[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (41)

Search Parameters:
Keywords = gramicidins

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 1572 KiB  
Article
Modifying Membranotropic Action of Antimicrobial Peptide Gramicidin S by Star-like Polyacrylamide and Lipid Composition of Nanocontainers
by Olga V. Vashchenko, Volodymyr P. Berest, Liliia V. Sviechnikova, Nataliya V. Kutsevol, Natalia A. Kasian, Dmitry S. Sofronov and Oleksii Skorokhod
Int. J. Mol. Sci. 2024, 25(16), 8691; https://doi.org/10.3390/ijms25168691 - 9 Aug 2024
Viewed by 269
Abstract
Gramicidin S (GS), one of the first discovered antimicrobial peptides, still shows strong antibiotic activity after decades of clinical use, with no evidence of resistance. The relatively high hemolytic activity and narrow therapeutic window of GS limit its use in topical applications. Encapsulation [...] Read more.
Gramicidin S (GS), one of the first discovered antimicrobial peptides, still shows strong antibiotic activity after decades of clinical use, with no evidence of resistance. The relatively high hemolytic activity and narrow therapeutic window of GS limit its use in topical applications. Encapsulation and targeted delivery may be the way to develop the internal administration of this drug. The lipid composition of membranes and non-covalent interactions affect GS’s affinity for and partitioning into lipid bilayers as monomers or oligomers, which are crucial for GS activity. Using both differential scanning calorimetry (DSC) and FTIR methods, the impact of GS on dipalmitoylphosphatidylcholine (DPPC) membranes was tested. Additionally, the combined effect of GS and cholesterol on membrane characteristics was observed; while dipalmitoylphosphatydylglycerol (DPPG) and cerebrosides did not affect GS binding to DPPC membranes, cholesterol significantly altered the membrane, with 30% mol concentration being most effective in enhancing GS binding. The effect of star-like dextran-polyacrylamide D-g-PAA(PE) on GS binding to the membrane was tested, revealing that it interacted with GS in the membrane and significantly increased the proportion of GS oligomers. Instead, calcium ions affected GS binding to the membrane differently, with independent binding of calcium and GS and no interaction between them. This study shows how GS interactions with lipid membranes can be effectively modulated, potentially leading to new formulations for internal GS administration. Modified liposomes or polymer nanocarriers for targeted GS delivery could be used to treat protein misfolding disorders and inflammatory conditions associated with free-radical processes in cell membranes. Full article
(This article belongs to the Section Molecular Microbiology)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of decapeptide gramicidin S (GS), dipalmitoylphosphatidylcholine (DPPC), dipalmitoylphosphatydylglycerol (DPPG), cholesterol (Chol), and cerebroside (Cer, R1, and R2 are fatty acid residues) and schematic representation of star-like polyanionic dextran-polyacrylamide copolymer D-g-PAA(PE).</p>
Full article ">Figure 2
<p>Dependences of melting temperature of DPPC (T<sub>m</sub>) on GS concentration (c, mol %). Solid circles are plotted for the initial DSC peak and GS monomer binding; data attributed to the binding of GS oligomers are marked with open circles. Means ± SDs for 3–5 independent preparations are shown.</p>
Full article ">Figure 3
<p>Effect of Chol content on thermodynamic parameters of GS binding with the DPPC membrane. The peak temperatures of Chol-enriched (lower four points with correspondent approximate curve, marked T<sub>m</sub>*) and Chol-depleted lipid phases (upper five points with correspondent approximate curve, marked T<sub>m</sub>) are plotted against Chol concentration in the DPPC membrane containing GS at 5 mol %. Data from one representative experiment out of three independent preparations.</p>
Full article ">Figure 4
<p>Values of enthalpies of DSC peaks of DPPC membrane (showed as “membrane”) in the presence of GS and/or polymer D-g-PAA(PE) (showed as “polymer”). Empty columns correspond to the enthalpies of GS “oligomer” peaks. Data are shown for GS content 5 mol %. Means ± SDs of 3–5 independent preparations are shown. Significant changes with <span class="html-italic">p</span> &lt; 0.05 are indicated with * over solid lines for the effect of GS oligomer binding and over dotted line for the effect of D-g-PAA(PE) polymer.</p>
Full article ">Figure 5
<p>Dependences of T<sub>m</sub> (panel <b>a</b>) and T<sub>m</sub>* (panel <b>b</b>) on GS content in DPPC membranes prepared on water subphase (■), 100 мM CaCl<sub>2</sub> (●), and 200 мM CaCl<sub>2</sub> (▲).</p>
Full article ">
12 pages, 1874 KiB  
Article
Morphology Observation of Two-Dimensional Monolayers of Model Proteins on Water Surface as Revealed by Dropping Method
by Yukie Asada, Shinya Tanaka, Hirotaka Nagano, Hiroki Noguchi, Akihiro Yoshino, Keijiro Taga, Yasushi Yamamoto and Zameer Shervani
Bioengineering 2024, 11(4), 366; https://doi.org/10.3390/bioengineering11040366 - 11 Apr 2024
Viewed by 819
Abstract
We have investigated the morphology of two-dimensional monolayers of gramicidin-D (GD) and alamethicin (Al) formed on the water surface by the dropping method (DM) using surface tension measurement (STm), Brewster angle microscopy (BAM), and atomic force microscopy (AFM). Dynamic light scattering (DLS) revealed [...] Read more.
We have investigated the morphology of two-dimensional monolayers of gramicidin-D (GD) and alamethicin (Al) formed on the water surface by the dropping method (DM) using surface tension measurement (STm), Brewster angle microscopy (BAM), and atomic force microscopy (AFM). Dynamic light scattering (DLS) revealed that GD in alcoholic solutions formed a dimeric helical structure. According to the CD and NMR spectroscopies, GD molecules existed in dimer form in methanol and lipid membrane environments. The STm results and BAM images revealed that the GD dimer monolayer was in a liquid expanded (LE) state, whereas the Al monolayer was in a liquid condensed (LC) state. The limiting molecular area (A0) was 6.2 ± 0.5 nm2 for the GD-dimer and 3.6 ± 0.5 nm2 for the Al molecule. The AFM images also showed that the molecular long axes of both the GD-dimer and Al were horizontal to the water surface. The stability of each monolayer was confirmed by the time dependence of the surface pressure (π) observed using the STm method. The DM monolayer preparation method for GD-dimer and Al peptide molecules is a useful technique for revealing how the model biological membrane’s components assemble in two dimensions on the water surface. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of gramicidin-D (GD) and alamethicin (Al).</p>
Full article ">Figure 2
<p><span class="html-italic">π-A</span> isotherm curves of GD and Al monolayers on the water surface at 26 °C using the DM. (<b>a</b>): GD (dimer) monolayer (○); (<b>b</b>): Al monolayer (☐). (- - -) in (<b>a</b>,<b>b</b>): <span class="html-italic">π-A</span> isotherm curves of each GD and Al monolayer using the CM for comparison, respectively.</p>
Full article ">Figure 3
<p><span class="html-italic">C<sub>s</sub></span><sup>−1</sup> profiles of GD and Al monolayers calculated from <a href="#bioengineering-11-00366-f002" class="html-fig">Figure 2</a>. (<b>a</b>): GD (dimer) monolayer (○); (<b>b</b>): Al monolayer (☐). (• • •) in (<b>a</b>,<b>b</b>): <span class="html-italic">C<sub>s</sub></span><sup>−1</sup> profiles of each GD and Al monolayer using the CM for comparison.</p>
Full article ">Figure 4
<p>BAM images of GD and Al monolayers on the water surface at 26 °C using the DM. (<b>a</b>–<b>c</b>): GD (dimer) monolayer; (<b>d</b>–<b>f</b>): Al monolayer. Inserted surface pressures correspond to the value in each <span class="html-italic">π-A</span> isotherm curve shown in <a href="#bioengineering-11-00366-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 5
<p>AFM images of GD and Al monolayers scooped on the HOPG surface after each monolayer formation at 26 °C using the DM. (<b>a</b>,<b>b</b>): GD (dimer) monolayer; (<b>c</b>): Al monolayer. Observation range: 2 µm × 2 µm. Inserted surface pressures correspond to the values in each π<span class="html-italic">-A</span> isotherm curve in <a href="#bioengineering-11-00366-f002" class="html-fig">Figure 2</a>. Red arrow and cross-section profile: thickness (height) of each GD and Al monolayer in the arrow range.</p>
Full article ">
24 pages, 3596 KiB  
Article
Intrinsic Lipid Curvature and Bilayer Elasticity as Regulators of Channel Function: A Comparative Single-Molecule Study
by Mohammad Ashrafuzzaman, Roger E. Koeppe and Olaf S. Andersen
Int. J. Mol. Sci. 2024, 25(5), 2758; https://doi.org/10.3390/ijms25052758 - 27 Feb 2024
Cited by 2 | Viewed by 855
Abstract
Perturbations in bilayer material properties (thickness, lipid intrinsic curvature and elastic moduli) modulate the free energy difference between different membrane protein conformations, thereby leading to changes in the conformational preferences of bilayer-spanning proteins. To further explore the relative importance of curvature and elasticity [...] Read more.
Perturbations in bilayer material properties (thickness, lipid intrinsic curvature and elastic moduli) modulate the free energy difference between different membrane protein conformations, thereby leading to changes in the conformational preferences of bilayer-spanning proteins. To further explore the relative importance of curvature and elasticity in determining the changes in bilayer properties that underlie the modulation of channel function, we investigated how the micelle-forming amphiphiles Triton X-100, reduced Triton X-100 and the HII lipid phase promoter capsaicin modulate the function of alamethicin and gramicidin channels. Whether the amphiphile-induced changes in intrinsic curvature were negative or positive, amphiphile addition increased gramicidin channel appearance rates and lifetimes and stabilized the higher conductance states in alamethicin channels. When the intrinsic curvature was modulated by altering phospholipid head group interactions, however, maneuvers that promote a negative-going curvature stabilized the higher conductance states in alamethicin channels but destabilized gramicidin channels. Using gramicidin channels of different lengths to probe for changes in bilayer elasticity, we found that amphiphile adsorption increases bilayer elasticity, whereas altering head group interactions does not. We draw the following conclusions: first, confirming previous studies, both alamethicin and gramicidin channels are modulated by changes in lipid bilayer material properties, the changes occurring in parallel yet differing dependent on the property that is being changed; second, isolated, negative-going changes in curvature stabilize the higher current levels in alamethicin channels and destabilize gramicidin channels; third, increases in bilayer elasticity stabilize the higher current levels in alamethicin channels and stabilize gramicidin channels; and fourth, the energetic consequences of changes in elasticity tend to dominate over changes in curvature. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p>Schematic models of gramicidin and alamethicin channels. (<b>A</b>) Top: sequence of [Val<sup>1</sup>]gA [<a href="#B34-ijms-25-02758" class="html-bibr">34</a>], the major gramicidin species in naturally occurring mixture of peptides [<a href="#B35-ijms-25-02758" class="html-bibr">35</a>]; f is formyl, ea ethanolamine and the D-amino acids are underlined. Bottom: gramicidin channels form and disappear, as indicated by the arrows, by a transmembrane association/dissociation [<a href="#B36-ijms-25-02758" class="html-bibr">36</a>]. Left, atomic resolution structures of the β<sup>6.3</sup>-helical monomers, the two subunits are depicted some distance apart; right, atomic resolution structure of the β<sup>6.3</sup>-helical conducting dimer. The carbons in the two subunits are colored green and yellow, respectively, with the carbon atoms in the Trp side chains emphasized. Blue is nitrogen, red is oxygen and white is hydrogen. (<b>B</b>) Top: sequence of alamethicin I [<a href="#B37-ijms-25-02758" class="html-bibr">37</a>], the major species of alamethicin; ac is acetate, Aib α-isobutyric acid and Pheol phenylalcohol. Bottom: different interconverting oligomeric states, as indicated by the arrows, of the bilayer-spanning channel. The number of subunits may change by the association/dissociation of bilayer-spanning subunits or oligomers or by the accretion of subunits at the bilayer/solution interface that inserts into the bilayer [<a href="#B33-ijms-25-02758" class="html-bibr">33</a>,<a href="#B38-ijms-25-02758" class="html-bibr">38</a>].</p>
Full article ">Figure 2
<p>Amphiphile-induced changes in alamethicin channel activity. Cpsn, TX100 and rTX100 increase Alm channel activity. Top four records: 40 s recorded before the addition of amphiphile and after the addition of the indicated amphiphile (the control traces were similar for each amphiphile trace). The calibration bars in the top trace apply to all four traces. Bottom four traces show the effect of the amphiphiles at higher resolution; calibration bars in the control trace segment apply to all the trace segments. The stippled lines denote different current levels; they do not vary with amphiphile addition (<a href="#ijms-25-02758-t001" class="html-table">Table 1</a>) (DOPC, 1.0 M NaCl, pH 7.0, 150 mV).</p>
Full article ">Figure 3
<p>Current level (all-point) histograms showing the effects of TX100 on Alm channel function, results from one experiment. Top: results from a 40 s recording before the addition of TX100. Bottom: results from a 40 s recording in the same membrane a few min after the addition of 10 μM TX100. The right panels show the same results as the left but at an expanded scale for the ordinate. nc denotes the no-channel current level; the plots were aligned such that the nc peak is centered at 0 pA. The numbers over the peaks denote the identity of the channel state; two numbers indicate that the peak results from the superposition of two different channels (DOPC, 1.0 M NaCl, pH 7.0, 150 mV).</p>
Full article ">Figure 4
<p>The variability of Alm channel activity as a function of time in the absence or presence of amphiphile. The ordinate denotes the channel activity, the time the channels reside in any conducting state relative to the no-channel state (<span class="html-italic">R</span><sub>Alm</sub>, Equation (2)) over a 10 s time interval, normalized to the average activity over the total 80 s recording time. Mean ± S.D. based on at least three independent experiments at each condition (DOPC, 1.0 M NaCl, pH 7.0, 150 mV).</p>
Full article ">Figure 5
<p>Effect of amphiphiles (TX100, rTX100 or Cpsn) on Alm channel activity. The ordinate displays the channel activity (Equation (2)) in the presence of amphiphile divided by the activity in the absence of amphiphile (<math display="inline"><semantics> <mrow> <msubsup> <mi>R</mi> <mrow> <mi>Alm</mi> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>R</mi> <mrow> <mi>Alm</mi> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> </mrow> </semantics></math>, cf. Equation (3)). Mean ± S.D. based on at least three independent experiments, with one to three measurements, at each condition (DOPC, 1.0 M NaCl, pH 7.0, 150 mV).</p>
Full article ">Figure 6
<p>Effect of TX100, rTX100 or Cpsn on the distribution of Alm channel current levels relative to the nc level. The ordinate depicts the changes in <math display="inline"><semantics> <mrow> <mi>ln</mi> <mfenced close="}" open="{"> <mrow> <mfenced> <mrow> <msubsup> <mi>A</mi> <mi mathvariant="normal">k</mi> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>A</mi> <mrow> <mi>nc</mi> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> </mrow> </mfenced> <mo>/</mo> <mfenced> <mrow> <msubsup> <mi>A</mi> <mi mathvariant="normal">k</mi> <mrow> <mi>cntl</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>A</mi> <mrow> <mi>nc</mi> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> </mrow> </mfenced> </mrow> </mfenced> </mrow> </semantics></math>, <span class="html-italic">k</span> = 0, 1, 2, 3, cf. Equation (6). Mean ± S.D. based on at least three independent experiments, each with one to three measurements, at each condition (DOPC, 1.0 M NaCl, pH 7.0, 150 mV).</p>
Full article ">Figure 7
<p>Effect of TX100, rTX100 or Cpsn on the distribution of time spent in different Alm current levels relative to the time spent in level 1. The ordinate shows (<math display="inline"><semantics> <mrow> <mi>ln</mi> <mfenced close="}" open="{"> <mrow> <mo stretchy="false">(</mo> <msubsup> <mi>A</mi> <mi mathvariant="normal">k</mi> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>A</mi> <mn>1</mn> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <mo stretchy="false">(</mo> <msubsup> <mi>A</mi> <mi mathvariant="normal">k</mi> <mrow> <mi>cntl</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>A</mi> <mn>1</mn> <mrow> <mi>cntl</mi> </mrow> </msubsup> <mo stretchy="false">)</mo> </mrow> </mfenced> </mrow> </semantics></math>, <span class="html-italic">k</span> = 2, 3cf. Equation (7)). Left, results for TX100 and rTX100. Right, results for Cpsn. Mean ± S.D. based on at least three independent experiments, each with one to three measurements, at each condition (DOPC, 1.0 M NaCl, pH 7.0, 150 mV).</p>
Full article ">Figure 8
<p>TX100 and Cpsn produce similar increases in gA channel activity. The three traces denote 60 s current traces recorded in the absence or presence of either 10 µM TX100 or 30 μM Cpsn (the control trace is from the TX100 experiment; similar single-channel activity was observed in the control trace for Cpsn). The experiments were performed using two different gA analogs, AgA(15) and gA<sup>−</sup>(13), which were added together to both sides of the bilayer. AgA(15) and gA<sup>−</sup>(13) channels can be distinguished by their current transition amplitudes (indicated by the horizontal dashed lines in the control current trace: blue for AgA(15) channels; red for gA<sup>−</sup>(13) channels). The calibration bars in the bottom trace apply to all traces (DOPC, 1.0 M NaCl, pH 7.0, 200 mV).</p>
Full article ">Figure 9
<p>Effect of TX100, rTX100 and Cpsn on the lifetimes, appearance rates, channel activities and the change in the free energies of formation (Equation (9)) of AgA(15) and gA<sup>−</sup>(13) channels. (Panel (<b>A</b>)) shows results for τ<sub>AM</sub>/τ<sub>cntl</sub>; (panel (<b>B</b>)) shows results for <span class="html-italic">f</span><sub>AM</sub>/<span class="html-italic">f</span><sub>cntl</sub>; (panel (<b>C</b>)) shows results for τ<sub>AM</sub>·<span class="html-italic">f</span><sub>AM</sub>/τ<sub>cntl</sub>·<span class="html-italic">f</span><sub>cntl</sub>. To facilitate comparison of the results for the 13-residue and 15-residue channels, the results are displayed using logarithmic y axes. In the control experiments for TX100, τ<sub>15</sub> and τ<sub>13</sub> were 160 ± 13 ms and 11.6 ± 1.4 ms, respectively; in the rTX100 experiments, τ<sub>15</sub> and τ<sub>13</sub> were 131 ± 7 ms and 11.0 ± 0.4 ms, respectively; in the Cpsn experiments, τ<sub>15</sub> and τ<sub>13</sub> were 206 ± 14 ms and 15.5 ± 0.2 ms, respectively. Filled symbols—results for AgA(15) channels; open symbols—results for gA<sup>−</sup>(13) channels. Mean ± S.D. based on at least three independent experiments, each with three or more measurements, at each condition (DOPC, 1.0 M NaCl, pH 7.0, 200 mV).</p>
Full article ">Figure 10
<p>Amphiphiles produce larger relative changes in the lifetimes of gA<sup>−</sup>(13) channels, <math display="inline"><semantics> <mrow> <mi>ln</mi> <mo>{</mo> <msubsup> <mo>τ</mo> <mrow> <mn>13</mn> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <mo> </mo> <msubsup> <mo>τ</mo> <mrow> <mn>13</mn> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> <mo>}</mo> </mrow> </semantics></math>, as compared to AgA(15) channels, <math display="inline"><semantics> <mrow> <mi>ln</mi> <mo>{</mo> <msubsup> <mo>τ</mo> <mrow> <mn>15</mn> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <mo> </mo> <msubsup> <mo>τ</mo> <mrow> <mn>15</mn> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> <mo>}</mo> </mrow> </semantics></math>, based on results in <a href="#ijms-25-02758-f009" class="html-fig">Figure 9</a>. The red, blue and green dashed lines denote linear fits to the result for TX100, rTX100 and Cpsn, respectively. For TX100, the slope was 1.64 ± 0.11, <span class="html-italic">r</span><sup>2</sup> = 0.986 (90% confidence interval for the slope, 1.29–1.99); for rTX100, the slope was 1.21 ± 0.04, <span class="html-italic">r</span><sup>2</sup> = 0.997 (90% confidence interval for the slope, 1.09–1.33); for Cpsn, the slope was 1.26 ± 0.05, <span class="html-italic">r</span><sup>2</sup> = 0.995 (90% confidence interval for the slope, 1.10–1.42). The black interrupted line has a slope of 1. (DOPC, 1.0 M NaCl, pH 7.0, 200 mV).</p>
Full article ">Figure 11
<p>Amphiphile-induced changes in Alm function as functions of the changes in AgA(15) channel lifetimes. (<b>A</b>): Effect of TX100 (4, 10, 30 µM), rTX100 (4, 10, 30 µM) or Cpsn (10, 30, 100 µM) on Alm channel activity, expressed as <math display="inline"><semantics> <mrow> <mi>ln</mi> <mfenced close="}" open="{"> <mrow> <msubsup> <mi>R</mi> <mrow> <mi>Alm</mi> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>R</mi> <mrow> <mi>Alm</mi> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> </mrow> </mfenced> </mrow> </semantics></math>, cf. Equation (3), as functions of the corresponding changes in <math display="inline"><semantics> <mrow> <mi>ln</mi> <mfenced close="}" open="{"> <mrow> <msubsup> <mo>τ</mo> <mrow> <mn>15</mn> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mo>τ</mo> <mrow> <mn>15</mn> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> </mrow> </mfenced> </mrow> </semantics></math>. Based on results in <a href="#ijms-25-02758-f005" class="html-fig">Figure 5</a>, <a href="#ijms-25-02758-f007" class="html-fig">Figure 7</a> and <a href="#ijms-25-02758-f009" class="html-fig">Figure 9</a>. The red, blue and green dashed lines denote linear fits to the results, including 0 µM, for TX100, rTX100 and Cpsn, respectively. For TX100, the slope was 1.74, ± 0.08; <span class="html-italic">r</span><sup>2</sup> = 0.994 (90% confidence interval for the slope, 1.15–3.60); for rTX100, the slope was 1.61 ± 0.22, <span class="html-italic">r</span><sup>2</sup> = 0.940 (90% confidence interval for the slope, 0.89–2.33); for Cpsn, the slope was 2.37 ± 0.40, <span class="html-italic">r</span><sup>2</sup> = 0.920 (90% confidence interval for the slope, 1.15–3.60) (DOPC, 1.0 M NaCl, pH 7.0). (<b>B</b>): Effect of TX100, rTX100 or Cpsn on the distribution between Alm current level 1 and 2, expressed as <math display="inline"><semantics> <mrow> <mi>ln</mi> <mfenced close="}" open="{"> <mrow> <mo stretchy="false">(</mo> <msubsup> <mi>A</mi> <mrow> <mrow> <mo> </mo> <mn>2</mn> </mrow> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>A</mi> <mrow> <mrow> <mo> </mo> <mn>1</mn> </mrow> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo stretchy="false">)</mo> <mo>/</mo> <mo stretchy="false">(</mo> <msubsup> <mi>A</mi> <mrow> <mrow> <mo> </mo> <mn>2</mn> </mrow> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>A</mi> <mrow> <mrow> <mo> </mo> <mn>1</mn> </mrow> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> <mo stretchy="false">)</mo> </mrow> </mfenced> </mrow> </semantics></math>, cf. Equation (7), as functions of the corresponding changes in <math display="inline"><semantics> <mrow> <mi>ln</mi> <mfenced close="}" open="{"> <mrow> <msubsup> <mo>τ</mo> <mrow> <mn>15</mn> </mrow> <mrow> <mi>AM</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mo>τ</mo> <mrow> <mn>15</mn> </mrow> <mrow> <mi>cntl</mi> </mrow> </msubsup> </mrow> </mfenced> </mrow> </semantics></math>. The dashed lines denote linear fits to the results, including 0 µM. For TX100, the slope was 0.51 ± 0.06, <span class="html-italic">r</span><sup>2</sup> = 0.959 (90% confidence interval for the slope, 0.33–0.70); for rTX100, the slope was 0.37 ± 0.08, <span class="html-italic">r</span><sup>2</sup> = 0872 (90% confidence interval for the slope, 0.13–0.61); for Cpsn, the slope was 0.59 ± 0.12, <span class="html-italic">r</span><sup>2</sup> = 0.920 (90% confidence interval for the slope, 0.24–0.95) (DOPC, 1.0 M NaCl, pH 7.0).</p>
Full article ">
21 pages, 3559 KiB  
Article
Dimeric Tubulin Modifies Mechanical Properties of Lipid Bilayer, as Probed Using Gramicidin A Channel
by Tatiana K. Rostovtseva, Michael Weinrich, Daniel Jacobs, William M. Rosencrans and Sergey M. Bezrukov
Int. J. Mol. Sci. 2024, 25(4), 2204; https://doi.org/10.3390/ijms25042204 - 12 Feb 2024
Viewed by 902
Abstract
Using the gramicidin A channel as a molecular probe, we show that tubulin binding to planar lipid membranes changes the channel kinetics—seen as an increase in the lifetime of the channel dimer—and thus points towards modification of the membrane’s mechanical properties. The effect [...] Read more.
Using the gramicidin A channel as a molecular probe, we show that tubulin binding to planar lipid membranes changes the channel kinetics—seen as an increase in the lifetime of the channel dimer—and thus points towards modification of the membrane’s mechanical properties. The effect is more pronounced in the presence of non-lamellar lipids in the lipid mixture used for membrane formation. To interpret these findings, we propose that tubulin binding redistributes the lateral pressure of lipid packing along the membrane depth, making it closer to the profile expected for lamellar lipids. This redistribution happens because tubulin perturbs the lipid headgroup spacing to reach the membrane’s hydrophobic core via its amphiphilic α-helical domain. Specifically, it increases the forces of repulsion between the lipid headgroups and reduces such forces in the hydrophobic region. We suggest that the effect is reciprocal, meaning that alterations in lipid bilayer mechanics caused by membrane remodeling during cell proliferation in disease and development may also modulate tubulin membrane binding, thus exerting regulatory functions. One of those functions includes the regulation of protein–protein interactions at the membrane surface, as exemplified by VDAC complexation with tubulin. Full article
(This article belongs to the Special Issue Modulation of Protein Structure and Function by Lipids)
Show Figures

Figure 1

Figure 1
<p>Tubulin increases the lifetime of grA channels and decreases their conductance in DOPE membranes but not in DOPC membranes: (<b>A</b>) Current traces of grA channels in DOPE membrane before (trace a) and after (trace b) addition of 30 nM tubulin. Tubulin notably increases grA lifetime and decreases channel conductance in the DOPE membrane. Tubulin also induces fast current flickering that can be better seen at a finer time scales in inset d in comparison with the control trace in inset c. (<b>B</b>) Current traces of grA channels in DOPC membrane before (trace a) and after (trace b) addition of 50 nM tubulin. The addition of 50 nM of tubulin does not appreciably change grA channel parameters in the DOPC membranes. The applied voltage was 100 mV. Tubulin was added to the <span class="html-italic">cis</span> compartment. Current records were digitally filtered using an averaging time of 10 ms. Dashed lines indicate zero current level and dotted lines indicate the currents through single grA channels. Here (and elsewhere) the medium consisted of 1 M KCl buffered with 5 mM HEPES at pH 7.4.</p>
Full article ">Figure 2
<p>Effect of tubulin on the lifetime and conductance of grA channels, which depends on DOPE content in DOPE/DOPC mixture. In DOPE membranes, tubulin increases grA lifetime (<b>A</b>) and decreases conductance (<b>B</b>) in a dose-dependent manner that displays saturation at about 20 nM tubulin concentration. There is virtually no effect of tubulin on the channel lifetime and conductance in DOPC membranes. Channel conductance is given as its ratio in the presence of tubulin to that in the absence of tubulin. (<b>C</b>,<b>D</b>) Effect of 30 nM tubulin on the channel lifetime (<b>C</b>) and conductance (<b>D</b>) increases with PE content in the PE/PC mixture. Lines are drawn to guide the eye. Data are the mean values obtained in 3–5 experiments ± S.E. Experimental conditions are as in <a href="#ijms-25-02204-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Tubulin induces fast flickering of the ionic current through grA channels in a dose- and voltage-dependent manner: (<b>A</b>) grA current traces were obtained at indicated tubulin concentrations in the DPhPC membrane at 200 mV applied voltage. The flickering increases with tubulin concentration and the corresponding decrease of the average conductance (black dotted line indicates conductance at 0 and 3 nM tubulin and red dotted line indicates conductance at 10 and 50 nM tubulin). Current records were filtered with a digital 8-pole Bessel filter at 1 kHz. (<b>B</b>) Power spectral densities of current fluctuations at 50 nM of tubulin (upper trace) can be approximated by a Lorentzian spectrum (smooth line through the data) with a corner frequency of <span class="html-italic">f<sub>c</sub></span> = 600 Hz. (<b>C</b>) Tubulin-induced current fluctuations in the grA channel increase with applied voltage. Power spectral densities of current fluctuations of the single grA channel obtained at applied voltages as indicated in the presence of 50 nM tubulin. Solid lines are fitted with Lorentzian spectra. Current records were filtered with a digital 8-pole Bessel filter at 2 kHz. The medium consisted of 1 M KCl buffered with 5 mM HEPES at pH 7.4.</p>
Full article ">Figure 4
<p>Tubulin-S changes the lifetime and conductance of grA channels in DOPE membranes. Similarly to tubulin, tubulin-S increases channel lifetime (<b>A</b>) and reduces its conductance (<b>B</b>) in comparison with the control. Channel lifetime and relative conductance (<span class="html-italic">G<sub>Tub</sub></span>/<span class="html-italic">G<sub>Cntr</sub></span>) were measured in the presence of 50 nM tubulin and 40 nM tubulin-S in the <span class="html-italic">cis</span> compartment. Data are the mean values obtained in 3–4 experiments ± S.E. Other experimental conditions were as in <a href="#ijms-25-02204-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 5
<p>Tubulin-induced fast flickering of grA channels in diC(22:1)PC bilayers. (<b>A</b>) Current traces of a single grA channel in a diC(22:1)PC bilayer before (trace a) and after (trace b) addition of 30 nM tubulin to the <span class="html-italic">cis</span> compartment. The addition of tubulin induces rapid events of grA channel closure to a zero-current level (indicated by black dashed line), as shown in trace c at a finer time scale. The dotted black line indicates open channel conductance. The applied voltage was 200 mV. Current records were filtered with a digital 8-pole Bessel filter at 2 kHz. (<b>B</b>) Power spectral density of tubulin-induced current fluctuations, which depends on the polarity of the applied voltage. Solid gray lines represent the fits to Lorentzian spectra.</p>
Full article ">Figure 6
<p>Illustration of stable α-tubulin binding to the DOPE membrane surfaces from ~800 ns of all-atom ANTON MD simulations. The insertion helical region (A330–W346) is colored according to chemical functionality (negative residues in red, positive residues in blue, and hydrophobic/aromatic in orange). The location of W346 is shown in sphere mode. β-tubulin, shown in the magenta-colored ribbon, is added based on RMSF alignment for α-tubulin in the dimer. MD simulations were performed for α-tubulin only. The unstructured C-terminal tails are not shown. Adapted with permission from Hoogerheide et al., <span class="html-italic">Proc. Natl. Acad. Sci. USA</span> (2017) [<a href="#B29-ijms-25-02204" class="html-bibr">29</a>]. Copyright (2017) National Academy of Sciences.</p>
Full article ">Figure 7
<p>α-Tubulin peptide increases grA lifetime: (<b>A</b>) Current traces of grA channels in DOPE membrane before (Control) and after the addition of 10 and 20 μM of α-tubulin peptide to both sides of the DOPE membrane. Current records were filtered with a digital 8-pole Bessel filter at 1 kHz. Dashed lines indicate zero current level. The membrane bathing solution contained 1 M KCl buffered with 5 mM HEPES at pH 7.4. The applied voltage was 100 mV. α-tubulin peptide was dissolved in DMSO. (<b>B</b>) α-tubulin peptide increases grA lifetime in a dose-dependent manner. Bars and error bars are the mean and standard deviation from the mean; the symbols represent data points of 4 independent experiments. Control measurements with the addition of DMSO aliquots corresponding to 20 μM of α-tubulin peptide addition do not show the effect on grA lifetime.</p>
Full article ">Figure 8
<p>Binding curves for α-tubulin peptide interaction with DOPE and DOPC membranes, as measured by BOA. The transmembrane potential ΔΨ changes due to α-tubulin peptide binding to the DOPE/DOPC (4:1) (reddish symbols) membranes and does not change for DOPC (bluish symbols) membranes. Membranes were formed in 150 mM KCl buffered with 5 mM HEPES at pH 7.4. α-Tubulin peptide dissolved in DMSO was added to the <span class="html-italic">cis</span> side of the membrane. Aliquots of DMSO were correspondingly added to the <span class="html-italic">trans</span> side. Large red and blue circles and error bars are the mean and standard deviation from the mean for DOPE/DOPC and DOPC membranes, respectively; they represent data points of 7 individual experiments for DOPE/DOPC membranes and 6 experiments for DOPC membranes. The solid line is a fit to the binding equation with <span class="html-italic">K<sub>d</sub></span> = 156 μM.</p>
Full article ">Figure 9
<p>Schematics of the effect of tubulin dimer on grA lifetime and conductance: (<b>A</b>) Binding of α-β-tubulin heterodimers to the DOPE membrane reduces packing stress of the lipid tails, which is observed as the increase of grA channel (shown in red) lifetime. (<b>B</b>) In the case of diC(22:1)PC membranes, binding of the tubulin dimers is limited to the regions of membranes where headgroup packing is distorted by grA channel presence in the region of the lipid funnel forming the entrance to the channel. This limitation leads to the unchanged integral properties of the membrane and unchanged grA lifetime (<a href="#ijms-25-02204-t001" class="html-table">Table 1</a>); however, the localized binding is clearly manifested via transient channel blockages (<a href="#ijms-25-02204-f005" class="html-fig">Figure 5</a>) from the bulky body of the tubulin dimer. Created with <a href="http://Biorender.com" target="_blank">Biorender.com</a>.</p>
Full article ">
9 pages, 2326 KiB  
Article
Gramicidin, a Bactericidal Antibiotic, Is an Antiproliferative Agent for Ovarian Cancer Cells
by Min Sung Choi, Chae Yeon Lee, Ji Hyeon Kim, Yul Min Lee, Sukmook Lee, Hyun Jung Kim and Kyun Heo
Medicina 2023, 59(12), 2059; https://doi.org/10.3390/medicina59122059 - 22 Nov 2023
Viewed by 1340
Abstract
Background and Objectives: Gramicidin, a bactericidal antibiotic used in dermatology and ophthalmology, has recently garnered attention for its inhibitory actions against cancer cell growth. However, the effects of gramicidin on ovarian cancer cells and the underlying mechanisms are still poorly understood. We [...] Read more.
Background and Objectives: Gramicidin, a bactericidal antibiotic used in dermatology and ophthalmology, has recently garnered attention for its inhibitory actions against cancer cell growth. However, the effects of gramicidin on ovarian cancer cells and the underlying mechanisms are still poorly understood. We aimed to elucidate the anticancer efficacy of gramicidin against ovarian cancer cells. Materials and Methods: The anticancer effect of gramicidin was investigated through an in vitro experiment. We analyzed cell proliferation, DNA fragmentation, cell cycle arrest and apoptosis in ovarian cancer cells using WST-1 assay, terminal deoxynucleotidyl transferase dUTP nick and labeling (TUNEL), DNA agarose gel electrophoresis, flow cytometry and western blot. Results: Gramicidin treatment induces dose- and time-dependent decreases in OVCAR8, SKOV3, and A2780 ovarian cancer cell proliferation. TUNEL assay and DNA agarose gel electrophoresis showed that gramicidin caused DNA fragmentation in ovarian cancer cells. Flow cytometry demonstrated that gramicidin induced cell cycle arrest. Furthermore, we confirmed via Western blot that gramicidin triggered apoptosis in ovarian cancer cells. Conclusions: Our results strongly suggest that gramicidin exerts its inhibitory effect on cancer cell growth by triggering apoptosis. Conclusively, this study provides new insights into the previously unexplored anticancer properties of gramicidin against ovarian cancer cells. Full article
(This article belongs to the Section Oncology)
Show Figures

Figure 1

Figure 1
<p>Gramicidin inhibited the proliferation of ovarian cancer (OC) cells. (<b>A</b>) OVCAR8, (<b>B</b>) SKOV3, and (<b>C</b>) A2780 cells were treated with the indicated concentrations of gramicidin for 72 h. (<b>D</b>) OVCAR8, (<b>E</b>) SKOV3, and (<b>F</b>) A2780 cells were treated with 0, 0.33, 1, or 3 μM gramicidin for 0, 24, 48, or 72 h. Relative cell proliferation rates were assessed using the WST-1 assay, which monitors mitochondrial succinate reductase activity at 450 nm. The data represent the mean ± SEM of duplicates from two independent experiments.</p>
Full article ">Figure 2
<p>Gramicidin-induced apoptosis, as indicated by DNA fragmentation in OC cells. Apoptotic cells were detected by DNA fragmentation using the TUNEL assay and DNA agarose gel electrophoresis. (<b>A</b>) OVCAR8 and (<b>B</b>) SKOV3 were treated with 0, 0.1, 0.3, or 1 μM gramicidin for 48 h and imaged using confocal microscopy. TUNEL-positive nuclei are shown in green, and total nuclei stained with Hoechst 33,342 are presented in blue (left panel). Relative TUNEL intensities, normalized to the negative control, were quantified in five randomly selected microscopic fields (right panel). (<b>C</b>) OVCAR8 and SKOV3 were treated with 0, 0.1, or 0.3 μM of gramicidin for 48 h, and genomic DNA was isolated. DNA fragmentation was assessed by agarose gel electrophoresis with ethidium bromide staining. The data are presented as mean ± SEM of a minimum of three independent experiments. Statistical significance was calculated using one-way ANOVA (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Gramicidin increased sub-G1 in OC cells. (<b>A</b>) OVCAR8 and (<b>B</b>) SKOV3 were treated with 0, 0.1, 0.3, or 1 μM of gramicidin for 48 h and subsequently stained with propidium iodide (PI) for cell cycle analysis using flow cytometry. The distribution of cells at different cell cycle phases was determined, and the percentages of cells in each phase relative to the total number of cells were calculated. The data are presented as mean ± SEM of three independent experiments.</p>
Full article ">Figure 4
<p>Gramicidin-induced cleavage of caspase-3 and PARP in OC cells. (<b>A</b>) OVCAR8 and (<b>B</b>) SKOV3 cells were treated with 0, 0.1, 0.3, or 1 μM of gramicidin for 24 h. Whole-cell lysates were analyzed via immunoblotting against caspase-3, cleaved caspase-3, PARP, cleaved PARP, or β-actin. The results represent a minimum of three independent experiments.</p>
Full article ">
20 pages, 5011 KiB  
Article
Genomic Based Analysis of the Biocontrol Species Trichoderma harzianum: A Model Resource of Structurally Diverse Pharmaceuticals and Biopesticides
by Suhad A. A. Al-Salihi and Fabrizio Alberti
J. Fungi 2023, 9(9), 895; https://doi.org/10.3390/jof9090895 - 31 Aug 2023
Cited by 3 | Viewed by 1535
Abstract
Fungi represents a rich repository of taxonomically restricted, yet chemically diverse, secondary metabolites that are synthesised via specific metabolic pathways. An enzyme’s specificity and biosynthetic gene clustering are the bottleneck of secondary metabolite evolution. Trichoderma harzianum M10 v1.0 produces many pharmaceutically important molecules; [...] Read more.
Fungi represents a rich repository of taxonomically restricted, yet chemically diverse, secondary metabolites that are synthesised via specific metabolic pathways. An enzyme’s specificity and biosynthetic gene clustering are the bottleneck of secondary metabolite evolution. Trichoderma harzianum M10 v1.0 produces many pharmaceutically important molecules; however, their specific biosynthetic pathways remain uncharacterised. Our genomic-based analysis of this species reveals the biosynthetic diversity of its specialised secondary metabolites, where over 50 BGCs were predicted, most of which were listed as polyketide-like compounds associated clusters. Gene annotation of the biosynthetic candidate genes predicted the production of many medically/industrially important compounds including enterobactin, gramicidin, lovastatin, HC-toxin, tyrocidine, equisetin, erythronolide, strobilurin, asperfuranone, cirtinine, protoilludene, germacrene, and epi-isozizaene. Revealing the biogenetic background of these natural molecules is a step forward towards the expansion of their chemical diversification via engineering their biosynthetic genes heterologously, and the identification of their role in the interaction between this fungus and its biotic/abiotic conditions as well as its role as bio-fungicide. Full article
(This article belongs to the Special Issue Genomics Analysis of Fungi)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Trichoderma</span> species SMs core enzymes and their associated BGCs.</p>
Full article ">Figure 2
<p>Maximum likelihood tree of five conserved genes (chitinase gene {chi18-5}, endochitinase1 {ech1}, β-tubulin, glyceraldehyde-3-phosphate dehydrogenase {gpdh}, and translation elongation factor {tef} of <span class="html-italic">T. harzianum</span> M10 v1.0, <span class="html-italic">T. harzianum</span> CBS226.95, <span class="html-italic">T. harzianum</span> TR274, <span class="html-italic">T. harzianum</span> T22, and <span class="html-italic">T. afroharzianum</span>. Nodes labels indicate species taxon-protein ID-gene function.</p>
Full article ">Figure 3
<p>Maximum likelihood tree of the core NRPS/NRPS-like protein sequences of <span class="html-italic">T. harzianum</span> M10 v1.0 and other experimentally described NRPSs of different microbial species. Nodes labels indicate species taxon-protein ID-chemical. <span class="html-italic">T. harzianum</span> M10 predicted proteins are in red.</p>
Full article ">Figure 4
<p>Organization of the genetic structure of the predicted non-ribosomal peptide BGCs of <span class="html-italic">Trichoderma harzianum</span> M10 v1.0 Sizes and directions of arrows represent different genes sizes and their 5′-3′ direction. Full description of gene function is provided in <a href="#app1-jof-09-00895" class="html-app">Tables S2–S20 in the supplementary information</a>.</p>
Full article ">Figure 5
<p>Maximum likelihood tree of the core PKS/PKS-like of <span class="html-italic">T. harzianum</span> M10 v1.0 and other experimentally described NRPS protein sequences of different microbial species Nodes labels indicate species taxon-protein ID-chemical. <span class="html-italic">T. harzianum</span> M10 predicted proteins are in red.</p>
Full article ">Figure 6
<p>Organization of the genetic structure of the predicted polyketide synthase (PKS/PKS-like) BGCs of <span class="html-italic">Trichoderma harzianum</span> M10 v1.0. Sizes and directions of arrows represent different gene sizes and their 5′-3′ direction. Full description of gene function is provided in <a href="#app1-jof-09-00895" class="html-app">Tables S21–S41 in the supplementary information</a>.</p>
Full article ">Figure 7
<p>Organization of the genetic structure of the predicted hybrid polyketide synthase (HrPKS) BGCs of <span class="html-italic">Trichoderma harzianum</span> M10 v1.0. Sizes and directions of arrows represent different genes sizes and their 5′-3′ direction. Full description of gene function is provided in <a href="#app1-jof-09-00895" class="html-app">Tables S42–S47 in the supplementary information</a>.</p>
Full article ">Figure 8
<p>Maximum likelihood tree of the core terpene cyclase (TC) of the <span class="html-italic">T. harzianum</span> M10 v1.0 and other experimentally described TC protein sequences of different microbial species. Nodes labels indicate Species taxon-protein ID-chemical. <span class="html-italic">T. harzianum</span> M10 predicted proteins are in red.</p>
Full article ">Figure 9
<p>Organization of the genetic structure of the predicted terpene cyclase (TC) BGCs of <span class="html-italic">Trichoderma harzianum</span> M10 v1.0. Sizes and directions of arrows represent different gene sizes and their 5′-3′ direction. Full description of gene function is provided in <a href="#app1-jof-09-00895" class="html-app">Tables S48–S52 in the supplementary information</a>.</p>
Full article ">Figure 10
<p>Organization of the genetic structure of the predicted dimethylallyltryptophan (DMAT) BGC of <span class="html-italic">Trichoderma harzianum</span> M10 v1.0. Sizes and directions of arrows represent different gene sizes and their 5′-3′ direction. Full description of gene function is provided in <a href="#app1-jof-09-00895" class="html-app">Table S53 in the supplementary information</a>.</p>
Full article ">Figure 11
<p>Cblaster analysis of three types of SMs enzymes that had high percentage matches with the <span class="html-italic">T. harzianum</span> M10 v1.0 SMs enzymes in our phylogenetic analysis. (<b>A</b>) Eight NRPS genes of <span class="html-italic">T. harzianum</span> were used as query, three of which had homologous sequence with <span class="html-italic">T. asperellum</span>. (<b>B</b>) Eight PKS genes of <span class="html-italic">T. harzianum</span> were used as query, five of which had homologous sequence with <span class="html-italic">T. gracile</span>. (<b>C</b>) Five TC genes of <span class="html-italic">T. harzianum</span> were used as query, none of which had sequences similarity with other organisms on NCBI database. A darker shade of blue denotes a higher percentage identity of the query in the output cluster, while the number within each box, resembles the counts of hits for a specific query sequence in the co-localized region. Orange and red borders indicate that similar genes found in multiple clusters.</p>
Full article ">
11 pages, 2904 KiB  
Article
Changes in Ion Transport across Biological Membranes Exposed to Particulate Matter
by Jakub Hoser, Adrianna Dabrowska, Miroslaw Zajac and Piotr Bednarczyk
Membranes 2023, 13(9), 763; https://doi.org/10.3390/membranes13090763 - 29 Aug 2023
Viewed by 971
Abstract
The cells of living organisms are surrounded by the biological membranes that form a barrier between the internal and external environment of the cells. Cell membranes serve as barriers and gatekeepers. They protect cells against the entry of undesirable substances and are the [...] Read more.
The cells of living organisms are surrounded by the biological membranes that form a barrier between the internal and external environment of the cells. Cell membranes serve as barriers and gatekeepers. They protect cells against the entry of undesirable substances and are the first line of interaction with foreign particles. Therefore, it is very important to understand how substances such as particulate matter (PM) interact with cell membranes. To investigate the effect of PM on the electrical properties of biological membranes, a series of experiments using a black lipid membrane (BLM) technique were performed. L-α-Phosphatidylcholine from soybean (azolectin) was used to create lipid bilayers. PM samples of different diameters (<4 (SRM-PM4.0) and <10 μm (SRM-PM10) were purchased from The National Institute of Standards and Technology (USA) to ensure the repeatability of the measurements. Lipid membranes with incorporated gramicidin A (5 pg/mL) ion channels were used to investigate the effect of PM on ion transport. The ionic current passing through the azolectin membranes was measured in ionic gradients (50/150 mM KCl on cis/trans side). In parallel, the electric membrane capacitance measurements, analysis of the conductance and reversal potential were performed. Our results have shown that PM at concentration range from 10 to 150 μg/mL reduced the basal ionic current at negative potentials while increased it at positive ones, indicating the interaction between lipids forming the membrane and PM. Additionally, PM decreased the gramicidin A channel activity. At the same time, the amplitude of channel openings as well as single channel conductance and reversal potential remained unchanged. Lastly, particulate matter at a concentration of 150 μg/mL did not affect the electric membrane capacity to any significant extent. Understanding the interaction between PM and biological membranes could aid in the search for effective cytoprotective strategies. Perhaps, by the use of an artificial system, we will learn to support the consequences of PM-induced damage. Full article
(This article belongs to the Special Issue Advances in Artificial and Biological Membranes, Volume II)
Show Figures

Figure 1

Figure 1
<p>Human bronchial epithelial cells were incubated in cell culture medium alone and in the presence of PM4.0 and PM10 for 24 h, 48 h and 72 h, respectively. Fragments of PM particles (&lt;4 μm and PM &lt; 10 μm) deposited on HBE cells are clearly visible. Blue arrows indicate healthy cell examples and black arrows indicate chosen particulate matter (PM). The pictures were taken with the DLTX1080PCMOSHDU2SD camera (DELTA optical) placed in inverted optical microscope (Olympus).</p>
Full article ">Figure 2
<p>Black lipid membrane technique. (<b>a</b>) Scheme of the system used in black lipid membrane (BLM) experiments including antivibration table, Faraday’s cage, chamber with <span class="html-italic">cis/trans</span> side, amplifier, converter and PC. Buffers were at pH = 7.2 and <span class="html-italic">trans</span> side was grounded. (<b>b</b>) Representative recording in 50/150 mM KCl (<span class="html-italic">cis/trans</span>) gradient before and after incorporation of gramicidin A (arrow) at 0 mV—indicates the closed state. (<b>c</b>) Multi-channel recording in 50/150 mM KCl (<span class="html-italic">cis/trans</span>) gradient after incorporation of gA. The graph shows the method of ionic current area (pA*s) calculation over 20 s through lipid membrane with incorporated gramicidin A ion channel.</p>
Full article ">Figure 3
<p>Effects of different PM concentrations on lipid membranes. (<b>a</b>) Registered current–time signals of ionic current flow through lipid bilayer membranes in control (0 μg/mL) and in the presence of 50, 100, 150 μg/mL PM4.0 and PM10 at a potential of +40 mV. (<b>b</b>,<b>c</b>) Analysis of the ionic current flow through lipid membrane at +40 and −40 mV without (control) and in the presence of PM4.0 and PM10 at concentrations of 10, 30, 50, 70, 100 and 150 μg/mL. The results are presented as mean ± SD, <span class="html-italic">n</span> = 5. Statistical significance was determined at <span class="html-italic">p</span> &lt; 0.001 (***) using one-way ANOVA.</p>
Full article ">Figure 4
<p>Electrical membrane capacitance changes in the presence of PM. Time-resolved effect of PM4.0 and PM10 (150 μg/mL) on electrical membrane capacitance changes. Results are presented as mean ± SD, <span class="html-italic">n</span> = 3. Statistical significance was determined at <span class="html-italic">p</span> &lt; 0.05 (*) using one-way ANOVA.</p>
Full article ">Figure 5
<p>Effects of PM on gramicidin A channels. (<b>a</b>) Registered signals of ionic current flow through azolectin bilayer membrane with incorporated gramicidin A (5 pg/mL) channels at −120 mV with and without presence 150 μg/mL of PM. (<b>b</b>,<b>c</b>) Effect of particulate matter on ionic current area through azolectin bilayer membrane with incorporated gramicidin A (5 pg/mL) channels with and without the presence of 150 μg/mL PM4.0 (<b>b</b>) and PM10 (<b>c</b>). (<b>d</b>,<b>e</b>) Effect of the particulate matter on ionic current amplitude of single gramicidin A (5 pg/mL) channel in the presence of PM4.0 (<b>d</b>) and PM10 (<b>e</b>) at a concentration of 150 μg/mL.</p>
Full article ">
15 pages, 3289 KiB  
Article
Experimental Design Approach for Development of HPLC Method for Simultaneous Analysis of Triamcinolone, Nystatin, and Gramicidin in Industrial Wastewater
by Loubna Elsharkawy, Maha A. Hegazy, Ahmed E. Elgendy and Rasha M. Ahmed
Separations 2023, 10(6), 342; https://doi.org/10.3390/separations10060342 - 1 Jun 2023
Cited by 1 | Viewed by 1975
Abstract
This study used an experimental design approach to optimize an HPLC method for the simultaneous determination of three pharmaceutical residues (triamcinolone, nystatin, and gramicidin) in industrial wastewater samples. The goal of using an experimental design approach was to maximize the method performance through [...] Read more.
This study used an experimental design approach to optimize an HPLC method for the simultaneous determination of three pharmaceutical residues (triamcinolone, nystatin, and gramicidin) in industrial wastewater samples. The goal of using an experimental design approach was to maximize the method performance through separation enhancement and shortening the time of analysis and/or minimizing the environmental effects through the reduction in wastes and sample treatment. To achieve this goal, two steps were performed: a full factorial screening design for the three chromatographic variables, and optimization design using central composite design to select the optimum conditions that accomplished the highest resolution between adjacent peaks within a minimum run time of less than 5 min. The optimal chromatographic conditions derived from Minitab software using the desirability function were applied. Separation was carried out on a Zorbax C18 column (250 mm × 4.6, 5 μm) with gradient elution of a mobile phase composed of methanol and 0.25 M potassium dihydrogen phosphate buffer (pH 3.6) at different UV detections. For the validation of the developed HPLC method, ICH guidelines were followed, and the obtained results were found to be in compliance with the acceptance criteria. Linearity was over the concentration range of 1.00–25.00 μg/mL for triamcinilone and nystatin and 10.00–50.00 µg/mL for gramicidin. The proposed method was successfully applied to quantify the three studied pharmaceutical compounds in rinsing wastewater samples. Full article
(This article belongs to the Section Analysis of Natural Products and Pharmaceuticals)
Show Figures

Figure 1

Figure 1
<p>Ishikawa diagram for identification of critical process variables.</p>
Full article ">Figure 2
<p>Main effect of the variables and their effect on response. (<b>a</b>) Main effect of the variables and their effect on R<sub>S</sub>2. (<b>b</b>) Main effect of the variables and their effect R<sub>S</sub>1. (<b>c</b>) Main effect of the variables and their effect on run time.</p>
Full article ">Figure 3
<p>Interaction plot between the variables and their effect on response. (<b>a</b>) Interaction plot between the variables and their effect on run time. (<b>b</b>) Interaction plot between the variables and their effect on R<sub>S</sub>1. (<b>c</b>) Interaction plot between the variables and their effect on R<sub>S</sub>2.</p>
Full article ">Figure 4
<p>Desirability plot displaying the optimum conditions of the method.</p>
Full article ">Figure 5
<p>HPLC chromatogram of triamcinolone (10.00 μg/mL), nystatin (10.00 μg/mL), and gramicidin (30.00 μg/mL) using gradient elution.</p>
Full article ">Figure 6
<p>Pareto charts of the variables and their effects on the response. (<b>a</b>) Effect of pH, organic ratio, and flow rate on resolution 1 (R<sub>S</sub>1). (<b>b</b>) Effect of pH, organic ratio, and flow rate on resolution 2 (R<sub>S</sub>2). (<b>c</b>) Effect of pH, organic ratio, and flow rate on run time.</p>
Full article ">Figure 7
<p>Chromatogram of HPLC determination of the 49.48 µg/mL triamcinolone, 0.62 µg/mL nystatin, and 4.41 µg/mL gramicidin in real environmental samples.</p>
Full article ">
15 pages, 5214 KiB  
Article
Transient Coatings from Nanoparticles Achieving Broad-Spectrum and High Antimicrobial Performance
by Rachel Zaia, Giovanna M. Quinto, Livia C. S. Camargo, Rodrigo T. Ribeiro and Ana M. Carmona-Ribeiro
Pharmaceuticals 2023, 16(6), 816; https://doi.org/10.3390/ph16060816 - 30 May 2023
Cited by 1 | Viewed by 1286
Abstract
Cationic and hydrophilic coatings based on casting and drying water dispersions of two different nanoparticles (NPs) onto glass are here described and evaluated for antimicrobial activity. Discoid cationic bilayer fragments (BF) surrounded by carboxy-methylcellulose (CMC) and poly (diallyl dimethyl ammonium) chloride (PDDA) NPs [...] Read more.
Cationic and hydrophilic coatings based on casting and drying water dispersions of two different nanoparticles (NPs) onto glass are here described and evaluated for antimicrobial activity. Discoid cationic bilayer fragments (BF) surrounded by carboxy-methylcellulose (CMC) and poly (diallyl dimethyl ammonium) chloride (PDDA) NPs and spherical gramicidin D (Gr) NPs dispersed in water solution were cast onto glass coverslips and dried, forming a coating quantitatively evaluated against Pseudomonas aeruginosa, Staphylococcus aureus and Candida albicans. From plating and colony forming units (CFU) counting, all strains interacting for 1 h with the coatings lost viability from 105 to 106, to zero CFU, at two sets of Gr and PDDA doses: 4.6 and 25 μg, respectively, or, 0.94 and 5 μg, respectively. Combinations produced broad spectrum, antimicrobial coatings; PDDA electrostatically attached to the microbes damaging cell walls, allowing Gr NPs interaction with the cell membrane. This concerted action promoted optimal activity at low Gr and PDDA doses. Further washing and drying of the deposited dried coatings showed that they were washed out so that antimicrobial activity was no longer present on the glass surface. Significant applications in biomedical materials can be foreseen for these transient coatings. Full article
(This article belongs to the Special Issue Self-Assembled Nanoparticles: An Emerging Delivery Platform for Drugs)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the current development of transient antimicrobial coatings on glass based on casting and drying of gramicidin (Gr) nanoparticles (NPs) [<a href="#B30-pharmaceuticals-16-00816" class="html-bibr">30</a>] and self−assembled discoidal NPs [<a href="#B41-pharmaceuticals-16-00816" class="html-bibr">41</a>] both in water dispersions (0.264 M D-glucose isotonic solution). The layer−by−layer discoid assemblies of dioctadecyldimethylammonium bromide bilayer fragments (DODAB BF) were covered by consecutive layers of carboxymethylcellulose (CMC) and poly diallyldimethylammonium chloride (PDDA) yielding the DODAB BF/CMC/PDDA disks in water dispersion [<a href="#B37-pharmaceuticals-16-00816" class="html-bibr">37</a>]. Scanning electron micrographs for the Gr NPs and discoid NPs are available from references [<a href="#B30-pharmaceuticals-16-00816" class="html-bibr">30</a>,<a href="#B41-pharmaceuticals-16-00816" class="html-bibr">41</a>], respectively.</p>
Full article ">Figure 2
<p>Effect of gramicidin concentration in Gr NPs on properties of small particles (SP) (<b>a</b>), or large particles (LP) (<b>b</b>). Properties were the hydrodynamic diameter (Dz), the zeta potential (ζ) and the polydispersity (P). Compositions for SP, LP or Gr NPs, are in <a href="#pharmaceuticals-16-00816-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 3
<p>Three independent experiments showing reproducible coating’s activities against <span class="html-italic">P. aeruginosa</span>. Each coating was obtained from casting and drying 50 µL of Gr NPs, SP, SP/Gr NPs, LP, and LP/Gr NPs dispersions in 0.264 M D-glucose water solution. Coatings yielded the same activity independent of the time they remained on the shelf before use. Statistical analysis revealed that SP, SP/Gr NPs, LP and LP/Gr NPs coatings significantly diminished cell viability (<span class="html-italic">p</span> &lt; 0.05) whereas coatings comprised of Gr NPs did not affect cell viability.</p>
Full article ">Figure 4
<p>Three independent experiments showing microbicidal activity of coatings on <span class="html-italic">S. aureus</span>. The coatings were obtained from casting and drying 50 µL of Gr NPs, SP, SP/Gr NPs, LP, and LP/Gr NPs dispersions in 0.264 M D-glucose water solution onto glass coverslips. Statistical analysis revealed that although each type of NPs on the coatings significantly decreased cell viability (<span class="html-italic">p</span> &lt; 0.05), the combinations of NPs such as SP/Gr NPs and LP/Gr NPs in the coatings brought cell viability to zero CFU count (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Three independent experiments showing antimicrobial activity of coatings on <span class="html-italic">C. albicans</span>. The coatings were obtained from casting and drying 50 µL of Gr NPs, SP, SP/Gr NPs, LP, and LP/Gr NPs dispersions in 0.264 M D—glucose water solution. Statistical analysis revealed that coatings comprised of NPs significantly decreased cell viability (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Three independent experiments showing the loss of antimicrobial activity of the coatings against <span class="html-italic">C. albicans</span> after immersing them in 0.264 M D-glucose aqueous solution for 1 h followed by drying.</p>
Full article ">Figure 7
<p>Cell viability of <span class="html-italic">C. albicans</span> after interacting with non-washed (<a href="#pharmaceuticals-16-00816-f005" class="html-fig">Figure 5</a>) and washed coatings (<a href="#pharmaceuticals-16-00816-f006" class="html-fig">Figure 6</a>). Due to the highly reproducible character of the non-washed and the washed coatings, only one experiment among three is shown.</p>
Full article ">
11 pages, 2520 KiB  
Article
In Silico and In Vitro Inhibition of SARS-CoV-2 PLpro with Gramicidin D
by Sara Protić, Nevena Kaličanin, Milan Sencanski, Olivera Prodanović, Jelena Milicevic, Vladimir Perovic, Slobodan Paessler, Radivoje Prodanović and Sanja Glisic
Int. J. Mol. Sci. 2023, 24(3), 1955; https://doi.org/10.3390/ijms24031955 - 19 Jan 2023
Cited by 3 | Viewed by 1928
Abstract
Finding an effective drug to prevent or treat COVID-19 is of utmost importance in tcurrent pandemic. Since developing a new treatment takes a significant amount of time, drug repurposing can be an effective option for achieving a rapid response. This study used a [...] Read more.
Finding an effective drug to prevent or treat COVID-19 is of utmost importance in tcurrent pandemic. Since developing a new treatment takes a significant amount of time, drug repurposing can be an effective option for achieving a rapid response. This study used a combined in silico virtual screening protocol for candidate SARS-CoV-2 PLpro inhibitors. The Drugbank database was searched first, using the Informational Spectrum Method for Small Molecules, followed by molecular docking. Gramicidin D was selected as a peptide drug, showing the best in silico interaction profile with PLpro. After the expression and purification of PLpro, gramicidin D was screened for protease inhibition in vitro and was found to be active against PLpro. The current study’s findings are significant because it is critical to identify COVID-19 therapies that are efficient, affordable, and have a favorable safety profile. Full article
(This article belongs to the Special Issue Molecular Interactions and Mechanisms of COVID-19 Inhibition 2.0)
Show Figures

Figure 1

Figure 1
<p>Gramicidin D docked into the catalytic site of PL<sup>pro</sup>. Green lines: hydrogen bonds; orange: electrostatic (including Pi-anion); yellow: Pi-sulfur; magenta: Pi-alkyl.</p>
Full article ">Figure 2
<p>Gramicidin S docked into the catalytic site of PL<sup>pro</sup>. Green lines: hydrogen bonds; orange: electrostatic (including Pi-anion); yellow: Pi-sulfur; magenta: Pi-alkyl.</p>
Full article ">Figure 3
<p>The ESP surfaces of gramicidin D (<b>a</b>) and gramicidin S (<b>b</b>). ESP potential is presented from a value of −20 (red) to −10 (blue).</p>
Full article ">Figure 4
<p>Inhibition (%) and activity (%) of PL<sup>pro</sup> with different concentrations of gramicidin D.</p>
Full article ">
16 pages, 7327 KiB  
Article
Effects of Calcium Ions on the Antimicrobial Activity of Gramicidin A
by Shang-Ting Fang, Shu-Hsiang Huang, Chin-Hao Yang, Jen-Wen Liou, Hemalatha Mani and Yi-Cheng Chen
Biomolecules 2022, 12(12), 1799; https://doi.org/10.3390/biom12121799 - 1 Dec 2022
Cited by 4 | Viewed by 1386
Abstract
Gramicidin A (gA) is a linear antimicrobial peptide that can form a channel and specifically conduct monovalent cations such as H+ across the lipid membrane. The antimicrobial activity of gA is associated with the formation of hydroxyl free radicals and the imbalance [...] Read more.
Gramicidin A (gA) is a linear antimicrobial peptide that can form a channel and specifically conduct monovalent cations such as H+ across the lipid membrane. The antimicrobial activity of gA is associated with the formation of hydroxyl free radicals and the imbalance of NADH metabolism, possibly a consequence caused by the conductance of cations. The ion conductivity of gramicidin A can be blocked by Ca2+ ions. However, the effect of Ca2+ ions on the antimicrobial activity of gA is unclear. To unveil the role of Ca2+ ions, we examined the effect of Ca2+ ions on the antimicrobial activity of gramicidin A against Staphylococcus aureus (S. aureus). Results showed that the antimicrobial mechanism of gA and antimicrobial activity by Ca2+ ions are concentration-dependent. At the low gA concentration (≤1 μM), the antimicrobial mechanism of gA is mainly associated with the hydroxyl free radical formation and NADH metabolic imbalance. Under this mode, Ca2+ ions can significantly inhibit the hydroxyl free radical formation and NADH metabolic imbalance. On the other hand, at high gA concentration (≥5 μM), gramicidin A acts more likely as a detergent. Gramicidin A not only causes an increase in hydroxyl free radical levels and NAD+/NADH ratios but also induces the destruction of the lipid membrane composition. At this condition, Ca2+ ions can no longer reduce the gA antimicrobial activity but rather enhance the bacterial killing ability of gramicidin A. Full article
(This article belongs to the Special Issue Functional Peptides and Their Interactions: From Molecules to Systems)
Show Figures

Figure 1

Figure 1
<p>The growth curve of <span class="html-italic">S. aureus</span> in the presence of gramicidin A 0, 1, 5, and 10 μM, and 1% TFE without gramicidin A, at the different growth phases, (<b>A</b>) lag phase, (<b>B</b>) elongation phase, and (<b>C</b>) stationary phase, respectively. (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.001, related to 0 μM gA).</p>
Full article ">Figure 1 Cont.
<p>The growth curve of <span class="html-italic">S. aureus</span> in the presence of gramicidin A 0, 1, 5, and 10 μM, and 1% TFE without gramicidin A, at the different growth phases, (<b>A</b>) lag phase, (<b>B</b>) elongation phase, and (<b>C</b>) stationary phase, respectively. (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.001, related to 0 μM gA).</p>
Full article ">Figure 2
<p>(<b>A</b>) Hydroxyl free radical level for <span class="html-italic">S. aureus</span> under the treatment of 0, 0.1, 1, and 5 μM of gramicidin A, and (<b>B</b>) the NAD<sup>+</sup>/NADH ratio measured for <span class="html-italic">S. aureus</span> under the treatment of 0, 1, and 5 μM of gramicidin A and 0, 30, 60, 90, and 120 min at the elongation state. (<span class="html-italic">n</span> = 9, ** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2 Cont.
<p>(<b>A</b>) Hydroxyl free radical level for <span class="html-italic">S. aureus</span> under the treatment of 0, 0.1, 1, and 5 μM of gramicidin A, and (<b>B</b>) the NAD<sup>+</sup>/NADH ratio measured for <span class="html-italic">S. aureus</span> under the treatment of 0, 1, and 5 μM of gramicidin A and 0, 30, 60, 90, and 120 min at the elongation state. (<span class="html-italic">n</span> = 9, ** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Effects of calcium ions on the growth rate of <span class="html-italic">S. aureus</span> in the presence of (<b>A</b>) 0–500 mM Ca<sup>2+</sup> concentrations only, (<b>B</b>) 0–500 mM of Ca<sup>2+</sup> ions and 1 μM gA treatment, and (<b>C</b>) 0–500 mM Ca<sup>2+</sup> and 5 μM gA treatment, respectively. (<span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> ≤ 0.001, related to 1 μM gA in (<b>B</b>)).</p>
Full article ">Figure 3 Cont.
<p>Effects of calcium ions on the growth rate of <span class="html-italic">S. aureus</span> in the presence of (<b>A</b>) 0–500 mM Ca<sup>2+</sup> concentrations only, (<b>B</b>) 0–500 mM of Ca<sup>2+</sup> ions and 1 μM gA treatment, and (<b>C</b>) 0–500 mM Ca<sup>2+</sup> and 5 μM gA treatment, respectively. (<span class="html-italic">n</span> = 3, ** <span class="html-italic">p</span> ≤ 0.001, related to 1 μM gA in (<b>B</b>)).</p>
Full article ">Figure 4
<p>Hydroxyl radical level in the presence of gA and 10 μM–500 mM of gramicidin A, (<b>A</b>) 1 μM gA/0–10 mM of Ca<sup>2+</sup> ions, (<b>B</b>) 1 μM gA/100–500 mM of Ca<sup>2+</sup> ions, (<b>C</b>) 5 μM gA/0–10 mM of Ca<sup>2+</sup> ions, and (<b>D</b>) 5 μM gA/100–500 mM of Ca<sup>2+</sup> ions, respectively. (<span class="html-italic">n</span> = 9, ** <span class="html-italic">p</span> &lt; 0.001, * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4 Cont.
<p>Hydroxyl radical level in the presence of gA and 10 μM–500 mM of gramicidin A, (<b>A</b>) 1 μM gA/0–10 mM of Ca<sup>2+</sup> ions, (<b>B</b>) 1 μM gA/100–500 mM of Ca<sup>2+</sup> ions, (<b>C</b>) 5 μM gA/0–10 mM of Ca<sup>2+</sup> ions, and (<b>D</b>) 5 μM gA/100–500 mM of Ca<sup>2+</sup> ions, respectively. (<span class="html-italic">n</span> = 9, ** <span class="html-italic">p</span> &lt; 0.001, * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>The NAD<sup>+</sup>/NADH ratios for <span class="html-italic">S. aureus</span> treated with (<b>A</b>) 1 and (<b>B</b>) 5 μM gramicidin A and 10, 100, 200, and 500 mM Ca<sup>2+</sup> ions at 0, 30, 60, 90, and 120 min, respectively. (<span class="html-italic">n</span> = 9, ** <span class="html-italic">p</span> &lt; 0.001, * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Atomic force microscopic morphologies of <span class="html-italic">S. aureus</span> under treatment with gramicidin A and CA<sup>2+</sup> ions, (<b>A</b>) <span class="html-italic">S. aureus</span> only without gA, (<b>B</b>) with 100 mM Ca<sup>2+</sup> ions, (<b>C</b>) with 1 μM gA, (<b>D</b>) with 5 μM gA, (<b>E</b>) with 1 μM gA and 100 mM Ca<sup>2+</sup> ions and (<b>F</b>) with 5 μM gA and 100 mM Ca<sup>2+</sup> ions.</p>
Full article ">
15 pages, 2257 KiB  
Article
Chromone-Containing Allylmorpholines Influence Ion Channels in Lipid Membranes via Dipole Potential and Packing Stress
by Svetlana S. Efimova, Vera A. Martynyuk, Anastasiia A. Zakharova, Natalia M. Yudintceva, Nikita M. Chernov, Igor P. Yakovlev and Olga S. Ostroumova
Int. J. Mol. Sci. 2022, 23(19), 11554; https://doi.org/10.3390/ijms231911554 - 30 Sep 2022
Cited by 2 | Viewed by 1439
Abstract
Herein, we report that chromone-containing allylmorpholines can affect ion channels formed by pore-forming antibiotics in model lipid membranes, which correlates with their ability to influence membrane boundary potential and lipid-packing stress. At 100 µg/mL, allylmorpholines 1, 6, 7, and 8 [...] Read more.
Herein, we report that chromone-containing allylmorpholines can affect ion channels formed by pore-forming antibiotics in model lipid membranes, which correlates with their ability to influence membrane boundary potential and lipid-packing stress. At 100 µg/mL, allylmorpholines 1, 6, 7, and 8 decrease the boundary potential of the bilayers composed of palmitoyloleoylphosphocholine (POPC) by about 100 mV. At the same time, the compounds do not affect the zeta-potential of POPC liposomes, but reduce the membrane dipole potential by 80–120 mV. The allylmorpholine-induced drop in the dipole potential produce 10–30% enhancement in the conductance of gramicidin A channels. Chromone-containing allylmorpholines also affect the thermotropic behavior of dipalmytoylphosphocholine (DPPC), abolishing the pretransition, lowering melting cooperativity, and turning the main phase transition peak into a multicomponent profile. Compounds 4, 6, 7, and 8 are able to decrease DPPC’s melting temperature by about 0.5–1.9 °C. Moreover, derivative 7 is shown to increase the temperature of transition of palmitoyloleoylphosphoethanolamine from lamellar to inverted hexagonal phase. The effects on lipid-phase transitions are attributed to the changes in the spontaneous curvature stress. Alterations in lipid packing induced by allylmorpholines are believed to potentiate the pore-forming ability of amphotericin B and gramicidin A by several times. Full article
(This article belongs to the Special Issue Ion and Molecule Transport in Membrane Systems 4.0)
Show Figures

Figure 1

Figure 1
<p>The relationships between the structures of allylmorpholines and their potential-modifying abilities: the dependence of dipole modifying efficiency of allylmorpholines on (<b>A</b>) the length of hydrophobic chain in <span class="html-italic">R<sub>5</sub></span> position; (<b>B</b>) lengths of hydrocarbon radicals in <span class="html-italic">R<sub>4</sub></span> and <span class="html-italic">R<sub>5</sub></span> positions; (<b>C</b>) the type of halogen in the <span class="html-italic">R<sub>1</sub></span> position; (<b>D</b>) the existence of halogen in <span class="html-italic">R<sub>1</sub></span> and <span class="html-italic">R<sub>3</sub></span> positions.</p>
Full article ">Figure 2
<p>(<b>A</b>) Current fluctuations corresponding to opening and closing single gramicidin A channels in the absence (<span class="html-italic">control</span>) and presence of <b>5</b> and <b>7</b> at 100 μg/mL. <span class="html-italic">V</span> = 150 mV. <span class="html-italic">C</span>—closed state of the channel, <span class="html-italic">O</span>—open state of the channel. (<b>B</b>) <span class="html-italic">g<sub>sc</sub></span>(<span class="html-italic">V</span>) curves of single gramicidin A channels in the absence and presence of chromone-containing allylmorpholines at 100 μg/mL. The relationship between the color of the symbol and the compound is given in the figure legend. (<b>C</b>) The effects of <b>5</b> and <b>7</b> on pore-forming activity of gramicidin A. A peptide was added into the bathing solution at both sides of the bilayers up to 1 nM. The moments when up to 100 μg/mL of <b>5</b> (<span class="html-italic">upper panel</span>) and <b>7</b> (<span class="html-italic">lower panel</span>) was added in the bilayer bathing solution are indicated by arrows. The membranes were composed of POPC and bathed in 2.0 M KCl (pH 7.4). <span class="html-italic">V</span> = 150 mV.</p>
Full article ">Figure 3
<p>The schematic representations of the mechanisms of the action of allylmorpholine <b>7</b> on the conductance of GrA channels via alterations in membrane dipole potential.</p>
Full article ">Figure 4
<p>Heating thermograms of DPPC liposomes in the absence and presence of various chromone-containing allylmorpholine derivatives at 100 μg/mL. (<b>A</b>) control, <b>1</b>, <b>2</b>, <b>3</b>, <b>4</b>, <b>5</b>, and <b>6</b>; (<b>B</b>) <b>7</b>, <b>8</b>, <b>9</b>, <b>10</b>, <b>11</b>, and <b>12</b>. The relationship between the profile and the compound is given in the figure.</p>
Full article ">Figure 5
<p>The effects of chromone-containing allylmorpholines (<b>4</b>, <b>6</b>, <b>7</b>, <b>8</b>) on the steady-state transmembrane current flowing through membranes modified by one-sided addition of AmB. The moments when 100 μg/mL of <b>4</b> (<b>A</b>), <b>6</b> (<b>B</b>), <b>7</b> (<b>C</b>), and <b>8</b> (<b>D</b>) was added to the bilayer bathing solution are indicated by arrows. The lipid bilayers were composed of POPC/CHOL (80/20 mol%) and were bathed in 2.0 M 4KCl, pH 7.4. <span class="html-italic">V</span> = 50 mV.</p>
Full article ">Figure 6
<p>The schematic representations of the mechanisms of action of allylmorpholine <b>7</b> on the pore-forming activity of AmB via alterations in lipid-packing stress.</p>
Full article ">
18 pages, 4166 KiB  
Article
Characterization and Differential Cytotoxicity of Gramicidin Nanoparticles Combined with Cationic Polymer or Lipid Bilayer
by Yunys Pérez-Betancourt, Rachel Zaia, Marina Franchi Evangelista, Rodrigo Tadeu Ribeiro, Bruno Murillo Roncoleta, Beatriz Ideriha Mathiazzi and Ana Maria Carmona-Ribeiro
Pharmaceutics 2022, 14(10), 2053; https://doi.org/10.3390/pharmaceutics14102053 - 27 Sep 2022
Cited by 2 | Viewed by 1826
Abstract
Gramicidin (Gr) nanoparticles (NPs) and poly (diallyl dimethyl ammonium) chloride (PDDA) water dispersions were characterized and evaluated against Gram-positive and Gram-negative bacteria and fungus. Dynamic light scattering for sizing, zeta potential analysis, polydispersity, and colloidal stability over time characterized Gr NPs/PDDA dispersions, and [...] Read more.
Gramicidin (Gr) nanoparticles (NPs) and poly (diallyl dimethyl ammonium) chloride (PDDA) water dispersions were characterized and evaluated against Gram-positive and Gram-negative bacteria and fungus. Dynamic light scattering for sizing, zeta potential analysis, polydispersity, and colloidal stability over time characterized Gr NPs/PDDA dispersions, and plating and colony-forming units counting determined their microbicidal activity. Cell viabilities of Staphylococcus aureus, Pseudomonas aeruginosa, and Candida albicans in the presence of the combinations were reduced by 6, 7, and 7 logs, respectively, at 10 μM Gr/10 μg·mL−1 PDDA, 0.5 μM Gr/0. 5μg·mL−1 PDDA, and 0.5 μM Gr/0.5 μg·mL−1 PDDA, respectively. In comparison to individual Gr doses, the combinations reduced doses by half (S. aureus) and a quarter (C. albicans); in comparison to individual PDDA doses, the combinations reduced doses by 6 times (P. aeruginosa) and 10 times (C. albicans). Gr in supported or free cationic lipid bilayers reduced Gr activity against S. aureus due to reduced Gr access to the pathogen. Facile Gr NPs/PDDA disassembly favored access of each agent to the pathogen: PDDA suctioned the pathogen cell wall facilitating Gr insertion in the pathogen cell membrane. Gr NPs/PDDA differential cytotoxicity suggested the possibility of novel systemic uses for the combination. Full article
(This article belongs to the Section Drug Delivery and Controlled Release)
Show Figures

Figure 1

Figure 1
<p>Scanning electron micrographs of gramicidin (Gr) dispersions at 0.05 mM Gr in pure water (<b>a</b>) or in 0.05 mg/mL poly (diallyldimethyl ammonium chloride) (PDDA) solution (<b>b</b>).</p>
Full article ">Figure 2
<p>Effect of gramicidin (Gr) concentration on the physical properties of Gr dispersions in water and 1% trifluoroethanol. Measurements were performed 30 min after preparation of the Gr dispersions. Physical properties were the macroscopic aspect, the mean z-average diameter (Dz), the zeta potential (ζ), and the polydispersity (P). All measurements represent a mean value ± the mean standard deviation.</p>
Full article ">Figure 3
<p>Effect of poly (diallyldimethyl ammonium chloride) (PDDA) concentration on the macroscopic aspect and physical properties of gramicidin (Gr)/PDDA dispersions at 0.05 mM Gr. Physical properties were the mean z-average diameter (Dz), the zeta potential (ζ), the polydispersity (P), and the conductance (G). All measurements represent a mean value ± the mean standard deviation.</p>
Full article ">Figure 4
<p>Colloidal stability of gramicidin (Gr) nanoparticles (NPs) in the absence or presence of poly (diallyldimethyl ammonium chloride) (PDDA). Mean z-average diameter (Dz), zeta potential (ζ), polydispersity (P), and turbidity at 400 nm (Abs 400 nm) of gramicidin (Gr) nanoparticles in water or in 0.05 mg/mL PDDA as a function of time.</p>
Full article ">Figure 5
<p>Circular dichroism spectra of 0.02 mM gramicidin (Gr) at 25 °C in different media. Aliquots (0.02 mL) of Gr stock solutions (6.4 mM) in trifluoroethanol (TFE) or ethanol were added to 2.0 mL of water, TFE, or ethanol. One should notice that in water or in aqueous PDDA solutions, Gr spherical NPs yielded the same spectrum. Alternatively, aliquots of poly (diallyldimethyl ammonium chloride) (PDDA) stock solution (10 mg/mL) were added to Gr dispersion under stirring by vortexing to yield 0.01, 0.02, and 0.05 mg/mL PDDA.</p>
Full article ">Figure 6
<p>Conductance of a 0.05 mM gramicidin (Gr) nanoparticles (NPs) dispersion in the absence (∆) or presence of poly (diallyldimethyl ammonium chloride) (PDDA) (□). PDDA conductance in absence of Gr was also measured over a range of [PDDA] (o).</p>
Full article ">Figure 7
<p>Cell viability of <span class="html-italic">Staphylococcus aureus</span> (10<sup>7</sup>–10<sup>8</sup> CFU/mL) after interacting for 1 h with gramicidin (Gr) nanoparticles or other Gr formulations in dioctadecyldimethylammonium bromide (DODAB) bilayers. Cell viability in the presence of Gr in DODAB supported bilayers on silica (SiO<sub>2</sub>/DODAB/Gr) [<a href="#B21-pharmaceutics-14-02053" class="html-bibr">21</a>], Gr in DODAB supported bilayers on polystyrene sulfate (PSS) nanoparticles (PSS/DODAB/Gr) [<a href="#B40-pharmaceutics-14-02053" class="html-bibr">40</a>], and DODAB bilayer fragments (DODAB BF [<a href="#B41-pharmaceutics-14-02053" class="html-bibr">41</a>] or DODAB BF/Gr [<a href="#B19-pharmaceutics-14-02053" class="html-bibr">19</a>]), all of them incorporating Gr. In the DODAB BF dispersions, Gr dimers in the channel conformation have been previously described [<a href="#B19-pharmaceutics-14-02053" class="html-bibr">19</a>,<a href="#B40-pharmaceutics-14-02053" class="html-bibr">40</a>]. The SiO<sub>2</sub>/DODAB/Gr stock dispersion was prepared at 2 mg/mL silica, 0.5 mM DODAB, and 0.05 mM gramicidin, yielding Dz = 280 ± 5 nm, P = 0.20 ± 0.02, and ζ = 45 ± 4.</p>
Full article ">Figure 8
<p>Cell viability of <span class="html-italic">Staphylococcus aureus</span> (10<sup>6</sup>–10<sup>8</sup> CFU/mL) after interacting for 1 h with gramicidin nanoparticles (Gr NPs), poly (diallyldimethyl ammonium chloride) PDDA/Gr NPs, or PDDA solutions over a range of Gr and/or PDDA concentrations. Data for cell viability over a range of [PDDA] were taken from [<a href="#B22-pharmaceutics-14-02053" class="html-bibr">22</a>].</p>
Full article ">Figure 9
<p>Cell viability of <span class="html-italic">C. albicans</span> (10<sup>6</sup> CFU/mL) after interacting for 1 h with gramicidin (Gr) nanoparticles (NPs), poly (diallyl dimethyl ammonium chloride) (PDDA) solutions, or Gr NPs/PDDA dispersions over a range of Gr and/or PDDA concentrations. Data for cell viability in the presence of PDDA only, over a range of PDDA concentrations, were reproduced from [<a href="#B22-pharmaceutics-14-02053" class="html-bibr">22</a>].</p>
Full article ">Figure 10
<p>Cell viability of <span class="html-italic">Pseudomonas aeruginosa</span> PA14 (10<sup>7</sup> CFU/mL) after interacting for 1 h with poly (diallyl dimethyl ammonium chloride) (PDDA) solutions (blue inverted triangles) or gramicidin nanoparticles Gr NPs/PDDA (black hollow circles) over a range of Gr and/or PDDA concentrations.</p>
Full article ">
21 pages, 3266 KiB  
Article
Antiplasmodial Cyclodecapeptides from Tyrothricin Share a Target with Chloroquine
by Adrienne N.-N. Leussa and Marina Rautenbach
Antibiotics 2022, 11(6), 801; https://doi.org/10.3390/antibiotics11060801 - 14 Jun 2022
Cited by 3 | Viewed by 2398
Abstract
Previous research found that the six major cyclodecapeptides from the tyrothricin complex, produced by Brevibacillus parabrevis, showed potent activity against chloroquine sensitive (CQS) Plasmodium falciparum. The identity of the aromatic residues in the aromatic dipeptide unit in cyclo-(D-Phe1-Pro2 [...] Read more.
Previous research found that the six major cyclodecapeptides from the tyrothricin complex, produced by Brevibacillus parabrevis, showed potent activity against chloroquine sensitive (CQS) Plasmodium falciparum. The identity of the aromatic residues in the aromatic dipeptide unit in cyclo-(D-Phe1-Pro2-(Phe3/Trp3)-D-Phe4/D-Trp4)-Asn5-Gln6-(Tyr7/Phe7/Trp7)-Val8-(Orn9/Lys9)-Leu10 was proposed to have an important role in activity. CQS and resistant (CQR) P. falciparum strains were challenged with three representative cyclodecapeptides. Our results confirmed that cyclodecapeptides from tyrothricin had significantly higher antiplasmodial activity than the analogous gramicidin S, rivaling that of CQ. However, the previously hypothesized size and hydrophobicity dependent activity for these peptides did not hold true for P. falciparum strains, other than for the CQS 3D7 strain. The Tyr7 in tyrocidine A (TrcA) with Phe3-D-Phe4 seem to be related with loss in activity correlating with CQ antagonism and resistance, indicating a shared target and/or resistance mechanism in which the phenolic groups play a role. Phe7 in phenycidine A, the second peptide containing Phe3-D-Phe4, also showed CQ antagonism. Conversely, Trp7 in tryptocidine C (TpcC) with Trp3-D-Trp4 showed improved peptide selectivity and activity towards the more resistant strains, without overt antagonism towards CQ. However, TpcC lead to similar parasite stage inhibition and parasite morphology changes than previously observed for TrcA. The disorganization of chromatin packing and neutral lipid structures, combined with amorphous hemozoin crystals, could account for halted growth in late trophozoite/early schizont stage and the nanomolar non-lytic activity of these peptides. These targets related to CQ antagonism, changes in neural lipid distribution, leading to hemozoin malformation, indicate that the tyrothricin cyclodecapeptides and CQ share a target in the malaria parasite. The differing activities of these cyclic peptides towards CQS and CQR P. falciparum strains could be due to variable target interaction in multiple modes of activity. This indicated that the cyclodecapeptide activity and parasite resistance response depended on the aromatic residues in positions 3, 4 and 7. This new insight on these natural cyclic decapeptides could also benefit the design of unique small peptidomimetics in which activity and resistance can be modulated. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The chemical structures of gramicidin S (GS), tyrocidine A (TrcA), phenycidine A (PhcA), tryptocidine C (TpcC), and chloroquine (CQ).</p>
Full article ">Figure 2
<p>Distribution of trophozoite, schizont and ring parasite stages over time from 0 to 48 h following treatment of synchronized <span class="html-italic">P. falciparum</span> D10 cultures at trophozoite stage incubated without (control) (<b>A</b>), with 200 nM of TpcC (<b>B</b>) or 200 ng/mL Trc mixture (<b>C</b>). Each data point represents the mean ± SEM of parasite counts made within 8–13 regions on the microscope slide from two cultures of each treatment.</p>
Full article ">Figure 3
<p>Super-resolution structured illumination fluorescence microscopy images of late intra-erythrocytic stages of <span class="html-italic">P. falciparum</span> D10 stained with the membrane impermeable fluorescent dye trypan blue (red) and the permeable nucleic acid fluorescent dye SYTO9 (green). Images (<b>A</b>–<b>D</b>) show untreated normal trophozoites and (<b>D</b>) a normal schizont. Images (<b>E</b>–<b>I</b>) show TpcC-treated trophozoites (post 6 h treatment) and (<b>J</b>–<b>L</b>) show TpcC-treated schizonts (post 6 h treatment). The arrows in (<b>I</b>,<b>H</b>) highlights the of dark irregular elongated structures, which could indicate hemozoin crystals, within or close to the nuclear material mass. Each image shows a single erythrocyte, with an average diameter of 7.5 μm. Refer to <a href="#antibiotics-11-00801-f004" class="html-fig">Figure 4</a> more detailed magnification scales.</p>
Full article ">Figure 4
<p>Super-resolution structured illumination fluorescence microscopy images of late intra-erythrocytic stages of <span class="html-italic">P. falciparum</span> D10 stained with trypan blue (red) and neutral lipid binding fluorescent dye LipidTOX (green). The large panels are compiled images from super position of both dye images, with four selected enlarged infected cell images are shown on the right. Images (<b>A</b>–<b>E</b>) show the control cultures (no peptide added) and images (<b>F</b>–<b>J</b>) show trophozoite stage cultures treated with TpcC after 5 h in culture.</p>
Full article ">
12 pages, 1889 KiB  
Article
Loss of Gramicidin Biosynthesis in Gram-Positive Biocontrol Bacterium Aneurinibacillus migulanus (Takagi et al., 1993) Shida et al. 1996 Emend Heyndrickx et al., 1997 Nagano Impairs Its Biological Control Ability of Phytophthora
by Faizah N. Alenezi, Ali Chenari Bouket, Hafsa Cherif-Silini, Allaoua Silini, Marcel Jaspars, Tomasz Oszako and Lassaȃd Belbahri
Forests 2022, 13(4), 535; https://doi.org/10.3390/f13040535 - 30 Mar 2022
Cited by 2 | Viewed by 1790
Abstract
The soil-borne species Aneurinibacillus migulanus (A. migulanus) strains Nagano and NCTC 7096 were shown to be potent biocontrol agents active against several plant diseases in agricultural and forest ecosystems. Both strains produce the cyclic peptide gramicidin S (GS) that was described as the [...] Read more.
The soil-borne species Aneurinibacillus migulanus (A. migulanus) strains Nagano and NCTC 7096 were shown to be potent biocontrol agents active against several plant diseases in agricultural and forest ecosystems. Both strains produce the cyclic peptide gramicidin S (GS) that was described as the main weapon inhibiting some gram-negative and gram-positive bacteria and fungus-like organisms along with the production of biosurfactant and hemolysis activities. However, the contribution of the cyclic peptide gramicidin S (GS) to the biocontrol ability of A. migulanus has never been studied experimentally. In this paper, using a mutant of the A. migulanus Nagano strain (E1 mutant) impaired in GS biosynthesis we evaluated the contribution of GS in the biocontrol potential of A. migulanus against Phytophthora spp. The two strains of A. migulanus, Nagano and NCTC 7096, were tested in a pilot study for the inhibition of the growth of 13 Phytophthora species in dual culture assays. A. migulanus Nagano was significantly more inhibitory than NCTC 7096 to all species. Additionally, using apple infection assays, P. rosacearum MKDF-148 and P. cryptogea E2 were shown to be the most aggressive on apple fruits displaying clear infection halos. Therefore, the three A. migulanus strains, Nagano, NCTC 7096, and E1, were used in apple infection experiments to check their effect on infection ability of these two Phytophthora species. Treatment with A. migulanus Nagano significantly reduced the severity of symptoms in apple fruits compared with NCTC 7096. A. migulanus E1 mutant showed total loss of biocontrol ability suggesting that GS is a major actor in the biocontrol ability of A. migulanus Nagano strain. Full article
(This article belongs to the Special Issue Biological Control in Forests Protection)
Show Figures

Figure 1

Figure 1
<p>The comparison of growth rates among the three strains of <span class="html-italic">A. migulanus.</span> (<b>A</b>) Optical density (OD<sub>600nm</sub>) measurements for (●) <span class="html-italic">A. migulanus</span> Nagano, <span style="color:#2D18A8">(</span><span style="color:#2D18A8">■</span><span style="color:#2D18A8">)</span> NCTC 7096, and (<span style="color:red">♦</span>) E1. (<b>B</b>) Colony forming units (CFU) per mL for (●) <span class="html-italic">A. migulanus</span> Nagano, <span style="color:#2D18A8">(</span><span style="color:#2D18A8">■</span><span style="color:#2D18A8">)</span> NCTC 7096, and (<span style="color:red">♦</span>) E1. Vertical bars represent standard errors of the means of three independent replicates. (<b>C</b>) <span class="html-italic">Aneurinibacillus migulanus</span> biosurfactant effects on surface tension of culture fluids and time course of changes in surface tension of (●) <span class="html-italic">A. migulanus</span> Nagano, <span style="color:#2D18A8">(</span><span style="color:#2D18A8">■</span><span style="color:#2D18A8">)</span> NCTC 7096, and (<span style="color:red">♦</span>) E1 in culture fluids. Vertical bars represent standard errors of the means (N = 3).</p>
Full article ">Figure 2
<p>(<b>A</b>) In vivo pathogenicity assay of <span class="html-italic">Phytophthora</span> species on apple fruits; (A) <span class="html-italic">P. cryptogea</span> E2, (B) <span class="html-italic">P. rosacearum</span> MKDF-148, (C) <span class="html-italic">P. quercina</span>, (D) <span class="html-italic">P. plurivora</span> MKDF-179, (E) <span class="html-italic">P. cinnamomi</span> SCRP 127, (F) <span class="html-italic">P. cambivora</span> P-075, (G) <span class="html-italic">P. taxon Pgchlamydo</span> P-126, (H) <span class="html-italic">P. psychrophila</span>, (I) <span class="html-italic">P. megasperma</span> MKDF-7, (J) <span class="html-italic">P. citrophthora</span> P-139, (K) <span class="html-italic">P. cactorum</span> P-138, (L) <span class="html-italic">P. gonapodyides,</span> (M) <span class="html-italic">P. ramorum</span> P-018 (7 days after inoculation at 20 °C). (<b>B</b>) Inhibition of <span class="html-italic">Phytophthora</span> species by <span class="html-italic">A. migulanus</span> in screening assay. Vertical bars represent standard errors of the means (N = 5).</p>
Full article ">Figure 3
<p>(<b>A</b>) Granny Smith apples inoculated with <span class="html-italic">P. cryptogea</span> (left side) and <span class="html-italic">P. rosacearum</span> (right side); (a) inoculated with <span class="html-italic">P. cryptogea</span> or <span class="html-italic">P. rosacearum</span> (positive control), (b) pre-treated with <span class="html-italic">A. migulanus</span> E1, (c) pre-treated with <span class="html-italic">A. migulanus</span> NCTC 7096, (d) pre-treated with <span class="html-italic">A. migulanus Nagano</span> (6 days after inoculation). (<b>B</b>) Effects of <span class="html-italic">A. migulanus Nagano</span>, NCTC 7096, and mutant E1 on lesion sizes in apples by <span class="html-italic">Phytophthora cryptogea</span> E2 and <span class="html-italic">P. rosacearum</span> MKDF-148 (6 days after inoculation). (<b>C</b>) Effects of <span class="html-italic">A. migulanus Nagano</span>, NCTC 7096, and mutant E1 on the amount of rot caused <span class="html-italic">by Phytophthora cryptogea</span> E2 and <span class="html-italic">P. rosacearum</span> MKDF-148 (6 days after inoculation). Bars labeled with different letters represent significant differences among the treatments at <span class="html-italic">p</span> &lt; 0.05 using the Tukey’s HSD test. In each bar groups, bars labeled with the same letter are not significantly different from each other according to Tukey’s HSD at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
Back to TopTop