits of the Kissinger dependencies. Dashed and dotted lines (matched with the right-hand Y-axis): <span class="html-italic">E</span>–<span class="html-italic">T</span> dependencies calculated based on the fits of the Kissinger dependencies.</p>
Full article ">
19 pages, 4011 KiB  
Article
Synthesis and Preliminary Studies for In Vitro Biological Activity of Two New Water-Soluble Bis(thio)carbohydrazones and Their Copper(II) and Zinc(II) Complexes
by Alessio Zavaroni, Elena Riva, Valentina Borghesani, Greta Donati, Federica Santoro, Vincenzo Maria D’Amore, Matteo Tegoni, Giorgio Pelosi, Annamaria Buschini, Dominga Rogolino and Mauro Carcelli
Int. J. Mol. Sci. 2024, 25(19), 10831; https://doi.org/10.3390/ijms251910831 - 9 Oct 2024
Viewed by 378
Abstract
Research in the field of metallodrugs is continually increasing. However, it is often limited by the poor solubility in water of the metal complexes. To try to overcome this problem, the two new ligands bis-(sodium 3-methoxy-5-sulfonate-salicylaldehyde)thiocarbohydrazone (bis-TCH, Na2H4 [...] Read more.
Research in the field of metallodrugs is continually increasing. However, it is often limited by the poor solubility in water of the metal complexes. To try to overcome this problem, the two new ligands bis-(sodium 3-methoxy-5-sulfonate-salicylaldehyde)thiocarbohydrazone (bis-TCH, Na2H4L1) and bis-(sodium 3-methoxy-5-sulfonate-salicylaldehyde)carbohydrazone (bis-CH, Na2H4L2) were synthesized and characterized, both achieving high solubility in water. The speciation of the ligands and their coordinating behaviour towards the biologically relevant Cu(II) and Zn(II) ions were studied spectroscopically and potentiometrically, determining the pKas of the ligands and the formation constants of the complex species. The monometallic and bimetallic Cu(II) and Zn(II) complexes were isolated, and the single-crystal X-ray structure of [Cu2(NaHL1)(H2O)7].3.5H2O was discussed. Finally, preliminary studies of the in vitro cytotoxic properties of the new compounds were started on normal (Hs27) and cancer (U937) cell lines. bis-TCH was able to induce a growth inhibition effect between 40% and 45% in both cell lines; bis-CH did not produce a reduction in cell viability in Hs27 cells but revealed mild antiproliferative activity after 72 h of treatment in U937 cancer cells (GI50 = 46.5 ± 4.94 μg/mL). Coordination of the Cu(II) ions increased the toxicity of the compounds, while, in contrast, Zn(II) complexes were not cytotoxic. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>UV–visible absorption spectra for (<b>A</b>) <b>bis-TCH</b> and (<b>B</b>) <b>bis-CH</b> in aqueous solution at different pH values (pH range: 3.98–10.00; <span class="html-italic">T</span> = 298.2 K, C<sub>L</sub> = 17 µM).</p>
Full article ">Figure 2
<p>Representative distribution diagram of (<b>A</b>) <b>bis-TCH</b> and (<b>B</b>) <b>bis-CH</b> in aqueous solution (C<sub>L</sub> = 3 mM, <span class="html-italic">I</span> = 0.1 M KCl, <span class="html-italic">T</span> = 298.2 K). For each pH value (x axis), the amount of ligand in its different protonation states is reported as a % of the total ligand in solution. At physiological pH, the major species are H<sub>4</sub>L<sup>2−</sup> and H<sub>3</sub>L<sup>3−</sup>. The former correspond to the ligand deprotonated on the sulfonic groups, while the latter to the ligand deprotonated on the two sulfonic groups and one phenolic functional group.</p>
Full article ">Figure 3
<p>UV–visible spectra for the titration of <b>bis-TCH</b> with Cu(II) in aqueous solution at pH 2.5 (C<sub>L</sub> = 17 µM; Cu(II):L = 0–5). Inset: plot of the observed (circles) and calculated (dots) absorbance values at 280 nm vs. Cu(II)/ligand molar ratio.</p>
Full article ">Figure 4
<p>UV–visible spectra for the titration of <b>bis-TCH</b> with Cu(II) in aqueous solution at pH 7.4 (25 mM HEPES buffer; C<sub>L</sub> = 17 µM; Cu(II):L = 0–3). Inset: plot of the observed absorbance values at 333 nm vs. Cu(II)/ligand molar ratio.</p>
Full article ">Figure 5
<p>UV–visible spectra for the titration of <b>bis-CH</b> with Cu(II) in aqueous solution at pH 2.5 (C<sub>L</sub> = 17 µM; Cu(II):L = 0–7). Inset: plot of the observed (circles) and calculated (dots) absorbance values at 257 nm vs. Cu(II)/ligand molar ratio.</p>
Full article ">Figure 6
<p>UV–visible spectra for the titration of <b>bis-CH</b> with Cu(II) in aqueous solution at pH 7.4 (25 mM HEPES buffer; C<sub>L</sub> = 17 µM; Cu(II):L = 0–3). Inset: plot of the observed absorbance values at 264 nm vs. Cu(II)/ligand molar ratio.</p>
Full article ">Figure 7
<p>UV–visible spectra for the titration of <b>bis-TCH</b> with Zn(II) in aqueous solution at pH 7.4 (25 mM HEPES buffer; C<sub>L</sub> = 17 µM; Zn(II):L = 0–3). Inset: plot of the observed (circles) and calculated (dots) absorbance values at 300 nm vs. Zn(II)/ligand molar ratio.</p>
Full article ">Figure 8
<p>UV–visible spectra for the titration of <b>bis-CH</b> with Zn(II) in aqueous solution at pH 7.4 (25 mM HEPES buffer; C<sub>L</sub> = 17 µM; Zn(II):L = 0–3). Inset: plot of the observed (circles) and calculated (dots) absorbance values at 307 nm vs. Zn(II)/ligand molar ratio.</p>
Full article ">Figure 9
<p>Crystal structure of the ligand arranged for chelation of the two Cu(II) ions in [Cu<sub>2</sub>(NaHL<sup>1</sup>)(H<sub>2</sub>O)<sub>7</sub>]<sup>.</sup>3.5H<sub>2</sub>O (oxygen atoms in red; sulphur in yellow; nitrogen in blue; carbon in grey and hydrogens in white; copper ions in orange).</p>
Full article ">Figure 10
<p>The bond of the sodium ion with the sulfonate of one ligand molecule and two oxygens from the salicylic moiety of another ligand molecule gives rise to dimers in the crystal structure of [Cu<sub>2</sub>(NaHL<sup>1</sup>)(H<sub>2</sub>O)<sub>7</sub>]<sup>.</sup>3.5H<sub>2</sub>O (oxygen atoms in red; sulphur in yellow; nitrogen in blue; carbon in grey and hydrogens in white; copper ions in orange; sodium ions in purple).</p>
Full article ">Figure 11
<p>Dose–response curves obtained in Hs27 (<b>left</b>) and U937 cells (<b>right</b>) after 24, 48 and 72 h treatment with (<b>A</b>) <b>bis-TCH</b>; (<b>B</b>) <b>bis-CH</b>; (<b>C</b>) <b>C1</b>; (<b>D</b>) <b>C2</b>. Data are expressed as cell proliferation percentage compared with control cells.</p>
Full article ">Scheme 1
<p>Synthetic pathway to obtain the ligands <b>bis-TCH</b> (Na<sub>2</sub>H<sub>4</sub>L<sup>1</sup>) and <b>bis-CH</b> (Na<sub>2</sub>H<sub>4</sub>L<sup>2</sup>). Reagents and conditions: (<b>i</b>) aniline, ethanol, reflux, 2 h; (<b>ii</b>) concentrated sulfuric acid, 105 °C, 2 h; (<b>iii</b>) Na<sub>2</sub>CO<sub>3</sub>, H<sub>2</sub>O, 115 °C, 2 h; then, CH<sub>3</sub>COOH until pH = 5; (<b>iv</b>) thiocarbohydrazide or carbohydrazide, H<sub>2</sub>O/ethanol 1/4, reflux, 1 h 30′.</p>
Full article ">Scheme 2
<p>Proton-dissociation equilibria for the two ligands (X = S, O); species in aqueous solution in the pH range 4–10 are shown in green circles.</p>
Full article ">
14 pages, 2584 KiB  
Article
Positive Effect of Camelina Intercropping with Legumes on Soil Microbial Diversity by Applying NGS Analysis and Mobile Fluorescence Spectroscopy
by Marina Marcheva, Mariana Petkova, Vanya Slavova and Vladislav Popov
Appl. Sci. 2024, 14(19), 9046; https://doi.org/10.3390/app14199046 - 7 Oct 2024
Viewed by 660
Abstract
Camelina (Camelina sativa (L.) Crantz) is a valuable source of essential amino acids, especially sulphur-containing ones, which are generally lacking in leguminous crops, thus representing an alternative source of protein for both humans and farm animals. Rhizosphere soil samples from five experimental [...] Read more.
Camelina (Camelina sativa (L.) Crantz) is a valuable source of essential amino acids, especially sulphur-containing ones, which are generally lacking in leguminous crops, thus representing an alternative source of protein for both humans and farm animals. Rhizosphere soil samples from five experimental plots with mono- and mixed cultivations of three camelina cultivars, including two introduced varieties Cs1.Pro (Luna) and Cs2.Pro (Lenka) and one Bulgarian variety Cs3.Pro (local Bulgarian landrace) with variety 666 of vetch (Vicia sativa L.) (Cs3-Vs.Pro) and variety Mir of pea (Pisum sativum L.) (Cs3-Ps.Pro), were collected and analysed. The total DNA was isolated from the rhizosphere soils and the presence of the 16S rRNA gene was confirmed by amplification with the universal primer 16SV34. In the present study, the structure of the soil bacterial community in five different plots (Cs1.S.Pro, Cs2.S.Pro, Cs3.S.Pro, Cs3.Vs.S.Pro, and Cs3.Ps.S.Pro) where camelina was grown alone and by being intercropped with pea and vetch was analysed via a metagenomic approach. The number of observed species was highest in the local genotype of the camelina Cs3 grown alone, followed by soil from the intercropped variants Cs3-Vs and CsS-Ps. The soil bacterial communities differed between the sole cultivation of camelina and that grown with joint cultivation with vetch and peas, indicating that legumes considerably affected the growth and development of beneficial microorganisms by aspects such as nitrogen fixing, levels of nitrifying bacteria, and levels of phosphorus-dissolving bacteria, thus helping to provide better plant nutrition. The α-diversity indicated that bacterial communities in the rhizosphere were higher in soils intercropped with vetch and pea. The optical properties of cereals and legumes were determined by their energy structure, which includes both their occupied and free electronic energy levels and the energy levels of the atomic vibrations of the molecules or the crystal lattice. Full article
(This article belongs to the Special Issue Food Microbiology Safety and Quality Control)
Show Figures

Figure 1

Figure 1
<p>Mobile experimental installation used by fluorescence spectroscopy.</p>
Full article ">Figure 2
<p>(<b>A</b>) Number of observed species (97% OTUs) from 16S amplicons for each of the five soil samples. (<b>B</b>) Rank Abundance curves of alpha diversity.</p>
Full article ">Figure 3
<p>A histogram of the relative abundance of taxa at the phylum level.</p>
Full article ">Figure 4
<p>Taxonomic abundance cluster heatmap at the class level.</p>
Full article ">Figure 5
<p>Beta diversity heatmap.</p>
Full article ">Figure 6
<p>NMDS based on Weighted Unifrac distance (left) and Unweighted Unifrac distance (right). Each data point in the graph stands for a sample. The distance between data points reflects the extent of the variation. Samples belonging to the same group are in the same colour. When the value of Stress factor is 0, it is considered that the NMDS is reliable to some extent.</p>
Full article ">Figure 7
<p>Difference spectral distribution in Cs1.S, Cs2.S, and Cs3.S (self-grown plants).</p>
Full article ">Figure 8
<p>Difference spectral distribution in Cs3.S, Cs3.Vs, and Cs3.Ps (intercropped samples).</p>
Full article ">
25 pages, 3993 KiB  
Article
Structural and Dynamical Effects of the CaO/SrO Substitution in Bioactive Glasses
by Margit Fabian, Matthew Krzystyniak, Atul Khanna and Zsolt Kovacs
Molecules 2024, 29(19), 4720; https://doi.org/10.3390/molecules29194720 - 5 Oct 2024
Viewed by 483
Abstract
Silicate glasses containing silicon, sodium, phosphorous, and calcium have the ability to promote bone regeneration and biodegrade as new tissue is generated. Recently, it has been suggested that adding SrO can benefit tissue growth and silicate glass dissolution. Motivated by these recent developments, [...] Read more.
Silicate glasses containing silicon, sodium, phosphorous, and calcium have the ability to promote bone regeneration and biodegrade as new tissue is generated. Recently, it has been suggested that adding SrO can benefit tissue growth and silicate glass dissolution. Motivated by these recent developments, the effect of SrO/CaO–CaO/SrO substitution on the local structure and dynamics of Si-Na-P-Ca-O oxide glasses has been studied in this work. Differential thermal analysis has been performed to determine the thermal stability of the glasses after the addition of strontium. The local structure has been studied by neutron diffraction augmented by Reverse Monte Carlo simulation, and the local dynamics by neutron Compton scattering and Raman spectroscopy. Differential thermal analysis has shown that SrO-containing glasses have lower glass transition, melting, and crystallisation temperatures. Moreover, the addition of the Sr2+ ions decreased the thermal stability of the glass structure. The total neutron diffraction augmented by the RMC simulation revealed that Sr played a similar role as Ca in the glass structure when substituted on a molar basis. The bond length and the coordination number distributions of the network modifiers and network formers did not change when SrO (x = 0.125, 0.25) was substituted for CaO (25-x). However, the network connectivity increased in glass with 12.5 mol% CaO due to the increased length of the Si-O-Si interconnected chain. The analysis of Raman spectra revealed that substituting CaO with SrO in the glass structure dramatically enhances the intensity of the high-frequency band of 1110–2000 cm−1. For all glasses under investigation, the changes in the relative intensities of Raman bands and the distributions of the bond lengths and coordination numbers upon the SrO substitution were correlated with the values of the widths of nuclear momentum distributions of Si, Na, P, Ca, O, and Sr. The widths of nuclear momentum distributions were observed to soften compared to the values observed and simulated in their parent metal-oxide crystals. The widths of nuclear momentum distributions, obtained from fitting the experimental data to neutron Compton spectra, were related to the amount of disorder of effective force constants acting on individual atomic species in the glasses. Full article
Show Figures

Figure 1

Figure 1
<p>DTA curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 1 Cont.
<p>DTA curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 2
<p>DTA and DTG curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 2 Cont.
<p>DTA and DTG curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 3
<p>Experimental and RMC calculated structure factors of Ca25 (red), Ca 12.5(blue), and Ca0 (green) glass samples. Curves are displaced by 1 unit successively for clarity. See text for details.</p>
Full article ">Figure 4
<p>Partial atomic pair correlations for Si-O (<b>a</b>), P-O (<b>b</b>), Ca-O (<b>c</b>), Sr-O (<b>d</b>), Na-O (<b>e</b>), and O-O (<b>f</b>), in glassy samples (Ca25 (red), Ca12.5 (blue), Ca0 (green)). The peak positions corresponding to key bond lengths are shown. See text for details.</p>
Full article ">Figure 5
<p>Si-O (<b>a</b>), P-O (<b>b</b>), CaO (<b>c</b>), and O-O (<b>d</b>) coordination number distributions in the glassy samples.</p>
Full article ">Figure 6
<p>Effects of replacement of CaO with SrO on the Raman spectra of phosphosilicate glasses. See text for details.</p>
Full article ">Figure 7
<p>Fits of the TOF spectra recorded at VESUVIO for bioactive glasses Ca25, Ca12.5, and Ca0. Recoil peaks of individual atomic species in the glasses have been colour-coded, with the recoil peaks due to the aluminium container marked in blue. See text for details.</p>
Full article ">Figure 8
<p>The disorder-induced softening of the widths of nuclear momentum distributions of individual atomic species present in bioactive glasses Ca0, Ca12.5, and Ca25. The bar charts show the adopted disorder scale: (i) the blue bars designate the Maxwell-Boltzmann distribution width limits for completely disordered gas of non-interacting particles without an underlying potential; (ii) the white bars show the upper distribution width limits calculated from atom-projected vibrational densities of states of parent metal oxides; and (iii) the red bars show the widths obtained from the analysis of the NCS experiments. See text for details.</p>
Full article ">
13 pages, 6132 KiB  
Article
Study on Microstructure and Texture of Fe-3%Si Ultra-Thin Ribbons Prepared by Planar Flow Casting
by Jiangjie Xu, Ning Zhang, Yang Tu, Li Meng, Xiaozhou Zhou and Chengzhou Niu
Materials 2024, 17(19), 4893; https://doi.org/10.3390/ma17194893 - 5 Oct 2024
Viewed by 423
Abstract
In this paper, Fe-3%Si ultra-thin ribbons prepared by the planar flow casting (PFC) technique were subjected to temper rolling and annealing treatments. The microstructure and texture evolution during this process were examined through experimental measurements coupled with crystal plasticity finite element (CPFE) simulation [...] Read more.
In this paper, Fe-3%Si ultra-thin ribbons prepared by the planar flow casting (PFC) technique were subjected to temper rolling and annealing treatments. The microstructure and texture evolution during this process were examined through experimental measurements coupled with crystal plasticity finite element (CPFE) simulation to assess the feasibility of preparing ultra-thin non-oriented silicon steel using PFC ribbons. The results indicate that the PFC ribbons exhibit a significant columnar crystal structure, and {001}-oriented grains comprise over 30%. After being annealed, the grains with different orientations grew uniformly, the texture components were basically unchanged, and the {001} texture was well preserved. When annealing was carried out after temper rolling with a reduction rate of 7%, uneven grain growth was observed, and the growth tendency of the {001} grains, especially, surpassed that of the {111} grains, with an elevated temperature which peaked at 950 °C, where the proportion of {001} grains was maximal. When being annealed after temper rolling to 15%, grains of other orientations showed significant growth at each temperature, while the {001} grains did not show an obvious growth advantage. Utilizing the CPFE, the deformation-stored energy distribution of each characteristic-oriented grain was simulated, and it was shown that compared to the 15% rolling reduction rate, the deformation-stored energy accumulation of {001}-oriented grains after being rolled to 7% reduction was significantly lower than that of {111}-oriented grains. It suggests that the larger stored energy difference makes {001} grains show a stronger growth advantage based on the SIBM mechanism during annealing, after being rolled with a reduction rate of 7%. Overall, for the synergistic optimization of microstructure and texture, rolling with a 7% reduction rate followed by annealing at 950 °C in a hydrogen atmosphere is most advantageous. Full article
Show Figures

Figure 1

Figure 1
<p>Different angle pictures of planar-flow-casted Fe-3%Si ribbons (<b>a</b>,<b>b</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) Process flow diagram, and (<b>b</b>) heat treatment diagram.</p>
Full article ">Figure 3
<p>Reconstructed microstructure.</p>
Full article ">Figure 4
<p>The microstructure and texture of planar-flow-casted ribbons. (<b>a</b>) EBSD inverse pole figure (IPF) map of microstructure, (<b>b</b>) φ<sub>2</sub> = 45° section of ODF. (The different colors in the figure represent different orientations).</p>
Full article ">Figure 5
<p>The microstructure and texture of planar-flow-casted ribbons after annealing at different temperatures. (<b>a</b>): initial state, (<b>b</b>) 850 °C, (<b>c</b>) 950 °C, (<b>d</b>) 1050 °C, (<b>e</b>) 1150 °C; in the map, (<b>a</b>–<b>e</b>): EBSD inverse pole figure (IPF) map of microstructure, (<b>a<sub>1</sub></b>–<b>e<sub>1</sub></b>): Grain distribution of the {001}, {110}, and {111} orientations (<b>a<sub>2</sub></b>–<b>e<sub>2</sub></b>): φ<sub>2</sub> = 45° section of ODF.</p>
Full article ">Figure 6
<p>The average grain size distribution and the proportion of typical texture components after annealing at different temperatures. (<b>a</b>): Average grain size diagram of typical texture components; (<b>b</b>) The proportion of {001}-, {110}- and {111}-oriented grains.</p>
Full article ">Figure 7
<p>The microstructure and texture of planar-flow-casted ribbons at different annealing temperatures after temper rolling with a 7% reduction rate. (<b>a</b>): 7% rolled state, (<b>b</b>) 850 °C, (<b>c</b>) 950 °C, (<b>d</b>) 1050 °C, (<b>e</b>) 1150 °C; in the map, (<b>a</b>–<b>e</b>): EBSD inverse pole figure (IPF) map of microstructure, (<b>a<sub>1</sub></b>–<b>e<sub>1</sub></b>): Grain distribution of the {001}, {110}, and {111} orientations, (<b>a<sub>2</sub></b>–<b>e<sub>2</sub></b>): φ<sub>2</sub> = 45° section of ODF.</p>
Full article ">Figure 8
<p>The grain size distribution and the proportion of typical texture components at different annealing temperatures after 7% reduction rolling of planar-flow-casted ribbons. (<b>a</b>) The largest grain size diagram of each typical texture component; (<b>b</b>) the proportion of {001}-, {110}-, and {111}-oriented grains.</p>
Full article ">Figure 9
<p>The microstructure and texture of planar-flow-casted ribbons at different annealing temperatures after temper rolling with 15% reduction rate. (<b>a</b>) 850 °C, (<b>b</b>) 950 °C, (<b>c</b>) 1050 °C, (<b>d</b>) 1150 °C; in the map, (<b>a</b>–<b>d</b>): EBSD inverse pole figure (IPF) map of microstructure, (<b>a<sub>1</sub></b>–<b>d<sub>1</sub></b>): Grain distribution of the {001}, {110} and {111} orientations, (<b>a<sub>2</sub></b>–<b>d<sub>2</sub></b>): φ<sub>2</sub> = 45° section of ODF.</p>
Full article ">Figure 10
<p>The grain size distribution and the proportion of typical texture components at different annealing temperatures after a 15% reduction rolling of the planar-flow-casted ribbon. (<b>a</b>) The largest grain size diagram of each typical texture component; (<b>b</b>) the proportion of {001}-, {110}-, and {111}-oriented grains.</p>
Full article ">Figure 11
<p>Mises stress of microstructure reconstructed by crystal plasticity finite element model (<b>a</b>): 7% reduction rate, the Mises stress distribution of grains 1–5 ({001}&lt;100&gt;, {001}&lt;110&gt;, {111}&lt;110&gt;,{111}&lt;112&gt;, {011}&lt;100&gt;) with different orientations in the reconstructed microstructure; (<b>b</b>): 15% reduction rate, the Mises stress distribution of grains 6–10 ({001}&lt;100&gt;, {001}&lt;110&gt;, {111}&lt;110&gt;,{111}&lt;112&gt;, {011}&lt;100&gt;) with different orientations in the reconstructed microstructure.</p>
Full article ">Figure 12
<p>Average Mises stress of grains at different rolling reduction rates. (The arrows in the figure represent the difference value between {001} and {111} grans).</p>
Full article ">
15 pages, 5387 KiB  
Article
Synergistic Effects of Ternary Microbial Self-Healing Agent Comprising Bacillus pasteurii, Saccharomyces cerevisiae, and Bacillus mucilaginosus on Self-Healing Performance in Mortar
by Zhaoyun Wu, Jiaxuan Li, Tianlei Wang, Lei Zhang, Ben Peng and Changsheng Yue
Materials 2024, 17(19), 4834; https://doi.org/10.3390/ma17194834 - 30 Sep 2024
Viewed by 615
Abstract
In order to prevent structural damage or high repair costs caused by concrete crack propagation, the use of microbial-induced CaCO3 precipitation to repair concrete cracks has been a hot topic in recent years. However, due to environmental constraints such as oxygen concentration, [...] Read more.
In order to prevent structural damage or high repair costs caused by concrete crack propagation, the use of microbial-induced CaCO3 precipitation to repair concrete cracks has been a hot topic in recent years. However, due to environmental constraints such as oxygen concentration, the width and depth of repaired cracks are seriously insufficient, which affects the further development of microbial self-healing agents. In this paper, a ternary microbial self-healing agent composed of different proportions of Bacillus pasteurii, Saccharomyces cerevisiae, and Bacillus mucilaginosus was designed, and its crack repair ability was evaluated. When the mixing ratio was 7:1:2, the cell concentration was the highest, the precipitation amount of CaCO3 was the highest, and the crystallinity of calcite crystal was the highest. Compared to the single microorganism, the mortar specimens with ternary microorganisms had the largest repair area (up to 100%) and the deepest repair depth (CaCO3 presents at 9–12 mm from the crack surface). This is because when the concrete breaks, all three microorganisms are activated by water, O2, and CO2. Saccharomyces cerevisiae and Bacillus mucilaginosus accelerated the growth of Bacillus pasteurii and more mineralized products; CaCO3 was rapidly formed and quickly filled on the crack surface. When CaCO3 seals the surface of the crack, the internal Saccharomyces cerevisiae and Bacillus mucilaginosus continue to play a role. Bacillus mucilaginosus can accelerate the dissolution of CO2 produced by the anaerobic fermentation of Saccharomyces cerevisiae and the hydrolysis of CO32−, thereby improving the repair of the crack depth direction. Full article
Show Figures

Figure 1

Figure 1
<p>The flow chart of ternary microbial self-healing agent on self-healing performance in mortar.</p>
Full article ">Figure 2
<p>Change in pH value of each microorganism suspension (<b>a</b>) and microorganism suspension with different mixing ratios (<b>b</b>); Changes in cell concentration of each microbial suspension (<b>c</b>) and microorganism suspension with different mixing ratios (<b>d</b>).</p>
Full article ">Figure 3
<p>Concentration of Ca<sup>2+</sup> in the supernatant of microorganism–substrate solution during mineralization of individual microorganism (<b>a</b>) and the ternary microorganisms (<b>b</b>) under the oxygen-rich condition; amount of CaCO<sub>3</sub> precipitate obtained after mineralization for 72 h (<b>c</b>).</p>
Full article ">Figure 4
<p>XRD patterns of mineralization products with different mixing ratios (<b>a</b>) and different microorganisms (<b>b</b>).</p>
Full article ">Figure 5
<p>SEM of mineralization products with different mixing ratios.</p>
Full article ">Figure 6
<p>Digital photos and binary images of mortar cracks before (<b>a</b>) and after (<b>b</b>) self-healing.</p>
Full article ">Figure 7
<p>SEM of powders in different crack depths.</p>
Full article ">Figure 8
<p>TG/DTG curves of powders with different crack depths.</p>
Full article ">
12 pages, 8469 KiB  
Article
The Creation of a Domain Structure Using Ultrashort Pulse NIR Laser Irradiation in the Bulk of MgO-Doped Lithium Tantalate
by Boris Lisjikh, Mikhail Kosobokov and Vladimir Shur
Photonics 2024, 11(10), 928; https://doi.org/10.3390/photonics11100928 - 30 Sep 2024
Viewed by 450
Abstract
The fabrication of stable, tailored domain patterns in ferroelectric crystals has wide applications in optical and electronic industries. All-optical ferroelectric poling by pulse laser irradiation has been developed recently. In this work, we studied the creation of the domain structures in MgO-doped lithium [...] Read more.
The fabrication of stable, tailored domain patterns in ferroelectric crystals has wide applications in optical and electronic industries. All-optical ferroelectric poling by pulse laser irradiation has been developed recently. In this work, we studied the creation of the domain structures in MgO-doped lithium tantalate by focused irradiation with a femtosecond near-infrared laser. Cherenkov-type second harmonic generation microscopy was used for domain imaging of the bulk. We have revealed the creation of enveloped domains around the induced microtracks under the action of the depolarization field. The domain growth is due to a pyroelectric field caused by a nonuniform temperature change. The domains in the bulk were revealed to have a three-ray star-shaped cross-section. It was shown that an increase in the field excess above the threshold leads to consequential changes in domain shape from a three-ray star to a triangular and a circular shape. The appearance of comb-like domains as a result of linear scanning was demonstrated. All effects were considered in terms of a kinetic approach, taking into account the domain wall motion by step generation and kink motion driven by excess of the local field over the threshold. The obtained knowledge is useful for the all-optical methods of domain engineering in ferroelectrics. Full article
(This article belongs to the Special Issue Ultrashort Laser Pulses)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Scheme of the experimental setup, (<b>b</b>) optical image of the microtracks that appeared after the local irradiation by femtosecond laser pulses with an energy of 6.0 µJ.</p>
Full article ">Figure 2
<p>SHGM side view of the domains formed by various numbers of pulses: (<b>a</b>) 1.4 µJ, (<b>b</b>) 4.0 µJ. Focusing depth was 300 µm.</p>
Full article ">Figure 3
<p>SHGM images of the domain XY cross-sections. (<b>a</b>) Domain XY cross-sections for various pulse numbers and energies. Focusing depth was 300 μm. (<b>b</b>) Domain XY cross-sections for different distances from the focusing point. Focusing depth was 500 µm, energy was 6.1 µJ, and pulse number was 512.</p>
Full article ">Figure 4
<p>SHGM images of the domains created at the various focusing depths: (<b>a</b>–<b>c</b>) 300 µm, (<b>d</b>,<b>e</b>) 500 µm. Pulse energy: (<b>a</b>–<b>c</b>) 5.4 µJ, (<b>d</b>,<b>e</b>) 6.1 µJ. Pulse number: 512. (<b>a</b>,<b>d</b>) Three-dimensional domain images, (<b>b</b>,<b>c</b>) and (<b>e</b>,<b>f</b>) two-dimensional images of the individual wings. The orange lines show the microtracks’ position.</p>
Full article ">Figure 5
<p>Dependencies of the wing length along the polar direction (<b>a</b>) on the energy and (<b>b</b>) the pulse number and the wing width along the Y+ direction (<b>c</b>) on the energy and (<b>d</b>) the pulse number.</p>
Full article ">Figure 6
<p>Results of the linear scanning along the X-axis. (<b>a</b>) Optical image of microtracks. (<b>b</b>–<b>d</b>) SHGM images of various cross-sections of comb-like domains: (<b>b</b>,<b>c</b>) XY at different depths: (<b>b</b>) at 400 µm, (<b>c</b>) at 320 µm. (<b>d</b>) XZ along scanning direction. Pulse energy was 2.7 µJ, focusing depth was 450 µm, and scanning rate was 2 mm/s.</p>
Full article ">Figure 7
<p>Scheme of the formation of the double-sided comb-like domain during linear scanning: (<b>a</b>) domain edge, (<b>b</b>) increase in CDW tilt, (<b>c</b>) appearance of the additional tooth, (<b>d</b>) elongated domain.</p>
Full article ">
20 pages, 3024 KiB  
Article
Secondary Metabolites from Australian Lichens Ramalina celastri and Stereocaulon ramulosum Affect Growth and Metabolism of Photobiont Asterochloris erici through Allelopathy
by Martin Bačkor, Dajana Kecsey, Blažena Drábová, Dana Urminská, Martina Šemeláková and Michal Goga
Molecules 2024, 29(19), 4620; https://doi.org/10.3390/molecules29194620 - 29 Sep 2024
Viewed by 338
Abstract
In the present work, the phytotoxic effects of secondary metabolites extracted from lichen Ramalina celastri (usnic acid) and lichen Stereocaulon ramulosum (a naturally occurring mixture of atranorin and perlatolic acid, approx. 3:1) on cultures of the aposymbiotically grown lichen photobiont Asterochloris erici were [...] Read more.
In the present work, the phytotoxic effects of secondary metabolites extracted from lichen Ramalina celastri (usnic acid) and lichen Stereocaulon ramulosum (a naturally occurring mixture of atranorin and perlatolic acid, approx. 3:1) on cultures of the aposymbiotically grown lichen photobiont Asterochloris erici were evaluated. Algae were cultivated on the surface of glass microfiber disks with applied crystals of lichen extracts for 14 days. The toxicity of each extract was tested at the two selected doses in quantities of 0.01 mg/disk and 0.1 mg/disk. Cytotoxicity of lichen extracts was assessed using selected physiological parameters, such as growth (biomass production) of photobiont cultures, content of soluble proteins, chlorophyll a fluorescence, chlorophyll a integrity, contents of chlorophylls and total carotenoids, hydrogen peroxide, superoxide anion, TBARS, ascorbic acid (AsA), reduced (GSH) and oxidized (GSSG) glutathione, and composition of selected organic acids of the Krebs cycle. The application of both tested metabolic extracts decreased the growth of photobiont cells in a dose-dependent manner; however, a mixture of atranorin and perlatolic acid was more effective when compared to usnic acid at the same dose tested. A higher degree of cytotoxicity of extracts from lichen S. ramulosum when compared to identical doses of extracts from lichen R. celastri was also confirmed by a more pronounced decrease in chlorophyll a fluorescence and chlorophyll a integrity, decreased content of chlorophylls and total carotenoids, increased production of hydrogen peroxide and superoxide anion, peroxidation of membrane lipids (assessed as TBARS), and a strong decrease in non-enzymatic antioxidants such as AsA, GSH, and GSSG. The cytotoxicity of lichen compounds was confirmed by a strong alteration in the composition of selected organic acids included in the Krebs cycle. The increased ratio between pyruvic acid and citric acid was a very sensitive parameter of phytotoxicity of lichen secondary metabolites to the algal partner of symbiosis. Secondary metabolites of lichens are potent allelochemicals and play significant roles in maintaining the balance between mycobionts and photobionts, forming lichen thallus. Full article
Show Figures

Figure 1

Figure 1
<p>Biomass production ((<b>A</b>); mg dw/disk), chlorophyll <span class="html-italic">a</span> fluorescence ((<b>B</b>); F<sub>V</sub>/F<sub>M</sub>), content of soluble proteins ((<b>C</b>); mg/g dw), and chlorophyll <span class="html-italic">a</span> integrity ((<b>D</b>); A 435 nm/415 nm) of 2 week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 1 Cont.
<p>Biomass production ((<b>A</b>); mg dw/disk), chlorophyll <span class="html-italic">a</span> fluorescence ((<b>B</b>); F<sub>V</sub>/F<sub>M</sub>), content of soluble proteins ((<b>C</b>); mg/g dw), and chlorophyll <span class="html-italic">a</span> integrity ((<b>D</b>); A 435 nm/415 nm) of 2 week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 2
<p>Chlorophyll <span class="html-italic">a</span> content ((<b>A</b>); mg/g dw), chlorophyll <span class="html-italic">b</span> content ((<b>B</b>); mg/g dw), chlorophyll <span class="html-italic">a</span>/<span class="html-italic">b</span> ((<b>C</b>), and chlorophyll <span class="html-italic">a</span>+<span class="html-italic">b</span> content ((<b>D</b>); mg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 2 Cont.
<p>Chlorophyll <span class="html-italic">a</span> content ((<b>A</b>); mg/g dw), chlorophyll <span class="html-italic">b</span> content ((<b>B</b>); mg/g dw), chlorophyll <span class="html-italic">a</span>/<span class="html-italic">b</span> ((<b>C</b>), and chlorophyll <span class="html-italic">a</span>+<span class="html-italic">b</span> content ((<b>D</b>); mg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 3
<p>Content of total carotenoids ((<b>A</b>); mg/g dw) and total carotenoids/total chlorophyll (<b>B</b>) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 4
<p>Content of hydrogen peroxide ((<b>A</b>); µmol/g dw), superoxide anion ((<b>B</b>); µg/g dw), TBARS ((<b>C</b>); nmol/g dw), and ascorbic acid (AsA, (<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 4 Cont.
<p>Content of hydrogen peroxide ((<b>A</b>); µmol/g dw), superoxide anion ((<b>B</b>); µg/g dw), TBARS ((<b>C</b>); nmol/g dw), and ascorbic acid (AsA, (<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 5
<p>Content of reduced glutathione (GSH, (<b>A</b>); µg/g dw), oxidized glutathione (GSSG, (<b>B</b>); µg/g dw), GSH/GSSG (<b>C</b>), and citric acid ((<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 5 Cont.
<p>Content of reduced glutathione (GSH, (<b>A</b>); µg/g dw), oxidized glutathione (GSSG, (<b>B</b>); µg/g dw), GSH/GSSG (<b>C</b>), and citric acid ((<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 6
<p>Content of fumaric acid ((<b>A</b>); µg/g dw), glutamic acid ((<b>B</b>); µg/g dw), ketoglutaric acid ((<b>C</b>); µg/g dw), and lactic acid ((<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 6 Cont.
<p>Content of fumaric acid ((<b>A</b>); µg/g dw), glutamic acid ((<b>B</b>); µg/g dw), ketoglutaric acid ((<b>C</b>); µg/g dw), and lactic acid ((<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 7
<p>Content of malic acid ((<b>A</b>); µg/g dw), pyruvic acid ((<b>B</b>); µg/g dw), quinic acid ((<b>C</b>); µg/g dw), and succinic acid ((<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 7 Cont.
<p>Content of malic acid ((<b>A</b>); µg/g dw), pyruvic acid ((<b>B</b>); µg/g dw), quinic acid ((<b>C</b>); µg/g dw), and succinic acid ((<b>D</b>); µg/g dw) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 8
<p>Content of tartaric acid ((<b>A</b>); µg/g dw) and pyruvic acid/citric acid (<b>B</b>) of 2-week-old photobiont <span class="html-italic">Asterochloris erici</span> cultures cultivated on disks with the addition of secondary metabolites extracts from lichens <span class="html-italic">Ramalina celastri</span> and <span class="html-italic">Stereocaulon ramulosum</span> (0.01 and 0.1 mg/disk). Values in vertical columns followed by the same letter(s) are not significantly different according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05), <span class="html-italic">n</span> = 3.</p>
Full article ">
Back to TopTop