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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (19)

Search Parameters:
Keywords = SLO channels

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 2233 KiB  
Article
Structure–Activity Relationship Studies in a Series of 2-Aryloxy-N-(pyrimidin-5-yl)acetamide Inhibitors of SLACK Potassium Channels
by Nigam M. Mishra, Brittany D. Spitznagel, Yu Du, Yasmeen K. Mohamed, Ying Qin, C. David Weaver and Kyle A. Emmitte
Molecules 2024, 29(23), 5494; https://doi.org/10.3390/molecules29235494 - 21 Nov 2024
Viewed by 660
Abstract
Epilepsy of infancy with migrating focal seizures (EIMFS) is a rare, serious, and pharmacoresistant epileptic disorder often linked to gain-of-function mutations in the KCNT1 gene. KCNT1 encodes the sodium-activated potassium channel known as SLACK, making small molecule inhibitors of SLACK channels a compelling [...] Read more.
Epilepsy of infancy with migrating focal seizures (EIMFS) is a rare, serious, and pharmacoresistant epileptic disorder often linked to gain-of-function mutations in the KCNT1 gene. KCNT1 encodes the sodium-activated potassium channel known as SLACK, making small molecule inhibitors of SLACK channels a compelling approach to the treatment of EIMFS and other epilepsies associated with KCNT1 mutations. In this manuscript, we describe a hit optimization effort executed within a series of 2-aryloxy-N-(pyrimidin-5-yl)acetamides that were identified via a high-throughput screen. We systematically prepared analogs in four distinct regions of the scaffold and evaluated their functional activity in a whole-cell, automated patch clamp (APC) assay to establish structure-activity relationships for wild-type (WT) SLACK inhibition. Two selected analogs were also profiled for selectivity versus other members of the Slo family of potassium channels, of which SLACK is a member, and versus a panel of structurally diverse ion channels. The same two analogs were evaluated for activity versus the WT mouse channel as well as two clinically relevant mutant human channels. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Examples of small molecule SLACK inhibitors that have been reported in the literature. Reported IC<sub>50</sub> values obtained using whole-cell, patch clamp electrophysiology versus WT SLACK. Refer to original references cited in the text for experimental details, as cell lines and protocols differ.</p>
Full article ">Figure 2
<p>Hit optimization plan for SLACK inhibitor <b>10</b> (VU0545326) in four regions.</p>
Full article ">Scheme 1
<p>Synthesis of analogs. <span class="html-italic">Reagents and Conditions:</span> (<b>a</b>) ArXH, K<sub>2</sub>CO<sub>3</sub>, CH<sub>3</sub>CN, µwave, 80 °C, 30 min (For X=NMe, an additional reaction with K<sub>2</sub>CO<sub>3</sub>, MeI, DMF was employed); (<b>b</b>) H<sub>2</sub>, Pt(sulfided)/C, MeOH, 70 °C, H-Cube<sup>®</sup>; (<b>c</b>) (4-chlorophenyl)-ZYCO<sub>2</sub>H, HATU, DIEA, DMF; (<b>d</b>) 2-bromo-2-methylpropionyl bromide, DIEA, CH<sub>2</sub>Cl<sub>2</sub>; (<b>e</b>) ArZH, CuBr·SMe<sub>2</sub>, PCy<sub>3</sub>, K<sub>3</sub>PO<sub>4</sub>, CH<sub>3</sub>CN.</p>
Full article ">Scheme 2
<p>Synthesis of analogs. Reagents and Conditions: (<b>a</b>) HN(<span class="html-italic">i</span>-Pr)<sub>2</sub>, <span class="html-italic">n</span>-BuLi (1.6M in hexanes), THF, −78 °C, then 4-chlorobenzyl bromide, 55%; (<b>b</b>) NaOH, MeOH, H<sub>2</sub>O, 60 °C, then 1N aq. HCl, 75%; (<b>c</b>) <b>13a</b> (X=O, R=2-F), HATU, DIEA, DMF; (<b>d</b>) 4-chlorophenol, DBU, DMF, 70 °C; (<b>e</b>) NaOH, THF, H<sub>2</sub>O; (<b>f</b>) CDI, MeCN, 0 °C to r.t.; (<b>g</b>) <b>13a</b> (X=O, R=2-F), MeCN, 0 °C to r.t.</p>
Full article ">Scheme 3
<p>Synthesis of analogs. Reagents and Conditions: (<b>a</b>) 2-fluorophenol, K<sub>2</sub>CO<sub>3</sub>, CH<sub>3</sub>CN, µwave, 80 °C, 30 min; (<b>b</b>) H<sub>2</sub>, Pt(sulfided)/C, MeOH, 70 °C, H-Cube<sup>®</sup>; (<b>c</b>) 3-(4-chlorophenyl)propanoic acid, HATU, DIEA, DMF.</p>
Full article ">
10 pages, 2004 KiB  
Article
Hispidol Regulates Behavioral Responses to Ethanol through Modulation of BK Channels: A Novel Candidate for the Treatment of Alcohol Use Disorder
by Wooin Yang, Hee Jae Goh, Young Taek Han, Myon-Hee Lee and Dong Seok Cha
Molecules 2024, 29(19), 4531; https://doi.org/10.3390/molecules29194531 - 24 Sep 2024
Viewed by 760
Abstract
Alcohol use disorder (AUD) is the most common substance use disorder and poses a significant global health challenge. Despite pharmacological advances, no single drug effectively treats all AUD patients. This study explores the protective potential of hispidol, a 6,4′-dihydroxyaurone, for AUD using the [...] Read more.
Alcohol use disorder (AUD) is the most common substance use disorder and poses a significant global health challenge. Despite pharmacological advances, no single drug effectively treats all AUD patients. This study explores the protective potential of hispidol, a 6,4′-dihydroxyaurone, for AUD using the Caenorhabditis elegans model system. Our findings demonstrate that hispidol-fed worms exhibited more pronounced impairments in thrashes, locomotory speed, and bending amplitude, indicating that hispidol exacerbated the detrimental effects of acute ethanol exposure. However, hispidol significantly improved ethanol withdrawal behaviors, such as locomotory speed and chemotaxis performance. These beneficial effects were absent in slo-1 worms (the ortholog of mammalian α-subunit of BK channel) but were restored with the slo-1(+) or hslo(+) transgene, suggesting the involvement of BK channel activity. Additionally, hispidol increased fluorescence intensity and puncta in the motor neurons of slo-1::mCherry-tagged worms, indicating enhanced BK channel expression and clustering. Notably, hispidol did not alter internal ethanol concentrations, suggesting that its action is independent of ethanol metabolism. In the mouse models, hispidol treatment also demonstrated anxiolytic activity against ethanol withdrawal. Overall, these findings suggest hispidol as a promising candidate for targeting the BK channel in AUD treatment. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of hispidol on ethanol-induced acute behaviors in <span class="html-italic">C. elegans</span>. (<b>A</b>) Chemical structure of hispidol; Age-synchronized worms were exposed to 300 mM ethanol for 10 min, and ethanol-induced behavioral phenotypes, including (<b>B</b>) thrashing, (<b>C</b>) locomotion speed, and (<b>D</b>) body amplitude were observed. Data are presented as the mean ± S.D., with results obtained from three independent experiments. Statistical significance is indicated as follows: <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with naïve animals; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with ethanol-exposed animals.</p>
Full article ">Figure 2
<p>Effects of hispidol on ethanol withdrawal-induced behaviors in <span class="html-italic">C. elegans</span>. (<b>A</b>) Age-synchronized worms were exposed to 150 mM ethanol for 24 h and subsequently transferred to fresh plates for a 1-h withdrawal period; (<b>B</b>) Locomotion speed of withdrawn worms was assessed under a dissecting microscope; (<b>C</b>) The chemotactic ability of withdrawn worms toward the attractant OP50 was monitored every 15 min for 1 h. Statistical significance is indicated as follows: <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with naïve animals; ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 compared with ethanol-exposed worms.</p>
Full article ">Figure 3
<p>Involvement of BK channel modulation in hispidol-mediated alterations of ethanol-induced behaviors. (<b>A</b>) Age-synchronized <span class="html-italic">slo-1(js379)</span>, <span class="html-italic">slo-1(RNAi)</span>, <span class="html-italic">slo-1(+/js379)</span>, and <span class="html-italic">slo-1(hslo-1+/js379)</span> worms were exposed to 300 mM ethanol for 10 min, and their locomotion speed was measured; (<b>B</b>) Locomotion speed of <span class="html-italic">slo-1(js379)</span>, <span class="html-italic">slo-1(RNAi)</span>, <span class="html-italic">slo-1(+/js379)</span>, and <span class="html-italic">slo-1(hslo-1+/js379)</span> worms was assessed following ethanol withdrawal; (<b>C</b>) Internal ethanol concentration in wild-type worms was spectrophotometrically analyzed at three different time points: after 10 min of ethanol exposure (acute), after 24 h of ethanol exposure (chronic), and after an additional 1-h withdrawal following 24 h of ethanol exposure). Statistical significance is indicated as follows: *** <span class="html-italic">p</span> &lt; 0.001 compared with hispidol-untreated worms; n.s. not significant.</p>
Full article ">Figure 4
<p>Effects of hispidol on BK channel clustering. (<b>A</b>) Expression patterns of mCherry in JPS572 (vxEx345 [slo-1p::slo-1(+)::mCherry::unc-54 3′UTR + myo-2p::mCherry]) worms, captured at 100× magnification using fluorescence microscopy; (<b>B</b>) Quantification of slo-1 puncta, representing clustered BK channels in cholinergic neurons; (<b>C</b>) Carbofuran-induced paralysis monitored every 15 min in wild-type and <span class="html-italic">slo-1</span>(<span class="html-italic">js379</span>) worms; (<b>D</b>) Locomotion speed of <span class="html-italic">ctn-1(RNAi)</span> worms observed after 10 min of ethanol exposure and following withdrawal induction. Statistical significance is indicated as follows: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with hispidol-untreated worms; n.s., not significant.</p>
Full article ">Figure 5
<p>Effects of hispidol on ethanol-induced behaviors in mice. (<b>A</b>) Body weights of C57BL/6 mice monitored over seven days during a series of ethanol intoxications; (<b>B</b>) Motion trajectories recorded during the open field test; (<b>C</b>) Total distance traveled and (<b>D</b>) entries of center zone by mice in the open field test; (<b>E</b>) Motion trajectories recorded during the elevated plus maze test; (<b>F</b>) Time spent in the open arms in the elevated plus maze test. Statistical significance is indicated as follows: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with naïve animals; ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 compared with ethanol-exposed worms.</p>
Full article ">
22 pages, 1607 KiB  
Review
pH Homeodynamics and Male Fertility: A Coordinated Regulation of Acid-Based Balance during Sperm Journey to Fertilization
by Pengyuan Dai, Meng Zou, Ziyi Cai, Xuhui Zeng, Xiaoning Zhang and Min Liang
Biomolecules 2024, 14(6), 685; https://doi.org/10.3390/biom14060685 - 12 Jun 2024
Cited by 1 | Viewed by 2298
Abstract
pH homeostasis is crucial for spermatogenesis, sperm maturation, sperm physiological function, and fertilization in mammals. HCO3 and H+ are the most significant factors involved in regulating pH homeostasis in the male reproductive system. Multiple pH-regulating transporters and ion channels localize [...] Read more.
pH homeostasis is crucial for spermatogenesis, sperm maturation, sperm physiological function, and fertilization in mammals. HCO3 and H+ are the most significant factors involved in regulating pH homeostasis in the male reproductive system. Multiple pH-regulating transporters and ion channels localize in the testis, epididymis, and spermatozoa, such as HCO3 transporters (solute carrier family 4 and solute carrier family 26 transporters), carbonic anhydrases, and H+-transport channels and enzymes (e.g., Na+-H+ exchangers, monocarboxylate transporters, H+-ATPases, and voltage-gated proton channels). Hormone-mediated signals impose an influence on the production of some HCO3 or H+ transporters, such as NBCe1, SLC4A2, MCT4, etc. Additionally, ion channels including sperm-specific cationic channels for Ca2+ (CatSper) and K+ (SLO3) are directly or indirectly regulated by pH, exerting specific actions on spermatozoa. The slightly alkaline testicular pH is conducive to spermatogenesis, whereas the epididymis’s low HCO3 concentration and acidic lumen are favorable for sperm maturation and storage. Spermatozoa pH increases substantially after being fused with seminal fluid to enhance motility. In the female reproductive tract, sperm are subjected to increasing concentrations of HCO3 in the uterine and fallopian tube, causing a rise in the intracellular pH (pHi) of spermatozoa, leading to hyperpolarization of sperm plasma membranes, capacitation, hyperactivation, acrosome reaction, and ultimately fertilization. The physiological regulation initiated by SLC26A3, SLC26A8, NHA1, sNHE, and CFTR localized in sperm is proven for certain to be involved in male fertility. This review intends to present the key factors and characteristics of pHi regulation in the testes, efferent duct, epididymis, seminal fluid, and female reproductive tract, as well as the associated mechanisms during the sperm journey to fertilization, proposing insights into outstanding subjects and future research trends. Full article
Show Figures

Figure 1

Figure 1
<p>The graphical representation of pH variation in male and female reproductive systems. pH in the testis is determined at 7.2~7.4, which is conducive to maintaining spermatogenesis. From the initial segment to the cauda epididymis, the pH is approximately 6.5~6.8, where the sperm gradually matures and finally enters the vas deferens with a pH of around 7.2~7.4. After ejaculation, from the vagina to the cervix, the pH gradually alkalinizes, increasing from 4.3 to 6.5~7.5; pH 7.0~7.8 is found in the uterus, and in such an alkaline environment, sperm is capacitated and hyperactivated. pH 7.3~7.7 is maintained in the fallopian tube and contributes to AR in sperm to further fertilization.</p>
Full article ">Figure 2
<p>The schematic diagram presents the transport of H<sup>+</sup> or HCO<sub>3</sub><sup>−</sup> in the testis and epididymis. (<b>A</b>) In the testis, SLC4A2 primarily transports Cl<sup>−</sup> and HCO<sub>3</sub><sup>−</sup> in germ and SCs, and MCT1 transports H<sup>+</sup>. NDCBE redistributes Na<sup>+</sup>, Cl<sup>−</sup> and HCO<sub>3</sub><sup>−</sup>. NBCn1 and NBCe1 regulate Na<sup>+</sup> and HCO<sub>3</sub><sup>−</sup> transportation. (<b>B</b>) In the epididymis, HCO<sub>3</sub><sup>−</sup> reabsorption plays a vital role in pH determination. NHE3, CA III, CA IV, and CA XIV collaborate to manage HCO<sub>3</sub><sup>−</sup> redistribution. In addition, H-ATPase in the caput and cauda epididymis regulates pH by participating in H<sup>+</sup> secretion in the epididymal lumen.</p>
Full article ">Figure 3
<p>The various transporters and ion channels cooperatively regulate the pH in the sperm plasma membrane. In the acrosome of the sperm head, H<sup>+</sup>-ATPase is present to regulate H<sup>+</sup> secretion, and SLC4A1 is involved in HCO<sub>3</sub><sup>−</sup> transport. In equatorial and midpieces, SLC26A3, SLC26A8 and cystic fibrosis transmembrane conductance regulator (CFTR) are found to participate in Cl<sup>−</sup> and HCO<sub>3</sub><sup>−</sup> transport, in which SLC26A8 and CFTR are co-localized. In the principal piece of the flagellum, NHA1, NHA2, sNHE, and HVCN1 are involved in the regulation of H<sup>+</sup> expulsion, among which HVCN1 is only present in human, bovine and pig spermatozoa. In addition, two sperm-specific cation channels, CatSper and SLO3, are located in the principal piece and activated by pH alkalization, mediating external calcium inflow and K<sup>+</sup> expulsion, respectively. The co-localization of CatSper and HVCN1 is revealed.</p>
Full article ">
25 pages, 2945 KiB  
Review
Cytosolic and Acrosomal pH Regulation in Mammalian Sperm
by Julio C. Chávez, Gabriela Carrasquel-Martínez, Sandra Hernández-Garduño, Arturo Matamoros Volante, Claudia L. Treviño, Takuya Nishigaki and Alberto Darszon
Cells 2024, 13(10), 865; https://doi.org/10.3390/cells13100865 - 17 May 2024
Cited by 1 | Viewed by 1732
Abstract
As in most cells, intracellular pH regulation is fundamental for sperm physiology. Key sperm functions like swimming, maturation, and a unique exocytotic process, the acrosome reaction, necessary for gamete fusion, are deeply influenced by pH. Sperm pH regulation, both intracellularly and within organelles [...] Read more.
As in most cells, intracellular pH regulation is fundamental for sperm physiology. Key sperm functions like swimming, maturation, and a unique exocytotic process, the acrosome reaction, necessary for gamete fusion, are deeply influenced by pH. Sperm pH regulation, both intracellularly and within organelles such as the acrosome, requires a coordinated interplay of various transporters and channels, ensuring that this cell is primed for fertilization. Consistent with the pivotal importance of pH regulation in mammalian sperm physiology, several of its unique transporters are dependent on cytosolic pH. Examples include the Ca2+ channel CatSper and the K+ channel Slo3. The absence of these channels leads to male infertility. This review outlines the main transport elements involved in pH regulation, including cytosolic and acrosomal pH, that participate in these complex functions. We present a glimpse of how these transporters are regulated and how distinct sets of them are orchestrated to allow sperm to fertilize the egg. Much research is needed to begin to envision the complete set of players and the choreography of how cytosolic and organellar pH are regulated in each sperm function. Full article
(This article belongs to the Special Issue The Cell Biology of Fertilization)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the location of the proteins related to pHi regulation in mammalian sperm.</p>
Full article ">Figure 2
<p>Structure of the sperm-specific Na<sup>+</sup>/H<sup>+</sup> exchanger, sNHE (SLC9C). (<b>A</b>) depicts the domain arrangement of sNHE with color-coded features representing different domains: transporter domain composed of 13 TMs (13TM-TD, red), first cytosolic helices (H1, green), voltage sensor domain (VSD, blue), second cytosolic helices (H2, yellow), cyclic nucleotide-binding domain (CNBD, pink), and the C-terminal β strand domain (β, orange). (<b>B</b>) illustrates the 3D structure of an inactive state of sea urchin sNHE (<span class="html-italic">Sp</span>sNHE) dimer, determined by cryo-EM analysis (PDB ID: 8OTX, [<a href="#B76-cells-13-00865" class="html-bibr">76</a>]). To highlight the interphase of the dimer in the cytosolic helix domain, one monomer is colored according to Scheme A, while the other monomer is represented in gray using the PyMOL program. Hyperpolarization (HyperP) of the membrane potential is expected to induce a downward movement of the positively charged S4 segment of the VSD, rendering the sNHE in an active state. Additionally, the binding of cAMP to the CNBD might facilitate the conformational change toward the active state.</p>
Full article ">Figure 3
<p>pHi changes in response to a valinomycin-induced hyperpolarization in mouse and human sperm. Sperm pHi was assessed using the pH-sensitive dual-emission fluorescence probe, SNARF-5F (Excitation= 530 nm; Emission= 575/640 nm) in single cell recordings. Panel (<b>A</b>) illustrates pHi changes in the midpiece of mouse sperm induced by 1 µM valinomycin (Val) followed by a 20 mM NH<sub>4</sub>Cl control addition. WTK4.7 (red trace) represents wild-type mouse sperm in a normal medium containing 4.7 mM K<sup>+</sup>, while KOK4.7 (black trace) indicates sperm from sNHE (SLC9C1) null mice in a normal medium. WTK40 (blue trace) indicates wild-type sperm in a medium with 40 mM K<sup>+</sup>. Panel (<b>B</b>) depicts pHi changes in the human sperm flagellar midpiece. HTF (black trace) indicates the addition of medium, serving as a negative control against the addition of 1 µM valinomycin (Val, red trace). The results are adapted from [<a href="#B18-cells-13-00865" class="html-bibr">18</a>], with some modifications.</p>
Full article ">Figure 4
<p>Indicator-dependent pHi changes in response to HCO<sub>3</sub><sup>−</sup> perfusion in human sperm. HCO<sub>3</sub><sup>−</sup> causes a pHi alkalinization in human sperm, as reported by the SNARF-5F dye, but a slight pHi acidification when pHrodo red is used. Representative pHi recordings using SNARF-5F (<b>A</b>,<b>B</b>) and pHrodo red (<b>C</b>,<b>D</b>) perfusing 15 or 30 mM HCO<sub>3</sub><sup>−</sup> (green rectangle) in a 5% CO<sub>2</sub> environment. As positive controls, perfusions of 10 mM NH<sub>4</sub>Cl (orange rectangle) and 5 mM HCl (purple rectangle) are shown in each panel. Traces in each panel show average responses from 104 cells (for SNARF-5F) and 101 cells (for pHrodo red), with S.E.M. in gray. Ratiometric SNARF-5F measurements are reported as pHi values, whereas for pHrodo red, the F/F<sub>0</sub> normalization is shown, and F = fluorescence intensity. ↑F/F<sub>0</sub> indicates pHi acidification.</p>
Full article ">Figure 5
<p>Model of the molecular entities that regulate pHa in human sperm. Under non-capacitated conditions (NC), the pHa is acidic, due mainly to the active pumping of H<sup>+</sup> mediated by the V-ATPase into the acrosomal lumen and the flow of counterions through transport such as ClC-3. As capacitation initiates, HCO<sub>3</sub><sup>−</sup> enters the cell through different channels and transporters, and/or it is produced inside by the conversion of CO<sub>2</sub>, H<sub>2</sub>O, and H<sup>+</sup>. sAC is stimulated by HCO<sub>3</sub><sup>−</sup>, elevating cAMP levels and activating PKA, allowing the phosphorylation of several proteins, including CFTR channels, which also may allow the entry of HCO<sub>3</sub><sup>−</sup>. During capacitation, pHi also increases, favoring Ca<sup>2+</sup> influx, which also enhances sAC activity. V-ATPase allows the acrosome to remain acidic during the first hours of capacitation. The continuous entry of HCO<sub>3</sub><sup>−</sup>, as well as the exit of H<sup>+</sup> from the cytosol, through the Hv1 channel in the case of human sperm or through NHEs in other mammals, stabilizes the cytosolic alkalinization, dissipates the H<sup>+</sup> gradient, decreases V-ATPase activity, and induces the alkalinization of the acrosome. Other mechanisms, not yet described, could also regulate the activity of the V-ATPase. The pHa increase destabilizes the acrosomal matrix, producing acrosome swelling and probably TPC1 channel activation, releasing acrosomal Ca<sup>2+,</sup> which in turn stimulates extracellular Ca<sup>2+</sup> uptake through ORAI channels (1 and 2). Both acrosome alkalinization and [Ca<sup>2+</sup>]i increases induce AR. Arrow indicates increase of the ion (↑, ↑↑). We place the sign (?) to highlight that some transporters or channels, although they have been detected, their exact location and identity has not been fully established (NBC), or their function in humans is unknown (TPC1, NHE11).</p>
Full article ">
15 pages, 2198 KiB  
Review
SLO3: A Conserved Regulator of Sperm Membrane Potential
by Maximilian D. Lyon, Juan J. Ferreira, Ping Li, Shweta Bhagwat, Alice Butler, Kelsey Anderson, Maria Polo and Celia M. Santi
Int. J. Mol. Sci. 2023, 24(13), 11205; https://doi.org/10.3390/ijms241311205 - 7 Jul 2023
Cited by 5 | Viewed by 3258
Abstract
Sperm cells must undergo a complex maturation process after ejaculation to be able to fertilize an egg. One component of this maturation is hyperpolarization of the membrane potential to a more negative value. The ion channel responsible for this hyperpolarization, SLO3, was first [...] Read more.
Sperm cells must undergo a complex maturation process after ejaculation to be able to fertilize an egg. One component of this maturation is hyperpolarization of the membrane potential to a more negative value. The ion channel responsible for this hyperpolarization, SLO3, was first cloned in 1998, and since then much progress has been made to determine how the channel is regulated and how its function intertwines with various signaling pathways involved in sperm maturation. Although Slo3 was originally thought to be present only in the sperm of mammals, recent evidence suggests that a primordial form of the gene is more widely expressed in some fish species. Slo3, like many reproductive genes, is rapidly evolving with low conservation between closely related species and different regulatory and pharmacological profiles. Despite these differences, SLO3 appears to have a conserved role in regulating sperm membrane potential and driving large changes in response to stimuli. The effect of this hyperpolarization of the membrane potential may vary among mammalian species just as the regulation of the channel does. Recent discoveries have elucidated the role of SLO3 in these processes in human sperm and provided tools to target the channel to affect human fertility. Full article
(This article belongs to the Special Issue Recent Advances in the Physiology of Ion Channels in Sperm Cells)
Show Figures

Figure 1

Figure 1
<p>Amino acid sequence homology of mouse (mSLO3), human (hSLO3), and bovine (bSLO3) SLO3. Conserved regions are highlighted in blue. Dark highlighting indicates conservation between three species, light highlighting indicates conservation between two species. Sequence alignment performed using Jalview Version 2 [<a href="#B47-ijms-24-11205" class="html-bibr">47</a>,<a href="#B48-ijms-24-11205" class="html-bibr">48</a>,<a href="#B49-ijms-24-11205" class="html-bibr">49</a>,<a href="#B50-ijms-24-11205" class="html-bibr">50</a>].</p>
Full article ">Figure 2
<p>Human SLO3 gating ring structure determined by X-ray crystallography. (<b>a</b>) Cartoon of domain topology of two opposing SLO3 α-subunits. (<b>b</b>) Crystal structure of the gating ring of a hSLO3 tetramer with RCK1 and RCK2 domains colored in blue and red, respectively. (<b>c</b>) A single subunit of the hSLO3 channel and (<b>d</b>) highlight of RCK1. (<b>e</b>) A closeup of the hSLO3 assembly interface and (<b>f</b>) the corresponding region of SLO1 bound to Ca<sup>2+</sup>. The RCK1 N-terminal residue that connects to the transmembrane pore is shown as a green sphere. Ca<sup>2+</sup> ion is shown as a yellow sphere. Reprinted/adapted with permission from [<a href="#B69-ijms-24-11205" class="html-bibr">69</a>].</p>
Full article ">Figure 3
<p>Models of mouse and human SLO3 activity. (<b>a</b>) Mouse: The exposure to a more alkaline pH and high [HCO<sub>3</sub><sup>−</sup>] concentrations in the female tract contribute to an increase in pH<sub>i</sub>, potentially through the activation of the sNHE. This rise in pH<sub>i</sub> leads to the activation of SLO3 channels, resulting in membrane hyperpolarization. This hyperpolarization enhances calcium influx through CatSper channels, possibly through two distinct mechanisms: Firstly, by increasing the inward driving force of calcium. Secondly, it may further activate sNHE to elevate intracellular pH even more. (<b>b</b>) Human: In human sperm, exposure to an elevated external pH could potentially activate the Hv1 channel, resulting in an increase in pH<sub>i</sub> and contributing to the activation of SLO3 and CatSper channels. However, it is important to note that in humans, SLO3 channels are primarily activated by calcium, while CatSper channels are activated by progesterone. On the other hand, activation of SLO3 leads to membrane hyperpolarization, which has been proposed to remove [Ca<sup>2+</sup>]<sub>i</sub> oscillations that inhibit CatSper activation. This raises the question of whether SLO3 is activated upstream or downstream of CatSper channels.</p>
Full article ">
13 pages, 5708 KiB  
Article
Effect of 2-Aminoethoxydiphenyl Borate (2-APB) on Heart Rate and Relation with Suppressed Calcium-Activated Potassium Channels: Larval Drosophila Model
by Nicole Hensley, Elizabeth R. Elliott, Maya O. Abul-Khoudoud and Robin L. Cooper
Appl. Biosci. 2023, 2(2), 236-248; https://doi.org/10.3390/applbiosci2020017 - 23 May 2023
Cited by 1 | Viewed by 2275
Abstract
Cardiac contractile cells depend on calcium in order to function. Understanding the regulation of calcium influx, efflux, and release from the sarcoplasmic reticulum is essential. The focus of this investigation is to address how a reduction of functional Ca2+-activated K+ [...] Read more.
Cardiac contractile cells depend on calcium in order to function. Understanding the regulation of calcium influx, efflux, and release from the sarcoplasmic reticulum is essential. The focus of this investigation is to address how a reduction of functional Ca2+-activated K+ (KCa) channels, via a mutational line, might impact the heart rate in larva when the SER is also modulated through Ca2+ loading and stimulation. The larval heart tube is exposed in situ and flushed with saline. With a known saline composition, a potential therapeutic pharmacological agent, 2-Aminoethyl diphenylborinate (2-APB), was examined for its effect on heart rate, as well as to determine the contribution from KCa channels. In this study, it was determined that mutation in the K(Ca) channel (i.e., Slo) showed a different trend than the wild-type CS strain. Exposure to high concentrations of 50 µM 2-APB decreased heart rate in the Slo strain and increased it in the wild-type CS strain. Serotonin increased heart rate in both thapsigargin- and 2-APB-treated larvae, with no significant difference between the strains. Full article
Show Figures

Figure 1

Figure 1
<p>Effect on heart rate with exposure to 2-APB for CS and <italic>Slo</italic> (K(Ca) mutant) strains. (<bold>A1</bold>,<bold>A2</bold>) Controls for exchanging the saline bath. Saline was exchanged for saline and again for fresh saline. (<bold>B1</bold>,<bold>B2</bold>) The effect of 2-APB (10 µM) on heart rate after 1 min of incubation followed by saline wash. (<bold>C1</bold>,<bold>C2</bold>) The effect of 2-APB (30 µM) on heart rate after 1 min of incubation followed by saline wash. (<bold>D1</bold>,<bold>D2</bold>) The effect of 2-APB (50 µM) on heart rate after 1 min of incubation followed by saline wash. The left column represents responses for the CS strain and the right column for the <italic>Slo</italic> (K(Ca) mutant) strain. There is a significant decrease in heart rate to exposure of 2-APB at 10, 30, and 50 µM for the <italic>Slo</italic> strain (<bold>B2</bold>,<bold>C2</bold>), (N = 10, N = 12 for 50 µM, <italic>p</italic> &gt; 0.05, Paired <italic>T</italic>-test; the CS strain presented with a significant effect for 10 µM and 50 µM, <italic>T</italic>-Test was used <italic>p</italic> &gt; 0.05). A percent change in the individuals from saline to 2-APB which normalized the differences in the initial values of the rates.</p>
Full article ">Figure 2
<p>The effect of exposure to 5-HT on heart rate. (<bold>A1</bold>,<bold>A2</bold>). The CS larvae and the (<bold>B1</bold>,<bold>B2</bold>) <italic>Slo</italic> both increased in heart rate immediately upon exposure to 5-HT and the rates remained elevated for over a minute (N = 10; <italic>p</italic> &gt; 0.05, Paired <italic>T</italic>-test). Upon exchanging the media back to fresh saline, the rates remained elevated. The * (asterisk) indicates a significant difference between saline and 5-HT.</p>
Full article ">Figure 3
<p>The effect of exposure to 5-HT on heartbeat rate after incubation with 2-APB. (<bold>A1</bold>) The CS larvae and the (<bold>B1</bold>) <italic>Slo</italic> both decreased the rate of the heartbeat with 2-APB after 7 min and increased in heartbeat rate immediately upon exposure to 5-HT and the rates remained elevated for over 2 min (N = 24 for each line; <italic>p</italic> &lt; 0.05, ANOVA and a post-hoc Bonferroni <italic>T</italic>-test). (<bold>A2</bold>) Represents the mean (+/− SEM) of values shown in A1. (<bold>B2</bold>) Represents the and mean (+/− SEM) of values shown in B1. The * (asterisk) indicates a significant difference between groups.</p>
Full article ">Figure 4
<p>The effect of exposure to 5-HT on heart rate after incubation with thapsigargin for 20 min. (<bold>A1</bold>,<bold>A2</bold>) Two batches of thapsigargin were examined. One batch (<bold>A1</bold>) was stored for over 6 months as a lyophilized solid at −20 °C; the other was a newer batch used within a week (<bold>A2</bold>). There were no significant differences in the responses between the two groups (N = 12 for each group, ANOVA). The rate increased in heart rate immediately upon exposure to thapsigargin and 5-HT as compared to 20 min of thapsigargin exposure (Paired <italic>T</italic>-test <italic>p</italic> &lt; 0.05, N = 24), and remined higher for the next 7 min. (<bold>B</bold>) The K(Ca) mutational line (<italic>Slo</italic>) increased in heart rate immediately upon exposure to thapsigargin and 5-HT as compared to 20 min of thapsigargin exposure (Paired <italic>T</italic>-test <italic>p</italic> &lt; 0.05, N = 12), and remined higher for the next 7 min. (<bold>C</bold>) A control for exchanging the bathing media altering the heart rate was performed for the K(Ca) mutational line (<italic>Slo</italic>) without any significant changes for 20 min of incubation. The * (asterisk) indicates a significant difference between thapsigargin after 20 min to the immediate application of 5-HT. Traces on the right are means (+/− SEM) of the traces on the left.</p>
Full article ">Figure 5
<p>Schematic model in the effects on heart rate in larvae when altering intracellular Ca<sup>2+</sup> by 2-APB in wild-type and <italic>slo</italic> strains of larval <italic>Drosophila</italic>.</p>
Full article ">
19 pages, 2901 KiB  
Article
Pharmacological Evidence Suggests That Slo3 Channel Is the Principal K+ Channel in Boar Spermatozoa
by Akila Cooray, Jeongsook Kim, Beno Ramesh Nirujan, Nishani Jayanika Jayathilake and Kyu Pil Lee
Int. J. Mol. Sci. 2023, 24(9), 7806; https://doi.org/10.3390/ijms24097806 - 25 Apr 2023
Viewed by 1952
Abstract
Sperm ion channels are associated with the quality and type of flagellar movement, and their differential regulation is crucial for sperm function during specific phases. The principal potassium ion channel is responsible for the majority of K+ ion flux, resulting in membrane [...] Read more.
Sperm ion channels are associated with the quality and type of flagellar movement, and their differential regulation is crucial for sperm function during specific phases. The principal potassium ion channel is responsible for the majority of K+ ion flux, resulting in membrane hyperpolarization, and is essential for sperm capacitation-related signaling pathways. The molecular identity of the principal K+ channel varies greatly between different species, and there is a lack of information about boar K+ channels. We aimed to determine the channel identity of boar sperm contributing to the primary K+ current using pharmacological dissection. A series of Slo1 and Slo3 channel modulators were used for treatment. Sperm motility and related kinematic parameters were monitored using a computer-assisted sperm analysis system under non-capacitated conditions. Time-lapse flow cytometry with fluorochromes was used to measure changes in different intracellular ionic concentrations, and conventional flow cytometry was used to determine the acrosome reaction. Membrane depolarization, reduction in acrosome reaction, and motility parameters were observed upon the inhibition of the Slo3 channel, suggesting that the Slo3 gene encodes the main K+ channel in boar spermatozoa. The Slo3 channel was localized on the sperm flagellum, and the inhibition of Slo3 did not reduce sperm viability. These results may aid potential animal-model-based extrapolations and help to ameliorate motility and related parameters, leading to improved assisted reproductive methods in industrial livestock production. Full article
(This article belongs to the Special Issue Recent Advances in the Physiology of Ion Channels in Sperm Cells)
Show Figures

Figure 1

Figure 1
<p>LDD175 treatment increased sperm motility parameters, while sperm treated with NS1619 exhibited a decreasing pattern. The left, middle, and right columns represent the progressive motility, rapid sperm%, and immotile sperm count, respectively. Row (<b>A</b>) depicts sperm subjected to the LDD175 treatment, while row (<b>B</b>) depicts sperm subjected to the NS1619 treatment. All samples were incubated for 5 min after treatment. Significant differences (<span class="html-italic">p</span> &lt; 0.05) compared with the negative control with no treatment are marked with asterisks (n = 10, average ± S.D.).</p>
Full article ">Figure 2
<p>The membrane potential of sperm is differently modulated by LDD175 and NS1619. Panel A depicts the relative changes in membrane potential signals upon treatment with LDD175 (■) and NS1619 (●) compared with the DMSO vehicle control (◆) (<b>A</b>). The bar graph (<b>B</b>) compares the relative signals at each concentration after the fluorescence reaches a stable point. (<b>C</b>–<b>F</b>) represent changes in membrane potential corresponding to DiSC3(5) fluorescence upon the sequential treatment of LDD175 with IbTx (<b>C</b>,<b>D</b>) and PAX (<b>E</b>,<b>F</b>). Data for each concentration contain recordings for 2 min and are presented as the average ± S.D. (n = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). Changes in membrane potential with Slo1-specific blockers (PAX and IbTx) are shown in the <a href="#app1-ijms-24-07806" class="html-app">Supplementary Data (Supplementary Figure S1)</a>.</p>
Full article ">Figure 3
<p>Both the LDD175 and NS1619 treatments increase the cytosolic calcium of boar sperm in an extracellular−calcium−dependent way. The upper row shows calcium signal changes in 1.8 mM of calcium medium, whereas the bottom row shows the corresponding signals in the Ca<sup>2+</sup>−free medium. Signals in LDD175− (<b>A</b>) and NS1619− (<b>B</b>) treated sperm are dark gray and blue, respectively. Data are presented as the average ± S.D. The bar graph (<b>C</b>) compares the maximum signals in 1.8 mM of Ca<sup>2+</sup>−medium (5 μM of LDD175, 50 μM of NS1619) and its counterparts corresponding to the same concentrations in Ca<sup>2+</sup>-free medium (n = 3, * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>CatSper and K<sup>+</sup> channel blockers reduce intracellular calcium signals. The top row shows (<b>A</b>,<b>B</b>) changes in Fluo4−AM signals after treatment with 1 μM of NNC55−0396 preceded by K<sup>+</sup> channel modulators. Signal changes after treatment with 60 mM of TEA followed by the respective K<sup>+</sup> channel modulators are illustrated in the bottom row (<b>C</b>,<b>D</b>). Blue traces represent [Ca<sup>2+</sup>]<sub>i</sub> changes after LDD175 treatment, while red traces represent changes after NS1619 treatment. The data for each LDD175 and NS1619 treatment had an acquisition duration of 2 min (n = 3).</p>
Full article ">Figure 5
<p>LDD175 and NS1619 acidify boar spermatozoa. This figure shows the changes in intracellular pH after treatment with LDD175 and NS1619. (<b>A</b>) shows the traces of normalized fluorescence during the data acquisition, with light red indicating the signals of the NS1619-treated group and black indicating the signals of the LDD175−treated group. The fluorescence of each concentration of LDD175 and NS1619 is compared in the bar graph (<b>B</b>). The effects of IbTx and PAX on sperm pH are shown in Figure (<b>C</b>). Data are presented as the average ± S.D. (n = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Activation of Slo3 increases the number of acrosome-reacted boar sperm. (<b>A</b>–<b>E</b>) show representative quadrant plots of the negative control (<b>A</b>), LDD175 ((<b>B</b>): 5 µM; (<b>C</b>): 10 µM), and NS1619 ((<b>D</b>): 50 µM; (<b>E</b>): 100 µM). The bar graph shows (<b>F</b>) the percentage of acrosome-reacted sperm cells in each treatment condition, where white, dark gray, and light gray boxes represent the negative control, the LDD175 treatment, and the NS1619 treatment, respectively. Data are presented as the average ± S.D. (n = 5, * <span class="html-italic">p</span> &lt; 0.05). Single staining conditions containing only PNA (<a href="#app1-ijms-24-07806" class="html-app">Supplementary Figure S2A</a>) and PI (<a href="#app1-ijms-24-07806" class="html-app">Supplementary Figure S2B</a>) that were used to obtain the necessary gating and the positive control containing 10 µM ionomycin (<a href="#app1-ijms-24-07806" class="html-app">Supplementary Figure S2C</a>) are provided in the <a href="#app1-ijms-24-07806" class="html-app">Supplementary Figures</a>.</p>
Full article ">Figure 7
<p>Localization of Slo channels in immunocytochemistry and boar sperm viability. (<b>A</b>) Anti-Slo1 antibody; (<b>B</b>) anti-Slo3 antibody; (<b>C</b>) control experiment using only the secondary antibody. White scale bar (<b>A</b>–<b>C</b>) represents 5 µm. (<b>D</b>) Percentage of intact live cells obtained from an apoptosis assay (n = 5, average ± S.D; * <span class="html-italic">p</span> &lt;0.05).</p>
Full article ">
24 pages, 7906 KiB  
Article
Interaction of Some Asymmetrical Porphyrins with U937 Cell Membranes–In Vitro and In Silico Studies
by Dragos Paul Mihai, Rica Boscencu, Gina Manda, Andreea Mihaela Burloiu, Georgiana Vasiliu, Ionela Victoria Neagoe, Radu Petre Socoteanu and Dumitru Lupuliasa
Molecules 2023, 28(4), 1640; https://doi.org/10.3390/molecules28041640 - 8 Feb 2023
Cited by 4 | Viewed by 1911
Abstract
The aim of the present study was to assess the effects exerted in vitro by three asymmetrical porphyrins (5-(2-hydroxyphenyl)-10,15,20-tris-(4-acetoxy-3-methoxyphenyl)porphyrin, 5-(2-hydroxyphenyl)-10,15,20-tris-(4-acetoxy-3-methoxyphenyl)porphyrinatozinc(II), and 5-(2-hydroxyphenyl)-10,15,20–tris-(4-acetoxy-3-methoxyphenyl)porphyrinatocopper(II)) on the transmembrane potential and the membrane anisotropy of U937 cell lines, using bis-(1,3-dibutylbarbituric acid)trimethine oxonol (DiBAC4(3)) and 1-(4-trimethylammoniumphenyl)-6-phenyl-1,3,5-hexatriene p-toluenesulfonate (TMA-DPH), [...] Read more.
The aim of the present study was to assess the effects exerted in vitro by three asymmetrical porphyrins (5-(2-hydroxyphenyl)-10,15,20-tris-(4-acetoxy-3-methoxyphenyl)porphyrin, 5-(2-hydroxyphenyl)-10,15,20-tris-(4-acetoxy-3-methoxyphenyl)porphyrinatozinc(II), and 5-(2-hydroxyphenyl)-10,15,20–tris-(4-acetoxy-3-methoxyphenyl)porphyrinatocopper(II)) on the transmembrane potential and the membrane anisotropy of U937 cell lines, using bis-(1,3-dibutylbarbituric acid)trimethine oxonol (DiBAC4(3)) and 1-(4-trimethylammoniumphenyl)-6-phenyl-1,3,5-hexatriene p-toluenesulfonate (TMA-DPH), respectively, as fluorescent probes for fluorescence spectrophotometry. The results indicate the hyperpolarizing effect of porphyrins in the concentration range of 0.5, 5, and 50 μM on the membrane of human U937 monocytic cells. Moreover, the tested porphyrins were shown to increase membrane anisotropy. Altogether, the results evidence the interaction of asymmetrical porphyrins with the membrane of U937 cells, with potential consequences on cellular homeostasis. Molecular docking simulations, and Molecular mechanics Poisson–Boltzmann surface area (MM/PBSA) free energy of binding calculations, supported the hypothesis that the investigated porphyrinic compounds could potentially bind to membrane proteins, with a critical role in regulating the transmembrane potential. Thus, both the free base porphyrins and the metalloporphyrins could bind to the SERCA2b (sarco/endoplasmic reticulum ATPase isoform 2b) calcium pump, while the metal complexes may specifically interact and modulate calcium-dependent (large conductance calcium-activated potassium channel, Slo1/KCa1.1), and ATP-sensitive (KATP), potassium channels. Further studies are required to investigate these interactions and their impact on cellular homeostasis and functionality. Full article
(This article belongs to the Special Issue Porphyrin-Based Compounds: Synthesis and Application)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The molecular structures of the investigated porphyrins. (<b>a</b>) 5-(2-hydroxyphenyl)-10,15,20-tris-(4-acetoxy-3-methoxyphenyl) porphyrin, (<b>b</b>) 5-(2-hydroxyphenyl)-10,15,20-tris-(4-acetoxy-3-methoxyphenyl)porphyrinatozinc(II), 5-(2-hydroxyphenyl)-10,15,20–tris-(4-acetoxy-3-methoxyphenyl)porphyrinatocopper(II).</p>
Full article ">Figure 2
<p>Representative effects of porphyrinic compounds on the transmembrane potential of U937 cells. (<b>a</b>)—Fluorescence spectra of U937 cells, with and without the potential-sensitive probe DiBAC<sub>4</sub>(3), when Zn(II)TMAPOHo was tested (<span class="html-fig-inline" id="molecules-28-01640-i001"><img alt="Molecules 28 01640 i001" src="/molecules/molecules-28-01640/article_deploy/html/images/molecules-28-01640-i001.png"/></span> control without probe, <span class="html-fig-inline" id="molecules-28-01640-i002"><img alt="Molecules 28 01640 i002" src="/molecules/molecules-28-01640/article_deploy/html/images/molecules-28-01640-i002.png"/></span> control with probe, <span class="html-fig-inline" id="molecules-28-01640-i003"><img alt="Molecules 28 01640 i003" src="/molecules/molecules-28-01640/article_deploy/html/images/molecules-28-01640-i003.png"/></span> 50 μM ZnTMAPOHo without probe, <span class="html-fig-inline" id="molecules-28-01640-i004"><img alt="Molecules 28 01640 i004" src="/molecules/molecules-28-01640/article_deploy/html/images/molecules-28-01640-i004.png"/></span> 50 μM ZnTMAPOHo with probe); (<b>b</b>)—Hyperpolarizing effect for TMAPOHo; (<b>c</b>)—Hyperpolarizing effect for Zn(II)TMAPOHo; (<b>d</b>)—Hyperpolarizing effect for Cu(II)TMAPOHo; (<b>e</b>)—Comparative effect of porphyrinic compounds (24 h incubation time and 0.5 µM, 5 µM, 50 µM concentrations) on the transmembrane potential of U937 cells. In <a href="#molecules-28-01640-f002" class="html-fig">Figure 2</a>b,c,d results are presented as mean hyperpolarizing effect (%) ± standard deviation (SD) for triplicate samples.</p>
Full article ">Figure 3
<p>Investigation of the effects of porphyrinic compounds on membrane anisotropy. (<b>a</b>)—Intensity of signals obtained in Time Drive mode for U937 cells labelled with TMA-DPH vs. unlabeled cells; (<b>b</b>)—Intensity of response signals obtained in Time Drive mode for TMA-DPH-labelled cells; (<b>c</b>)—The effect of porphyrinic compounds (24 h incubation time and 0.5 µM, 5 µM, 50 µM concentrations) on the membrane anisotropy of U937 cells. Data of a representative experiments are presented.</p>
Full article ">Figure 4
<p>Variation of free energy of binding in relation to the distance from the membrane center (Z).</p>
Full article ">Figure 5
<p>(<b>a</b>)—Superposition of SERCA2b model generated with YASARA (blue) on experimental (green) structure (PDB ID: 6LN9); (<b>b</b>)—superposition of Slo1 model (monomer) generated with SWISS-MODEL (blue) on experimental (green) structure (PDB ID: 6V35); (<b>c</b>)—superposition of SUR2 model generated with AlphaFold (blue) on experimental (green) structure (PDB ID: 7MIT); (<b>d</b>)—Ramachandran plot for SERCA2b model; (<b>e</b>)—Ramachandran plot for Slo1 model; (<b>f</b>)—Ramachandran plot for SUR2 model.</p>
Full article ">Figure 6
<p>Predicted binding modes between positive controls and full-length structures of SERCA2b, Slo1, and the SUR2 subunit of K<sub>ATP</sub>. (<b>a</b>)—predicted conformation of BHQ-SERCA2b complex; (<b>b</b>)—interaction diagram for BHQ-SERCA2b complex; (<b>c</b>)—predicted conformation of heme-Slo1 complex; (<b>d</b>)—interaction diagram for heme-Slo1 complex; (<b>e</b>)—predicted conformation of heme-SUR2 complex; (<b>f</b>)—interaction diagram for heme-SUR2 complex. Metal coordination bonds are represented as purple dashes.</p>
Full article ">Figure 7
<p>Predicted binding modes between porphyrin derivatives and SERCA2b. (<b>a</b>)—predicted conformation of TMAPOHo-SERCA2b complex; (<b>b</b>)—interaction diagram for TMAPOHo-SERCA2b complex; (<b>c</b>)—predicted conformation of Zn(II)TMAPOHo-SERCA2b complex; (<b>d</b>)—interaction diagram for Zn(II)TMAPOHo-SERCA2b complex; (<b>e</b>)—predicted conformation of Cu(II)TMAPOHo-SERCA2b complex; (<b>f</b>)—interaction diagram for Cu(II)TMAPOHo-SERCA2b complex. Metal coordination bonds are represented as purple dashes.</p>
Full article ">Figure 8
<p>Predicted binding modes between porphyrin derivatives and Slo1. (<b>a</b>)—predicted conformation of TMAPOHo-Slo1 complex; (<b>b</b>)—interaction diagram for TMAPOHo-Slo1 complex; (<b>c</b>)—predicted conformation of Zn(II)TMAPOHo-Slo1 complex; (<b>d</b>)—interaction diagram for Zn(II)TMAPOHo-Slo1 complex; (<b>e</b>)—predicted conformation of Cu(II)TMAPOHo-Slo1 complex; (<b>f</b>)—interaction diagram for Cu(II)TMAPOHo-Slo1 complex. Metal coordination bonds are represented as purple dashes.</p>
Full article ">Figure 9
<p>Predicted binding modes between porphyrin derivatives and the SUR2 subunit of K<sub>ATP</sub>. (<b>a</b>)—predicted conformation of TMAPOHo-SUR2 complex; (<b>b</b>)—interaction diagram for TMAPOHo-SUR2 complex; (<b>c</b>)—predicted conformation of Zn(II)TMAPOHo-SUR2 complex; (<b>d</b>)—interaction diagram for Zn(II)TMAPOHo-SUR2 complex; (<b>e</b>)—predicted conformation of Cu(II)TMAPOHo-SUR2 complex; (<b>f</b>)—interaction diagram for Cu(II)TMAPOHo-SUR2 complex. Metal coordination bonds are represented as purple dashes.</p>
Full article ">Figure 10
<p>Electrostatic potential maps of the binding sites in complex with Zn(II)TMAPOHo. (<b>a</b>) – binding site of SERCA2b; (<b>b</b>) – binding site of Slo1; (<b>c</b>) – binding site of SUR2.</p>
Full article ">Figure 11
<p>(<b>a</b>)—Superposition of simulated conformation (green) of Cu(II)TMAPOHo-SERCA2b complex on original structure (blue); (<b>b</b>)—superposition of simulated conformation (green) of Cu(II)TMAPOHo-Slo1 complex on original structure (blue); (<b>c</b>)—superposition of simulated conformation (green) of Cu(II)TMAPOHo-SUR2 complex on original structure (blue); (<b>d</b>)—exponential relationship between free binding energy values and tested porphyrin concentrations; (<b>e</b>)—correlation diagram between free binding energies and negative logarithmic values of tested porphyrin concentrations. Solvent molecules are hidden for clarity.</p>
Full article ">Figure 12
<p>RMSF values per amino acid residue after short MD simulations. (<b>a</b>) – SERCA2b; (<b>b</b>) – Slo1; (<b>c</b>) – SUR2.</p>
Full article ">Figure 13
<p>Illustration of hypothesized molecular mechanisms related to the interaction of the investigated porphyrinic compounds with cellular membranes.</p>
Full article ">
13 pages, 2308 KiB  
Article
Structure of the Human BK Ion Channel in Lipid Environment
by Lige Tonggu and Liguo Wang
Membranes 2022, 12(8), 758; https://doi.org/10.3390/membranes12080758 - 31 Jul 2022
Cited by 10 | Viewed by 3247
Abstract
Voltage-gated and ligand-modulated ion channels play critical roles in excitable cells. To understand the interplay among voltage sensing, ligand binding, and channel opening, the structures of ion channels in various functional states and in lipid membrane environments need to be determined. Here, the [...] Read more.
Voltage-gated and ligand-modulated ion channels play critical roles in excitable cells. To understand the interplay among voltage sensing, ligand binding, and channel opening, the structures of ion channels in various functional states and in lipid membrane environments need to be determined. Here, the random spherically constrained (RSC) single-particle cryo-EM method was employed to study human large conductance voltage- and calcium-activated potassium (hBK or hSlo1) channels reconstituted into liposomes. The hBK structure was determined at 3.5 Å resolution in the absence of Ca2+. Instead of the common fourfold symmetry observed in ligand-modulated ion channels, a twofold symmetry was observed in hBK in liposomes. Compared with the structure of isolated hSlo1 Ca2+ sensing gating rings, two opposing subunits in hBK unfurled, resulting in a wider opening towards the transmembrane region of hBK. In the pore gate domain, two opposing subunits also moved downwards relative to the two other subunits. Full article
(This article belongs to the Special Issue Ion Channel in Lipid Environment)
Show Figures

Figure 1

Figure 1
<p>Subtraction of modeled liposome images from an hBK proteoliposome micrograph. (<b>A</b>) A representative cryo-EM image of hBK proteoliposomes at −3.8 μm defocus; (<b>B</b>) liposomes are subtracted from (<b>A</b>). hBK particles are marked with red boxes (15 nm); (<b>C</b>,<b>D</b>) 2D class averages of BK particles (<b>C</b>) before and (<b>D</b>) after liposome subtraction. Box size was 27 nm, and the circular mask was 17 nm in diameter.</p>
Full article ">Figure 2
<p>Flowchart for data processing.</p>
Full article ">Figure 3
<p>The structure of hBK in liposomes. (<b>A</b>) hBK cryo-EM density map. Lipid membrane is shown in gray mesh. For clarity, only half of the lipid membrane is shown; (<b>B</b>) view of the gating ring in hBK from the extracellular side. The diagonal distances between the C<sub>α</sub> atoms of Val 785 are indicated; (<b>C</b>) superposition of hBK high (orange red) and low (blue) subunits. The shoulder helix J is indicated; (<b>D</b>) stereo view of the superimposed hBK high and low subunits; (<b>E</b>–<b>G</b>) side views of hBK high € (red), hBK low (<b>F</b>) (dark blue), and hSlo1-GR (<b>G</b>) (light blue, PDB: 3NAF). To distinguish RCK1 and RCK2 domains, the RCK1 domains in all three models were colored green. The RCK2 domains in hBK high subunits are aligned to those in hSlo1-GR; (<b>H</b>) cartoon to show the relative rotation among hBK high and low, and hSlo1-GR.</p>
Full article ">Figure 4
<p>Stereo views of assembly and flexible interfaces in hBK. (<b>A</b>) Flexible interfaces in hBK. hBK high (orange red) and hBK low (dark blue) subunits are superimposed. To show the flexible interface, RCK2 in hBK high subunit is colored green; (<b>B</b>) Same as (<b>A</b>) except the replacement of hBK low with hSlo1-GR (light blue, PDB: 3NAF) subunit; (<b>C</b>) Assembly interface in hBK as defined in <a href="#membranes-12-00758-f003" class="html-fig">Figure 3</a>B. hBK high (orange red) and low (blue) subunits are superimposed. The interacting residues are annotated; (<b>D</b>) Same as (<b>C</b>) except the replacement of hBK low with hSlo1-GR (light blue, PDB: 3NAF) subunit. The αD and αE helices in RCK2 of hSlo1-GR were aligned to those of hBK.</p>
Full article ">Figure 5
<p>Stereo views of Ca<sup>2+</sup>-binding sites in hBK. (<b>A</b>) Ca<sup>2+</sup> bowl in metal-free hBK (orange-red and blue) and metal-free hSlo1-GR (light blue, PDB: 3NAF); (<b>B</b>) same as (<b>A</b>), except metal-free hSlo1-GR was replaced with metal-free aSlo1 (green, PDB: 5TJI); (<b>C</b>) RCK1 Ca<sup>2+</sup>-binding site in metal-free hBK and metal-free hSlo1-GR (light blue, PDB: 3NAF); (<b>D</b>) same as (<b>C</b>), except metal-free hSlo1-GR was replaced with metal-free aSlo1 (green, PDB: 5TJI). Ca<sup>2+</sup> coordinating residues were labeled, and the position of Ca<sup>2+</sup> ion in liganded aSlo1 is represented by a green sphere.</p>
Full article ">Figure 6
<p>Comparison of the TM region in hBK in liposomes in the absence of Ca<sup>2+</sup> and metal-free aSlo1 (PDB: 5TJI). (<b>A</b>) TM region in hBK (orange red) rotated by 12 degrees clockwise with respect to the main 3D class from the aSlo1 dataset (cyan); (<b>B</b>) aSlo1 rotated by 12 degrees to overlay with hBK. All views are from the extracellular side; (<b>C</b>) superposition of the intracellular helix bundle crossing region of hBK and aSlo1; (<b>D</b>,<b>E</b>) intracellular helix bundle crossing region of hBK and aSlo1, respectively.</p>
Full article ">
18 pages, 2428 KiB  
Article
Slack Potassium Channels Modulate TRPA1-Mediated Nociception in Sensory Neurons
by Fangyuan Zhou, Katharina Metzner, Patrick Engel, Annika Balzulat, Marco Sisignano, Peter Ruth, Robert Lukowski, Achim Schmidtko and Ruirui Lu
Cells 2022, 11(10), 1693; https://doi.org/10.3390/cells11101693 - 19 May 2022
Cited by 5 | Viewed by 2802
Abstract
The transient receptor potential (TRP) ankyrin type 1 (TRPA1) channel is highly expressed in a subset of sensory neurons where it acts as an essential detector of painful stimuli. However, the mechanisms that control the activity of sensory neurons upon TRPA1 activation remain [...] Read more.
The transient receptor potential (TRP) ankyrin type 1 (TRPA1) channel is highly expressed in a subset of sensory neurons where it acts as an essential detector of painful stimuli. However, the mechanisms that control the activity of sensory neurons upon TRPA1 activation remain poorly understood. Here, using in situ hybridization and immunostaining, we found TRPA1 to be extensively co-localized with the potassium channel Slack (KNa1.1, Slo2.2, or Kcnt1) in sensory neurons. Mice lacking Slack globally (Slack−/−) or conditionally in sensory neurons (SNS-Slack−/−) demonstrated increased pain behavior after intraplantar injection of the TRPA1 activator allyl isothiocyanate. By contrast, pain behavior induced by the TRP vanilloid 1 (TRPV1) activator capsaicin was normal in Slack-deficient mice. Patch-clamp recordings in sensory neurons and in a HEK cell line transfected with TRPA1 and Slack revealed that Slack-dependent potassium currents (IKS) are modulated in a TRPA1-dependent manner. Taken together, our findings highlight Slack as a modulator of TRPA1-mediated, but not TRPV1-mediated, activation of sensory neurons. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Slack<sup>−/−</sup> mice display increased allyl isothiocyanate (AITC)-evoked but normal capsaicin-evoked pain behavior. (<b>A</b>) Time course of paw licking and biting (left) and the sum of licking and biting time over 30 min (right; <span class="html-italic">p</span> = 0.0394; <span class="html-italic">n</span> = 7 mice per group) after intraplantar AITC injection in wildtype (WT) and Slack<sup>−/−</sup> littermates. (<b>B</b>) Time course of mechanical hypersensitivity after intraplantar AITC injection. Two-way analysis of variance (ANOVA), effect of genotype (<span class="html-italic">p</span> = 0.0007) with Sidak’s multiple comparison test (<span class="html-italic">p</span> values represent comparisons between genotypes for each time point: 3 h, <span class="html-italic">p</span> = 0.0421; 24 h, <span class="html-italic">p</span> = 0.0076; 48 h, <span class="html-italic">p</span> = 0.0064); <span class="html-italic">n</span> = 8 mice per group. Note that both licking/biting and mechanical hypersensitivity are significantly increased in Slack<sup>−/−</sup> mice after AITC injection. (<b>C</b>) Quantitative RT-PCR in DRGs of WT and Slack<sup>−/−</sup> mice revealed that the transient receptor potential (TRP) ankyrin 1 (TRPA1) mRNA expression is not compensatorily regulated in the absence of Slack (<span class="html-italic">p</span> = 0.3727; <span class="html-italic">n</span> = 3 mice per group). (<b>D</b>) Time course of paw licking and biting (left) and the sum of licking and biting time over 10 min (right; <span class="html-italic">p</span> = 0.9621; <span class="html-italic">n</span> = 7–8 mice per group) after intraplantar capsaicin injection. (<b>E</b>) Time course of mechanical hypersensitivity after intraplantar capsaicin injection. Two-way ANOVA, effect of genotype (<span class="html-italic">p</span> = 0.5893; <span class="html-italic">n</span> = 7–8 mice per group). Note that the capsaicin-induced pain behavior was unaltered in Slack<sup>−/−</sup> mice. (<b>F</b>) Quantitative RT-PCR in DRGs of WT and Slack<sup>−/−</sup> mice revealed that TRP vanilloid 1 (TRPV1) mRNA expression is not compensatory regulated in the absence of Slack (<span class="html-italic">p</span> = 0.4874; <span class="html-italic">n</span> = 3 mice per group) (<b>G</b>) Double in situ hybridization of Slack mRNA and TRPA1 mRNA in DRGs. Scale bar, 50 µm. (<b>H</b>) Double-labeling immunostaining of Slack and TRPV1 in DRGs. A quantitative summary of co-expression in G (180 Slack-positive cells from 3 mice were counted) and H (204 Slack-positive cells from 4 mice were counted) is shown on the right. (<b>I</b>) Expression of TRPA1, TRPV1, and Slack (gene <span class="html-italic">Kcnt1</span>) across sensory neuron subsets from published scRNA-seq data. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>SNS-Slack<sup>−/−</sup> mice display increased AITC-evoked pain behavior. (<b>A</b>) Quantitative RT-PCR in lumbar DRGs, lumbar spinal cord, cerebellum and prefrontal cortex revealed that Slack mRNA is selectively reduced in DRGs of SNS-Slack<sup>−/−</sup> mice (<span class="html-italic">p</span> = 0.0002; <span class="html-italic">n</span> = 3 mice per group). (<b>B</b>,<b>C</b>) Expression pattern and percentages of DRG neurons binding IB4 (2207 cells from 4 mice per group were counted), or immunoreactive for CGRP (2110 cells from 4 mice per group were counted), TH (839 cells from 4 mice per group were counted), NF200 (2751 cells from 4 mice per group were counted), are similar in SNS-Slack<sup>−/−</sup> and control mice. Scale bar, 50 µm. (<b>D</b>) The distribution of central terminals of primary afferents immunoreactive for CGRP or binding IB4 in the dorsal horn appears normal in SNS-Slack<sup>−/−</sup> mice. Scale bar, 50 µm. (<b>E</b>) Quantitative RT-PCR revealed that TRPA1 mRNA expression in lumbar DRGs is similar in control and SNS-Slack<sup>−/−</sup> mice (<span class="html-italic">p</span> = 0.8638; <span class="html-italic">n</span> = 3 mice per group). (<b>F</b>) Time course of paw licking and biting (left; <span class="html-italic">p</span> = 0.0148 for the 0–5 min period) and the sum of licking and biting time over 30 min (right; <span class="html-italic">p</span> = 0.0238; <span class="html-italic">n</span> = 8 mice per group) after intraplantar injection of AITC in control and SNS-Slack<sup>−/−</sup> littermates. (<b>G</b>) Time course of mechanical hypersensitivity after intraplantar AITC injection. Two-way ANOVA, effect of genotype (<span class="html-italic">p</span> = 0.0126) with Sidak’s multiple comparisons test (<span class="html-italic">p</span> = 0.0346, representing comparisons between genotypes for the 24 h time point); <span class="html-italic">n</span> = 8 mice per group. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Lbx1-Slack<sup>−/−</sup> mice display normal AITC-evoked pain behavior. (<b>A</b>) Quantitative RT-PCR in lumbar DRGs, lumbar spinal cord, cerebellum, and prefrontal cortex revealed that Slack mRNA levels are selectively reduced in the spinal cord of Lbx1-Slack<sup>−/−</sup> mice (<span class="html-italic">p</span> &lt; 0.0001; <span class="html-italic">n</span> = 6 mice per group). (<b>B</b>) The distribution of GAD67<sup>+</sup> inhibitory interneurons and PKCγ<sup>+</sup> excitatory interneurons in the dorsal horn appears normal in Lbx1-Slack<sup>−/−</sup> mice. Scale bar, 50 µm. (<b>C</b>) Time course of paw licking and biting (left) and the sum of licking and biting time over 30 min (right; <span class="html-italic">p</span> = 0.7319; <span class="html-italic">n</span> = 6 mice per group) after intraplantar AITC injection in Lbx1-Slack<sup>−/−</sup> and control littermates. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>AITC-mediated calcium influx is normal in sensory neurons of Slack<sup>−/−</sup> mice. (<b>A</b>) Representative examples of Fura-2-ratiometric calcium traces evoked by AITC and KCl in cultured lumbar DRG neurons of WT and Slack<sup>−/−</sup> mice. (<b>B</b>) Magnitude of the calcium response to AITC stimulation (WT, <span class="html-italic">n</span> = 452 neurons in 3 mice; Slack<sup>−/−</sup>, <span class="html-italic">n</span> = 476 neurons in 3 mice; <span class="html-italic">p</span> = 0.7000). (<b>C</b>) Percentage of responsive neurons to AITC stimulation (<span class="html-italic">p</span> = 0.9739). These data show that AITC-evoked calcium responses are normal in DRG neurons from Slack<sup>−/−</sup> mice. Data in B and C are presented as mean ± SEM.</p>
Full article ">Figure 5
<p>AITC-mediated modulation of potassium currents in IB4<sup>+</sup> sensory neurons from WT and Slack<sup>−/−</sup> mice. (<b>A</b>) IV relations of outward potassium currents (I<sub>K</sub>) obtained in whole-cell patch-clamp recordings in IB4<sup>+</sup> sensory neurons from 4 WT (<span class="html-italic">n</span> = 18 cells) and 4 Slack<sup>−/−</sup> mice (<span class="html-italic">n</span> = 13 cells) before and after AITC (200 µM) application in the physiological extracellular buffer. Note that in this experimental setting (which includes 2 mM Ca<sup>2+</sup> and 140 mM Na<sup>+</sup> in the external solution), TRPA1 activation led to a significant reduction in I<sub>K</sub> in sensory neurons from WT but not Slack<sup>−/−</sup> mice. (<b>B</b>) IV relations of I<sub>K</sub> in sensory neurons from 3 WT (<span class="html-italic">n</span> = 13 cells) and 3 Slack<sup>−/−</sup> mice (<span class="html-italic">n</span> = 12 cells) before and after application of the TRPA1 antagonist A-967079 (10 µM) in the physiological extracellular buffer. The TRPA1 antagonist significantly reduced I<sub>K</sub> in sensory neurons from both WT and Slack<sup>−/−</sup> mice. (<b>C</b>) IV relations of I<sub>K</sub> in sensory neurons from 3 WT (<span class="html-italic">n</span> = 9 cells) and 3 Slack<sup>−/−</sup> mice (<span class="html-italic">n</span> = 8 cells) before and after AITC (200 µM) application in a Na<sup>+</sup> free extracellular buffer. In this experimental setting (which includes 2 mM Ca<sup>2+</sup> but no Na<sup>+</sup> in the external solution), TRPA1 activation did not alter I<sub>K</sub> in sensory neurons from both WT and Slack<sup>−/−</sup> mice. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>AITC-mediated modulation of potassium currents in transfected HEK293 cells. (<b>A</b>,<b>B</b>) IV relations of I<sub>K</sub> in HEK-Slack cells ((<b>A</b>); <span class="html-italic">n</span> = 5 cells) and HEK-Slack-TRPA1 cells ((<b>B</b>); <span class="html-italic">n</span> = 14 cells) obtained in whole-cell patch-clamp recordings using a Ca<sup>2+</sup>-free external solution before and after TRPA1 activation by 200 µM AITC. (<b>C</b>) IV relations of I<sub>K</sub> in HEK-Slack-TRPA1 cells before and after TRPA1 inhibition by 10 µM A-967079 (<span class="html-italic">n</span> = 9 cells) in whole-cell patch-clamp recordings using a Ca<sup>2+</sup>-free external solution. (<b>D</b>) IV relations of I<sub>K</sub> in HEK-Slack-TRPA1 cells before and after 200 µM AITC application (<span class="html-italic">n</span> = 9 cells) in whole-cell patch-clamp recordings in physiological extracellular buffer Note that in a Ca<sup>2+</sup>-free external solution, Slack-mediated I<sub>K</sub> is increased after TRPA1 activation. Data are presented as mean ± SEM. Paired <span class="html-italic">t</span> test, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
18 pages, 5524 KiB  
Article
Identification of the Large-Conductance Ca2+-Regulated Potassium Channel in Mitochondria of Human Bronchial Epithelial Cells
by Aleksandra Sek, Rafal P. Kampa, Bogusz Kulawiak, Adam Szewczyk and Piotr Bednarczyk
Molecules 2021, 26(11), 3233; https://doi.org/10.3390/molecules26113233 - 27 May 2021
Cited by 16 | Viewed by 3268
Abstract
Mitochondria play a key role in energy metabolism within the cell. Potassium channels such as ATP-sensitive, voltage-gated or large-conductance Ca2+-regulated channels have been described in the inner mitochondrial membrane. Several hypotheses have been proposed to describe the important roles of mitochondrial [...] Read more.
Mitochondria play a key role in energy metabolism within the cell. Potassium channels such as ATP-sensitive, voltage-gated or large-conductance Ca2+-regulated channels have been described in the inner mitochondrial membrane. Several hypotheses have been proposed to describe the important roles of mitochondrial potassium channels in cell survival and death pathways. In the current study, we identified two populations of mitochondrial large-conductance Ca2+-regulated potassium (mitoBKCa) channels in human bronchial epithelial (HBE) cells. The biophysical properties of the channels were characterized using the patch-clamp technique. We observed the activity of the channel with a mean conductance close to 285 pS in symmetric 150/150 mM KCl solution. Channel activity was increased upon application of the potassium channel opener NS11021 in the micromolar concentration range. The channel activity was completely inhibited by 1 µM paxilline and 300 nM iberiotoxin, selective inhibitors of the BKCa channels. Based on calcium and iberiotoxin modulation, we suggest that the C-terminus of the protein is localized to the mitochondrial matrix. Additionally, using RT-PCR, we confirmed the presence of α pore-forming (Slo1) and auxiliary β3-β4 subunits of BKCa channel in HBE cells. Western blot analysis of cellular fractions confirmed the mitochondrial localization of α pore-forming and predominately β3 subunits. Additionally, the regulation of oxygen consumption and membrane potential of human bronchial epithelial mitochondria in the presence of the potassium channel opener NS11021 and inhibitor paxilline were also studied. In summary, for the first time, the electrophysiological and functional properties of the mitoBKCa channel in a bronchial epithelial cell line were described. Full article
(This article belongs to the Special Issue Compounds Modulating Mitochondrial Ion Channels)
Show Figures

Figure 1

Figure 1
<p>Outline of the mitochondrial patch-clamp experiments and comparison of the two types of recorded channel activities. (<b>a</b>) Scheme of the preparation of human bronchial epithelial mitochondria, mitoplasts, mitoplast patching, patch-clamp inside out mode and single-channel recordings. (<b>b</b>) Comparison of the two types of single-channel current-time recordings of the mitoBK<sub>Ca</sub> channel: control and high activity in a symmetric 150/150 mM KCl isotonic solution (in the presence of 100 µM Ca<sup>2+</sup>) at different voltages (<span class="html-italic">n</span> = 10). “—” indicates a closed channel state. “<span class="html-italic">Po</span>” represents open probability analysis of single channel recordings. (<b>c</b>) Comparison of the numbers of mitoBK<sub>Ca</sub> channels observed with control (<span class="html-italic">n</span> = 32) and high activity (<span class="html-italic">n</span> = 18) in the analyzed patches.</p>
Full article ">Figure 2
<p>Biophysical properties of both mitoBK<sub>Ca</sub> channels (channels with control and high activity) present in the inner mitochondrial membranes of human bronchial epithelial cells. (<b>a</b>) The current-voltage relationship based on single-channel recordings in a symmetric 150/150 mM KCl isotonic solution (100 µM Ca<sup>2+</sup>) of the channel with control and high activity (<span class="html-italic">n</span> = 4). (<b>b</b>) Analysis of the open probability of the mitoBK<sub>Ca</sub> channels with control and high activities in the presence of 100 µM Ca<sup>2+</sup> at different voltages (<span class="html-italic">n</span> = 4). (<b>c</b>,<b>e</b>) Distribution of the mean times of closure and opening of the mitoBK<sub>Ca</sub> channel with control activity (<span class="html-italic">n</span> = 4). (<b>d</b>,<b>f</b>) Distribution of the mean times of closure and opening of the mitoBK<sub>Ca</sub> channel with high activity (<span class="html-italic">n</span> = 3). Notes: (<b>a</b>,<b>b</b>) data are presented as the means ± SD under control conditions (100 µM Ca<sup>2+</sup>); (<b>c</b>–<b>f</b>) the boxes include values for the mean time, the line across the box indicates the median, and whiskers show the minimum and maximum values for mean time of closure/opening.</p>
Full article ">Figure 3
<p>Regulation of the two populations of mitoBK<sub>Ca</sub> channels by calcium ions. (<b>a</b>,<b>b</b>) Representative single-channel current-time recordings of the mitoBK<sub>Ca</sub> channel in the presence of different calcium concentrations (1, 10, 30, 50, 70 and 100 μM Ca<sup>2+</sup>). Both types of channel activities recorded in 150/150 mM KCl isotonic solution at +40 mV are presented (<span class="html-italic">n</span> = 3). “—” indicates a closed channel state. “<span class="html-italic">Po</span>” represents the average channel open probability. (<b>c</b>,<b>d</b>) Quantification of the open probability of channels in the presence of 100, 70, 50, 30, 10 and 1 µM Ca<sup>2+</sup>. Data are presented as the means ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Effects of the potassium channel opener NS11021 on the BK<sub>Ca</sub> channel present in human bronchial epithelial mitochondria. (<b>a</b>,<b>b</b>) Representative traces of single-channel recordings of mitoBK<sub>Ca</sub> channels from human bronchial epithelial mitochondria in symmetric 150/150 mM KCl control solution (100 µM Ca<sup>2+</sup>) under different conditions. Sequences of the applied drugs (NS11021 and paxilline) are presented in the graph. Left panel, (<b>a</b>) mitoBK<sub>Ca</sub> channel with control activity, and right panel, (<b>b</b>) mitoBK<sub>Ca</sub> channel with high activity. Single-channel activities from different experiments (<span class="html-italic">n</span> = 3) recorded at −40 mV. “—” indicates a closed channel state. The <span class="html-italic">Po</span> analysis (%) of single channel events is shown. (<b>c</b>,<b>d</b>) Analysis of the mitoBK<sub>Ca</sub> channel open probability under different conditions: control (100 µM Ca<sup>2+</sup>), different concentrations of NS11021, washout and 3 µM NS11021 plus 1 µM paxilline. Left panel, (<b>c</b>) mitoBK<sub>Ca</sub> channel with control activity, and right panel, (<b>d</b>) mitoBK<sub>Ca</sub> channel with high activity. Notes: the boxes include values for <span class="html-italic">Po</span>, the dots are single counted traces, the line across the box indicates the median, and whiskers show the minimum and maximum values for <span class="html-italic">Po</span>. The data are presented as the means ± SD (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span> &lt; 0.0001 (****).</p>
Full article ">Figure 5
<p>Effect of iberiotoxin on mitoBK<sub>Ca</sub> channel activity. (<b>a</b>) Scheme of the mitochondrial matrix direction of iberiotoxin administration to the mitoBK<sub>Ca</sub> channel with control activity. Representative single-channel current-time recordings of the mitoBK<sub>Ca</sub> channel incubated with 300 nM iberiotoxin (<span class="html-italic">n</span> = 3) for different times at −40 mV and +40 mV. ‘—’ indicates a closed channel state. The <span class="html-italic">Po</span> analysis (%) of single channel events is shown. (<b>b</b>) Scheme of the intermembrane space direction of iberiotoxin administration to the mitoBK<sub>Ca</sub> channel with control activity. Representative traces showing the effects of the time of incubation with 300 nM iberiotoxin (<span class="html-italic">n</span> = 3) recorded at −40 mV and +40 mV. “<b>—</b>” indicates a closed channel state, and the <span class="html-italic">Po</span> analysis (%) of single channel events is shown. The charts below show the quantification of the mitoBK<sub>Ca</sub> channel open probability in the presence of the control (100 µM Ca<sup>2+</sup>) and 300 nM iberiotoxin after different incubation times. <span class="html-italic">p</span> &lt; 0.05 (*).</p>
Full article ">Figure 6
<p>Analysis of gene expression and the immunodetection of the BK<sub>Ca</sub> channel α pore forming and auxiliary β subunits. (<b>a</b>) PCR amplification of BK<sub>Ca</sub> channel subunits. α, Product of the amplification of <span class="html-italic">KCNMA1</span> gene (<span class="html-italic">n</span> = 3); β1–4, products of the amplification of transcripts encoded by KCNMB1-KCNMB4 (<span class="html-italic">n</span> = 3). (<b>b</b>) qPCR detection of the BK<sub>Ca</sub> channel subunits; Ct: cycle threshold; (<span class="html-italic">n</span> = 4). (<b>c</b>) Representative images of immunostaining for BK<sub>Ca</sub> channel subunits. Western blot analysis of proteins from the cell homogenate (H; 20 µg of total protein) and mitochondrial fractions (M at two concentrations, 20 µg and 40 µg), loaded into each lane (<span class="html-italic">n</span> = 3). (<b>d</b>) Quantification of band intensities was performed using ImageJ software. The number of Western blots analyzed was three per group. <span class="html-italic">p</span> &lt; 0.05 (*), <span class="html-italic">p</span> &lt; 0.01 (**).</p>
Full article ">Figure 7
<p>Effects of mitoBK<sub>Ca</sub> channel modulators on the membrane potential and respiration rate. (<b>a</b>) Mitochondrial respiratory capacity; a final density of 1.5 × 10<sup>6</sup> cells/mL was plated in each chamber and the NS-induced changes in respiratory rate in the presence (•) or absence (<span style="color:#808080">•</span>) of 10 µM paxilline were detected. Data are reported as the means ± SD (<span class="html-italic">n</span> = 3), and each dot represents a separate experiment. (<b>b</b>) Measurement of the mitochondrial membrane potential using the JC-10 dye. Changes in red vs. green fluorescence were used to measure the mitochondrial membrane potential after treatment of 16HBE14o- cells with BK<sub>Ca</sub> channel modulators. FCCP-treated cells were used as positive controls. (<b>c</b>) Representative dot plots of stained 16HBE14o- cells obtained using a flow cytometry analysis. Cells that appear in quadrant 4 represent cells with depolarized mitochondrial membranes. <span class="html-italic">p</span> &lt; 0.05 (*).</p>
Full article ">
16 pages, 2549 KiB  
Article
Effective Activation of BKCa Channels by QO-40 (5-(Chloromethyl)-3-(Naphthalen-1-yl)-2-(Trifluoromethyl)Pyrazolo [1,5-a]pyrimidin-7(4H)-one), Known to Be an Opener of KCNQ2/Q3 Channels
by Wei-Ting Chang and Sheng-Nan Wu
Pharmaceuticals 2021, 14(5), 388; https://doi.org/10.3390/ph14050388 - 21 Apr 2021
Cited by 2 | Viewed by 1990
Abstract
QO-40 (5-(chloromethyl)-3-(naphthalene-1-yl)-2-(trifluoromethyl) pyrazolo[1,5-a]pyrimidin-7(4H)-one) is a novel and selective activator of KCNQ2/KCNQ3 K+ channels. However, it remains largely unknown whether this compound can modify any other type of plasmalemmal ionic channel. The effects of QO-40 on ion channels in pituitary GH [...] Read more.
QO-40 (5-(chloromethyl)-3-(naphthalene-1-yl)-2-(trifluoromethyl) pyrazolo[1,5-a]pyrimidin-7(4H)-one) is a novel and selective activator of KCNQ2/KCNQ3 K+ channels. However, it remains largely unknown whether this compound can modify any other type of plasmalemmal ionic channel. The effects of QO-40 on ion channels in pituitary GH3 lactotrophs were investigated in this study. QO-40 stimulated Ca2+-activated K+ current (IK(Ca)) with an EC50 value of 2.3 μM in these cells. QO-40-stimulated IK(Ca) was attenuated by the further addition of GAL-021 or paxilline but not by linopirdine or TRAM-34. In inside-out mode, this compound added to the intracellular leaflet of the detached patches stimulated large-conductance Ca2+-activated K+ (BKCa) channels with no change in single-channel conductance; however, there was a decrease in the slow component of the mean closed time of BKCa channels. The KD value required for the QO-40-mediated decrease in the slow component at the mean closure time was 1.96 μM. This compound shifted the steady-state activation curve of BKCa channels to a less positive voltage and decreased the gating charge of the channel. The application of QO-40 also increased the hysteretic strength of BKCa channels elicited by a long-lasting isosceles-triangular ramp voltage. In HEK293T cells expressing α-hSlo, QO-40 stimulated BKCa channel activity. Overall, these findings demonstrate that QO-40 can interact directly with the BKCa channel to increase the amplitude of IK(Ca) in GH3 cells. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of QO-40.</p>
Full article ">Figure 2
<p>Stimulatory effect of QO-40 on the magnitude of whole-cell Ca<sup>2+</sup>-activated K<sup>+</sup> current (<span class="html-italic">I</span><sub>K(Ca)</sub>) recorded from GH<sub>3</sub> pituitary tumor cells. This set of voltage-clamp experiments was undertaken in cells which were kept immersed in normal Tyrode’s solution containing 1.8 mM CaCl<sub>2</sub>; the recording pipet used was backfilled with a K<sup>+</sup>-containing solution. We elicited <span class="html-italic">I</span><sub>K(Ca)</sub> from a holding potential of 0 mV to test potentials in the range of 0 and +60 mV (10 mV increments) at a rate of 0.1 Hz. (<b>A</b>) Representative <span class="html-italic">I</span><sub>K(Ca)</sub> traces activated in response to a series of voltage steps (indicated in the uppermost part). Current traces in the upper part are controls (i.e., QO-40 was not present), while those in the lower part were obtained in the presence of 3 μM QO-40. Arrowhead in each panel denotes the zero-current level, calibration mark in the right lower corner applies to all current traces illustrated, and the duration of square voltage command pulse applied was set in the range of 300 and 180 ms (30 ms decrements). (<b>B</b>) Mean current–voltage (<span class="html-italic">I-V</span>) relationship of <span class="html-italic">I</span><sub>K(Ca)</sub> obtained in the control, during the exposure to 3 μM QO-40, or after washout of QO-40. Each point represents the mean SEM (<span class="html-italic">n</span> = 7–9). The statistical analyses were undertaken by ANOVA-2 for repeated measures, <span class="html-italic">p</span> (factor 1, groups among data taken at different levels of voltage) &lt; 0.05, <span class="html-italic">p</span> (factor 2, groups between the absence and presence of 3 μM QO-40) &lt; 0.05, <span class="html-italic">p</span> (interaction) &lt; 0.05, followed by post hoc Fisher’s least-significant difference test, <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) Concentration–response relationship for QO-40-induced stimulation of <span class="html-italic">I</span><sub>K(Ca)</sub>. Current amplitude was taken at the end of depolarizing pulse from 0 to +50 mV. Data analysis was performed by ANOVA-1 (<span class="html-italic">p</span> &lt; 0.05). The smooth dashed line is fitted to the Hill equation. The values for EC<sub>50</sub>, maximal percentage increase in <span class="html-italic">I</span><sub>K(Ca)</sub> amplitude, and Hill coefficient were yielded to be 2.3 μM, 100%, and 1.3, respectively. Each point represents the mean ± SEM (<span class="html-italic">n</span> = 8).</p>
Full article ">Figure 3
<p>Effects of linopirdine, TRAM-34, GAL-021, and paxilline on QO-40-stimulated <span class="html-italic">I</span><sub>K(Ca)</sub> in GH<sub>3</sub> cells. In this set of experiments, whole-cell current recordings were undertaken in cells bathed in normal Tyrode’s solution, and the pipet was backfilled with K<sup>+</sup>-containing internal solution. (<b>A</b>) Representative <span class="html-italic">I</span><sub>K(Ca)</sub> traces in the absence (a, blue color) and presence of either QO-40 (b, red color), QO-40 plus linopirdine (c, upper panel, green color), or QO-40 plus paxilline (c, lower panel, brown color). The uppermost part shows the voltage-clamp protocol used. (<b>B</b>) Vertical scatter graph showing effects of linopirdine, TRAM-34, GAL-021, or paxilline on QO-40-induced stimulation of <span class="html-italic">I</span><sub>K(Ca)</sub> (mean ± SEM; <span class="html-italic">n</span> = 6–8 for each point). QO-40: 3 μM QO-40; Lino: 10 μM linopirdine; TRAM-34: 3 μM TRAM-34; GAL-021: 3 μM GAL-021; Pax: 1 μM paxilline. Data analysis was performed by ANOVA-1 (<span class="html-italic">p</span> &lt; 0.05). * Significantly different from control (<span class="html-italic">p</span> &lt; 0.05) and † significantly different from QO-40 (3 μM) alone group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Stimulatory effect of QO-40 on the large-conductance Ca<sup>2+</sup>-activated K<sup>+</sup> (BK<sub>Ca</sub>) channels identified in GH<sub>3</sub> cells. The single-channel experiments in an excised inside-out membrane patch were undertaken with symmetrical K<sup>+</sup> concentration (145 mM). The bath medium contained 0.1 μM Ca<sup>2+</sup>, and we kept the patch in voltage clamp at the level of +60 mV. (<b>A</b>) Representative current trace (upper, blue color) and the open probability (lower, red color) showing changes in the activity of BK<sub>Ca</sub> channels after addition of QO-40 (3 μM). Channel openings are indicated as upward deflections, and the horizontal bar shown above either current tracings or time course of single open probability corresponds to the application of QO-40 to the bath. The lower parts in (<b>A</b>) depict expanded records obtained from the dashed boxes in the uppermost part. Current traces in the left or right side indicate the absence or presence of 3 μM QO-40, respectively. Note that the presence of QO-40 leads to an increase in channel open-state probability of BK<sub>Ca</sub> channels. (<b>B</b>) BK<sub>Ca</sub>-channel activity obtained in the control period (i.e., neither QO-40 nor paxilline was present) (upper) and QO-40 (3 μM) plus paxilline (1 μM) (lower). In the experiments on QO-40 plus paxilline, paxilline (1 μM) was further added, as patch was continually exposed to QO-40 (3 μM). (<b>C</b>) Vertical scatter graph showing effects of QO-40, QO-40 plus TRAM-34, QO-40 plus linopirdine, or QO-40 plus paxilline on channel open-state probability of BK<sub>Ca</sub> channels (mean ± SEM; <span class="html-italic">n</span> = 8 for each point). Lino: 10 μM linopirdine; TRAM-34: 3 μM TRAM-34; Pax: 1 μM paxilline. Data analysis was performed by ANOVA-1 (<span class="html-italic">p</span> &lt; 0.05). * Significantly different from control (<span class="html-italic">p</span> &lt; 0.05) and † significantly different from QO-40 (3 μM) alone group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of QO-40 on the mean closed time of BK<sub>Ca</sub> channels identified in GH<sub>3</sub> cells (<b>A</b>) and the relationship of the reciprocal of slow component in the mean closed time of the channel versus the QO-40 concentration (<b>B</b>). In (<b>A</b>), mean closed-time histogram of BK<sub>Ca</sub> channels in the absence (left) or presence (right) of QO-40 (10 μM) in GH<sub>3</sub> cells is illustrated, respectively. Under symmetrical K<sup>+</sup> concentrations (145 mM) in which bath medium contained 0.1 μM Ca<sup>2+</sup>, the potential was voltage-clamped at +60 mV, and the inside-out configuration was performed. The closed-time histogram in the absence or presence of 10 μM QO-40 was least-squares fitted by a sum of two-exponential function (indicated by nonlinear continuous curve, pink color) with a mean closed time of 13 and 146 ms, or 12 and 36 ms, respectively. The x- or y-axis indicates the logarithm of mean closed time (ms) or the square root of the event number, respectively, and the broken line in each lifetime distribution is pointed at the values of the fast or slow component of time constant in closed (resting) states of the channel. Data in the control were obtained from a measurement of 232 channel openings with a total recording time of 1 min, whereas those in the presence of 10 μM QO-40 were from 439 channel openings with a total record time of 30 s. In (<b>B</b>), the reciprocal of slow component in the mean closed time of the channel (i.e., 1/t) versus the QO-40 concentration was derived and plotted. Data points indicated in filled squares were fitted by a linear regression (red color); hence, a molecularity of one was inferred. According to the first-order binding scheme elaborated in <a href="#sec4-pharmaceuticals-14-00388" class="html-sec">Section 4</a>, forward (on, <span class="html-italic">k</span><sub>+1</sub><sup>*</sup>) or backward (off, <span class="html-italic">k</span><sub>−1</sub>) rate constant for QO-40-induced decrease in the slow component of the mean closed time of the channel was calculated to be 2.298 s<sup>−1</sup>μM<sup>−1</sup> or 4.512 s<sup>−1</sup>, respectively. Mean ± SEM (<span class="html-italic">n</span> = 8–10 for each point).</p>
Full article ">Figure 6
<p>Effect of OD-40 on BK<sub>Ca</sub>-channel activity measured at the different levels of membrane potentials. Inside-out current recordings were performed in these experiments; cells were bathed in high-K<sup>+</sup> solution (145 mM) containing 0.1 μM Ca<sup>2+</sup>. (<b>A</b>) Representative current traces obtained in the absence (left, blue color) and presence (right, red color) of 3 μM OD-40. The number shown in each panel indicates the membrane potential held, and the upper deflection is the opening event of the channel. (<b>B</b>) Relationship of single-channel current versus membrane potential (mean ± SEM; <span class="html-italic">n</span> = 9 for each point). The dashed lines were pointed to the reversal potential with 0 mV. Notice that the two lines are virtually superimposed, indicating the single-channel conductance of BK<sub>Ca</sub> channels did not differ between the absence (filled symbol, blue color) and presence (open circles, red color) of OD-40. (<b>C</b>) The steady-state activation curve of BK<sub>Ca</sub> obtained with or without addition of 3 μM OD-40 (mean ± SEM; <span class="html-italic">n</span>-7 for each point). The statistical analyses were undertaken by Student’s <span class="html-italic">t</span>-tests (<span class="html-italic">p</span> &lt; 0.05). Continuous sigmoidal lines were best fit to the modified Boltzmann equation as described under <a href="#sec4-pharmaceuticals-14-00388" class="html-sec">Section 4</a>.</p>
Full article ">Figure 7
<p>Effect of QO-40 on the voltage-dependent hysteresis of BK<sub>Ca</sub> channels identified from GH<sub>3</sub> cells. In this set of inside-out current recordings, we bathed cells in symmetrical K<sup>+</sup> solution (145 mM). (<b>A</b>) Representative current traces obtained in the presence of QO-40 (3 μM). Channel activities were activated in response to long isosceles-triangular ramp pulse with a duration of 2.8 sec (indicated in the Inset). The dashed arrow indicates the direction of current flow through the channel in which time passes. (<b>B</b>) The relationship of the relative channel open probability versus membrane potential of BK<sub>Ca</sub> channels in response to the forward (blue color) or backward (pink color) limb of triangular ramp pulse. (<b>C</b>) Vertical scatter graph showing effect of varying QO-40 concentrations on the hysteresis of BK<sub>Ca</sub> channels. Hysteresis was measured at the voltage separation between the forward and backward limb at 50% of the relative channel open probability. Inside-out configuration was made, and a ramp speed of ±93 mV/s was applied to the patch. Each point indicates the mean ± SEM (<span class="html-italic">n</span> = 7). Data analysis was performed by ANOVA-1 (<span class="html-italic">p</span> &lt; 0.05). * Significantly different from 1 μM QO-40 alone group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
24 pages, 2344 KiB  
Review
Structure and Function of Ion Channels Regulating Sperm Motility—An Overview
by Karolina Nowicka-Bauer and Monika Szymczak-Cendlak
Int. J. Mol. Sci. 2021, 22(6), 3259; https://doi.org/10.3390/ijms22063259 - 23 Mar 2021
Cited by 47 | Viewed by 5766
Abstract
Sperm motility is linked to the activation of signaling pathways that trigger movement. These pathways are mainly dependent on Ca2+, which acts as a secondary messenger. The maintenance of adequate Ca2+ concentrations is possible thanks to proper concentrations of other [...] Read more.
Sperm motility is linked to the activation of signaling pathways that trigger movement. These pathways are mainly dependent on Ca2+, which acts as a secondary messenger. The maintenance of adequate Ca2+ concentrations is possible thanks to proper concentrations of other ions, such as K+ and Na+, among others, that modulate plasma membrane potential and the intracellular pH. Like in every cell, ion homeostasis in spermatozoa is ensured by a vast spectrum of ion channels supported by the work of ion pumps and transporters. To achieve success in fertilization, sperm ion channels have to be sensitive to various external and internal factors. This sensitivity is provided by specific channel structures. In addition, novel sperm-specific channels or isoforms have been found with compositions that increase the chance of fertilization. Notably, the most significant sperm ion channel is the cation channel of sperm (CatSper), which is a sperm-specific Ca2+ channel required for the hyperactivation of sperm motility. The role of other ion channels in the spermatozoa, such as voltage-gated Ca2+ channels (VGCCs), Ca2+-activated Cl-channels (CaCCs), SLO K+ channels or voltage-gated H+ channels (VGHCs), is to ensure the activation and modulation of CatSper. As the activation of sperm motility differs among metazoa, different ion channels may participate; however, knowledge regarding these channels is still scarce. In the present review, the roles and structures of the most important known ion channels are described in regard to regulation of sperm motility in animals. Full article
(This article belongs to the Special Issue Ion Channels in Sperm Physiology 2.0)
Show Figures

Figure 1

Figure 1
<p>Voltage-gated Ca<sup>2+</sup> channel (VGCC) structure scheme. (<b>a</b>) The topology of the α1 subunit is made up of four homologous domains that each consist of six transmembrane α helices (TM1–6). TM4 from each homologous domain serves as the voltage sensor moving outward and rotates under the influence of the electric field, thereby initiating a conformational change that opens the respective pore. TM5, TM6, and the loop between them (P-loop) from each domain form a pore. The C-terminal tail contains a Ca<sup>2+</sup> binding domain (CBD) and in some types of VGCCs a site for calmodulin (calcium-modulated protein; CaM) binding. The binding of Ca<sup>2+</sup> to CBD or via CaM inactivates the channels. (<b>b</b>) A schematic presentation of the VGCC subunits (α1, α2δ, β, and γ) with their spatial localizations. (<b>c</b>) Overview of the types of VGCCs in relation to Vm-dependent activation – high voltage activation (HVA) and low voltage activation (LVA) (based on References [<a href="#B26-ijms-22-03259" class="html-bibr">26</a>,<a href="#B27-ijms-22-03259" class="html-bibr">27</a>]).</p>
Full article ">Figure 2
<p>Topology of a store-operated Ca<sup>2+</sup> channel (SOCC) created by ORAI1. Each ORAI protein has four TMs. TM2 and TM3 create a pore. There are two sites for STIM1 binding at the N- and C-termini. The interaction between STIM1 and ORAI activates the channel and the release of Ca<sup>2+</sup> from the endoplasmic reticulum (ER). The binding of Ca<sup>2+</sup> by the Ca<sup>2+</sup> binding domain (CBD) localized on the central loop inactivates the channel [<a href="#B51-ijms-22-03259" class="html-bibr">51</a>]. Additionally, it can also be inactivated upon CaM binding [<a href="#B52-ijms-22-03259" class="html-bibr">52</a>].</p>
Full article ">Figure 3
<p>A topological and spatial structure of CatSper. (<b>a</b>) The α1 subunit created by CatSper1. Like most voltage-gated channels, each α subunit contains six transmembrane domains (TM1–TM6) creating two physiologically distinctive regions, namely the voltage-sensing domain (VSD; TM1–4) and pore-forming region (TM5–6). Each TM4 contains several (two to six) positively charged amino acid residues that serve as voltage sensors (reviewed in Reference [<a href="#B57-ijms-22-03259" class="html-bibr">57</a>]). Voltage slopes move TM4, resulting in conformational changes that open and close the channel pore [<a href="#B64-ijms-22-03259" class="html-bibr">64</a>]. Additionally, a short and hydrophobic cyclic structure linking TM5–6 contains a conserved homologous amino acid sequence (T × D × W), which selectively permits Ca<sup>2+</sup> influx. The N-terminus of CatSper 1 contains a specific histidine-rich region that might be involved in the pH regulation of CatSper activity. (<b>b</b>) The topological localizations of all auxiliary subunits are not randomly organized. The auxiliary CatSperβ subunit has two predicted TMs that are separated by a large (ca. 1000 amino acids) extracellular loop [<a href="#B64-ijms-22-03259" class="html-bibr">64</a>], whereas CatSperγ, CatSperδ, and CatSperε feature only one TM. Brown et al. [<a href="#B69-ijms-22-03259" class="html-bibr">69</a>] suggested that CatSperζ is a late evolutionary adaptation to maximize fertilization success inside the female mammalian reproductive tract. The predicted topology of Hwang et al. [<a href="#B62-ijms-22-03259" class="html-bibr">62</a>] situates the CatSperζ and EFCAB9 subunits as a cytoplasm complex that is located just below the CatSper 1–4 subunits. This complex interacts with the channel pore as a gatekeeper. The increase in pH<sub>i</sub> causes Ca<sup>2+</sup> binding to highly conserved EF-hands of EFCAB9, leading to dissociation of the EFCAB9-CatSperζ complex and full activation of the channel. Accordingly, EFCAB9-CatSperζ appears to be responsible for both modulation of the channel activity and organization of the CatSper domains [<a href="#B62-ijms-22-03259" class="html-bibr">62</a>]. The scheme has been prepared based on Reference [<a href="#B62-ijms-22-03259" class="html-bibr">62</a>].</p>
Full article ">Figure 4
<p>A simplified topology of the TMEM16A monomer. Each monomer has 10 TMs. The ion conduction pore of TMEM16A is formed by TMs three to seven in each subunit, and thus the CaCC features two pores [<a href="#B96-ijms-22-03259" class="html-bibr">96</a>,<a href="#B97-ijms-22-03259" class="html-bibr">97</a>]. As summarized in a review of Ji et al. [<a href="#B97-ijms-22-03259" class="html-bibr">97</a>], the activation of TMEM16A is gated by two main mechanisms: voltage (Vm) and low concentrations of Ca<sup>2+</sup> (&lt;600 nM) via the EEEEEAVK motif in the TM2–TM3 loop. Contreras-Vite et al. [<a href="#B98-ijms-22-03259" class="html-bibr">98</a>] proposed a gating mechanism model where TMEM16A is directly activated by the Vm-dependent binding of two Ca<sup>2+</sup> ions coupled by a Vm-dependent binding of one external Cl<sup>−</sup> ion. The scheme was prepared based on Reference [<a href="#B97-ijms-22-03259" class="html-bibr">97</a>].</p>
Full article ">Figure 5
<p>SLO1 structure scheme. (<b>a</b>) A topology of the α subunit. Each α subunit consists of seven (0–6) TMs, where TM4 is a typical voltage-sensing domain (VSD). An extracellular loop between TM5 and TM6 forms the pore. The N-tail is located extracellularly but the C-end is a long tail containing the RCK1 (regulator of K<sup>+</sup> conductance 1) and RKC2 domains [<a href="#B135-ijms-22-03259" class="html-bibr">135</a>]. The structural difference between SLO1 and SLO3 is that there are “Ca<sup>2+</sup>-bowl” structures within the RKC domains of SLO1, making the channel sensitive to [Ca<sup>2+</sup>]<sub>i</sub>. (<b>b</b>) In the tetrameric structure of the channel, the cytoplasmic C-termini creates a gating ring. According to the literature, SLO1 has five auxiliary subunits: one β subunit (with two transmembrane domains) and four Leucine-rich repeat-containing membrane proteins (LRRCs, also named γ subunits), LRRC26, LRCC52, LRRC55, and LRRC38, which modulate SLO1 sensitivity to Vm and [Ca<sup>2+</sup>]<sub>i</sub> (revised by Reference [<a href="#B144-ijms-22-03259" class="html-bibr">144</a>]). In murine testes and spermatozoa, two auxiliary subunits of the SLO3 channel have been identified: Lrrc52 and Lrrc26. Both of them are involved in the regulation of SLO3, and the expression of Lrrc52 is critically dependent on the presence of SLO3 [<a href="#B143-ijms-22-03259" class="html-bibr">143</a>]. The schemes are adapted from References [<a href="#B136-ijms-22-03259" class="html-bibr">136</a>,<a href="#B144-ijms-22-03259" class="html-bibr">144</a>].</p>
Full article ">Figure 6
<p>A structure of a voltage-gated Na<sup>+</sup> channel (VGNC) based on a SCN2A isoform. (<b>a</b>) The α subunit is created by four repeat domains (RD1–RD4) that each have six TMs. Classically, TM1–TM4 of each domain form a VSD where TM4 acts as a positively charged sensor. During depolarization, TM4 is believed to move toward the extracellular surface, allowing the channel to become permeable to ions. Na<sup>+</sup> is transported inside a cell through a pore (P-loop) formed between TM5 and TM6 of each RD. The RDs are connected with long intracytoplasmic loops with sites for protein phosphorylation via PKA and PKC [<a href="#B157-ijms-22-03259" class="html-bibr">157</a>]. The cytoplasmic loop between RD3 and RD4 contains an “h” (I × F × M sequence) motif, which stands for a hydrophobic triad of amino acids, namely, isoleucine, phenylalanine, and methionine (I1488, F1489, and M1490). The IFM motif is involved in the inactivation of VGNC, serving as a hydrophobic latch for a hinged lid formed by the loop between RD3 and RD4 [<a href="#B159-ijms-22-03259" class="html-bibr">159</a>]. Phosphorylation in the RD1/RD2 and RD3/RD4 loops modulates the channel inactivation (adapted from Reference [<a href="#B157-ijms-22-03259" class="html-bibr">157</a>], revised in Reference [<a href="#B160-ijms-22-03259" class="html-bibr">160</a>]). (<b>b</b>) A cartoon of VGNC created by the pore-forming α subunit and the two auxiliary β subunits.</p>
Full article ">Figure 7
<p>A voltage-gated H<sup>+</sup> channel (VGHC) and its structure. (<b>a</b>) A VGHC monomer is created by four TMs which in classical voltage-gated channels comprise VSD. Accordingly, VGHCs do not possess a pore-forming domain (TM5-TM6) and the extrusion of H<sup>+</sup> ions probably takes place via a water wire spanning the VSD [<a href="#B168-ijms-22-03259" class="html-bibr">168</a>]. According to Boonamnaj et al. [<a href="#B169-ijms-22-03259" class="html-bibr">169</a>], in VGHC dimers, C-terminal tails interact by forming a coiled structure that stabilizes the channel. Sites of phosphorylation in the N-termini may enhance the selectivity of the channel. (<b>b</b>) A dimeric structure of a VGHC. As the VGHC has no pore-forming domains, H<sup>+</sup> diffuses through each monomer.</p>
Full article ">
17 pages, 4761 KiB  
Article
HVCN1 but Not Potassium Channels Are Related to Mammalian Sperm Cryotolerance
by Ariadna Delgado-Bermúdez, Yentel Mateo-Otero, Marc Llavanera, Sergi Bonet, Marc Yeste and Elisabeth Pinart
Int. J. Mol. Sci. 2021, 22(4), 1646; https://doi.org/10.3390/ijms22041646 - 6 Feb 2021
Cited by 3 | Viewed by 2285
Abstract
Little data exist about the physiological role of ion channels during the freeze–thaw process in mammalian sperm. Herein, we determined the relevance of potassium channels, including SLO1, and of voltage-gated proton channels (HVCN1) during mammalian sperm cryopreservation, using the pig as a model [...] Read more.
Little data exist about the physiological role of ion channels during the freeze–thaw process in mammalian sperm. Herein, we determined the relevance of potassium channels, including SLO1, and of voltage-gated proton channels (HVCN1) during mammalian sperm cryopreservation, using the pig as a model and through the addition of specific blockers (TEA: tetraethyl ammonium chloride, PAX: paxilline or 2-GBI: 2-guanidino benzimidazole) to the cryoprotective media at either 15 °C or 5 °C. Sperm quality of the control and blocked samples was performed at 30- and 240-min post-thaw, by assessing sperm motility and kinematics, plasma and acrosome membrane integrity, membrane lipid disorder, intracellular calcium levels, mitochondrial membrane potential, and intracellular O2⁻ and H2O2 levels. General blockade of K+ channels by TEA and specific blockade of SLO1 channels by PAX did not result in alterations in sperm quality after thawing as compared to control samples. In contrast, HVCN1-blocking with 2-GBI led to a significant decrease in post-thaw sperm quality as compared to the control, despite intracellular O2⁻ and H2O2 levels in 2-GBI blocked samples being lower than in the control and in TEA- and PAX-blocked samples. We can thus conclude that HVCN1 channels are related to mammalian sperm cryotolerance and have an essential role during cryopreservation. In contrast, potassium channels do not seem to play such an instrumental role. Full article
Show Figures

Figure 1

Figure 1
<p>Percentages of viable spermatozoa (SYBR14<sup>+</sup>/PI<sup>−</sup>) in extended and frozen–thawed (FT) samples at 30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 2
<p>Percentages of total (<b>a</b>,<b>b</b>) and progressive (<b>c</b>,<b>d</b>) motile spermatozoa in extended and frozen–thawed (FT) samples at 30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>,<b>c</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>,<b>d</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 2 Cont.
<p>Percentages of total (<b>a</b>,<b>b</b>) and progressive (<b>c</b>,<b>d</b>) motile spermatozoa in extended and frozen–thawed (FT) samples at 30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>,<b>c</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>,<b>d</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 3
<p>Curvilinear velocity (VCL; <b>a</b>,<b>b</b>) and straight-line velocity (VSL; <b>c</b>,<b>d</b>) in extended and frozen–thawed (FT) samples at 30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>,<b>c</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>,<b>d</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 4
<p>Percentages of viable spermatozoa with an intact acrosome in extended and frozen-thawed (FT) samples at 30 min and 240 min post-thaw. In experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>), whereas in experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 5
<p>Percentages of viable spermatozoa with high membrane lipid disorder in extended and frozen–thawed (FT) samples at 30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 6
<p>Percentages of viable spermatozoa with high levels of intracellular calcium in extended samples and frozen—30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 7
<p>JC1<sub>agg</sub>/JC1<sub>mon</sub> ratios in extended and frozen–thawed (FT) samples at 30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">Figure 8
<p>Mitochondrial superoxide levels expressed as percentages of viable spermatozoa Mito-E<sup>+</sup>/YO-PRO-1⁻ (<b>a</b>,<b>b</b>), and peroxide levels expressed as percentages of viable spermatozoa DCF<sup>+</sup>/PI⁻, (<b>c</b>,<b>d</b>) in extended and frozen–thawed (FT) samples at 30 min and 240 min post-thaw. In Experiment 1 (<span class="html-italic">n</span> = 8), TEA, PAX, and 2-GBI blockers were added to LEY medium at 15 °C (<b>a</b>,<b>c</b>), whereas in Experiment 2 (<span class="html-italic">n</span> = 9), they were added to LEYGO medium at 5 °C (<b>b</b>,<b>d</b>). Different superscripts indicate significant differences (<span class="html-italic">P</span> &lt; 0.05) between samples within the same time point. Results are given as mean ± SEM.</p>
Full article ">
18 pages, 2900 KiB  
Article
Functional Coupling of Slack Channels and P2X3 Receptors Contributes to Neuropathic Pain Processing
by Ruirui Lu, Katharina Metzner, Fangyuan Zhou, Cathrin Flauaus, Annika Balzulat, Patrick Engel, Jonas Petersen, Rebekka Ehinger, Anne Bausch, Peter Ruth, Robert Lukowski and Achim Schmidtko
Int. J. Mol. Sci. 2021, 22(1), 405; https://doi.org/10.3390/ijms22010405 - 2 Jan 2021
Cited by 10 | Viewed by 4423
Abstract
The sodium-activated potassium channel Slack (KNa1.1, Slo2.2, or Kcnt1) is highly expressed in populations of sensory neurons, where it mediates the sodium-activated potassium current (IKNa) and modulates neuronal activity. Previous studies suggest that Slack is involved in the processing [...] Read more.
The sodium-activated potassium channel Slack (KNa1.1, Slo2.2, or Kcnt1) is highly expressed in populations of sensory neurons, where it mediates the sodium-activated potassium current (IKNa) and modulates neuronal activity. Previous studies suggest that Slack is involved in the processing of neuropathic pain. However, mechanisms underlying the regulation of Slack activity in this context are poorly understood. Using whole-cell patch-clamp recordings we found that Slack-mediated IKNa in sensory neurons of mice is reduced after peripheral nerve injury, thereby contributing to neuropathic pain hypersensitivity. Interestingly, Slack is closely associated with ATP-sensitive P2X3 receptors in a population of sensory neurons. In vitro experiments revealed that Slack-mediated IKNa may be bidirectionally modulated in response to P2X3 activation. Moreover, mice lacking Slack show altered nocifensive responses to P2X3 stimulation. Our study identifies P2X3/Slack signaling as a mechanism contributing to hypersensitivity after peripheral nerve injury and proposes a potential novel strategy for treatment of neuropathic pain. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Neuropathic Pain and Nerve Injury)
Show Figures

Figure 1

Figure 1
<p>Neuropathic pain behavior is increased in Slack<sup>-/-</sup> mice. (<b>A</b>) paw withdrawal latencies of Slack<sup>-/-</sup> and wild-type (WT) mice after mechanical stimulation with von Frey filaments (up-and-down method) in the spared nerve injury (SNI) model of neuropathic pain (<span class="html-italic">n</span> = 10 animals per group). Thirteen days after SNI, Slack<sup>-/-</sup> mice showed increased mechanical hypersensitivity compared to WT littermates (two-way ANOVA; <span class="html-italic">p</span> = 0.0312; WT versus Slack<sup>-/-</sup>). (<b>B</b>) percentage of weight bearing on the ipsilateral hind paw relative to both hind paws in Slack<sup>-/-</sup> and WT mice (<span class="html-italic">n</span> = 10 animals per group), as assessed using a dynamic weight-bearing device. Twelve days after SNI, the weight-bearing reduction was more pronounced in Slack<sup>-/-</sup> mice compared to WT littermates (two-way ANOVA; <span class="html-italic">p</span> = 0.0464; WT versus Slack<sup>-/-</sup>) Bars denote mean ± SEM. * <span class="html-italic">p</span> ˂ 0.05.</p>
Full article ">Figure 2
<p>Slack-mediated potassium currents in sensory neurons are reduced after SNI. (<b>A</b>,<b>B</b>) Representative outward K<sup>+</sup> current (I<sub>K</sub>) traces (<b>A</b>) and associated current-voltage (I-V) curves (<b>B</b>) from whole-cell voltage recordings on IB4-positive lumbar (L4-L5) dorsal root ganglion (DRG) neurons of WT (black) and Slack<sup>-/-</sup> mice (red) 14–19 days after spared nerve injury (SNI). Contralateral DRG neurons were used as control. Recordings shown in (<b>A</b>) and (<b>B</b>) were performed in the presence of 140 mM NaCl in the external solution, i.e., under physiological conditions. <span class="html-italic">n</span> = 21–29 cells per group. Repeated ANOVA measures followed by Fisher’s Least Significant Difference test; WT control versus WT SNI: <span class="html-italic">p</span> = 0.0062; Slack<sup>-/-</sup> control versus Slack<sup>-/-</sup> SNI: <span class="html-italic">p</span> = 0.4374. (<b>C</b>,<b>D</b>) Representative I<sub>K</sub> traces (<b>C</b>) and associated I-V curves (<b>D</b>) in the same experimental setting as shown in (<b>A</b>) and (<b>B</b>), however, after replacement of NaCl by 140 mM choline chloride in the external solution to obtain Na<sup>+</sup>-free conditions. <span class="html-italic">n</span> = 7–9 cells per group. Repeated ANOVA measures: WT control versus WT SNI: <span class="html-italic">p</span> = 0.1825; Slack<sup>-/-</sup> control versus Slack<sup>-/-</sup> SNI: <span class="html-italic">p</span> = 0.6125. The data show that Na<sup>+</sup>-activated I<sub>K</sub> (I<sub>KNa</sub>) in sensory neurons is carried by Slack channels and reduced after SNI. Data in (<b>B</b>) and (<b>D</b>) are mean ± SEM. * <span class="html-italic">p</span> ˂ 0.05.</p>
Full article ">Figure 3
<p>Unaltered Slack expression in WT mice after peripheral nerve injury. (<b>A</b>,<b>B</b>) Double-labeling immunostaining of Slack and the neuronal marker anti-βIII-tubulin (TUBB3) in DRGs of naive WT mice and 14 days after SNI. Colocalization of Slack and TUBB3 appears in yellow. The staining suggests that Slack is exclusively localized to neurons and that its distribution is not altered in response to the injury. The absence of Slack immunoreactivity in DRGs of Slack<sup>-/-</sup> mice confirms the antibody specificity (<b>A</b>). The percentage of Slack-immunoreactive DRG neurons of all βIII-tubulin-stained neurons is similar in naive mice and 14 days after SNI ((<b>B</b>); 1833 and 1630 cells counted, respectively; <span class="html-italic">n</span> = 3 mice per group). Student’s t-test: <span class="html-italic">p</span> = 0.9130. (<b>C</b>), Quantitative RT-PCR experiments showed that Slack mRNA levels are not altered in DRGs 7 or 14 days after SNI as compared to naive control animals (<span class="html-italic">n</span> = 8 mice per group). One-way ANOVA: <span class="html-italic">p =</span> 0.1433. (<b>D</b>), Immunostaining in the lumbar spinal cord of WT mice 14 d after SNI shows similar Slack expression in the ipsilateral and contralateral dorsal horn. The absence of Slack immunoreactivity in the spinal cord of Slack<sup>-/-</sup> mice confirms the antibody specificity. (<b>E</b>,<b>F</b>) A Western blot of spinal cord extracts shows similar Slack protein (140 kDa) expression in naive mice and 7 or 14 days after SNI surgery (<b>E</b>). Uncropped original image is shown in <a href="#app1-ijms-22-00405" class="html-app">Figure S2A</a>. Quantification is shown in (<b>F</b>). Alpha-tubulin was used as a loading control. One-way ANOVA: <span class="html-italic">p =</span> 0.4446. Bars denote mean ± SEM. Scale bars: 50 µm (<b>A</b>), 200 µm (<b>D</b>).</p>
Full article ">Figure 4
<p>Slack channels co-localize with P2X3 receptors in sensory neurons. (<b>A</b>–<b>C</b>) Double-labeling immunostaining of Slack and P2X3 in sensory neurons (<b>A</b>) revealed that 97.1% ± 0.2% of Slack-positive DRG neurons co-stained with P2X3 ((<b>B</b>); 1203 cells counted, <span class="html-italic">n</span> = 3 mice) and that 94.6% ± 2.0% of P2X3-positive DRG neurons co-stained with Slack ((<b>C</b>); 1203 cells counted, <span class="html-italic">n</span> = 3 mice). (<b>D</b>) Double-labeling immunostaining of Slack and P2X3 in the spinal cord indicates a high degree of co-localization in the superficial dorsal horn. (<b>E</b>,<b>F</b>) Western blot of P2X3 in spinal cord (SC) and DRGs from WT and Slack<sup>-/-</sup> mice demonstrates identical abundance of P2X3 in both genotypes. The uncropped original image is shown in <a href="#app1-ijms-22-00405" class="html-app">Figure S2B</a>. Student’s t-test: <span class="html-italic">p</span> = 0.5986 in the spinal cord and <span class="html-italic">p</span> = 0.7631 in DRGs. Alpha-tubulin was used as a loading control. (<b>G</b>) Immunostaining revealed that the percentage of DRG neurons positive for P2X3 is similar in WT and Slack<sup>-/-</sup> mice. Student’s t-test: <span class="html-italic">p</span> = 0.4046. Bars denote mean ± SEM. Scale bars: 50 µm (<b>A</b>) and 100 µm (<b>D</b>).</p>
Full article ">Figure 5
<p>Slack-mediated potassium currents are altered by P2X3 activation in vitro. Representative I<sub>K</sub> traces from whole-cell voltage-clamp recordings on HEK-Slack and HEK-Slack-P2X3 cells are shown. The current traces presented in (<b>A</b>–<b>C</b>) were recorded in the presence of 2 mM CaCl<sub>2</sub> in the external solution (<span class="html-italic">n</span> = 9–10 cells per group), whereas those depicted in (<b>D</b>–<b>F</b>) were recorded after the replacement of CaCl<sub>2</sub> by MgCl<sub>2</sub> (<span class="html-italic">n</span> = 5–8 cells per group). Experiments were performed without (control) or with the addition of the P2X3 agonist α,β-methylene ATP (α,β-meATP; 30 µM) to the external solution. Note that the P2X3 agonist exerted opposite effects in HEK-Slack-P2X3 cells dependent on the Ca<sup>2+</sup> concentration: α,β-meATP reduced I<sub>K</sub> in the presence of Ca<sup>2+</sup>, whereas it increased I<sub>K</sub> under Ca<sup>2+</sup>-free conditions. Data are shown as mean ± SEM. Student’s t-test, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>P2X3-mediated Ca<sup>2+</sup> traces are normal in sensory neurons of Slack<sup>-/-</sup> mice. (<b>A</b>) Representative examples of Fura-2 ratiometric Ca<sup>2+</sup> traces in DRG neurons of WT and Slack<sup>-/-</sup> mice evoked by α,β-meATP (30 µM, 30 s application) and KCl (75 mM, 20 s application). Responses to KCl were used to test neuron viability. Experiments were performed in lumbar (L4-L5) DRG neurons (<span class="html-italic">n</span> = 402–518 neurons per group). (<b>B</b>,<b>C</b>) Quantification of the magnitude of the Ca<sup>2+</sup> response to α,β-meATP stimulation with percentage above baseline ((<b>B</b>) <span class="html-italic">p</span> = 0.7858) and ratio peak ((<b>C</b>) <span class="html-italic">p</span> = 0.7858). (<b>D</b>) Quantification of the percentage of responsive neurons to α,β-meATP stimulation (<span class="html-italic">p</span> = 0.5966). The data show that α,β-meATP-evoked Ca<sup>2+</sup> responses are similar in DRG neurons from Slack<sup>-/-</sup> and WT mice. Bars denote mean ± SEM and circles show data from each neuron in (<b>B</b>) and from each mouse in (<b>D</b>). Student’s t-tests were performed.</p>
Full article ">Figure 7
<p>P2X3-dependent nocifensive behavior is altered in Slack<sup>-/-</sup> mice. Paw-licking responses induced by paw injection of drugs are shown. (<b>A</b>) The immediate paw licking in the first 2 min after injection of α,β-meATP (12 nmol) is increased in naive Slack<sup>-/-</sup> mice compared with WT mice (<span class="html-italic">n</span> = 7/genotype; <span class="html-italic">p</span> = 0.0398 in 0–2 min; <span class="html-italic">p</span> = 0.9800 in 2–6 min; <span class="html-italic">p</span> = 0.9352 in 6–10 min). (<b>B</b>) No significant differences between groups occurred when the P2X3 receptor antagonist AF353 (70 nmol intraplantar) was injected 10 min prior to α,β-meATP (<span class="html-italic">n</span> = 5–7/genotype; <span class="html-italic">p</span> = 0.9664 in 0–2 min; <span class="html-italic">p</span> = 0.9732 in 2–6 min; <span class="html-italic">p</span> = 0.7546 in 6–10 min). (<b>C</b>) Paw-licking responses after injection of the vehicle were comparable in Slack<sup>-/-</sup> and WT mice (<span class="html-italic">n</span> = 6/genotype; <span class="html-italic">p</span> = 0.9999 in 0–2 min; <span class="html-italic">p</span> = 0.8934 in 2–6 min; <span class="html-italic">p</span> = 0.8709 in 6–10 min) and similar to the licking behavior after combined injection of α,β-meATP and AF353 (<b>B</b>). (<b>D</b>) When α,β-meATP (12 nmol) was injected in the ipsilateral hind paw after SNI, the paw-licking response persisted over the 10 min observation period in WT mice. In Slack<sup>-/-</sup> mice, the paw licking was increased in the first 2 min and decreased from 6–10 min compared with WT mice (<span class="html-italic">n</span> = 16/genotype; <span class="html-italic">p</span> = 0.0262 in 0–2 min; <span class="html-italic">p</span> = 0.9993 in 2–6 min; and <span class="html-italic">p</span> = 0.0362 in 6–10 min). Two-way ANOVA tests were performed. Bars denote mean ± SEM. * <span class="html-italic">p</span> ˂ 0.05.</p>
Full article ">Figure 8
<p>Proposed model demonstrating the functional coupling of Slack channels and P2X3 receptors in IB4-positive sensory neurons. Activation of P2X3 receptors by ATP may lead to an influx of Na<sup>+</sup>, Ca<sup>2+</sup>, or both into sensory neurons. (<b>A</b>) P2X3-mediated Na<sup>+</sup> influx activates Slack, which results in K<sup>+</sup> efflux and thus partial inhibition of neuronal activity. (<b>B</b>) P2X3-mediated Ca<sup>2+</sup> influx inhibits Slack, which leads to increased neuronal activity due to the lack of inhibition.</p>
Full article ">
Back to TopTop