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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
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
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

Search Results (7,284)

Search Parameters:
Keywords = proton

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 1609 KiB  
Article
Rapid Detection of Aluminium and Iron Impurities in Lithium Carbonate Using Water-Soluble Fluorescent Probes
by Hong-Mei Wu, Huai-Gang Cheng, Zi-Wen Zhu and Li Cui
Molecules 2025, 30(1), 135; https://doi.org/10.3390/molecules30010135 - 31 Dec 2024
Viewed by 158
Abstract
The real-time measurement of the content of impurities such as iron and aluminium ions is one of the keys to quality evaluation in the production process of high-purity lithium carbonate; however, impurity detection has been a time-consuming process for many years, which limits [...] Read more.
The real-time measurement of the content of impurities such as iron and aluminium ions is one of the keys to quality evaluation in the production process of high-purity lithium carbonate; however, impurity detection has been a time-consuming process for many years, which limits the optimisation of the production of high-purity lithium carbonate. In this context, this work explores the possibility of using water-soluble fluorescent probes for the rapid detection of impurity ions. Salicylaldehyde was modified with the hydrophilic group dl-alanine to synthesise a water-soluble Al3+ fluorescent probe (Probe A). Moreover, a water-soluble Fe3+ fluorescent probe (Probe B) was synthesised from coumarin-3-carboxylic acid and 3-hydroxyaminomethane. Probe A and Probe B exhibited good stability in the pH range of 4–9 in aqueous solutions, high sensitivity, as well as high selectivity for Al3+ and Fe3+; the detection limits for Al3+ and Fe3+ were 1.180 and 1.683 μmol/L, whereas the response times for Al3+ and Fe3+ were as low as 10 and 30 s, respectively. Electrostatic potential (ESP) analysis and density functional theory calculations identified the binding sites and fluorescence recognition mechanism; theoretical calculations showed that the enhanced fluorescence emission of Probe A when detecting Al3+ was due to the excited intramolecular proton transfer (ESIPT) effect, whereas the fluorescence quenching of Probe B when detecting Fe3+ was due to the electrons turning off fluorescence when binding through the photoelectron transfer (PET) mechanism. Full article
19 pages, 5238 KiB  
Article
In Situ Raman Spectroscopy for Early Corrosion Detection in Coated AA2024-T3
by Adrienne K. Delluva, Ronald L. Cook, Matt Peppel, Sami Diaz, Rhia M. Martin, Vinh T. Nguyen, Jeannine E. Elliott and Joshua R. Biller
Sensors 2025, 25(1), 179; https://doi.org/10.3390/s25010179 - 31 Dec 2024
Viewed by 207
Abstract
Here we describe the synthesis and evaluation of a molecular corrosion sensor that can be applied in situ in aerospace coatings, then used to detect corrosion after the coating has been applied. A pH-sensitive molecule, 4-mercaptopyridin (4-MP), is attached to a gold nanoparticle [...] Read more.
Here we describe the synthesis and evaluation of a molecular corrosion sensor that can be applied in situ in aerospace coatings, then used to detect corrosion after the coating has been applied. A pH-sensitive molecule, 4-mercaptopyridin (4-MP), is attached to a gold nanoparticle to allow surface-enhanced Raman-scattering (SERS) for signal amplification. These SERS nanoparticles, when combined with an appropriate micron-sized carrier system, are incorporated directly into an MIL-SPEC coating and used to monitor the process onset and progression of corrosion using pH changes occurring at the metal–coating interface. The sensor can track corrosion spatially as it proceeds underneath the coating, due to the mobility of the proton front generated during corrosion and the homogeneous distribution of the sensor in the coating layer. To our knowledge, this report is the first time a 4-MP functionalized gold nanoparticle has been used, along with SERS spectroscopy, to monitor corrosion in an applied commercial coating in a fast, non-contact way. Full article
(This article belongs to the Special Issue Nanotechnology Applications in Sensors Development)
Show Figures

Figure 1

Figure 1
<p>Illustration of corrosion occurring near the metal surface on AA-2024. An acidic environment is a hallmark of severe corrosion.</p>
Full article ">Figure 2
<p>SEM of sensor powder in backscattering mode: bright spots are gold.</p>
Full article ">Figure 3
<p>Raman spectrum of powder corrosion sensor loaded into a primer at 150 ppm, compared to the Raman spectra of the primer alone and the powder sensor alone. (Inset) The Raman spectra of the primary peaks of interest for 4-MP as a function of changing pH, from 1.2 to 12.6.</p>
Full article ">Figure 4
<p>Percentage change in PRR for (<b>A</b>) the sensor dispersed in aqueous solution as a function of pH and (<b>B</b>) loaded into the primer coating, with pH solution applied to the coatings. Both graphs share the same legend.</p>
Full article ">Figure 5
<p>Corrosion exposure results for a scribed AA-2024 panel coated in MIL-DTL-53030 primer, loaded with 150 ppm of corrosion sensor. (<b>A</b>) Photos of the scribe center as a function of time. (<b>B</b>) Percentage change in the PRR as a function of time.</p>
Full article ">Figure 6
<p>MIL-DTL-53030 primer loaded with corrosion sensor was applied to panels with no pretreatment, or those that had been treated with an alodine conversion coating. (<b>A</b>) Raman signal as a function of time in the salt fog. (<b>B</b>) A representative “bare” (no pretreatment) panel after 1500 h of exposure to ASTM B117. (<b>C</b>) Photographs of the alodine panel, stripped after 2000 h in ASTM B117. (Inset) Deep corrosion damage is present, which was indicated by the corrosion sensor prior to stripping.</p>
Full article ">Figure 7
<p>(<b>A</b>) Photos of non-ideal test panel surfaces covered in (left to right) hydraulic fluid, pristine, covered in dirt, and curved. (<b>B</b>) Raman spectra of the different panel states.</p>
Full article ">Figure 8
<p>Comparison of accelerated corrosion on a 3″ × 3″ AA-2024 panel coated in MIL-DTL-53030, as assessed by the Raman corrosion sensor or electrical impedance spectroscopy (EIS). The decrease in the charge transfer resistance tracks well with the decrease in the Raman sensor, brought on by a decrease in pH due to active and severe corrosion.</p>
Full article ">Figure 9
<p>(<b>A</b>) 12″ × 12″ panel used for spatial resolution testing. The scribe is in the bottom right and circles are marked with pen on the surface of the panel to measure the same locations at each time point in ASTM B117. (<b>B</b>) Zoomed-in scribe after 2500 h in ASTM-B117. (<b>C</b>) Zoomed-in scribe after coating was stripped at 6000 h in ASTM B117. Note the holes where it has corroded clean through the panel. (<b>D</b>) PRR values at the spots labeled A in the salt fog.</p>
Full article ">
15 pages, 4833 KiB  
Article
Fluorescent Polymers via Coordination of bis-Terpyridine Ligands with Transition Metals and Their pH Response Properties
by Tao Zhang, Fengxue Liu, Yongxin Liu, Kaixiu Li, Zhengguang Li, Yaqin Li, Fan Fu, Mingliang Liu, Yiming Li, Die Liu and Pingshan Wang
Polymers 2025, 17(1), 87; https://doi.org/10.3390/polym17010087 (registering DOI) - 31 Dec 2024
Viewed by 240
Abstract
Stimulus-responsive luminescent materials are pivotal in the field of sensing. Fluorescent transition metal complexes with a charge transfer excited state, especially terpyridine-coordinated polymers, are of particular interest due to their tunable emission. In this paper, a novel bis-terpyridine ligand was synthesized and assembled [...] Read more.
Stimulus-responsive luminescent materials are pivotal in the field of sensing. Fluorescent transition metal complexes with a charge transfer excited state, especially terpyridine-coordinated polymers, are of particular interest due to their tunable emission. In this paper, a novel bis-terpyridine ligand was synthesized and assembled into a coordination polymer, which showed intense visible light absorption and fluorescence emission in the solid state that could be regulated by an acidic or basic pH. After being protonated by acid, the fluorescence of the polymer P2 was quenched. The emission of the polymer split from 635 nm to two peaks of 674 and 440 nm, and then stabilized at 728 nm for 7 days, which showed a significant red-shift and good protonation stability. The fluorescence emission wavelength of the protonated polymers recovered after alkalization, and the fluorescence intensity of the polymer was greatly improved after alkalization, showing interesting acid–base-response luminescence characteristics. The sensitive response of the synthesized coordination polymers to acids and bases will contribute to expanding the application of linear coordination polymers in sensing and other fields. Full article
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H NMR spectra of (<b>a</b>) <b>L2</b> (CDCl<sub>3</sub>), (<b>b</b>) <b>P2</b> (DMSO-d<sub>6</sub>), (<b>c</b>) <b>L1</b> (CDCl<sub>3</sub>) and (<b>d</b>) <b>P1</b> (DMSO-d<sub>6</sub>).</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM and mapping (EDS) images of <b>P2</b>; (<b>b</b>) SEM image of <b>P1</b>; (<b>c</b>)TEM image of <b>P1</b>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Abs of mixture of <b>L3</b> (9.316 μM) and [Zn<sup>2+</sup>] (0.0–1.0 eq) stirred for 4 h. (<b>b</b>) Plot of (A<sub>0</sub>/(A − A<sub>0</sub>) as a function of 1/[Zn<sup>2+</sup>]. The apparent association constant Ka corresponding to the interaction between <b>L3</b> (9.316 μM) and [Zn<sup>2+</sup>] (0.0–1.0 eq) was determined using the Benesi–Hilderbrand equation A<sub>0</sub>/(A − A<sub>0</sub>) = (A<sub>0</sub>/(Amax − A<sub>0</sub>))((1/Ka)[Zn<sup>2+</sup>]<sup>−1</sup> + 1).</p>
Full article ">Figure 4
<p>(<b>a</b>) UV-Vis and (<b>b</b>) Fluorescence spectrum of <b>L1</b>, <b>L2</b>, <b>P1</b>, <b>P2</b> (1 × 10<sup>−5</sup> mol·L<sup>−1</sup> in CH<sub>3</sub>CN), LOD and LOQ of <b>L1</b>, <b>L2</b>, <b>P1</b>, <b>P2</b> can be seen in <a href="#app1-polymers-17-00087" class="html-app">Figure S29</a>; (<b>c</b>) Fluorescence spectrum of solid <b>L1</b>, <b>L2</b>, <b>P1</b>, <b>P2</b>; (<b>d</b>) Acid–base response of <b>P2</b>, the Fluorescence spectrum of solid <b>P2</b>, <b>P2</b> fumigated with concentrated hydrochloric acid (30 s), fumigated <b>P2</b> after 7 days and fumigated <b>P2</b> fumed with ammonium hydroxide. Inset: Photos of ligands and polymers under sunlight (<b>top</b>) and ultraviolet light (<b>bottom</b>).</p>
Full article ">Figure 5
<p>(<b>a</b>) Infrared spectrum of solid <b>P2</b>, <b>P2</b> fumigated with concentrated hydrochloric acid (30 s), fumigated <b>P2</b> after 7 days and fumigated <b>P2</b> fumed with ammonium hydroxide; (<b>b</b>) Thermogravimetric and differential thermogravimetric analysis diagram of <b>P1</b> and <b>P2</b> (the red line represents <b>P2</b>, the black line represents <b>P1</b>, the solid line represents thermogravimetric analysis, and the dashed line represents differential thermogravimetric analysis).</p>
Full article ">Scheme 1
<p>Synthesis route of the terpyridine ligands <b>L1</b>/<b>L2</b> and coordination polymers <b>P1</b>/<b>P2</b> (I: a. CH<sub>3</sub>CH<sub>2</sub>OH, NaOH; b. NH<sub>3</sub>·H<sub>2</sub>O, refluxed; II: KOH, THF, CH<sub>3</sub>OH, Ar, refluxed; III: Zn(NO<sub>3</sub>)<sub>2</sub>·6H<sub>2</sub>O, CHCl<sub>3</sub>/CH<sub>3</sub>OH).</p>
Full article ">
27 pages, 4802 KiB  
Review
Impact of Coolant Operation on Performance and Heterogeneities in Large Proton Exchange Membrane Fuel Cells: A Review
by Marine Cornet, Erwan Tardy, Jean-Philippe Poirot-Crouvezier and Yann Bultel
Energies 2025, 18(1), 111; https://doi.org/10.3390/en18010111 - 30 Dec 2024
Viewed by 290
Abstract
PEMFCs’ operation entails the presence of heterogeneities in the generation of current, heat and water along the active surface area. Indeed, PEMFCs are open systems, and as such, operating heterogeneities are inherent to their operation. A review of the literature reveals numerous attempts [...] Read more.
PEMFCs’ operation entails the presence of heterogeneities in the generation of current, heat and water along the active surface area. Indeed, PEMFCs are open systems, and as such, operating heterogeneities are inherent to their operation. A review of the literature reveals numerous attempts to achieve uniform current density distribution. These attempts are primarily focused on bipolar plate design and operating conditions, with the underlying assumption that uniform current density correlates with enhanced performance. Most studies focus on the influence of gas flow-field design and inlet hydrogen and air flow conditioning, and less attention has been paid to the coolant operating condition. However, uncontrolled temperature distribution over a large cell active surface area can lead to performance loss and localized degradations. On this latter point, we notice that studies to date have been confined to a narrow range of operating conditions. It appears that complementary durability studies are needed in order to obtain in-depth analyses of the coupled influence of temperature distribution and gas humidification in large PEMFCs. Full article
(This article belongs to the Section D2: Electrochem: Batteries, Fuel Cells, Capacitors)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Five different cooling flow-field designs: straight parallel channels (case A), straight parallel channels filled with metal foam (Case B), multi-channel serpentine (Case C), novel serpentine channels (Case D) and integrated metal foam (Case E); (<b>b</b>) average temperature (membrane), maximum and minimum temperature of the cathode catalyst layer and (<b>c</b>) comparison of the polarization curve and power density curves for the various fuel cells [<a href="#B106-energies-18-00111" class="html-bibr">106</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 2
<p>(<b>a</b>) Computational domains of four configurations of coolant channel of a large-size proton exchange membrane fuel cell including straight + straight (S + S), straight + wavy (S + W), same waveform (SW) and different waveform (DW) channels and (<b>b</b>) pressure drops and temperature differences of the four configurations [<a href="#B116-energies-18-00111" class="html-bibr">116</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 3
<p>(<b>a</b>) Photos of the air-cooled PEMFC of the experimental test system and BPs and (<b>b</b>) schematic diagram of the computational domain and meshing details in numerical simulations [<a href="#B119-energies-18-00111" class="html-bibr">119</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 4
<p>Fuel cell geometry: (<b>a</b>) schematic diagram of the PEMFC and (<b>b</b>) tree-shaped fractal cooling flow field [<a href="#B121-energies-18-00111" class="html-bibr">121</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 5
<p>(<b>a</b>) PEMFC model; (<b>b</b>) smooth cooling channel; (<b>c</b>) circular dimpled cooling channel and (<b>d</b>) elliptical dimple cooling channel [<a href="#B123-energies-18-00111" class="html-bibr">123</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 6
<p>(<b>a</b>) Location of the spacer between two bipolar plates; (<b>b</b>) the design steps of a spacer from the first design to the optimized design; and (<b>c</b>) IUT obtained according to the different spacer designs [<a href="#B127-energies-18-00111" class="html-bibr">127</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 7
<p>(<b>a</b>) Design scheme of flow field. Effect of inlet flow on heat transfer characteristics; (<b>b</b>) maximum temperature and (<b>c</b>) temperature difference [<a href="#B114-energies-18-00111" class="html-bibr">114</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 8
<p>(<b>a</b>) Three coolant flow-field designs for large-scale PEMFCs; (<b>b</b>) average cathode liquid saturation and index of uniform temperature profile for the PEMFC under different cooling water inlet volume flow rates; and (<b>c</b>) average current density profile and pumping power profile under different coolant volume flow rates [<a href="#B115-energies-18-00111" class="html-bibr">115</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 9
<p>(<b>a</b>) Temperature distribution at the reference cut line in the membrane, cathode microporous layer, cathode gas diffusion layer and bipolar plate for a current density of 0.42 A cm<sup>−2</sup> and (<b>b</b>) current density and cooling flow velocity at the cut line [<a href="#B129-energies-18-00111" class="html-bibr">129</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 10
<p>(<b>a</b>) Measured and (<b>b</b>) simulated total liquid water thickness for a stack of 5 cells for a current density of 0.25 A cm<sup>−2</sup> and for an inlet cooling water temperature of 64 °C [<a href="#B69-energies-18-00111" class="html-bibr">69</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 11
<p>Cooling schemes of each model: (<b>a</b>) traditional unidirectional cooling flow (Model A); (<b>b</b>) reverse flow cooling of the interlayer flow channel (Model B); (<b>c</b>) reverse flow cooling of the adjacent channel (Model C;) and (<b>d</b>) bidirectional circulation cooling (Model D) [<a href="#B130-energies-18-00111" class="html-bibr">130</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 12
<p>(<b>a</b>) I–V curves and (<b>b</b>) averaged water content obtained for the 108 simulations with the calibrated spatially averaged P3D model featuring a variation in coolant outlet temperature (60 °C, 70 °C and 80 °C), in coolant temperature gradient between inlet and outlet ΔT (2 °C, 6 °C, 10 °C and 20 °C) and in anode and cathode RH (30%, 50% and 60%). The best (red) and the worst (blue) performances at high current density have been highlighted.</p>
Full article ">Figure 13
<p>Example of single cell hardware using real bipolar plates and current scan sensor: (<b>a</b>) sectional view of the cell and (<b>b</b>) air side of the bipolar plate. The airflow direction in the straight channels is indicated by white arrows [<a href="#B138-energies-18-00111" class="html-bibr">138</a>]. Reproduced with permission from Elsevier.</p>
Full article ">Figure 14
<p>Evolution of the fluorine release rate [<a href="#B156-energies-18-00111" class="html-bibr">156</a>].</p>
Full article ">Figure 15
<p>Overview of the observed degradations. [<a href="#B67-energies-18-00111" class="html-bibr">67</a>]. Reproduced with permission from Elsevier.</p>
Full article ">
19 pages, 564 KiB  
Article
Comparative Computational Study of Frequency Shifts and Infrared Intensity Changes in Model Binary Complexes with Red- and Blue-Shifting Hydrogen Bonds
by Sean A. C. McDowell
Molecules 2025, 30(1), 106; https://doi.org/10.3390/molecules30010106 - 30 Dec 2024
Viewed by 180
Abstract
A computational study of X-H···Y binary hydrogen-bonded complexes was undertaken to examine the red- and blue-shifting behavior of three model X-H proton donors interacting with a series of Lewis bases: Y = NH3, NCLi, NCH, NCF, C2H2, [...] Read more.
A computational study of X-H···Y binary hydrogen-bonded complexes was undertaken to examine the red- and blue-shifting behavior of three model X-H proton donors interacting with a series of Lewis bases: Y = NH3, NCLi, NCH, NCF, C2H2, BF, CO, N2 and Ne. Two of these proton donors, FArH and F3CH, have blue-shifting tendencies, while the third, FH, has red-shifting tendencies. A perturbation theory model for frequency shifts that was derived many years ago was employed to partition the predicted frequency shift into the sum of two components, one dependent on the second derivative of the interaction energy with respect to X-H displacement and the other dependent on the X-H bond length change in the binary complex. The predicted shifts were found to be in good agreement with standard ab initio computations, but they were obtained at much lower computational cost. The change in the infrared intensity of the X-H stretching frequency, expressed as a ratio of complex to monomer intensities, was also investigated, along with its relation to the X-H permanent dipole moment derivative and total induced dipole moment derivative with respect to X-H displacement, and used to rationalize the observed infrared intensity changes in the red- and blue-shifted X-H···Y complexes. Full article
Show Figures

Figure 1

Figure 1
<p>Optimized geometries of some representative X-H···Y dimers: (<b>a</b>) FH···C<sub>2</sub>H<sub>2</sub>, (<b>b</b>) FArH···N<sub>2</sub>, (<b>c</b>) F<sub>3</sub>CH···NH<sub>3</sub>, (<b>d</b>) F<sub>3</sub>CH···C<sub>2</sub>H<sub>2</sub>.</p>
Full article ">
12 pages, 1587 KiB  
Article
Investigation of Pre-Pulse Effects on Ultrashort-Pulse Laser Interaction with Structured Targets
by Artem Kim, Indranuj Dey, Alexander Bespaly, Pavel Komm, Assaf Shaham, Jenya Papeer, Mordechai Botton and Arie Zigler
Appl. Sci. 2025, 15(1), 237; https://doi.org/10.3390/app15010237 - 30 Dec 2024
Viewed by 299
Abstract
The role of pre-plasma in the efficient generation of protons by intense laser-matter interaction from structured targets is investigated. Optimal energy coupling between laser and plasma is found by varying the fluence and arrival time of an independently controllable ultrashort pre-pulse with respect [...] Read more.
The role of pre-plasma in the efficient generation of protons by intense laser-matter interaction from structured targets is investigated. Optimal energy coupling between laser and plasma is found by varying the fluence and arrival time of an independently controllable ultrashort pre-pulse with respect to the main interaction pulse. The coupling is evaluated based on the energy of the accelerated protons. The accelerated proton energy is maximized at optimal pre-pulse delay and fluence conditions. Plasma emission spectrum and Particle-in-Cell simulations provide a possible explanation of the obtained experiment results. Full article
(This article belongs to the Special Issue Ultrafast and Nonlinear Laser Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Block diagram schematic of the experimental setup. MP: main pulse, APP: artificial pre-pulse, SA: saturable absorber, PBS: pellicle beam splitter, IC: interaction chamber. (<b>b</b>) The experimental setup inside the vacuum interaction chamber. TN: target normal, CRN: normal to the CR39 proton detector. (<b>c</b>) SEM image of the printed structured target. Red-ellipse: MP projection, yellow-dashed-ellipse: APP projection, dark-red arrow: laser polarization axis. (<b>d</b>) Enlarged part of planned target design with focus on spike. (<b>e</b>) The laser focusing geometry on the spike target showing the typical feature size on which the laser is focused, as well as the error in the alignment due to the depth of field of the imaging objective with respect to its focal plane. The red dashed lines demonstrate the FWHM of APP and MP.</p>
Full article ">Figure 2
<p>Plasma emission optical spectrum (intensity vs wavelength) plots at artificial pre-pulse (APP) time delays (<math display="inline"><semantics> <mi mathvariant="sans-serif">Δ</mi> </semantics></math>t) of, (<b>a</b>) −2 ns, (<b>b</b>) −5 ns, (<b>c</b>) −14 ns, with fluence ranges: (I) no-APP (black dashed line), (II) 0.1–0.2 (orange dash-dot line), (III) 0.4–0.6 (violet solid line), and (IV) 2–5 (teal dash-dot-dot line) J/cm<sup>2</sup> in each plot (<b>a</b>–<b>c</b>). (<b>d</b>) Normalized intensity versus wavelength plot for the optimum APP condition of <math display="inline"><semantics> <mi mathvariant="sans-serif">Δ</mi> </semantics></math>t ≈−5 ns in the fluence range ≈0.4–0.6 J/cm<sup>2</sup>.</p>
Full article ">Figure 3
<p>(<b>a</b>) A typical aluminum net filter on top of the CR39 plate used in the experiments: (I) is the base filter (thickness, 28 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m), (II) and (III) are the vertical and horizontal strips of 14 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and 22 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m thickness, respectively. (<b>b</b>) CR39 with signals from 2.5 ≤ PE &lt; 2.8 MeV protons. The adjacent image is obtained by scanning the area of interest in a SEM. (<b>c</b>) Variation in proton energy with the fluence of the artificial pre-pulse (APP) at selected time delays (<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>t</mi> </mrow> </semantics></math>) of: (i) −2 ns (red square with dashed line), (ii) −5 ns (blue circle with solid line), (iii) −8 ns (green diamond with dash-dot line). (<b>d</b>) Variation in average and maximum proton energy with the delay of the APP in the fluence range of 0.4–0.6 J/cm<sup>2</sup>. The data points are connected by Bezier curves to help discern the trend.</p>
Full article ">Figure 4
<p>(<b>a</b>) Structured target and pre-plasma geometry shown on a 2D electron number density map normalized to critical density <math display="inline"><semantics> <mrow> <mi>l</mi> <mi>o</mi> <msub> <mi>g</mi> <mn>10</mn> </msub> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <msub> <mi>n</mi> <mi>e</mi> </msub> <msub> <mi>n</mi> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> </mrow> </msub> </mfrac> </mstyle> </mrow> </semantics></math>. The red arrow indicates direction of the laser. (<b>b</b>) Electron energy spectrum of a flat target, structured target without pre-plasma, and structured target pre-plasma. (<b>c</b>) Time evolution of average accelerating field around the tips of the spikes with various targets. Time = 0 fs corresponds to the interaction of the peak of the pulse with target. (<b>d</b>) Proton energy cutoff in simulations with various pre-plasma expansion.</p>
Full article ">
12 pages, 2493 KiB  
Article
Tungsten Carbide/Tungsten Oxide Catalysts for Efficient Electrocatalytic Hydrogen Evolution
by Jian Ouyang, Yu Sun, Yiqiong Zhang, Juzhe Liu, Xin Bo and Zenglin Wang
Molecules 2025, 30(1), 84; https://doi.org/10.3390/molecules30010084 (registering DOI) - 29 Dec 2024
Viewed by 353
Abstract
Catalyzing hydrogen evolution reaction (HER) is a key process in high-efficiency proton exchange membrane water electrolysis (PEMWE) devices. To replace the use of Pt-based HER catalyst, tungsten carbide (W2C) is one of the most promising non-noble-metal-based catalysts with low cost, replicable [...] Read more.
Catalyzing hydrogen evolution reaction (HER) is a key process in high-efficiency proton exchange membrane water electrolysis (PEMWE) devices. To replace the use of Pt-based HER catalyst, tungsten carbide (W2C) is one of the most promising non-noble-metal-based catalysts with low cost, replicable catalytic performance, and durability. However, the preparation access to scalable production of W2C catalysts is inevitable. Herein, we introduced a facile protocol to achieve the tungsten carbide species by plasma treatment under a CH4 atmosphere from the WO3 precursor. Moreover, the heterogeneous structure of the tungsten carbide/tungsten oxide nanosheets further enhances the catalytic activity for HER with the enlarged specific surface area and the synergism on the interfaces. The prepared tungsten carbide/tungsten oxide heterostructure nanosheets (WO3-x-850-P) exhibit exceptional HER catalytic activity and stable longevity in acid electrolytes. This work provides a facile and effective method to construct high-performance and non-precious-metal-based electrocatalysts for industrial-scale water electrolysis. Full article
Show Figures

Figure 1

Figure 1
<p>Preparation diagram of tungsten carbide/tungsten oxide heterostructure nanosheets.</p>
Full article ">Figure 2
<p>SEM images of WO<sub>3</sub> (<b>a</b>), WO<sub>3-x</sub>-850-P (<b>b</b>), Raman and (<b>c</b>), XRD patterns of WO<sub>3</sub> (<b>d</b>), WO<sub>3-x</sub>-850 (<b>e</b>), WO<sub>3-x</sub>-850-P (<b>f</b>), respectively.</p>
Full article ">Figure 3
<p>XPS spectra of (<b>a</b>,<b>d</b>,<b>g</b>) W 4f, (<b>b</b>,<b>e</b>,<b>h</b>) O 1s, and (<b>c</b>,<b>f</b>,<b>i</b>) C 1s for WO<sub>3-x</sub>-850-P, WO<sub>3-x</sub>-850, and WO<sub>3</sub>. Cyan, W<sup>6+</sup>; Orange, satellite peaks of W<sup>6+</sup>; Violet, W<sup>4+</sup>; Green, W<sup>2+</sup>.</p>
Full article ">Figure 4
<p>HER performance of WO<sub>3-x</sub>-850-P, WO<sub>3-x</sub>-850, and WO<sub>3</sub> catalysts in 0.5 M H<sub>2</sub>SO<sub>4</sub>. (<b>a</b>) Polarization curves; (<b>b</b>) the corresponding Tafel slopes; (<b>c</b>) durability measurement of the WO<sub>3-x</sub>-850-P catalyst at the applied potential of −0.17 V vs. RHE; (<b>d</b>) Nyquist of WO<sub>3</sub>, WO<sub>3-x</sub>-850 and WO<sub>3-x</sub>-850-P.</p>
Full article ">
13 pages, 1408 KiB  
Article
Potential Chemopreventive Role of Proton Pump Inhibitors in Head and Neck Cancer: Insights from a Nested Case–Control Analysis of a National Health Screening Cohort
by Joong Seob Lee, Soomin Jo, Ho Suk Kang, Mi Jung Kwon, Jee Hye Wee, Jeong Wook Kang, Hyo Geun Choi and Heejin Kim
J. Pers. Med. 2025, 15(1), 8; https://doi.org/10.3390/jpm15010008 (registering DOI) - 28 Dec 2024
Viewed by 177
Abstract
Background/Objectives: This study investigated the potential chemopreventive role of proton pump inhibitor (PPI) use in relation to the occurrence of head and neck cancer (HNC) within a national cohort amid concerns of PPI overprescription. Methods: From a cohort of 1,137,861 individuals [...] Read more.
Background/Objectives: This study investigated the potential chemopreventive role of proton pump inhibitor (PPI) use in relation to the occurrence of head and neck cancer (HNC) within a national cohort amid concerns of PPI overprescription. Methods: From a cohort of 1,137,861 individuals and 219,673,817 medical claim records collected between 2005 and 2019, 1677 HNC patients were identified and matched 1:4 with 6708 controls after adjusting for covariates. Odds ratios (ORs) for PPI use and its duration in relation to HNC and its subsites were estimated using propensity score overlap-weighted multivariable logistic regression. Additional subgroup analyses were performed based on age, sex, income level, and geographic region. Results: In the crude model, both current (OR 7.85 [95% CI 6.52–9.44]) and past PPI (OR 1.44 [95% CI 1.23–1.70]) use were associated with increased odds for HNC. However, after overlap weighting, this association reversed for both current (aOR 0.14 [95% CI 0.11–0.17]) and past PPI (aOR 0.69 [95% CI 0.60–0.79]). Subsite analysis showed reduced odds for hypopharyngeal (aOR 0.33, [95% CI 0.25–0.43]) and laryngeal cancer (aOR 0.19 [95% CI 0.16–0.22]) in current PPI users and similar results for past users. Conclusions: This study suggests a potential chemopreventive effect of PPIs, particularly in hypopharyngeal and laryngeal cancers. Additional studies are required to investigate the mechanisms underlying the association of the development of HNC with PPI use. Full article
Show Figures

Figure 1

Figure 1
<p>A flowchart depicting the participant selection process in this study. Out of 1,136,184 participants, 1677 individuals with head and neck cancer were matched to 6708 control participants based on age, sex, income, and geographic region.</p>
Full article ">Figure 2
<p>Subgroup analyses of proton pump inhibitor (PPI) use (nonuser [ref] vs. user) for head and neck cancer (HNC) based on age, sex, income, and region of residence presented in a forest plot (<b>a</b>). (<b>b</b>) Subgroup analyses of PPI use for HNC according to comorbidities (Charlson comorbidity index [CCI]), history of gastroesophageal reflux disease (GERD), and H<sub>2</sub> blocker use), also visualized in a forest plot.</p>
Full article ">Figure 3
<p>Subgroup analyses of the duration of proton pump inhibitor (PPI) use (&lt;30 day [ref] versus ≥30 days) for head and neck cancer (HNC) according to age, sex, income, and region of residence (<b>a</b>), as well as comorbidities (Charlson comorbidity index [CCI]), history of gastroesophageal reflux disease (GERD), and H<sub>2</sub> blocker use (<b>b</b>), visualized as a forest plot.</p>
Full article ">
12 pages, 1776 KiB  
Article
Are Collagen Protons Visible with the Zero Echo Time (ZTE) Magnetic Resonance Imaging Sequence: A D2O Exchange and Freeze-Drying Study
by Tan Guo, Dina Moazamian, Arya A. Suprana, Saeed Jerban, Eric Y. Chang, Yajun Ma, Michael Carl, Min Chen and Jiang Du
Bioengineering 2025, 12(1), 16; https://doi.org/10.3390/bioengineering12010016 - 28 Dec 2024
Viewed by 262
Abstract
It is known that ultrashort echo time (UTE) magnetic resonance imaging (MRI) sequences can detect signals from water protons but not collagen protons in short T2 species such as cortical bone and tendons. However, whether collagen protons are visible with the zero echo [...] Read more.
It is known that ultrashort echo time (UTE) magnetic resonance imaging (MRI) sequences can detect signals from water protons but not collagen protons in short T2 species such as cortical bone and tendons. However, whether collagen protons are visible with the zero echo time (ZTE) MRI sequence is still unclear. In this study, we investigated the potential of the ZTE MRI sequence on a clinical 3T scanner to directly image collagen protons via D2O exchange and freeze-drying experiments. ZTE and UTE MRI sequences were employed to image fully hydrated bovine cortical bone (n = 10) and human patellar tendon (n = 1) specimens. Then, each specimen was kept in a 30 mL syringe filled with D2O solution for two days. Fresh D2O was flushed every 2 h to reach a more complete D2O–H2O exchange. Later, the samples were lyophilized for over 40 h and then sealed in tubes. Finally, the samples were brought to room temperature and visualized using the identical 3D ZTE and UTE sequences. All hydrated bone and tendon specimens showed high signals with ZTE and UTE sequences. However, all specimens showed zero signal after the D2O exchange and freeze-drying procedures. Therefore, similar to UTE imaging, the signal source in ZTE imaging is water. The ZTE sequence cannot directly detect signals from collagen protons in bone and tendons. Full article
Show Figures

Figure 1

Figure 1
<p>The 3D ZTE sequence utilizes a short rectangular RF pulse (duration = 8 µs, flip angle = 4°) for nonselective excitation, followed by 3D center-out radial sampling during fully ramped-up readout gradients (<b>A</b>). The 3D UTE sequence employs a short rectangular RF pulse (duration = 32 µs, flip angle = 10°) for nonselective excitation, followed by 3D center-out radial ramp sampling (<b>B</b>).</p>
Full article ">Figure 2
<p>A flow diagram for the experimental procedure. PBS, phosphate-buffered saline.</p>
Full article ">Figure 3
<p>Fully hydrated bovine cortical bone samples were imaged with 3D ZTE (<b>A</b>) and UTE sequences (<b>B</b>), along with ZTE (<b>C</b>) and UTE (<b>D</b>) imaging of the same bone specimens after two days of repeated D<sub>2</sub>O exchange followed by freeze-drying for over 40 h. Both the ZTE and the UTE sequences show high signals for the hydrated cortical bone samples but zero signals after the repeated D<sub>2</sub>O exchange and freeze-drying procedure.</p>
Full article ">Figure 4
<p>A fully hydrated cadaveric human patellar tendon sample was imaged with 3D ZTE (<b>A</b>) and UTE sequences (<b>B</b>), along with ZTE (<b>C</b>) and UTE (<b>D</b>) imaging of the same patellar tendon specimen after 2 days of repeated D<sub>2</sub>O exchange followed by freeze-drying for over 40 h. The hydrated patellar tendon sample shows a high signal with both ZTE and UTE sequences. After the repeated D<sub>2</sub>O exchange and freeze-drying procedure, only thin bright lines were observed, which showed typical fat/water in-phase and out-phase behaviors.</p>
Full article ">
17 pages, 1408 KiB  
Article
Ab Initio Study of Electron Capture in Collisions of Protons with CO2 Molecules
by Luis Méndez and Ismanuel Rabadán
Molecules 2025, 30(1), 74; https://doi.org/10.3390/molecules30010074 (registering DOI) - 28 Dec 2024
Viewed by 201
Abstract
Ab initio calculations of cross sections for electron capture by protons in collisions with CO2 are carried out at energies between 100 eV/u and 50 keV/u, employing a semiclassical method within the Franck–Condon framework. The scattering wave function is expanded in a [...] Read more.
Ab initio calculations of cross sections for electron capture by protons in collisions with CO2 are carried out at energies between 100 eV/u and 50 keV/u, employing a semiclassical method within the Franck–Condon framework. The scattering wave function is expanded in a set of ab initio electronic wave functions of the HCO2+ supermolecule. The calculation is performed on several trajectory orientations to obtain orientation-averaged total cross sections. A two-state model with an exponential interaction between the entrance and the lowest charge transfer channel is proposed to describe the main aspects of the charge transfer process and to estimate the precision of the molecular expansion. The symmetry of the HOMO πg of CO2 is relevant to choose the signs of the molecular functions and to set up the orientation average of the cross sections. Very good agreement is found with the experimental charge transfer cross sections. Full article
Show Figures

Figure 1

Figure 1
<p>Potential energy curves (top panels) and dynamical couplings (bottom panels) along trajectories (<b>a</b>) t<sub>‖</sub>, (<b>b</b>) t<sub>⊥0</sub>, and (<b>c</b>) t<sub>⊥90</sub>, with <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>4.0</mn> <mo> </mo> <mrow> <mi>bohr</mi> </mrow> </mrow> </semantics></math>. The dynamical couplings shown in the bottom panels are coupling <math display="inline"><semantics> <msub> <mi>M</mi> <mn>12</mn> </msub> </semantics></math>, black lines; <math display="inline"><semantics> <msub> <mi>M</mi> <mn>13</mn> </msub> </semantics></math>, red lines; and <math display="inline"><semantics> <msub> <mi>M</mi> <mn>23</mn> </msub> </semantics></math>, green lines. The adiabatic states are numbered in increasing energy order, as shown in the top panels. The two green lines in the energies of panel (<b>c</b>) correspond to the two molecular states dissociating in H(1s) + CO<sub>2</sub><sup>+</sup> (<math display="inline"><semantics> <mrow> <msup> <mi>A</mi> <mn>2</mn> </msup> <msub> <mi mathvariant="normal">Π</mi> <mi>u</mi> </msub> </mrow> </semantics></math>).</p>
Full article ">Figure 2
<p>Color maps of the dynamical couplings, <math display="inline"><semantics> <mfenced separators="" open="&#x2329;" close="&#x232A;"> <msub> <mi>χ</mi> <mn>1</mn> </msub> <mfenced separators="" open="|" close="|"> <mi mathvariant="normal">d</mi> <mo>/</mo> <mi mathvariant="normal">d</mi> <mi>Z</mi> </mfenced> <msub> <mi>χ</mi> <mn>2</mn> </msub> </mfenced> </semantics></math>, between the first two electronic states of (H + CO<sub>2</sub>)<sup>+</sup> system for trajectories parallel (top) and perpendicular (bottom) to the CO<sub>2</sub> molecular axis. Comparison between the Demkov model (Equation (<a href="#FD8-molecules-30-00074" class="html-disp-formula">8</a>)) with <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.006</mn> </mrow> </semantics></math> hartree and the ab initio data.</p>
Full article ">Figure 3
<p>Energy difference and coupling used in the 2-state model of <a href="#sec2dot2-molecules-30-00074" class="html-sec">Section 2.2</a> with <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>0</mn> </msub> <mo>=</mo> <mo>+</mo> <mn>0.004</mn> </mrow> </semantics></math> hartree (black lines) and <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>0</mn> </msub> <mo>=</mo> <mo>−</mo> <mn>0.006</mn> </mrow> </semantics></math> hartree (red lines) to estimate the error produced by the swapping of asymptotic energies, while keeping the same interaction, on the CT cross-section.</p>
Full article ">Figure 4
<p>Charge transfer cross sections obtained with the two-state models and the ab initio results for trajectories parallel to the CO<sub>2</sub> axis (t<sub>‖</sub>). The full lines are the results from the models with <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>0</mn> </msub> <mo>=</mo> <mo>+</mo> <mn>0.004</mn> </mrow> </semantics></math> hartree (black line) and <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>0</mn> </msub> <mo>=</mo> <mo>−</mo> <mn>0.006</mn> </mrow> </semantics></math> hartree of (red line). The dashed lines are the ab initio results for production of CO<sub>2</sub><sup>+</sup>(<math display="inline"><semantics> <mrow> <msup> <mi>X</mi> <mn>2</mn> </msup> <msub> <mi mathvariant="normal">Π</mi> <mi>g</mi> </msub> </mrow> </semantics></math>) (long-dashed green) and for production of CO<sub>2</sub><sup>+</sup>(<math display="inline"><semantics> <mrow> <msup> <mi>A</mi> <mn>2</mn> </msup> <msub> <mi mathvariant="normal">Π</mi> <mi>u</mi> </msub> </mrow> </semantics></math>) (short-dashed orange). Symbols are experimental data: • [<a href="#B5-molecules-30-00074" class="html-bibr">5</a>]; ▴ [<a href="#B6-molecules-30-00074" class="html-bibr">6</a>]; ⧫ [<a href="#B7-molecules-30-00074" class="html-bibr">7</a>]; ▾ [<a href="#B8-molecules-30-00074" class="html-bibr">8</a>]; ◂ [<a href="#B10-molecules-30-00074" class="html-bibr">10</a>].</p>
Full article ">Figure 5
<p>Color maps of the charge transfer probability in proton collisions with CO<sub>2</sub> as a function of the impact parameter <span class="html-italic">b</span> and the impact energy <span class="html-italic">E</span> with the 2-state model of <a href="#sec2dot2-molecules-30-00074" class="html-sec">Section 2.2</a> with <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>0</mn> </msub> <mo>=</mo> <mo>−</mo> <mn>0.006</mn> </mrow> </semantics></math> hartree (top) and <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mn>0</mn> </msub> <mo>=</mo> <mo>+</mo> <mn>0.004</mn> </mrow> </semantics></math> hartree (bottom).</p>
Full article ">Figure 6
<p>Color maps showing the probability of electron capture as function of the position of the impact parameter vector, <math display="inline"><semantics> <mi mathvariant="bold-italic">b</mi> </semantics></math> in the Cartesian plane, with <span class="html-italic"><b>v</b></span> perpendicular to this plane. Each panel corresponds to a different impact energy, as labeled. The position of the molecule in the coordinate system is also shown.</p>
Full article ">Figure 7
<p>Orientation-averaged CT cross-sections in proton collisions with CO<sub>2</sub> as functions of the impact energy. The lines are the present calculations of the total OAXS, obtained by averaging over different number of trajectory orientations: 3t, (<a href="#FD20-molecules-30-00074" class="html-disp-formula">20</a>); 4t, (<a href="#FD21-molecules-30-00074" class="html-disp-formula">21</a>); 5t, (<a href="#FD22-molecules-30-00074" class="html-disp-formula">22</a>); and 6t, (<a href="#FD23-molecules-30-00074" class="html-disp-formula">23</a>), as indicated in the figure. Experimental results: • [<a href="#B5-molecules-30-00074" class="html-bibr">5</a>]; ▴ [<a href="#B6-molecules-30-00074" class="html-bibr">6</a>]; ⧫ [<a href="#B7-molecules-30-00074" class="html-bibr">7</a>]; ▾ [<a href="#B8-molecules-30-00074" class="html-bibr">8</a>]; ◂ [<a href="#B10-molecules-30-00074" class="html-bibr">10</a>].</p>
Full article ">Figure 8
<p>Orientation-averaged charge-transfer cross section in proton collisions with CO<sub>2</sub> as a function of the impact energy. Lines are the present calculations: solid line, total charge transfer; red-dashed line, charge transfer to CO<sub>2</sub><sup>+</sup>(X); and green dashed-dotted line, charge transfer to CO<sub>2</sub><sup>+</sup>(A). Symbols are experimental data: • [<a href="#B5-molecules-30-00074" class="html-bibr">5</a>]; ▴ [<a href="#B6-molecules-30-00074" class="html-bibr">6</a>]; ⧫ [<a href="#B7-molecules-30-00074" class="html-bibr">7</a>]; ▾ [<a href="#B8-molecules-30-00074" class="html-bibr">8</a>]; ◂ [<a href="#B10-molecules-30-00074" class="html-bibr">10</a>].</p>
Full article ">Figure 9
<p>Scheme with the two families of projectile trajectories employed to study H<sup>+</sup> + CO<sub>2</sub> collisions: on the left, the t<sub>⊥</sub> family with <math display="inline"><semantics> <mi mathvariant="bold-italic">v</mi> </semantics></math> perpendicular to the molecular axis, <math display="inline"><semantics> <mi mathvariant="bold-italic">ρ</mi> </semantics></math>; on the right, the t<sub>‖</sub> family, with <math display="inline"><semantics> <mi mathvariant="bold-italic">v</mi> </semantics></math> parallel to <math display="inline"><semantics> <mi mathvariant="bold-italic">ρ</mi> </semantics></math>. The t<sub>⊥</sub> family includes subfamilies with different values of the angle <math display="inline"><semantics> <mi>α</mi> </semantics></math> between the impact parameter vector, <span class="html-italic"><b>b</b></span>, and <math display="inline"><semantics> <mi mathvariant="bold-italic">ρ</mi> </semantics></math> being <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mn>30</mn> <mo>°</mo> </semantics></math>, <math display="inline"><semantics> <mn>45</mn> <mo>°</mo> </semantics></math>, <math display="inline"><semantics> <mn>60</mn> <mo>°</mo> </semantics></math>, and <math display="inline"><semantics> <mn>90</mn> <mo>°</mo> </semantics></math>.</p>
Full article ">Figure A1
<p>Potential energy curves (middle diagram) of the first two electronic states of (H-CO<sub>2</sub>)<sup>+</sup> along a projectile trajectory perpendicular to the CO<sub>2</sub> internuclear axis (top left) with impact parameter <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>8</mn> </mrow> </semantics></math> bohr, and the non-adiabatic coupling between the two states (bottom) with the molecular orbitals depicted at five stages of the collisions, drawn in color squares and marked next to the energies and coupling lines at approximately their corresponding value of <span class="html-italic">Z</span>. At the first four stages, the electronic density for the two electronic states is shown in the top-right diagram. The two molecular orbitals depicted resemble a <math display="inline"><semantics> <msub> <mi>π</mi> <mi>g</mi> </msub> </semantics></math> on the CO<sub>2</sub> molecule and the <math display="inline"><semantics> <mrow> <mn>1</mn> <mi>s</mi> </mrow> </semantics></math> on the projectile.</p>
Full article ">
46 pages, 1601 KiB  
Review
Drug-Drug Interactions Between HIV Antivirals and Concomitant Drugs in HIV Patients: What We Know and What We Need to Know
by Emanuela De Bellis, Danilo Donnarumma, Adele Zarrella, Salvatore Maria Mazzeo, Annarita Pagano, Valentina Manzo, Ines Mazza, Francesco Sabbatino, Graziamaria Corbi, Pasquale Pagliano, Amelia Filippelli and Valeria Conti
Pharmaceutics 2025, 17(1), 31; https://doi.org/10.3390/pharmaceutics17010031 - 28 Dec 2024
Viewed by 524
Abstract
Highly active antiretroviral therapy has led to a significant increase in the life expectancy of people living with HIV. The trade-off is that HIV-infected patients often suffer from comorbidities that require additional treatment, increasing the risk of Drug-Drug Interactions (DDIs), the clinical relevance [...] Read more.
Highly active antiretroviral therapy has led to a significant increase in the life expectancy of people living with HIV. The trade-off is that HIV-infected patients often suffer from comorbidities that require additional treatment, increasing the risk of Drug-Drug Interactions (DDIs), the clinical relevance of which has often not been determined during registration trials of the drugs involved. Therefore, it is important to identify potential clinically relevant DDIs in order to establish the most appropriate therapeutic approaches. This review aims to summarize and analyze data from studies published over the last two decades on DDI-related adverse clinical outcomes involving anti-HIV drugs and those used to treat comorbidities. Several studies have examined the pharmacokinetics and tolerability of different drug combinations. Protease inhibitors, followed by nonnucleoside reverse transcriptase inhibitors and integrase inhibitors have been recognized as the main players in DDIs with antivirals used to control co-infection, such as Hepatitis C virus, or with drugs commonly used to treat HIV comorbidities, such as lipid-lowering agents, proton pump inhibitors and anticancer drugs. However, the studies do not seem to be consistent with regard to sample size and follow-up, the drugs involved, or the results obtained. It should be noted that most of the available studies were conducted in healthy volunteers without being replicated in patients. This hampered the assessment of the clinical burden of DDIs and, consequently, the optimal pharmacological management of people living with HIV. Full article
Show Figures

Figure 1

Figure 1
<p>Number of studies published between 2000 and 2024, conducted in healthy volunteers and <span class="html-italic">HIV</span>+ patients that investigated potential DDI-related adverse outcomes involving drugs belonging to HAART and drugs co-administered for the treatment of comorbidities. The percentages of studies, calculated from the total number of studies published year by year, are marked within each histogram.</p>
Full article ">Figure 2
<p>Number of studies published between 2000 and 2024, that investigated potential DDI-related adverse outcomes involving drugs belonging to HAART and drugs co-administered for the treatment of co-infections (malaria, tuberculosis and HCV). The percentages of studies, calculated from the total number of studies published year by year, are marked within each histogram. (Panel <b>A</b>) shows the number of studies, conducted in HIV mono-infected and HIV/Malaria co-infected patients, that investigated potential DDI-related adverse outcomes involving drugs belonging to HAART and drugs co-administered for the treatment of malaria. (Panel <b>B</b>) shows the number of studies, conducted in HIV/Tuberculosis co-infected patients, that investigated potential DDI-related adverse outcomes involving drugs belonging to HAART and drugs co-administered for the treatment of Tuberculosis. (Panel <b>C</b>) shows the number of studies, in healthy volunteers and HIV/HCV co-infected patients, that investigated potential DDI-related adverse outcomes involving drugs belonging to HAART and drugs co-administered for the treatment of HCV.</p>
Full article ">
17 pages, 2061 KiB  
Article
Development of a Polygenic Risk Score for Metabolic Dysfunction-Associated Steatotic Liver Disease Prediction in UK Biobank
by Panagiota Giardoglou, Ioanna Gavra, Athina I. Amanatidou, Ioanna Panagiota Kalafati, Panagiotis Symianakis, Maria Kafyra, Panagiotis Moulos and George V. Dedoussis
Genes 2025, 16(1), 33; https://doi.org/10.3390/genes16010033 - 28 Dec 2024
Viewed by 331
Abstract
Background: Metabolic dysfunction-associated steatotic liver disease (MASLD) is the leading cause of liver-related morbidity and mortality. Although the invasive liver biopsy remains the golden standard for MASLD diagnosis, Magnetic Resonance Imaging-derived Proton Density Fat Fraction (MRI-PDFF) is an accurate, non-invasive method for the [...] Read more.
Background: Metabolic dysfunction-associated steatotic liver disease (MASLD) is the leading cause of liver-related morbidity and mortality. Although the invasive liver biopsy remains the golden standard for MASLD diagnosis, Magnetic Resonance Imaging-derived Proton Density Fat Fraction (MRI-PDFF) is an accurate, non-invasive method for the assessment of treatment response. This study aimed at developing a Polygenic Risk Score (PRS) to improve MRI-PDFF prediction using UK Biobank data to assess an individual’s genetic liability to MASLD. Methods: We iteratively sequestered 10% of MRI-PDFF samples as a validation set and split the rest of each dataset into base and target partitions, containing GWAS summary statistics and raw genotype data, respectively. PRSice2 was deployed to derive PRS candidates. Based on the frequency of SNP appearances along the PRS candidates, we generated different SNP sets according to variable frequency cutoffs. By applying the PRSs to the validation set, we identified the optimal SNP set, which was then applied to a Greek nonalcoholic fatty liver disease (NAFLD) study. Results: Data from 3553 UK Biobank participants yielded 49 different SNP sets. After calculating the PRS on the validation set for every SNP set, an optimal PRS with 75 SNPs was selected (incremental R2 = 0.025, p-value = 0.00145). Interestingly, 43 SNPs were successfully mapped to MASLD-related known genes. The selected PRS could predict traits, like LDL cholesterol and diastolic blood pressure in the UK Biobank, as also disease outcome in the Greek NAFLD study. Conclusions: Our findings provide strong evidence that PRS is a powerful prediction model for MASLD, while it can also be applied on populations of different ethnicity. Full article
(This article belongs to the Section Bioinformatics)
Show Figures

Figure 1

Figure 1
<p>Flow chart of UKBB participants who met the inclusion/exclusion criteria for the study.</p>
Full article ">Figure 2
<p>After completion of iterative and aggregation PRS derivation process, different sets of SNP contents were generated depending on frequency cutoffs. Validation of the 68 generated SNP sets resulted in 20 optimal SNP sets that were statistically significant and exhibited explanatory power for the MRI-PDFF validation dataset. X-axis: number of SNPs per SNP set; y-axis: incremental R<sup>2</sup>, color-coding denotes the <span class="html-italic">p</span>-value. For the graph illustration, ggplot2 (v 3.5.1) [<a href="#B28-genes-16-00033" class="html-bibr">28</a>] was used.</p>
Full article ">Figure 3
<p>A final 75 SNPs containing PRS was selected based on high predictive value and low SNP content. The selected PRS was also evaluated for its predictive ability of NASH score in the Greek NAFLD study. The model that was generated when the 75 SNP set of the optimal PRS was applied yielded notable metrics, including <span class="html-italic">p</span>-value = 0.009 and incremental R<sup>2</sup> value = 0.003.</p>
Full article ">Figure 4
<p>A final 75 SNPs containing PRS was selected based on high predictive value and low SNP content. The optimal PRS was applied to the whole set of UK Biobank samples and evaluated for its predicted ability of MRI-PDFF (<span class="html-italic">p</span>-value = 0.001 and incremental R<sup>2</sup> value = 0.025).</p>
Full article ">Figure 5
<p>Phenotypic variance of UKBB explained for MASLD-related markers by the optimal 75 SNP-containing PRS. The graph illustration was performed using the software GraphPad Prism version 5.03.</p>
Full article ">Figure 6
<p>Enrichment analysis of the gene set in order to investigate molecular functions in which genes associated with the SNP set of the optimal PRS are involved.</p>
Full article ">
18 pages, 5622 KiB  
Article
Dimer Is Not Double: The Unexpected Behavior of Two-Floor Peptide Nanosponge
by Grazia Maria Lucia Messina, Marta De Zotti, Alvaro S. Siano, Claudia Mazzuca, Giovanni Marletta and Antonio Palleschi
Molecules 2025, 30(1), 47; https://doi.org/10.3390/molecules30010047 - 26 Dec 2024
Viewed by 217
Abstract
Using the framework of an investigation of the stimuli-responsive behavior of peptide assembly on a solid surface, this study on the behavior of a chemisorbed peptide on a gold surface was performed. The studied peptide is a dimeric form of the antimicrobial peptide [...] Read more.
Using the framework of an investigation of the stimuli-responsive behavior of peptide assembly on a solid surface, this study on the behavior of a chemisorbed peptide on a gold surface was performed. The studied peptide is a dimeric form of the antimicrobial peptide Trichogin GAIV, which was also modified by substituting the glycine with lysine residues, while the N-terminus octanoyl group was replaced by a lipoic one that was able to bind to the gold surface. In this way, a chemically linked peptide assembly that is pH-responsive was obtained because of the protonation/deprotonation of the sidechains of the Lys residues. Information about the effect of protonation/deprotonation equilibria switching the pH from acid (pH = 3) to basic (pH = 11) conditions was obtained macroscopically by performing Quartz crystal microbalance with dissipation monitoring (QCM-D), Surface Plasmon Resonance (SPR), Nanoplasmonic Sensing (NPS), and FTIR techniques. Using molecular dynamics (MD) simulations, it is possible to explain, at the molecular level, our main experimental results: (1) pH changes induce a squeezing behavior in the system, consisting in thickness and mass variations in the peptide layer, which are mainly due to the pH-driven hydrophilic/hydrophobic character of the lysine residues, and (2) the observed hysteresis is due to small conformational rearrangements from helix to beta sheets occurring mainly on the first half of the peptide, closer to the surface, while the second half remains almost unaffected. The latter result, together with the evidence that the layer thickness is not simply double the assembly of the monomeric analog, indicates that the dimeric peptide does not behave as a double monomer, but assumes very peculiar features. Full article
(This article belongs to the Section Computational and Theoretical Chemistry)
Show Figures

Figure 1

Figure 1
<p>NPS data showing the change in resonance plasmon, that is, the thickness change, for the L2 peptide layer with the change from pH = 3 (starting baseline) to pH = 11.</p>
Full article ">Figure 2
<p>QCM-D data showing the cyclic behavior of frequency increase (basic condition) and recovery (acid condition) of the L2 monolayer (red) and of gold (yellow) by changing the pH. The baseline is set for the sample in an acid environment at pH = 3.</p>
Full article ">Figure 3
<p>Weight loss/gain after each pH variation obtained using the QCM-D experimental data (blue circles) and MD simulations (red circles). See <a href="#sec3-molecules-30-00047" class="html-sec">Section 3</a> for details.</p>
Full article ">Figure 4
<p>Peptide density as a function of distance from the surface of (<b>A</b>) for the first cycle (pH = 3 (red) and pH = 11 (blue)), (<b>B</b>,<b>C</b>) for the subsequent cycles (I: red; II: green; III: blue; IV: brown), (<b>B</b>) at pH = 3 and (<b>C</b>) at pH = 11. Arrows indicate the trend of peptide thickness.</p>
Full article ">Figure 5
<p>Water density as a function of distance from the surface for the simulations at pH = 3 (<b>A</b>) and at pH = 11 (<b>B</b>); (<b>C</b>) ion density as a function of distance from the surface for the simulations at pH = 3 (continuous lines) or pH = 11 (dashed lines). Colors of the cycles: I: red; II: green; III: blue; IV: brown. Arrow is a guide for the eyes.</p>
Full article ">Figure 6
<p>Top view snapshots of the last frame of the eight MD simulations concerning the four pH change cycles: (<b>A</b>,<b>B</b>) I cycle; (<b>C</b>,<b>D</b>) II cycle; (<b>E</b>,<b>F</b>) III cycle; (<b>G</b>,<b>H</b>) IV cycle). (<b>A</b>,<b>C</b>,<b>D</b>,<b>F</b>) are related to the monolayer at pH = 3; the others are related to the monolayer at pH = 11. Oxygen atoms are in red, hydrogen atoms are in white, nitrogen atoms are in blue, and carbon atoms are in cyan.</p>
Full article ">Figure 7
<p>SASA values of the last 10 ns of MD simulations, during several cycles of pH variation (pH = 3 → pH = 11). (<b>A</b>) The peptide moieties; (<b>B</b>) the LYS side chains; (<b>C</b>) hydrophobic (Leu, Ile, and Aib) side chains.</p>
Full article ">Figure 8
<p>(<b>A</b>) Percentage of secondary structures (black: helix, and green: β-sheets) upon the eight simulations mimicking the four pH cycles (pH = 3 → pH= 11). (<b>B</b>) Comparison of the percentages obtained using MD simulations and FTIR experiments of secondary structures at pH = 3 (red and violet, respectively) and at pH = 11 (blue and light blue, respectively).</p>
Full article ">Figure 9
<p>Percentage of secondary structures (red: helix; blue: β-sheets) upon the eight MD simulations corresponding to the four pH cycles (pH = 3 → pH = 11) for (<b>A</b>) the first half (closer to the surface) and (<b>B</b>) the second half (further from the surface) of the L2 peptide.</p>
Full article ">Figure 10
<p>In each panel, a top view (left) and side view (right) of the last frame of the eight MD simulations (<b>A</b>–<b>H</b>) corresponding to the four pH cycles (pH = 3 → pH = 11) of the L2 half portion closer to the surface are presented. In the snapshots, only β-sheets and helix structures (side view) and coils (top view only) in peptides are shown. Helical tracts are in red, β-sheets are in green, and coils are in cyan.</p>
Full article ">Figure 11
<p>In each panel, a top view (left) and side view (right) of the last frame of the eight MD simulations (<b>A</b>–<b>H</b>) corresponding to the four pH cycles (pH = 3 → pH = 11) of the L2 half portion further from the surface are presented. In the snapshots, only β-sheet and helix structures (side view) and coils (top view only) in peptides are shown. Helical tracts are in red, β-sheets are in green, and coils are in cyan.</p>
Full article ">Scheme 1
<p>The chemical structure of the peptide L2.</p>
Full article ">
28 pages, 7839 KiB  
Review
Progress in Continuous Flow Synthesis of Hydrogen-Bonded Organic Framework Material Synthons
by Xingjun Yao, Sanmiao Wen, Ningning Ji, Qiulin Deng, Zhiliang Li, Hongbing Wang and Qianqian Shang
Molecules 2025, 30(1), 41; https://doi.org/10.3390/molecules30010041 - 26 Dec 2024
Viewed by 373
Abstract
Hydrogen-bonded organic framework (HOF) materials are typically formed by the self-assembly of small organic units (synthons) with specific functional groups through hydrogen bonding or other interactions. HOF is commonly used as an electrolyte for batteries. Well-designed HOF materials can enhance the proton exchange [...] Read more.
Hydrogen-bonded organic framework (HOF) materials are typically formed by the self-assembly of small organic units (synthons) with specific functional groups through hydrogen bonding or other interactions. HOF is commonly used as an electrolyte for batteries. Well-designed HOF materials can enhance the proton exchange rate, thereby boosting battery performance. This paper reviews recent advancements in the continuous synthesis of HOF synthons, in the continuous synthesis of HOF’s unit small molecules enabling the multi-step, rapid, and in situ synthesis of synthons, such as carboxylic acid, diaminotriazine (DAT), urea, guanidine, imidazole, pyrazole, pyridine, thiazole, triazole, and tetrazole, with online monitoring. Continuous flow reactors facilitate fast chemical reactions and precise microfluidic control, offering superior reaction speed, product yield, and selectivity compared to batch processes. Integrating the continuous synthesis of synthons with the construction of HOF materials on a single platform is essential for achieving low-cost, safe, and efficient processing, especially for reactions involving toxic, flammable, or explosive substances. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>Continuous flow reactor for synthesis of Hof synthons and common hydrogen bonding connection methods for synthons: carboxyl group; diaminotriazine; pyridine; carboxyl and pyridine; amidinium and carboxylate; benzimidazolone; sulfonic acid group and guanidine group; pyrazole.</p>
Full article ">Figure 2
<p>Continuous flow carboxylation yields a central fragment in the synthesis of Lifitegrast [<a href="#B33-molecules-30-00041" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>LDA-mediated carboxylation of heterocycles in flow (Adapted from Ref. [<a href="#B34-molecules-30-00041" class="html-bibr">34</a>]).</p>
Full article ">Figure 4
<p>Reactions and flow system consisting of two microreactors and in-line FT-IR spectrometer. Reprinted with permission from Ref. [<a href="#B39-molecules-30-00041" class="html-bibr">39</a>]. Copyright 2017 Copyright American Chemical Society.</p>
Full article ">Figure 5
<p>Two libraries of urea compounds based on the piperidine-4-one scaffold have been developed in continuous flow and twenty drug-like compounds were synthesized in moderate to high yields [<a href="#B40-molecules-30-00041" class="html-bibr">40</a>].</p>
Full article ">Figure 6
<p>The synthesis of urea and thioureas in continuous flow (Reprinted from Refs. [<a href="#B41-molecules-30-00041" class="html-bibr">41</a>,<a href="#B42-molecules-30-00041" class="html-bibr">42</a>]).</p>
Full article ">Figure 6 Cont.
<p>The synthesis of urea and thioureas in continuous flow (Reprinted from Refs. [<a href="#B41-molecules-30-00041" class="html-bibr">41</a>,<a href="#B42-molecules-30-00041" class="html-bibr">42</a>]).</p>
Full article ">Figure 7
<p>Schematic representation of the self-assembly of macrocycle 1 into crystalline host 1, containing columnar channels. The porous crystals can reversibly absorb a variety of guests including 2-cyclohexenone. UV-irradiation of included guests yields a photo dimer in high conversion and selectivity. Reprinted with permission from Ref. [<a href="#B44-molecules-30-00041" class="html-bibr">44</a>]. Copyright 2008 American Chemical Society.</p>
Full article ">Figure 8
<p>Tuning the binding strength of even and uneven hydrogen-bonded arrays with remote substituents (Reprinted from Ref. [<a href="#B46-molecules-30-00041" class="html-bibr">46</a>]). ED (electron-donating group), EW (electron-withdrawing group).</p>
Full article ">Figure 9
<p>Formal continuous-flow synthesis of phentolamine. Reprinted with permission from Ref. [<a href="#B48-molecules-30-00041" class="html-bibr">48</a>]. Copyright 2020 Copyright John Wiley and Sons.</p>
Full article ">Figure 10
<p>Schematic Molecular Structure of the HOF-6 Building Block; (<b>a</b>) building block of H<sub>2</sub>TDDP, (<b>b</b>) hydrogen-bonding interaction between DAT moieties, (<b>c</b>) three neighboring 2D supramolecular grids in different colors, (<b>d</b>,<b>e</b>) 3D packing supramolecular structure along the [1,0,1] and [1,0,0] directions with a channel size of ∼6.4 and 7.5 Å, respectively. Reprinted with permission from Ref. [<a href="#B49-molecules-30-00041" class="html-bibr">49</a>]. Copyright 2016 Copyright American Chemical Society.</p>
Full article ">Figure 11
<p>Scope of triazine formation in flow (Reprinted from Ref. [<a href="#B50-molecules-30-00041" class="html-bibr">50</a>]).</p>
Full article ">Figure 12
<p>Schematic representation of the continuous-flow synthesis of 8. P20 and P21 are pumps (Reprinted from Ref. [<a href="#B51-molecules-30-00041" class="html-bibr">51</a>]).</p>
Full article ">Figure 13
<p>1-GH and HMDA aqueous solution; 2-medium-pressure syringe pump module; 3-PC; 4-microreactor; 5-cyclone; 6-flask-receiver of ammonia; 7-rotary evaporator (Reprinted from Ref. [<a href="#B55-molecules-30-00041" class="html-bibr">55</a>]).</p>
Full article ">Figure 14
<p>The BrCN generator set-up. PFA coil (0.8 mm i.d.) at 0–58 °C: 5.2 min residence time. Glass microreactor chip: 20 s residence time; liquid/liquid separator, FTIR. Reprinted with permission from Ref. [<a href="#B56-molecules-30-00041" class="html-bibr">56</a>]. Copyright 2017 Copyright John Wiley and Sons.</p>
Full article ">Figure 15
<p>Continuous flow synthesis of AS-136A. A continuous flow telescoped synthesis with modules 1, 2, 4, and 5 to yield measles therapeutic AS-136A in a 34% isolated yield. Using modules individually yields AS-136A in a 75% yield. Reprinted with permission from Ref. [<a href="#B57-molecules-30-00041" class="html-bibr">57</a>]. Copyright 2017 Copyright John Wiley and Sons.</p>
Full article ">Figure 16
<p>Continuous-flow synthesis of <b>3i</b> as a key intermediate. Reprinted with permission from Ref. [<a href="#B59-molecules-30-00041" class="html-bibr">59</a>]. Copyright 2019 Copyright John Wiley and Sons.</p>
Full article ">Figure 17
<p>Schematic setup for the epoxide ring-opening reaction under flow conditions. Enzymatic kinetic Resolution of Racemic trans-2-(1H-imidazol-1-yl) cycloalkanols trans-3a-b in fixed bed. Reprinted with permission from Ref. [<a href="#B62-molecules-30-00041" class="html-bibr">62</a>]. Copyright 2012 Copyright American Chemical Society.</p>
Full article ">Figure 18
<p>Continuous Flow Formation of Imidazole 4b from Acyloxy Ketone <b>3b</b>. Reprinted with permission from Ref. [<a href="#B64-molecules-30-00041" class="html-bibr">64</a>]. Copyright 2015 Copyright American Chemical Society.</p>
Full article ">Figure 19
<p>Continuous synthesis of 2,4,5-trisubstituted imidazoles using a microreactor system under pressure. Reprinted with permission from Ref. [<a href="#B65-molecules-30-00041" class="html-bibr">65</a>]. Copyright 2010 Copyright American Chemical Society.</p>
Full article ">Figure 20
<p>Synthesis of annulated pyridines in flow. Reprinted with permission from Ref. [<a href="#B71-molecules-30-00041" class="html-bibr">71</a>]. Copyright 2012 Copyright John Wiley and Sons.</p>
Full article ">Figure 21
<p>Flow synthesis of thiadiazoles. Reprinted with permission from Ref. [<a href="#B76-molecules-30-00041" class="html-bibr">76</a>]. Copyright 2017 Copyright Elsevier.</p>
Full article ">Figure 22
<p>Fully telescoped synthesis of the thiazole target <b>3</b> (Reprinted from Ref. [<a href="#B78-molecules-30-00041" class="html-bibr">78</a>]).</p>
Full article ">Figure 23
<p>Flow synthesis of 5-amino-4-cyano-1,2,3-triazole. Reprinted with permission from Ref. [<a href="#B83-molecules-30-00041" class="html-bibr">83</a>]. Copyright 2024 Copyright American Chemical Society.</p>
Full article ">Figure 24
<p>Continuous flow synthesis of 1,2,4-oxadiazole 2 (Reprinted from Ref. [<a href="#B84-molecules-30-00041" class="html-bibr">84</a>]).</p>
Full article ">Figure 25
<p>This work is to develop a three-step flow preparation of nonafyl azide, diazo transfer to primary amines, and subsequent CuAAC formation of triazoles (Reprinted from Ref. [<a href="#B86-molecules-30-00041" class="html-bibr">86</a>]).</p>
Full article ">Figure 26
<p>Evaluation of the conditions for step A and B-cyclization in continuous flow regime [<a href="#B88-molecules-30-00041" class="html-bibr">88</a>].</p>
Full article ">Figure 27
<p>Schematic diagram for the continuous flow tetrazole synthesis performed in a FlowSyn reactor: static glass mixer block (2.0 mL internal volume); C: coil reactor (Sulfinert, 10.7 mL internal volume, 1.0 mm i.d.); HE: heat-exchanger; BPR: back pressure regulator (34 bar) [<a href="#B89-molecules-30-00041" class="html-bibr">89</a>].</p>
Full article ">Figure 28
<p>Laboratory setup for process development at Roche, including continuous aqueous quench and extraction. Product 1 was isolated in high purity following the steps outlined here. Impurity 4 was avoided by preventing the reaction of product 1 with azide anion. Reprinted with permission from Ref. [<a href="#B90-molecules-30-00041" class="html-bibr">90</a>]. Copyright 2021 Copyright American Chemical Society.</p>
Full article ">Figure 29
<p>Safe and efficient tetrazole synthesis in a continuous flow microreactor [<a href="#B91-molecules-30-00041" class="html-bibr">91</a>].</p>
Full article ">Figure 30
<p>Continuous flow alkylation of 1H-Tz via amine diazotization [<a href="#B94-molecules-30-00041" class="html-bibr">94</a>].</p>
Full article ">Scheme 1
<p>Improved synthetic protocol to prepare 7-bromopyrrolo [2,1-f] [1,2,4] triazin-4-amine (4) (Reprinted from Ref. [<a href="#B51-molecules-30-00041" class="html-bibr">51</a>]).</p>
Full article ">Scheme 2
<p>One-step continuous flow synthesis of <span class="html-italic">N</span>-alkyl imidazoles (Reprinted from Ref. [<a href="#B63-molecules-30-00041" class="html-bibr">63</a>]).</p>
Full article ">Scheme 3
<p>[1.3]-prototropic shift of ethylpropylideneamine and consecutive formation of 1-ethyl-2, 4, 5-trimethylimidazole in alkaline media [<a href="#B48-molecules-30-00041" class="html-bibr">48</a>].</p>
Full article ">Scheme 4
<p>Synthesis of sugar-containing pyrimidine compounds in continuous-flow microreactors (Reprinted from Ref. [<a href="#B70-molecules-30-00041" class="html-bibr">70</a>]).</p>
Full article ">Scheme 5
<p>Scheme of benzonitrile azidation and 5-phenyltetrazole methylation [<a href="#B93-molecules-30-00041" class="html-bibr">93</a>].</p>
Full article ">
15 pages, 5849 KiB  
Article
Damage on a Solid–Liquid Interface Induced by the Dynamical Behavior of Injected Gas Bubbles in Flowing Mercury
by Hiroyuki Kogawa, Takashi Wakui and Masatoshi Futakawa
Fluids 2025, 10(1), 3; https://doi.org/10.3390/fluids10010003 - 26 Dec 2024
Viewed by 206
Abstract
Microbubbles have been applied in various fields. In the mercury targets of spallation neutron sources, where cavitation damage is a crucial issue for life estimation, microbubbles are injected into the mercury to absorb the thermal expansion of the mercury caused by the pulsed [...] Read more.
Microbubbles have been applied in various fields. In the mercury targets of spallation neutron sources, where cavitation damage is a crucial issue for life estimation, microbubbles are injected into the mercury to absorb the thermal expansion of the mercury caused by the pulsed proton beam injection and reduce the macroscopic pressure waves, which results in reducing the damage. Recently, when the proton beam power was increased and the number of injected gas bubbles was increased, unique damage morphologies were observed on the solid–liquid interface. Detailed observation and numerical analyses revealed that the microscopic pressure emitted from the gas bubbles contracting is sufficient to form pit damage, i.e., the directions of streak-like defects which are formed by connecting the pit damage coincides with the direction of the gas bubble trajectories, and the distances between the pits was understandable when taking the natural period of gas bubble vibration into account. This indicates that gas microbubbles, used to reduce macroscopic pressure waves, have the potential to be inceptions of cavitation damage due to the microscopic pressure emitted from these gas bubbles. To completely mitigate the damage, we have to consider the two effects of injecting gas bubbles: reducing macroscopic pressure waves and reducing the microscopic pressure due to bubble dynamics. Full article
(This article belongs to the Special Issue Cavitation and Bubble Dynamics)
Show Figures

Figure 1

Figure 1
<p>Cross sectional view of a mercury target vessel. Pulsed proton beams were injected at a beam window. The mercury flowed from one side to another side. At the horizontal center, mercury flowed in, perpendicular to the proton beam’s injected direction. The bubble generator was set on the inlet side.</p>
Full article ">Figure 2
<p>The gas bubble population near the beam window calculated by CFD. Many gas bubbles were distributed in the upper area because of buoyancy. The mean bubble radius was larger in the upper area than that in the lower area.</p>
Full article ">Figure 3
<p>Damage at the beam window in contact with mercury in the cases of (<b>a</b>) low gas bubble injection and (<b>b</b>) relatively high gas bubble injection. Damage was mitigated in (<b>b</b>) but inclined streak-like defects were observed.</p>
Full article ">Figure 4
<p>Gas bubble trajectories obtained by CFD analysis. Swirl flow, generated in a bubble generator, affected the slopes of the trajectories, which were similar with those of the streak-like defects observed in <a href="#fluids-10-00003-f003" class="html-fig">Figure 3</a>b.</p>
Full article ">Figure 5
<p>Detailed observation of the streak-like defects in (<b>a</b>) Area 1 and (<b>b</b>) Area 2 in <a href="#fluids-10-00003-f003" class="html-fig">Figure 3</a>b. Beaded pits were observed in the streak-like defects. The distance to the next pit, ∆<span class="html-italic">D</span>, was relatively narrow in the lower area in (<b>b</b>).</p>
Full article ">Figure 6
<p>Comparison of the <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>C</mi> <mi>F</mi> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> distribution based on the bubble distribution by CFD and the observed <math display="inline"><semantics> <mrow> <mo>∆</mo> <mi>D</mi> </mrow> </semantics></math> distribution for the (<b>a</b>) upper area (<a href="#fluids-10-00003-f005" class="html-fig">Figure 5</a>a) and (<b>b</b>) lower area (<a href="#fluids-10-00003-f005" class="html-fig">Figure 5</a>b). The observed distribution of <math display="inline"><semantics> <mrow> <mo>∆</mo> <mi>D</mi> </mrow> </semantics></math> was similar to the distribution of <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>C</mi> <mi>F</mi> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>. The correlation between bubble size and damage was recognized.</p>
Full article ">Figure 7
<p>Relationship between the maximum pit depth, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math>, and the pit radius, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>d</mi> </mrow> </msub> </mrow> </semantics></math>, of the isolated pits observed in Area 1 and Area 2 of <a href="#fluids-10-00003-f003" class="html-fig">Figure 3</a>b. The <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> in Area 1 and Area 2 were distributed with similar variations, but the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>d</mi> </mrow> </msub> </mrow> </semantics></math> tended to be larger in Area 1, where larger gas bubbles existed.</p>
Full article ">Figure 8
<p>The pressure change around a gas bubble. This time history was used for bubble dynamics calculations in Equation (3).</p>
Full article ">Figure 9
<p>Time behaviors of the gas bubble radius and the pressure at the bubble interface at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>g</mi> <mo>,</mo> <mn>0</mn> </mrow> </msub> <mo>=</mo> <mn>100</mn> <mtext> </mtext> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mo>∞</mo> <mo>,</mo> <mn>1</mn> </mrow> </msub> <mo>=</mo> <mo>−</mo> <mn>0.02</mn> <mtext> </mtext> <mi mathvariant="normal">M</mi> <mi mathvariant="normal">P</mi> <mi mathvariant="normal">a</mi> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>t</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math> = 5 ms. An impact pressure exceeding the 0.2% proof stress of the wall material was generated, which indicated that the local impact pressure generated from the gas bubble had the potential to form a pit on the wall surface.</p>
Full article ">Figure 10
<p>Schematic diagram of the analysis for the calculation of pit morphology. For analysis, the applied pressure on the wall surface was assumed that the local impact pressure, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>L</mi> <mo>,</mo> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math>, generated on the bubble surface propagated as a spherical wave. The applied pressure on the solid wall was expressed by <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>i</mi> <mi>m</mi> <mi>p</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Relationship between the initial gas bubble radius and the maximum applied pressure on the wall surface, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>P</mi> </mrow> <mrow> <mi>i</mi> <mi>m</mi> <mi>p</mi> <mo>,</mo> <mi>r</mi> <mo>=</mo> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math>, for different distances between the gas bubble and the solid wall, <math display="inline"><semantics> <mrow> <mi>h</mi> </mrow> </semantics></math>, in the case of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mo>∞</mo> <mo>,</mo> <mn>1</mn> </mrow> </msub> <mo>=</mo> <mo>−</mo> <mn>0.02</mn> <mtext> </mtext> <mi mathvariant="normal">M</mi> <mi mathvariant="normal">P</mi> <mi mathvariant="normal">a</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>t</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> <mo>=</mo> <mn>5</mn> <mtext> </mtext> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>. The maximum applied pressure, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>P</mi> </mrow> <mrow> <mi>i</mi> <mi>m</mi> <mi>p</mi> <mo>,</mo> <mi>r</mi> <mo>=</mo> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math>, increased with increases in <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>g</mi> <mo>,</mo> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> and decreases in <math display="inline"><semantics> <mrow> <mi>h</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Photograph of a series of the isolated pits together with the pressure history emitted from the gas bubble (Enlarged <a href="#fluids-10-00003-f009" class="html-fig">Figure 9</a>).</p>
Full article ">Figure 13
<p>Formation mechanism of the streak-like defect with beaded pits.</p>
Full article ">Figure 14
<p>Distribution in the radial direction of the MICP applied on the solid wall, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>P</mi> </mrow> <mrow> <mi>i</mi> <mi>m</mi> <mi>p</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> </mrow> </semantics></math>, while varying <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>g</mi> <mo>,</mo> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>h</mi> </mrow> </semantics></math>.</p>
Full article ">
Back to TopTop