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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (114)

Search Parameters:
Keywords = giant vesicle

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 3237 KiB  
Article
Electroformation of Giant Unilamellar Vesicles from Damp Films in Conditions Involving High Cholesterol Contents, Charged Lipids, and Saline Solutions
by Ivan Mardešić, Zvonimir Boban and Marija Raguz
Membranes 2024, 14(10), 215; https://doi.org/10.3390/membranes14100215 - 12 Oct 2024
Viewed by 1032
Abstract
Giant unilamellar vesicles (GUVs) are frequently used as membrane models in studies of membrane properties. They are most often produced using the electroformation method. However, there are a number of parameters that can influence the success of the procedure. Some of the most [...] Read more.
Giant unilamellar vesicles (GUVs) are frequently used as membrane models in studies of membrane properties. They are most often produced using the electroformation method. However, there are a number of parameters that can influence the success of the procedure. Some of the most common conditions that have been shown to have a negative effect on GUV electroformation are the presence of high cholesterol (Chol) concentrations, the use of mixtures containing charged lipids, and the solutions with an elevated ionic strength. High Chol concentrations are problematic for the traditional electroformation protocol as it involves the formation of a dry lipid film by complete evaporation of the organic solvent from the lipid mixture. During drying, anhydrous Chol crystals form. They are not involved in the formation of the lipid bilayer, resulting in a lower Chol concentration in the vesicle bilayer compared to the original lipid mixture. Motivated primarily by the issue of artifactual Chol demixing, we have modified the electroformation protocol by incorporating the techniques of rapid solvent exchange (RSE), ultrasonication, plasma cleaning, and spin-coating for reproducible production of GUVs from damp lipid films. Aside from decreasing Chol demixing, we have shown that the method can also be used to produce GUVs from lipid mixtures with charged lipids and in ionic solutions used as internal solutions. A high yield of GUVs was obtained for Chol/1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) samples with mixing ratios ranging from 0 to 2.5. We also succeeded in preparing GUVs from mixtures containing up to 60 mol% of the charged lipid 1-palmitoyl-2-oleoyl-sn-glycero-3-phospho-L-serine (POPS) and in NaCl solutions with low ionic strength (<25 mM). Full article
(This article belongs to the Section Membrane Fabrication and Characterization)
Show Figures

Figure 1

Figure 1
<p>Examples of Chol crystals formed during the dry film phase when using the traditional electroformation protocol. The observed crystal structures are in good agreement with those observed in the study of Park et al. [<a href="#B21-membranes-14-00215" class="html-bibr">21</a>] on the phases of Chol crystallization. The scale bar represents 50 µm.</p>
Full article ">Figure 2
<p>The modified protocol is for electroformation of GUVs from a damp lipid film. (<b>a</b>) The RSE method is used to produce an MLV solution. The RSE method is used to produce an MLV solution. An organic solvent (blue) containing lipids (red) is mixed with an aqueous solution (tube 1). The organic solvent is removed by vortexing the solution under vacuum in order to form MLVs (tube 3). (<b>b</b>) The suspension of MLVs is sonicated to produce LUVs. (<b>c</b>) A plasma cleaner is used to hydrophilize the ITO electrode. (<b>d</b>) The LUV suspension is deposited onto a plasma-cleaned ITO-coated glass and spin-coated to obtain a damp lipid film. (<b>e</b>) The coated electrode is used to assemble the electroformation chamber and connected to a voltage source to enable the growth of GUVs.</p>
Full article ">Figure 3
<p>The effect of sonication parameters on the size of produced LUVs. (<b>a</b>) The effect of sonication duration using different sonication amplitudes for MLVs produced from a mixture with a Chol/POPC ratio of 1.5. (<b>b</b>) The effect of sonication duration for different Chol/POPC mixing ratios at a sonication amplitude of 60%.</p>
Full article ">Figure 4
<p>The effect of increasing the Chol content on the electroformation of GUVs. (<b>a</b>) Size of GUVs as a function of the different Chol mixing ratios. The points and bars represent mean values and standard errors. The mean values were calculated by averaging the mean diameters from three independent samples for each concentration. (<b>b</b>) Size distribution densities for different Chol contents. Each distribution density represents 300 vesicles (100 vesicles from each of the three samples). (<b>c</b>) Fluorescence microscopy images for each sample. The scale bar represents 50 µm.</p>
Full article ">Figure 5
<p>Electroformation from damp lipid films using different concentrations of POPS. (<b>a</b>) GUV size as a function of POPS concentrations. The points and bars represent mean values and standard errors. Mean values were calculated by averaging the mean diameters of three independent samples for each concentration. (<b>b</b>) Size distribution densities of GUVs for different POPS concentrations. Each distribution density represents 300 vesicles (100 vesicles from each of the three samples). (<b>c</b>) Fluorescence microscopy images for each sample. The scale bar represents 50 µm.</p>
Full article ">Figure 6
<p>Electroformation from damp lipid films using saline solutions for the Chol/POPC mixture with a fixed Chol concentration of 10 mol%. (<b>a</b>) Size distribution of GUVs for different salt concentrations. Each distribution density represents 300 vesicles (100 vesicles from each of the three samples). (<b>b</b>) Fluorescence microscopy images for each sample. The scale bar represents 50 µm.</p>
Full article ">
20 pages, 4992 KiB  
Article
Membrane Activity and Viroporin Assembly for the SARS-CoV-2 E Protein Are Regulated by Cholesterol
by Marta V. Volovik, Zaret G. Denieva, Polina K. Gifer, Maria A. Rakitina and Oleg V. Batishchev
Biomolecules 2024, 14(9), 1061; https://doi.org/10.3390/biom14091061 - 26 Aug 2024
Viewed by 948
Abstract
The SARS-CoV-2 E protein is an enigmatic viral structural protein with reported viroporin activity associated with the acute respiratory symptoms of COVID-19, as well as the ability to deform cell membranes for viral budding. Like many viroporins, the E protein is thought to [...] Read more.
The SARS-CoV-2 E protein is an enigmatic viral structural protein with reported viroporin activity associated with the acute respiratory symptoms of COVID-19, as well as the ability to deform cell membranes for viral budding. Like many viroporins, the E protein is thought to oligomerize with a well-defined stoichiometry. However, attempts to determine the structure of the protein complex have yielded inconclusive results, suggesting several possible oligomers, ranging from dimers to pentamers. Here, we combined patch-clamp, confocal fluorescence microscopy on giant unilamellar vesicles, and atomic force microscopy to show that E protein can exhibit two modes of membrane activity depending on membrane lipid composition. In the absence or the presence of a low content of cholesterol, the protein forms short-living transient pores, which are seen as semi-transmembrane defects in a membrane by atomic force microscopy. Approximately 30 mol% cholesterol is a threshold for the transition to the second mode of conductance, which could be a stable pentameric channel penetrating the entire lipid bilayer. Therefore, the E-protein has at least two different types of activity on membrane permeabilization, which are regulated by the amount of cholesterol in the membrane lipid composition and could be associated with different types of protein oligomers. Full article
(This article belongs to the Special Issue Genetic and Structural Analyses of SARS-CoV-2 and Its Variants)
Show Figures

Figure 1

Figure 1
<p>Schematic structure of the SARS-CoV-2 E protein monomer with its H3 peptide and SUMO-1 N-terminal tag. The structure was combined from the TMD structure obtained by NMR (PDB 7K3G [<a href="#B26-biomolecules-14-01061" class="html-bibr">26</a>]), structures of the H2 and H3 helices predicted in [<a href="#B20-biomolecules-14-01061" class="html-bibr">20</a>], and SUMO-1 structure (PDB 1A5R).</p>
Full article ">Figure 2
<p>Typical fluorescence images of deformations of the GUV after the addition of 1 μM of the E protein in the vicinity of the GUV from lipid mixtures (<b>A</b>–<b>E</b>). The GUVs were labeled with 0.5 mol% of fluorescent dye Rho-DOPE. The numbers above each GUV indicate the time in min from the start of the protein addition. Scale bar is 10 µm.</p>
Full article ">Figure 3
<p>Typical fluorescence images of deformations of a GUV after the addition of 20 μM of the H3 peptide in the vicinity of the GUV from lipid mixture D. The GUVs were labeled with 0.5 mol% of fluorescent dye Rho-DOPE. The numbers above each GUV image indicate the time in min from the start of the peptide addition. Scale bar is 10 µm.</p>
Full article ">Figure 4
<p>Typical fluorescence images of the calcein leakage from the GUVs from lipid mixtures (<b>A</b>–<b>E</b>) after the addition of 1 μM E protein. The numbers above each GUV indicate the time in min from the start of the protein addition. Scale bar is 10 µm.</p>
Full article ">Figure 5
<p>Typical kinetics of fluorescence intensity changes inside the GUV after the addition of 1 μM E protein in the vicinity of the GUVs from lipid mixture A (orange), mixture B (black), mixture C (blue), mixture D (red), and mixture E (green); and after the addition of 20 μM H3 peptide in the vicinity of the GUVs from lipid mixture D (pink). The results are summarized from 3 experiments of 30 min each for each lipid mixture (error bars for each lipid mixture are shown as SD in corresponding translucent color).</p>
Full article ">Figure 6
<p>Typical fluorescence images of the calcein leakage from the GUVs from lipid mixture D after the addition of 20 μM H3 peptide. The numbers above each GUV indicate the time in min from the start of the protein addition. Scale bar is 10 µm.</p>
Full article ">Figure 7
<p>Histogram of the relative frequency of occurrence of conductance changes over the course of the experiments (12 min) for 1 µM of the E protein, summarized from 5 experiments for four lipid mixture A (orange), mixture B (black), mixture C (blue), and mixture D (red).</p>
Full article ">Figure 8
<p>Typical conductance traces of the three types of electrical signals obtained by the patch-clamp experiments: spike (<b>A</b>), multi-level (<b>B</b>), and square-top (<b>C</b>) signals. The total duration of each record displayed is 30 s.</p>
Full article ">Figure 9
<p>Histograms of the relative frequency of occurrence of each type of electrical signal, spike, multi-level, and square-top signals, over the course of the experiments (12 min) for 1 µM of the E protein, summarized from 5 experiments for lipid mixtures A–D (panels (<b>A</b>–<b>D</b>).</p>
Full article ">Figure 10
<p>Typical AFM images of the supported lipid bilayers of lipid mixtures A–D (panels (<b>A</b>–<b>D</b>), respectively) exposed to 100 nM E protein. White frame in panel D highlights an example of the smallest defect, which is further shown at higher resolution in <a href="#biomolecules-14-01061-f011" class="html-fig">Figure 11</a>. Image size is 3 × 3 μm<sup>2</sup>. Scale bar is 10 nm.</p>
Full article ">Figure 11
<p>High-resolution AFM image of an example of the smallest defect, highlighted by the white frame in <a href="#biomolecules-14-01061-f010" class="html-fig">Figure 10</a>D (<b>A</b>). Image size is 0.1 × 0.1 μm<sup>2</sup>. Scale bar is 10 nm. The profile is taken along the white line across the defect (<b>B</b>).</p>
Full article ">Figure 12
<p>Comparison of the characteristic values of the depths of the membrane defects formed in a supported lipid bilayer upon adsorption of 100 nM E protein for lipid mixture A (orange), mixture B (black), mixture C (blue), and mixture D (red).</p>
Full article ">Figure 13
<p>Schematic representation of the SARS-CoV-2 E protein activity depending on the cholesterol content: lipidic pore induced by the E protein dimer in a membrane with low (15%) or zero cholesterol (<b>A</b>); E protein pentameric channel in the membrane stabilized by high amount of cholesterol (30–40 mol%) (<b>B</b>).</p>
Full article ">
18 pages, 2713 KiB  
Article
Engineering Phosphatidylserine Containing Asymmetric Giant Unilamellar Vesicles
by Jake McDonough, Trevor A. Paratore, Hannah M. Ketelhohn, Bella C. DeCilio, Alonzo H. Ross and Arne Gericke
Membranes 2024, 14(9), 181; https://doi.org/10.3390/membranes14090181 - 23 Aug 2024
Viewed by 1270
Abstract
The plasma membrane lipid distribution is asymmetric, with several anionic lipid species located in its inner leaflet. Among these, phosphatidylserine (PS) plays a crucial role in various important physiological functions. Over the last decade several methods have been developed that allow for the [...] Read more.
The plasma membrane lipid distribution is asymmetric, with several anionic lipid species located in its inner leaflet. Among these, phosphatidylserine (PS) plays a crucial role in various important physiological functions. Over the last decade several methods have been developed that allow for the fabrication of large or giant unilamellar vesicles (GUVs) with an asymmetric lipid composition. Investigating the physicochemical properties of PS in such asymmetric lipid bilayers and studying its interactions with proteins necessitates the reliable fabrication of asymmetric GUVs (aGUVs) with a high degree of asymmetry that exhibit PS in the outer leaflet so that the interaction with peptides and proteins can be studied. Despite progress, achieving aGUVs with well-defined PS asymmetry remains challenging. Recently, a Ca2+-initiated hemifusion method has been introduced, utilizing the fusion of symmetric GUVs (sGUVs) with a supported lipid bilayer (SLB) for the fabrication of aGUVs. We extend this approach to create aGUVs with PS in the outer bilayer leaflet. Comparing the degree of asymmetry between aGUVs obtained via Ca2+ or Mg2+ initiated hemifusion of a phosphatidylcholine (PC) sGUVwith a PC/PS-supported lipid bilayer, we observe for both bivalent cations a significant number of aGUVs with near-complete asymmetry. The degree of asymmetry distribution is narrower for physiological salt conditions than at lower ionic strengths. While Ca2+ clusters PS in the SLB, macroscopic domain formation is absent in the presence of Mg2+. However, the clustering of PS upon the addition of Ca2+ is apparently too slow to have a negative effect on the quality of the obtained aGUVs. We introduce a data filtering method to select aGUVs that are best suited for further investigation. Full article
(This article belongs to the Special Issue Advances in Symmetric and Asymmetric Lipid Membranes)
Show Figures

Figure 1

Figure 1
<p>Time evolution of Ca<sup>2+</sup>- or Mg<sup>2+</sup>-induced domain formation in DOPC/POPS/TF-PS (69.9/30/0.1 mol%) SLBs. (<b>A</b>) 63× magnification SLB images at the same location just before and at 20 to 180 min after Ca<sup>2+</sup> was introduced to the aqueous phase. A large domain rich in TF-PS appears as a bright green spot that increases in size over time. Images were taken at a low detector gain and are digitally brightened. (<b>B</b>) 10× magnification images of the SLB from panel (<b>A</b>) at two separate areas 180 or more minutes after adding Ca<sup>2+</sup>. With this lower magnification we can observe the formation of many such domains as presented in Panel (<b>A</b>). (<b>C</b>,<b>D</b>) Same as in panels (<b>A</b>,<b>B</b>), but with 3 mM Mg<sup>2+</sup> present in the aqueous phase in place of Ca<sup>2+</sup>. After the addition of the bivalent cation, the buffer composition is 3 mM MCl<sub>2</sub> (M= Ca<sup>2+</sup> or Mg<sup>2+</sup>), 102 mM NaCl, 92 mM KCl, 3 mM sucrose, 25 mM HEPES, pH = 7.4.</p>
Full article ">Figure 2
<p>Comparison of sGUV and corresponding aGUV images. (<b>A</b>) Images of a DOPC/DiD (99.9/0.1 mol%) and DOPC/POPS/TF-PS (69.9/30/0.1 mol%) sGUV. sGUVs contain 100 mM sucrose, 100 mM NaCl, 50 mM KCl, 25 mM HEPES, pH = 7.4 and are imaged in a 100 mM glucose buffer of equal ionic strength, pH, and osmolality. (<b>B</b>) Images of aGUVs formed via hemifusion with Ca<sup>2+</sup> or Mg<sup>2+</sup>. For each aGUV, the DiD (red) and TF (green) channels are shown separately. aGUVs are in 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 100 mM NaCl, 84 mM KCl, 2 mM sucrose, 25 mM HEPES, pH = 7.4. (<b>C</b>) Composite images of Z-stacks taken of a Ca<sup>2+</sup>- and Mg<sup>2+</sup>-formed aGUV in the same buffer as those in panel (<b>A</b>). Please note that the reduced intensity around the equator line in the DID (red) channel is due to polarization effects. Each fusion experiment was repeated four times. “DiD” sGUV: DOPC/DID 99.9/0.1 mol%; “TF” sGUV: DOPC/POPS/TF-PS 69.9/30/0.1 mol%.</p>
Full article ">Figure 3
<p>Intensity comparisons between sGUVs and aGUVs. (<b>A</b>) Intensities of Ca<sup>2+</sup>- and Mg<sup>2+</sup>-formed aGUVs compared to those of sGUVs with compositions equal to the theoretical inner leaflet composition (DiD sGUVs, graphed in red) or theoretical outer leaflet composition (TF sGUVs, graphed in green) of the aGUVs (all graphed in blue). Both symmetric and asymmetric GUVs contain 100 mM sucrose, 100 mM NaCl, 50 mM KCl, 25 mM HEPES, pH = 7.4. Unlike sGUVs, which are imaged in an isotonic glucose buffer of equal ionic strength and pH, aGUVs are imaged in a solution of 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 100 mM NaCl, 84 mM KCl, 2 mM sucrose, 25 mM HEPES, pH = 7.4. All intensities are normalized by the average of the corresponding sGUVs. Averages are shown as squares on each boxplot. (<b>B</b>) Same as panel (<b>A</b>), but after removing all range-failing aGUVs (spread of aGUV intensities is half of the spread observed for the sGUVs). “DiD” sGUV: DOPC/DID 99.9/0.1 mol%; “TF” sGUV: DOPC/POPS/TF-PS 69.9/30/0.1 mol%.</p>
Full article ">Figure 4
<p>Correlation between DID and TF-PS outer leaflet exchange. (<b>A</b>) Outer leaflet DiD % exchange vs. TF % exchange for Ca<sup>2+</sup>- and Mg<sup>2+</sup>-formed aGUVs. aGUVs contain 100 mM sucrose, 100 mM NaCl, 50 mM KCl, 25 mM HEPES, pH = 7.4 and are imaged in 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 100 mM NaCl, 84 mM KCl, 2 mM sucrose, 25 mM HEPES, pH = 7.4. Error bars are omitted for clarity. (<b>B</b>) Same plot as in panel (<b>A</b>), but after removing all range-failing aGUVs. Error bars are calculated via a propagation of uncertainty using standard deviations.</p>
Full article ">Figure 5
<p>Diameter comparisons between sGUVs and aGUVs. (<b>A</b>) Average diameter of all Ca<sup>2+</sup>- and Mg<sup>2+</sup>-formed aGUVs, those that are range-passing, and those that are range-failing compared to that of sGUVs whose compositions match the theoretical inner leaflet composition (DiD sGUVs) or theoretical outer leaflet composition (TF sGUVs) of the aGUVs. Error bars show standard deviations. All vesicles contain 200 mM sucrose, 5 mM HEPES, pH 7.4. sGUVs are imaged in 200 mM glucose, 5 mM HEPES, pH = 7.4. aGUVs are measured in low-salt buffer with a final composition of 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 49 mM NaCl, 34 mM KCl, 5 mM sucrose, 25 mM HEPES, pH = 7.4. (<b>B</b>) Same as panel (<b>A</b>), but for vesicles in physiological ionic strength salt buffer. All vesicles contain 100 mM sucrose, 100 mM NaCl, 50 mM KCl, 25 mM HEPES, pH = 7.4. sGUVs are imaged in an isotonic glucose buffer of equal ionic strength and pH. aGUVs are imaged in 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 100 mM NaCl, 84 mM KCl, 2 mM sucrose, 25 mM HEPES, pH = 7.4. “DiD” sGUV: DOPC/DID 99.9/0.1 mol%; “TF” sGUV: DOPC/POPS/TF-PS 69.9/30/0.1 mol%.</p>
Full article ">Figure A1
<p>Time evolution of Ca<sup>2+</sup> or Mg<sup>2+</sup> induced domain formation in DOPC/POPS/TF-PS (69.9/30/0.1 mol%) SLBs at low salt concentration. (<b>A</b>) 63× magnification SLB images at the same location just before and at 20 to 180 min after Ca<sup>2+</sup> was introduced to the aqueous phase. A large domain rich in TF-PS appears as a bright green spot that increases in size over time. Images were taken at a low detector gain and digitally brightened. (<b>B</b>) 10× magnification images of the SLB from panel A at two separate areas 180 or more minutes after adding Ca<sup>2+</sup>. With this lower magnification we can observe the formation of many such domains as presented in Panel (<b>A</b>). (<b>C</b>,<b>D</b>) Same as in panels (<b>A</b>,<b>B</b>), but with 3 mM Mg<sup>2+</sup> present in the aqueous phase in place of Ca<sup>2+</sup>. After the addition of the bivalent cation, the buffer composition is 3 mM MCl<sub>2</sub> (M = Ca<sup>2+</sup> or Mg<sup>2+</sup>), 56 mM NaCl, 36 mM KCl, 7 mM sucrose, 24 mM HEPES, pH = 7.4.</p>
Full article ">Figure A2
<p>Images of sGUVs and aGUVs in low-salt buffer. (<b>A</b>) Images of a DOPC/DiD (99.9/0.1 mol%) and DOPC/POPS/TF-PS (69.9/30/0.1 mol%) All vesicles contain 200 mM sucrose, 5 mM HEPES, pH 7.4. sGUVs are imaged in 200 mM glucose, 5 mM HEPES, pH = 7.4. (<b>B</b>) Images of aGUVs formed via hemifusion with Ca<sup>2+</sup> or Mg<sup>2+</sup>. For each aGUV, the DiD (red) and TF (green) channels are shown separately. aGUVs are measured in low-salt buffer with a final composition of 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 49 mM NaCl, 34 mM KCl, 5 mM sucrose, 25 mM HEPES, pH = 7.4. (<b>C</b>) Composite images of Z-stacks taken of a Ca<sup>2+</sup>- and Mg<sup>2+</sup>-formed aGUV in the same buffer as those in panel (<b>A</b>). Please note that the reduced intensity around the equator line in the DID (red) channel is due to polarization effects. This fusion experiment was performed eight times with Ca<sup>2+</sup> and 18 times with Mg<sup>2+</sup>.</p>
Full article ">Figure A3
<p>Intensity comparisons between sGUVs and aGUVs in low-salt buffer. (<b>A</b>) Intensities of Ca<sup>2+</sup>- and Mg<sup>2+</sup>-formed aGUVs compared to those of sGUVs with compositions equal to the theoretical inner leaflet composition (DiD sGUVs) or theoretical outer leaflet composition (TF sGUVs) of the aGUVs. All vesicles contain 200 mM sucrose, 5 mM HEPES, pH 7.4. sGUVs are imaged in 200 mM glucose, 5 mM HEPES, pH = 7.4. aGUVs are measured in low-salt buffer with a final composition of 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 49 mM NaCl, 34 mM KCl, 5 mM sucrose, 25 mM HEPES, pH = 7.4. All intensities are normalized by the average of the corresponding sGUVs. Averages are shown as squares on each boxplot. (<b>B</b>) Same as panel (<b>A</b>), but after removing all range-failing aGUVs (spread of aGUV intensities is half of the spread observed for the sGUVs).</p>
Full article ">Figure A4
<p>Correlation between DID and TF outer leaflet exchange in low-salt buffer. (<b>A</b>) Outer leaflet DiD % exchange vs. TF % exchange for Ca<sup>2+</sup>- and Mg<sup>2+</sup>-formed aGUVs. aGUVs are measured in low-salt buffer with a final composition of 4 mM CaCl<sub>2</sub> (or MgCl<sub>2</sub>), 6 mM Na<sub>2</sub>EDTA, 49 mM NaCl, 34 mM KCl, 5 mM sucrose, 25 mM HEPES, pH = 7.4. Error bars are omitted for clarity. (<b>B</b>) Same plot as in panel (<b>A</b>), but after removing all range-failing aGUVs. Error bars are calculated via a propagation of uncertainty using standard deviations.</p>
Full article ">
32 pages, 7726 KiB  
Review
Droplet Microfluidics for High-Throughput Screening and Directed Evolution of Biomolecules
by Goran T. Vladisavljević
Micromachines 2024, 15(8), 971; https://doi.org/10.3390/mi15080971 - 29 Jul 2024
Viewed by 2207
Abstract
Directed evolution is a powerful technique for creating biomolecules such as proteins and nucleic acids with tailor-made properties for therapeutic and industrial applications by mimicking the natural evolution processes in the laboratory. Droplet microfluidics improved classical directed evolution by enabling time-consuming and laborious [...] Read more.
Directed evolution is a powerful technique for creating biomolecules such as proteins and nucleic acids with tailor-made properties for therapeutic and industrial applications by mimicking the natural evolution processes in the laboratory. Droplet microfluidics improved classical directed evolution by enabling time-consuming and laborious steps in this iterative process to be performed within monodispersed droplets in a highly controlled and automated manner. Droplet microfluidic chips can generate, manipulate, and sort individual droplets at kilohertz rates in a user-defined microchannel geometry, allowing new strategies for high-throughput screening and evolution of biomolecules. In this review, we discuss directed evolution studies in which droplet-based microfluidic systems were used to screen and improve the functional properties of biomolecules. We provide a systematic overview of basic on-chip fluidic operations, including reagent mixing by merging continuous fluid streams and droplet pairs, reagent addition by picoinjection, droplet generation, droplet incubation in delay lines, chambers and hydrodynamic traps, and droplet sorting techniques. Various microfluidic strategies for directed evolution using single and multiple emulsions and biomimetic materials (giant lipid vesicles, microgels, and microcapsules) are highlighted. Completely cell-free microfluidic-assisted in vitro compartmentalization methods that eliminate the need to clone DNA into cells after each round of mutagenesis are also presented. Full article
(This article belongs to the Special Issue μ-TAS: A Themed Issue in Honor of Professor Andreas Manz)
Show Figures

Figure 1

Figure 1
<p>Typical HTS strategies based on fluorescence-based assays [<a href="#B6-micromachines-15-00971" class="html-bibr">6</a>]. The droplet-based screening strategies shown in (<b>c</b>,<b>d</b>) will be discussed in this review.</p>
Full article ">Figure 2
<p>Fabrication methods for lab-on-a-chip devices.</p>
Full article ">Figure 3
<p>Main strategies for on-chip mixing of cells and reagent solutions.</p>
Full article ">Figure 4
<p>Main microfluidic strategies of droplet formation (DP = dispersed phase; CP = continuous phase). Adapted from [<a href="#B82-micromachines-15-00971" class="html-bibr">82</a>].</p>
Full article ">Figure 5
<p>Strategies for on-chip droplet incubation: (<b>a</b>) serpentine channel [<a href="#B90-micromachines-15-00971" class="html-bibr">90</a>]; (<b>b</b>) wide channel with and without constrictions [<a href="#B84-micromachines-15-00971" class="html-bibr">84</a>]; (<b>c</b>) rectangular chamber with a pneumatic valve [<a href="#B85-micromachines-15-00971" class="html-bibr">85</a>]; (<b>d</b>) wavy chamber with pillars along the side walls to remove the excess oil [<a href="#B86-micromachines-15-00971" class="html-bibr">86</a>]; (<b>e</b>) flow resistance-based microwells for trapping droplets [<a href="#B87-micromachines-15-00971" class="html-bibr">87</a>].</p>
Full article ">Figure 6
<p>Fluorescence-activated droplet sorting (FADS) in microfluidic chips based on the following phenomena: (<b>a</b>) dielectrophoresis (DEP); (<b>b</b>) electrocoalescence [<a href="#B101-micromachines-15-00971" class="html-bibr">101</a>].</p>
Full article ">Figure 7
<p>Microfluidic compartmentalization of cell lysate in W/O emulsions [<a href="#B111-micromachines-15-00971" class="html-bibr">111</a>,<a href="#B112-micromachines-15-00971" class="html-bibr">112</a>].</p>
Full article ">Figure 8
<p>The continuous platform for the directed evolution of alditol oxidase [<a href="#B113-micromachines-15-00971" class="html-bibr">113</a>].</p>
Full article ">Figure 9
<p>Distribution of cells in monodisperse droplets: (<b>a</b>) <math display="inline"> <semantics> <mrow> <mi>P</mi> <mfenced separators="|"> <mrow> <mi>λ</mi> <mo>,</mo> <mi>k</mi> </mrow> </mfenced> </mrow> </semantics> </math> vs. <math display="inline"> <semantics> <mrow> <mi>λ</mi> </mrow> </semantics> </math> at <math display="inline"> <semantics> <mrow> <mi>k</mi> </mrow> </semantics> </math> = 1; (<b>b</b>) <math display="inline"> <semantics> <mrow> <mi>P</mi> <mfenced separators="|"> <mrow> <mi>λ</mi> <mo>,</mo> <mi>k</mi> </mrow> </mfenced> </mrow> </semantics> </math> vs. <math display="inline"> <semantics> <mrow> <mi>k</mi> </mrow> </semantics> </math> at three different <math display="inline"> <semantics> <mrow> <mi>λ</mi> </mrow> </semantics> </math> values. Cells are randomly distributed in the feed stream.</p>
Full article ">Figure 10
<p>Evolution of catalytic RNAs (ribosomes) [<a href="#B128-micromachines-15-00971" class="html-bibr">128</a>] and light-up RNA aptamers [<a href="#B129-micromachines-15-00971" class="html-bibr">129</a>] by microfluidic-assisted in vitro compartmentalization (μIVC): (<b>a</b>) gene compartmentalization in W/O droplets; (<b>b1</b>) pairwise fusion of droplets containing amplified genes and IVT mixture/fluorogen, respectively; (<b>b2</b>) pairwise fusion of droplets containing amplified genes and IVT mixture; (<b>c</b>) picoinjection of activity assay mixture; (<b>d</b>) droplet sorting by dielectrophoretic-based FADS. Off-chip operations are shown by grey boxes.</p>
Full article ">Figure 11
<p>The HTS of single-cell lysate using double emulsion produced in microfluidic chips: (<b>i</b>) the generation of gene libraries from an enzyme-encoding plasmid; (<b>ii</b>) the introduction of plasmid DNA into <span class="html-italic">E. coli</span>; (<b>iii</b>) the compartmentalization of single cells in droplets together with substrate and lysis agents; (<b>iv</b>) cell lysis to allow enzymes to be released from the cytoplasm and react with the product; (<b>v</b>) droplet incubation for a desired period; (<b>vi</b>) the formation of double emulsions in a second microfluidic device; (<b>vii</b>) the sorting of enzyme variants in a standard flow cytometer [<a href="#B134-micromachines-15-00971" class="html-bibr">134</a>].</p>
Full article ">Figure 12
<p>Directed evolution of polymerase using double emulsion produced in a microfluidic device: (<b>a</b>): a water-in-oil (W/O) droplet that contains a single <span class="html-italic">E. coli</span> cell and a fluorescence-based polymerase activity assay; (<b>b</b>) a W/O droplet containing a functional polymerase that extends a primer–template complex with RNA and releases a quencher probe; (<b>c</b>) a W/O/W droplet containing a non-functional polymerase [<a href="#B138-micromachines-15-00971" class="html-bibr">138</a>].</p>
Full article ">Figure 13
<p>The preparation of GUVs by the phase transfer method: (<b>a</b>) The off-chip transfer of lipid monolayer-coated water-in-oil droplets from the oil phase into an outer aqueous solution by density difference. In this case, the density of the oil phase is smaller than the densities of both aqueous solutions [<a href="#B140-micromachines-15-00971" class="html-bibr">140</a>]; (<b>b</b>) The on-chip transfer of droplets across the interface via a microfabricated post placed in a microchannel [<a href="#B144-micromachines-15-00971" class="html-bibr">144</a>]. In both cases, lipids residing at the interface are deposited on the droplet-adsorbed lipids, forming the outer- and inner-leaflets of the membrane bilayer.</p>
Full article ">Figure 14
<p>Microfluidic chips for production of GUVs: (<b>a</b>) on-chip emulsion transfer [<a href="#B144-micromachines-15-00971" class="html-bibr">144</a>,<a href="#B145-micromachines-15-00971" class="html-bibr">145</a>]; (<b>b</b>) double emulsion templating using two consecutive flow-focusing junctions, followed by off-chip de-wetting [<a href="#B147-micromachines-15-00971" class="html-bibr">147</a>]; (<b>c</b>) de-wetting-induced formation of GUV; (<b>d</b>) flow-focusing junction, followed by a multi-V-shaped microfluidic droplet splitting channel network [<a href="#B148-micromachines-15-00971" class="html-bibr">148</a>].</p>
Full article ">Figure 15
<p>The formation of GUVs by the charge-induced fusion of internally entrapped small unilamellar vesicles (SUVs) at the surfactant-loaded droplet interface [<a href="#B157-micromachines-15-00971" class="html-bibr">157</a>].</p>
Full article ">Figure 16
<p>A directed evolution strategy for screening glucose oxidase (GOx) enzyme variants in a homogeneous solution. GOx expressed on the yeast surface triggers the crosslinking of alginate grafted with phenol moieties and aminofluorescein using hydrogen peroxide in the presence of horse radish peroxidase (HRP) and glucose (Glc) [<a href="#B163-micromachines-15-00971" class="html-bibr">163</a>].</p>
Full article ">Figure 17
<p>Directed evolution in (gel core)-(polyelectrolyte shell) (GSB) beads: (1) Cel lysis; (2) Enzymatic reaction (3) Droplet gelation upon cooling; (4) Mixing gel beads with PAH-containing W/O emulsion; (5) Surfactant displacement and polyelectrolyte film formation; (6) Shell decomposition at high pH and the release of microgel content [<a href="#B162-micromachines-15-00971" class="html-bibr">162</a>].</p>
Full article ">Figure 18
<p>Microfluidic HTS of filamentous fungi within core–shell microcapsules consisting of aqueous cores and hydrogel shells [<a href="#B172-micromachines-15-00971" class="html-bibr">172</a>].</p>
Full article ">
21 pages, 4551 KiB  
Article
Benefits of Combined Fluorescence Lifetime Imaging Microscopy and Fluorescence Correlation Spectroscopy for Biomedical Studies Demonstrated by Using a Liposome Model System
by Kristina Bruun, Hans-Gerd Löhmannsröben and Carsten Hille
Biophysica 2024, 4(2), 207-226; https://doi.org/10.3390/biophysica4020015 - 25 Apr 2024
Viewed by 1250
Abstract
Drug delivery systems play a pivotal role in targeted pharmaceutical transport and controlled release at specific sites. Liposomes, commonly used as drug carriers, constitute a fundamental part of these systems. Moreover, the drug–liposome model serves as a robust platform for investigating interaction processes [...] Read more.
Drug delivery systems play a pivotal role in targeted pharmaceutical transport and controlled release at specific sites. Liposomes, commonly used as drug carriers, constitute a fundamental part of these systems. Moreover, the drug–liposome model serves as a robust platform for investigating interaction processes at both cellular and molecular levels. To advance our understanding of drug carrier uptake mechanisms, we employed fluorescence lifetime imaging microscopy (FLIM) and fluorescence correlation spectroscopy (FCS), leveraging the unique benefits of two-photon (2P) excitation. Our approach utilized giant unilamellar vesicles (GUVs) as a simplified model system for cell membranes, labelled with the amphiphilic fluorescent dye 3,3′-dioctadecyloxa-carbocyanine (DiOC18(3)). Additionally, large unilamellar vesicles (LUVs) functioned as a drug carrier system, incorporating the spectrally distinct fluorescent sulforhodamine 101 (SRh101) as a surrogate drug. The investigation emphasized the diverse interactions between GUVs and LUVs based on the charged lipids employed. We examined the exchange kinetics and structural alterations of liposome carriers during the uptake process. Our study underscores the significance of employing 2P excitation in conjunction with FLIM and FCS. This powerful combination offers a valuable methodological approach for studying liposome interactions, positioning them as an exceptionally versatile model system with a distinct technical advantage. Full article
Show Figures

Figure 1

Figure 1
<p>Spectroscopic properties of the used dyes. (<b>a</b>) Normalised fluorescence spectra of the LUV membrane-bound DiOC<sub>18</sub>(3) (green line) and LUV-encapsulated SRh101 (red line) in phosphate-buffered saline (pH 7.0). The transmission characteristics of dichroic mirror (FF605) and filters (BP514/44 and BP700/75) used for the spectral separation of the emission signal are also indicated (dotted lines). (<b>b</b>) The 2P fluorescence excitation action cross-section spectra of DiOC<sub>18</sub>(3) (green dots) and SRh101 (red dots) in LUVs (means ± SEM, <span class="html-italic">N</span> = 3). Fluorescein in NaOH (pH 13) and rhodamine 6G in methanol served as references [<a href="#B24-biophysica-04-00015" class="html-bibr">24</a>]. The 2P action cross-section is the product of the fluorescence quantum yield <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Φ</mi> </mrow> <mrow> <mi mathvariant="normal">F</mi> </mrow> </msub> <mtext> </mtext> </mrow> </semantics></math> and the absolute 2P absorption cross-section <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math> [<a href="#B25-biophysica-04-00015" class="html-bibr">25</a>]; 1 GM = 10<sup>−50</sup> cm<sup>4</sup> s/photon [<a href="#B29-biophysica-04-00015" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Comparative in vitro characterisation of the fluorescent dyes SRh101 and DiOC<sub>18</sub>(3) free in solution and in LUVs. Representative fluorescence decay curves of (<b>a</b>) SRh101 in PBS (black line) and SRh101-encapsulated LUVs (red line) as well as (<b>b</b>) DiOC<sub>18</sub>(3) in methanol (MeOH, black line) and DiOC<sub>18</sub>(3) incorporated into LUVs (red line), were recorded at <span class="html-italic">λ</span><sub>ex,2P</sub> = 780 nm; the corresponding biexponential reconvolution fits (black dashed lines) of the SRh101-encapsulated LUVs and DiOC<sub>18</sub>(3) incorporated into LUVs are also shown. The weighted residuals of mono-, bi- and triexponential fits and their corresponding <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>χ</mi> </mrow> <mrow> <mi mathvariant="normal">R</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msubsup> </mrow> </semantics></math> values are displayed below. IRF: instrument response function (grey line).</p>
Full article ">Figure 3
<p>2P-FCS data analysis obtained by measuring SRh101 in phosphate-buffered saline (PBS) and in LUVs. Normalised fluorescence autocorrelation curves and the corresponding weighted residuals of free SRh101 (black circles) and SRh101-encapsulated LUVs (red circles) fitted by using a two-component diffusion model according to Equation (5) (solid lines) [<a href="#B29-biophysica-04-00015" class="html-bibr">29</a>,<a href="#B30-biophysica-04-00015" class="html-bibr">30</a>]. The curves were normalized by multiplying with the total number of fluorescent molecules in the detection volume, <span class="html-italic">N</span>, derived from the fits (<span class="html-italic">G</span>(0) = 1/<span class="html-italic">N</span>).</p>
Full article ">Figure 4
<p>Interaction of cationic LUVs (DOPC:DOPE:DOTAP, 1:1:2) containing SRh101 with an immobilised neutral GUV (DOPC:DOPE, 3:1) labelled with 0.1 mol% DiOC<sub>18</sub>(3). Representative 2P-FLIM images recorded 30 min after the addition of the LUV solution showed the fluorescence signals from DiOC<sub>18</sub>(3) in the green detection channel ((<b>left</b>), BP 514/44) and from SRh101 in the red detection channel ((<b>middle</b>), BP 700/75), as well as merged images of DiOC<sub>18</sub>(3) and SRh101 direct excitation and emission channels (<b>right</b>). The images represent an equatorial cross-section of the GUV. Recording parameters: image acquisition time~60 s, pixel dwell time 2.3 ms/pixel.</p>
Full article ">Figure 5
<p>Analysis of the LUV-GUV interaction by 2P-FLIM recordings. GUVs contained 0.1 mol% DiOC<sub>18</sub>(3) and were composed of (<b>a</b>) DOPC:DOPE (3:1), (<b>b</b>) DOPC:DOPE:DOPS (8:3:2) or (<b>c</b>) DOPC:DOPS (3:1). GUVs were treated for 120 min with SRh101-encapsulated and positively charged LUVs with the lipid composition DOPC:DOTAP (1:1). Representative 2P-FLIM images are overlays of the fluorescence signals from the green (BP 514/44) and red (BP 700/75) detection channels at distinct time points. The results are representative for <span class="html-italic">N</span> = 3–4 independent repeat measurements.</p>
Full article ">Figure 6
<p>Analysis of the LUV-GUV interactions by 2P-FLIM recordings. GUVs contained 0.1 mol% DiOC<sub>18</sub>(3) and were composed of (<b>a</b>) DOPC:DOPE (3:1), (<b>b</b>) DOPC:DOPE:DOPS (8:3:2) and (<b>c</b>) DOPC:DOPS (3:1). GUVs were treated for 120 min with SRh101-encapsulated and positively charged LUVs with the lipid composition DOPC:DOPE:DOTAP (1:1:2). Representative 2P-FLIM images are overlays of the fluorescence signals in the green (BP 514/44) and red (BP 700/75) detection channels at distinct time points. The results are representative for <span class="html-italic">N</span> = 3–4 independent repeat measurements.</p>
Full article ">Figure 7
<p>Fluorescence recordings of LUVs obtained outside the GUVs in the surrounding medium. Representative 2P-excited fluorescence autocorrelation curves (<b>a</b>) and 2P-excited fluorescence decay curves (<b>b</b>) obtained at distinct time points during the incubation of an immobilised anionic GUV (DOPC:DOPE:DOPS, 8:3:2) with cationic LUVs (DOPC:DOTAP, 1:1) for 120 min. The colour code for (<b>a</b>) and (<b>b</b>) is defined within the legend of (<b>a</b>). The measuring point in the medium with immobilized GUV and added LUVs is shown schematically in (<b>b</b>). Determined diffusion times (<b>c</b>), concentrations of SRh101 in LUVs, (<b>d</b>) and intensity-weighted average decay times (<b>e</b>) vs. the measurement time; means ± SEM of <span class="html-italic">N</span> = 6 (<b>c</b>,<b>d</b>) and <span class="html-italic">N</span> = 10–13 (<b>e</b>).</p>
Full article ">Figure 8
<p>Fluorescence recordings of LUVs obtained inside the GUV lumen. Representative 2P-excited fluorescence autocorrelation curves (<b>a</b>) and 2P-excited fluorescence decay curves (<b>b</b>) obtained at distinct time points during the incubation of an immobilised anionic GUV (DOPC:DOPE:DOPS, 8:3:2) with cationic LUVs (DOPC:DOTAP, 1:1) for 120 min. The colour code for (<b>a</b>) and (<b>b</b>) is defined within the legend of (<b>a</b>). The measuring point in the GUV lumen and added LUVs are shown schematically in (<b>b</b>). Determined diffusion times (<b>c</b>), the concentrations of SRh101 in LUVs (<b>d</b>) and intensity-weighted average decay times (<b>e</b>) vs. the measurement time; means ± SEM of <span class="html-italic">N</span> = 6 (<b>c</b>,<b>d</b>) and <span class="html-italic">N</span> = 12 (<b>e</b>).</p>
Full article ">Figure 9
<p>Analysis of the internalisation of SRh101-encapsulated LUVs into the GUV interior. (<b>a</b>) Internalisation efficiency of DOPC:DOTAP (1:1) LUVs (red columns) and DOPC:DOPE:DOTAP (1:1:2) LUVs (black columns) into the GUVs of three different lipid compositions determined after a 60 min incubation period; means ± SEM of <span class="html-italic">N</span> = 3–4. (<b>b</b>) Time-dependent relative fluorescence intensity changes inside the lumen of neutral GUVs (DOPC:DOPE, 3:1) recorded during an incubation period of 120 min for DOPC:DOTAP (1:1) LUVs (red) and DOPC:DOPE:DOTAP (1:1:2) LUVs (black). Intensity changes relative to the starting point (<span class="html-italic">t</span> = 0) were analysed from 2P-FLIM images recorded in the red detection channel; means ± SEM of <span class="html-italic">N</span> = 4; data were fit to the first-order kinetics model [<a href="#B56-biophysica-04-00015" class="html-bibr">56</a>].</p>
Full article ">
11 pages, 6992 KiB  
Article
Glass Microdroplet Generator for Lipid-Based Double Emulsion Production
by Alessandra Zizzari and Valentina Arima
Micromachines 2024, 15(4), 500; https://doi.org/10.3390/mi15040500 - 5 Apr 2024
Viewed by 1449
Abstract
Microfluidics offers a highly controlled and reproducible route to synthesize lipid vesicles. In recent years, several microfluidic approaches have been introduced for this purpose, but double emulsions, such as Water-in-Oil-in-Water (W/O/W) droplets, are preferable to produce giant vesicles that are able to maximize [...] Read more.
Microfluidics offers a highly controlled and reproducible route to synthesize lipid vesicles. In recent years, several microfluidic approaches have been introduced for this purpose, but double emulsions, such as Water-in-Oil-in-Water (W/O/W) droplets, are preferable to produce giant vesicles that are able to maximize material encapsulation. Flow focusing (FF) is a technique used to generate double emulsion droplets with high monodispersity, a controllable size, and good robustness. Many researchers use polydimethylsiloxane as a substrate material to fabricate microdroplet generators, but it has some limitations due to its hydrophobicity, incompatibility with organic solvents, and the molecular adsorption on the microchannel walls. Thus, specific surface modification and functionalization steps, which are uncomfortable to perform in closed microchannels, are required to overcome these shortcomings. Here, we propose glass as a material to produce a chip with a six-inlet junction geometry. The peculiar geometry and the glass physicochemical properties allow for W/O/W droplet formation without introducing microchannel wall functionalization and using a variety of reagents and organic solvents. The robust glass chip can be easily cleaned and used repeatedly, bringing advantages in terms of cost and reproducibility in emulsion preparation. Full article
(This article belongs to the Special Issue μ-TAS: A Themed Issue in Honor of Professor Andreas Manz)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The device scheme with a single six-channel FF junction. The reported measurements are the nominal microchannels’ widths; 6 μm is the orifice nominal width. (<b>b</b>) The optical image of the junction area where the orifice, indicated by the red arrow, is clearly visible.</p>
Full article ">Figure 2
<p>(<b>a</b>) A scheme of double emulsion droplet production and collection in a vial where the de-wetting process takes place. We, LO, and Wi are the external aqueous, oil with lipids, and internal aqueous phases, respectively. (<b>b</b>) De-wetting steps: from double emulsion to giant vesicle (GV) generation.</p>
Full article ">Figure 3
<p>A “tandem system” device. (<b>a</b>) The sketch process: Two flow focusing chips were connected using a microtube. In the first droplet generator (Chip1), Step1 takes place, and Water-in-Oil (Wi/LO) droplets are produced. These droplets are the inputs of Step2 in the second droplet generator (Chip2), where they are sheared by the outer aqueous phase (We), thus generating double emulsions. (<b>b</b>) A glass “tandem system”: Without an adequate functionalization of the post-junction area, the Wi interacts with the hydrophilic walls, thereby resulting in the formation of large plugs indicated by green arrows. (<b>c</b>) The large plugs are visible in the connecting tube and, since they are inlets of the second chip, they compromise the success of the whole double emulsion formation process.</p>
Full article ">Figure 4
<p>Fluids’ behaviors in microchannels observed by using low-resolution camera. (<b>a</b>) Co-flow work conditions (Wi = 5 μL/min, LO = 5 μL/min, and We = 50 μL/min). (<b>b</b>) Flow rate conditions for double emulsion droplet generation (Wi = 0.1 μL/min, LO = 0.1 μL/min, and We = 4 μL/min). Black arrows indicate Wi and LO phases’ meniscus. Oil phase enveloping Wi is sheared into droplets by We.</p>
Full article ">Figure 5
<p>Fluorescence images of sample produced at Wi = 0.1 μL/min, LO = 0.1 μL/min, and We = 4 μL/min. Images were acquired under TRITC filter (<b>a</b>) and FITC filter (<b>b</b>). Both images show mixture of oil droplet (1), double emulsion droplet (2), GV (3), and partially de-wetting droplet (4). Legend is reported in (<b>c</b>). (<b>d</b>) GVs’ size distribution was obtained from a statistical analysis on several samples produced under same experimental conditions.</p>
Full article ">Figure 6
<p>Optical images of the de-wetting phenomenon acquired at different times are shown in (<b>a</b>,<b>b</b>). The starting step is the formation of an oil pocket inside the double emulsion (some are indicated with white arrows), and then the oil drop protrudes (some are indicated with yellow arrows) until its final detachment, thus generating a giant vesicle (GV) (some are indicated with green arrows).</p>
Full article ">Figure 7
<p>Images of sample produced at Wi = 1 μL/min, LO = 0.1 μL/min, and We = 1.5 μL/min. Images show only GVs since complete de-wetting took place for almost all double emulsion droplets. (<b>a</b>) Optical image of vesicles: oil with lipid residue remains inside some GVs. (<b>b</b>) Image of GVs acquired under TRITC filter. (<b>c</b>) GV size distribution. (<b>d</b>) GVs are quite stable even after five days, as shown in red fluorescence image. Here, lipid residues attached to GVs are clearly visible.</p>
Full article ">
12 pages, 1674 KiB  
Article
Electroformation of Giant Unilamellar Vesicles from Damp Lipid Films with a Focus on Vesicles with High Cholesterol Content
by Ivan Mardešić, Zvonimir Boban and Marija Raguz
Membranes 2024, 14(4), 79; https://doi.org/10.3390/membranes14040079 - 27 Mar 2024
Cited by 1 | Viewed by 1623
Abstract
Giant unilamellar vesicles (GUVs) are membrane models used to study membrane properties. Electroformation is one of the methods used to produce GUVs. During electroformation protocol, dry lipid film is formed. The drying of the lipid film induces the cholesterol (Chol) demixing artifact, in [...] Read more.
Giant unilamellar vesicles (GUVs) are membrane models used to study membrane properties. Electroformation is one of the methods used to produce GUVs. During electroformation protocol, dry lipid film is formed. The drying of the lipid film induces the cholesterol (Chol) demixing artifact, in which Chol forms anhydrous crystals which do not participate in the formation of vesicles. This leads to a lower Chol concentration in the vesicle bilayers compared to the Chol concentration in the initial lipid solution. To address this problem, we propose a novel electroformation protocol that includes rapid solvent exchange (RSE), plasma cleaning, and spin-coating methods to produce GUVs. We tested the protocol, focusing on vesicles with a high Chol content using different spin-coating durations and vesicle type deposition. Additionally, we compared the novel protocol using completely dry lipid film. The optimal spin-coating duration for vesicles created from the phosphatidylcholine/Chol mixture was 30 s. Multilamellar vesicles (MLVs), large unilamellar vesicles (LUVs) obtained by the extrusion of MLVs through 100 nm membrane pores and LUVs obtained by extrusion of previously obtained LUVs through 50 nm membrane pores, were deposited on an electrode for 1.5/1 Chol/phosphatidylcholine (POPC) lipid mixture, and the results were compared. Electroformation using all three deposited vesicle types resulted in a high GUV yield, but the deposition of LUVs obtained by the extrusion of MLVs through 100 nm membrane pores provided the most reproducible results. Using the deposition of these LUVs, we produced high yield GUVs for six different Chol concentrations (from 0% to 71.4%). Using a protocol that included dry lipid film GUVs resulted in lower yields and induced the Chol demixing artifact, proving that the lipid film should never be subjected to drying when the Chol content is high. Full article
(This article belongs to the Special Issue Artificial Models of Biological Membranes—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Adapted electroformation protocol. Lipids dissolved in organic solvent (pink) are mixed with an aqueous solution (light blue). MLVs are formed using the RSE method. MLV solution is extruded through a filter with membrane pores in order to obtain LUVs. LUV solution is spin-coated on a plasma threated electrode, where these vesicle rupture and form a damp lipid film. An electrode with damp lipid film is used to build an electroformation chamber, where lipid bilayers detach and form GUVs under the influence of osmotic pressure and an alternating electric field.</p>
Full article ">Figure 2
<p>Electroformation of GUVs for deposition of different vesicle types. (<b>a</b>) Fluorescence microscopic images of GUVs formed when different vesicle types were deposited. MLV represents the deposition of MLVs on the electrode, LUV 100 represents the deposition of LUVs formed by extruding MLVs through 100 nm membrane pores, and LUV 50 represents the deposition of LUVs formed by extruding LUV 100 vesicles through 50 nm membrane pores. The scale bar represents 50 µm. (<b>b</b>) Comparison of average diameters and standard deviations of GUVs for preparations from three different vesicle types. The averages and standard deviations were calculated by averaging the mean diameters from three independent samples for each condition. (<b>c</b>) Size distribution densities of GUVs for deposition of different vesicle types for each sample. Each distribution density represents one independent sample (100 vesicles).</p>
Full article ">Figure 3
<p>(<b>a</b>) GUVs mean diameters for different Chol/POPC mixing ratios. Points and bars represent mean values and their standard errors. (<b>b</b>) Size distribution densities of GUVs for different Chol/POPC mixing ratios. All experiments were performed using the deposition of LUV 100 vesicles spin-coated onto electrode for 30 s.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic depiction of differences obtained using our novel protocol for three different experiments. In the first experiment, damp lipid film was formed by depositing LUV 100 vesicles on an electrode and spinning it for 30 s at 600 rpm angular velocity (Damp 30 s). In the second experiment, Damp 30 s lipid film was held under vacuum for 30 min (Dry 30 s). In the third experiment, instead of spinning for 30 s, the electrode was spun for 240 s, and afterwards held under vacuum for 30 min (Dry 240 s). (<b>b</b>) Success of electroformation using Damp 30 s, Dry 30 s, and Dry 240 s protocols for obtaining lipid films, depending on different Chol/POPC mixing ratios. Success is based on the population homogeneity, yield, and number of defects. It is displayed through circle fullness, where a fuller circle indicates greater success. Empty circles (white circle) denote that no GUVs were formed or that their number was negligible; quarter circles represent low, half and three-quarter circles indicate medium, and full circles (gray circle) indicate high success. (<b>c</b>) Fluorescence microscopy images of GUVs for the different Chol/POPC mixing ratios using three different approaches for obtaining lipid films. The scale bar represents 50 µm.</p>
Full article ">
16 pages, 2357 KiB  
Article
Flavonoid-Labeled Biopolymer in the Structure of Lipid Membranes to Improve the Applicability of Antioxidant Nanovesicles
by Patrick D. Mathews, Gabriella S. Gama, Hector M. Megiati, Rafael R. M. Madrid, Bianca B. M. Garcia, Sang W. Han, Rosangela Itri and Omar Mertins
Pharmaceutics 2024, 16(1), 141; https://doi.org/10.3390/pharmaceutics16010141 - 20 Jan 2024
Cited by 3 | Viewed by 1877
Abstract
Nanovesicles produced with lipids and polymers are promising devices for drug and bioactive delivery and are of great interest in pharmaceutical applications. These nanovesicles can be engineered for improvement in bioavailability, patient compliance or to provide modified release or enhanced delivery. However, their [...] Read more.
Nanovesicles produced with lipids and polymers are promising devices for drug and bioactive delivery and are of great interest in pharmaceutical applications. These nanovesicles can be engineered for improvement in bioavailability, patient compliance or to provide modified release or enhanced delivery. However, their applicability strongly depends on the safety and low immunogenicity of the components. Despite this, the use of unsaturated lipids in nanovesicles, which degrade following oxidation processes during storage and especially during the proper routes of administration in the human body, may yield toxic degradation products. In this study, we used a biopolymer (chitosan) labeled with flavonoid (catechin) as a component over a lipid bilayer for micro- and nanovesicles and characterized the structure of these vesicles in oxidation media. The purpose of this was to evaluate the in situ effect of the antioxidant in three different vesicular systems of medium, low and high membrane curvature. Liposomes and giant vesicles were produced with the phospholipids DOPC and POPC, and crystalline cubic phase with monoolein/DOPC. Concentrations of chitosan–catechin (CHCa) were included in all the vesicles and they were challenged in oxidant media. The cytotoxicity analysis using the MTT assay (3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide) revealed that concentrations of CHCa below 6.67 µM are non-toxic to HeLa cells. The size and zeta potential of the liposomes evidenced the degradation of their structures, which was minimized by CHCa. Similarly, the membrane of the giant vesicle, which rapidly deteriorated in oxidative solution, was protected in the presence of CHCa. The production of a lipid/CHCa composite cubic phase revealed a specific cubic topology in small-angle X-ray scattering, which was preserved in strong oxidative media. This study demonstrates the specific physicochemical characteristics introduced in the vesicular systems related to the antioxidant CHCa biopolymer, representing a platform for the improvement of composite nanovesicle applicability. Full article
Show Figures

Figure 1

Figure 1
<p>Assessment of HeLa cell viability via MTT assay following CHCa (<b>A</b>) and HAc (<b>B</b>) treatment. Two independent experiments were carried out, each performed in triplicate. One-way ANOVA was used to compare different concentrations of CHCa and HAc with control group (0 mg/mL or mM, respectively). ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.0005 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Snapshots of microscopy images in the phase contrast mode of characteristic time sequences of (<b>a</b>) the damage on POPC giant vesicles membrane by photoactivation of MB (665 nm; 40 μM) from left to right: stable vesicle right before MB irradiation; vesicle flickers, increase in membrane surface and bud release; phase contrast decay and large transient pore take place; complete loss of phase contrast; (<b>b</b>) GUV covered with chitosan–catechin (1.8 wt%) showing higher membrane stability and prolonged phase contrast decay with MB irradiation. Start time: start of Hg lamp on the microscope. Bars span 20 μm for all.</p>
Full article ">Figure 3
<p>Profiles of phase contrast fading over time for POPC giant vesicles (<b>a</b>) and the same containing chitosan–catechin (<b>b</b>) at 0.9 (▲) and 1.8 wt% (●) submitted to photoactivation of MB (665 nm; 40 μM). Each point was obtained from the average of at least 10 vesicles and the uncertainty is shown by the error bar.</p>
Full article ">Figure 4
<p>SAXS patterns of monoolein/DOPC (40/60 <span class="html-italic">w</span>/<span class="html-italic">w</span>) liquid crystalline phase (<b>a</b>), the same in oxidative solution (<b>b</b>), hybrid lipid–biopolymer phase (0.8 μM chitosan–catechin) (<b>c</b>) and the same in oxidative solution (<b>d</b>). The indexing of the Bragg peaks is shown in upper panels along the respective models for cubic phase of Im3m space group and inverse hexagonal H<sub>II</sub>.</p>
Full article ">Scheme 1
<p>Coordination of chitosan–catechin segment over a layer of phospholipids by means of known electrostatic interaction [<a href="#B40-pharmaceutics-16-00141" class="html-bibr">40</a>] and mechanism of singlet oxygen scavenging by O-H dissociation at phenolic ring of chemically bonded catechin depicting the antioxidant activity of the flavonoid. The biopolymer provides a protecting shield on lipids membranes.</p>
Full article ">
14 pages, 994 KiB  
Review
Heat Shock Proteins Mediate Intercellular Communications within the Tumor Microenvironment through Extracellular Vesicles
by Renata F. Saito, Camila Maria Longo Machado, Ana Luiza Oliveira Lomba, Andréia Hanada Otake and Maria Cristina Rangel
Appl. Biosci. 2024, 3(1), 45-58; https://doi.org/10.3390/applbiosci3010003 - 1 Jan 2024
Cited by 2 | Viewed by 1959
Abstract
From an evolutive perspective, tumor cells endure successive turnover upon stress conditions and pressure to adapt to new environments. These cells use exceptional communication skills to share biological information to “survive upon every metabolic cost”. The tumor microenvironment (TME) is a miscellaneous collection [...] Read more.
From an evolutive perspective, tumor cells endure successive turnover upon stress conditions and pressure to adapt to new environments. These cells use exceptional communication skills to share biological information to “survive upon every metabolic cost”. The tumor microenvironment (TME) is a miscellaneous collection of cells, factors, and extracellular vesicles (EVs). EVs are small lipid bilayer-delimited particles derived from cells with sizes ranging from 100 to 1000 nm. Exosomes (<160 nm) are the minor subtype of EVs, originating from the endosomal pathways. The TME also contains “giant” vesicles, microvesicles (100–1000 nm, MV), originated from membrane blebbing. EVs can act as intercellular communication mediators, contributing to many biological processes, by carrying different biomolecules, such as proteins, lipids, nucleic acids, and metabolites. EV secretion can promote either tumor cell survival or manage their stress to death. Tumor-derived EVs transfer adaptative stress signaling to recipient cells, reprograming these cells. Heat shock proteins (HSP) are prominent stress response regulators, specifically carried by exosomes. HSP-loaded EVs reprogram tumor and TME cells to acquire mechanisms contributing to tumor progression and therapy resistance. The intercellular communication mediated by HSP-loaded EVs favors the escape of tumor cells from the endoplasmic reticulum stress, hypoxia, apoptosis, and anticancer therapies. Extracellular HSPs activate and deactivate the immune response, induce cell differentiation, change vascular homeostasis, and help to augment the pre-metastatic niche formation. Here we explore EVs’ mechanisms of HSP transmission among TME cells and the relevance of these intercellular communications in resistance to therapy. Full article
Show Figures

Figure 1

Figure 1
<p>The dual role of EV-HSPs inside the tumor microenvironment. (HSP: heat shock protein; MDSC: myeloid-derived suppressor cell; NK: natural killer). Created with BioRender.com (accessed on 10 May 2023).</p>
Full article ">Figure 2
<p>Roles of extracellular vesicles (EVs) derived from tumor microenvironment (TME) after injuries acting in tumor repopulation (ER: endoplasmic reticulum; HSP: heat shock protein; MDSC: myeloid-derived suppressor cell; NK: natural killer). Created with BioRender.com (accessed on 15 April 2023).</p>
Full article ">
13 pages, 1611 KiB  
Article
Membrane Tubulation with a Biomembrane Force Probe
by Lancelot Pincet and Frédéric Pincet
Membranes 2023, 13(12), 910; https://doi.org/10.3390/membranes13120910 - 18 Dec 2023
Viewed by 1942
Abstract
Tubulation is a common cellular process involving the formation of membrane tubes ranging from 50 nm to 1 µm in diameter. These tubes facilitate intercompartmental connections, material transport within cells and content exchange between cells. The high curvature of these tubes makes them [...] Read more.
Tubulation is a common cellular process involving the formation of membrane tubes ranging from 50 nm to 1 µm in diameter. These tubes facilitate intercompartmental connections, material transport within cells and content exchange between cells. The high curvature of these tubes makes them specific targets for proteins that sense local geometry. In vitro, similar tubes have been created by pulling on the membranes of giant unilamellar vesicles. Optical tweezers and micromanipulation are typically used in these experiments, involving the manipulation of a GUV with a micropipette and a streptavidin-coated bead trapped in optical tweezers. The interaction forms streptavidin/biotin bonds, leading to tube formation. Here, we propose a cost-effective alternative using only micromanipulation techniques, replacing optical tweezers with a Biomembrane Force Probe (BFP). The BFP, employing a biotinylated erythrocyte as a nanospring, allows for the controlled measurement of forces ranging from 1 pN to 1 nN. The BFP has been widely used to study molecular interactions in cellular processes, extending beyond its original purpose. We outline the experimental setup, tube formation and characterization of tube dimensions and energetics, and discuss the advantages and limitations of this approach in studying membrane tubulation. Full article
(This article belongs to the Special Issue Artificial Models of Biological Membranes—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>BFP formation. The three snapshots are extracted from <a href="#app1-membranes-13-00910" class="html-app">Video S1</a>. In panel (<b>a</b>), a streptavidin-coated bead is securely held by the left micropipette, while a biotinylated erythrocyte is held by the right micropipette. In panel (<b>b</b>), the bead is brought into contact with the erythrocyte, facilitating the formation of streptavidin/biotin bonds. Finally, in panel (<b>c</b>), the bead is released from the left pipette, resulting in the assembly of the probe comprising the erythrocyte and the bead. This assembled probe is now ready for use in force measurements.</p>
Full article ">Figure 2
<p>Tube diameter from tube length variations. After tube formation, the tube diameter is gauged by altering the tube length through a controlled approach and separation of the two micropipettes (panel (<b>a</b>)). In the upper image, the microscope is focused on the tube, while in the lower two images, it is focused on the end of the vesicle extension within the micropipette. Given that the vesicle surface area and the volume of the vesicle lumen remain constant, changes in tube length (<math display="inline"><semantics> <mrow> <mi>δ</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math>) and vesicle extension in the micropipette (<math display="inline"><semantics> <mrow> <mi>δ</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math>) directly quantify the tube diameter using Equation (4). In panel (<b>b</b>), data points are presented for two approaches and one separation, where the slope is proportionate to the tube diameter. In this specific instance, the tube diameter measures 82 nm, with an accuracy of approximately 10%.</p>
Full article ">Figure 3
<p>Tube diameter from fluorescence intensity. For a specific fluorescent lipid, the ratio, <span class="html-italic">r</span>, between the intensity of the tube and the intensity at the equator of the GUV is directly proportional to the tube diameter. The method outlined in <a href="#membranes-13-00910-f002" class="html-fig">Figure 2</a> is employed to measure the diameter for various tubes, enabling the determination of the proportionality coefficient. Once established, this coefficient allows for a reliable estimation of the tube diameter based on a measurement of the ratio <span class="html-italic">r</span>. The presented curve is the outcome of 87 tube diameter measurements. These diameters were grouped in increments of 5 nm, and the error bars represent standard deviations.</p>
Full article ">Figure 4
<p>Force measurement. In (<b>a</b>), the three upper panels are screenshots extracted using Leica LasX software (version 4.5.0.25531) from the experiment featured in <a href="#app1-membranes-13-00910" class="html-app">Video S4</a>. The GUV comes into contact with the BFP and is subsequently moved away. The force is determined throughout this process by measuring the bead displacement and utilizing the predetermined stiffness of the BFP. Upon tube formation, a negative force is observed, indicating that the bead is being drawn towards the GUV. The three lower panels provide the intensities (a.u) observed along the green horizontal line (µm) in the bright-field channel (bead, top curve) and fluorescence channel (GUV, bottom curve) for the three screenshots in the upper panels. The red (resp. white) dashed lines between the screenshots and the intensity plots indicate the end of the tube on the GUV side (resp. the bead side) indicated by the maximum (resp. minimum) intensity at each timepoint. In the example illustrated in panel (<b>b</b>), the force is plotted against the relative position of the left pipette (green line). The leftmost panel in (<b>a</b>) corresponds to the point where the attractive force is maximal (approximately −20 pN), while the subsequent two panels correspond to the plateau around −15 pN at two distinct tube lengths. In this example, the surface tension of the GUV was 28.9 µN/m and the bending modulus 20.5 <span class="html-italic">k<sub>B</sub>T</span>. Hence, the tube force predicted by Equation (10) is 13.9 pN. This predicted tube force, indicated by the orange dashed line, is very close to the experimental value.</p>
Full article ">
17 pages, 3806 KiB  
Article
Antifungal Activity and Molecular Mechanisms of Copper Nanoforms against Colletotrichum gloeosporioides
by Mun’delanji C. Vestergaard, Yuki Nishida, Lihn T. T. Tran, Neha Sharma, Xiaoxiao Zhang, Masayuki Nakamura, Auriane F. Oussou-Azo and Tomoki Nakama
Nanomaterials 2023, 13(23), 2990; https://doi.org/10.3390/nano13232990 - 22 Nov 2023
Cited by 2 | Viewed by 1976
Abstract
In this work, we have synthesized copper nanoforms (Cu NFs) using ascorbic acid as a reducing agent and polyvinylpyrrolidone as a stabilizer. Elemental characterization using EDS has shown the nanostructure to be of high purity and compare well with commercially sourced nanoforms. SEM [...] Read more.
In this work, we have synthesized copper nanoforms (Cu NFs) using ascorbic acid as a reducing agent and polyvinylpyrrolidone as a stabilizer. Elemental characterization using EDS has shown the nanostructure to be of high purity and compare well with commercially sourced nanoforms. SEM images of both Cu NFs show some agglomeration. The in-house NFs had a better even distribution and size of the nanostructures. The XRD peaks represented a face-centered cubic structure of Cu2O. The commercially sourced Cu NFs were found to be a mixture of Cu and Cu2O. Both forms had a crystalline structure. Using these two types of Cu NFs, an antimicrobial study against Colletotrichum gloeosporioides, a devastating plant pathogen, showed the in-house Cu NFs to be most effective at inhibiting growth of the pathogen. Interestingly, at low concentrations, both Cu NFs increased fungal growth, although the mycelia appeared thin and less dense than in the control. SEM macrographs showed that the in-house Cu NFs inhibited the fungus by flattening the mycelia and busting some of them. In contrast, the mycelia were short and appeared clustered when exposed to commercial Cu NFs. The difference in effect was related to the size and/or oxidation state of the Cu NFs. Furthermore, the fungus produced a defense mechanism in response to the NFs. The fungus produced melanin, with the degree of melanization directly corresponding to the concentration of the Cu NFs. Localization of aggregated Cu NFs could be clearly observed outside of the model membranes. The large agglomerates may only contribute indirectly by a hit-and-bounce-off effect, while small structures may adhere to the membrane surface and/or internalize. Spatio-temporal membrane dynamics were captured in real time. The dominant dynamics culminated into large fluctuations. Some of the large fluctuations resulted in vesicular transformation. The major transformation was exo-bud/exo-cytosis, which may be a way to excrete the foreign object (Cu NFs). Full article
(This article belongs to the Special Issue Morphological Design and Synthesis of Nanoparticles)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction (XRD) patterns of (<b>A</b>) in-house Cu NFs and (<b>B</b>) commercial Cu NFs. Patterns were obtained with a PANalytical X’PERT PRO MPD diffractometer (Malvern Panalytical, Japan) operating at 45 kV and 40 mA with Cu-Kα radiation (λ = 1.54060 Å) measurement being performed at a scan speed of 0.01(units) from the 2θ 20° to 80° range.</p>
Full article ">Figure 1 Cont.
<p>X-ray diffraction (XRD) patterns of (<b>A</b>) in-house Cu NFs and (<b>B</b>) commercial Cu NFs. Patterns were obtained with a PANalytical X’PERT PRO MPD diffractometer (Malvern Panalytical, Japan) operating at 45 kV and 40 mA with Cu-Kα radiation (λ = 1.54060 Å) measurement being performed at a scan speed of 0.01(units) from the 2θ 20° to 80° range.</p>
Full article ">Figure 2
<p>Dispersive energy X-ray spectra of (<b>A</b>) Cu NFs synthesized in house using ascorbic acid and polyvinyl pyrrolidone and (<b>B</b>) commercial Cu NFs. The elemental (% weight and atomic %), values and SEM images of the analysis area can be seen (inset). The analysis was conducted in triplicate at an accelerating voltage of 25.0 kV and magnification of 1000×. Elemental analysis data are the mean of three replicates.</p>
Full article ">Figure 3
<p>Typical SEM images of Cu NFs (<b>A</b>) synthesized in house using ascorbic acid and polyvinylpyrrolidone and (<b>B</b>) commercially sourced Cu NFs. Bars indicate 20 μM and 2 μM (insets). The white bars represent 100 μm size; the red bars (inset images) represent 10 μm size Samples were imaged using an accelerating voltage of 5.0 kV.</p>
Full article ">Figure 4
<p>Effect of Cu NFs on <span class="html-italic">C. gloeosporioides</span>, a fungus that attacks food crops. (<b>A</b>) Diameter growth of the fungus in the presence and absence (control) of (<b>i</b>) in-house Cu NFs and (<b>ii</b>) commercial Cu NFs; Percent inhibition in fungal growth in the presence of Cu NSFs (<b>iii</b>); (<b>B</b>) SEM images of the of the fungus (<b>i</b>) alone; in the presence of (<b>ii</b>) inhouse Cu NFs and (<b>iii</b>) commercial Cu NFs. The white bars represent 50 μm size; the white bars (inset images) represent 10 μm size. Images were obtained in low-vacuum mode with water vapor and an accelerating voltage of 15.0 kV.</p>
Full article ">Figure 5
<p>Interaction of <span class="html-italic">C. gloeosporioides</span> with copper nanoforms analyzed using FTIR. FTIR spectra of fungal mycelium (<b>A</b>) alone (control), (<b>B</b>) with Cu NPs, (<b>C</b>) with Cu<sub>2</sub>O NPs (FTIR analyses were conducted using a JASCO FT/IR-4200 instrument with an attenuated total reflection (ATR) set-up). Absorption spectra were obtained in the region between 4000 and 400 cm<sup>−1</sup>.</p>
Full article ">Figure 6
<p>(<b>A</b>) Aggregated Cu NFs in aqueous lipid media (black dots). The outline of a lipid vesicle can be seen. Giant unilamellar vesicles were prepared using the 1,2-dioleoyl-<span class="html-italic">sn</span>-glycero-3-phosphocholine (DOPC, 0.2 mM) by the natural swelling method. Top: original image; Bottom: image after artistic modification to enhance contrast; (<b>B</b>) membrane dynamics induced by 0.5 mg/mL of Cu NPs and Cu<sub>2</sub>O NPs. Vesicles were prepared as in (<b>A</b>).</p>
Full article ">Figure 7
<p>Endo-bud formation. The most dominant lipid vesicular transformation pathway induced by Cu NFs. Top: Original image; Bottom: stylized image to clearly show what was observable under the objective lens. Giant unilamellar vesicles were prepared using 1,2-dioleoyl-<span class="html-italic">sn</span>-glycero-3-phosphocholine (DOPC, 0.2 mM) by the natural swelling method. Size bar = 20 μm.</p>
Full article ">Figure 8
<p>Melanization of <span class="html-italic">C. gloeosporioides</span> in the presence of in-house Cu NFs (<b>A</b>). <span class="html-italic">C. gloeosporioides</span> was cultured in potato dextrose agar for 14 d at 25 °C in the presence of in-house Cu NFs. Standard curve of melanin (<b>B</b>).</p>
Full article ">
13 pages, 1020 KiB  
Article
Membrane-Mediated Cooperative Interactions of CD47 and SIRPα
by Long Li, Chen Gui, Jinglei Hu and Bartosz Różycki
Membranes 2023, 13(11), 871; https://doi.org/10.3390/membranes13110871 - 2 Nov 2023
Cited by 3 | Viewed by 2159
Abstract
The specific binding of the ubiquitous ‘marker of self’ protein CD47 to the SIRPα protein anchored in the macrophage plasma membrane results in the inhibition of the engulfment of ‘self’ cells by macrophages and thus constitutes a key checkpoint of our innate [...] Read more.
The specific binding of the ubiquitous ‘marker of self’ protein CD47 to the SIRPα protein anchored in the macrophage plasma membrane results in the inhibition of the engulfment of ‘self’ cells by macrophages and thus constitutes a key checkpoint of our innate immune system. Consequently, the CD47–SIRPα protein complex has been recognized as a potential therapeutic target in cancer and inflammation. Here, we introduce a lattice-based mesoscale model for the biomimetic system studied recently in fluorescence microscopy experiments where GFP-tagged CD47 proteins on giant plasma membrane vesicles bind to SIRPα proteins immobilized on a surface. Computer simulations of the lattice-based mesoscale model allow us to study the biomimetic system on multiple length scales, ranging from single nanometers to several micrometers and simultaneously keep track of single CD47–SIRPα binding and unbinding events. Our simulations not only reproduce data from the fluorescence microscopy experiments but also are consistent with results of several other experiments, which validates our numerical approach. In addition, our simulations yield quantitative predictions on the magnitude and range of effective, membrane-mediated attraction between CD47–SIRPα complexes. Such detailed information on CD47–SIRPα interactions cannot be obtained currently from experiments alone. Our simulation results thus extend the present understanding of cooperative effects in CD47–SIRPα interactions and may have an influence on the advancement of new cancer treatments. Full article
(This article belongs to the Section Biological Membrane Dynamics and Computation)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Cartoon of the system under study: SIRP<math display="inline"><semantics> <mi>α</mi> </semantics></math> receptors (dark blue) immobilized on a planar surface (gray) can bind CD47 ligands (dark green) on a GPMV (light green). The surface is coated with BSA (light blue) to prevent non-specific adhesion of the membrane to the surface. (<b>B</b>) Lattice-based mesoscale model that takes into account (i) diffusion of the membrane-anchored ligands, (ii) binding and unbinding of the receptors and their ligands, and (iii) elastic deformations and thermal undulations of the membrane. The color code is as in panel A. The lateral size of the system is 6 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">Figure 2
<p>Two-dimensional binding affinity <math display="inline"><semantics> <msub> <mi>K</mi> <mrow> <mn>2</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math> times the area concentration of receptors, <math display="inline"><semantics> <mrow> <mo>[</mo> <mi>RL</mi> <mo>]</mo> <mo>=</mo> <mn>4000</mn> <mo>/</mo> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math><math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math>, versus the average area concentration <math display="inline"><semantics> <mrow> <mo>[</mo> <mi>RL</mi> <mo>]</mo> </mrow> </semantics></math> of receptor–ligand complexes. The data points in black correspond to the experimental FRAP data taken from Figure 2A in Reference [<a href="#B9-membranes-13-00871" class="html-bibr">9</a>]. The points in blue, orange, and purple represent the results of MC simulations with <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mi mathvariant="normal">b</mi> </msub> <mo>=</mo> <mn>5.4</mn> <mo>,</mo> <mo> </mo> <mn>6.6</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mn>7.8</mn> </mrow> </semantics></math> nm, respectively. The dashed lines are to guide the eye.</p>
Full article ">Figure 3
<p>Two-dimensional binding affinity <math display="inline"><semantics> <msub> <mi>K</mi> <mrow> <mn>2</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math> times the area concentration of receptors, <math display="inline"><semantics> <mrow> <mo>[</mo> <mi>RL</mi> <mo>]</mo> <mo>=</mo> <mn>4000</mn> <mo>/</mo> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math><math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math>, versus the thermal roughness <math display="inline"><semantics> <msub> <mo>ξ</mo> <mo>⊥</mo> </msub> </semantics></math> of the membrane. The different colors represent the results of MC simulations with <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mi mathvariant="normal">b</mi> </msub> <mo>=</mo> <mn>5.4</mn> <mo>,</mo> <mo> </mo> <mn>6.6</mn> <mo>,</mo> <mo>…</mo> <mn>9</mn> <mo>,</mo> <mo> </mo> <mn>10.2</mn> </mrow> </semantics></math> nm. The solid line in black shows the relation <math display="inline"><semantics> <mrow> <msub> <mi>K</mi> <mrow> <mn>2</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> <mrow> <mo>[</mo> <mi>RL</mi> <mo>]</mo> </mrow> <mo>=</mo> <msub> <mo>ℓ</mo> <mn>1</mn> </msub> <mo>/</mo> <msub> <mo>ξ</mo> <mo>⊥</mo> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mo>ℓ</mo> <mn>1</mn> </msub> <mo>=</mo> <mn>5.45</mn> <mspace width="0.166667em"/> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> being a fitting parameter. This relation is equivalent to <math display="inline"><semantics> <mrow> <msub> <mi>K</mi> <mrow> <mn>2</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>K</mi> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>c</mi> <mn>1</mn> </msub> <mo>/</mo> <msub> <mo>ξ</mo> <mo>⊥</mo> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>c</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>1.22</mn> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <msub> <mi>K</mi> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> <mspace width="0.166667em"/> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">M</mi> </mrow> </semantics></math> is the dissociation constant of the soluble variants of CD47 and SIRP<math display="inline"><semantics> <mi>α</mi> </semantics></math>.</p>
Full article ">Figure 4
<p>CD47–SIRP<math display="inline"><semantics> <mi>α</mi> </semantics></math> binding rate constants <math display="inline"><semantics> <msub> <mi>k</mi> <mi>off</mi> </msub> </semantics></math> (<b>A</b>) and <math display="inline"><semantics> <msub> <mi>k</mi> <mi>on</mi> </msub> </semantics></math> (<b>B</b>) as a function of membrane roughness <math display="inline"><semantics> <msub> <mo>ξ</mo> <mo>⊥</mo> </msub> </semantics></math>. The different colors represent the results of MC simulations with <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mi mathvariant="normal">b</mi> </msub> <mo>=</mo> <mn>5.4</mn> <mo>,</mo> <mo> </mo> <mn>6.6</mn> <mo>,</mo> <mo>…</mo> <mn>9</mn> <mo>,</mo> <mo> </mo> <mn>10.2</mn> </mrow> </semantics></math> nm. The dashed line in panel (<b>A</b>) indicates <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>off</mi> </msub> <mo>=</mo> <mn>1.6</mn> </mrow> </semantics></math> s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, determined in surface plasmon resonance experiments [<a href="#B14-membranes-13-00871" class="html-bibr">14</a>]. The solid line in panel (<b>B</b>) shows <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>on</mi> </msub> <mo>=</mo> <msub> <mi>c</mi> <mn>2</mn> </msub> <mo>/</mo> <msub> <mo>ξ</mo> <mo>⊥</mo> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <msub> <mi>c</mi> <mn>2</mn> </msub> </semantics></math> being a fit parameter.</p>
Full article ">Figure 5
<p>Potential of mean force <math display="inline"><semantics> <mrow> <mi>w</mi> <mrow> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> <mo>=</mo> <mo>−</mo> <msub> <mi>k</mi> <mi mathvariant="normal">B</mi> </msub> <mi>T</mi> <mo form="prefix">ln</mo> <mi>g</mi> <mrow> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </semantics></math> is the two-dimensional pair correlation function for receptor–ligand complexes. Panels (<b>A</b>,<b>B</b>) correspond to <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mi mathvariant="normal">b</mi> </msub> <mo>=</mo> <mn>7.8</mn> </mrow> </semantics></math> nm and <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mi mathvariant="normal">b</mi> </msub> <mo>=</mo> <mn>6.6</mn> </mrow> </semantics></math> nm, respectively. The lines in orange, red, purple, and blue correspond to 30, 50, 70, and 90 receptor–ligand complexes per square micron.</p>
Full article ">Figure 6
<p>Second virial coefficient, <math display="inline"><semantics> <msub> <mi>B</mi> <mn>2</mn> </msub> </semantics></math>, as a function of the area concentration of the receptor–ligand complexes, <math display="inline"><semantics> <mrow> <mo>[</mo> <mi>RL</mi> <mo>]</mo> </mrow> </semantics></math>, computed for <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mi mathvariant="normal">b</mi> </msub> <mo>=</mo> <mn>5.4</mn> </mrow> </semantics></math> nm (blue), <math display="inline"><semantics> <mrow> <mn>6.6</mn> </mrow> </semantics></math> (orange), and <math display="inline"><semantics> <mrow> <mn>7.8</mn> </mrow> </semantics></math> (purple). The color code and symbols are as in <a href="#membranes-13-00871-f002" class="html-fig">Figure 2</a>.</p>
Full article ">
13 pages, 3704 KiB  
Article
QS21-Initiated Fusion of Liposomal Small Unilamellar Vesicles to Form ALFQ Results in Concentration of Most of the Monophosphoryl Lipid A, QS21, and Cholesterol in Giant Unilamellar Vesicles
by Erwin G. Abucayon, Mangala Rao, Gary R. Matyas and Carl R. Alving
Pharmaceutics 2023, 15(9), 2212; https://doi.org/10.3390/pharmaceutics15092212 - 26 Aug 2023
Cited by 2 | Viewed by 1687
Abstract
Army Liposome Formulation with QS21 (ALFQ), a vaccine adjuvant preparation, comprises liposomes containing saturated phospholipids, with 55 mol% cholesterol relative to the phospholipids, and two adjuvants, monophosphoryl lipid A (MPLA) and QS21 saponin. A unique feature of ALFQ is the formation of giant [...] Read more.
Army Liposome Formulation with QS21 (ALFQ), a vaccine adjuvant preparation, comprises liposomes containing saturated phospholipids, with 55 mol% cholesterol relative to the phospholipids, and two adjuvants, monophosphoryl lipid A (MPLA) and QS21 saponin. A unique feature of ALFQ is the formation of giant unilamellar vesicles (GUVs) having diameters >1.0 µm, due to a remarkable fusion event initiated during the addition of QS21 to nanoliposomes containing MPLA and 55 mol% cholesterol relative to the total phospholipids. This results in a polydisperse size distribution of ALFQ particles, with diameters ranging from ~50 nm to ~30,000 nm. The purpose of this work was to gain insights into the unique fusion reaction of nanovesicles leading to GUVs induced by QS21. This fusion reaction was probed by comparing the lipid compositions and structures of vesicles purified from ALFQ, which were >1 µm (i.e., GUVs) and the smaller vesicles with diameter <1 µm. Here, we demonstrate that after differential centrifugation, cholesterol, MPLA, and QS21 in the liposomal phospholipid bilayers were present mainly in GUVs (in the pellet). Presumably, this occurred by rapid lateral diffusion during the transition from nanosize to microsize particles. While liposomal phospholipid recoveries by weight in the pellet and supernatant were 44% and 36%, respectively, higher percentages by weight of the cholesterol (~88%), MPLA (94%), and QS21 (96%) were recovered in the pellet containing GUVs, and ≤10% of these individual liposomal constituents were recovered in the supernatant. Despite the polydispersity of ALFQ, most of the cholesterol, and almost all of the adjuvant molecules, were present in the GUVs. We hypothesize that the binding of QS21 to cholesterol caused new structural nanodomains, and subsequent interleaflet coupling in the lipid bilayer might have initiated the fusion process, leading to creation of GUVs. However, the polar regions of MPLA and QS21 together have a “sugar lawn” of ten sugars, the hydrophilicity of which might have provided a driving force for rapid lateral diffusion and concentration of the MPLA and QS21 in the GUVs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Physical properties of ALFQ in comparison with AS01-like liposomal adjuvant. Unlike AS01-like liposomes, centrifugation of ALFQ at 15,000 RPM (2696× <span class="html-italic">g</span>) for 15 min at 22 °C resulted in the formation of a pellet.</p>
Full article ">Figure 2
<p>Characterization of the pellets and supernatant from ALFQ centrifugation at 8000 RPM (767× <span class="html-italic">g</span>). (<b>A</b>) Visualization by phase-contrast light microscopy studies showed that the pellet is made of GUVs with diameter &gt;1 μm. (<b>B</b>) Majority of the vesicles in the supernatant have diameters &lt;1 μm. (<b>C</b>) DLS studies also show that the modal sizes of the vesicles in the supernatant are ~80 nm (SUVs) and ~550 nm (LUVs).</p>
Full article ">Figure 3
<p>Confocal microscopy images of the pellet (<b>A</b>), with the corresponding image in gray scale (<b>B</b>), and supernatant (<b>C</b>), with the corresponding image in gray scale (<b>D</b>) after centrifugation of ALFQ labelled with TopFluor<sup>®</sup> cholesterol at 8000 RPM (767× <span class="html-italic">g</span>) for 10 min at 22 °C. The pellet was washed 3× before conducting confocal microscopy studies.</p>
Full article ">Figure 4
<p>Molar amounts of QS21 and MPLA in the pellet and supernatant after centrifugation, relative to the original ALFQ.</p>
Full article ">Figure 5
<p>Molar ratios of cholesterol relative to total phospholipids (DMPC + DMPG) in ALFQ (<b>A</b>), centrifuged pellet (<b>B</b>), and supernatant (<b>C</b>). The pellet, consisting of GUVs, has substantially higher cholesterol content relative to total phospholipids compared to the original ALFQ.</p>
Full article ">Figure 6
<p>Biophysical characterization and visualization of ALF55 and ALFQ without MPLA, in comparison with the original ALFQ suspension. (<b>A</b>) DLS analysis of ALF55 without MPLA showing a modal peak of ~50 nm. (<b>B</b>) Visualization of ALFQ without MPLA using confocal microscopy, (<b>C</b>) with the corresponding grayscale image. (<b>D</b>) DLS analysis of ALF55 with MPLA showing a modal peak of ~50 nm. (<b>E</b>) Visualization of the original ALFQ suspension using confocal microscopy, (<b>F</b>) with the corresponding grayscale image.</p>
Full article ">
9 pages, 2979 KiB  
Article
Microscopic Observation of Membrane Fusion between Giant Liposomes and Baculovirus Budded Viruses Activated by the Release of a Caged Proton
by Misako Nishigami, Yuki Uno and Kanta Tsumoto
Membranes 2023, 13(5), 507; https://doi.org/10.3390/membranes13050507 - 11 May 2023
Viewed by 1644
Abstract
Baculovirus (Autographa californica multiple nucleopolyhedrovirus, AcMNPV) is an envelope virus possessing a fusogenic protein, GP64, which can be activated under weak acidic conditions close to those in endosomes. When the budded viruses (BVs) are bathed at pH 4.0 to 5.5, they can [...] Read more.
Baculovirus (Autographa californica multiple nucleopolyhedrovirus, AcMNPV) is an envelope virus possessing a fusogenic protein, GP64, which can be activated under weak acidic conditions close to those in endosomes. When the budded viruses (BVs) are bathed at pH 4.0 to 5.5, they can bind to liposome membranes with acidic phospholipids, and this results in membrane fusion. In the present study, using the caged-proton reagent 1-(2-nitrophenyl)ethyl sulfate, sodium salt (NPE-caged-proton), which can be uncaged by irradiation with ultraviolet light, we triggered the activation of GP64 by lowering the pH and observed membrane fusion on giant liposomes (giant unilamellar vesicles, GUVs) by visualizing the lateral diffusion of fluorescence emitted from a lipophilic fluorochrome (octadecyl rhodamine B chloride, R18) that stained viral envelopes of BVs. In this fusion, entrapped calcein did not leak from the target GUVs. The behavior of BVs prior to the triggering of membrane fusion by the uncaging reaction was closely monitored. BVs appeared to accumulate around a GUV with DOPS, implying that BVs preferred phosphatidylserine. The monitoring of viral fusion triggered by the uncaging reaction could be a valuable tool for revealing the delicate behavior of viruses affected by various chemical and biochemical environments. Full article
(This article belongs to the Special Issue Analytical Sciences of/with Bio(mimetic) Membranes (Volume II))
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the experimental procedure for preparing GUVs that could be observed using a fluorescence confocal laser scanning microscope. Initially, the solution of the bottom phase, which would, in turn, be an outer solution, contained both BVs and NPE-caged-protons. After the preparation of GUVs by the transfer of aqueous droplets through an oil/water interface, the chamber was irradiated by UV light leading to the release of protons from the caged molecules.</p>
Full article ">Figure 2
<p>Time-course of pH vs. the duration of UV irradiation in the aqueous solution containing NPE-caged protons at the indicated concentrations.</p>
Full article ">Figure 3
<p>Typical CLSM images of GUVs (DOPC/DOPS = 1:1) mixed with BVs that have been fluorescently labeled with the fluorochrome R18. The indicated concentrations, 0 mM (none) and 2 mM, refer to the concentration of NPE-caged-protons added to the outer solution of the GUVs. The time shows the elapsed time after the observation started. The sample began to be illuminated by UV light at 5.0 min. Each panel consists of R18 fluorescent images (upper panels) and differential interference contrast (DIC) images (lower panels) of the same GUVs. Bar is 50 µm.</p>
Full article ">Figure 4
<p>Changes in fluorescence intensity of calcein encapsulated inside GUVs in the presence (+) or absence (−) of BVs and NPE-caged-protons. (<b>A</b>) The fluorescence intensity of calcein was monitored after the start of UV irradiation. Each line represents the time-course data of a single GUV. (<b>B</b>) The difference in relative fluorescence intensity with each GUV was plotted under the indicated conditions. The difference is calculated by dividing the intensity at 14.5 min minus that at 0 min by that at 0 min. Typical fluorescence images of GUVs entrapping calcein are provided in <a href="#app1-membranes-13-00507" class="html-app">Figure S1 in the Supplementary Material</a>. The lines and dots are corresponding to individual GUVs.</p>
Full article ">Figure 5
<p>Typical CLSM images of GUVs (DOPC only, DOPC/DOPG = 1:1, DOPC/DOPS = 1:1 molar ratio) mixed with BVs that were fluorescently labeled with the fluorochrome R18 at 2 mM NPE-caged-proton. The solutions were irradiated by UV for the indicated time. Each panel consists of the top profiles, which show the time-lapse change in R18 fluorescence along the dashed line (yellow) crossing the GUV (the intersections are indicated by red dashed lines in the upper panel), and the bottom images, which show the corresponding GUVs at 0 min and 14.5 min. In the case of DOPC/DOPG, the GUV that was smaller compared to the others was observed using the objective lens with low magnification, and the image looks more pixelated. Bar is 20 µm.</p>
Full article ">
22 pages, 7519 KiB  
Article
Morphological Diversity in Diblock Copolymer Solutions: A Molecular Dynamics Study
by Senyuan Liu and Radhakrishna Sureshkumar
Colloids Interfaces 2023, 7(2), 40; https://doi.org/10.3390/colloids7020040 - 9 May 2023
Cited by 2 | Viewed by 2350
Abstract
Coarse-grained molecular dynamics simulations that incorporate explicit water-mediated hydrophilic/hydrophobic interactions are employed to track spatiotemporal evolution of diblock copolymer aggregation in initially homogeneous solutions. A phase portrait of the observed morphologies and their quantitative geometric features such as aggregation numbers, packing parameters, and [...] Read more.
Coarse-grained molecular dynamics simulations that incorporate explicit water-mediated hydrophilic/hydrophobic interactions are employed to track spatiotemporal evolution of diblock copolymer aggregation in initially homogeneous solutions. A phase portrait of the observed morphologies and their quantitative geometric features such as aggregation numbers, packing parameters, and radial distribution functions of solvent/monomers are presented. Energetic and entropic measures relevant to self-assembly such as specific solvent accessible surface area (SASA) and probability distribution functions (pdfs) of segmental stretch of copolymer chains are analyzed. The simulations qualitatively capture experimentally observed morphological diversity in diblock copolymer solutions. Topologically simpler structures predicted include spherical micelles, vesicles (polymersomes), lamellae (bilayers), linear wormlike micelles, and tori. More complex morphologies observed for larger chain lengths and nearly symmetric copolymer compositions include branched wormlike micelles with Y-shaped junctions and cylindrical micelle networks. For larger concentrations, vesicle strands, held together by hydrogen bonds, and “giant” composite aggregates that consist of lamellar, mixed hydrophobic/hydrophilic regions and percolating water cores are predicted. All structures are dynamic and exhibit diffuse domain boundaries. Morphology transitions across topologically simpler structures can be rationalized based on specific SASA measurements. PDFs of segmental stretch within vesicular assemblies appear to follow a log-normal distribution conducive for maximizing configuration entropy. Full article
Show Figures

Figure 1

Figure 1
<p>Atomistic (<b>a</b>) and coarse-grained (<b>b</b>) descriptions of a single PEO-PB copolymer chain composed of 6 butadiene monomers (red) and 6 ethylene oxide monomers (blue). See <a href="#app1-colloids-07-00040" class="html-app">Tables S1–S4</a> for force field parameter values used in the simulations.</p>
Full article ">Figure 2
<p>Schematic diagram illustrating interface of self-assembled micelles constructed using the alpha-shape algorithm: (<b>a</b>) mixed structure (<b>left</b>) and mesh representation of the water−aggregate interface (<b>right</b>), (<b>b</b>) bilayer spherical micelle (<b>left</b>), mesh representation of the interface (<b>middle</b>, Only PB beads are shown), and cross sections showing the inner and outer hydrophobic layers (<b>right</b>).</p>
Full article ">Figure 3
<p>Sectional view (inset of a) and RDF (<b>a</b>) of a 5PB6PEO (<span class="html-italic">x<sub>PEO</sub></span> = 0.55) vesicle. Yellow, red, and blue beads represent water, PB, and PEO, respectively. RDF of vesicles formed by different copolymers: (<b>b</b>): 4PB3PEO (<span class="html-italic">x<sub>PEO</sub></span> = 0.43), (<b>c</b>): 3PB4PEO (<span class="html-italic">x<sub>PEO</sub></span> = 0.57), and (<b>d</b>): 5PB5PEO (<span class="html-italic">x<sub>PEO</sub></span> = 0.50).</p>
Full article ">Figure 4
<p>Probability distribution functions of <span class="html-italic">s</span> ≡ <span class="html-italic">Q</span><sub>max</sub> − <span class="html-italic">Q</span> of internal (<b>a</b>) and external (<b>b</b>) PEO segments as well as PB segments (<b>c</b>) of vesicles corresponding to <a href="#colloids-07-00040-f003" class="html-fig">Figure 3</a>b (5PB6PEO). The solid lines represent the best fit to a log-normal distribution.</p>
Full article ">Figure 5
<p>Sectional view and RDFs of representative spherical micelles made of relatively short chains (<span class="html-italic">N</span> = 6). 2PB4PEO (<b>a</b>,<b>b</b>), 3PB3PEO (<b>c</b>,<b>d</b>), 4PB2PEO (<b>e</b>,<b>f</b>). These structures are found close to the micelle-vesicle transition region in the <span class="html-italic">N</span> vs. <span class="html-italic">x<sub>PEO</sub></span> phase diagram.</p>
Full article ">Figure 6
<p>Sectional view and RDFs of representative PB-dominated spherical micelles (11PB2PEO, panels (<b>a</b>,<b>b</b>)) and spherical micelles (7PB5PEO, panels (<b>c</b>,<b>d</b>); 5PB6PEO, panels (<b>e</b>,<b>f</b>)).</p>
Full article ">Figure 7
<p>Sectional view (<b>a</b>) and RDFs (<b>b</b>) of a PEO-dominated micelle of 3PB8PEO chains. For comparison, sectional view (<b>c</b>) and RDFs (<b>d</b>) of an 8PB3PEO micelle are shown.</p>
Full article ">Figure 8
<p>Side view (<b>a</b>) and RDFs (<b>b</b>) of a flexible, wormlike micelle (5PB8PEO), side views of branched cylindrical micelles ((<b>c</b>), 3PB11PEO), a micelle network ((<b>d</b>), 6PB8PEO), and wormlike micelle with Y branch ((<b>e</b>), 3PB8PEO). RDFs of (<b>a</b>) are along the radial direction. Persistence length and <span class="html-italic">n</span> of the wormlike micelle are 0.74 ± 0.09 nm and 1429, respectively.</p>
Full article ">Figure 9
<p>(<b>a</b>) Orientation angle <span class="html-italic">θ</span> and end-to-end distance <span class="html-italic">Q</span> vs. time for the rodlike micelle are shown in <a href="#colloids-07-00040-f008" class="html-fig">Figure 8</a>a. Snapshots of instantaneous micelle configurations are shown in blue. (<b>b</b>) ACF of <span class="html-italic">θ</span> between 400 ns and 515 ns (0 ns in <b>(b)</b> corresponds to 400 ns in <b>(a)</b>). The exponential fit yields an orientation relaxation time of 40 ± 3 ns.</p>
Full article ">Figure 10
<p>Top view (all beads (<b>a</b>), only PB beads (<b>b</b>)) and RDFs (<b>c</b>) of a torus structure (6PB8PEO). Sectional view (<b>d</b>) and RDFs (<b>e</b>) along the direction shown by the arrow of lamella micelle (7PB7PEO). <span class="html-italic">n</span> = 543 (<b>a</b>), 517 (<b>d</b>).</p>
Full article ">Figure 11
<p>Phase diagram of equilibrium morphologies at 300 K and polymer concentration of 10 mol%. H: homogeneous solution, V: vesicles (polymersomes, <a href="#colloids-07-00040-f003" class="html-fig">Figure 3</a>), SM: short-chain micelles (<a href="#colloids-07-00040-f005" class="html-fig">Figure 5</a>), PB-S: PB-dominated spherical micelles (<a href="#colloids-07-00040-f006" class="html-fig">Figure 6</a>a), S: spherical micelles (<a href="#colloids-07-00040-f006" class="html-fig">Figure 6</a>c,e), PEO-M: PEO-dominated micelles (<a href="#colloids-07-00040-f007" class="html-fig">Figure 7</a>a), R: rodlike/wormlike micelles (linear or branched, <a href="#colloids-07-00040-f008" class="html-fig">Figure 8</a>a,c,e), MN: micelle network (<a href="#colloids-07-00040-f008" class="html-fig">Figure 8</a>d), T: tori (<a href="#colloids-07-00040-f010" class="html-fig">Figure 10</a>a), L: lamellae (<a href="#colloids-07-00040-f010" class="html-fig">Figure 10</a>d), Y: V + S + MN + T, Z: V + S + MN. The phase portrait is inferred from the results of 237 distinct CGMD simulations excluding replicates. The dotted lines connect the midpoint of two adjacent coordinates at which different structures are observed. The red circle corresponds to <span class="html-italic">N</span> = 12 and <span class="html-italic">x<sub>PEO</sub></span> = 0.5, used in studies of polymer concentration effect on morphology changes.</p>
Full article ">Figure 12
<p>Radius of gyration <span class="html-italic">R<sub>g</sub></span> of the largest cluster vs. chain composition (<span class="html-italic">N</span>: inset).</p>
Full article ">Figure 13
<p>Specific SASA depends on composition and increases rapidly as sphere to rod transition occurs (<span class="html-italic">N</span>: inset).</p>
Full article ">Figure 14
<p>Number of clusters <span class="html-italic">N</span><sub>C</sub> and <span class="html-italic">R<sub>g</sub></span> depend on concentration (6PB6PEO at 300 K).</p>
Full article ">Figure 15
<p>Representative structures realized for copolymer concentrations (in wt%) of 4.0 (<b>a</b>), 7.7 (<b>b</b>), 11.1 (<b>c</b>), 14.3 (<b>d</b>), 20.0 (<b>e</b>), 29.5 (<b>f</b>). (<b>a</b>) Small spherical and irregular micelles, and lamellae; (<b>b</b>) spherical micelles, lamellae, and vesicles; (<b>c</b>) spherical micelles, lamellae, and vesicles; (<b>d</b>) spherical micelles, vesicle strands, rodlike micelles; (<b>e</b>) vesicle strands and thick multilayered rodlike micelles; (<b>f</b>) “giant” composite structure with water-filled hole with diameters ranging from 10 nm to 30 nm.</p>
Full article ">Figure 16
<p>Probability distribution functions of <span class="html-italic">Q</span> of internal (<b>a</b>,<b>b</b>) and external (<b>c</b>,<b>d</b>) PEO segments as well as PB segments (<b>e</b>,<b>f</b>) vesicles corresponding to wt% = 7.7 (<b>a</b>,<b>c</b>,<b>e</b>) and 11.1 (<b>b</b>,<b>d</b>,<b>f</b>). Average <span class="html-italic">Q</span> values (in nm) are (<b>a</b>) 1.0 ± 0.3, (<b>b</b>) 1.1 ± 0.3, (<b>c</b>) 1.1 ± 0.3, (<b>d</b>) 1.0 ± 0.3, (<b>e</b>) 1.8 ± 0.2, and (<b>f</b>) 1.8 ± 0.2.</p>
Full article ">Figure 17
<p>Vesicle strands observed for copolymer concentrations of 14.3 wt% ((<b>a</b>), cross section) and 20.0% ((<b>b</b>), side view). Panels (<b>c</b>–<b>e</b>) show PB monomers in the structure depicted in (<b>b</b>), RDFs along the orange arrow, and RDFs along the radial direction of the vesicle marked by the purple dotted circle in (<b>b</b>), respectively. Green arrows indicate linear dimensions of the structures.</p>
Full article ">Figure 18
<p>Probability distribution functions of <span class="html-italic">Q</span> of PEO (<b>a</b>,<b>c</b>) and PB segments (<b>b</b>,<b>d</b>) for copolymers within the molecular bridge (<b>a</b>,<b>b</b>) and those farther from it (<b>c</b>,<b>d</b>). Average <span class="html-italic">Q</span> values (in nm) are (<b>a</b>) 1.1 ± 0.2, (<b>b</b>) 1.8 ± 0.3, (<b>c</b>) 1.1 ± 0.2, and (<b>d</b>) 1.8 ± 0.2.</p>
Full article ">
Back to TopTop