[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

Journals

Article Types

Countries / Regions

Search Results (61)

Search Parameters:
Keywords = DLVO theory

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 3180 KiB  
Article
Influence of Particle Size on Flotation Separation of Ilmenite and Forsterite
by Senpeng Zhang, Yaohui Yang, Donghui Wang, Weiping Yan and Weishi Li
Minerals 2024, 14(10), 1041; https://doi.org/10.3390/min14101041 (registering DOI) - 17 Oct 2024
Abstract
In addition to bubble–particle interaction, particle–particle interaction also has a significant influence on mineral flotation. Fine particles that coat the mineral surface prevent direct contact with collectors and/or air bubbles, thereby lowering flotation recovery. Calculating the particle interaction energy can help in evaluating [...] Read more.
In addition to bubble–particle interaction, particle–particle interaction also has a significant influence on mineral flotation. Fine particles that coat the mineral surface prevent direct contact with collectors and/or air bubbles, thereby lowering flotation recovery. Calculating the particle interaction energy can help in evaluating the interaction behavior of particles. In this study, the floatability of coarse ilmenite (−151 + 74 μm) and different particle sizes (−45 + 25, −25 + 19, −19 μm) of forsterite with NaOL as a collector was investigated. The results showed that forsterite sizes of −45 + 25 and −25 + 19 μm had no effect on the ilmenite floatability, whereas −19 μm forsterite significantly reduced ilmenite floatability. A particle size analysis of artificially mixed minerals and a scanning electron microscopy (SEM) analysis of the flotation products showed that heterogeneous aggregation occurred between ilmenite and −19 μm forsterite particles. The extended DLVO (Derjaguin–Landau–Verwey–Overbeek) theory was applied to calculate the interaction energy between mineral particles using data from zeta potential and contact angle measurements. The results showed that the interaction barriers between ilmenite (−151 + 74 μm) and forsterite (−45 + 25, −25 + 19, and −19 μm) were 11.94 × 103 kT, 8.23 × 103 kT and 4.09 × 103 kT, respectively. Additionally, the interaction barrier between forsterite particles smaller than 19 μm was 0.51 × 103 kT. The strength of the barrier decreased as the size of the forsterite decreased. Therefore, fine forsterite particles and aggregated forsterite can easily overcome the energy barrier, coating the ilmenite particle surface. This explains the effect of different forsterite sizes on the floatability of ilmenite and the underlying mechanism of particle interaction. Full article
Show Figures

Figure 1

Figure 1
<p>XRD spectra of samples: (<b>a</b>) ilmenite; (<b>b</b>) forsterite.</p>
Full article ">Figure 2
<p>Effect of solution chemical environment on the floatability of ilmenite and forsterite: (<b>a</b>) NaOL concentration; (<b>b</b>) pH.</p>
Full article ">Figure 3
<p>Effect of different forsterite contents on ilmenite floatability.</p>
Full article ">Figure 4
<p>Effect of pH on zeta potential of ilmenite and forsterite in the presence and absence of NaOL.</p>
Full article ">Figure 5
<p>Effect of fine forsterite on particle size distribution of mixed minerals.</p>
Full article ">Figure 6
<p>SEM images of flotation products: (<b>a</b>) −151 + 74 μm ilmenite and −45 + 25 μm forsterite; (<b>b</b>) −151 + 74 μm ilmenite and −25 + 19 μm forsterite; (<b>c</b>) −151 + 74 μm ilmenite and −19 μm forsterite; (<b>d</b>) partial enlargement of the <a href="#minerals-14-01041-f006" class="html-fig">Figure 6</a>c.</p>
Full article ">Figure 7
<p>DLVO and the E-DLVO interaction energy profiles for ilmenite and forsterite in the presence and absence of NaOL: (<b>a</b>) −151 + 74 μm ilmenite and −45 + 25 μm forsterite; (<b>b</b>) −151 + 74 μm ilmenite and −25 + 19 μm forsterite; (<b>c</b>) −151 + 74 μm ilmenite and −19 μm forsterite; (<b>d</b>) −19 μm forsterite and −19 μm forsterite.</p>
Full article ">
23 pages, 13110 KiB  
Article
Boron Removal in Aqueous Solutions Using Adsorption with Sugarcane Bagasse Biochar and Ammonia Nanobubbles
by Lianying Liao, Hao Chen, Chunlin He, Gjergj Dodbiba and Toyohisa Fujita
Materials 2024, 17(19), 4895; https://doi.org/10.3390/ma17194895 - 6 Oct 2024
Viewed by 623
Abstract
Boron is a naturally occurring trace chemical element. High concentrations of boron in nature can adversely affect biological systems and cause severe pollution to the ecological environment. We examined a method to effectively remove boron ions from water systems using sugarcane bagasse biochar [...] Read more.
Boron is a naturally occurring trace chemical element. High concentrations of boron in nature can adversely affect biological systems and cause severe pollution to the ecological environment. We examined a method to effectively remove boron ions from water systems using sugarcane bagasse biochar from agricultural waste with NH3 nanobubbles (10% NH3 and 90% N2). We studied the effects of the boron solution concentration, pH, and adsorption time on the adsorption of boron by the modified biochar. At the same time, the possibility of using magnesium chloride and NH3 nanobubbles to enhance the adsorption capacity of the biochar was explored. The carbonization temperature of sugarcane bagasse was investigated using thermogravimetric analysis. It was characterized using XRD, SEM, and BET analysis. The boron adsorption results showed that, under alkaline conditions above pH 9, the adsorption capacity of the positively charged modified biochar was improved under the double-layer effect of magnesium ions and NH3 nanobubbles, because the boron existed in the form of negatively charged borate B(OH)4 anion groups. Moreover, cations on the NH3 nanobubble could adsorb the boron. When the NH3 nanobubbles with boron and the modified biochar with boron could coagulate each other, the boron was removed to a significant extent. Extended DLVO theory was adopted to model the interaction between the NH3 nanobubble and modified biochar. The boron adsorption capacity was 36 mg/g at room temperature according to a Langmuir adsorption isotherm. The adsorbed boron was investigated using FT-IR and XPS analysis. The ammonia could be removed using zeolite molecular sieves and heating. Boron in an aqueous solution can be removed via adsorption with modified biochar with NH3 nanobubbles and MgCl2 addition. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Speciation diagram for boron.</p>
Full article ">Figure 2
<p>Preparation method of char A.</p>
Full article ">Figure 3
<p>Preparation method. (<b>a</b>) Schematic procedure for the preparation of nanocomposite Ws–N–La [<a href="#B63-materials-17-04895" class="html-bibr">63</a>]. (<b>b</b>) Preparation method of char B.</p>
Full article ">Figure 4
<p>Preparation method for NH<sub>3</sub> and N<sub>2</sub> nanobubbles, using ultrasonic atomization equipment.</p>
Full article ">Figure 5
<p>TGA and DTG curves of sugarcane bagasse under N<sub>2</sub> air atmosphere at a heating rate of 5 °C min<sup>−1</sup>.</p>
Full article ">Figure 6
<p>XRD patterns of char A and char B.</p>
Full article ">Figure 7
<p>SEM photographs of char A and char B. Char A: magnitude of 500× (<b>a</b>) and magnitude of 2000× (<b>b</b>). Char B: magnitude of 500× (<b>c</b>) and magnitude of 2000× (<b>d</b>).</p>
Full article ">Figure 8
<p>Pore size distribution depending on pore diameter.</p>
Full article ">Figure 9
<p>Nanobubble size distributions and bubble concentrations of NH<sub>3</sub> and N<sub>2</sub> nanobubbles. (<b>a</b>) Size distribution of NH<sub>3</sub> nanobubbles (10% NH<sub>3</sub> and 90% N<sub>2</sub>, pH 11.2), (<b>b</b>) size distribution of N<sub>2</sub> nanobubbles (100% N<sub>2</sub>, pH 10.4), (<b>c</b>) changes in concentrations of NH<sub>3</sub> nanobubbles and N<sub>2</sub> nanobubbles over time.</p>
Full article ">Figure 10
<p>Zeta potential as a function of pH for adsorbents and nanobubbles. Zeta potential in MgCl<sub>2</sub> solutions of different concentrations of char A (<b>a</b>); char B (<b>b</b>). Zeta potential in MgCl<sub>2</sub> solutions of different concentrations of N<sub>2</sub> nanobubbles (<b>c</b>) and NH<sub>3</sub> nanobubbles (<b>d</b>).</p>
Full article ">Figure 11
<p>Total potential energy V<sub>T</sub> as a function of the distance between the NH<sub>3</sub> nanobubbles and char B at pH 9. (<b>a</b>) No addition of MgCl<sub>2</sub>, NH<sub>3</sub> nanobubbles −18 mV and char B −8 mV; (<b>b</b>) addition of 5 ppm MgCl<sub>2</sub> and 200 ppm B, NH<sub>3</sub> nanobubbles −5 mV and char B −13 mV. Here, the NH<sub>3</sub> nanobubble size is 200 nm, and the char B size is 10 mm.</p>
Full article ">Figure 12
<p>Boron adsorption types. (<b>a</b>) No additives, (<b>b</b>) addition of MgCl<sub>2</sub>, (<b>c</b>) addition of MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles.</p>
Full article ">Figure 13
<p>Boron adsorption capacity in different conditions. (<b>a</b>) Boron adsorption in different adsorbents with 5 ppm MgCl<sub>2</sub>; (<b>b</b>) Comparison of N<sub>2</sub> and NH<sub>3</sub> nanobubbles with 5 ppm MgCl<sub>2</sub> for char B adsorbent. (<b>c</b>) Boron adsorption capacity at different temperatures and contact times.</p>
Full article ">Figure 13 Cont.
<p>Boron adsorption capacity in different conditions. (<b>a</b>) Boron adsorption in different adsorbents with 5 ppm MgCl<sub>2</sub>; (<b>b</b>) Comparison of N<sub>2</sub> and NH<sub>3</sub> nanobubbles with 5 ppm MgCl<sub>2</sub> for char B adsorbent. (<b>c</b>) Boron adsorption capacity at different temperatures and contact times.</p>
Full article ">Figure 14
<p>Boron adsorption isotherm and kinetics for char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles at pH 9 (adsorbent mass/volume of solution (m/V) is 0.4 g/L, measured temperature is 298 K, adsorption time is 2 h). (<b>a</b>) Langmuir adsorption isotherm, (<b>b</b>) Freundlich adsorption isotherm, (<b>c</b>) effect of adsorption capacity depending on time.</p>
Full article ">Figure 15
<p>FT-IR spectra of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption at pH 9. (Adsorbent mass/volume of solution (m/V) is 0.4 g/L, measured temperature is 298 K, adsorption time is 2 h).</p>
Full article ">Figure 16
<p>XPS spectra of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption at pH 9. (Adsorbent mass/volume of solution (m/V) is 0.4 g/L, measured temperature is 298 K, adsorption time is 2 h). (<b>a</b>–<b>d</b>) XPS spectra of C 1s of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption. (<b>e</b>–<b>h</b>) XPS spectra of O 1s of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption. (<b>i</b>) XPS survey scans of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption. (<b>j</b>) XPS spectra of B 1s of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption.</p>
Full article ">Figure 16 Cont.
<p>XPS spectra of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption at pH 9. (Adsorbent mass/volume of solution (m/V) is 0.4 g/L, measured temperature is 298 K, adsorption time is 2 h). (<b>a</b>–<b>d</b>) XPS spectra of C 1s of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption. (<b>e</b>–<b>h</b>) XPS spectra of O 1s of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption. (<b>i</b>) XPS survey scans of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption. (<b>j</b>) XPS spectra of B 1s of char A and char B before boron adsorption and those of char B with NH<sub>3</sub> nanobubbles and char B with 5 ppm MgCl<sub>2</sub> and NH<sub>3</sub> nanobubbles after boron adsorption.</p>
Full article ">Figure 17
<p>NH<sub>4</sub><sup>+</sup> ion removal using zeolite molecular sieves for adsorption and boiling.</p>
Full article ">
16 pages, 2529 KiB  
Article
Primary Study on Influence of Conventional Hydrochemical Components on Suspension of Endogenous Fine Loess Particles in Groundwater over Loess Regions
by Zherui Zhang, Xinshuo Wang, Zuoyi Wang, Haiqiang Lan, Ran Sun, Sihai Hu, Xiaofeng Sun and Yaoguo Wu
Appl. Sci. 2024, 14(19), 8809; https://doi.org/10.3390/app14198809 - 30 Sep 2024
Viewed by 366
Abstract
To ascertain the effects of conventional hydrochemical components on the presence of endogenous fine loess particles (EFLPs) in groundwater over loess regions, Na+, NO3 and Cu2+, as conventional hydrochemical components, were employed in batch tests with EFLPs [...] Read more.
To ascertain the effects of conventional hydrochemical components on the presence of endogenous fine loess particles (EFLPs) in groundwater over loess regions, Na+, NO3 and Cu2+, as conventional hydrochemical components, were employed in batch tests with EFLPs from a typical loess as aquifer media in Guanzhong Plain, China. The results showed that EFLPs had high zeta potential (ζ) and remained suspended over 40 h, indicating their good dispersity and potential to be suspended in groundwater. ζ was employed to replace electrostatic repulsion in the DLVO equation to determine the critical coagulation concentrations for Cu(NO3)2 and NaF as 0.1 mmol/L and 50 mmol/L for 1.1 µm D50 EFLPs, which were almost consistent with the batch test results and greater than those in the groundwater, respectively, further implying that EFLPs are likely to be suspended in groundwater. The multi-factor tests showed that the key factors including particle size, hydro-chemical component and concentration interacted with each other and their relative magnitudes varied in the test processes, where the effects of concentration strengthened while those of the component weakened. So, hydrogeochemical conditions were beneficial to the suspension of EFLPs and the benefit got strong along the groundwater flow path, which is conducive to the cotransport of EFLPs with pollutants in groundwater over loess regions. Full article
(This article belongs to the Special Issue Advances in Soil and Water Pollution Control)
Show Figures

Figure 1

Figure 1
<p>Effects of particle diameter size on the AS rate of EFLPs.</p>
Full article ">Figure 2
<p>Effects of Cu<sup>2+</sup> concentration (<b>I</b>) and Na<sup>+</sup> concentration (<b>II</b>) on the AS rate of EFLPs.</p>
Full article ">Figure 3
<p>Effect of hydrochemical components on the AS rate of EFLPs.</p>
Full article ">Figure 4
<p>Normal distribution plot of residuals at 3-h reaction.</p>
Full article ">Figure 5
<p>Three-dimensional surface plots and contour maps for the interaction of hydrochemical component and concentration (<b>I-a</b>,<b>I-A</b>), hydrochemical component and particle size (<b>I-b</b>,<b>I-B</b>), and hydrochemical concentration and particle size (<b>I-c</b>,<b>I-C</b>) on the AS and suspension of 1.1 µm EFLPs with a 3-h reaction, respectively.</p>
Full article ">Figure 6
<p>Three-dimensional surface plots and contour maps for the interaction of hydro-chemical component and concentration (<b>II-a</b>,<b>II-A</b>), hydro-chemical component and particle size (<b>II-b</b>,<b>II-B</b>), and component concentration and particle size (<b>II-c</b>,<b>II-C</b>) on the AS and suspension of 1.1 µm EFLPs in a 36-h reaction, respectively.</p>
Full article ">Figure 6 Cont.
<p>Three-dimensional surface plots and contour maps for the interaction of hydro-chemical component and concentration (<b>II-a</b>,<b>II-A</b>), hydro-chemical component and particle size (<b>II-b</b>,<b>II-B</b>), and component concentration and particle size (<b>II-c</b>,<b>II-C</b>) on the AS and suspension of 1.1 µm EFLPs in a 36-h reaction, respectively.</p>
Full article ">
13 pages, 4199 KiB  
Article
Utilization of Lead Nitrate to Enhance the Impact of Hydroxamic Acids on the Hydrophobic Aggregation and Flotation Behavior of Cassiterite
by Saizhen Jin, Xiaobo Liu, Yun Feng, Yanfei Chen, Mengtao Wang and Qingfei Xiao
Molecules 2024, 29(15), 3692; https://doi.org/10.3390/molecules29153692 - 4 Aug 2024
Viewed by 695
Abstract
Lead nitrate (LN) is frequently employed as an activator in the flotation of cassiterite using hydroxamic acids as the collectors. This study investigated the effect of LN on the hydrophobic aggregation of cassiterite when benzohydroxamic acid (BHA), hexyl hydroxamate (HHA), and octyl hydroxamate [...] Read more.
Lead nitrate (LN) is frequently employed as an activator in the flotation of cassiterite using hydroxamic acids as the collectors. This study investigated the effect of LN on the hydrophobic aggregation of cassiterite when benzohydroxamic acid (BHA), hexyl hydroxamate (HHA), and octyl hydroxamate (OHA) were used as the collectors through micro-flotation, focused beam reflectance measurement (FBRM) and a particle video microscope (PVM), zeta potential, and the extended DLVO theory. Micro-flotation tests confirmed that LN activated the flotation of cassiterite using the hydroxamic acids as collectors. Focused beam reflectance measurement (FBRM) and a particle video microscope (PVM) were used to capture in situ data on the changes in size distribution and morphology of cassiterite aggregates during stirring. The FBRM and PVM image results indicated that the addition of LN could promote the formation of hydrophobic aggregates of fine cassiterite, when BHA or HHA was used as the collector, and reduce the dosage of OHA needed to induce the formation of hydrophobic aggregates of cassiterite. The extended DLVO theory interaction energies indicated that the presence of LN could decrease the electrostatic interaction energies (Vedl) and increase the hydrophobic interaction energies (Vhy) between cassiterite particles, resulting in the disappearance of the high energy barriers that existed between the particles in the absence of LN. Thus, cassiterite particles could aggregate in the presence of LN when BHA, HHA, or a low concentration of OHA was used as the collector. Full article
(This article belongs to the Special Issue Molecular Structure of Minerals)
Show Figures

Figure 1

Figure 1
<p>Recovery of fine cassiterite as a function of pH in the presence and absence of LN using (<b>a</b>) BHA, (<b>b</b>) HHA, and (<b>c</b>) OHA as collectors.</p>
Full article ">Figure 2
<p>Aggregation of fine cassiterite in the presence of various BHA concentrations. (<b>a</b>) Counts and square-weighted mean chord length of cassiterite suspension as a function of time after adding (<b>a1</b>) 3 × 10<sup>−4</sup> mol/L, (<b>a2</b>) 5 × 10<sup>−4</sup> mol/L, and (<b>a3</b>) 2 × 10<sup>−3</sup> mol/L of BHA; (<b>b</b>) non-weighted and square-weighted chord-length distributions of the cassiterite suspension before and after adding various concentrations of BHA at 20:00; (<b>c</b>) PVM images of cassiterite before (<b>c1</b>) and after (<b>c2</b>) adding 2 × 10<sup>−3</sup> mol/L BHA at 20:00. (400 rpm; pH = 8.5~9.0).</p>
Full article ">Figure 3
<p>Aggregation of fine cassiterite in the presence of various BHA concentrations and Pb<sup>2+</sup>. (<b>a</b>) Counts and square-weighted mean chord length of cassiterite suspension as a function of time; (<b>b</b>) non-weighted and square-weighted chord-length distributions of cassiterite suspension at different times; (<b>c</b>) PVM images of cassiterite after adding BHA at 15:00. (C<sub>(BHA)</sub> = 2 × 10<sup>−4</sup> mol/L (<b>a1</b>,<b>b1</b>,<b>c1</b>), C<sub>(BHA)</sub> = 5 × 10<sup>−4</sup> mol/L (<b>a2</b>,<b>b2</b>,<b>c2</b>), C<sub>(BHA)</sub> = 2 × 10<sup>−3</sup> mol/L (<b>a3</b>,<b>b3</b>,<b>c3</b>); 400 rpm; pH = 8.5~9.0; C<sub>(LN)</sub> = 6 × 10<sup>−5</sup> mol/L).</p>
Full article ">Figure 4
<p>Aggregation of cassiterite under different HHA concentrations in the presence of LN. (<b>a</b>) Counts and square-weighted mean chord length of cassiterite suspension as a function of time. (<b>b</b>) Non-weighted and square-weighted CLDs of cassiterite suspension at different times. (C<sub>(LN)</sub> = 6 × 10<sup>−5</sup> mol/L; C<sub>(HHA)</sub> = 5 × 10<sup>−5</sup> mol/L (<b>a1</b>,<b>b1</b>), C<sub>(HHA)</sub> = 1 × 10<sup>−4</sup> mol/L (<b>a2</b>,<b>b2</b>), C<sub>(HHA)</sub> = 4 × 10<sup>−4</sup> mol/L (<b>a3</b>,<b>b3</b>); N = 400 rpm; pH = 8.5~9.0).</p>
Full article ">Figure 5
<p>PVM images of cassiterite after adding 6 × 10<sup>−5</sup> mol/L LN and 5 × 10<sup>−5</sup> mol/L HHA at the time points of 4:30, 10:30, and 20:00.</p>
Full article ">Figure 6
<p>Aggregation of cassiterite by different OHA concentrations in the presence of LN. (C<sub>(LN)</sub> = 6 × 10<sup>−5</sup> mol/L; C<sub>(OHA)</sub> = 5 × 10<sup>−5</sup> mol/L (<b>a1</b>,<b>b1</b>), C<sub>(OHA)</sub> = 1 × 10<sup>−4</sup> mol/L (<b>a2</b>,<b>b2</b>), C<sub>(OHA)</sub>= 4 × 10<sup>−4</sup> mol/L (<b>a3</b>,<b>b3</b>); N = 400 rpm; pH = 8.5~9.0).</p>
Full article ">Figure 7
<p>PVM images of cassiterite after adding OHA and LN. (C<sub>(OHA)</sub> = 5 × 10<sup>−5</sup> mol/L (<b>a</b>), C<sub>(OHA)</sub> = 1 × 10<sup>−4</sup> mol/L (<b>b</b>), C<sub>(OHA)</sub> = 4 × 10<sup>−4</sup> mol/L (<b>c</b>)).</p>
Full article ">Figure 8
<p>The EDLVO interaction energy diagram as a function of separation distance between cassiterite particles in the presence of BHA or BHA + Pb<sup>2+</sup>. (<b>a</b>) V<sub>edl</sub> and V<sub>hy</sub> in the absence of LN; (<b>b</b>) presence of <sub>LN</sub>; and (<b>c</b>) V<sub>DT</sub> in the absence and presence of LN.</p>
Full article ">Figure 9
<p>The EDLVO interaction energy diagram as a function of separation distance between cassiterite particles in the presence of LN and HHA (<b>a</b>) or OHA (<b>b</b>). HHA + LN (<b>a1</b>,<b>a2</b>) and OHA+LN (<b>b1</b>,<b>b2</b>).</p>
Full article ">
20 pages, 8551 KiB  
Article
Migration Rules and Mechanisms of Nano-Biochar in Soil Columns under Various Transport Conditions
by Peng Li, Meifang Yan, Min Li, Tao Zhou, Huijie Li and Bingcheng Si
Nanomaterials 2024, 14(12), 1035; https://doi.org/10.3390/nano14121035 - 15 Jun 2024
Viewed by 857
Abstract
Compared to traditional biochar (BC), nano-biochar (NBC) boasts superior physicochemical properties, promising extensive applications in agriculture, ecological environments, and beyond. Due to its strong adsorption and migration properties, NBC may carry nutrients or pollutants to deeper soil layers or even groundwater, causing serious [...] Read more.
Compared to traditional biochar (BC), nano-biochar (NBC) boasts superior physicochemical properties, promising extensive applications in agriculture, ecological environments, and beyond. Due to its strong adsorption and migration properties, NBC may carry nutrients or pollutants to deeper soil layers or even groundwater, causing serious environmental risks. Nevertheless, the migration rules and mechanisms of NBC in soil are still unclear. Therefore, this study employed soil column migration experiments to systematically explore the migration rules and mechanisms of NBC under various flow rates, initial soil water contents, soil depths, and soil textures. The results showed that regulated by smaller particle size differences and greater surface charges, NBC exhibited a stronger migration ability compared with traditional BC. As the soil texture transitioned from fine to coarse, the migration capability of NBC significantly improved, driven by both pore structure and interaction forces as described by the DLVO theory. The migration ability of NBC was also greatly boosted as the soil transitioned from saturated to unsaturated conditions, primarily because of preferential flow. When the flow rate increased from 70% KS to 100% KS and 130% KS, the migration ability of NBC also increased accordingly, as changes in injection flow rates altered the velocity distribution of pore water. NBC in 25 cm soil columns was more prone to shallow retention compared with 10 cm soil columns, resulting in weaker overall migration ability. In addition, through fitting of the two-site kinetic model and related parameters, the penetration curves of NBC under various variable conditions were effectively characterized. These findings could offer valuable insights for NBC’s future efficient, rational, and sustainable utilization, facilitating the evaluation and mitigation of its potential environmental risks. Full article
Show Figures

Figure 1

Figure 1
<p>Optimization steps of parameters for the preparation of nano-biochar by ball milling.</p>
Full article ">Figure 2
<p>Size distribution of nano-biochar particles.</p>
Full article ">Figure 3
<p>Diagram of soil column migration experiment installation.</p>
Full article ">Figure 4
<p>The scanning electron microscopy images of bulk biochar (<b>a</b>) and nano-biochar (<b>b</b>).</p>
Full article ">Figure 5
<p>FTIR spectra of bulk biochar and nano-biochar.</p>
Full article ">Figure 6
<p>Br<sup>-</sup> penetration curves and fitting results.</p>
Full article ">Figure 7
<p>Penetration curves of biochar with different particle sizes.</p>
Full article ">Figure 8
<p>Migration penetration curves and fitting results of nano-biochar in soil under different soil properties (<b>a</b>), injection flow rates in sandy loam (<b>b</b>), different initial water contents in sandy loam (<b>c</b>) and soil column lengths in sandy loam (<b>d</b>).</p>
Full article ">Figure 9
<p>Potential energy of DLVO interaction between nano-biochar and soil of different textures.</p>
Full article ">Figure A1
<p>The calibration curves for Br detection of (<b>a</b>) 0.2–2 ppm and (<b>b</b>) 2–20 ppm.</p>
Full article ">
16 pages, 4615 KiB  
Article
The Effect of Lysozyme on the Aggregation and Charging of Oxidized Carbon Nanohorn (CNHox) in Aqueous Solution
by Zhengjian Tian, Maolin Li, Takuya Sugimoto and Motoyoshi Kobayashi
Appl. Sci. 2024, 14(6), 2645; https://doi.org/10.3390/app14062645 - 21 Mar 2024
Cited by 1 | Viewed by 795
Abstract
To clarify the effect of proteins on the charging and aggregation–dispersion characteristics of oxidized carbon nanohorn (CNHox), we measured the electrophoretic mobility and stability ratios as a function of concentrations of a model protein, lysozyme (LSZ), and KCl. The zeta potential from the [...] Read more.
To clarify the effect of proteins on the charging and aggregation–dispersion characteristics of oxidized carbon nanohorn (CNHox), we measured the electrophoretic mobility and stability ratios as a function of concentrations of a model protein, lysozyme (LSZ), and KCl. The zeta potential from the electrophoretic mobility of CNHox was neutralized and reversed by the addition of oppositely charged LSZ. Electrical and hydrophobic interactions between CNHox and LSZ can be attributed to the adsorption and charge reversal of CNHox. The stability ratio of CNHox in the presence or absence of LSZ showed Derjaguin–Landau and Verwey–Overbeek (DLVO) theory-like behavior. That is, the slow aggregation regime, fast aggregation regime, and critical coagulation concentration (CCC) were identified. At the isoelectric point, only the fast aggregation regime was shown. The existence of patch-charge attraction due to the charge heterogeneity on the surface was inferred to have happened due to the enhanced aggregation of CNHox at high LSZ dosage and low electrolyte concentration. The relationship between critical coagulation ionic strength and surface charge density at low LSZ dosage showed that the aggregation of CNHox is in line with the DLVO theory. An obvious decrement in the Hamaker constant at high LSZ dosage can probably be found due to an increased interaction of LSZ-covered parts. Full article
(This article belongs to the Section Chemical and Molecular Sciences)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the oxidation of a CNH.</p>
Full article ">Figure 2
<p>Transmission electron micrograph images of CNHox particles [<a href="#B15-applsci-14-02645" class="html-bibr">15</a>] reused with permission from Elsevier. Irreversible aggregates of CNHox of approximately 100 nm can be seen.</p>
Full article ">Figure 3
<p>The relationship between the zeta potential of the CNHox and KCl concentration.</p>
Full article ">Figure 4
<p>Hydrodynamic diameter of bare CNHox particles as a function of time at different KCl concentrations.</p>
Full article ">Figure 5
<p>The reciprocal stability ratio 1/<span class="html-italic">W</span> of CNHox as a function of KCl concentration.</p>
Full article ">Figure 6
<p>Zeta potential of lysozyme-coated CNHox in 10 mM KCl as a function of the mass ratio of lysozyme to CNHox (g/g).</p>
Full article ">Figure 7
<p>Hydrodynamic diameter changes in lysozyme-coated CNHox particles with time for the mass ratios of (<b>a</b>) 0.0875 g/g, (<b>b</b>) 0.01 g/g, and (<b>c</b>) 50 g/g and various KCl concentrations.</p>
Full article ">Figure 8
<p>Reciprocal stability ratio 1/<span class="html-italic">W</span> against KCl concentration for bare CNHox (blue circles), CNHox with LSZ at the IEP (red triangles), CNHox with LSZ at low mass ratio (green squares), and CNHox with LSZ at high mass ratio (yellow circles).</p>
Full article ">Figure 9
<p>Sketch map of patch-charge attraction between two LSZ-saturated CNHox particles.</p>
Full article ">Figure 10
<p>Comparison between surface charge density (σ) of LSZ-coated CNHox (red filled circles), bare CNHox [<a href="#B15-applsci-14-02645" class="html-bibr">15</a>] (white open triangle), and critical coagulation ionic strength (CCIS) with DLVO prediction with different Hmaker constant <span class="html-italic">H</span> (dash-dot line, solid line, dotted line, and broken line).</p>
Full article ">Figure 11
<p>Zeta potential of bare CNHox or CNHox with the presence of LSZ at different mass ratios depending on the electrolyte concentration.</p>
Full article ">
25 pages, 3272 KiB  
Article
Counter-Ion Effect on the Surface Potential of Foam Films and Foams Stabilized by 0.5 mmol/L Sodium Dodecyl Sulfate
by Nidelina Petkova, Dilyana Ivanova-Stancheva, Nikolay A. Grozev, Kristina Mircheva and Stoyan I. Karakashev
Coatings 2024, 14(1), 51; https://doi.org/10.3390/coatings14010051 - 28 Dec 2023
Viewed by 1136
Abstract
It is well known that the type of counter-ion affects the state of the adsorption layer of ionic surfactants and, consequently, its surface potential. Yet, it is not clear how they affect the foamability, the rate of foam decay or foam production. How [...] Read more.
It is well known that the type of counter-ion affects the state of the adsorption layer of ionic surfactants and, consequently, its surface potential. Yet, it is not clear how they affect the foamability, the rate of foam decay or foam production. How is the surface potential of the air/water interface related to the properties of the foam? This work aims to answer these questions. Foam films, stabilized by 0.5 mmol/L sodium dodecyl sulfate (SDS) in the presence of added LiCl, NaCl, and KCl, were studied by means of the interferometric experimental setup of Scheludko–Exerowa. The surface potential values were derived from the equilibrium film thickness by means of the DLVO theory. A linear relation between the values of the surface potential and specific adsorption energy of the counter-ions on the air/water interface was established. The slope of this linear relation depends on the salt concentration. The foamability, the rate of foam decay, and the foam production of the same aqueous solutions of SDS and added salts were studied by means of the shaking method. A correlation was found between the derived surface potential of the foam film’s surfaces and the properties of the foam. The foam production, which is the ratio between the initial foam volume and the rate of foam decay, increases with the decrease in the surface potential. Previous studies in the literature confirm that the lower surface potential promotes higher surfactant adsorption, thus boosting more foam and vice versa. It was also confirmed that the dual effect of KCl on foam production involves converting the best foam stabilizer into a foam suppressor at the highest salt concentration. Full article
(This article belongs to the Special Issue Nanostructured Films and Their Multi-scale Applications)
Show Figures

Figure 1

Figure 1
<p>Surface tension values of 0.5 mmol/L SDS versus the concentration of the added LiCl, NaCl, KCl.</p>
Full article ">Figure 2
<p>Equilibrium thickness of foam films stabilized by 0.5 mmol/L SDS versus the concentration of the added LiCl, NaCl, and KCl.</p>
Full article ">Figure 3
<p>Surface potential of foam films stabilized by 0.5 mmol/L SDS versus the concentration of the added LiCl, NaCl, KCl; the red line represents the surface potential versus the concentration of added salt but with excluded adsorption of the counter-ions.</p>
Full article ">Figure 4
<p>Dimensionless surface potential <math display="inline"><semantics> <mrow> <mi>F</mi> <msub> <mi>φ</mi> <mi>s</mi> </msub> <mo>/</mo> <msub> <mi>R</mi> <mi>g</mi> </msub> <mi>T</mi> </mrow> </semantics></math> of the foam films at different concentrations of added salts versus the specific energy of adsorption <math display="inline"><semantics> <mrow> <mo>−</mo> <msub> <mi>u</mi> <mn>0</mn> </msub> <mo>/</mo> <msub> <mi>k</mi> <mi>B</mi> </msub> <mi>T</mi> </mrow> </semantics></math> of Li<sup>+</sup>, Na<sup>+</sup> and K<sup>+</sup> counter-ions.</p>
Full article ">Figure 5
<p>Total disjoining pressure versus distance between the film’s surfaces for the cases of (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt.</p>
Full article ">Figure 5 Cont.
<p>Total disjoining pressure versus distance between the film’s surfaces for the cases of (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt.</p>
Full article ">Figure 6
<p>Initial foam volume versus the specific adsorption energy of the counter-ions on the air/water interface for (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 6 Cont.
<p>Initial foam volume versus the specific adsorption energy of the counter-ions on the air/water interface for (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 7
<p>Initial rate of foam decay versus the specific adsorption energy of the counter-ions on the air/water interface for (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 7 Cont.
<p>Initial rate of foam decay versus the specific adsorption energy of the counter-ions on the air/water interface for (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 8
<p>Foam production versus the specific adsorption energy of the counter-ions on the air/water interface for each one of the concentrations of added salt.</p>
Full article ">Figure 9
<p>Initial foam volume versus equilibrium surface potential values of the foam films for: (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 9 Cont.
<p>Initial foam volume versus equilibrium surface potential values of the foam films for: (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 10
<p>Initial foam volume versus equilibrium surface potential values of the foam films for all of the concentrations of the added salt.</p>
Full article ">Figure 11
<p>Initial rate of foam decay versus equilibrium surface potential values of the foam films for (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 11 Cont.
<p>Initial rate of foam decay versus equilibrium surface potential values of the foam films for (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 12
<p>Initial rate of foam decay versus equilibrium surface potential values of the foam films for all of the concentrations of the added salt.</p>
Full article ">Figure 13
<p>Foam production versus equilibrium surface potential values of the foam films for: (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 13 Cont.
<p>Foam production versus equilibrium surface potential values of the foam films for: (<b>A</b>) 5.5 mmol/L added salt; (<b>B</b>) 11 mmol/L added salt; (<b>C</b>) 20 mmol/L added salt; (<b>D</b>) 40 mmol/L added salt. The inverted Debye length for each particular case is presented as well.</p>
Full article ">Figure 14
<p>Foam production versus equilibrium surface potential values of the foam films for each one of the concentrations of the added salt.</p>
Full article ">
22 pages, 3523 KiB  
Review
Dynamic Light Scattering and Its Application to Control Nanoparticle Aggregation in Colloidal Systems: A Review
by Jesus Rodriguez-Loya, Maricarmen Lerma and Jorge L. Gardea-Torresdey
Micromachines 2024, 15(1), 24; https://doi.org/10.3390/mi15010024 - 22 Dec 2023
Cited by 9 | Viewed by 3252
Abstract
Colloidal systems and their control play an essential role in daily human activities, but several drawbacks lead to an avoidance of their extensive application in some more productive areas. Some roadblocks are a lack of knowledge regarding how to influence and address colloidal [...] Read more.
Colloidal systems and their control play an essential role in daily human activities, but several drawbacks lead to an avoidance of their extensive application in some more productive areas. Some roadblocks are a lack of knowledge regarding how to influence and address colloidal forces, as well as a lack of practical devices to understand these systems. This review focuses on applying dynamic light scattering (DLS) as a powerful tool for monitoring and characterizing nanoparticle aggregation dynamics. We started by outlining the core ideas behind DLS and how it may be used to examine colloidal particle size distribution and aggregation dynamics; then, in the last section, we included the options to control aggregation in the chemically processed toner. In addition, we pinpointed knowledge gaps and difficulties that obstruct the use of DLS in real-world situations. Although widely used, DLS has limits when dealing with complicated systems, including combinations of nanoparticles, high concentrations, and non-spherical particles. We discussed these issues and offered possible solutions and the incorporation of supplementary characterization approaches. Finally, we emphasized how critical it is to close the gap between fundamental studies of nanoparticle aggregation and their translation into real-world applications, recognizing challenges in colloidal science. Full article
(This article belongs to the Special Issue Self-Assembly of Nanoparticles)
Show Figures

Figure 1

Figure 1
<p>Relevant articles per country. The United States has a higher number of articles per country from the years 2018–2023 relating to control agglomeration.</p>
Full article ">Figure 2
<p>This figure depicts the breakdown of relevant articles published between 2018 and 2023 concerning controlled agglomeration, categorized by their respective subject areas. As shown, the field of Chemistry dominates the distribution with the highest number of articles published, followed by Physics and Astronomy. This suggests a significant focus on the application of controlled agglomeration within the realm of chemical research.</p>
Full article ">Figure 3
<p>DLS principle of operation: laser light hits particles in suspension, and a high-time-resolution detector measures the intensity of the scattered light at a specific angle <span class="html-italic">θ</span>; scattered light intensity is due to particles’ Brownian motion fluctuating over time. Reproduced with permission from reference [<a href="#B73-micromachines-15-00024" class="html-bibr">73</a>].</p>
Full article ">Figure 4
<p>Incident light that strikes dispersed nanoparticles (NPs) is scattered to the 6th power of NP’s radii. Particle size &lt;1/10th of the incident wavelength (λ/10), scatter light with same energy (Rayleigh scattering). Particle size &gt;1/10th of the incident wavelength scatter light is angle dependent (Mie scattering). Reproduced with permission from reference [<a href="#B76-micromachines-15-00024" class="html-bibr">76</a>].</p>
Full article ">Figure 5
<p>Zeta potential and the electrical double layer. Reproduced with permission from reference [<a href="#B76-micromachines-15-00024" class="html-bibr">76</a>].</p>
Full article ">Figure 6
<p>Isoelectric point latex sample.</p>
Full article ">Figure 7
<p>Analyzing nanoplastics in environmental samples. Reproduced with permission from reference [<a href="#B99-micromachines-15-00024" class="html-bibr">99</a>].</p>
Full article ">Figure 8
<p>Parameters and variables in a colloidal system and their particle agglomeration performance are explained using the DLVO (Derjaguin, Landau, Vervey, and Overbeek) theory (<b>a</b>). Two main contributors to the total free energy per unit area are the additive forces of (<span class="html-italic">V<sub>vdw</sub></span>) van der Waals and (<span class="html-italic">V<sub>dl</sub></span>) double-layer interactions (<b>a</b>). To evaluate the <span class="html-italic">V<sub>vdw</sub></span> forces (<b>b</b>), the Hamaker constant can be measured using the Atomic Force Microscope (AFM), as described in (<b>d</b>). The <span class="html-italic">V<sub>dl</sub></span> forces (<b>c</b>) can be defined using any of the two equations. At the same time, the ionic strength (<b>f</b>) can be measured to replace the value in equation (<b>e</b>).</p>
Full article ">Figure 9
<p>Major variables in a colloidal system: qi = particle charge, Ri = particle radius, r = surface particle distance, h = particle center distance, Єm = dielectric permittivity from the medium.</p>
Full article ">Figure 10
<p>Emulsion aggregation. Nucleation types. Depicts the nucleation mechanism types in emulsion polymerization, mainly represented as free-radical polymerization [<a href="#B121-micromachines-15-00024" class="html-bibr">121</a>]. Open access.</p>
Full article ">
20 pages, 5967 KiB  
Review
Protein Association in Solution: Statistical Mechanical Modeling
by Vojko Vlachy, Yurij V. Kalyuzhnyi, Barbara Hribar-Lee and Ken A. Dill
Biomolecules 2023, 13(12), 1703; https://doi.org/10.3390/biom13121703 - 24 Nov 2023
Cited by 2 | Viewed by 1385
Abstract
Protein molecules associate in solution, often in clusters beyond pairwise, leading to liquid phase separations and high viscosities. It is often impractical to study these multi-protein systems by atomistic computer simulations, particularly in multi-component solvents. Instead, their forces and states can be studied [...] Read more.
Protein molecules associate in solution, often in clusters beyond pairwise, leading to liquid phase separations and high viscosities. It is often impractical to study these multi-protein systems by atomistic computer simulations, particularly in multi-component solvents. Instead, their forces and states can be studied by liquid state statistical mechanics. However, past such approaches, such as the Derjaguin-Landau-Verwey-Overbeek (DLVO) theory, were limited to modeling proteins as spheres, and contained no microscopic structure–property relations. Recently, this limitation has been partly overcome by bringing the powerful Wertheim theory of associating molecules to bear on protein association equilibria. Here, we review these developments. Full article
(This article belongs to the Section Biomacromolecules: Proteins)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Driving forces of association as deduced from experiments. From left to right: proteins cluster beyond pairwise; electrostatics can be important; clustering is enthalpic; and hydration and hydrophobicity are important.</p>
Full article ">Figure 2
<p>Model of the interactions of two globular proteins. Protein spheres interact at M × M pairs of binding sites on the surfaces, one pair of which (A and B) is indicated here. Reprinted by permission from [<a href="#B103-biomolecules-13-01703" class="html-bibr">103</a>].</p>
Full article ">Figure 3
<p>Liquid–liquid phase equilibria: theory and experiment. Top: <math display="inline"><semantics> <mi>γ</mi> </semantics></math> IIIa-crystallin. Bottom: lysozyme. Solid curves are calculated from the model; see [<a href="#B103-biomolecules-13-01703" class="html-bibr">103</a>] for details. Experimental data shown by symbols taken from Taratuta [<a href="#B117-biomolecules-13-01703" class="html-bibr">117</a>] (upper symbols) and Broide [<a href="#B118-biomolecules-13-01703" class="html-bibr">118</a>] (lower symbols). Reprinted by permission from [<a href="#B103-biomolecules-13-01703" class="html-bibr">103</a>].</p>
Full article ">Figure 4
<p>Experimental data modulation of protein interactions by salts in lysozyme solutions. <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>c</mi> <mi>l</mi> <mi>o</mi> <mi>u</mi> <mi>d</mi> </mrow> </msub> </semantics></math> for lysozyme as a function of ionic strength of the added alkali-halide salts <math display="inline"><semantics> <msub> <mi>I</mi> <mrow> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </msub> </semantics></math>; the symbols denote experimental data [<a href="#B117-biomolecules-13-01703" class="html-bibr">117</a>]. The lines are the results of Equation (12) from [<a href="#B103-biomolecules-13-01703" class="html-bibr">103</a>] (from top to bottom: KBr, NaBr, KCl, and NaCl salts). Reprinted by permission from [<a href="#B103-biomolecules-13-01703" class="html-bibr">103</a>].</p>
Full article ">Figure 5
<p>The ions that bind most weakly to water most strongly affect the protein-protein attraction. Correlation between slope <span class="html-italic">a</span> in Equation (12) [<a href="#B103-biomolecules-13-01703" class="html-bibr">103</a>] and the Gibbs free energy of hydration <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>G</mi> <mrow> <mi>h</mi> <mi>y</mi> <mi>d</mi> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math> for the corresponding anions. Reprinted by permission from [<a href="#B103-biomolecules-13-01703" class="html-bibr">103</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Seven-bead model molecules. Each Y-shaped molecule first assembles from seven individual beads via strong forces, which act only between the sticky spots of the same color. Next, these molecules associate into non-covalent clusters. A and B denote the Fab fragments (the region of the antibody that binds to antigens), while C denotes the Fc arm (called the fragment crystallizable region, which interacts with the cell surface receptors). Figure reprinted by permission from [<a href="#B56-biomolecules-13-01703" class="html-bibr">56</a>]. Copyright Elsevier (2017).</p>
Full article ">Figure 7
<p>Relative viscosity <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>/</mo> <msub> <mi>η</mi> <mn>0</mn> </msub> </mrow> </semantics></math> as a function of the protein concentration [<a href="#B56-biomolecules-13-01703" class="html-bibr">56</a>]. From bottom to top: (i) model of bispecific antibodies, green curve; (ii) symmetric Fab–Fab model of antibodies, red curve; and (iii) model of interacting Fab–Fc terminals, blue curve. For more details, see the original paper. Figure reprinted by permission from [<a href="#B56-biomolecules-13-01703" class="html-bibr">56</a>]. Copyright Elsevier (2017).</p>
Full article ">Figure 8
<p>Three types of antibody clustering studied in this work. The types of antibody clustering studied in this work are: (<b>a</b>) monospecific two-arm binding; (<b>b</b>) bispecific single-arm binding; and (<b>c</b>) arms-to-Fc binding. <a href="#biomolecules-13-01703-f008" class="html-fig">Figure 8</a> reprinted by permission from [<a href="#B56-biomolecules-13-01703" class="html-bibr">56</a>]. Copyright Elsevier (2017).</p>
Full article ">Figure 9
<p>The liquid-liquid phase equilibria of mAb solutions. Temperature <span class="html-italic">T</span> vs. mAb concentration: calculated (full line) and symbols (experimental data) [<a href="#B121-biomolecules-13-01703" class="html-bibr">121</a>,<a href="#B128-biomolecules-13-01703" class="html-bibr">128</a>]. The interaction between sites A, B, and C is modeled using the short-range square-well attraction. The two-phase region is indicated by the colored area; for more details, see [<a href="#B123-biomolecules-13-01703" class="html-bibr">123</a>]. Reprinted with permission from [<a href="#B123-biomolecules-13-01703" class="html-bibr">123</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 10
<p>Liquid-liquid phase diagrams of antibody solutions in the presence of attractive obstacles. Phase diagrams <math display="inline"><semantics> <msup> <mi>T</mi> <mo>*</mo> </msup> </semantics></math> vs. <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mi>π</mi> <msubsup> <mi>ρ</mi> <mn>1</mn> <mo>*</mo> </msubsup> <mo>/</mo> <mn>6</mn> </mrow> </semantics></math> coordinate frame for model of monoclonal antibodies in Yukawa hard-sphere porous media at bonding distance <math display="inline"><semantics> <mrow> <mn>0.05</mn> <msub> <mi>σ</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and for <math display="inline"><semantics> <mrow> <msubsup> <mi>ϵ</mi> <mrow> <mi>A</mi> <mi>A</mi> </mrow> <mrow> <mo>(</mo> <mi>a</mi> <mi>s</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <msubsup> <mi>ϵ</mi> <mrow> <mi>B</mi> <mi>B</mi> </mrow> <mrow> <mo>(</mo> <mi>a</mi> <mi>s</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <msubsup> <mi>ϵ</mi> <mrow> <mi>A</mi> <mi>B</mi> </mrow> <mrow> <mo>(</mo> <mi>a</mi> <mi>s</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <mi>ϵ</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mi>ϵ</mi> <mrow> <mi>C</mi> <mi>C</mi> </mrow> <mrow> <mo>(</mo> <mi>a</mi> <mi>s</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mi>ϵ</mi> <mrow> <mi>A</mi> <mi>C</mi> </mrow> <mrow> <mo>(</mo> <mi>a</mi> <mi>s</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <msubsup> <mi>ϵ</mi> <mrow> <mi>B</mi> <mi>C</mi> </mrow> <mrow> <mo>(</mo> <mi>a</mi> <mi>s</mi> <mo>)</mo> </mrow> </msubsup> <mo>=</mo> <mi>ϵ</mi> </mrow> </semantics></math>, obstacles packing fraction <math display="inline"><semantics> <mrow> <msub> <mi>η</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>, and different strengths of the Yukawa interaction: <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>Y</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (blue (2) line), <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>Y</mi> </msub> <mo>=</mo> <mn>0.06</mn> <mi>ϵ</mi> </mrow> </semantics></math> (black (5) line), and <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>Y</mi> </msub> <mo>=</mo> <mn>0.1</mn> <mi>ϵ</mi> </mrow> </semantics></math> (red (7) line). The green (1) line denotes the result for the neat fluid (no obstacles present, <math display="inline"><semantics> <mrow> <msub> <mi>η</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>). Reproduced in part from Ref. [<a href="#B135-biomolecules-13-01703" class="html-bibr">135</a>]. Reproduced with permission of the Royal Society of Chemistry.</p>
Full article ">
20 pages, 5610 KiB  
Article
Atomic Force Microscopy of Hydrolysed Polyacrylamide Adsorption onto Calcium Carbonate
by Jin Hau Lew, Omar K. Matar, Erich A. Müller, Paul F. Luckham, Adrielle Sousa Santos and Maung Maung Myo Thant
Polymers 2023, 15(20), 4037; https://doi.org/10.3390/polym15204037 - 10 Oct 2023
Cited by 1 | Viewed by 1351
Abstract
In this work, the interaction of hydrolysed polyacrylamide (HPAM) of two molecular weights (F3330, 11–13 MDa; F3530, 15–17 MDa) with calcium carbonate (CaCO3) was studied via atomic force microscopy (AFM). In the absence of polymers at 1.7 mM and 1 M [...] Read more.
In this work, the interaction of hydrolysed polyacrylamide (HPAM) of two molecular weights (F3330, 11–13 MDa; F3530, 15–17 MDa) with calcium carbonate (CaCO3) was studied via atomic force microscopy (AFM). In the absence of polymers at 1.7 mM and 1 M NaCl, good agreement with DLVO theory was observed. At 1.7 mM NaCl, repulsive interaction during approach at approximately 20 nm and attractive adhesion of approximately 400 pN during retraction was measured, whilst, at 1 M NaCl, no repulsion during approach was found. Still, a significantly larger adhesion of approximately 1400 pN during retraction was observed. In the presence of polymers, results indicated that F3330 displayed higher average adhesion (450–625 pN) and interaction energy (43–145 aJ) with CaCO3 than F3530’s average adhesion (85–88 pN) and interaction energy (8.4–11 aJ). On the other hand, F3530 exerted a longer steric repulsion distance (70–100 nm) than F3330 (30–70 nm). This was likely due to the lower molecular weight. F3330 adopted a flatter configuration on the calcite surface, creating more anchor points with the surface in the form of train segments. The adhesion and interaction energy of both HPAM with CaCO3 can be decreased by increasing the salt concentration. At 3% NaCl, the average adhesion and interaction energy of F3330 was 72–120 pN and 5.6–17 aJ, respectively, while the average adhesion and interaction energy of F3530 was 11.4–48 pN and 0.3–2.98 aJ, respectively. The reduction of adhesion and interaction energy was likely due to the screening of the COO charged group of HPAM by salt cations, leading to a reduction of electrostatic attraction between the negatively charged HPAM and the positively charged CaCO3. Full article
(This article belongs to the Special Issue Surfaces and Interfaces of Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>Schematics of polymer adsorption model with attached segments (trains) separated by unattached segments (tails and loops), adapted from Al-Hashmi et al. [<a href="#B18-polymers-15-04037" class="html-bibr">18</a>].</p>
Full article ">Figure 2
<p>Molecular formula of HPAM.</p>
Full article ">Figure 3
<p>Force–distance curve of calcite crystal incubated in 1.7 mM NaCl solution.</p>
Full article ">Figure 4
<p>Force–distance curve of calcite crystal incubated in 1 M NaCl solution. Inset: extend force distance curve in red-squared region.</p>
Full article ">Figure 5
<p>The two commonest forms of retraction profiles observed for F3330 in 0.1% NaCl. (<b>a</b>) Type I; (<b>b</b>) Type II.</p>
Full article ">Figure 6
<p>The two commonest forms of retraction force curve shape at F3530 in 0.1% NaCl. (<b>a</b>) Type III; (<b>b</b>) Type IV.</p>
Full article ">Figure 7
<p>Histogram plot of adhesion (<b>left</b>) and interaction energy (<b>right</b>) of calcite immersed in F3330S in 0.1% NaCl. Green curve shows normal distribution plot.</p>
Full article ">Figure 8
<p>Histogram plot of adhesion (<b>left</b>) and interaction energy (<b>right</b>) of calcite immersed in F3530S in 0.1% NaCl. Green curve shows normal distribution plot.</p>
Full article ">Figure 9
<p>Cartoon representation of the possible configurations adopted by F3330 (<b>left</b>) and F3530 (<b>right</b>) on the calcite surface and their interactions with a CaCO<sub>3</sub> particle attached to an AFM tip.</p>
Full article ">Figure 10
<p>Force–distance curve of a single HPAM molecule (black) with probable HPAM detachment mechanisms from the substrate surface, adapted from Zhang et al. [<a href="#B51-polymers-15-04037" class="html-bibr">51</a>]. Anchor points between polymers with the surface are emphasized in red.</p>
Full article ">Figure 11
<p>Force–distance curve for multiple detachment from more than one F3330 molecule (<b>top</b>) with proposed illustrations (<b>bottom</b>). Here two polymer molecules are coloured differently (blue and black) for the sake of clarity. Anchor points are emphasized in red. Different stages of polymer detachment are as followed: (<b>a</b>) Contact of CaCO<sub>3</sub> particles with both polymer molecules at rest; (<b>b</b>) detachment of majority of the anchor points of polymer molecules from the surface; (<b>c</b>) even detachment of anchor points of polymer molecules from the surface; (<b>d</b>) full detachment of polymer molecules from the surface.</p>
Full article ">Figure 12
<p>Illustration of possible single HPAM molecule (black) detachment that forms zigzag retraction force curve. Adapted from Zhang et al. [<a href="#B51-polymers-15-04037" class="html-bibr">51</a>]. Anchor points are emphasized in red. Different stages of polymer detachment are as followed: (<b>a</b>) tension due to detachment of polymer’s first anchor point from the surface produced the first adhesion peak; (<b>b</b>) stretching of the polymer molecule before second anchor point with no tension observed; (<b>c</b>) tension due to detachment of polymer’s third anchor point from the surface produced the third adhesion peak; (<b>d</b>) complete detachment of polymer molecule from the surface.</p>
Full article ">Figure 13
<p>Specific viscosity plot of F3330 in 0 M (DI water) and 0.5 M salinity. CEC of F3330 in 0 M and 0.5 M are approximately 75 ppm and 2250 ppm, respectively.</p>
Full article ">Figure 14
<p>Histogram plot of adhesion (<b>left</b>) and interaction energy (<b>right</b>) of calcite immersed in F3330 in 3% NaCl. Green curve shows normal distribution plot.</p>
Full article ">Figure 15
<p>Histogram plot of adhesion (<b>left</b>) and interaction energy (<b>right</b>) of calcite immersed in F3530 in 3% NaCl. Green curve shows normal distribution plot.</p>
Full article ">
15 pages, 16310 KiB  
Article
The Mechanism of Viscosity-Enhancing Admixture in Backfill Slurry and the Evolution of Its Rheological Properties
by Liuhua Yang, Hengwei Jia, Huazhe Jiao, Mengmeng Dong and Tongyi Yang
Minerals 2023, 13(8), 1045; https://doi.org/10.3390/min13081045 - 6 Aug 2023
Cited by 11 | Viewed by 1566
Abstract
Since filling slurry is a cement-based material, viscosity-enhancing admixture exerts a significant effect on its rheological performance and mechanical properties. Viscosity-enhancing admixture can improve pipeline transportation performance and reduce pipeline wear during the filling process of a kilometer-deep mine by changing the plastic [...] Read more.
Since filling slurry is a cement-based material, viscosity-enhancing admixture exerts a significant effect on its rheological performance and mechanical properties. Viscosity-enhancing admixture can improve pipeline transportation performance and reduce pipeline wear during the filling process of a kilometer-deep mine by changing the plastic viscosity and yield stress of high-concentration filling slurries. In order to reveal the influence mechanism of viscosity-enhancing admixture on rheological performance in slurry, the influence of viscosity-enhancing admixture on the rheological performance of slurry is explored by adjusting viscosity-enhancing admixture dosage and conducting bleeding test, liquidity test, and rheological performance test. The extended DLVO theory is employed to analyze the mechanism of HPMC on the stability of filling slurry. The results show that compared with ordinary slurry, after adding HPMC and XG, the particles of filling slurry are prone to link to form a mesh structure. Besides, the increasing frictional force between particles results in a significant decrease in the bleeding rate and liquidity of the slurry. Such an effect becomes more obvious with the increase of viscosity-enhancing admixture dosage. Meanwhile, the overall effect of HPMC molecules is better than that of XG molecules since HPMC can reduce inter-particle repulsion and facilitate particle aggregation. The optimal dosage is about 0.1%, at which time the yield stress of the filling slurry increases from 89.236 to 160.06 Pa, the plastic viscosity increases from 0.296 to 1.063 Pa·s, and the compressive strength increases from 2.58 to 3.59 MPa in 28 days. The study reveals the influence of viscosity-enhancing admixture on the rheological performance of filling slurry and its evolution characteristics, which provides theoretical support for the development of filling resistance and wear reduction technology. Full article
(This article belongs to the Special Issue Backfilling Materials for Underground Mining, Volume III)
Show Figures

Figure 1

Figure 1
<p>Particle size distribution of tailings.</p>
Full article ">Figure 2
<p>Bleeding rate under different thickeners: (<b>a</b>) XG, (<b>b</b>) HPMC.</p>
Full article ">Figure 3
<p>Influence of thickener dosing on slurry ductility: (<b>a</b>) XG, (<b>b</b>) HPMC.</p>
Full article ">Figure 4
<p>Relationship between paste shear stress and shear rate under various schemes: (<b>a</b>) XG, (<b>b</b>) HPMC.</p>
Full article ">Figure 5
<p>Relationship between apparent viscosity and shear rate under various schemes: (<b>a</b>) XG, (<b>b</b>) HPMC.</p>
Full article ">Figure 6
<p>Effect of different dosages of HPMC on total potential energy between cement particles: (<b>a</b>) 0%, (<b>b</b>) 0.1% HPMC.</p>
Full article ">Figure 7
<p>Influence of different polymer viscosity-enhanced agents on flocculation particles in cement paste: (<b>a</b>) Aggregation of particles, (<b>b</b>) growth in bubbles.</p>
Full article ">Figure 8
<p>Sectional view of curing test specimens.</p>
Full article ">Figure 9
<p>Action mechanism of thickener.</p>
Full article ">Figure 10
<p>Influence of different polymer viscosity-enhanced agents on compressive strength: (<b>a</b>) XG, (<b>b</b>) HPMC.</p>
Full article ">
19 pages, 30122 KiB  
Article
Effect of Nanoparticles on Rheological Properties of Water-Based Drilling Fluid
by Yuan Lin, Qizhong Tian, Peiwen Lin, Xinghui Tan, Huaitao Qin and Jiawang Chen
Nanomaterials 2023, 13(14), 2092; https://doi.org/10.3390/nano13142092 - 18 Jul 2023
Cited by 7 | Viewed by 2131
Abstract
Nano-water-based drilling fluids (NWBDFs) are prepared using nano-copper oxide (CuO) and multiwalled carbon nanotubes (MWCNTs) as modification materials. The effects of the temperature and concentration of the nanoparticles (NPs) on the rheological properties are studied using a rotational rheometer and viscometer. Also, the [...] Read more.
Nano-water-based drilling fluids (NWBDFs) are prepared using nano-copper oxide (CuO) and multiwalled carbon nanotubes (MWCNTs) as modification materials. The effects of the temperature and concentration of the nanoparticles (NPs) on the rheological properties are studied using a rotational rheometer and viscometer. Also, the influence of two NPs on the filtration properties is studied using a low-pressure and low-temperature filtration apparatus, as well as a scanning electron microscope (SEM). It is found that MWCNTs with a concentration of 0.05 w/v% have the most obvious influence on the NWBDFs, which improve the stability of the gel structure against temperature and also decrease the filtration rate. Finally, a theoretical model predicating the yield point (YP) and the plastic viscosity (PV) as a function of the temperature considering the influence of the NPs is developed based on DLVO theory. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images: (<b>a</b>) copper oxide nanoparticles; (<b>b</b>) multiwalled carbon nanotubes.</p>
Full article ">Figure 2
<p>The mixing design of the NWBDFs.</p>
Full article ">Figure 3
<p>Variation in the shear stress with the shear rate at different temperatures: (<b>a</b>) WBDF without NPs; (<b>b</b>) NWBDF with 0.15 wt% nano-CuO; (<b>c</b>) NWBDF with 0.15 wt% MWCNTs.</p>
Full article ">Figure 4
<p>Change in the rheological properties of the NWBDFs at different temperatures with the addition of various concentrations of CuO NPs: (<b>a</b>) AV varies with different nanoparticle concentrations and different temperatures; (<b>b</b>) PV varies with different nanoparticle concentrations and different temperatures; (<b>c</b>) YP varies with different nanoparticle concentrations and different temperatures; (<b>d</b>) YP/PV varies with different nanoparticle concentrations and different temperatures.</p>
Full article ">Figure 5
<p>Change in the rheological properties of the NWBDFs at different temperatures with the addition of various concentrations of MWCNTs: (<b>a</b>) AV varies with different nanoparticle concentrations and different temperatures; (<b>b</b>) PV varies with different nanoparticle concentrations and different temperatures; (<b>c</b>) YP varies with different nanoparticle concentrations and different temperatures; (<b>d</b>) YP/PV varies with different nanoparticle concentrations and different temperatures.</p>
Full article ">Figure 6
<p>Changes in the gel strength of the two NWBDFs with the concentration of the NPs.</p>
Full article ">Figure 7
<p>Changes in the modulus and phase angle with the concentration of the NPs at a frequency of <math display="inline"><semantics><mrow><mn>1</mn><mo> </mo><mi>Hz</mi></mrow></semantics></math> for the two NWBDFs at <math display="inline"><semantics><mrow><mn>20</mn><mo> </mo><mo>°</mo><mi mathvariant="normal">C</mi></mrow></semantics></math>.</p>
Full article ">Figure 8
<p>Changes in the modulus and phase angle with the temperature at a frequency of <math display="inline"><semantics><mrow><mn>1</mn><mo> </mo><mi>Hz</mi></mrow></semantics></math> for the two NWBDFs at a NP concentration of <math display="inline"><semantics><mrow><mn>0.05</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>.</p>
Full article ">Figure 9
<p>API filtration of the two NWBDFs’ change with the concentration of the NPs.</p>
Full article ">Figure 10
<p>Filter cake obtained from the filtration experiments with the CuO-NWBDF and different NP concentrations: (<b>a</b>) no NPs, (<b>b</b>) <math display="inline"><semantics><mrow><mn>0.025</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>; (<b>c</b>) <math display="inline"><semantics><mrow><mn>0.05</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>, (<b>d</b>) <math display="inline"><semantics><mrow><mn>0.075</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>; (<b>e</b>) <math display="inline"><semantics><mrow><mn>0.1</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>; (<b>f</b>) <math display="inline"><semantics><mrow><mn>0.15</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>.</p>
Full article ">Figure 11
<p>Filter cake obtained from the filtration experiments with the MWCNT-NWBDF and different NP concentrations: (<b>a</b>) no NPs; (<b>b</b>) <math display="inline"><semantics><mrow><mn>0.025</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>; (<b>c</b>) <math display="inline"><semantics><mrow><mn>0.05</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>; (<b>d</b>) <math display="inline"><semantics><mrow><mn>0.075</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>; (<b>e</b>) <math display="inline"><semantics><mrow><mn>0.1</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>; (<b>f</b>) <math display="inline"><semantics><mrow><mn>0.15</mn><mo> </mo><mi>w</mi><mo>/</mo><mi>v</mi><mo>%</mo></mrow></semantics></math>.</p>
Full article ">Figure 12
<p>SEM photographs of filter cake of WBDF without NPs: (<b>a</b>)<math display="inline"><semantics><mrow><mo> </mo><mn>200</mn><mrow><mo> </mo><mi>nm</mi></mrow></mrow></semantics></math>, (<b>b</b>)<math display="inline"><semantics><mrow><mo> </mo><mn>300</mn><mrow><mo> </mo><mi>nm</mi></mrow></mrow></semantics></math>; (<b>c</b>) <math display="inline"><semantics><mrow><mn>3</mn><mrow><mo> </mo><mi mathvariant="sans-serif">μ</mi><mi mathvariant="normal">m</mi></mrow></mrow></semantics></math>.</p>
Full article ">Figure 13
<p>SEM photographs of filter cake of CuO-NWBDF: (<b>a</b>)<math display="inline"><semantics><mrow><mo> </mo><mn>200</mn><mrow><mo> </mo><mi>nm</mi></mrow></mrow></semantics></math>, (<b>b</b>) <math display="inline"><semantics><mrow><mn>300</mn><mrow><mo> </mo><mi>nm</mi></mrow></mrow></semantics></math>; (<b>c</b>) <math display="inline"><semantics><mrow><mn>2</mn><mrow><mo> </mo><mi mathvariant="sans-serif">μ</mi><mi mathvariant="normal">m</mi></mrow></mrow></semantics></math>.</p>
Full article ">Figure 14
<p>SEM photographs of filter cake of MWCNT-NWBDF: (<b>a</b>)<math display="inline"><semantics><mrow><mo> </mo><mn>200</mn><mrow><mo> </mo><mi>nm</mi></mrow></mrow></semantics></math>, (<b>b</b>) <math display="inline"><semantics><mrow><mn>300</mn><mrow><mo> </mo><mi>nm</mi></mrow></mrow></semantics></math> and (<b>c</b>) <math display="inline"><semantics><mrow><mn>1</mn><mrow><mo> </mo><mi mathvariant="sans-serif">μ</mi><mi mathvariant="normal">m</mi></mrow></mrow></semantics></math>.</p>
Full article ">Figure 15
<p>Schematic diagram of the influence mechanism of CuO nanoparticles on filtration: (<b>a</b>) Moderate concentration of CuO; (<b>b</b>) Excessive concentrations of CuO.</p>
Full article ">Figure 16
<p>Schematic diagram of the influence mechanism of MWCNTs on filtration: (<b>a</b>) Moderate concentration of MWCNT; (<b>b</b>) Excessive concentrations of MWCNT.</p>
Full article ">Figure 17
<p>Change in the relative dielectric constant of water with temperature (experimental data from Dass [<a href="#B43-nanomaterials-13-02092" class="html-bibr">43</a>].</p>
Full article ">Figure 18
<p>YP changes of temperatures with different nanoparticles (<b>a</b>) CuO and (<b>b</b>) MWCNT and different NP concentrations. The lines indicate the theoretical curves obtained by fitting Equation (12).</p>
Full article ">Figure 19
<p>PV changes of YP with different nanoparticles (<b>a</b>) CuO and (<b>b</b>) MWCNT and different NP concentrations. The lines indicate the theoretical curves obtained by fitting Equation (14).</p>
Full article ">Figure 20
<p>Experimental and theoretical curves of PV versus temperature for different nanoparticles (<b>a</b>) CuO and (<b>b</b>) MWCNT.</p>
Full article ">
16 pages, 5348 KiB  
Article
Hydrophobic Flocculation of Fine Cassiterite Using Alkyl Hydroxamic Acids with Different Carbon Chain Lengths as Collectors
by Saizhen Jin, Qing Shi and Leming Ou
Molecules 2023, 28(9), 3911; https://doi.org/10.3390/molecules28093911 - 5 May 2023
Cited by 3 | Viewed by 1620
Abstract
This work investigated the hydrophobic flocculation of cassiterite using four alkyl hydroxamic acids with varying carbon chain lengths, i.e., hexyl hydroxamate (C6), octyl hydroxamate (C8), decyl hydroxamate (C10) and dodecyl hydroxamate (C12), as collectors. Microflotation [...] Read more.
This work investigated the hydrophobic flocculation of cassiterite using four alkyl hydroxamic acids with varying carbon chain lengths, i.e., hexyl hydroxamate (C6), octyl hydroxamate (C8), decyl hydroxamate (C10) and dodecyl hydroxamate (C12), as collectors. Microflotation tests were performed to investigate the flotation behaviour of cassiterite in the presence of the four alkyl hydroxamic acids. Focused beam reflectance measurement (FBRM) and a particle video microscope (PVM) were used to analyse and monitor the real-time evolution of the particle size distribution of cassiterite and the images of flocs during flocculation. The extended DLVO theory interaction energies between the cassiterite particles were calculated on the basis of the measured contact angle and the zeta potential of cassiterite to determine the aggregation and dispersion behaviour of the cassiterite particles. The microflotation test results suggested that the floatability of cassiterite improved with the increase in the carbon chain length of hydroxamates. FBRM, PVM images and extended DLVO theory calculation results indicated that when C6 was used as the collector, the cassiterite particles could not form hydrophobic flocs because the total potential energy between them was repulsive. When C8, C10 and C12 were used as collectors, the energy barrier amongst particles decreased with increasing hydroxamate concentration. The lowest concentrations of C8, C10 and C12 that could cause the hydrophobic aggregation of cassiterite were approximately 1 × 10−3, 1 × 10−4 and 2 × 10−5 mol/L, respectively. The aggregation growth rate and apparent floc size increased with an increasing collector concentration. Hydroxamic acid with a longer carbon chain could induce the cassiterite particles to form larger flocs at a lower concentration in a shorter time. Full article
Show Figures

Figure 1

Figure 1
<p>Cassiterite recovery as a function of the alkyl hydroxamic concentration (pH = 8.5).</p>
Full article ">Figure 2
<p>Cassiterite recovery as a function of pH.</p>
Full article ">Figure 3
<p>Aggregation of cassiterite by C<sub>6</sub> at different concentrations and pH 8.5–9.0. (<b>a</b>) No-weighted and square-weighted CLDs of cassiterite suspensions before and after the addition of different concentrations of C<sub>6</sub> at 20:00. (<b>b</b>,<b>c</b>) PVM images of cassiterite before and after the addition of 2 × 10<sup>−3</sup> mol/L C<sub>6</sub> at 20:00.</p>
Full article ">Figure 4
<p>Aggregation of cassiterite by different C<sub>8</sub> concentrations at pH 8.5–9.0. (<b>a</b>,<b>b<sub>2</sub></b>,<b>c<sub>2</sub></b>) No-weighted and square-weighted CLDs of the cassiterite suspension before and after the addition of different concentrations of C<sub>8</sub> at different times. (<b>b<sub>1</sub></b>,<b>c<sub>1</sub></b>) Counts and square-weighted mean chord length of the cassiterite suspension as a function of time after the addition of 1 × 10<sup>−3</sup> and 1.5 ×10<sup>−3</sup> mol/L C<sub>8</sub>. (<b>b<sub>3</sub></b>,<b>c<sub>3</sub></b>) PVM images of cassiterite before and after the addition of 1 × 10<sup>−3</sup> and 1.5 × 10<sup>−3</sup> mol/L C<sub>6</sub> at 40:00 and 20:00, respectively.</p>
Full article ">Figure 4 Cont.
<p>Aggregation of cassiterite by different C<sub>8</sub> concentrations at pH 8.5–9.0. (<b>a</b>,<b>b<sub>2</sub></b>,<b>c<sub>2</sub></b>) No-weighted and square-weighted CLDs of the cassiterite suspension before and after the addition of different concentrations of C<sub>8</sub> at different times. (<b>b<sub>1</sub></b>,<b>c<sub>1</sub></b>) Counts and square-weighted mean chord length of the cassiterite suspension as a function of time after the addition of 1 × 10<sup>−3</sup> and 1.5 ×10<sup>−3</sup> mol/L C<sub>8</sub>. (<b>b<sub>3</sub></b>,<b>c<sub>3</sub></b>) PVM images of cassiterite before and after the addition of 1 × 10<sup>−3</sup> and 1.5 × 10<sup>−3</sup> mol/L C<sub>6</sub> at 40:00 and 20:00, respectively.</p>
Full article ">Figure 5
<p>Aggregation of cassiterite by different C<sub>10</sub> concentrations at pH 8.5–9.0. (<b>a</b>,<b>b<sub>2</sub></b>,<b>c<sub>2</sub></b>) No-weighted and square-weighted CLDs of the cassiterite suspension before and after the addition of different concentrations of C<sub>10</sub> at different times. (<b>b<sub>1</sub></b>,<b>c<sub>1</sub></b>) Counts and square-weighted mean chord length of the cassiterite suspension as a function of time after adding 1 × 10<sup>−4</sup> and 2 × 10<sup>−4</sup> mol/L of C<sub>10</sub>. (<b>b<sub>3</sub></b>,<b>c<sub>3</sub></b>) PVM images of cassiterite before and after the addition of 1 × 10<sup>−4</sup> and 2 ×10<sup>−4</sup> mol/L C<sub>10</sub> at 40:00 and 30:00, respectively.</p>
Full article ">Figure 6
<p>Aggregation of cassiterite by different C<sub>12</sub> concentrations at pH 8.5–9.0. (<b>a</b>) Counts and square-weighted mean chord length of the cassiterite suspension as a function of time. (<b>b</b>) No-weighted and square-weighted CLDs of the cassiterite suspension at different times. (C = 1 × 10<sup>−5</sup> mol/L—(<b>a<sub>1</sub></b>,<b>b<sub>1</sub></b>), C = 2 × 10<sup>−5</sup> mol/L—(<b>a<sub>2</sub></b>,<b>b<sub>2</sub></b>), C = 4 × 10<sup>−5</sup> mol/L—(<b>a<sub>3</sub></b>,<b>b<sub>3</sub></b>), C = 8 × 10<sup>−5</sup> mol/L—(<b>a<sub>4</sub></b>,<b>b<sub>4</sub></b>).).</p>
Full article ">Figure 6 Cont.
<p>Aggregation of cassiterite by different C<sub>12</sub> concentrations at pH 8.5–9.0. (<b>a</b>) Counts and square-weighted mean chord length of the cassiterite suspension as a function of time. (<b>b</b>) No-weighted and square-weighted CLDs of the cassiterite suspension at different times. (C = 1 × 10<sup>−5</sup> mol/L—(<b>a<sub>1</sub></b>,<b>b<sub>1</sub></b>), C = 2 × 10<sup>−5</sup> mol/L—(<b>a<sub>2</sub></b>,<b>b<sub>2</sub></b>), C = 4 × 10<sup>−5</sup> mol/L—(<b>a<sub>3</sub></b>,<b>b<sub>3</sub></b>), C = 8 × 10<sup>−5</sup> mol/L—(<b>a<sub>4</sub></b>,<b>b<sub>4</sub></b>).).</p>
Full article ">Figure 7
<p>PVM images of cassiterite in the presence of different C<sub>12</sub> concentrations at different times. (C = 2 × 10<sup>−5</sup> mol/L—(<b>a<sub>1</sub></b>), C = 4 × 10<sup>−5</sup> mol/L—(<b>a<sub>2</sub></b>), C = 8 × 10<sup>−5</sup> mol/L—(<b>a<sub>3</sub></b>).).</p>
Full article ">Figure 8
<p>Diagram of the extended DLVO interaction energy as a function of the separation distance between cassiterite particles in the presence of alkyl hydroxamic acids. (C<sub>6</sub>—(<b>a<sub>1</sub></b>,<b>a<sub>2</sub></b>), C<sub>8</sub>—(<b>b<sub>1</sub></b>,<b>b<sub>2</sub></b>), C<sub>10</sub>—(<b>c<sub>1</sub></b>,<b>c<sub>2</sub></b>), C<sub>12</sub>—(<b>d<sub>1</sub></b>,<b>d<sub>2</sub></b>)).</p>
Full article ">Figure 8 Cont.
<p>Diagram of the extended DLVO interaction energy as a function of the separation distance between cassiterite particles in the presence of alkyl hydroxamic acids. (C<sub>6</sub>—(<b>a<sub>1</sub></b>,<b>a<sub>2</sub></b>), C<sub>8</sub>—(<b>b<sub>1</sub></b>,<b>b<sub>2</sub></b>), C<sub>10</sub>—(<b>c<sub>1</sub></b>,<b>c<sub>2</sub></b>), C<sub>12</sub>—(<b>d<sub>1</sub></b>,<b>d<sub>2</sub></b>)).</p>
Full article ">
19 pages, 2025 KiB  
Article
Quantitative Assessment of Interfacial Interactions Governing Ultrafiltration Membrane Fouling by the Mixture of Silica Nanoparticles (SiO2 NPs) and Natural Organic Matter (NOM): Effects of Solution Chemistry
by Yuqi Sun, Runze Zhang, Chunyi Sun, Zhipeng Liu, Jian Zhang, Shuang Liang and Xia Wang
Membranes 2023, 13(4), 449; https://doi.org/10.3390/membranes13040449 - 21 Apr 2023
Cited by 2 | Viewed by 1555
Abstract
Mixtures of silica nanoparticles (SiO2 NPs) and natural organic matter (NOM) are ubiquitous in natural aquatic environments and pose risks to organisms. Ultrafiltration (UF) membranes can effectively remove SiO2 NP–NOM mixtures. However, the corresponding membrane fouling mechanisms, particularly under different solution [...] Read more.
Mixtures of silica nanoparticles (SiO2 NPs) and natural organic matter (NOM) are ubiquitous in natural aquatic environments and pose risks to organisms. Ultrafiltration (UF) membranes can effectively remove SiO2 NP–NOM mixtures. However, the corresponding membrane fouling mechanisms, particularly under different solution conditions, have not yet been studied. In this work, the effect of solution chemistry on polyethersulfone (PES) UF membrane fouling caused by a SiO2 NP–NOM mixture was investigated at different pH levels, ionic strengths, and calcium concentrations. The corresponding membrane fouling mechanisms, i.e., Lifshitz–van der Waals (LW), electrostatic (EL), and acid–base (AB) interactions, were quantitatively evaluated using the extended Derjaguin–Landau–Verwey–Overbeek (xDLVO) theory. It was found that the extent of membrane fouling increased with decreasing pH, increasing ionic strength, and increasing calcium concentration. The attractive AB interaction between the clean/fouled membrane and foulant was the major fouling mechanism in both the initial adhesion and later cohesion stages, while the attractive LW and repulsive EL interactions were less important. The change of fouling potential with solution chemistry was negatively correlated with the calculated interaction energy, indicating that the UF membrane fouling behavior under different solution conditions can be effectively explained and predicted using the xDLVO theory. Full article
(This article belongs to the Section Membrane Applications)
Show Figures

Figure 1

Figure 1
<p>Variation of normalized flux with permeate volume during UF of SiO<sub>2</sub> NP–NOM mixture under different solution conditions: (<b>a</b>) pH, (<b>b</b>) ionic strength, and (<b>c</b>) calcium concentration.</p>
Full article ">Figure 2
<p>Variation of zeta potential of UF membrane and SiO<sub>2</sub> NP–NOM mixture as a function of solution chemistry: (<b>a</b>) pH, (<b>b</b>) ionic strength, and (<b>c</b>) calcium concentration.</p>
Full article ">Figure 3
<p>Profiles of the total interaction energies of membrane–foulant and foulant–foulant combinations with separation distance under different solution conditions: (<b>a</b>,<b>b</b>) pH, (<b>c</b>,<b>d</b>) IS, and (<b>e</b>,<b>f</b>) Ca<sup>2+</sup> concentration.</p>
Full article ">Figure 4
<p>Correlation between fouling potential and interaction energy under different solution conditions: (<b>a<sub>1</sub></b>,<b>a<sub>2</sub></b>) pH, (<b>b<sub>1</sub></b>,<b>b<sub>2</sub></b>) ionic strength, and (<b>c<sub>1</sub></b>,<b>c<sub>2</sub></b>) calcium concentration. Subscript 1 represents initial fouling (adhesion) stage, subscript 2 represents later fouling (cohesion) stage.</p>
Full article ">
11 pages, 1481 KiB  
Article
Forces Governing the Transport of Pathogenic and Nonpathogenic Escherichia coli in Nitrogen and Magnesium Doped Biochar Amended Sand Columns
by Katherine Quinn, Sohrab Haghighi Mood, Elizabeth Cervantes, Manuel Garcia Perez and Nehal I. Abu-Lail
Microbiol. Res. 2023, 14(1), 218-228; https://doi.org/10.3390/microbiolres14010018 - 7 Feb 2023
Cited by 1 | Viewed by 1937
Abstract
Background: Access to safe drinking water remains a global issue with fecal indicator bacteria being major pollutants. Biochars offer low-cost adsorbents for bacterial pathogens. A fundamental understanding of how biochars interact with bacterial pathogens is essential to designing effective biofilters. Methods: Water-saturated sand [...] Read more.
Background: Access to safe drinking water remains a global issue with fecal indicator bacteria being major pollutants. Biochars offer low-cost adsorbents for bacterial pathogens. A fundamental understanding of how biochars interact with bacterial pathogens is essential to designing effective biofilters. Methods: Water-saturated sand columns amended with Magnesium and Nitrogen-doped biochars produced by pyrolysis at 400, 500, 600, and 700 °C were used to Quantify the transport of pathogenic Escherichia coli O157:H7 and nonpathogenic E. coli k12 strains in porous media. Measured data were modeled using DLVO theory of colloidal stability. were explored. Results: (1) Biochar is hydrophobic while sand and bacteria are hydrophilic; (2) all Gibbs free energy values quantified between E. coli O157:H7 and biochar were negative except for biochar produced at 700 °C; (3) all types of forces investigated (van der Waals, electrostatic, and acid-base interactions) played a role in governing the interactions between bacteria and biochar. Conclusions: (1) Adding doped biochar to sand at a 2% weight ratio enhanced the retention of bacterial cells in the sand/biochar columns; (2) bacterial transport is strain-dependent and mediated by various types of forces resulting from interactions between the various functional groups displayed on bacteria and biochar/sand. Our findings emphasize the importance of monitoring biochar’s functionality to eliminate bacterial pollutants from contaminated water. Full article
Show Figures

Figure 1

Figure 1
<p>Breakthrough curves (BTCs) for negative controls: (<b>A</b>) cotton ball, (<b>B</b>) sand, and (<b>C</b>–<b>F</b>) biochars 400, 500, 600, and 700 °C. Each curve represents a triplicate with error bars representing the standard deviation. Data shown in (<b>A</b>–<b>C</b>) quantified between cotton, sand, and biochar 400 and each of the two bacterial strains were statistically similar with <span class="html-italic">p</span> values of 0.984, 0.268 and 0.061, respectively. Data shown in (<b>D</b>–<b>F</b>) quantified between biochars 500, 600, and 700 °C and each of the two bacterial strains were statistically different with <span class="html-italic">p</span> values of 0.038, 0.017, and 0.024, respectively.</p>
Full article ">Figure 1 Cont.
<p>Breakthrough curves (BTCs) for negative controls: (<b>A</b>) cotton ball, (<b>B</b>) sand, and (<b>C</b>–<b>F</b>) biochars 400, 500, 600, and 700 °C. Each curve represents a triplicate with error bars representing the standard deviation. Data shown in (<b>A</b>–<b>C</b>) quantified between cotton, sand, and biochar 400 and each of the two bacterial strains were statistically similar with <span class="html-italic">p</span> values of 0.984, 0.268 and 0.061, respectively. Data shown in (<b>D</b>–<b>F</b>) quantified between biochars 500, 600, and 700 °C and each of the two bacterial strains were statistically different with <span class="html-italic">p</span> values of 0.038, 0.017, and 0.024, respectively.</p>
Full article ">Figure 2
<p>Comparison of the Gibbs free energies calculated for both bacteria strains and biochar types (400–700 °C) or sand in DIW. No statistical differences were measured between the various biochars and bacterial strains investigated.</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>E</b>) Electrostatic, van der Waal and total energy profiles as a function of distance for interactions between <span class="html-italic">E. coli</span> k12 and biochars 400, 500, 600, and 700 °C and sand in DIW. (<b>F</b>) Comparison of all total energy profiles between <span class="html-italic">E. coli</span> k12 and all biochars and sand in water. The total energy profiles quantified between biochars and <span class="html-italic">E. coli</span> k12 were statistically different (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>E</b>) Electrostatic, van der Waal and total energy profiles as a function of distance for interactions between <span class="html-italic">E. coli</span> O157:H7 and biochars 400, 500, 600, and 700 °C and sand in DIW. (<b>F</b>) Comparison of all total energy profiles between <span class="html-italic">E. coli</span> O157:H7 and all biochars and sand in water. The total energy profiles quantified between biochars and <span class="html-italic">E. coli</span> O157:H7 were statistically different (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4 Cont.
<p>(<b>A</b>–<b>E</b>) Electrostatic, van der Waal and total energy profiles as a function of distance for interactions between <span class="html-italic">E. coli</span> O157:H7 and biochars 400, 500, 600, and 700 °C and sand in DIW. (<b>F</b>) Comparison of all total energy profiles between <span class="html-italic">E. coli</span> O157:H7 and all biochars and sand in water. The total energy profiles quantified between biochars and <span class="html-italic">E. coli</span> O157:H7 were statistically different (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
Back to TopTop