[go: up one dir, main page]

 
 
ijms-logo

Journal Browser

Journal Browser

Membrane Channels: Mechanistic Insights

A special issue of International Journal of Molecular Sciences (ISSN 1422-0067). This special issue belongs to the section "Molecular Biophysics".

Deadline for manuscript submissions: closed (30 December 2023) | Viewed by 11053

Special Issue Editors


E-Mail Website
Guest Editor
Department of Biology, University of Maryland, College Park, MD 20842, USA
Interests: biophysics of membrane channels and voltage-gated channels in bacteria; ceramide channels; VDAC channels; the function of the mitochondrial outer membrane

E-Mail Website
Co-Guest Editor
Program in Physical Biology, Eunice Kennedy Shriver National Institute of Child Health and Human Development, National Institutes of Health, Bethesda, MD 20892, USA
Interests: membrane channels

Special Issue Information

Dear Colleagues,

Membrane channels are amazing biological machines that work in a very complex environment that includes: at least 2 phases of matter, surface potentials, dipole potentials, strong electric fields, lateral pressures/tension, unphysiological values of pH and ionic strength, etc. Many of these factors influence or control the states of the channels and thus the flux of matter through the pores. The mechanisms by which membrane channels are regulated vary widely perhaps far more than other biological machines. This special issue will focus on research that proves insight into the molecular mechanisms responsible for the regulation of membrane channels. The contributions could be either original research papers or reviews that bring together recent advances in understanding the molecular mechanism used in a specific membrane channel that underlies the observed phenomenology. Although new insights into well-studied channels are welcome, even more desirable are contributions that provide mechanistic insights into unusual channel-formers. We would like to include as much as possible of the entire spectrum of molecular mechanisms that are known to exist in nature.

Prof. Dr. Marco Colombini
Dr. Sergey M. Bezrukov
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. International Journal of Molecular Sciences is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. There is an Article Processing Charge (APC) for publication in this open access journal. For details about the APC please see here. Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • voltage gating mechanism
  • chemical gating mechanism
  • tension gating mechanism
  • channel forming antibiotic
  • pore forming mechanism
  • cochlear channel mechanism
  • sensory channel mechanism
  • porin gating mechanism
  • ion channel mechanism
  • metabolite channel mechanism
  • ATP channel mechanism
  • calcium channel gating mechanism
  • sodium channel gating mechanism
  • potassium channel gating mechanism
  • connexin gating mechanism
 
 

Published Papers (8 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Review

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 II 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 ">
18 pages, 2909 KiB  
Article
Beta-Barrel Channel Response to High Electric Fields: Functional Gating or Reversible Denaturation?
by Ekaterina M. Nestorovich and Sergey M. Bezrukov
Int. J. Mol. Sci. 2023, 24(23), 16655; https://doi.org/10.3390/ijms242316655 - 23 Nov 2023
Cited by 1 | Viewed by 1025
Abstract
Ion channels exhibit gating behavior, fluctuating between open and closed states, with the transmembrane voltage serving as one of the essential regulators of this process. Voltage gating is a fundamental functional aspect underlying the regulation of ion-selective, mostly α-helical, channels primarily found in [...] Read more.
Ion channels exhibit gating behavior, fluctuating between open and closed states, with the transmembrane voltage serving as one of the essential regulators of this process. Voltage gating is a fundamental functional aspect underlying the regulation of ion-selective, mostly α-helical, channels primarily found in excitable cell membranes. In contrast, there exists another group of larger, and less selective, β-barrel channels of a different origin, which are not directly associated with cell excitability. Remarkably, these channels can also undergo closing, or “gating”, induced by sufficiently strong electric fields. Once the field is removed, the channels reopen, preserving a memory of the gating process. In this study, we explored the hypothesis that the voltage-induced closure of the β-barrel channels can be seen as a form of reversible protein denaturation by the high electric fields applied in model membranes experiments—typically exceeding twenty million volts per meter—rather than a manifestation of functional gating. Here, we focused on the bacterial outer membrane channel OmpF reconstituted into planar lipid bilayers and analyzed various characteristics of the closing-opening process that support this idea. Specifically, we considered the nearly symmetric response to voltages of both polarities, the presence of multiple closed states, the stabilization of the open conformation in channel clusters, the long-term gating memory, and the Hofmeister effects in closing kinetics. Furthermore, we contemplate the evolutionary aspect of the phenomenon, proposing that the field-induced denaturation of membrane proteins might have served as a starting point for their development into amazing molecular machines such as voltage-gated channels of nerve and muscle cells. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p>Voltage-dependent closing-opening transitions of OmpF were observed at the single-channel (<b>A</b>) and multichannel (<b>B</b>,<b>C</b>) levels. (<b>A</b>) Typical recordings of ion current through a single trimeric OmpF channel reconstituted into PLMs taken at ±75 mV and ±150 mV applied voltage. Notably, higher voltages lead to the channel closing in approximately three equal but not identical steps. Dashed lines indicate zero current (<span class="html-italic">top</span>) and voltage (<span class="html-italic">bottom</span>) levels. Short-dashed lines designate fully open (L3), one monomer closed (L2), two monomers closed (L1), and residual current (L0) states, with two asterisks (*) highlighting current variations in the L1 state. (<b>B</b>) Raw ion current data from a multichannel system containing around 75 OmpF channels (<span class="html-italic">bottom</span>), subjected to a three-fold repetition of a −200 mV to +200 mV voltage ramp at 1 mHz frequency (<span class="html-italic">top</span>). (<b>C</b>) I-V curves, replotted from the experiments presented in panel (<b>B</b>), at three frequencies (5 mHz, 2 mHz, and 1 mHz), averaged over three ramp cycles, elucidating voltage gating at both polarities. The ion channel recordings were additionally filtered using a 1000 Hz (<b>A</b>) and 100 Hz low-pass Bessel filter (<b>B</b>,<b>C</b>), and data reduction (reduction factor 100, Clampfit, Molecular Devices, San Jose, CA, USA) was applied. Here, the membrane bathing solutions contained 1 M KCl and 5 mM HEPES at pH 7.4.</p>
Full article ">Figure 2
<p>The conductance of a multichannel membrane, responding to a voltage ramp from 0.25 mHz to 10 mHz, reveals frequency-dependent sensitivity in its hysteresis curves. (<b>A</b>) Raw data showing ion current across approximately 75 OmpF channels (<span class="html-italic">bottom</span>) in response to a series of repeated 0 to 150 mV voltage ramps at a frequency of 1 mHz (<span class="html-italic">top</span>). (<b>B</b>) Conductance hysteresis curves at different ramp frequencies. Arrows indicate voltage change directions during channel closing and opening within the hysteresis loop. The data represents an average across 3–6 ramp periods. (<b>C</b>) Areas encircled by hysteresis curves from panel (<b>B</b>) as a function of ramp frequency. The data averaged over 3–5 independent experiments with 60–120 OmpF channels. (<b>D</b>) Conductance hysteresis curves from panel (<b>B</b>) after transformation according to Equation (1) in the text. Normalized hysteresis curves show that both the closing and opening branches of the hysteresis curves nearly overlap. (<b>E</b>,<b>F</b>) Gating parameters, the voltage of half-effect <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> (<b>E</b>), and the effective “gating charge” <math display="inline"><semantics> <mrow> <mi>n</mi> </mrow> </semantics></math> (<b>F</b>) obtained by fitting the opening and closing branches of the panel (<b>D</b>) curves to Equation (2) in the text, showing hardly any dependence on the ramp frequency.</p>
Full article ">Figure 3
<p>Multichannel conductance relaxation experiment. The experiment was started at the applied voltage of 2.5 mV, then, for durations of 50 s, it was switched to 120 mV, 180 mV, back to 120 mV, and then returned to 2.5 mV as shown in the voltage protocol at the bottom part. Solid lines depict the best-fit single exponentials in the 50 s intervals (the fast component of the opening branch at 100 s is ignored), with the equations displayed in the figure. The lines do not converge to the same value, illustrating the long-term memory of the OmpF voltage-induced closing and opening. The data represent averages from 10 runs. Recordings were filtered using a 100 Hz low-pass Bessel filter, and data reduction (reduction factor 100, Clampfit, Molecular Devices) was applied.</p>
Full article ">Figure 4
<p>The current distribution of voltage-induced closing events reveals the variety of conformational states within the closed states of OmpF. (<b>A</b>) An experimental procedure for quantifying the current of the open OmpF ion channel (<b>B</b>), voltage-induced closing event currents (<b>C</b>–<b>E</b>), and residual conductance current (<b>F</b>). To mitigate current variations between channels, data were collected from the same single OmpF channel by repeatedly applying a 150 mV voltage and recording its 3-step closings, similar to those depicted in panel (<b>A</b>), until sufficient statistical data were accumulated. The current histogram from a single open OmpF channel (L3) shown in panel (<b>B</b>), acquired after closings under 150 mV and the subsequent reopenings at 0 mV, displays a narrow current distribution of 651 ± 9 pA. The current histograms in panels C-E of the 1st (L3 → L2), 2nd (L2 → L1), and 3rd (L1 → L0) monomer closings, with the arrows showing transitions, as well as of the residual current (L0) in panel (<b>F</b>), exhibit significantly broader ranges compared to the narrow distribution of the open state current. To ensure easy comparison, panels (<b>B</b>–<b>F</b>) are graphed with the equivalent 250-pA range on the <span class="html-italic">X</span>-axis. Interestingly, the closing event current histograms in panels (<b>C</b>–<b>E</b>) extend beyond one-third of the total channel current, suggesting that the consequences of a single monomer closing could impact nearby monomer conformations and their associated currents. Note that in panel (<b>F</b>), the single Gaussian fit is drawn to guide the eye only as it significantly extends into the unrealistic negative current values.</p>
Full article ">Figure 5
<p>Hofmeister effect in OmpF voltage-induced closing. (<b>A</b>) Examples of raw currents through a single OmpF channel at the application of 150 mV. (<b>B</b>) The histograms of the first monomer closing times are fitted by a single exponential function. For illustrative purposes, only the 1 M LiCl (<span class="html-italic">left panel</span>), KCl (<span class="html-italic">middle panel</span>), and RbCl (<span class="html-italic">right panel</span>) data are shown. (<b>C</b>) There is an order of magnitude difference in closing times in the LiCl, NaCl, KCl, RbCl, and CsCl series. (<b>D</b>,<b>E</b>) Multichannel conductance hysteresis curves at 1 mHz ramp frequency (<b>D</b>) show an increase of more than four times in the areas encircled by hysteresis curves (<b>E</b>) in the LiCl, NaCl, KCl, RbCl, CsCl series. (<b>F</b>,<b>G</b>) Multichannel conductance 1 mHz hysteresis curves for different anions. The influence of anions on the OmpF voltage-induced closing and opening (<b>F</b>) is seen as a more than fourfold increase in the areas encircled by hysteresis curves (<b>G</b>) in the KF, KCl, and KBr series.</p>
Full article ">Figure 6
<p>Impact of protein cluster insertion on voltage-induced closing. (<b>A</b>,<b>B</b>) A typical multichannel experiment wherein OmpF is exclusively inserted into the PLM as trimeric units, resulting in an average current of 0.23 ± 0.01 nA at 50 mV applied voltage. Notably, upon the application of 150 mV voltage, the channels in this multichannel membrane display a relatively swift and nearly complete closing. OmpF was added into the <span class="html-italic">cis</span> chamber of the bilayer chamber using a 0.22 μg/mL solution following a 1:10,000 stock solution dilution performed approximately eight weeks prior to the experiment. Panel (<b>B</b>) displays a cluster size distribution graph exclusively depicting trimeric OmpF single channel insertions. (<b>C</b>,<b>D</b>) A typical experiment in which about 35% of the total multichannel current was a result of the insertion of OmpF clusters. These clusters comprised thirteen dimers, three trimers, four tetramers, and one octamer, as seen from the cluster distribution graph in panel (<b>D</b>). The remaining current was generated by the insertion of 88 single-channel OmpF trimers. Notably, a significant reduction in the degree of channel closing is observed at 150 mV ((<b>C</b>), <span class="html-italic">right panel</span>). OmpF was added to the <span class="html-italic">cis</span> size of the bilayer chamber using a freshly diluted 0.22 μg/mL solution. (<b>E</b>,<b>F</b>) A typical experiment wherein the insertion of OmpF clusters accounted for over 70% of the total multichannel membrane conductance. These clusters consisted of eight dimers, three trimers, two pentamers, and one decamer (<b>F</b>). Upon applying a voltage of 150 mV ((<b>E</b>), <span class="html-italic">right panel</span>), the effect of voltage on channel closing noticeably diminishes, resulting in a substantial residual conductance that corresponds to approximately 74% of the initial current. OmpF was added to the <span class="html-italic">cis</span> size of the bilayer chamber using a freshly diluted 22 µg/mL solution. The lower panel illustrates the voltage protocol employed for panels (<b>A</b>,<b>C</b>,<b>E</b>). For illustrative purposes, in panel (<b>C</b>), a yellow arrow marks one octamer cluster insertion. In panel (<b>E</b>), orange arrows indicate two pentamer cluster insertions, while a magenta arrow marks one decamer cluster insertion. The same color code is used in panels (<b>B</b>,<b>D</b>,<b>F</b>). The ion channel recordings were filtered using a 50 Hz Bessel filter and data reduction (reduction factor 100, Clampfit, Molecular Devices) was applied.</p>
Full article ">Figure 7
<p>One-dimensional cross-section of an energy landscape illustrating voltage-induced channel protein denaturation. Hofmeister series ions are able to change the height of the barrier separating the functional folded state F and the multitude of (partially) unfolded ones U.</p>
Full article ">
15 pages, 2071 KiB  
Article
Modulation of Voltage-Gating and Hysteresis of Lysenin Channels by Cu2+ Ions
by Andrew Bogard, Pangaea W. Finn, Aviana R. Smith, Ilinca M. Flacau, Rose Whiting and Daniel Fologea
Int. J. Mol. Sci. 2023, 24(16), 12996; https://doi.org/10.3390/ijms241612996 - 20 Aug 2023
Cited by 1 | Viewed by 982
Abstract
The intricate voltage regulation presented by lysenin channels reconstituted in artificial lipid membranes leads to a strong hysteresis in conductance, bistability, and memory. Prior investigations on lysenin channels indicate that the hysteresis is modulated by multivalent cations which are also capable of eliciting [...] Read more.
The intricate voltage regulation presented by lysenin channels reconstituted in artificial lipid membranes leads to a strong hysteresis in conductance, bistability, and memory. Prior investigations on lysenin channels indicate that the hysteresis is modulated by multivalent cations which are also capable of eliciting single-step conformational changes and transitions to stable closed or sub-conducting states. However, the influence on voltage regulation of Cu2+ ions, capable of completely closing the lysenin channels in a two-step process, was not sufficiently addressed. In this respect, we employed electrophysiology approaches to investigate the response of lysenin channels to variable voltage stimuli in the presence of small concentrations of Cu2+ ions. Our experimental results showed that the hysteretic behavior, recorded in response to variable voltage ramps, is accentuated in the presence of Cu2+ ions. Using simultaneous AC/DC stimulation, we were able to determine that Cu2+ prevents the reopening of channels previously closed by depolarizing potentials and the channels remain in the closed state even in the absence of a transmembrane voltage. In addition, we showed that Cu2+ addition reinstates the voltage gating and hysteretic behavior of lysenin channels reconstituted in neutral lipid membranes in which lysenin channels lose their voltage-regulating properties. In the presence of Cu2+ ions, lysenin not only regained the voltage gating but also behaved like a long-term molecular memory controlled by electrical potentials. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p>Lysenin inserts uniform channels in artificial lipid membranes. The insertion of individual lysenin channels in a planar bilayer lipid membrane was monitored from the stepwise variation of the ionic currents at −60 mV transmembrane voltage. Each inserted channel adjusted the ionic current by ~20 pA.</p>
Full article ">Figure 2
<p>Cu<sup>2+</sup> ions adjust the voltage gating and hysteresis in the positive voltage range. (<b>a</b>) The I-V plots recorded for forward and reverse voltage ramps before and after Cu<sup>2+</sup> addition indicate major adjustments in channel closing and reopening. Cu<sup>2+</sup> addition reduces the voltage required to initiate gating during ascending (Fwd) voltage ramps and elicit resistance to reopening during descending voltage ramps (Rev). (<b>b</b>) The negligible ionic currents recorded for a consecutive voltage ramp applied to Cu<sup>2+</sup> -exposed channels indicates the persistency of the closed state of the channels. The panels show experimental data from single traces, with the symbols added to facilitate identification.</p>
Full article ">Figure 3
<p>Cu<sup>2+</sup> ions modulate the voltage gating and hysteresis of lysenin channels for an extended voltage range. In the absence of Cu<sup>2+</sup> ions, the hysteresis in conductance in response to ascending and descending voltage ramps is observed for I-V (<b>a</b>) and open probability (P<sub>open</sub>) (<b>b</b>) plots. The changes in the macroscopic currents, P<sub>open</sub>, and midway voltage of activation for channels in a previously open state indicate a history-dependent response to applied voltages. Cu<sup>2+</sup> addition influences the I-V (<b>c</b>) and P<sub>open</sub> (<b>d</b>) plots recorded in response to the oscillatory voltage stimuli. The addition of Cu<sup>2+</sup> ions induces a strong leftward shift in gating during ascending voltage ramps and the previously closed channels resist reopening. The traces represent experimental data, from single traces, with the symbols added to facilitate identification.</p>
Full article ">Figure 4
<p>Determination of lysenin channels’ macroscopic conductance from combined AC/DC stimulation. The application of 0 mV and +60 mV DC voltages is indicated in the figure. The fully open state at 0 mV at the beginning of the recording is indicated by the large value of the AC current amplitude. The decreasing amplitude observed after the application of +60 mV indicates channel closure. The re-application of +60 mV reinstates the fully conducting state, which is indicative of channel reopening.</p>
Full article ">Figure 5
<p>Cu<sup>2+</sup> ions adjust the response of lysenin channels to voltages in a history-dependent manner. A large macroscopic conductance at 0 mV for previously open channels is indicated by the large amplitude of the AC current. The channels close rapidly at +60 mV but removal of the DC bias voltage does not lead to reopening. The channels reopen upon application of a negative bias voltage step (−60 mV); subsequent removal of the DC stimulus (i.e., reapplication of 0 mV) reinstates the current prior to channel closure, demonstrating a bistable system.</p>
Full article ">Figure 6
<p>The power spectrum recorded for all the voltage and conformation conditions indicate the presence of the 10 Hz AC signal for channels open at 0 mV (squares), closed at +60 mV (circles), closed at 0 mV (up triangles), reopen at −60 mV (down triangles), and again at 0 mV after reopening by the negative step voltage (diamonds). The symbols were added to traces constructed from all the experimental data to facilitate identification.</p>
Full article ">Figure 7
<p>Cu<sup>2+</sup> reinstates the voltage gating and hysteresis features of lysenin channels in neutral membranes. The linear I-V plot recorded for lysenin channels reconstituted in neutral membranes ((<b>a</b>), open squares) indicates the absence of voltage-induced gating. The voltage-induced gating feature is reinstated upon Cu<sup>2+</sup> addition ((<b>a</b>), open up triangles). The I-V plot recorded after Cu<sup>2+</sup> addition (<b>b</b>) for ascending (full squares) and descending (full down triangles) voltage ramps indicates the hysteresis in conductance and history-dependent response to applied voltages. The plots represent experimental data from single traces, with the symbols added to facilitate identification.</p>
Full article ">Figure 8
<p>The experimental setup for electrophysiology measurements. Lysenin channels are reconstituted in a bilayer lipid membrane (BLM) bathed by electrolyte solutions. The electrical connections to the Axopatch 200B electrophysiology amplifier are ensured by agarose salt bridges and Ag/AgCl electrodes wired to the headstage. The diagram is not to scale.</p>
Full article ">Figure 9
<p>Experimental setup for investigating the status of the channels at any applied DC voltage. A combined AC/DC signal is applied from the electrophysiology amplifier to the membrane depicted as a capacitor C<sub>m</sub> in parallel to a variable resistor R (the conducting pathway created by inserted lysenin channels).</p>
Full article ">
13 pages, 2145 KiB  
Article
The Complex Proteolipidic Behavior of the SARS-CoV-2 Envelope Protein Channel: Weak Selectivity and Heterogeneous Oligomerization
by Wahyu Surya, Ernesto Tavares-Neto, Andrea Sanchis, María Queralt-Martín, Antonio Alcaraz, Jaume Torres and Vicente M. Aguilella
Int. J. Mol. Sci. 2023, 24(15), 12454; https://doi.org/10.3390/ijms241512454 - 5 Aug 2023
Viewed by 1348
Abstract
The envelope (E) protein is a small polypeptide that can form ion channels in coronaviruses. In SARS coronavirus 2 (SARS-CoV-2), the agent that caused the recent COVID-19 pandemic, and its predecessor SARS-CoV-1, E protein is found in the endoplasmic reticulum–Golgi intermediate compartment (ERGIC), [...] Read more.
The envelope (E) protein is a small polypeptide that can form ion channels in coronaviruses. In SARS coronavirus 2 (SARS-CoV-2), the agent that caused the recent COVID-19 pandemic, and its predecessor SARS-CoV-1, E protein is found in the endoplasmic reticulum–Golgi intermediate compartment (ERGIC), where virion budding takes place. Several reports claim that E protein promotes the formation of “cation-selective channels”. However, whether this term represents specificity to certain ions (e.g., potassium or calcium) or the partial or total exclusion of anions is debatable. Herein, we discuss this claim based on the available data for SARS-CoV-1 and -2 E and on new experiments performed using the untagged full-length E protein from SARS-CoV-2 in planar lipid membranes of different types, including those that closely mimic the ERGIC membrane composition. We provide evidence that the selectivity of the E-induced channels is very mild and depends strongly on lipid environment. Thus, despite past and recent claims, we found no indication that the E protein forms cation-selective channels that prevent anion transport, and even less that E protein forms bona fide specific calcium channels. In fact, the E channel maintains its multi-ionic non-specific neutral character even in concentrated solutions of Ca2+ ions. Also, in contrast to previous studies, we found no evidence that SARS-CoV-2 E channel activation requires a particular voltage, high calcium concentrations or low pH, in agreement with available data from SARS-CoV-1 E. In addition, sedimentation velocity experiments suggest that the E channel population is mostly pentameric, but very dynamic and probably heterogeneous, consistent with the broad distribution of conductance values typically found in electrophysiological experiments. The latter has been explained by the presence of proteolipidic channel structures. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p><b>Comparison of SARS E sequences.</b> Sequence comparison between SARS-1 E (accession number P59637) and SARS-2 E (accession number QHZ00401). Transmembrane domain (TMD) is shown in blue. SARS-2 E only differs from SARS-1 E at three mutations and one deletion (underlined in SARS-1 E sequence). The samples used in the present paper (gray rectangle) correspond to SARS-2: full length (E-FL), truncated with a His-tag (His-E-TR) or E-TM. Both E-FL and E-TR contain additional three N-terminal residues (SNA) arising from the N-terminal fusion tag that was cleaved in E-FL, and native Cys residues were mutated to Ala (underlined). For reference, charged residues are indicated as blue (positive) or red (negative) in bold.</p>
Full article ">Figure 2
<p><b>Ohmic character of SARS-2 E-FL in KCl and CaCl<sub>2</sub>.</b> Example of stable current traces (<b>A</b>,<b>C</b>) and corresponding I–V curves (<b>B</b>,<b>D</b>) for SARS-2 E-FL channels in 500/50 mM KCl at the applied voltages shown in gray below each current trace; (<b>C</b>,<b>D</b>) same as (<b>A</b>,<b>B</b>) but using 500/50 mM CaCl<sub>2</sub>. The solid line in (<b>B</b>,<b>D</b>) is a linear fit to the experimental data.</p>
Full article ">Figure 3
<p><b>Selectivity (RP) of E protein in different lipid compositions.</b> (<b>A</b>) RP measured in KCl with E-TM 7–38, SARS-1 E-FL or SARS-2 E-FL in neutral DPhPC (black) or negatively charged DPhPS (blue) membranes; (<b>B</b>) RP measured in CaCl<sub>2</sub> solutions with SARS-1 E-FL or SARS-2 E-FL in membranes with 20% negatively charged lipids (ERGIC lipid composition). Numbers represent permeability ratios, P<sub>+</sub>/P<sub>−</sub> (<b>A</b>) or P<sub>−</sub>/P<sub>+</sub> (<b>B</b>). The horizontal lines indicate the RP of a neutral channel (V<sub>diff</sub> = 0.93 mV, which corresponds to P<sub>+</sub>/P<sub>−</sub> = 0.95 (<b>A</b>) or V<sub>diff</sub> = 17.8 mV corresponding to P<sub>−</sub>/P<sub>+</sub> = 3.49 (<b>B</b>)). The concentration gradient was 500/50 mM in all selectivity measurements and all RP values were corrected using the corresponding liquid junction potential from Henderson’s equation to eliminate the contribution of the electrode’s agarose bridges [<a href="#B27-ijms-24-12454" class="html-bibr">27</a>]. Number of experiments was <span class="html-italic">n</span> = 20 (black circle), 15 (black triangle), 16 (black square), 10 (blue circle), 10 (blue triangle), 26 (blue square), 10 (green triangle) and 19 (green square). Error bars indicate ± standard deviation. Data for SARS-1 were reported in [<a href="#B10-ijms-24-12454" class="html-bibr">10</a>,<a href="#B11-ijms-24-12454" class="html-bibr">11</a>].</p>
Full article ">Figure 4
<p><b>AUC-SV profile of E protein in C14SB detergent micelles.</b> Comparison of <span class="html-italic">c</span>(<span class="html-italic">s</span>) size distribution of (<b>A</b>) E-TM, (<b>B</b>) His-E-TR and (<b>C</b>) E-FL at the indicated detergent-to-protein ratios (DPR, right). The predicted protein molecular weight according to SEDFIT using a common <span class="html-italic">f</span>/<span class="html-italic">f</span><sub>0</sub> for all species is shown with a number (kDa) above each band. Alternatively, the predicted range of S values for various n-oligomeric states (n is shown on the right of the bar) are shown as colored bars. These ranges were calculated as described in the <a href="#sec4-ijms-24-12454" class="html-sec">Section 4</a>; (<b>D</b>–<b>G</b>) two-dimensional plots <span class="html-italic">c</span>(<span class="html-italic">s</span>, <span class="html-italic">f</span>/<span class="html-italic">f</span><sub>0</sub>) corresponding to E-TM (<b>D</b>), E-TR (<b>E</b>,<b>F</b>) and E-FL (<b>G</b>) at the DPRs indicated.</p>
Full article ">
14 pages, 1835 KiB  
Article
Triplin: Mechanistic Basis for Voltage Gating
by Marco Colombini, Patrick Liu and Chase Dee
Int. J. Mol. Sci. 2023, 24(14), 11473; https://doi.org/10.3390/ijms241411473 - 14 Jul 2023
Cited by 1 | Viewed by 879
Abstract
The outer membrane of Gram-negative bacteria contains a variety of pore-forming structures collectively referred to as porins. Some of these are voltage dependent, but weakly so, closing at high voltages. Triplin, a novel bacterial pore-former, is a three-pore structure, highly voltage dependent, with [...] Read more.
The outer membrane of Gram-negative bacteria contains a variety of pore-forming structures collectively referred to as porins. Some of these are voltage dependent, but weakly so, closing at high voltages. Triplin, a novel bacterial pore-former, is a three-pore structure, highly voltage dependent, with a complex gating process. The three pores close sequentially: pore 1 at positive potentials, 2 at negative and 3 at positive. A positive domain containing 14 positive charges (the voltage sensor) translocates through the membrane during the closing process, and the translocation is proposed to take place by the domain entering the pore and thus blocking it, resulting in the closed conformation. This mechanism of pore closure is supported by kinetic measurements that show that in the closing process the voltage sensor travels through most of the transmembrane voltage before reaching the energy barrier. Voltage-dependent blockage of the pores by polyarginine, but not by a 500-fold higher concentrations of polylysine, is consistent with the model of pore closure, with the sensor consisting mainly of arginine residues, and with the presence, in each pore, of a complementary surface that serves as a binding site for the sensor. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p>Voltage gating in a single Triplin. On the left side, a triangular voltage wave (+75/−77mV; 30 mHz) results in no pore closure. In region “A”, a +69 mV potential was applied to the <span class="html-italic">cis</span> compartment. At point “B”, pore 1 closed. The resumption of the triangular wave caused gating of pores 2 and 3. “C” indicates the location of pore 2 closure and “D” the locations of pore 2 reopening. “E” is the point at which pore 3 closed and at “F” pores 2 and 3 opened simultaneously.</p>
Full article ">Figure 2
<p>Model of the gating of Triplin. The top of the structure is the <span class="html-italic">cis</span> side of the membrane, the side from which Triplin is inserted. The bottom of the structure is the <span class="html-italic">trans</span> side and that is the side maintained at virtual ground by the amplifier. The numbers refer to pores 1, 2 and 3. All indicated voltages refer to the <span class="html-italic">cis</span> side. For simplicity, the closed state of the pore is illustrated as a result of blockage by a single loop of the beta barrel, but, of course, multiple loops may be involved. Blue regions are positively charged whereas red are negatively charged. From Figure 14 in ref [<a href="#B1-ijms-24-11473" class="html-bibr">1</a>].</p>
Full article ">Figure 3
<p>Pore 2 closing time at three different voltages, labeled (<b>A</b>–<b>C</b>). In all records, the initial voltage was 10 mV (short segment on left). The voltage was then switched to the indicated value, and at the point indicated by the arrow, pore 2 closed. The voltage was then switched to 10 mV, and shortly thereafter, the pore reopened. Zero current is indicated by the short line on the left side. The inset shows a plot of the average closure time (the time constant) as a function of voltage for that experiment.</p>
Full article ">Figure 4
<p>Voltage dependence of the opening and closing time constants for pore 2. The error bars are standard errors of the mean of 20 measurements.</p>
Full article ">Figure 5
<p>Pore 3 closure interferes with the opening of pore 2. Four sample records are illustrated, two taken at 25 mV (<b>C</b>,<b>D</b>) and two at 30 mV (<b>A</b>,<b>B</b>). Pore 2 was closed at −40 mV, and then the positive voltage indicated was applied. The downward events are pore 3 closures: some are transient and other are long-lived. In the case of record (<b>C</b>), the reopening of pore 3 after its closure in the middle of the record, took place at a time beyond the end of the record shown. Hence, pore 2 opening is not visible in the record shown.</p>
Full article ">Figure 6
<p>Pore 3 closure blocks pore 2 opening. Pore 2 opening at various applied voltages was measured either by ignoring pore 3 closure (circle and square symbols) or by subtracting the time during which pore 3 was closed (triangles). Error bars are SEM of 18 to 22 measurements.</p>
Full article ">Figure 7
<p>Transient blocking events by the addition of the indicated amount of polyarginine to the cis side of a membrane containing a single Triplin with all pores open. The three records shown (<b>A</b>–<b>C</b>) are typical samples of the recorded blockage events. The zero current level is indicated just below the record. All records were collected in the presence of a 40 mV applied potential.</p>
Full article ">Figure 8
<p>Long-lived pore blockage in the presence of 0.4 µg/ml polyarginine. A single Triplin was still gating as demonstrated by the triangular voltage wave (+72/−71 mV) on the far left side. Pore 2 closure was followed by pore 3 closure and then simultaneous pore 2 and 3 opening at high positive voltages allowing polyarginine to block transiently. This was followed by applying a constant voltage (+50 mV) resulting in both transient and long-lived blockages. The expanded regions show that often both pores were blocked simultaneously but at times unblocking took place in two separate events. These regions were only expanded in the time axis.</p>
Full article ">Figure 9
<p>Voltage-dependent block of Triplin by polyarginine. A triangular voltage wave (30 mHz; +72 to −71 mV) was applied to a membrane containing a single Triplin activated by closing pore 1. The horizontal lines are the zero current levels. Polyarginine was added sequentially, and sample records are illustrated. Record (<b>A</b>) was taken before polyarginine addition, followed later by (<b>B</b>) and much later by (<b>C</b>). The amount of polyarginine present in the cis compartment during each recording is indicated.</p>
Full article ">Figure 10
<p>Voltage dependence of the formation of long-lived pore blocks by polyaginine at the indicated concentration. The power dependence of the exponential fit to the data beginning at 30 mV is indicated next to each curve.</p>
Full article ">Figure 11
<p>Model of the gating process used by each pore. Left is the open state with the sensor (blue) out of the pore. The red negative domain is close to the other end of the pore and proposed to be responsible for the selectivity of the pore and the rectification. Right shows the obstructed pore with the sensor (blue) interacting with a negative domain (red).</p>
Full article ">

Review

Jump to: Research

17 pages, 2708 KiB  
Review
Counter-Intuitive Features of Particle Dynamics in Nanopores
by Alexander M. Berezhkovskii and Sergey M. Bezrukov
Int. J. Mol. Sci. 2023, 24(21), 15923; https://doi.org/10.3390/ijms242115923 - 3 Nov 2023
Cited by 1 | Viewed by 739
Abstract
Using the framework of a continuous diffusion model based on the Smoluchowski equation, we analyze particle dynamics in the confinement of a transmembrane nanopore. We briefly review existing analytical results to highlight consequences of interactions between the channel nanopore and the translocating particles. [...] Read more.
Using the framework of a continuous diffusion model based on the Smoluchowski equation, we analyze particle dynamics in the confinement of a transmembrane nanopore. We briefly review existing analytical results to highlight consequences of interactions between the channel nanopore and the translocating particles. These interactions are described within a minimalistic approach by lumping together multiple physical forces acting on the particle in the pore into a one-dimensional potential of mean force. Such radical simplification allows us to obtain transparent analytical results, often in a simple algebraic form. While most of our findings are quite intuitive, some of them may seem unexpected and even surprising at first glance. The focus is on five examples: (i) attractive interactions between the particles and the nanopore create a potential well and thus cause the particles to spend more time in the pore but, nevertheless, increase their net flux; (ii) if the potential well-describing particle-pore interaction occupies only a part of the pore length, the mean translocation time is a non-monotonic function of the well length, first increasing and then decreasing with the length; (iii) when a rectangular potential well occupies the entire nanopore, the mean particle residence time in the pore is independent of the particle diffusivity inside the pore and depends only on its diffusivity in the bulk; (iv) although in the presence of a potential bias applied to the nanopore the “downhill” particle flux is higher than the “uphill” one, the mean translocation times and their distributions are identical, i.e., independent of the translocation direction; and (v) fast spontaneous gating affects nanopore selectivity when its characteristic time is comparable to that of the particle transport through the pore. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p>In the simplest, idealized case, a single transmembrane nanopore is represented by the right cylinder of a constant radius <math display="inline"><semantics> <mi>R</mi> </semantics></math> and length <math display="inline"><semantics> <mi>L</mi> </semantics></math>. What is the probability for a particle, starting at distance <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>0</mn> </msub> </mrow> </semantics></math> from the pore opening, to pass through the pore and escape to infinity on the opposite side of otherwise impermeable membrane?</p>
Full article ">Figure 2
<p>A more realistic representation of the channel structure (panel (<b>A</b>)) includes a coordinate-dependent radius <math display="inline"><semantics> <mrow> <mi>R</mi> <mfenced> <mi>x</mi> </mfenced> </mrow> </semantics></math>. The particle-channel interactions are characterized by a coordinate-dependent potential of mean force <math display="inline"><semantics> <mrow> <mi>U</mi> <mfenced> <mi>x</mi> </mfenced> </mrow> </semantics></math>, which, in addition to the obvious entropy contributions, may also include Coulomb, van der Waals, and other types of specific interactions between the translocating particle and pore walls (panel (<b>B</b>)). A simplified version of the potential of mean force (panel (<b>C</b>)) that allows us to obtain simple analytical expressions for quantities characterizing particle transport through the pore containing a rectangular potential well of depth <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>U</mi> </mrow> </semantics></math> occupying a fraction <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>/</mo> <mi>L</mi> </mrow> </semantics></math> of the total channel length <math display="inline"><semantics> <mi>L</mi> </semantics></math>.</p>
Full article ">Figure 3
<p>Particle flux as a function of the depth <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>U</mi> </mrow> </semantics></math> of a rectangular potential well occupying the entire channel pore (<a href="#ijms-24-15923-f002" class="html-fig">Figure 2</a>C, with <math display="inline"><semantics> <mrow> <mi>l</mi> <mo>=</mo> <mi>L</mi> </mrow> </semantics></math>). Flux is calculated according to Equation (10) for <math display="inline"><semantics> <mrow> <msub> <mi>c</mi> <mi>R</mi> </msub> </mrow> </semantics></math> = 0 and other parameters chosen as described in the text.</p>
Full article ">Figure 4
<p>Channel-facilitated flux as a function of the potential well depth for different lengths of the pore <span class="html-italic">L</span>. Flux is calculated according to Equation (10) at <span class="html-italic">c<sub>L</sub></span> = 100 μM (6.02 × 10<sup>22</sup> m<sup>−3</sup>), <span class="html-italic">c<sub>R</sub></span> = 0; other parameters are as those for <a href="#ijms-24-15923-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>The probability for a particle entering the channel pore to translocate (Equation (7)) as a function of the potential well depth. The parameters are the same as those for <a href="#ijms-24-15923-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 6
<p>The ratio of the mean translocation time <math display="inline"><semantics> <mrow> <msub> <mi>τ</mi> <mrow> <mi>t</mi> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math> in Equation (13) to its value at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> as a function of the fraction of the channel pore occupied by the symmetric square potential well, <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mi>l</mi> <mo>/</mo> <mi>L</mi> </mrow> </semantics></math> (<a href="#ijms-24-15923-f002" class="html-fig">Figure 2</a>C), for <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>, and well depths <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>Δ</mo> <mi>U</mi> </mrow> </semantics></math> = 2, 2.5, and 3 (from bottom to top).</p>
Full article ">Figure 7
<p>Depending on the molecule intra-channel diffusivity (indicated by the numbers near the curves), the flux through a stochastically gated channel can be significantly different from its conventional estimate. The flux ratio, Equation (19), is the ratio of the flux <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mi>g</mi> </msub> </mrow> </semantics></math>, calculated with gating/diffusion interference taken into account its conventional counterpart <math display="inline"><semantics> <mrow> <msubsup> <mi>J</mi> <mi>g</mi> <mrow> <mi>c</mi> <mi>o</mi> <mi>n</mi> <mi>v</mi> </mrow> </msubsup> </mrow> </semantics></math> given by Equation (18).</p>
Full article ">
12 pages, 1324 KiB  
Review
Mechanisms of PIEZO Channel Inactivation
by Zijing Zhou and Boris Martinac
Int. J. Mol. Sci. 2023, 24(18), 14113; https://doi.org/10.3390/ijms241814113 - 14 Sep 2023
Cited by 3 | Viewed by 2598
Abstract
PIEZO channels PIEZO1 and PIEZO2 are the newly identified mechanosensitive, non-selective cation channels permeable to Ca2+. In higher vertebrates, PIEZO1 is expressed ubiquitously in most tissues and cells while PIEZO2 is expressed more specifically in the peripheral sensory neurons. PIEZO channels [...] Read more.
PIEZO channels PIEZO1 and PIEZO2 are the newly identified mechanosensitive, non-selective cation channels permeable to Ca2+. In higher vertebrates, PIEZO1 is expressed ubiquitously in most tissues and cells while PIEZO2 is expressed more specifically in the peripheral sensory neurons. PIEZO channels contribute to a wide range of biological behaviors and developmental processes, therefore driving significant attention in the effort to understand their molecular properties. One prominent property of PIEZO channels is their rapid inactivation, which manifests itself as a decrease in channel open probability in the presence of a sustained mechanical stimulus. The lack of the PIEZO channel inactivation is linked to various mechanopathologies emphasizing the significance of studying this PIEZO channel property and the factors affecting it. In the present review, we discuss the mechanisms underlying the PIEZO channel inactivation, its modulation by the interaction of the channels with lipids and/or proteins, and how the changes in PIEZO inactivation by the channel mutations can cause a variety of diseases in animals and humans. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p><b>Essential domains and residues for PIEZO inactivation.</b> (<b>A</b>) Cap domain and inner helix are highlighted in the 3D structure of mouse PIEZO1 (PDB: 6BPZ). Critical amino acids in mouse PIEZO1, together with their positions, are highlighted in red; positions of amino acids in human PIEZO1 are in grey. The right panel presents a zoom-in diagram of the inner helix with the critical amino acids. (<b>B</b>) Structure overview of wild-type PIEZO1 and PIEZO2, or chimeric PIEZO1 or PIEZO2 fused with the ‘Cap’ domain (C terminal extracellular domain, CED) from each other, is shown in the left panel. Inactivation of PIEZO1 is faster than PIEZO2 (right panel, up). Swapping the CED exchanges the inactivation constant of the channels, as P1-P2<sub>CED</sub> inactivates faster than the P2-P1<sub>CED</sub> (right panel, down). The figure is adapted from (Wu et al., 2017) under CC BY-NC-ND 4.0. (<b>C</b>) Manipulating the hydrophobic gate by mutating L2475/V2476 of mouse PIEZO1 significantly reduces inactivation compared to the wild-type. Figure is adopted from [<a href="#B28-ijms-24-14113" class="html-bibr">28</a>] under CC BY-NC-ND 4.0.</p>
Full article ">Figure 2
<p><b>Interacting proteins may explain the slow-inactivating PIEZO1 currents in native cells.</b> (<b>A</b>) Indentation-induced whole-cell currents in N2A cells overexpressing mouse PIEZO1 (Left) or native PIEZO1 current in the mouse embryonic stem cells (Right). Native PIEZO1 shows a slow inactivating kinetics. (<b>B</b>) 3D structure of PIEZO1-MDFIC Complex. The C terminus of MDFIC is shown in green. MDFIC inserts into PIEZO1′s pore module and stays near the inner helix, as detailed in the right panel. (<b>C</b>) HEK cells overexpressing PIEZO1 show a rapid inactivating current (Left). Upon co-expression of MDFIC, the inactivation is largely removed (Right). Figures are adopted from [<a href="#B54-ijms-24-14113" class="html-bibr">54</a>] and [<a href="#B55-ijms-24-14113" class="html-bibr">55</a>] under CC BY-NC-ND 4.0.</p>
Full article ">
26 pages, 3737 KiB  
Review
Gating of β-Barrel Protein Pores, Porins, and Channels: An Old Problem with New Facets
by Lauren A. Mayse and Liviu Movileanu
Int. J. Mol. Sci. 2023, 24(15), 12095; https://doi.org/10.3390/ijms241512095 - 28 Jul 2023
Cited by 5 | Viewed by 1706
Abstract
β barrels are ubiquitous proteins in the outer membranes of mitochondria, chloroplasts, and Gram-negative bacteria. These transmembrane proteins (TMPs) execute a wide variety of tasks. For example, they can serve as transporters, receptors, membrane-bound enzymes, as well as adhesion, structural, and signaling elements. [...] Read more.
β barrels are ubiquitous proteins in the outer membranes of mitochondria, chloroplasts, and Gram-negative bacteria. These transmembrane proteins (TMPs) execute a wide variety of tasks. For example, they can serve as transporters, receptors, membrane-bound enzymes, as well as adhesion, structural, and signaling elements. In addition, multimeric β barrels are common structural scaffolds among many pore-forming toxins. Significant progress has been made in understanding the functional, structural, biochemical, and biophysical features of these robust and versatile proteins. One frequently encountered fundamental trait of all β barrels is their voltage-dependent gating. This process consists of reversible or permanent conformational transitions between a large-conductance, highly permeable open state and a low-conductance, solute-restrictive closed state. Several intrinsic molecular mechanisms and environmental factors modulate this universal property of β barrels. This review article outlines the typical signatures of voltage-dependent gating. Moreover, we discuss recent developments leading to a better qualitative understanding of the closure dynamics of these TMPs. Full article
(This article belongs to the Special Issue Membrane Channels: Mechanistic Insights)
Show Figures

Figure 1

Figure 1
<p>β-barrel proteins of Gram-negative bacteria. (<b>a</b>) OmpA (PDB:1QJP; [<a href="#B15-ijms-24-12095" class="html-bibr">15</a>]). (<b>b</b>) OmpT (PDB:6EHD; [<a href="#B39-ijms-24-12095" class="html-bibr">39</a>]). (<b>c</b>) OprD (OccD1) (PDB:3SY7; [<a href="#B40-ijms-24-12095" class="html-bibr">40</a>]). (<b>d</b>) OmpG(PDB:2F1C; [<a href="#B31-ijms-24-12095" class="html-bibr">31</a>]). (<b>e</b>) OmpF(PDB:2ZFG; [<a href="#B41-ijms-24-12095" class="html-bibr">41</a>]). (<b>f</b>) FhuA(PDB:1BY3; [<a href="#B16-ijms-24-12095" class="html-bibr">16</a>,<a href="#B17-ijms-24-12095" class="html-bibr">17</a>]). (<b>g</b>) PapC(PDB:3FIP; [<a href="#B18-ijms-24-12095" class="html-bibr">18</a>]). (<b>h</b>) TolC(PDB:7NG9; [<a href="#B42-ijms-24-12095" class="html-bibr">42</a>]).</p>
Full article ">Figure 2
<p>Mitochondrial β-barrel proteins. (<b>a</b>) VDAC-1 (porin) from <span class="html-italic">H. sapiens</span> (PDB:6TIQ; [<a href="#B43-ijms-24-12095" class="html-bibr">43</a>]). (<b>b</b>) TOM complex from <span class="html-italic">H. sapiens</span> (PDB:7VD2; [<a href="#B44-ijms-24-12095" class="html-bibr">44</a>]). (<b>c</b>) FhaC from <span class="html-italic">E. coli</span> (PDB:4QKY; [<a href="#B45-ijms-24-12095" class="html-bibr">45</a>,<a href="#B46-ijms-24-12095" class="html-bibr">46</a>,<a href="#B47-ijms-24-12095" class="html-bibr">47</a>]).</p>
Full article ">Figure 3
<p>β-barrel pore-forming toxins. (<b>a</b>) α-hemolysin of <span class="html-italic">S. aureus</span> (PDB:4ANZ; [<a href="#B48-ijms-24-12095" class="html-bibr">48</a>,<a href="#B49-ijms-24-12095" class="html-bibr">49</a>,<a href="#B50-ijms-24-12095" class="html-bibr">50</a>]). (<b>b</b>) Anthrax toxin with lethal factor side and top view (PDB:6PSN; [<a href="#B51-ijms-24-12095" class="html-bibr">51</a>]). (<b>c</b>) γ–hemolysin from <span class="html-italic">S. aureus</span> (PDB:3B07; [<a href="#B9-ijms-24-12095" class="html-bibr">9</a>]). (<b>d</b>) Aerolysin prepore side and top view from <span class="html-italic">A. hydrophila</span> (PDB: 5JZH/5JZW; [<a href="#B12-ijms-24-12095" class="html-bibr">12</a>]). (<b>e</b>) MspA of <span class="html-italic">M. smegmatis</span> (PDB:1UUN; [<a href="#B52-ijms-24-12095" class="html-bibr">52</a>]).</p>
Full article ">Figure 4
<p>Loop 6 is crucial for the gating dynamics of OmpG. (<b>a</b>) This is a cartoon representation of loop L6 of OmpG being anchored into the lipid bilayer via dodecylation at Cys226. (<b>b</b>) Representative single-channel electrical recordings using the wild-type OmpG (<b>left</b> panel) and an OmpG mutant with the loop L6 immobilized onto the lipid bilayer, as shown in (<b>a</b>) (<b>right</b> panel). This figure was adapted from Zhuang and Tamm (2014) [<a href="#B128-ijms-24-12095" class="html-bibr">128</a>].</p>
Full article ">Figure 5
<p>Gating evaluations of OmpG using single-molecule electrophysiology and high-speed AFM height spectroscopy (HS-AFM-HS). (<b>a</b>) A representative single-channel electrical trace of OmpG acquired at a transmembrane potential of +40 mV and pH 7.6. The schematic on the right side provides a scheme of the single-channel electrical recording experimental formulation. OmpG (yellow) is functionally reconstituted into a lipid bilayer (green). Potassium and chloride ions are indicated as red and blue spheres, respectively. The red arrow shows the direction of the ionic flow of cations at a positive applied potential. (<b>b</b>) A semilogarithmic dwell time histogram of the open and closed states, as determined by single-molecule electrophysiology. (<b>c</b>) A representative 60 ms long HS-AFM-HS recording that probes an OmpG protein functionally reconstituted into a lipid bilayer, which was suspended on mica at pH 7.6. The schematic on the right side is the HS-AFM-HS experimental setup. An AFM tip monitors conformational fluctuations of loop L6. (<b>d</b>) A semilogarithmic dwell time histogram of the open and closed states, as determined by HS-AFM-HS. Here, the low state indicates the open state, where the tip navigates within the pore lumen. The high state corresponds to the closed state, precluding the partitioning of the tip into the pore lumen. This figure was adapted from Sanganna Gari and coworkers (2021) [<a href="#B129-ijms-24-12095" class="html-bibr">129</a>].</p>
Full article ">Figure 6
<p>A proposed model for voltage sensing of VDAC1. VDAC1 (in blue) remains in a 4 nS-conductance open state at a zero transmembrane potential (upper, left). Yet, at an amplified applied transmembrane potential greater than 30 mV, regardless of its polarity, an electric force is exerted on the N-terminal helix that acts as a voltage sensor (in red; center). L10 (in green) is the contact residue of the N-terminal helix with the V143 residue on the barrel wall. The reversible dissociation of the rigid N-terminal helix from the pore wall results in a more flexible structure, which is likely to switch the channel into a semi-collapsed, elliptical conformation that leads to a 2 nS conductance closed state (upper, right). The lower panel indicates the correlated values in the open and closed state unitary conductance and ionic selectivity. This figure was adapted from Zachariae and coworkers (2013) [<a href="#B136-ijms-24-12095" class="html-bibr">136</a>].</p>
Full article ">Figure 7
<p>Direct experimental evidence for the implication of a key charged residue in the voltage-dependent gating of VDAC1. (<b>a</b>) Single-channel electrical recordings of mVDAC1 reveal the intense gating activity of the wild-type channel (left traces) but the drastically declined gating activity of the charge-reversal K12E mutant (right traces). Horizontal dashed lines show the zero current. (<b>b</b>) These panels indicate quantitative assessments of the gating activity of different VDAC1 proteins using a multichannel system. The vertical axis indicates the overall multichannel current normalized to the value corresponding to open-state multichannel conductance. The left panel compares the wild-type (WT) protein and the charge-reversal K12E mutant. The right panel compares the WT protein as well as the K12A and K12S mutants. This figure was adapted from Ngo and coworkers (2022) [<a href="#B161-ijms-24-12095" class="html-bibr">161</a>].</p>
Full article ">Figure 8
<p>Temperature dependence of conductance substates of OpdK. (<b>a</b>) Single-channel electrical traces collected with the native OpdK at various temperatures. (<b>b</b>) A free energy landscape model illustrating the kinetic transitions among the O<sub>1</sub>, O<sub>2</sub>, and O<sub>3</sub> open substates. This model shows the activation free energies characterizing various kinetic transitions (Δ<span class="html-italic">G</span><sub>O1→O2</sub><sup>‡</sup>, Δ<span class="html-italic">G</span><sub>O2→O1</sub><sup>‡</sup>, Δ<span class="html-italic">G</span><sub>O2→O3</sub><sup>‡</sup>, and Δ<span class="html-italic">G</span><sub>O3→O2</sub><sup>‡</sup>). This figure was adapted from Cheneke and coworkers (2015) [<a href="#B221-ijms-24-12095" class="html-bibr">221</a>].</p>
Full article ">
Back to TopTop