[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

Article Types

Countries / Regions

Search Results (165)

Search Parameters:
Keywords = nanobubbles

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 3612 KiB  
Article
Effect of CO2 Nanobubble Water on the Fracture Properties of Cemented Backfill Materials under Different Aggregate Fractal Dimensions
by Xiaoxiao Cao, Akihiro Hamanaka, Hideki Shimada and Takashi Sasaoka
Appl. Sci. 2024, 14(17), 7792; https://doi.org/10.3390/app14177792 - 3 Sep 2024
Viewed by 368
Abstract
In order to cope with climate change and achieve the goal of carbon neutrality, the use of carbonization technology to enhance the performance of cement-based materials and achieve the purpose of carbon sequestration has become a very promising research direction. This paper considers [...] Read more.
In order to cope with climate change and achieve the goal of carbon neutrality, the use of carbonization technology to enhance the performance of cement-based materials and achieve the purpose of carbon sequestration has become a very promising research direction. This paper considers the use of CO2NBW as mixing water for cement-based materials, aiming to improve the carbonization efficiency of materials to achieve the goal of carbon neutrality. This time, the effect of CO2NBW on cementitious filling materials under different aggregate fractal dimensions was studied through uniaxial compression tests and acoustic emission technology. The effect of CO2NBW on the mechanical properties and crack evolution of the material was discussed. The results showed that CO2 nanobubbles significantly improved the strength of cemented filling materials under different fractal dimensions, and the uniaxial compressive strength was most significantly improved by 23.04% when the fractal dimension was 2.7824. In addition, the characteristics of acoustic emission ring counts and energy parameters indicate that CO2 nanobubbles help improve the overall pore structure of the sample, affecting the macroscopic strength. However, the addition of CO2 nanobubbles reduces the limit energy storage ratio of elastic strain energy, which indicates that excessive CO2 concentration may affect the hydration reaction of the cementing material. Full article
Show Figures

Figure 1

Figure 1
<p>The distribution of particle size: cement and nanobubbles: (<b>a</b>) Cement; (<b>b</b>) nanobubble.</p>
Full article ">Figure 2
<p>Aggregate particle size distribution with different fractal dimensions (Solid lines: Interval proportion; Dotted lines: cumulative proportion).</p>
Full article ">Figure 3
<p>Sample preparation process.</p>
Full article ">Figure 4
<p>The variation in UCS with fractal dimension: (<b>a</b>) UCS; (<b>b</b>) UCS increase value and change ratio.</p>
Full article ">Figure 5
<p>Variation law of acoustic emission ringing of natural water mixed CBM samples with aggregate fractal dimension: (<b>a</b>) <span class="html-italic">D</span> = 2.2106; (<b>b</b>) <span class="html-italic">D</span> = 2.4150; (<b>c</b>) <span class="html-italic">D</span> = 2.6084; (<b>d</b>) <span class="html-italic">D</span> = 2.7824.</p>
Full article ">Figure 6
<p>Variation law of acoustic emission ringing of CO<sub>2</sub>NBW mixed CBM samples with aggregate fractal dimension: (<b>a</b>) <span class="html-italic">D</span> = 2.2106; (<b>b</b>) <span class="html-italic">D</span> = 2.4150; (<b>c</b>) <span class="html-italic">D</span> = 2.6084; (<b>d</b>) <span class="html-italic">D</span> = 2.7824.</p>
Full article ">Figure 6 Cont.
<p>Variation law of acoustic emission ringing of CO<sub>2</sub>NBW mixed CBM samples with aggregate fractal dimension: (<b>a</b>) <span class="html-italic">D</span> = 2.2106; (<b>b</b>) <span class="html-italic">D</span> = 2.4150; (<b>c</b>) <span class="html-italic">D</span> = 2.6084; (<b>d</b>) <span class="html-italic">D</span> = 2.7824.</p>
Full article ">Figure 7
<p>Variation in cumulative acoustic emission ringing counts with fractal dimension.</p>
Full article ">Figure 8
<p>Variation law of energy parameters of natural water mixed CBM samples with aggregate fractal dimension: (<b>a</b>) <span class="html-italic">D</span> = 2.2106; (<b>b</b>) <span class="html-italic">D</span> = 2.4150; (<b>c</b>) <span class="html-italic">D</span> = 2.6084; (<b>d</b>) <span class="html-italic">D</span> = 2.7824.</p>
Full article ">Figure 8 Cont.
<p>Variation law of energy parameters of natural water mixed CBM samples with aggregate fractal dimension: (<b>a</b>) <span class="html-italic">D</span> = 2.2106; (<b>b</b>) <span class="html-italic">D</span> = 2.4150; (<b>c</b>) <span class="html-italic">D</span> = 2.6084; (<b>d</b>) <span class="html-italic">D</span> = 2.7824.</p>
Full article ">Figure 9
<p>Variation law of energy parameters of CO<sub>2</sub>NBW mixed CBM samples with aggregate fractal dimension: (<b>a</b>) <span class="html-italic">D</span> = 2.2106; (<b>b</b>) <span class="html-italic">D</span> = 2.4150; (<b>c</b>) <span class="html-italic">D</span> = 2.6084; (<b>d</b>) <span class="html-italic">D</span> = 2.7824.</p>
Full article ">Figure 10
<p>Relationship between energy parameters and fractal dimensions under different mixing conditions: (<b>a</b>) natural water; (<b>b</b>) CO<sub>2</sub>NBW.</p>
Full article ">
13 pages, 580 KiB  
Article
The Quest for Industrially and Environmentally Efficient Nanobubble Engineering: Electric-Field versus Mechanical Generation Approaches
by Niall J. English
Appl. Sci. 2024, 14(17), 7636; https://doi.org/10.3390/app14177636 - 29 Aug 2024
Viewed by 336
Abstract
Nanobubbles (NBs) are gaseous domains at the nanoscale that can exist in bulk liquid or on solid surfaces. They are noteworthy for their high potential for real-world applications and their long (meta)stability. “Platform-wide” applications abound in medicine, wastewater treatment, hetero-coagulation, boundary-slip control in [...] Read more.
Nanobubbles (NBs) are gaseous domains at the nanoscale that can exist in bulk liquid or on solid surfaces. They are noteworthy for their high potential for real-world applications and their long (meta)stability. “Platform-wide” applications abound in medicine, wastewater treatment, hetero-coagulation, boundary-slip control in microfluidics, and nanoscopic cleaning. Here, we compare and contrast the industrial NB-generation performance of various types of commercial NB generators in both water-flow and submerged-in-water settings—in essence, comparing electric-field NB-generation approaches versus mechanical ones—finding that the former embodiments are superior from a variety of perspectives. It was found that the electric-field approach for NB generation surpasses traditional mechanical approaches for clean-water NB generation, especially when considering the energy running cost. In particular, more passive electric-field approaches are very operationally attractive for NB generation, where water and gas flow can be handled at little to no cost to the end operator, and/or submersible NB generators can be deployed, allowing for the use of photovoltaic approaches (with backup batteries for night-time and “low-sun” scenarios and air-/CO2-pumping paraphernalia). Full article
Show Figures

Figure 1

Figure 1
<p>Graphical summary of the scope of non-equilibrium nano-dispersed fluids, with the bottom-right inset section showing the electrostriction concept. (Image credit: Jon Tallon).</p>
Full article ">
18 pages, 5624 KiB  
Article
Investigating the Potential of CO2 Nanobubble Systems for Enhanced Oil Recovery in Extra-Low-Permeability Reservoirs
by Liyuan Cai, Jingchun Wu, Miaoxin Zhang, Keliang Wang, Bo Li, Xin Yu, Yangyang Hou and Yang Zhao
Nanomaterials 2024, 14(15), 1280; https://doi.org/10.3390/nano14151280 - 30 Jul 2024
Viewed by 862
Abstract
Carbon Capture, Utilization, and Storage (CCUS) stands as one of the effective means to reduce carbon emissions and serves as a crucial technical pillar for achieving experimental carbon neutrality. CO2-enhanced oil recovery (CO2-EOR) represents the foremost method for CO [...] Read more.
Carbon Capture, Utilization, and Storage (CCUS) stands as one of the effective means to reduce carbon emissions and serves as a crucial technical pillar for achieving experimental carbon neutrality. CO2-enhanced oil recovery (CO2-EOR) represents the foremost method for CO2 utilization. CO2-EOR represents a favorable technical means of efficiently developing extra-low-permeability reservoirs. Nevertheless, the process known as the direct injection of CO2 is highly susceptible to gas scrambling, which reduces the exposure time and contact area between CO2 and the extra-low-permeability oil matrix, making it challenging to utilize CO2 molecular diffusion effectively. In this paper, a comprehensive study involving the application of a CO2 nanobubble system in extra-low-permeability reservoirs is presented. A modified nano-SiO2 particle with pro-CO2 properties was designed using the Pickering emulsion template method and employed as a CO2 nanobubble stabilizer. The suitability of the CO2 nanobubbles for use in extra-low-permeability reservoirs was evaluated in terms of their temperature resistance, oil resistance, dimensional stability, interfacial properties, and wetting-reversal properties. The enhanced oil recovery (EOR) effect of the CO2 nanobubble system was evaluated through core experiments. The results indicate that the CO2 nanobubble system can suppress the phenomena of channeling and gravity overlap in the formation. Additionally, the system can alter the wettability, thereby improving interfacial activity. Furthermore, the system can reduce the interfacial tension, thus expanding the wave efficiency of the repellent phase fluids. The system can also improve the ability of CO2 to displace the crude oil or water in the pore space. The CO2 nanobubble system can take advantage of its size and high mass transfer efficiency, among other advantages. Injection of the gas into the extra-low-permeability reservoir can be used to block high-gas-capacity channels. The injected gas is forced to enter the low-permeability layer or matrix, with the results of core simulation experiments indicating a recovery rate of 66.28%. Nanobubble technology, the subject of this paper, has significant practical implications for enhancing the efficiency of CO2-EOR and geologic sequestration, as well as providing an environmentally friendly method as part of larger CCUS-EOR. Full article
Show Figures

Figure 1

Figure 1
<p>Janus functional particle synthesis program.</p>
Full article ">Figure 2
<p>MFDA molecular formula.</p>
Full article ">Figure 3
<p>Schematic diagram of CO<sub>2</sub> nanobubble preparation process. 1—CO<sub>2</sub> nanobubbles; 2–Masking plate; 3–Nanobubbles generator; 4—CO<sub>2</sub> cylinder.</p>
Full article ">Figure 4
<p>High-temperature and high-pressure visualization foam analyzer. 1—ISCO high-pressure piston pump; 2—Intermediate vessel; 3—Precision pressure gauge; 4—High-temperature and high-pressure visualization reactor; 5—Return valve; 6—Nanobubbles generator; 7—Gas cylinder.</p>
Full article ">Figure 5
<p>Flowchart of the experimental foam drive in two-pipe parallel core. 1—ISCO high-pressure piston pump; 2-4—Intermediate vessel; 5—One-way valve; 6–Hand pump; 7—Core gripper; 8—Return valve; 9—Gas-liquid separator.</p>
Full article ">Figure 6
<p>FTIR spectra of nano-SiO<sub>2</sub> before and after modification: (<b>a</b>) FTIR spectra of SiO<sub>2</sub> nanoparticles; (<b>b</b>) FTIR spectra of modified nano-SiO<sub>2</sub>.</p>
Full article ">Figure 7
<p>Distribution of nano-SiO<sub>2</sub> particle sizes before and after modification.</p>
Full article ">Figure 8
<p>Scanning electron micrograph of nano-SiO<sub>2</sub> before and after modification: (<b>a</b>) nano-SiO<sub>2</sub> nanoparticles; (<b>b</b>) modified nano-SiO<sub>2</sub>.</p>
Full article ">Figure 9
<p>Thermo gravimetric analysis diagram.</p>
Full article ">Figure 10
<p>The relationship between the average count rate obtained from 90 plus PLAS and the number of suspended silica particles in Milli-Q water.</p>
Full article ">Figure 11
<p>Images of CO<sub>2</sub> nanobubbles obtained from NTA at different times: (<b>a</b>) 10 min; (<b>b</b>) 20 min.</p>
Full article ">Figure 12
<p>Effect of different surfactant types on foaming performance of CO<sub>2</sub> nanobubble system.</p>
Full article ">Figure 13
<p>Effect of different temperatures on the foaming performance of CO<sub>2</sub> nanobubble system.</p>
Full article ">Figure 14
<p>Effect of different temperatures on size of CO<sub>2</sub> nanobubble system.</p>
Full article ">Figure 15
<p>Effect of oil content on the foaming performance of CO<sub>2</sub> nanobubble system.</p>
Full article ">Figure 16
<p>Effect of oil content on size of CO<sub>2</sub> nanobubble system.</p>
Full article ">Figure 17
<p>Contact angle variation with time for CO<sub>2</sub> nanobubble systems.</p>
Full article ">Figure 18
<p>Interfacial tension of CO<sub>2</sub> nanobubble systems.</p>
Full article ">Figure 19
<p>Variation in pressure and gas–oil ratio in different injection systems: (<b>a</b>) Variation in pressure; (<b>b</b>) Variation in gas–oil ratio.</p>
Full article ">Figure 20
<p>Variation in recovery rates for different injection systems.</p>
Full article ">
19 pages, 5872 KiB  
Article
Diclofenac Degradation in Aqueous Solution Using Electron Beam Irradiation and Combined with Nanobubbling
by Yongxia Sun, Joana Madureira, Gonçalo C. Justino, Sandra Cabo Verde, Dagmara Chmielewska-Śmietanko, Marcin Sudlitz, Sylwester Bulka, Ewelina Chajduk, Andrzej Mróz, Shizong Wang and Jianlong Wang
Appl. Sci. 2024, 14(14), 6028; https://doi.org/10.3390/app14146028 - 10 Jul 2024
Viewed by 553
Abstract
Diclofenac (DCF) degradation in aqueous solution under electron beam (EB) irradiation after nanobubbling treatment was studied and compared with treatments using nanobubbling or EB irradiation alone. It was found that the removal efficiency of DCF increased by increasing the adsorbed dose, and it [...] Read more.
Diclofenac (DCF) degradation in aqueous solution under electron beam (EB) irradiation after nanobubbling treatment was studied and compared with treatments using nanobubbling or EB irradiation alone. It was found that the removal efficiency of DCF increased by increasing the adsorbed dose, and it depended on the initial concentration of DCF in solution, being higher for the lower concentrations. Furthermore, when using the nanobubbling treatment alone, about 16% of the DCF was removed from the aqueous solution due to the OH radicals generated during the process. On the other hand, using EB treatment at 0.5 kGy, the degradation of DCF increased from 36% to 51% when adding a nanobubbling pretreatment before the EB radiation. At higher doses (5 kGy), the degradation of DCF was 96% using EB radiation and 99% using nanobubbling before EB radiation, indicating that the nanobubbling effect was not synergistic. With an increase in the adsorbed doses, EB radiation seemed to play a more important role on the degradation of DCF, probably due to the reactive species generated. Moreover, the solutions treated with nanobubbling and EB radiation presented higher COD values and radiolytic by-products with aromatic rings with chlorine. This work can support the development of innovative strategies to treat municipal wastewaters using ionizing radiation technologies. Full article
(This article belongs to the Special Issue Application of Radiation in Wastewater Treatment)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of experimental setup for diclofenac solution nanobubbling and irradiation.</p>
Full article ">Figure 2
<p>DCF degradation in aqueous solution under EB irradiation.</p>
Full article ">Figure 3
<p>Concentration of DCF and dissolved O<sub>2</sub> vs. nanobubbling time ([DCF]<sub>0</sub> = 10 mg/L).</p>
Full article ">Figure 4
<p>DCF removal efficiency in aqueous solution under electron-beam irradiation with (Nano + EB) and without nanobubbling pretreatment (EB) ([DCF]<sub>0</sub> = 125 mg/L).</p>
Full article ">Figure 5
<p>Variation in Cl<sup>−</sup> concentration with the absorbed dose for DCF degradation under electron-beam irradiation with (Nano + EB) and without nanobubble pretreatment (EB) ([DCF]<sub>0</sub> = 125 mg/L).</p>
Full article ">Figure 6
<p>Variation in N concentration with the adsorbed dose for DCF degradation under electron-beam irradiation with (Nano + EB) and without nanobubble pretreatment (EB) ([DCF]<sub>0</sub> = 125 mg/L).</p>
Full article ">Figure 7
<p>Variation in inorganic carbon (IC) concentration with the adsorbed dose under electron-beam irradiation with (Nano + EB) and without nano-bubble pretreatment (EB) ([DCF]<sub>0</sub> = 125 mg/L).</p>
Full article ">Figure 8
<p>Total carbon (TC) variation with the absorbed dose under electron-beam irradiation with (Nano + EB) and without nano-bubble pretreatment (EB) ([DCF]<sub>0</sub> = 125 mg/L).</p>
Full article ">Figure 9
<p>Possible bond cleavage of DCF under EB irradiation.</p>
Full article ">Figure 10
<p>DCF degradation in aqueous solution in flow system under nanobubbling (Nano), electron-beam radiation (EB), and EB with nanobubbling (Nano + EB) ([DCF]<sub>0</sub> = 125 mg/L).</p>
Full article ">Figure 11
<p>Chemical oxygen demand (COD) of DCF aqueous solutions under electron-beam radiation (EB) and EB with nanobubbling (Nano + EB) pretreatment processes in batch and flow systems ([DCF]<sub>0</sub> = 125 mg/L).</p>
Full article ">Figure 12
<p>Chromatographic profile of DCF solution under electron-beam radiation after nanobubbling process in flow system, recorded at 280 nm.</p>
Full article ">Figure 13
<p>Identified products and proposed mechanisms of degradation of DCF by combination treatment of nanobubbling and electron-beam irradiation in a flow system.</p>
Full article ">Figure A1
<p>Isotope profiles of precursor ions listed in the main text (<a href="#applsci-14-06028-t002" class="html-table">Table 2</a>). Representative experimental and calculated (insets) isotope profiles of ions at <span class="html-italic">m</span>/<span class="html-italic">z</span> 296 (<b>A</b>), 106 (<b>B</b>), 134 (<b>C</b>), 260 (<b>D</b>), 312 (<b>E</b>), 152 (<b>F</b>), 294 (<b>G</b>) and 126 (<b>H</b>) are shown; the insets show the calculated profiles for each protonated precursor ion.</p>
Full article ">Figure A1 Cont.
<p>Isotope profiles of precursor ions listed in the main text (<a href="#applsci-14-06028-t002" class="html-table">Table 2</a>). Representative experimental and calculated (insets) isotope profiles of ions at <span class="html-italic">m</span>/<span class="html-italic">z</span> 296 (<b>A</b>), 106 (<b>B</b>), 134 (<b>C</b>), 260 (<b>D</b>), 312 (<b>E</b>), 152 (<b>F</b>), 294 (<b>G</b>) and 126 (<b>H</b>) are shown; the insets show the calculated profiles for each protonated precursor ion.</p>
Full article ">
24 pages, 918 KiB  
Review
Systematic Review of Poultry Slaughterhouse Wastewater Treatment: Unveiling the Potential of Nanobubble Technology
by Ephraim Kaskote, Moses Basitere, Vusi Vincent Mshayisa and Marshall Sheerene Sheldon
Water 2024, 16(13), 1933; https://doi.org/10.3390/w16131933 - 8 Jul 2024
Viewed by 961
Abstract
Aeration is crucial for the biological decomposition of organic compounds in wastewater treatment. However, it is a highly energy-intensive process in traditional activated sludge systems, accounting for 50% to 75% of a plant’s electricity consumption and making it a major cost driver for [...] Read more.
Aeration is crucial for the biological decomposition of organic compounds in wastewater treatment. However, it is a highly energy-intensive process in traditional activated sludge systems, accounting for 50% to 75% of a plant’s electricity consumption and making it a major cost driver for wastewater treatment plants. Nanobubbles (NBs), characterized by their tiny size with diameters less than 200 nm, have emerged as a potential alternative to the low efficiency of aeration and high sludge production in aeration systems. NBs proved effective in removing COD and other pollutants from wastewater. For example, when applied in flotation, aeration, and advanced oxidation, NBs achieved up to 95%, 85%, and 92.5% COD removal, respectively. Considering the recent advancements in wastewater treatment, a compelling need arises for a thorough investigation of the effectiveness and mechanisms of nanobubbles in this field. This systematic review summarizes recent advancements in understanding nanobubbles (NBs) and their unique properties that enhance physical, chemical, and biological water and wastewater treatment processes. Moreover, this study reviews various methods for generating NBs and provides an in-depth review of their applications in wastewater treatment, with a particular focus on poultry slaughterhouse wastewater (PSW) treatment. Full article
(This article belongs to the Section Wastewater Treatment and Reuse)
Show Figures

Figure 1

Figure 1
<p>Flowchart of study selection based on PRISMA.</p>
Full article ">Figure 2
<p>Co-occurrence analysis of the authors’ keywords.</p>
Full article ">
22 pages, 5140 KiB  
Article
Assessing the Role of Air Nanobubble-Saturated Water in Enhancing Soil Moisture, Nutrient Retention, and Plant Growth
by Yeganeh Arablousabet and Arvydas Povilaitis
Sustainability 2024, 16(13), 5727; https://doi.org/10.3390/su16135727 - 4 Jul 2024
Viewed by 881
Abstract
Nanobubble-saturated water (NBSW) has received significant attention in water management in recent years. Therefore, three parallel experiments (E1, E2, and E3) were conducted on two silty loam soils (one with 12.11% higher clay) and sandy loam soil, with additional biochar amendments in each [...] Read more.
Nanobubble-saturated water (NBSW) has received significant attention in water management in recent years. Therefore, three parallel experiments (E1, E2, and E3) were conducted on two silty loam soils (one with 12.11% higher clay) and sandy loam soil, with additional biochar amendments in each soil type, to assess air NBSW’s impact on soil moisture, nutrient retention, and plant growth. The results revealed increased soil moisture retention in the sandy loam and silty loam soils with a lower clay content. It reduced the K+ input compared to conventional watering without highly affecting the amount of leached-out substances. Biochar amendment significantly reduced the TDS losses from silty loam with a higher clay content and reduced the leaching of NO3, Ca2+, and K+ from sandy loam soil. Air NBSW enhanced the stomatal conductance in California pepper plants in silty loam and sandy loam soils but had no effect on silty loam with a higher clay content. A decrease in chlorophyll concentrations and stomatal conductance was observed when air NBSW was combined with biochar in sandy loam soil. The study highlighted that air NBSW alone does not significantly affect water and nutrient retention or key plant parameters. However, its combination with biochar can enhance agricultural water management and sustainability by increasing soil moisture retention and reducing nutrient leaching. Full article
(This article belongs to the Section Pollution Prevention, Mitigation and Sustainability)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Test samples arranged for experimentation. (<b>b</b>) Plant development stage. (<b>c</b>) The experimental setup for leached-out water collection.</p>
Full article ">Figure 2
<p>Soil moisture dynamics, water input, and leaching-out in E1 scenarios.</p>
Full article ">Figure 3
<p>Soil moisture dynamics, water input, and leaching-out in E2 scenarios.</p>
Full article ">Figure 4
<p>Soil moisture dynamics, water input, and leaching-out in E3 scenarios.</p>
Full article ">Figure 5
<p>Comparison of various substances in conventional water and NBSW used for watering (Boxplots show the interquartile range from Q1 to Q3 quartiles, with whiskers extending to the smallest and largest values. The means are represented by horizontal lines within the boxes.).</p>
Full article ">Figure 6
<p>Comparative substance losses from the soil in experiments E1, E2, and E3.</p>
Full article ">Figure 7
<p>Comparison of chlorophyll concentration in California pepper plants in scenarios E1 (<b>a</b>), E2 (<b>b</b>), and E3 (<b>c</b>) (Each boxplot displays the interquartile range from the first quartile (Q1) to the third quartile (Q3), with whiskers extending to the smallest and largest values. The means are indicated by crosses and medians by horizontal lines within the boxes.).</p>
Full article ">Figure 8
<p>Soil moisture–total chl. relationship in experiments E1 (<b>a</b>), E2 (<b>b</b>), and E3 (<b>c</b>).</p>
Full article ">Figure 9
<p>Comparison of stomatal conductance in California pepper plants across various scenarios, E1 (<b>a</b>), E2 (<b>b</b>), and E3 (<b>c</b>). (Each boxplot displays the interquartile range from the first quartile (Q1) to the third quartile (Q3), with whiskers extending to the smallest and largest values. The means are indicated by crosses and medians by horizontal lines within the boxes.)</p>
Full article ">Figure 10
<p>Soil moisture–stomatal conductance relationship in experiments E1 (<b>a</b>), E2 (<b>b</b>), and E3 (<b>c</b>).</p>
Full article ">
16 pages, 2804 KiB  
Article
Insight into the Effect of Nanobubbles on Fine Muscovite Powder Flotation in Different Dodecylamine Concentrations and Stirring Intensities: Kinetics and Mechanism
by Xinyu Zhang, Liuyi Ren, Shenxu Bao, Yimin Zhang, Guohao Chen and Bo Chen
Minerals 2024, 14(7), 694; https://doi.org/10.3390/min14070694 - 3 Jul 2024
Viewed by 822
Abstract
Flotation-introduced nanobubbles were expected to be an efficient and economical method to recover fine muscovite. This study aimed to explore the mechanism of the change appearing in flotation after introducing nanobubbles through micro-flotation, particle vision and measurement, flotation kinetics, and induction time measurement. [...] Read more.
Flotation-introduced nanobubbles were expected to be an efficient and economical method to recover fine muscovite. This study aimed to explore the mechanism of the change appearing in flotation after introducing nanobubbles through micro-flotation, particle vision and measurement, flotation kinetics, and induction time measurement. The results of micro-flotation, which respectively feed muscovite or muscovite pretreated with nanobubbles in different concentrations of dodecylamine (DDA), were fitted with four flotation kinetic models using Origin. Different methods were used to examine how the introduction of nanobubbles affected the flotation process. The results showed that nanobubbles improved both the flotation rate and recovery of muscovite. Nanobubbles played different roles in different stirring intensities. At low stirring intensity, nanobubbles did not perform well. In suitable stirring intensity, nanobubbles helped particles aggregate and improved the collision probability between bubbles and minerals. However, at high stirring intensity, shear forces caused by ultra-high fluid velocities could disrupt particle aggregation. Full article
(This article belongs to the Special Issue Advances on Fine Particles and Bubbles Flotation, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Particle size distribution curve of the final sample.</p>
Full article ">Figure 2
<p>Schematic diagram of nanobubble-generating device.</p>
Full article ">Figure 3
<p>Bubble size distribution.</p>
Full article ">Figure 4
<p>(<b>a</b>) XFGC hanging trough flotation machine (Jilin Prospecting Machinery Factory); (<b>b</b>) Flow chart of micro-flotation; (<b>c</b>) Flow chart of batches scraping.</p>
Full article ">Figure 5
<p>Bubble-particle attachment picture.</p>
Full article ">Figure 6
<p>Recovery-time profile in different DDA concentrations with and without NBs (DDA: 5 mg/L, 10 mg/L, 15 mg/L, 20 mg/L); fitting results in different DDA concentrations with and without NBs (DDA: 5 mg/L, 10 mg/L, 15 mg/L, 20 mg/L).</p>
Full article ">Figure 7
<p>Recovery-time profile in different stirring intensity with and without NBs (1000 r/min, 1500 r/min, 2000 r/min, 2500 r/min); fitting results in different stirring intensity with and without NBs (1000 r/min, 1500 r/min, 2000 r/min, 2500 r/min).</p>
Full article ">Figure 8
<p>Images of bubble and particle behaviors at different stirring intensities with and without NBs.</p>
Full article ">Figure 9
<p>Induction time test results of muscovite flakes (pre-muscovite; muscovite flakes soaked in DDA solution for 30 min).</p>
Full article ">
13 pages, 1181 KiB  
Review
Modifications of Nanobubble Therapy for Cancer Treatment
by Katarzyna M. Terlikowska, Bozena Dobrzycka and Slawomir J. Terlikowski
Int. J. Mol. Sci. 2024, 25(13), 7292; https://doi.org/10.3390/ijms25137292 - 2 Jul 2024
Viewed by 969
Abstract
Cancer development is related to genetic mutations in primary cells, where 5–10% of all cancers are derived from acquired genetic defects, most of which are a consequence of the environment and lifestyle. As it turns out, over half of cancer deaths are due [...] Read more.
Cancer development is related to genetic mutations in primary cells, where 5–10% of all cancers are derived from acquired genetic defects, most of which are a consequence of the environment and lifestyle. As it turns out, over half of cancer deaths are due to the generation of drug resistance. The local delivery of chemotherapeutic drugs may reduce their toxicity by increasing their therapeutic dose at targeted sites and by decreasing the plasma levels of circulating drugs. Nanobubbles have attracted much attention as an effective drug distribution system due to their non-invasiveness and targetability. This review aims to present the characteristics of nanobubble systems and their efficacy within the biomedical field with special emphasis on cancer treatment. In vivo and in vitro studies on cancer confirm nanobubbles’ ability and good blood capillary perfusion; however, there is a need to define their safety and side effects in clinical trials. Full article
Show Figures

Figure 1

Figure 1
<p>Low-intensity ultrasound-mediated nanobubble cancer therapy.</p>
Full article ">Figure 2
<p>Nanobubble architecture for biomedicine.</p>
Full article ">Figure 3
<p>Modifications of nanobubbles.</p>
Full article ">
21 pages, 357 KiB  
Review
Use of Nanobubbles to Improve Mass Transfer in Bioprocesses
by Javier Silva, Laura Arias-Torres, Carlos Carlesi and Germán Aroca
Processes 2024, 12(6), 1227; https://doi.org/10.3390/pr12061227 - 15 Jun 2024
Viewed by 972
Abstract
Nanobubble technology has emerged as a transformative approach in bioprocessing, significantly enhancing mass-transfer efficiency for effective microbial activity. Characterized by their nanometric size and high internal pressure, nanobubbles possess distinct properties such as prolonged stability and minimal rise velocities, allowing them to remain [...] Read more.
Nanobubble technology has emerged as a transformative approach in bioprocessing, significantly enhancing mass-transfer efficiency for effective microbial activity. Characterized by their nanometric size and high internal pressure, nanobubbles possess distinct properties such as prolonged stability and minimal rise velocities, allowing them to remain suspended in liquid media for extended periods. These features are particularly beneficial in bioprocesses involving aerobic strains, where they help overcome common obstacles, such as increased culture viscosity and diffusion limitations, that traditionally impede efficient mass transfer. For instance, in an experimental setup, nanobubble aeration achieved 10% higher soluble chemical oxygen demand (sCOD) removal compared to traditional aeration methods. Additionally, nanobubble-aerated systems demonstrated a 55.03% increase in caproic acid concentration when supplemented with air nanobubble water, reaching up to 15.10 g/L. These results underscore the potential of nanobubble technology for optimizing bioprocess efficiency and sustainability. This review delineates the important role of the mass-transfer coefficient (kL) in evaluating these interactions and underscores the significance of nanobubbles in improving bioprocess efficiency. The integration of nanobubble technology in bioprocessing not only improves gas exchange and substrate utilization but also bolsters microbial growth and metabolic performance. The potential of nanobubble technology to improve the mass-transfer efficiency in biotechnological applications is supported by emerging research. However, to fully leverage these benefits, it is essential to conduct further empirical studies to specifically assess their impacts on bioprocess efficacy and scalability. Such research will provide the necessary data to validate the practical applications of nanobubbles and identify any limitations that need to be addressed in industrial settings. Full article
(This article belongs to the Special Issue Micro–Nano Bubble Technology and Its Applications)
17 pages, 5159 KiB  
Article
Increased Absorption of Thyroxine in a Murine Model of Hypothyroidism Using Water/CO2 Nanobubbles
by Maria Cecilia Opazo, Osvaldo Yañez, Valeria Márquez-Miranda, Johana Santos, Maximiliano Rojas, Ingrid Araya-Durán, Daniel Aguayo, Matías Leal, Yorley Duarte, Jorge Kohanoff and Fernando D. González-Nilo
Int. J. Mol. Sci. 2024, 25(11), 5827; https://doi.org/10.3390/ijms25115827 - 27 May 2024
Viewed by 635
Abstract
Thyroxine (T4) is a drug extensively utilized for the treatment of hypothyroidism. However, the oral absorption of T4 presents certain limitations. This research investigates the efficacy of CO2 nanobubbles in water as a potential oral carrier for T4 administration to C57BL/6 hypothyroid [...] Read more.
Thyroxine (T4) is a drug extensively utilized for the treatment of hypothyroidism. However, the oral absorption of T4 presents certain limitations. This research investigates the efficacy of CO2 nanobubbles in water as a potential oral carrier for T4 administration to C57BL/6 hypothyroid mice. Following 18 h of fasting, the formulation was administered to the mice, demonstrating that the combination of CO2 nanobubbles and T4 enhanced the drug’s absorption in blood serum by approximately 40%. To comprehend this observation at a molecular level, we explored the interaction mechanism through which T4 engages with the CO2 nanobubbles, employing molecular simulations, semi-empirical quantum mechanics, and PMF calculations. Our simulations revealed a high affinity of T4 for the water–gas interface, driven by additive interactions between the hydrophobic region of T4 and the gas phase and electrostatic interactions of the polar groups of T4 with water at the water–gas interface. Concurrently, we observed that at the water–gas interface, the cluster of T4 formed in the water region disassembles, contributing to the drug’s bioavailability. Furthermore, we examined how the gas within the nanobubbles aids in facilitating the drug’s translocation through cell membranes. This research contributes to a deeper understanding of the role of CO2 nanobubbles in drug absorption and subsequent release into the bloodstream. The findings suggest that utilizing CO2 nanobubbles could enhance T4 bioavailability and cell permeability, leading to more efficient transport into cells. Additional research opens the possibility of employing lower concentrations of this class of drugs, thereby potentially reducing the associated side effects due to poor absorption. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Induction of hypothyroidism in C57BL/6 adult mice. C57BL/6 mice were subjected to a methimazole treatment to induce a decrease in T4 levels. As observed in (<b>A</b>), T4 levels were significantly reduced in the methimazole-treated mice (Hypo). (<b>B</b>) Increased weight was observed in the Hypo group during the second week of treatment (Hypo-W2); this observation is consistent with what has been observed in hypothyroid patients. Thyroid-stimulating hormone (TSH) levels were also evaluated. (<b>C</b>) An increase in plasmatic levels is observed, consistent with a hypothyroid phenotype, but no significant differences were observed. Statistics were carried out by Student’s <span class="html-italic">t</span>-test or one-way ANOVA test and Tukey’s post-test; ns: non-significant * <span class="html-italic">p</span> &gt; 0.05; <span class="html-italic">N</span> = 3 mice per group.</p>
Full article ">Figure 2
<p>Administration of T4 combined with CO<sub>2</sub> nanobubbles increases T4 levels in mice plasma. C57BL/6 mice subjected (Hypo) or not (Control) to a methimazole treatment were treated with CO<sub>2</sub> nanobubbles plus T4 (Hypo + NBH<sub>2</sub>O + T4) or normal water with T4 (Hypo + H<sub>2</sub>O + T4). As observed in (<b>A</b>), plasma T4 levels were significantly increased in mice treated with the combination of T4 and CO<sub>2</sub> nanobubbles (Hypo + NBH<sub>2</sub>O + T4). (<b>B</b>) A decrease in weight was observed in the NB+T4-treated group as expected for a hypothyroid phenotype recovery. Thyroid-stimulating hormone (TSH) levels were also evaluated. (<b>C</b>) The NB + T4 group presented a decrease in plasmatic levels, but no significant differences were observed. Statistics were carried out by one-way ANOVA test and Tukey’s post-test; * <span class="html-italic">p</span> &gt; 0.05; <span class="html-italic">N</span> = 3 mice per group.</p>
Full article ">Figure 3
<p>T4 molecules interacting with the water/CO<sub>2</sub> interface. Initial (<b>A</b>) and final (<b>B</b>) snapshot (100 ns) depicting the behavior of a T4 (zwitterionic) cluster in water (top) and how the cluster disassembles at the water/CO<sub>2</sub> interface (<b>B</b>). Water molecules appear in cyan, while CO<sub>2</sub> appears in red. As expected, several CO<sub>2</sub> molecules escape to the water interface.</p>
Full article ">Figure 4
<p>Interaction energy plot. (<b>A</b>) T4zw and (<b>B</b>) T4 molecules interacting in different phases. CO<sub>2</sub> in black liquorice representation and water in light blue liquorice representation. (<b>C</b>) Chemical structure depiction. Zwitterionic T4 and canonical T4 molecules, highlighted to the C<sub>α</sub>.</p>
Full article ">Figure 5
<p>Diffusion of a CO<sub>2</sub> nanobubble with cell membrane. (<b>A</b>) CO<sub>2</sub> nanobubble placed above the cell membrane at the beginning of the simulation, depicted in red for oxygen atoms and cyan for carbons. Membrane is depicted in grey. (<b>B</b>) Snapshot of the last frame of the simulation, showing that CO<sub>2</sub> molecules from the nanobubbles tend to diffuse towards the center of the membrane.</p>
Full article ">Figure 6
<p>(<b>A</b>) Potential of Mean Force of the translocation of a T4 (zwitterionic) molecule through pure POPC and POPC-CO<sub>2</sub> membranes. (<b>B</b>) Snapshots of the conformations of the T4 molecule in different stages of the reaction coordinate: (1) the molecule starts in the water (membrane COM distance = 30 Å), (2) the molecule falls into an energy well in the water–lipid interface (membrane COM distance ~15 Å), (3) T4 is in the middle of the membrane facing a high free energy barrier (membrane COM distance = 0 Å), and T4 falls again into an energy well (membrane COM distance ~−15 Å). Lipid heads appear depicted in orange, in cyan the T4 molecule and in green some CO<sub>2</sub> molecules around T4.</p>
Full article ">
19 pages, 5401 KiB  
Article
Enhancing Root Distribution, Nitrogen, and Water Use Efficiency in Greenhouse Tomato Crops Using Nanobubbles
by Fernando del Moral Torres, Rafael Hernández Maqueda and David Erik Meca Abad
Horticulturae 2024, 10(5), 463; https://doi.org/10.3390/horticulturae10050463 - 1 May 2024
Viewed by 1382
Abstract
The aim of this work was to determine the effect of saturating the irrigation solution with air (MNBA) or oxygen nanobubbles (MNBO) on relevant agronomic, productive, and postharvest parameters of tomato crops (Solanum lycopersicum L.) in greenhouses. As a control, conventional management [...] Read more.
The aim of this work was to determine the effect of saturating the irrigation solution with air (MNBA) or oxygen nanobubbles (MNBO) on relevant agronomic, productive, and postharvest parameters of tomato crops (Solanum lycopersicum L.) in greenhouses. As a control, conventional management was established, without nanobubbles, under the best possible agronomic conditions used in commercial greenhouses in southeastern Spain. No significant differences were found in the soil properties analysed or in the ionic concentration of the pore water extracted with Rhizon probes. Both MNBA and MNBO modified the root distribution and improved the N uptake efficiency and field water uptake efficiency compared to the control. MNBA had the highest harvest index. The total or marketable production was not affected, although it did increase the overall size of the fruit and the earliness with which they were produced compared to the control. MNBA significantly decreased titratable acidity and soluble solids content compared to the control in the last harvests. Both nanobubble treatments improved postharvest storage under room-temperature (20–25 °C) conditions. Full article
(This article belongs to the Section Vegetable Production Systems)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic layout of the nanobubble generation system and distribution of fertigation solutions for the different treatments. EV: solenoid valve; P: pump; NB: nanobubble.</p>
Full article ">Figure 2
<p><b>Left column</b>: evolution over time of pore water N-NO<sub>3</sub><sup>−</sup> obtained via Rhizon MOM samplers and sap N-NO<sub>3</sub><sup>−</sup> concentrations in each treatment. <b>Right column</b>: evolution over time of K<sup>+</sup> concentrations. Data show means ± MSE. DAT stands for the date after transplanting.</p>
Full article ">Figure 3
<p>Root length density per soil layer sampled in the two sampling positions (P1, 10 cm from the plant; P2 30 cm from the plant). C, control; MNBA: micro-nanobubbles of air; MNBO micro-nanobubbles of oxygen. Different lowercase letters in each graph show significant differences among treatments (<span class="html-italic">p</span> &lt; 0.05). Data show means ± MSE.</p>
Full article ">Figure 4
<p><b>Left:</b> relative root length distribution by soil layer depth. <b>Right:</b> relative root length distribution by sampling position.</p>
Full article ">Figure 5
<p><b>Left:</b> mean above-ground (leaves and stems) fresh and dry matter weights for each treatment. Different capital letters indicate differences (<span class="html-italic">p</span> &lt; 0.05) in fresh weight between treatments. Different lowercase letters indicate significant differences in dry weight between treatments. <b>Right:</b> evolution of cumulative above-ground (leaves and stems) dry weight over the growing season for each treatment. DAT stands for the date after transplanting. Data show means ± MSE.</p>
Full article ">Figure 6
<p>Cumulative commercial production expressed by category for each treatment. Bottom right, average fruit weight at each harvest for each treatment. DAT: days after transplanting.</p>
Full article ">Figure 7
<p>Average percentage (n = 10) of postharvest weight loss during conservation at room temperature (<b>left</b>) and in the cooling chamber (<b>right</b>). Data show means ± MSE.</p>
Full article ">
14 pages, 9071 KiB  
Article
Synergistic Effect of He for the Fabrication of Ne and Ar Gas-Charged Silicon Thin Films as Solid Targets for Spectroscopic Studies
by Asunción Fernández, Vanda Godinho, José Ávila, M. Carmen Jiménez de Haro, Dirk Hufschmidt, Jennifer López-Viejobueno, G. Eduardo Almanza-Vergara, F. Javier Ferrer, Julien L. Colaux, Stephane Lucas and M. Carmen Asensio
Nanomaterials 2024, 14(8), 727; https://doi.org/10.3390/nano14080727 - 21 Apr 2024
Viewed by 959
Abstract
Sputtering of silicon in a He magnetron discharge (MS) has been reported as a bottom-up procedure to obtain He-charged silicon films (i.e., He nanobubbles encapsulated in a silicon matrix). The incorporation of heavier noble gases is demonstrated in this work with a synergistic [...] Read more.
Sputtering of silicon in a He magnetron discharge (MS) has been reported as a bottom-up procedure to obtain He-charged silicon films (i.e., He nanobubbles encapsulated in a silicon matrix). The incorporation of heavier noble gases is demonstrated in this work with a synergistic effect, producing increased Ne and Ar incorporations when using He–Ne and He–Ar gas mixtures in the MS process. Microstructural and chemical characterizations are reported using ion beam analysis (IBA) and scanning and transmission electron microscopies (SEM and TEM). In addition to gas incorporation, He promotes the formation of larger nanobubbles. In the case of Ne, high-resolution X-ray photoelectron and absorption spectroscopies (XPS and XAS) are reported, with remarkable dependence of the Ne 1s photoemission and the Ne K-edge absorption on the nanobubble’s size and composition. The gas (He, Ne and Ar)-charged thin films are proposed as “solid” targets for the characterization of spectroscopic properties of noble gases in a confined state without the need for cryogenics or high-pressure anvils devices. Also, their use as targets for nuclear reaction studies is foreseen. Full article
Show Figures

Figure 1

Figure 1
<p>Cross-section SEM images of samples 1 to 7. Samples 1, 4 and 6 correspond to Si-Ne. Samples 2, 3, 5 and 7 correspond to Si-Ne(He). Notes: (i) Scales are not equal to visualize entire layers maximizing magnification. (ii) For the Si-Ne samples zoom images show columnar microstructure. (iii) Dash lines mark the interface between the silicon substrate and the coating.</p>
Full article ">Figure 2
<p>Cross-section SEM images at high magnification for representative S4 (Ne) and S5 (Ne + He) samples.</p>
Full article ">Figure 3
<p>TEM cross-section images for representative samples: 1, 4 and 6 correspond to Si-Ne; 2, 5 and 7 correspond to Si-Ne(He).</p>
Full article ">Figure 4
<p>Bar diagram plots of atomic content ratio for the Si-Ne and Si-Ne(He) samples. For uncertainty values refer to <a href="#nanomaterials-14-00727-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 5
<p>Proton beam EBS-spectra for representative samples: S1 (Ne) and S2 (Ne + He).</p>
Full article ">Figure 6
<p>Cross-section SEM images for investigated S8 (Ar) and S9 (Ar + He) samples. Notes: (i) Scales are not equal to visualize entire layers maximizing magnification. (ii) Dash lines mark the interface between the silicon substrate and the coating.</p>
Full article ">Figure 7
<p>TEM cross-section images for the S8 (Ar) and S9 (Ar + He)) samples.</p>
Full article ">Figure 8
<p>Bar diagram plots of atomic content ratios for the S8 (Ar) and S9 (Ar + He) samples. For uncertainty values refer to <a href="#nanomaterials-14-00727-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 9
<p>Normalized Ne1s XPS spectra for the as prepared amples S6 (Ne) and S7 (Ne + He). The spectra after successive Ar<sup>+</sup> sputtering for 2 and 10 are also shown. The red line is a reference to visualize peaks shifts.</p>
Full article ">Figure 10
<p>Normalized (<b>a</b>,<b>b</b>) and first derivative (<b>c</b>,<b>d</b>) of Ne 1s XAS spectra for samples S1 (Ne), S2 (Ne + He), S6 (Ne) and S7 (Ne + He).</p>
Full article ">
12 pages, 11291 KiB  
Article
Effect of Bulk Nanobubbles on the Flocculation and Filtration Characteristics of Kaolin Using Cationic Polyacrylamide
by Yihong Li, Guangxi Ma, Muhammad Bilal, Jie Sha and Xiangning Bu
Minerals 2024, 14(4), 405; https://doi.org/10.3390/min14040405 - 15 Apr 2024
Viewed by 724
Abstract
This study investigated the influence of bulk nanobubbles (NBs) on the flocculation and filtration behavior of kaolin suspensions treated with cationic polyacrylamide (CPAM). Traditionally, flocculation relies on bridging mechanisms by polymers like CPAM. The present work examines the possibility of combining NBs with [...] Read more.
This study investigated the influence of bulk nanobubbles (NBs) on the flocculation and filtration behavior of kaolin suspensions treated with cationic polyacrylamide (CPAM). Traditionally, flocculation relies on bridging mechanisms by polymers like CPAM. The present work examines the possibility of combining NBs with CPAM to achieve more efficient kaolin separation. The settling behavior of kaolin suspensions with and without bulk nanobubbles was compared. The results with 2 mL CPAM and 300 s settling time revealed that bulk NBs significantly enhanced flocculation efficiency, with supernatant zone height reductions exceeding 50% compared to CPAM alone, indicating a faster settling rate resulting from bulk NBs. This improvement in the settling rate is attributed to NBs’ ability to reduce inter-particle repulsion (as evidenced by a shift in zeta potential from −20 mV to −10 mV) and bridge kaolin particles, complementing the action of CPAM. Additionally, the study demonstrated that bulk NBs improved dewatering characteristics by lowering the medium resistance and specific cake resistance during filtration. These findings pave the way for the utilization of bulk NBs as a novel and efficient strategy for kaolin separation in mineral processing, potentially leading to reduced processing times and lower operational costs. Full article
(This article belongs to the Section Mineral Processing and Extractive Metallurgy)
Show Figures

Figure 1

Figure 1
<p>Effects of bulk NBs on supernatants’ heights and settling rates of kaolin suspension under different CPAM dosages: (<b>a</b>) 0.5 and 2 mL CPAM; (<b>b</b>) 6, 10, and 12 mL CPAM.</p>
Full article ">Figure 2
<p>Settling rates of kaolin suspension under different CPAM dosages.</p>
Full article ">Figure 3
<p>The floc images under different flocculant dosages in the absence and presence of bulk NBs.</p>
Full article ">Figure 4
<p>Zeta potentials of kaolin particles in the absence and presence of bulk NBs.</p>
Full article ">Figure 5
<p>EDLVO interaction energy calculation results in the absence and presence of bulk NBs.</p>
Full article ">Figure 6
<p>Filter cake moisture contents with different flocculant dosages in the absence and presence of bulk NBs.</p>
Full article ">Figure 7
<p>Images of CPAM solution prepared by lyophilization of the absence (<b>a</b>) and presence (<b>b</b>) of bulk NBs.</p>
Full article ">Figure 8
<p>Plots of t/V versus V data in the absence and presence of bulk NBs under 6 mL CPAM dosage.</p>
Full article ">Figure 9
<p>Comparisons of distribution curves of transverse relaxation time of filter cakes in the absence and presence of bulk NBs under 6 mL CPAM dosage.</p>
Full article ">
21 pages, 5967 KiB  
Article
Studying the Effects of Dissolved Noble Gases and High Hydrostatic Pressure on the Spherical DOPC Bilayer Using Molecular Dynamic Simulations
by Eugeny Pavlyuk, Irena Yungerman, Alice Bliznyuk and Yevgeny Moskovitz
Membranes 2024, 14(4), 89; https://doi.org/10.3390/membranes14040089 - 12 Apr 2024
Viewed by 1267
Abstract
Fine-grained molecular dynamics simulations have been conducted to depict lipid objects enclosed in water and interacting with a series of noble gases dissolved in the medium. The simple point-charge (SPC) water system, featuring a boundary composed of 1,2-Dioleoyl-sn-glycero-3-phosphocholine (DOPC) molecules, maintained stability throughout [...] Read more.
Fine-grained molecular dynamics simulations have been conducted to depict lipid objects enclosed in water and interacting with a series of noble gases dissolved in the medium. The simple point-charge (SPC) water system, featuring a boundary composed of 1,2-Dioleoyl-sn-glycero-3-phosphocholine (DOPC) molecules, maintained stability throughout the simulation under standard conditions. This allowed for the accurate modeling of the effects of hydrostatic pressure at an ambient pressure of 25 bar. The chosen pressure references the 240 m depth of seawater: the horizon frequently used by commercial divers, who comprise the primary patient population of the neurological complication of inert gas narcosis and the consequences of high-pressure neurological syndrome. To quantify and validate the neurological effects of noble gases and discriminate them from high hydrostatic pressure, we reduced the dissolved gas molar concentration to 1.5%, three times smaller than what we previously tested for the planar bilayer (3.5%). The nucleation and growth of xenon, argon and neon nanobubbles proved consistent with the data from the planar bilayer simulations. On the other hand, hyperbaric helium induces only a residual distorting effect on the liposome, with no significant condensed gas fraction observed within the hydrophobic core. The bubbles were distributed over a large volume—both in the bulk solvent and in the lipid phase—thereby causing substantial membrane distortion. This finding serves as evidence of the validity of the multisite distortion hypothesis for the neurological effect of inert gases at high pressure. Full article
(This article belongs to the Section Biological Membrane Dynamics and Computation)
Show Figures

Figure 1

Figure 1
<p>The initial frames of the liposome at the beginning of the molecular dynamics production cycle at t = 0 ns are shown for 6 simulations conducted at 25 bar of ambient pressure. The small droplets of xenon in the bulk solvent are designated by the violet color.</p>
Full article ">Figure 1 Cont.
<p>The initial frames of the liposome at the beginning of the molecular dynamics production cycle at t = 0 ns are shown for 6 simulations conducted at 25 bar of ambient pressure. The small droplets of xenon in the bulk solvent are designated by the violet color.</p>
Full article ">Figure 2
<p>The density (kg·m<sup>−3</sup>) of DOPC lipids (<span class="html-italic">y</span>-axis of the plots) is presented for 6 simulations at 0, 20, 100 and 160 ns. The <span class="html-italic">x</span>-axis is the radius of the liposome (nm) centered at the center of mass (x = 0) comprising DOPC molecules. (<b>A</b>) 1 bar, (<b>B</b>) 25 bar, (<b>C</b>) He 25 bar, (<b>D</b>) Ne 25 bar, (<b>E</b>) Ar 25 bar, and (<b>F</b>) Xe 25 bar.</p>
Full article ">Figure 2 Cont.
<p>The density (kg·m<sup>−3</sup>) of DOPC lipids (<span class="html-italic">y</span>-axis of the plots) is presented for 6 simulations at 0, 20, 100 and 160 ns. The <span class="html-italic">x</span>-axis is the radius of the liposome (nm) centered at the center of mass (x = 0) comprising DOPC molecules. (<b>A</b>) 1 bar, (<b>B</b>) 25 bar, (<b>C</b>) He 25 bar, (<b>D</b>) Ne 25 bar, (<b>E</b>) Ar 25 bar, and (<b>F</b>) Xe 25 bar.</p>
Full article ">Figure 3
<p>The density (kg·m<sup>−3</sup>) of noble gas atoms (<span class="html-italic">y</span>-axis of the plots) is presented for 4 simulations at 0, 20, 100 and 160 ns. The <span class="html-italic">x</span>-axis is the radius of the liposome (nm) centered at the center of mass (x = 0) comprising DOPC molecules. (<b>A</b>) He 25 bar, (<b>B</b>) Ne 25 bar, (<b>C</b>) Ar 25 bar, and (<b>D</b>) Xe 25 bar.</p>
Full article ">Figure 4
<p>The dynamics of gas percolations in time. The number of effective contacts is shown between gas atoms and the bulk solvent; DOPC atoms in the outer and inner monolayers of the liposome; the interlayer gap and confined solvent molecules. The contacts’ sampling frequency—1ns. A contact cutoff of 1.0 nm has been applied. (<b>A</b>) He 25 bar, (<b>B</b>) Ne 25 bar, (<b>C</b>) Ar 25 bar, and (<b>D</b>) Xe 25 bar.</p>
Full article ">Figure 5
<p>The representative last frames of the liposome trajectory for 6 simulations. (<b>A</b>) 1 bar, (<b>B</b>) 25 bar, (<b>C</b>) He 25 bar, (<b>D</b>) Ne 25 bar, (<b>E</b>) Ar 25 bar, and (<b>F</b>) Xe 25 bar. The xenon free-floating bubble in periodic boundary conditions is presented in a violet color.</p>
Full article ">Figure 6
<p>The representative last frames of MD trajectory: (<b>A</b>) helium 25 bar; (<b>B</b>) neon 25 bar; and (<b>C</b>) argon 25 bar. The continuous gas fractions are depicted by the <sup>3</sup>V server using probe radius 6.0 Å [<a href="#B30-membranes-14-00089" class="html-bibr">30</a>]. The color spectrum represents the three-dimensionality of the molecular volumes.</p>
Full article ">Figure 7
<p>External monosphere DOPC acyl chains Sn1 and Sn2 prochiral order parameters: (<b>A</b>)—Sn1: Pro-S; (<b>B</b>)—Sn1: Pro-R; (<b>C</b>)—Sn2: Pro-S; and (<b>D</b>)—Sn2: Pro-R.</p>
Full article ">Figure 8
<p>Internal monosphere DOPC acyl chains Sn1 and Sn2 prochiral order parameters: (<b>A</b>)—Sn1: Pro-S; (<b>B</b>)—Sn1: Pro-R; (<b>C</b>)—Sn2: Pro-S; and (<b>D</b>)—Sn2: Pro-R.</p>
Full article ">
30 pages, 4807 KiB  
Review
Generation Mechanism of Hydroxyl Free Radicals in Micro–Nanobubbles Water and Its Prospect in Drinking Water
by Tianzhi Wang, Ci Yang, Peizhe Sun, Mingna Wang, Fawei Lin, Manuel Fiallos and Soon-Thiam Khu
Processes 2024, 12(4), 683; https://doi.org/10.3390/pr12040683 - 28 Mar 2024
Cited by 2 | Viewed by 1479
Abstract
Micro–nanobubbles (MNBs) can generate ·OH in situ, which provides a new idea for the safe and efficient removal of pollutants in water supply systems. However, due to the difficulty in obtaining stable MNBs, the generation efficiency of ·OH is low, and the removal [...] Read more.
Micro–nanobubbles (MNBs) can generate ·OH in situ, which provides a new idea for the safe and efficient removal of pollutants in water supply systems. However, due to the difficulty in obtaining stable MNBs, the generation efficiency of ·OH is low, and the removal efficiency of pollutants cannot be guaranteed. This paper reviews the application research of MNB technology in water security from three aspects: the generation process of MNBs in water, the generation rule of ·OH during MNB collapse, and the control mechanisms of MNBs on pollutants and biofilms. We found that MNB generation methods are divided into chemical and mechanical (about 10 kinds) categories, and the instability of the bubble size restricts the application of MNB technology. The generation of ·OH by MNBs is affected by the pH, gas source, bubble size, temperature, and external stimulation. And the pH and external stimulus have more influence on ·OH generation in situ than the other factors. Adjusting the pH to alkaline or acidic conditions and selecting ozone or oxygen as the gas source can promote ·OH generation. MNB collapse also releases a large amount of energy, during which the temperature and pressure can reach 3000 K and 5 Gpa, respectively, making it efficient to remove ≈90% of pollutants (i.e., trichloroethylene, benzene, and chlorobenzene). The biofilm can also be removed by physical, chemical, and thermal effects. MNB technology also has great application potential in drinking water, which can be applied to improve water quality, optimize household water purifiers, and enhance the taste of bottled water. Under the premise of safety, after letting people of different ages taste water samples, we found that compared with ordinary drinking water, 85.7% of people think MNB water is softer, and 73.3% of people think MNB water is sweeter. This further proves that MNB water has a great prospect in drinking water applications. This review provides innovative theoretical support for solving the problem of drinking water safety. Full article
(This article belongs to the Special Issue Micro–Nano Bubble Technology and Its Applications)
Show Figures

Figure 1

Figure 1
<p>Differences among microbubbles, microbubbles, and nanobubbles. (<b>A</b>) The generation process of micro–nanobubbles in water. (<b>B</b>) Mechanism of hydroxyl radical generation by micro–nanobubbles. Note: HTHP, high temperature and high pressure.</p>
Full article ">Figure 2
<p>Relationships among the characteristics of MNBs. (<b>A</b>) Degradation of phenol by MNBW. (<b>B</b>) The relationship between the bubble rising velocity and bubble radius indirectly reflects the relationship between the bubble rising velocity and existence time. (<b>C</b>) The relationship between the bubble zeta potential and bubble radius. (<b>D</b>) The relationship between the DO level and bubble radius indirectly reflects the relationship between the bubble mass transfer efficiency and bubble radius. (Ref. [<a href="#B61-processes-12-00683" class="html-bibr">61</a>], (<b>B</b>): reproduced with permission from [Li Hengzhen et al.], [Water Environment Research]; published by [WILEY], [2014]; Ref. [<a href="#B100-processes-12-00683" class="html-bibr">100</a>], (<b>C</b>): reproduced with permission from [Li Hao], [Jiangsu University]; published by [Jiangsu University], [2020]; Ref. [<a href="#B83-processes-12-00683" class="html-bibr">83</a>], (<b>D</b>): reproduced with permission from [Li Hengzhen et al.], [International Journal of Environmental Research and Public Health]; published by [MDPI], [2014]). Note: Solid line means established relationship, dotted line means relationship unknown. (i) TB means for tert-butanol, ·OH quencher. Experimental conditions: temperature 15 °C, pH 7.3, and initial concentration of phenol 10 mg/L.</p>
Full article ">Figure 3
<p>Influence of different factors on the generation of ·OH by MNBs. (<b>A</b>) Factors affecting the generation of ·OH by MNBs. (<b>B</b>) Effect of pH on ·OH exposure. (<b>C</b>) Effect of gas source type on the generation of free radicals by MNBs. (<b>D</b>) Relationship between active oxygen concentration and bubble size. (<b>E</b>) Relationship between active oxygen concentration and temperature. (Ref. [<a href="#B109-processes-12-00683" class="html-bibr">109</a>], (<b>B</b>): reproduced with permission from [Snigdha Khuntia et al.], [Chemical Engineering Research and Design]; published by [ELSEVIER], [2015]; Ref. [<a href="#B93-processes-12-00683" class="html-bibr">93</a>], (<b>C</b>): reproduced with permission from [Masahiro Kohno et al.], [Journal of Clinical Biochemistry and Nutrition]; published by [The Society for Free Radical Research Japan], [2011]; Ref. [<a href="#B110-processes-12-00683" class="html-bibr">110</a>], (<b>D</b>,<b>E</b>): reproduced with permission from [Yu Xiaobin et al.], [Journal of Chemical Technology &amp; Biotechnology]; published by [WILEY], [2017]).</p>
Full article ">Figure 4
<p>Effect of pH on the degradation of organic pollutants by MNBs. (<b>A</b>) Degradation rates of different pollutants by MNBs. (<b>B</b>) Degradation rate constants of different pollutants by MNBs. (Ref. [<a href="#B117-processes-12-00683" class="html-bibr">117</a>], reproduced with permission from [Wang Xikui et al.], [Ultrasonics sonochemistry]; published by [ELSEVIER], [2008]; Ref. [<a href="#B118-processes-12-00683" class="html-bibr">118</a>], reproduced with permission from [Wang Xikui and Zhang Yong], [Journal of hazardous materials]; published by [Elsevier], [2009]; Ref. [<a href="#B127-processes-12-00683" class="html-bibr">127</a>], reproduced with permission from [Xia Zhiran and Hu Liming], [Water]; published by [MDPI], [2018]; Ref. [<a href="#B130-processes-12-00683" class="html-bibr">130</a>], reproduced with permission from [Abdisa Jabesa and Pallab Ghosh], [Journal of Environmental Management]; published by [ELSEVIER], [2016]; Ref. [<a href="#B132-processes-12-00683" class="html-bibr">132</a>], reproduced with permission from [Li Pan et al.], [Industrial &amp; engineering chemistry research]; published by [ACS], [2009]; Ref. [<a href="#B136-processes-12-00683" class="html-bibr">136</a>], reproduced with permission from [Il-Kyu Kim and Chin-Pao Huang], [Journal of the Chinese Institute of Engineers]; published by [Springer Nature Singapore Pte Ltd.], [2005]).</p>
Full article ">Figure 5
<p>Control mechanism of physical, chemical, and thermal effects of MNBs on biofilm.</p>
Full article ">Figure 6
<p>Application prospects of MNBs in drinking water. (<b>A</b>) Application of MNBs in waterwork processes. (<b>B</b>) Application of MNBs in second water supply tanks. (<b>C</b>) Application of MNBs in household water purifiers. (<b>D</b>) Application of MNBs in high quality bottled water.</p>
Full article ">Figure 7
<p>Survey results of MNB drinking water and ordinary drinking water. (<b>A</b>) MNB drinking water and ordinary drinking water. (<b>B</b>) Investigation results of hardness of MNB drinking water. (<b>C</b>) Investigation results of hardness of ordinary drinking water. (<b>D</b>) Investigation result of sweetness of MNB drinking water. (<b>E</b>) Investigation results of sweetness of ordinary drinking water.</p>
Full article ">
Back to TopTop