Next Article in Journal
Synthesis and Structural Characterization of Selenium Nanoparticles–Bacillus sp. MKUST-01 Exopolysaccharide (SeNPs–EPS) Conjugate for Biomedical Applications
Next Article in Special Issue
ClC-1 Chloride Channel: Inputs on the Structure–Function Relationship of Myotonia Congenita-Causing Mutations
Previous Article in Journal
The Efficacy of Anthropometric Indicators in Predicting Non-Alcoholic Fatty Liver Disease Using FibroScan® CAP Values among the Taiwanese Population
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Signaling Roleplay between Ion Channels during Mammalian Sperm Capacitation

1
Institute of Biotechnology, Faculty of Biotechnology and Food Sciences, Slovak University of Agriculture in Nitra, Tr. A. Hlinku 2, 949 76 Nitra, Slovakia
2
AgroBioTech Research Centre, Slovak University of Agriculture in Nitra, Tr. A. Hlinku 2, 949 76 Nitra, Slovakia
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(9), 2519; https://doi.org/10.3390/biomedicines11092519
Submission received: 30 July 2023 / Revised: 31 August 2023 / Accepted: 6 September 2023 / Published: 12 September 2023
(This article belongs to the Special Issue Novel Insight into Ion Channel and Ion-Related Signaling)

Abstract

:
In order to accomplish their primary goal, mammalian spermatozoa must undergo a series of physiological, biochemical, and functional changes crucial for the acquisition of fertilization ability. Spermatozoa are highly polarized cells, which must swiftly respond to ionic changes on their passage through the female reproductive tract, and which are necessary for male gametes to acquire their functional competence. This review summarizes the current knowledge about specific ion channels and transporters located in the mammalian sperm plasma membrane, which are intricately involved in the initiation of changes within the ionic milieu of the sperm cell, leading to variations in the sperm membrane potential, membrane depolarization and hyperpolarization, changes in sperm motility and capacitation to further lead to the acrosome reaction and sperm–egg fusion. We also discuss the functionality of selected ion channels in male reproductive health and/or disease since these may become promising targets for clinical management of infertility in the future.

Graphical Abstract

1. Introduction

Ion channels play a significant role in the regulation of membrane potential by maintaining intracellular pH (pHi), osmotic balance as well as sperm physiological responses associated with fertilization such as hyperactivated motility, capacitation, chemotaxis, and acrosome reaction. An increase in pHi and alkalization of sperm cytosol is essential for hyperpolarization of the plasmatic membrane and later hyperactivation of sperm motility via a Ca2+-dependent pathway. Membrane potential (Em) or resting membrane potential (Emr) reflects different concentrations of ions (mmol/L) between intra- and extra-cellular spaces [1,2,3]. During capacitation, sperm cells change membrane potential by the process called hyperpolarization, which increases a negative charge of membrane through to reduced permeability for Na+ and increased permeability for K+. In non-capacitated mammalian spermatozoa, the value of Em varies from −35 to −45 mV while the Em of capacitated cells is around −65 mV [4,5].
Sperm ion channels are pore-forming proteins which can be found in the whole surface of the cell including principal piece or midpiece of flagellum and head (Figure 1). They are classified based on their opening or closing into voltage-gated and ligand-gated ion channels [6]. The regulation of voltage-gated ion channels depends on the voltage gradient across the plasmatic membrane, charge (cation/anion) or species of ions (Na+, Ca2+, H+, Cl, K+). On the other hand, the activity of ligand-gated channels is managed through to the specific bind of primary signaling transmitters such as cyclic nucleotides [7,8]. The ability of spermatozoa to undergo capacitation depends on numerous factors like membrane potential, pH homeostasis and balanced ion environment. Mutual cooperation between ion channels, pumps and transporters is required for proper sperm motility. From all ions, Ca2+ is one of the most crucial because its higher concentrations (100–300 nM) start the hyperactivation of spermatozoa. In mammals, Ca2+ participate in the activation of the signaling pathways of capacitation as a secondary messenger. Before capacitation itself, sperm cells received specific signals from environment in the female genital tract. It all starts with the increase in intracellular pH and uptake of bicarbonate, which stimulate the sAC/cAMP/PKA pathway as we mention below. From the point of view of ion channel activity, intracellular alkalinization managed the activity of CatSper and KSper channels, which is required for capacitation [9,10]. In this paper, we provide an overview of the most important ion channels occurring in spermatozoa and analyze their involvement in sperm activation following their entry into the female reproductive system, including the process of capacitation, hyperactivation and acrosome reaction. The review methodology is available in Supplementary Material.

2. Bicarbonate Transporters

Bicarbonate transporters or acid extruders (Table 1) can maintain the intracellular pH homeostasis by transporting HCO3, which induce phosphorylation of functional flagellar proteins in serine, threonine, and tyrosine residues through to the sAC/cAMP/PKA pathway. Soluble adenylate cyclase (sAC) catalyzes the synthesis of cAMP, which is sensitive for higher concentration of HCO3. In general, HCO3 ions are responsible for (1) the initiation of sperm motility right after ejaculation and (2) the activation of sperm capacitation in the female reproductive tract. The range of pH in the seminal fluid is between 7.2 and 8.4; it is believed that seminal fluid works as a buffer which controls the acidic environment of the vagina. Bicarbonate membrane transporters are represented by two major protein groups, solute carrier 4 (SLC4) and solute carrier 26 (SLC26) [11,12,13].
The principal group of the SLC4 family can be divided based on affinity to Na+-independent and Na+-dependent HCO3 exchangers (Na+/HCO3 cotransporters—NBCs). Na+-independent transporters maintain electroneutral exchange of Cl into HCO3, which is ensured by three anion exchangers (AE) SLC4A1 (AE1), SLC4A2 (AE2), SLC4A3 (AE3) and two Na+-coupled exchangers SLCA48 (NDCBE), SLCA49 (AE4). The subfamily of Na+-dependent HCO3 exchangers includes two electrogenic SLC4A4 (NBCe1), SLC4A5 (NBCe2) and two electroneutral SLC4A7 (NBCn1), SLC4A10 (NCBE) exchangers. The functionality of electroneutral transporters depends on the chemical gradient between intra- and extra-cellular space compared to electrogenic transporters, which depends on the negative electrical potential of the membrane [14,15,16]. The main core of the SLC4 channel is made from 14 loop-connected transmembrane α-helices (TM1–14) and amphipathic helices (H1–6) which form the core domain (TM1–4/TM8–11) and the gate domain (TM5–7/TM12–14). The active site of SLC4 is localized on the N-terminus of half-helices TM3 and TM10 while the blocker of the active side takes place at segment TM8 [17]. Several sources confirm that NBC channels, especially electrogenic ones, play an essential part in sperm capacitation, hyperactivated motility and membrane hyperpolarization through the initial fast increase in bicarbonate, which is necessary for cAMP/PKA pathway activation and later redistribution of cholesterol [3,18].
A group of SLC26 contains 11 electrogenic anion channels, but only 5 of them are able to transport HCO3, represented by SLC26A3, SLC26A4, SLC26A6, SLC26A7 and SLC26A9. Like their cousins, the SLC26 channels are made from 14 transmembrane α-helices connected by loops, which occasionally contain amphipathic helices (H). The active site is also in the N-termini of TM3 and TM10 half-helices, but the C-termini region contains STAS (sulphate transporter anti-sigma factor antagonist domain) or the dimerization domain, involved in the expression and protein interaction, which can interact with the regulatory R-domain of the CFTR (cystic fibrosis transmembrane conductance regulator) channel. TM1–4 together with TM8–11 helices form the core domain, while TM5–7 and TM12–14 represent the gate [19,20]. The functional interaction between SLC26 and CFTR channels in capacitated spermatozoa regulates and maintains the high bicarbonate entrance, the sAC/PKA pathway and acrosomal exocytosis together with the ions of Ca2+. On the other hand, the inhibition of SLC26, especially SLC26A3 and SLC26A6 localized in the midpiece of the flagellum, can cause a decrease in the Cl influx and a blockage of intracellular alkalization and membrane hyperpolarization [21,22].
Table 1. Bicarbonate transporters.
Table 1. Bicarbonate transporters.
ChannelSpeciesLocalizationFunctionalityReferences
SLCA26A3/A6human, mice, guinea pigsperm plasma membrane, acrosomal region of sperm headHCO3 transport, pHi alkalinization, protein phosphorylation, hyperactivation, CFTR channel interactions[19,23,24]
SLCA26A8human, miceequatorial segment of spermregulation of sperm motility and acrosome exocytosis during capacitation, CFTR channel interactions[25]
SLC4A1human, micesperm head and flagellumactin depolymerization and regulation of acrosome reaction[26]

3. Sodium Channels

Overall, spermatozoa are exposed to high sodium concentrations in the female reproductive tract, which directly regulate sperm membrane potential and electrogenic Na+/K+ ATPase gradient pump. An increased Na+ influx has a great influence on membrane polarization and improves linear sperm motility. The presence of voltage-gated Na+ channels (VGNCs) was confirmed in the human and bovine sperm cells (Table 2). Their activation depends on the depolarization of the sperm plasma membrane and the conduction of sodium ions into the cell, which directly supports the action potential [27,28].
Based on the structure, the core of the VGNC channel includes a big α subunit made from four repeat homologous domains (RD1–RD4) with six transmembrane helices (TM1–TM6) as well as one or more auxiliary β subunits (β1–β4), each consisting of a big extracellular N-terminal domain and an intracellular tail depending on the isoform. Furthermore, TM4, known as a positively charged sensor, manages the channel ion permeability by moving into the extracellular space and Na+ is transported via the pore P-loop between TM5 and TM6. Repeat domains play an important role in the PKA/PKC protein phosphorylation due to the connection with long intracytoplasmic loops. The level of protein phosphorylation of RD loops regulates the inactivation of the channel. The members of the VGNC family were localized particularly in the sperm flagellum (NaV1.2, NaV1.6, NaV1.8 and NaX) and the connecting piece (NaV1.4, NaV1.7 and NaV1.9) due to the immunostaining of human spermatozoa. In the case of NaV1.8, its presence was confirmed in bull and ram sperm. These channels have the ability to maintain progressive motility rather than hyperactivation or acrosome reaction [29,30,31].
The electrogenic transport of Na+ as well as membrane potential in spermatozoa can be also regulated by heteromultimeric epithelial sodium selective channels (ENaCs). They come from a superfamily of non-voltage-gated and amiloride-sensitive DEG/ENaC ion channels [32]. The channel itself is formed from four subunits (α, β, γ and δ), where α and δ are characterized as pore forming subunits which can be found in the midpiece or the principal piece of the sperm flagellum (ENaC-α) and the acrosome region (ENaC-δ). The activity of ENaC is controlled by intracellular pH, Ca2+, Cl, phosphorylation or amiloride. Previous findings support the involvement of this channel in sperm movement and regulation of sperm resting potential, which becomes more negative during capacitation-associated hyperpolarization [33,34,35].
Table 2. Sodium channels.
Table 2. Sodium channels.
ChannelSpeciesLocalizationFunctionalityReferences
VGNC, Navhuman, bullsperm connecting piece and flagellummaintaining of progressive motility, tyrosine phosphorylation[27,28]
ENaChumancentral region of sperm flagellum, acrosomeregulation of sperm motility and sperm resting potential during capacitation, activator of CFTR channel, membrane hyperpolarization[33]

4. Calcium Channels

As secondary intracellular messengers of capacitation, ions of calcium (Ca2+) participate in acrosome reaction, tyrosine phosphorylation, modulation of the cAMP-dependent pathway and maintenance of functionality of mitochondria as well as synthesis of ATP, which is necessary for hyperactivated motility [7]. The concentration of Ca2+ can be regulated by alcium channels (Table 3), transporters, or exchangers like CatSper (cation channel of sperm), VGCCs (voltage-gated Ca2+ channels), TRPVs (transient receptor potential vanilloids), SOCCs (store-operated Ca2+ channels) and CNGs (cyclic nucleotide-gated channels) [36]. Intracellular Ca2+ levels are also regulated by Ca2+ pumps of the plasma and outer acrosomal membranes, as well as mitochondrial transporters.
In general, the heterotetrameric complex of the CatSper channel is composed from four main pore-forming alpha subunits (CatSper 1–4) and six minor ancillary subunits CatSper β (beta), γ (gamma), δ (delta), ε (epsilon), ζ (zeta) and EFCAB9 (calcium-binding domain-containing protein 9). The activity of these channels is related to hyperactivation, sperm chemotaxis and thermotaxis as well as acrosome reaction. Every α subunit is made from six transmembrane domains (TM 1–6), which form a voltage-sensing domain (TM1–4) containing voltage sensors and a pore-forming region (TM 5–6), which particularly coordinate the Ca2+ influx. The CatSper channel can be found in the principal piece of the sperm flagellum in the form of four-sided longitudinal nanodomains responsible for sperm motility [37]. However, the principal piece of flagellum is missing organelles, so there is the theory that the CatSper complex is part of the membrane in the principal piece, and it is involved in the regulation of flagellar movement [38]. The pH sensible small voltage-gated CatSper channel is primary activated by the intracellular alkalization and the pH is regulated via a special histidine-rich region of the N-terminus part of CatSper1, but its activity can be also controlled by cyclic nucleotides, phosphorylation, progesterone, prostaglandins, glycoproteins of zona pellucida or bovine serum albumin as part of the oviductal fluid [39,40]. According to Hwang et al. [41], the cytoplasm complex CatSper ζ together with the EFCAB9 subunit work as gatekeepers and activity modulators of the other CatSper domains. The genes, which coded the functions of CatSper channels, are expressed in the testis during spermatogenesis.
Another major functional subclass of Ca2+ permeable channels are voltage-gated Ca2+ channels (VGCC’s) or voltage-sensitive Ca2+ channels (CaV), which can be divided into high- and low-voltage-activated (HVA/LVA) channels localized in the sperm tail. Their activation depends on the changes in membrane potential and strong/low depolarization. Regulatory mechanisms of CaV channels could be activated by protein kinases (protein kinases A/C, Ca2+/calmodulin (CaM)-dependent protein kinase II) or ions of Ca2+ itself. Similar to CatSper channels, VGCCs contain four main transmembrane domains (1–4), but every central pore is surrounded by six transmembrane α helixes (S1–6). The α1 subunit of the central pore is encoded at least 10 genes from three subfamilies CaV1–3. Based on their different physiological properties, several types of VGCCs were identified in the spermatozoa, including the L-type (CaV1.1–1.4), T-type (CaV3.1–3.3), R-type (CaV2.3), and P/Q-type (CaV2.1), which are involved in motility hyperactivation, capacitation, acrosome reaction as well as phosphorylation of the cAMP-dependent protein kinase A [42,43].
Transient receptor potential vanilloid channels (TRPVs) belong to a big branch of cation polymodal voltage-gated and Ca2+ permeable cellular channels (TRPV1–6), which are activated by the increased intracellular Ca2+ concentration via acrosomal inositol triphosphate (IP3) receptors or phospholipase C through receptor-mediated messenger phosphatidylinositol 4,5-bisphosphate (PIP2). These channels are responsible for the regulation of sperm functionality at different levels like basal and hyperactivated motility, thermotaxis, or acrosome reaction [44,45]. Interestingly, TRPV channels work as modulators of the Ca2+ signaling pathway of other Ca2+-permeable channels, which include regulations of the cytosolic cation flux and electrical activity because of the unique gate mechanisms and a wide range of ion selectiveness. Based on their structure, TRPV channels consist of six transmembrane segments (S1–6) and a pore-creating intracellular loop between the S5 and 6 segments [46,47].
Non-voltage-dependent store-operated Ca2+ channels (SOCCs) can operate under a negative membrane potential when VGCCs stay inactive. Their basal structure is formed by ORAI proteins (ORAI1–3) found in the sperm head and flagellum where each protein contains four TMs which create a pore (between TM2 and TM3)- and Ca2+-binding domain (CBD) localized in the center of the pore. This domain basically regulates the activity of the channel by binding Ca2+ or CaM, which inactivate the channel. The main role of the SOCC channel is to bind and store the extracellular Ca2+ in sperm mitochondria as well as regulate the motility and the acrosome reaction [48,49,50]. SOCC is activated when the intraacrosomal Ca2+ levels are reduced due to Ca2+ efflux from the acrosome via the inositol trisphosphate receptor (IP3-R). The activity of SOCCs can be inhibited with the induction of AMPK (5′ AMP-activated protein kinase) phosphorylation, which leads into reduction in the asymmetrical flagellar beating necessary for chemotaxis [51].
Cyclic nucleotide-gated cation channels (CNGs) are ligand voltage-gated channels, which use the free binding energy of second messengers like cAMP (cyclic adenosine monophosphate) or cGMP (cyclic guanosine monophosphate) for the regulation of pore opening [8,52]. Their structure is made from a heterotetrameric complex including homologous A (CNGA1–4) subunits, which defines principal channel properties and B (CNGB1/3) subunits necessary for gating kinetics. Similar to other voltage-gated ion channels, A as well as B subunits contain six transmembrane α helices (S1–6) and an ion-selective pore loop between S5 and S6. A cyclic nucleotide-binding domain (CNDB) formed at the cytosolic C-terminus of a channel is responsible for its activation [53]. In mammalian spermatozoa, CNG channels are important modulators of motility, capacitation, and acrosome reaction because of high permeability to Ca2+ and cGMP. The activity of these channels can be inhibited by Mg2+ or Ca2+ itself by binding to calmodulin, which also acts as a voltage-dependent blocker of permeability for monovalent cations including Na+ and K+ [54].
Table 3. Calcium channels.
Table 3. Calcium channels.
ChannelSpeciesLocalizationFunctionalityReferences
CatSpermice, human, boar, bull, sea urchinprincipal piecepromotion of hyperactivated motility, sperm chemotaxis and thermotaxis, late acrosome reaction, Ca2+ uptake, alkalinization[55,56,57]
VGCC, Cavnewt, marine fish, equine, mice, bull, humansperm neck and tailregulation of total and progressive motility, hyperactivation, capacitation, phosphorylation of protein kinase A, acrosome reaction[58,59,60,61]
TRPVvertebrateswhole surface of spermatozoaactivation of basic and hyperactivated motility, capacitation, membrane depolarization, opening of other channels (CatSper, Hv1)[45,62,63]
SOOCmice, chicken, ascidiansperm head and flagellumregulation of sperm motility and acrosome reaction, induction of 5′ AMP-activated protein kinase (AMPK) phosphorylation[50,51,64]
CNGmammals, sea urchinflagellum of spermatozoaeffectors for CNG induced Ca2+ response, sperm hyperactivation[65]

5. Proton Channels

A specific group of channels involved in carrier-mediated proton transport are combined membrane sodium–hydrogen exchangers/antiporters (NHE’s) and the Hv1 voltage-gated ion channel. NHEs are membrane proteins encoded by the solute carrier 9 (SLC9) gene family which transport Na+ into the cell and H+ out of the cell across the lipid bilayer accompanied with the maintenance of intracellular pHi and Na+ homeostasis (Table 4). Several members of NHEs were identified, especially in spermatozoa. NHE1 and NHE10/sNHE (sperm-specific NHE isoform with a binding site for cAMP) expressed in the principal piece of the sperm flagellum are important for normal sperm motility and capacitation, while NHE8 ensures the formation of the acrosome. Knock-out of any of these NHEs results in male infertility characterized by a lower expression of sAC and intracellular cAMP, which confirms the mutual relationship with the cAMP signal pathway [66,67]. NHEs are formed from 12 TM helices connected together with six extracellular (EL1–6) and five intracellular loops (IL1–5). The channel also contains cytosolic N- and C-terminal domini with an extracellular N-linked glycosylation site [68].
The Hv1 pH-sensitive channel is co-localized together with other flagellar channels like CatSper and KSper in the principal piece, and it is activated by membrane depolarization, alkaline changes in the extracellular environment or the removal of zinc, which works as a potential blocker of Hv1 activity. Based on their similar subcellular location, proton exchange through Hv1 promotes intraflagellar alkalinization and stimulates CatSper channels in human or bovine spermatozoa compared to murine spermatozoa, which lacks the presence of Hv1; its role in pH regulation is replaced by an Na+-dependent Cl/HCO3 exchanger or sNHE. The synergy pathway between the Hv1/CatSper channels elevates intracellular pH (5.5–6.5) as well as the Ca2+ uptake, which is necessary for the activation of pH-dependent axonemal proteins and the maintenance of sperm motility, chemotaxis, capacitation, and later acrosome reaction [61,69,70]. The architecture of an Hv1 dimeric channel complex resembles voltage-gated channels, with the presence of voltage sensor domain (VSD) but missing a separate pore domain. Instead of that, the Hv1 channel contains internal selective proton transfer water wire located in the center of VSD [71].
Table 4. Proton channels.
Table 4. Proton channels.
ChannelSpeciesLocalizationFunctionalityReferences
NHE1human, ram, mice, rat, boarmid- and principal piece of sperm flagellumactivation of sperm motility upon ejaculation, co-activation of Ca2+ and SLO channels[72,73,74]
NHE10/sNHEhuman, miceprincipal piece of flagellumregulation of cAMP signal pathway, hyperactivated motility, sAC regulation[66,75]
Hv1human, bull, boarprincipal and terminal part of sperm tail close to CatSperpromotion of fast intracellular alkalinization via pHi regulation, synergic pathway with CatSper channels for the regulation of capacitation and late acrosome reaction[3,61,76]

6. Potassium Channels

The primary role of potassium (K+) channels (Table 5) in male gametes is the hyperpolarization of the plasma membrane during the capacitation process, which is a crucial step for the sperm motility hyperactivation [7]. The opening of K+ channels leads to changes in the membrane potential in the capacitated sperm [4], resulting in membrane hyperpolarization, whilst the closure of these channels causes its depolarization. The first indication of their existence was reported by Arnoult et al. [77], who observed that membrane hyperpolarization that accompanies sperm capacitation is affected by the external K+ concentration and K+-channel blockers, assuming that the hyperpolarization process can be executed by the opening of these channels.
Four major classes of K+ channels are currently recognized, specifically (1) voltage-gated K+ channels which open or close depending on fluctuations of the membrane potential; (2) ion-activated K+ channels that are stimulated by the presence of Ca2+ or other intracellular signaling molecules; (3) inwardly rectifying K+ channels that transfer K+ more easily into the cell than out of it; (4) tandem pore domain K+ channels which may be constitutively open [78]. At the same time, voltage-gated and ion-activated K+ channels may share overlapping properties. Evidence gathered from previous studies supports the presence of several K+ channels in spermatogenic cells as well as in spermatozoa [79,80,81,82,83,84,85].
The calcium-activated potassium channel (SLO1) and the potassium channel subfamily U member 1 (SLO3) are the most frequently described K+ channels in male gametes and are considered as primary regulating channels of K+ currents. They share characteristics of voltage-gated and ion-activated K+ channels since they may be triggered by membrane depolarization, Ca2+ and Mg2+ [86,87]. Besides the K+ current regulation, SLO channels are involved in the regulation of osmolality and membrane potential of the plasma membrane [88]. The structure of both consists of four pore-forming α subunits and several auxiliary subunits [86,89]. SLO1 channels are present in all multicellular, mitochondrial eukaryotes, primarily in muscle and neural cells, while SLO3 is exclusive to mammalian testes and spermatozoa [83,86,90], although its isoforms may be found in the kidneys, brain, and eyes [91]. Although both channels are primarily defined as voltage gated, SLO1 is also activated by Ca2+ [92], while SLO3 is triggered by intracellular alkalinization [86,93].
SLO3 is the principal K+ channel in mammalian spermatozoa [88,94] and is localized in the principal piece of the sperm flagellum [95]. SLO3 channels are responsible for K+ efflux, and subsequent membrane hyperpolarization, which then affect other voltage-sensitive ion channels such as CatSper and voltage-gated calcium channels (VGCCs) [96]. Additionally, it has been speculated that SLO3 may participate in volume control and the ability of spermatozoa to respond to osmotic challenges during their transit within the female reproductive tract [97]. Because of rapid evolution, the channel presents with high structural divergence and various functional properties amongst mammals, resulting in different voltage ranges for the activation, sensitivity to pH, Ca2+ or phosphatidylinositol 4,5-bisphosphate and subsequent dynamics of SLO3 in different mammalian species [95,98,99].
Male SLO3 knockout mice produce spermatozoa with a reduced motility and abnormal morphology, most likely due to a lack of membrane hyperpolarization, activation of other voltage-gated channels and subsequent disturbance of osmotic homeostasis, since apparently no other channels can compensate for the loss of SLO3 [94]. Although an SLO3 mutant sperm can, to some degree, undergo spontaneous AR, it fails to undergo this exocytotic event when exposed to solubilized zona pellucida (ZP). This result supports the hypothesis that membrane hyperpolarization during capacitation is a key factor required for the induction of the AR [100]. As such, these animals are neither capable of producing offspring even during extended mating periods nor achieve fertilization in vitro [101].
Since K+ concentrations vary significantly in aquatic environments [102], the primary channels responsible for K+ transport and subsequent sperm plasma membrane hyperpolarization in aquatic animals are the cyclic nucleotide-gated K+ channels (CNGKs). Depending on the species, these tetrameric channels may be located at the head [103] or flagellum [104] and are activated by oocyte-derived chemoattractants [7]. The presence of a chemoattractant leads to an increase in cGMP [105], which opens CNGK channels, to trigger membrane hyperpolarization, followed by a continued depolarization [104,105]. Furthermore, CNGK channels play a pivotal role in the induction of sperm motility, as their activation leads to Ca2+ influx carried out by voltage-sensitive Ca2+ channels [103,104,105].
Inward rectifier K+ channels conduct larger inward currents at membrane voltages negative to the K+ equilibrium potential than outward currents at positive voltages, which enables them to be active at negative voltages [106,107]. Two types of inward rectifiers have been identified in male reproductive cells, specifically K+ channels with strong inward rectification and weakly rectifying K+ channels. K+ channels with strong inward rectification are highly selective to K+ and may be inhibited by intracellular acidification and Ba2+, which is a known inhibitor of sperm capacitation and acrosome reaction [108]. Weakly rectifying K+ channels sensitive to ATP (KATP channels) comprise Kir 6.1, Kir 6.2 (subunits of the KATP channel), SUR1 and SUR2 (sulfonylurea receptor) channels which were detected in both spermatogenic cells and mature spermatozoa, specifically in the flagellum principal piece (SUR2) and midpiece (Kir 6.1, Kir 6.2, SUR1, SUR2) as well as in the postacrosomal region of the sperm head (Kir 6.2, SUR1) [79,109]. These channels are particularly sensitive to the KATP channel blockers, tolbutamide and glibeclamide, and the loss of glucose leading to the reduction in ATP [79].
Delayed outward voltage-dependent K+ channels are a family of K+ channels that enable a sustained K+ efflux with a delay following membrane depolarization, which leads to a rapid membrane repolarization. There are two types of delayed rectifiers in spermatogenic cells depending on their sensitivity to tetraethyl ammonium (TEA). The most prominent TEA-sensitive channel found in spermatogenic cells and mature spermatozoa is KV3.1, whilst the predominant delayed rectifier less sensitive to TEA is hypothesized to correspond to the SLO3 K+ channels which have been discussed previously. Other less prominent delayed rectifiers identified in male reproductive cells are KV1.1, KV1.2, and G protein-coupled inwardly rectifying potassium (GIRK1) channels [80].
Table 5. Potassium channels.
Table 5. Potassium channels.
ChannelSpeciesLocalizationFunctionalityReferences
SLO1/3mammals including human, reptiles, birds, fishprincipal piece of sperm flagellumregulation of K+ efflux, osmolality and membrane potential—hyperpolarization[86,90,94]
GNGK’ssea urchin, zebrafishsperm head and flagelluminduction of sperm motility, mediators of voltage sensitive Ca2+ channels[103,106]
KATPmice, ratpostacrosomal region of sperm head, midpiece and principal piece of sperm tailcapacitation-associated hyperpolarization[78,79]
KVmice (Kv1.1, Kv1.2 and Kv1.3), bull (Kv1.1)equatorial segment of spermatozoamembrane hyperpolarization, capacitation and acrosome reaction, maintaining of sperm osmotic resistance[34,110]

7. Chloride Channels

In comparison to previously discussed ion channels and transporters, information on chloride (Cl) channels is relatively sparse (Table 6). Nevertheless, it has been previously reported that sperm capacitation, hyperactivation and fertilization is blocked in media lacking Cl channels, which supports the hypothesis that regulation of Cl homeostasis is necessary for a proper membrane hyperpolarization, tyrosine phosphorylation and cytosolic alkalization [7,111,112]. According to Santi [78], Cl transporters may be divided into two categories: (1) Cl channels located in the plasma membrane with four structural families, specifically γ-aminobutyric (GABA)-gated and related glycine-gated neurotransmitter receptors, cystic fibrosis transmembrane conductance regulator (CFTR), Ca2+-activated Cl channels (CaCCs) and ClC-3 channels; and (2) Cl transporters which enable Cl passage with another ion in either the same or the opposite direction (symporters, cotransporters or antiporters). The most prominent Cl transporters include the electroneutral cation-Cl cotransporters such as the Na+/Cl cotransporter, the Na+/K+/2Cl cotransporter or the Na+-independent K+/Cl cotransporter.
Gamma-aminobutyric acid (GABA) receptors are Cl channels that are most known to mediate inhibitory neurotransmission in the central nervous system. In spermatozoa, GABA receptors have been shown to be involved in the induction of acrosome reaction [113], regulation of sperm motility and hyperactivation [114,115], as well as modulation of the response of male gametes to progesterone [114,115,116,117].
The cystic fibrosis transmembrane conductance regulator (CFTR) is a unique member of the ATP-binding cassette (ABC) transporter family that acts as an ion channel modulated by cAMP/PKA and ATP [118]. The channel is in charge of Cl and HCO3 transport in an electrochemical gradient contrary to other members of the ABC family that transport substrates against their chemical gradients [119]. Structurally, CFTR contains two membrane-spanning domains (MSDs), two nucleotide-binding domains (NBDs), and one regulatory (R) domain. MSDs constitute the channel pore; phosphorylation of the R domain determines the activity of CFTR and ATP hydrolysis by NBDs affects the channel-gating properties [120]. CFTRs have been localized in the midpiece of mammalian spermatozoa [23,111,121] where they are involved in the regulation of sperm motility, cAMP production and membrane hyperpolarization [121]; their specific roles depend on the species. In humans, CFTRs seem to be involved in Cl removal from spermatozoa upon capacitation [33], whilst in mice and guinea pigs, CFTRs are suggested to transport Cl to spermatozoa [21,23]. In addition to its role as an Cl channel, CFTR is also known to interact with, cooperate with and regulate other ion channels such as chloride anion exchanger SLC26A3, a HCO3 transporter of the SLC26 family, or epithelial Na+ channels (ENaC) [122,123,124].
Ca2+-activated Cl channels (CaCCs) are stimulated by increases in intracellular Ca2+ levels caused either by its influx through the plasma membrane or release from intracellular stores. These are anionic channels belonging to the anoctamin family (ANO/TMEM16). Depending on the species, CaCCs may be found in the sperm head (humans), the apical part of the acrosome or the middle piece of the flagella (guinea pigs) [125,126]. Evidence gathered from currently available reports suggests that CaCCs play an important role in the process of capacitation and acrosome reaction, as well as in the regulation of sperm motion, particularly in the acquisition of hyperactivated motility [7,127].
Chloride channels (ClCs) are an evolutionary conserved voltage-gated channel family of nine members found in prokaryotic as well as eukaryotic organisms [128,129]. Out of these, ClC-3, an intracellular voltage-dependent electrogenic 2Cl/H+-exchanger [130,131], has been detected in the sperm flagellum (humans, rhesus monkeys) [90,94], as well as in the acrosome and midpiece (bulls) [132,133]. Chloride channel-3 (ClC-3) regulates outwardly rectifying Cl currents that are inhibited by protein kinase C (PKC) activation [134], chloride channels thus playing important roles in the regulation of sperm volume and motility [135]. At the same time, ClC-3 can bind protein phosphatase PP1γ2, which is crucial for sperm maturation and motility [132]. Accordingly, spermatozoa from asthenozoospermic patients present with a lower cell volume and mobility, which correlates with lower expression levels of ClC-3 [135].
Table 6. Chloride channels.
Table 6. Chloride channels.
ChannelSpeciesLocalizationFunctionalityReferences
CFTRmice, guinea pig, humanequatorial segment of spermatozoatransportation of Cl to (rodents) and out (humans) of spermatozoa, intracellular alkalinization, cAMP synthesis, membrane hyperpolarization, cooperation with SLC26 channels[21,23,112]
CaCChuman, guinea pigsperm head, apical part of the acrosome, middle piece of sperm tailregulation of sperm movement and acquisition of hyperactivation[125,126]
ClChuman, monkey, bullsperm flagella, acrosomeregulation of cell volume, capacity and mobility[131,132,135]
Previous research has unraveled that CFTR inhibitors affect the plasma membrane hyperpolarization without compromising other aspects of capacitation (such as tyrosine phosphorylation), suggesting the presence of other Cl transporters in spermatozoa. Cl may enter the cell with the help of electroneutral carriers; specifically, the sodium–chloride symporter and the sodium–potassium–chloride carriers transport Cl into the cell, while potassium–chloride cotransporters (KCCs) transport Cl out of the cell under physiological conditions [136]. Cl levels have been shown to be increased during capacitation [111,137], indicating that NCC and NKCC may be involved in the regulation of Cl homeostasis during sperm activation and preparation for the physiologically induced acrosome reaction. Further research has revealed the presence of NKCC1 in spermatids, and null mutants of this protein present with a defective spermatogenesis and infertility [138].
Cl may be carried through the plasma membrane with the help of molecules that exchange Cl for HCO3 in either direction. The role of HCO3 in the activation of cAMP synthesis via soluble adenylyl cyclase is undeniable [139,140]; specific HCO3 carriers have not yet been fully defined. Whilst previous research has identified that Na+/HCO3 cotransporters are responsible for initial HCO3 influx into the sperm cell [141], Cl/HCO3 exchangers have been suggested to be involved in the regulation of HCO3 homeostasis. In addition, through their contribution to the Cl gradient, they are important players in the regulation of cell volume, intracellular pH, and membrane potential. The most relevant Cl/HCO3 exchangers include two evolutionary independent gene superfamilies, SLC4 and SLC26, with specific patterns of anion selectivity and tissue distribution. From the SLC4 superfamily, only AE2 is found in testicular germ cells with suggested roles either in the spermatogenic process or later in sperm function. With respect to the SLC26 superfamily, recent studies have identified SLC26A3 and SLC26A6 in the sperm midpiece [21,23,78]. Accordingly, a SLC26A3-specific inhibitor blocked the capacitation-associated membrane hyperpolarization and the ZP-induced acrosome reaction, without affecting the cAMP pathway or tyrosine phosphorylation [78].

8. Aquaporins

Aquaporins (AQPs) are a ubiquitous transmembrane protein family (Table 7) playing pivotal roles in cellular fluid homeostasis, facilitating bidirectional water diffusion across the membrane [142,143]. Structurally, AQPs are formed by four monomers, each with their own permeable pore and with one central pore inside the tetramer whose function is currently unknown [144]. Thirteen AQPs are currently known with different permeability properties, structural features, and localization [145]:
  • Classical or orthodox AQPs present with the smallest channel size and hydrophilic nature. Orthodox AQPs are located in the plasma membrane and are considered primarily selective to water (with the exception of AQP6 present in the membranes of cellular organelles that acts as a selective anion channel). This group comprises AQP0, AQP1, AQP2, AQP4, AQP5, AQP6, and AQP8, considered primarily selective to water [144,146,147].
  • Aquaglyceroporins (GLPs) present with larger pore size and lower hydrophilicity. As such, these proteins are able to permeate glycerol (preferably), arsenite, urea, polyols, purines, or pyrimidines. AQP3, AQP7, AQP9, and AQP10 belong to this group [142,148,149].
  • Nonorthodox AQPs or superaquaporins (superAQPs) include AQP11 and AQP12. These are expressed in intracellular membranes, primarily in the endoplasmatic reticulum. While it has been reported that superAQPs are involved in the transport of water and glycerol, their specific pore selectivity and function are currently unknown [142,150].
All AQPs, except for AQP6 and AQP12, have been found in different locations of the male reproductive system as well as in spermatozoa, out of which AQP3, AQP7, AQP8 and AQP11 are the most prominent ones [144,151]. Their specific location by and large differs among species. AQP3 is present in the sperm mid-piece in bulls and boars [152,153], while in human and murine sperm, the principal piece is its prime location [154,155]. AQP7 is present in the tail of ejaculated spermatozoa in bulls, stallions, boars, mice, and rats, as well as in certain regions of the sperm head of men [153,156,157,158,159,160]. AQP8 has been detected in the tail of human, mouse, and rat spermatozoa, additionally to the mitochondria from the mid-piece in men [159,161]. Finally, AQP11 is present in the intracellular tail structures of human, boar, stallion, mouse, and rat spermatozoa. In the meantime, its presence was also confirmed in the sperm head of humans and the terminal piece of the sperm tail in rats [157,162,163,164].
The most important role of AQPs in male gametes is related to osmoregulation to counteract hypoosmotic stress that spermatozoa must withstand upon entering the female reproductive tract [142]. The resulting hypoosmotic shock causes spermatozoa to uptake excessive amounts of water, leading to swelling, membrane ruptures and the loss of proper movement of the sperm tail [143]. Under physiological conditions, hypoosmotic stress triggers an osmolyte efflux that drives a rapid water trafficking via AQPs and restores cell volume [76]. Nevertheless, osmotic changes that occur in response to the hypotonic shock that spermatozoa experience are crucial to initiate the capacitation process. Changes in the sperm volume trigger the opening of calcium channels which enable calcium influx, which is the first event occurring during capacitation [144]. At the same time, acrosomal swelling is an essential prerequisite for a physiological acrosome reaction [145]. In the meantime, evidence from recent years points to the necessity of strictly regulated hydrogen peroxide (H2O2) levels, which are essential for sperm capacitation, hyperactivation and acrosome reaction [146]. Since some AQPs (particularly AQP8) have been suggested to be involved in the diffusion of H2O2, their possible involvement in ROS-mediated sperm activation seems plausible. Accordingly, experimental blockade of AQPs has been reported to impede ROS detoxification, leading to excessive intracellular ROS accumulation with an inhibitory effect on the sperm capacitation accompanied by plasma membrane ruptures, low membrane hyperpolarization and premature acrosome exocytosis. Finally, the permeability of AQPs to glycerol has been revealed to be important for the use of this molecule in metabolic pathways and as a source of energy [147].
Table 7. Aquaporins.
Table 7. Aquaporins.
ChannelSpeciesLocalizationFunctionalityReferences
AQP3bulls, boars, human, murinemidpiece and principal piece of sperm tailosmotic balance as a defense to hypoosmotic stress, restoring of cell volume for Ca2+ influx and triggering of capacitation process, diffusion of H2O2[153,156,157,158,160,161,163,164,165,166,167,168,169,170]
AQP7bulls, stallions, boars, mice, rats, humansperm flagella and head (human)
AQP8boars, stallions, mice, rats, humanspermatozoa tail
AQP11boars, stallions, rodents, humanintracellular tail structure, terminal piece (rats) and sperm head (human)

9. Ion Channels Relevant to the Sperm Redundant Nuclear Envelope (RNE)

The redundant nuclear envelope (RNE) is defined as a residual nuclear membrane that accumulates at the sperm neck due to nuclear condensation. This membrane is considered as a continuum of the membrane covering the endoplasmic reticulum of the sperm cell before entering spermiogenesis, and occasionally RNE carries the remains of a functional endoplasmic membrane [171]. From a functional point of view, RNE has been proposed to contribute to the generation of Ca2+ signals necessary for sperm activation [172]. Accordingly, Ho and Suarez [171] unraveled the presence of a receptor-operated Ca2+ channel (IP3R) as well as a Ca2+-binding and storage protein called calreticulin in the region occupied by RNE. Subsequent functional experiments have revealed that IP3R mobilizes Ca2+ in the sperm neck, leading to an efflux of Ca2+ from RNE to trigger hyperactivated motility [173,174]. Moreover, Naaby-Hansen et al. [175] localized IP3R and calreticulin in the equatorial segment of the acrosome, in vesicles in the sperm neck close to the nucleus and in the cytoplasmic droplet.

10. Ion Channels Relevant to the Sperm Mitochondria

The integrity of mitochondrial membranes is a crucial prerequisite for proper mitochondrial function. Accordingly, mitochondrial ion channels present in the outer as well as inner mitochondrial membrane are recognized as essential regulators of mitochondrial function [176,177] and may be divided into two major types. Channels of the first type, including most K+ channels, present with properties similar to those located in the plasma membrane. Channels of the second type are exclusive for mitochondria, such as mitochondrial calcium channels [155] or mitochondrial porins [178,179].
Mitochondrial porins are voltage-dependent anion channels (VDACs) located in the outer mitochondrial membrane whose primary function is to regulate the exchange of ATP/ADP, Ca2+, and other metabolites and/or ions between the cytoplasm and mitochondria [180]. Two isomers, specifically VDAC2 and VDAC3, have been identified in the head, acrosome, and outer dense fibers of the flagellum in bovine spermatozoa [181] where they control the cross-talk between mitochondria and the rest of the cell. Their crucial involvement in the regulation of the sperm fertilization ability is furthermore evidenced by Kwon et al. who observed that locking VDAC significantly decreases motility, viability, acrosome reaction, capacitation, tyrosine phosphorylation, fertilization, and embryo development, regardless of Ca2+ levels [182].
Mitochondria enable a rapid uptake of Ca2+ into the matrix and thus are involved in the regulation of cytosolic Ca2+ signals [183]. The molecular machinery underlying Ca2+ uptake into energized mitochondria is driven through the mitochondrial calcium uniporter (mCU) complex (mCUC). It is a highly selective ion channel, mediating the Ca2+ influx across the inner mitochondrial membrane driven by negative mitochondrial membrane potential (ΔΨm) [184]. Experimental studies have revealed that a proper mCUC function is a prerequisite for a desirable sperm viability, motility, and ATP levels while sustaining a proper ΔΨm and ROS production [185].
Potassium is a crucial element for the mitochondrial integrity as its cycle regulates the mitochondrial volume and homeostasis [186]. The K+ influx into mitochondria is driven by a negative ΔΨm, accompanied by anion and water flux, which leads to mitochondrial swelling. The first K+ selective mitochondrial channel to be described is mitoKATP, driven by ATP and localized in the inner mitochondrial membrane [187]. The channel shows similarities with channels regulated by ATP present in the plasma membrane [188]. In the meantime, the most known mitochondrial Ca2+-activated K+ channels include the small conductance K+ (mitoSKCa) channel [189], the intermediate-conductance Ca2+-activated K+ channel (mitoIKCa) [190] and the large-conductance Ca2+-activated K+ channel (mitoBKCa) [191]. The latter has received increased attention for its wide occurrence in various cell types and suggested participation in cytoprotection [192]. Activation of the channel leads to an influx of K+ into the mitochondrial matrix, followed by membrane depolarization, and a decrease in ROS production [193]. Among voltage-gated potassium channels, the inner mitochondrial membrane holds (a) the mitochondrial 1.3 voltage-gated potassium (mitoKv1.3) channel [192]; (b) the mitochondrial 1.5 voltage-gated potassium (mitoKv1.5) channel; and (c) the mitochondrial 7.4 voltage-gated potassium (mitoKv7.4) channel [194]. In isolated mitochondria, modulation of both mitoKv1.3 and mitoKv7.4 leads to changes in ΔΨm and ROS levels, suggesting that both channels are open under physiological conditions [192]. The closure of mitoKv1.3 by either inhibitors or apoptotic regulators leads to excessive ROS release, most likely due to interactions between mitoKv1.3 and the respiratory complex I [195].
Besides cation-selective channels, the inner mitochondrial membrane also contains anion-selective transport systems [196], such as the inner membrane anion channel IMAC, also called the mitochondrial Centum picoSiemens (mCS), which is involved in the regulation of the mitochondrial volume homeostasis [197]. More information is available on channels belonging to chloride intracellular channel proteins (CLIC), specifically CLIC4 and CLIC5, which were recently detected in mitochondrial membranes [198] and are likely to play significant roles in sperm function [199].
Since mitochondria are able to uptake large amounts of Ca2+, a fast pathway for Ca2+ release is associated with mitochondrial permeability transition pore (MPTP), which contributes to cellular homeostasis and prevents mitochondria from Ca2+ overload [200]. MPTP is a multiprotein complex whose opening triggers a massive increase in the inner mitochondrial membrane permeability to solutes of up to 1.5 kDa [201]. According to existing evidence, several proteins present in both mitochondrial membranes and the matrix have been associated with MPTP, including the VDA channels, BCL-2 proteins, adenine nucleotide translocator (ANT) or hexokinase [202,203,204]. In the meantime, the opening of the pore is accelerated by the loss of the inner ΔΨm [205]; the alkalinization of the mitochondrial matrix [206]; and the increase in Ca2+ levels [207]. A transient MPTP opening in intact and healthy cells contributes to cellular homeostasis since it provides a fast release of ions or toxic compounds accumulated in the mitochondrial matrix [208]. Moreover, an MPTP opening may also regulate the activity of some mitochondrial Ca2+-dependent enzymes [209].

11. Ion Channels Relevant to the Sperm Acrosome

The acrosome is a membrane-derived organelle that covers the sperm head of numerous species. Functional acrosomal structures are a critical component of the sperm–egg fusion that is triggered by physiological inducers released from the female gamete or by exposure to specific pharmacological stimuli. Acrosome reaction is defined by a strictly regulated and irreversible process during which the acrosomal contents including Ca2+ and enzymes are released to the extracellular medium [210]. The activation of the acrosome reaction relies on the concurrence of several transduction pathways, most notably ion permeability changes leading to increased intracellular pH, Ca2+ and cAMP, G protein activation, changes to the membrane lipids and protein phosphorylation. Since several ion channel blockers inhibit the progress of acrosome reaction, the role of membrane transporters in these secretory events is indisputable [43,211]. Hence, this section briefly discusses the most important ion channels that are relevant for a proper acrosome functionality.
Voltage-gated Ca2+ channels (VGCCs) have been found in the sperm acrosome of a broad range of species including fish [58], newts [57], and mammals [60,212]. Several types of VGCCs have been identified in spermatozoa, specifically the “long-lasting” L-type that is activated by high voltage and is resistant to ω-conotoxin and ω-agatoxin [58,60,212]; the “transient” T-type channel operated by low voltage [212,213]; and the “Purkinje” P/Q-type that is activated by high voltage, is resistant to ω-conotoxin and blocked by ω-agatoxin [212]. Both the L- and T-type channels have been reported to participate in human sperm acrosome reactions [60] by activating membrane depolarization and mediating Ca2+ influx in response to changes in the action potential and depolarizing signals [214].
Transient receptor potential vanilloids (TRPVs) consist of six subtypes divided into two groups depending on their Ca2+ permeability and sensitivity to temperature, specifically TRPV1/TRPV2/TRPV3/TRPV4 and TRPV5/TRPV6 [215]. All six TRPV subtypes have been found in the spermatozoa of vertebrates, even though TRPV1 and TRPV4 seem to be present more frequently. TRPV1 is a voltage/heat/lipid/pH-modulated channel localized in the sperm head and the acrosome [45,216,217] which is desensitized by internal Ca2+; however, the channel is not activated by Ca2+-store reduction [215]. The intensity of the TRPV1 current rises with increasingly acidic pH and is regulated by intracellular phosphatidylinositol 4,5-bisphosphate [215]. In the meantime, the TRPV4 channel activation depends on the extracellular osmolality, pH, lipids, and mechanical triggers, such as shear stress or membrane stretching [215,218].
Store-operated Ca2+ channels (SOCCs) are inward rectifiers, and their primary role is to supply the cellular compartments with Ca2+ from extracellular environment after Ca2+ is released and pumped out across the plasma membrane [43]. Since SOCCs are not voltage-dependent channels, they are functional even at negative membrane potentials at which depolarization-sensitive channels (such as VGCCs) are not engaged in action. SOCCs are assembled by ORAI1–3 proteins, where one channel is created by one ORAI protein. The channel is inactivated by Ca2+ binding [48]. SOCs have been suggested to play an important role in the regulation of sperm physiology as the channel inhibition reduces sperm motility and acrosome reaction [51,219]. This phenomenon is most likely regulated through the induction of 5′ AMP-activated protein kinase phosphorylation [51].
Similar to the sperm plasma membrane, CatSper channels are the most studied sperm Ca2+ channels in the acrosome structures because of their sperm specificity and crucial roles in sperm–egg chemotaxis, capacitation, and acrosome reaction [38]. Besides Ca2+, CatSper also facilitates the entry of monovalent (Na+ and Cs+) and bivalent cations (Ba2+) to spermatozoa if extracellular Ca2+ is absent. The channel is pH-sensitive and triggered by alkaline pH [38]. Its activity is furthermore regulated by cyclic nucleotides, membrane voltage, phosphorylation, biomolecules (such as prostaglandin, BSA and progesterone) and zona pellucida glycoproteins [38,220].
A Ca2+-activated Cl channel (CaCC) opening is stimulated by increases in intracellular Ca2+ levels resulting from its influx through the plasma membrane channels or its release from intracellular stores. CaCCs have been reported to be present in the heads of mature human spermatozoa where they may contribute to Ca2+-dependent Cl currents necessary for a proper acrosome reaction [130]. In the meantime, chloride channels (ClCs), specifically ClC-3, have been detected in the acrosome and midpiece of bovine spermatozoa [133], playing important roles in the regulation of sperm volume, motility, and fertilization ability [135].
The main role of a voltage-gated H+ channel (VGHC) is to extrude H+ from the cell, leading to an increase in intracellular pH [221]. These channels exhibit a highly selective H+-conductance, opening with membrane depolarization, high extracellular pH and decreased intracellular pH [221]. Besides in humans [76,222,223], VGHCs have also been found in macaque [224], boar [225], and bull spermatozoa [61]. VGHC activation leads to intracellular alkalinization, which is accompanied by CatSper activation, Ca2+ influx and the induction of hypermotility and acrosome activation [213,222,226]. As such, VGHC has been shown to be involved in the induction of capacitation, progressive motility, and acrosome reaction through induced NADPH oxidase 5 activation and ROS generation [69,222,227]. Correspondingly, their inhibition leads to a reduced sperm motility and progesterone-induced acrosome reaction [225,226].

12. Conclusions

Regulation machinery behind the whole capacitation process is still not fully understood. However, what we can say for sure is that ion channels of the sperm plasma membrane are responsible for the maintenance of various biological and biochemical changes such as alkalinization, hyperpolarization, hyperactivation as well as capacitation. These channels and transporters supported the adaptation of sperm cells to a constantly changing environment during their fertilization journey in the female genital tract. Accordingly, since their dysfunction has been frequently correlated with sub- or infertility, a more profound understanding of their involvement in the regulation of sperm behavior in future studies may contribute to the evolution of new strategies for the management of male reproductive dysfunction.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/biomedicines11092519/s1.

Author Contributions

Conceptualization, F.B., E.T. and D.U.; writing—original draft preparation, F.B., E.T. and M.Ď.; writing—review and editing, F.B. and E.T.; supervision, E.T. and D.U.; project administration, E.T.; funding acquisition, E.T. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was supported by the Slovak Research and Development Agency, grant number APVV-15-0544 and APVV-21-0095; and by the Cultural and Educational Grant Agency Ministry of Education, Science, Research and Sport of the Slovak Republic under the grant number 008SPU-4/2021.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We wish to thank the Center for Animal Reproduction (CeRA) Team of Excellence for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gogoi, J.; Mondal, M.; Pal, P.; Upadhyay, V.R.; Kuri, P.; Mukherjee, I.; Tripura, S. Potential role of CatSper and proton ion channels in bull fertility. Pharma Innov. 2022, 11, 313–321. [Google Scholar]
  2. Delgado-Bermúdez, A.; Yeste, M.; Bonet, S.; Pinart, E. A review on the role of Bicarbonate and Proton Transporters during Sperm Capacitation in Mammals. Int. J. Mol. Sci. 2022, 23, 6333. [Google Scholar] [CrossRef] [PubMed]
  3. Matamoros-Volante, A.; Treviño, C.L. Capacitation-associated alkalization in human sperm is differentially controlled at the subcellular level. J. Cell Sci. 2020, 133, jcs238816. [Google Scholar] [CrossRef]
  4. López-González, I.; Torres-Rodríguez, P.; Sánchez-Carranza, O.; Solís-López, A.; Santi, C.M.; Darzson, A.; Treviño, C.L. Membrane hyperpolarization during human sperm capacitation. Mol. Hum. Reprod. 2014, 20, 619–629. [Google Scholar] [CrossRef]
  5. Baro Graf, C.; Ritagliati, C.; Stival, C.; Balestrini, P.A.; Buffone, M.G.; Krapf, D. Determination of a robust assay for human sperm membrane potential analysis. Front. Cell Develop. Biol. 2019, 7, 101. [Google Scholar] [CrossRef]
  6. Shukla, K.K.; Mahdi, A.A.; Rajender, S. Ion Channels in Sperm Physiology and Male Fertility and Infertility. J. Androl. 2013, 33, 777–788. [Google Scholar] [CrossRef] [PubMed]
  7. Nowicka-Bauer, K.; Szymczak-Cendlak, M. Structure and Function of Ion Channels Regulating Sperm Motility—An Overview. Int. J. Mol. Sci. 2021, 22, 3259. [Google Scholar] [CrossRef]
  8. Michalakis, S.; Becirovic, E.; Biel, M. Retinal Cyclic Nucleotide-Gated Channels: From Pathophysiology to Therapy. Int. J. Mol. Sci. 2018, 19, 749. [Google Scholar] [CrossRef]
  9. Balbach, M.; Ghanem, L.; Rossetti, T.; Kaur, N.; Ritagliati, C.; Ferreira, J.; Krapf, D.; Puga Molina, L.C.; Santi, C.M.; Hansen, J.N.; et al. Soluble adenylyl cyclase inhibition prevents human sperm functions essential for fertilization. Mol. Hum. Reprod. 2021, 27, gaab054. [Google Scholar] [CrossRef]
  10. Nishigaki, T.; José, O.; González-Cota, A.L.; Romero, F.; Trevino, C.L.; Darzson, A. Intracellular pH in sperm physiology. Biochem. Biophys. Res. Commun. 2014, 450, 1149–1158. [Google Scholar] [CrossRef] [PubMed]
  11. Pastor-Soler, N.; Piétrement, C.; Breton, S. Role of acid/base tansporters in the male reproductive tract and potential consequences of their malfunction. Physiology 2005, 20, 417–428. [Google Scholar] [CrossRef]
  12. Chen, Y.; Cann, M.J.; Litvin, T.N.; Iourgenko, V.; Sinclair, M.L.; Levin, L.R.; Buck, J. Soluble adenylyl cyclase as an evolutionary conserved bicarbonate sensor. Science 2000, 289, 481–485. [Google Scholar] [CrossRef]
  13. Bernardino, R.L.; Carrageta, D.F.; Sousa, M.; Alves, M.G.; Oliveira, P.F. pH and male fertility: Making sense on pH homeodynamics throughout the male reproductive tract. Cell. Mol. Life Sci. 2019, 76, 3783–3800. [Google Scholar] [CrossRef] [PubMed]
  14. Bonar, P.T.; Casey, J.R. Plasma membrane Cl/HCO3 exchangers: Structure, mechanism and physiology. Channels 2008, 2, 337–345. [Google Scholar] [CrossRef] [PubMed]
  15. Apler, S.L. Molecular physiology and genetics of Na+-independent SLC4 anion exchangers. J. Exp. Biol. 2009, 212, 1672–1683. [Google Scholar]
  16. Peña-Münzenmayer, G.; George, A.T.; Shull, G.E.; Melvin, J.E.; Catalán, M.A. Ae4 (Slc4a9) is an electroneutral monovalent cation-dependent Cl/HCO3 exchanger. J. Gen. Phys. 2016, 147, 423–436. [Google Scholar] [CrossRef]
  17. Liu, Y.; Yang, J.; Chen, L.M. Structure and Function of SLC4 Family HCO3- Transporters. Front. Physiol. 2015, 6, 355. [Google Scholar] [CrossRef]
  18. Puga Molina, L.C.; Lugue, G.M.; Balestrini, P.A.; Marín-Briggiler, C.I.; Romarowski, A.; Buffone, M.G. Molecular basis of human sperm capacitation. Front. Cell Dev. Biol. 2018, 6, 72. [Google Scholar] [CrossRef]
  19. Rapp, C.; Bai, X.; Reithmeier, R.A.F. Molecular analysis of human solute carrier SLC26 anion transporter disease-causing mutations using 3-dimensional homology modeling. Biochim. Et Biophys. Acta (BBA)—Biomembr. 2017, 1859, 2420–2434. [Google Scholar] [CrossRef]
  20. Wedenoja, S.; Khamaysi, A.; Shimshilashvili, L.; Anbtawe-Jomaa, S.; Elomaa, O.; Toppari, J.; Höglund, P.; Aittomäki, K.; Holmberg, C.; Hovatta, O.; et al. A missense mutation in SLC26A3 is associated with human male subfertility and impaired activation of CFTR. Sci. Rep. 2017, 7, 14208. [Google Scholar] [CrossRef]
  21. Chávez, J.C.; Hernández-González, E.O.; Wertheimer, E.; Visconti, P.E.; Darzson, A.; Treviño, C.L. Participation of the Cl/HCO3 exchangers SLC26A3 and SLC26A6, the Cl channel CFTR, and the regulatory factor SLC9A3R1 in mouse sperm capacitation. Biol. Reprod. 2012, 86, 1–14. [Google Scholar] [CrossRef]
  22. Seidler, U.; Nikolovska, K. Slc26 family of anion transporters in the gastrointestinal tract: Expression, function, regulation, and role in disease. Compr. Physiol. 2019, 9, 839–872. [Google Scholar]
  23. Chen, W.Y.; Xu, W.M.; Chen, Z.H.; Ni, Y.; Yuan, Y.Y.; Zhou, S.C.; Zhou, W.W.; Tsang, L.L.; Chung, Y.W.; Höglund, P.; et al. Cl is required for HCO3 entry necessary for sperm capacitation in guinea pig: Involvement of a Cl/HCO3 exchanger (SLC26A3) and CFTR. Biol. Reprod. 2009, 80, 115–123. [Google Scholar] [CrossRef]
  24. Touré, A. Importance of SLC26 transmembrane anion exchangers in sperm post-testicular maturation and fertilization potential. Front. Cell Dev. Biol. 2019, 7, e230. [Google Scholar] [CrossRef]
  25. Rode, B.; Dirami, T.; Bakouh, N.; Rizk-Rabin, M.; Norez, C.; Lhuillier, P.; Lorès, P.; Jollivet, M.; Melin, P.; Zvetkova, I.; et al. The testis anion transporter TAT1 (SLC26A8) physically and functionally interacts with the cystic fibrosis transmembrane conductance regulator channel: A potential role during sperm capacitation. Hum. Mol. Gen. 2012, 21, 1287–1298. [Google Scholar] [CrossRef]
  26. Breitbart, H.; Finkelstein, M. Regulation of Sperm Capacitation and the Acrosome Reaction by PIP2 and Actin Modulation. Asian J. Androl. 2015, 17, 597. [Google Scholar] [CrossRef]
  27. Chauhan, D.S.; Swain, D.K.; Shah, N.; Yadav, H.P.; Nakade, U.P.; Singh, V.K.; Nigam, R.; Yadav, S.; Garg, S.K. Functional and molecular characterization of voltage gated Nav 1.8 in bull spermatozoa. Theriogenology 2017, 90, 210–218. [Google Scholar] [CrossRef] [PubMed]
  28. Pinto, F.M.; Ravina, C.G.; Fernández-Sánchez, M.; Gallardo-Castro, M.; Cejudo-Román, A.; Candenas, L. Molecular and functional characterization of voltage-gated sodium channels in human sperm. Reprod. Biol. Endocrinol. 2009, 7, 71. [Google Scholar] [CrossRef] [PubMed]
  29. Catterall, W.A.; Goldin, A.L.; Waxman, S.G. International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol. Rev. 2005, 57, 397–409. [Google Scholar] [PubMed]
  30. Cejudo-Roman, A.; Pinto, F.M.; Subirán, N.; Ravina, C.G.; Fernández-Sánchez, M.; Pérez-Hernámdez, N.; Pérez, R.; Pacheco, A.; Irazusta, J.; Candenas, L. The Voltage-Gated Sodium Channel NaV1.8 Is Expressed in Human Sperm. PLoS ONE 2013, 8, e76084. [Google Scholar] [CrossRef]
  31. Ureña, I.; González, C.; Ramón, M.; Gódia, M.; Clop, A.; Calvo, J.H.; Carabaño, M.J.; Serrano, M. Exploring the ovine sperm transcriptome by RNAseq techniques. Effect of seasonal conditions on transcripts abundance. PLoS ONE 2022, 17, e0264978. [Google Scholar] [CrossRef] [PubMed]
  32. Eastwood, A.L.; Goodman, M.B. Insight into DEG/ENaC Channel Gating from Genetics and Structure. Physiology 2012, 27, 282–290. [Google Scholar] [CrossRef] [PubMed]
  33. Puga Molina, L.C.; Pinto, N.A.; Torres, N.I.; González-Cota, A.L.; Luque, G.M.; Balestrini, P.A.; Romarowski, A.; Krapf, D.; Santi, C.M.; Treviño, C.L.; et al. CFTR/ENaC-dependent regulation of membrane potential during human sperm capacitation is initiated by bicarbonate uptake through NBC. J. Biol. Chem. 2018, 293, P9924–P9936. [Google Scholar] [CrossRef] [PubMed]
  34. Pinto, F.M.; Odriozola, A.; Candenas, l.; Subirán, N. The Role of Sperm Membrane Potential and ion Channels in Regulating Sperm Function. Int. J. Mol. Sci. 2023, 24, 6995. [Google Scholar] [CrossRef]
  35. Hernández-González, E.O.; Sosnik, J.; Edwards, J.; Acevedo, J.J.; Mendoza-Lujambio, I.; López-González, I.; Demarco, I.; Wertheimer, E.; Darzson, A.; Visconti, P.E. Sodium and epithelial sodium channels participate in the regulation of the capacitation-associated hyperpolarization in mouse sperm. J. Biol. Chem. 2006, 281, 5623–5633. [Google Scholar] [CrossRef]
  36. Darzson, A.; Nishigaki, T.; Wood, C.; Treviño, C.L.; Felix, R.; Beltrán, C. Calcium Channels and Ca2+ Fluctuations in Sperm Physiology. Int. Rev. Cytol. 2005, 243, 79–172. [Google Scholar]
  37. Vyklicka, L.; Lishko, P.V. Dissecting the signaling pathways involved in the function of sperm flagellum. Curr. Opin. Cell Biol. 2020, 63, 154–161. [Google Scholar] [CrossRef]
  38. Sun, X.H.; Zhu, Y.Y.; Wang, L.; Liu, H.L.; Ling, Y.; Li, Z.L.; Sun, L.B. The CatSper channel and its roles in male fertility: A systematic review. Reprod. Biol. Endocrinol. 2017, 15, 65. [Google Scholar] [CrossRef]
  39. Navarro, B.; Kirichok, Y.; Chung, J.J.; Clapham, D.E. Ion channels that control fertility in mammalian spermatozoa. Int. J. Dev. Biol. 2008, 52, 607. [Google Scholar] [CrossRef]
  40. Brenker, C.; Goodwin, N.; Weyand, I.; Kashikar, N.D.; Naruse, M.; Krahling, M.; Muller, A.; Kaupp, U.B.; Strunker, T. The CatSper channel: A polymodal chemosensor in human sperm. EMBO J. 2012, 31, 1654–1665. [Google Scholar] [CrossRef]
  41. Hwang, J.Y.; Mannowetz, N.; Zhang, Y.; Everley, R.A.; Gygi, S.P.; Bewersdorf, J.; Lishko, P.V.; Chung, J.J. Dual Sensing of Physiologic pH and Calcium by EFCAB9 Regulates Sperm Motility. Cell 2019, 177, 1480–1494. [Google Scholar] [CrossRef] [PubMed]
  42. Swain, D.K.; Sharma, P.; Shah, N.; Sethi, M.; Mahajan, A.; Gupta, S.; Mishra, A.K.; Yadav, S. Introduction to the pathways involved in the activation and regulation of sperm motility: A review of the relevance of ion channels. Anim. Reprod. Sci. 2022, 246, 107052. [Google Scholar] [CrossRef] [PubMed]
  43. Darzson, A.; Nishigaki, T.; Beltran, C.; Treviño, C.L. Calcium Channels in the Development, Maturation, and Function of Spermatozoa. Physiol. Rev. 2011, 91, 1305–1355. [Google Scholar] [CrossRef]
  44. Kumar, A.; Mishra, A.K.; Swain, D.K.; Singh, V.; Yadav, S.; Saxena, A. Role of transient receptor potential channels in regulating spermatozoa functions: A mini-review. Vet. World. 2018, 11, 1618–1623. [Google Scholar] [CrossRef]
  45. Kumar, A.; Mishra, A.K.; Singh, V.; Yadav, S.; Saxena, A.; Garg, S.K.; Swain, D.K. Molecular and functional insights into transient receptor potential vanilloid 1 (TRPV1) in bull spermatozoa. Theriogenology 2019, 128, 207–217. [Google Scholar] [CrossRef]
  46. Latorre, R.; Zaelzer, C.; Brauchi, S. Structure-functional intimacies of transient receptor potential channels. Q. Rev. Biophys. 2009, 42, 201–246. [Google Scholar] [CrossRef]
  47. Vangeel, L.; Voets, T. Transient Receptor Potential Channels and Calcium Signaling. Cold Spring Harb. Perspect. Biol. 2019, 11, a035048. [Google Scholar] [CrossRef]
  48. Prakriya, M.; Lewis, R.S. Store-Operated Calcium Channels. Physiol. Rev. 2015, 95, 1383–1436. [Google Scholar] [CrossRef] [PubMed]
  49. Costello, S.; Michelangeli, F.; Nash, K.; Lefievre, L.; Morris, J.; Machado-Oliveira, G.; Barrat, C.; Kirkman-Brown, J.; Publicover, S. Ca2+-stores in sperm: Their identities and functions. Reproduction 2009, 138, 425–437. [Google Scholar] [CrossRef]
  50. Darzson, A.; Sánchez-Cárdenas, C.; Orta, G.; Sánchez-Tusie, A.A.; Beltrán, C.; López-González, I.; Granados-González, G.; Treviño, C.L. Are TRP channels involved in sperm development and function? Cell Tissue Res. 2012, 349, 749–764. [Google Scholar] [CrossRef]
  51. Nguyen, T.M.; Duittoz, A.; Praud, C.; Combarnous, Y.; Blesbois, E. Calcium channels in chicken sperm regulate motility and the acrosome reaction. FEBS J. 2016, 283, 1902–1920. [Google Scholar] [CrossRef] [PubMed]
  52. Mazzolini, M.; Arcangeletti, M.; Marchesi, A.; Napolitano, L.M.R.; Grosa, D.; Maity, S.; Anselmi, C.; Torre, V. The gating mechanism in cyclic nucleotide-gated ion channels. Sci. Rep. 2018, 8, 45. [Google Scholar] [CrossRef] [PubMed]
  53. Kaupp, B.U.; Seifert, R. Cyclic nucleotide-gated ion channels. Physiol Rev. 2002, 82, 769–824. [Google Scholar] [CrossRef] [PubMed]
  54. Matulef, k.; Zagotta, W.N. Cyclic nucleotide-gated ion channels. Annu. Rev. Cell Dev. Biol. 2003, 19, 23–44. [Google Scholar] [CrossRef] [PubMed]
  55. Espinal-Enríquez, J.; Priego-Espinosa, D.A.; Darszon, A.; Beltrán, C.; Martínez-Mekler, G. Network model predicts that CatSper is the main Ca2+ channel in the regulation of sea urchin sperm motility. Sci. Rep. 2017, 7, 4236. [Google Scholar] [CrossRef]
  56. Loux, S.C.; Crawford, K.R.; Ing, N.H.; González-Fernández, L.; Macías-García, B.; Love, C.C.; Varner, D.D.; Velez, I.C.; Choi, Y.H.; Hinrichs, K. CatSper and the Relationship of Hyperactivated Motility to Intracellular Calcium and pH Kinetics in Equine Sperm1. Biol. Reprod. 2013, 89, 123. [Google Scholar] [CrossRef]
  57. Takayama-Watanabe, E.; Ochiai, H.; Tanino, S.; Watanabe, A. Contribution of different Ca2+ channels to the acrosome reaction-mediated initiation of sperm motility in the newt Cynops pyrrhogaster. Zygote 2015, 23, 342–351. [Google Scholar] [CrossRef]
  58. Lissabet, J.F.B.; Belén, L.H.; Lee-Estevez, M.; Farias, J.G. Role of voltage-gated L-type calcium channel in the spermatozoa motility of Atlantic salmon (Salmo salar). Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2020, 241, 110633. [Google Scholar] [CrossRef]
  59. Albrizio, M.; Moramarco, A.M.; Nicassio, M.; Micera, E.; Zarrilli, A.; Lacalandra, G.M. Localization and functional modification of L-type voltage-gated calcium channels in equine spermatozoa from fresh and frozen semen. Theriogenology 2015, 83, 421–429. [Google Scholar] [CrossRef]
  60. José, O.; Hernández-Hernández, O.; Chirinos, M.; González-González, M.E.; Larrea, F.; Almanza, A.; Felix, R.; Darszon, A.; Treviño, C.L. Recombinant human ZP3-induced sperm acrosome reaction: Evidence for the involvement of T- and L-type voltage-gated calcium channels. Biochem. Biophys. Res. Commun. 2010, 395, 530–534. [Google Scholar] [CrossRef]
  61. Mishra, A.K.; Kumar, A.; Yadav, S.; Anand, M.; Yadav, B.; Nigam, R.; Garg, S.K.; Swain, D.K. Functional insights into voltage gated proton channel (Hv1) in bull spermatozoa. Theriogenology 2019, 136, 118–130. [Google Scholar] [CrossRef]
  62. Majhi, R.K.; Kumar, A.; Yadav, M.; Swain, N.; Kumari, S.; Saha, A.; Pradhan, A.; Goswami, L.; Saha, S.; Samanta, L.; et al. Thermosensitive ion channel TRPV1 is endogenously expressed in the sperm of a fresh water teleost fish (Labeo rohita) and regulates sperm motility. Channels 2013, 7, 483–492. [Google Scholar] [CrossRef] [PubMed]
  63. Hamano, K.; Kawanishi, T.; Mizuno, A.; Suzuki, M.; Takagi, Y. Involvement of Transient Receptor Potential Vanilloid (TRPV) 4 in mouse sperm thermotaxis. J. Reprod. Dev. 2016, 62, 415–422. [Google Scholar] [CrossRef] [PubMed]
  64. Yoshida, K.; Shiba, K.; Sakamoto, A.; Ikenaga, J.; Matsunaga, S.; Inaba, K.; Yoshida, M. Ca2+ efflux via plasma membrane Ca2+-ATPase mediates chemotaxis in ascidian sperm. Sci. Rep. 2018, 8, 16622. [Google Scholar] [CrossRef] [PubMed]
  65. Cisneros-Mejorado, A.; Sánchez Herrera, D.P. cGMP and cyclic nucleotide-gated channels participate in mouse sperm capacitation. FEBS Lett. 2012, 586, 149–153. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, Z.; Yang, Y.; Wu, H.; Zhang, H.; Zhang, H.; Mao, J.; Liu, D.; Zhao, L.; Lin, H.; Tang, W.; et al. Sodium-Hydrogen-Exchanger expression in human sperm and its relationship with semen parameters. J. Assist. Reprod. Genet. 2017, 34, 795–801. [Google Scholar] [CrossRef]
  67. Chen, S.R.; Chen, M.; Deng, S.L.; Hao, X.X.; Liu, Y.X. Sodium-hydrogen exchanger NHA1 and NHA2 control sperm motility amd male fertility. Cell Death Dis. 2016, 7, e2152. [Google Scholar] [CrossRef]
  68. Wakabayashi, S.; Pang, T.; Su, T.; Shigekawa, M. A Novel Topology Model of the Human Na+/H+ Exchanger Isoform 1. J. Biol. Chem. 2000, 275, P7942–P7949. [Google Scholar] [CrossRef]
  69. Lishko, P.V.; Kirichok, Y. The role of Hv1 and CatSper channels in sperm activation. J. Physiol. 2010, 588, 4667–4672. [Google Scholar] [CrossRef]
  70. Florman, H.M.; Jungnickel, M.K.; Sutton, K.A. Shedding light on sperm pHertility. Cell 2010, 140, 310–312. [Google Scholar] [CrossRef]
  71. Ramsey, I.S.; Mokrab, Y.; Carvacho, I.; Sands, Z.A.; Sansom, M.S.P.; Clapham, D.E. An aqueous H+ permeation in the voltage-gated proton channel Hv1. Nat. Struct. Mol. Biol. 2010, 17, 869–875. [Google Scholar] [CrossRef]
  72. Muzzachi, S.; Guerra, L.; Martino, N.A.; Favia, M.; Punzi, G.; Silvestre, F.; Guaricci, A.C.; Roscino, M.T.; Pierri, C.L.; Dell’Aquila, M.E.; et al. Effect of cariporide on ram sperm pH regulation and motility: Possible role of NHE1. Reproduction 2018, 155, 433–445. [Google Scholar] [CrossRef]
  73. Martins, A.D.; Bernardino, R.L.; Neuhaus-Oliveira, A.; Sousa, M.; Sá, R.; Alves, M.G.; Oliveira, P.F. Physiology of Na+/H+ Exchangers in the Male Reproductive Tract: Relevance for Male Fertility. Biol. Reprod. 2014, 91, 11. [Google Scholar] [CrossRef]
  74. Yeste, M.; Recuero, S.; Maside, C.; Salas-Huetos, A.; Bonet, S.; Pinart, E. Blocking NHE Channels Reduces the Ability of In Vitro Capacitated Mammalian Sperm to Respond to Progesterone Stimulus. Int. J. Mol. Sci. 2021, 22, 12646. [Google Scholar] [CrossRef] [PubMed]
  75. Cavarocchi, E.; Whitfield, M.; Chargui, A.; Stouvenel, L.; Lorès, P.; Coutton, C.; Arnoult, C.; Santulli, P.; Patrat, C.; Thierry-Mieg, N.; et al. The sodium/proton exchanger SLC9C1 (sNHE) is essential for human sperm motility and fertility. Clin. Genet. 2021, 99, 684–693. [Google Scholar] [CrossRef] [PubMed]
  76. Berger, T.K.; Fußhöller, D.M.; Goodwin, N.; Bönigk, W.; Müller, A.; Dokani Khesroshahi, N.; Brenker, C.; Wachten, D.; Krause, E.; Kaupp, U.B.; et al. Post-translational cleavage of Hv1 in human sperm tunes pH- and voltage-dependent gating. J. Physiol. 2017, 595, 1533–1546. [Google Scholar] [CrossRef]
  77. Arnoult, C.; Kazam, I.G.; Visconti, P.E.; Kopf, G.S.; Villaz, M.; Florman, H.M. Control of the low voltage-activated calcium channel of mouse sperm by egg ZP3 and by membrane hyperpolarization during capacitation. Proc. Natl. Acad. Sci. USA 1999, 96, 6757–6762. [Google Scholar] [CrossRef]
  78. Santi, C.M.; Orta, G.; Salkoff, L.; Visconti, P.E.; Darszon, A.; Treviño, C.L. K+ and Cl Channels and Transporters in Sperm Function. Gametogenesis 2013, 102, 385–421. [Google Scholar]
  79. Acevedo, J.J.; Mendoza-Lujambio, I.; de la Vega-Beltran, J.L.; Trevino, C.L.; Felix, R.; Darszon, A. KATP channels in mouse spermatogenic cells and sperm, and their role in capacitation. Dev. Biol. 2006, 289, 395–405. [Google Scholar] [CrossRef] [PubMed]
  80. Felix, R.; Serrano, C.J.; Treviño, C.L.; Muñoz-Garay, C.; Bravo, A.; Navarro, A.; Pacheco, J.; Tsutsumi, V.; Darszon, A. Identification of distinct K+ channels in mouse spermatogenic cells and sperm. Zygote 2002, 10, 183–188. [Google Scholar] [CrossRef] [PubMed]
  81. Hagiwara, S.; Kawa, K. Calcium and potassium currents in spermatogenic cells dissociated from rat seminiferous tubules. J. Physiol. 1984, 356, 135–149. [Google Scholar] [CrossRef] [PubMed]
  82. Jacob, A.; Hurley, I.R.; Goodwin, L.O.; Cooper, G.W.; Benoff, S. Molecular characterization of a voltage-gated potassium channel expressed in rat testis. Mol. Hum. Reprod. 2000, 6, 303–313. [Google Scholar] [CrossRef] [PubMed]
  83. Martinez-Lopez, P.; Santi, C.M.; Trevino, C.L.; Ocampo-Gutierrez, A.Y.; Acevedo, J.J.; Alisio, A.; Salkoff, L.B.; Darszon, A. Mouse sperm K+ currents stimulated by pH and cAMP possibly coded by Slo3 channels. Biochem. Biophys. Res. Commun. 2009, 381, 204–209. [Google Scholar] [CrossRef] [PubMed]
  84. Salvatore, L.; D’Adamo, M.C.; Polishchuk, R.; Salmona, M.; Pessia, M. Localization and age-dependent expression of the inward rectifier K+ channel subunit Kir 5.1 in a mammalian reproductive system. FEBS Lett. 1999, 449, 146–152. [Google Scholar] [CrossRef]
  85. Schreiber, M.; Wei, A.; Yuan, A.; Gaut, J.; Saito, M.; Salkoff, L. Slo3, a novel pH-sensitive K+ channel from mammalian spermatocytes. J. Biol. Chem. 1998, 273, 3509–3516. [Google Scholar] [CrossRef]
  86. Leonetti, M.D.; Yuan, P.; Hsiung, Y.; Mackinnon, R. Functional and structural analysis of the human SLO3 pH- and voltage-gated K+ channel. Proc. Natl. Acad. Sci. USA 2012, 109, 19274–19279. [Google Scholar] [CrossRef]
  87. Zhou, Y.; Yang, H.; Cui, J.; Lingle, C.J. Threading the biophysics of mammalian Slo1 channels onto structures of an invertebrate Slo1 channel. J. Gen. Physiol. 2017, 149, 985–1007. [Google Scholar] [CrossRef]
  88. Zeng, X.H.; Yang, C.; Kim, S.T.; Lingle, C.J.; Xia, X.M. Deletion of the Slo3 gene abolishes alkalization-activated K+ current in mouse spermatozoa. Proc. Natl. Acad. Sci. USA 2011, 108, 5879–5884. [Google Scholar] [CrossRef]
  89. Yang, C.; Zeng, X.H.; Zhou, Y.; Xia, X.M.; Lingle, C.J. LRRC52 (leucine-rich-repeat-containing protein 52), a testis-specific auxiliary subunit of the alkalization-activated Slo3 channel. Proc. Natl. Acad. Sci. USA 2011, 108, 19419–19424. [Google Scholar] [CrossRef]
  90. Vicens, A.; Andrade-López, K.; Cortez, D.; Gutiérrez, R.M.; Treviño, C.L. Premammalian origin of the sperm-specific Slo3 channel. FEBS Open Bio. 2017, 7, 382–390. [Google Scholar] [CrossRef]
  91. Chávez, J.C.; Vicens, A.; Wrighton, D.C.; Andrade-López, K.; Beltrán, C.; Gutiérrez, R.M.; Lippiat, J.D.; Treviño, C.L. A cytoplasmic Slo3 isoform is expressed in somatic tissues. Mol. Biol. Rep. 2019, 46, 5561–5567. [Google Scholar] [CrossRef]
  92. Schreiber, M.; Salkoff, L. A novel calcium-sensing domain in the BK channel. Biophys. J. 1997, 73, 1355–1363. [Google Scholar] [CrossRef]
  93. Zhang, X.; Zeng, X.; Lingle, C.J. Slo3 K+ channels: Voltage and pH dependence of macroscopic currents. J. Gen. Physiol. 2006, 128, 317–336. [Google Scholar] [CrossRef] [PubMed]
  94. Geng, Y.; Ferreira, J.J.; Dzikunu, V.; Butler, A.; Lybaert, P.; Yuan, P.; Magleby, K.L.; Salkoff, L.; Santi, C.M. A genetic variant of the sperm-specific SLO3 K+ channel has altered pH and Ca2+ sensitivities. J. Biol. Chem. 2017, 292, 8978–8987. [Google Scholar] [CrossRef] [PubMed]
  95. Brenker, C.; Zhou, Y.; Müller, A.; Echeverry, F.A.; Trötschel, C.; Poetsch, A.; Xia, X.M.; Bönigk, W.; Lingle, C.J.; Kaupp, U.B.; et al. The Ca2+-activated K+ current of human sperm is mediated by Slo3. eLife 2014, 3, e01438. [Google Scholar] [CrossRef]
  96. Kumar, S.; Ying, Y.K.; Hong, P.; Maddaiah, V.T. Potassium increases intracellular calcium simulating progesterone action in human sperm. Arch. Androl. 2000, 44, 93–101. [Google Scholar] [PubMed]
  97. Yeung, C.H.; Barfield, J.P.; Cooper, T.G. Physiological volume regulation by spermatozoa. Mol. Cell. Endocrinol. 2006, 250, 98–105. [Google Scholar] [CrossRef]
  98. Santi, C.M.; Butler, A.; Kuhn, J.; Wei, A.; Salkoff, L. Bovine and mouse SLO3 K+ channels: Evolutionary divergence points to an RCK1 region of critical function. J. Biol. Chem. 2009, 284, 21589–21598. [Google Scholar] [CrossRef]
  99. Balla, T. Phosphoinositides: Tiny lipids with giant impact on cell regulation. Physiol. Rev. 2013, 93, 1019–1137. [Google Scholar] [CrossRef]
  100. Zeng, Y.; Clark, E.N.; Florman, H.M. Sperm membrane potential: Hyperpolarization during capacitation regulates zona pellucida-dependent acrosomal secretion. Dev. Biol. 1995, 171, 554–563. [Google Scholar] [CrossRef]
  101. Santi, C.M.; Martínez-López, P.; de la Vega-Beltrán, J.L.; Butler, A.; Alisio, A.; Darszon, A.; Salkoff, L. The SLO3 sperm-specific potassium channel plays a vital role in male fertility. FEBS Lett. 2010, 584, 1041–1046. [Google Scholar] [CrossRef] [PubMed]
  102. Alavi, S.M.; Cosson, J. Sperm motility in fishes. (II) Effects of ions and osmolality: A review. Cell Biol. Int. 2006, 30, 1–14. [Google Scholar] [CrossRef] [PubMed]
  103. Fechner, S.; Alvarez, L.; Bönigk, W.; Müller, A.; Berger, T.K.; Pascal, R.; Trötschel, C.; Poetsch, A.; Stölting, G.; Siegfried, K.R.; et al. A K+-selective CNG channel orchestrates Ca2+ signalling in zebrafish sperm. eLife 2015, 4, e07624. [Google Scholar] [CrossRef] [PubMed]
  104. Strünker, T.; Weyand, I.; Bönigk, W.; Van, Q.; Loogen, A.; Brown, J.E.; Kashikar, N.; Hagen, V.; Krause, E.; Kaupp, U.B. A K+-selective cGMP-gated ion channel controls chemosensation of sperm. Nat. Cell. Biol. 2006, 8, 1149–1154. [Google Scholar] [CrossRef] [PubMed]
  105. Bönigk, W.; Loogen, A.; Seifert, R.; Kashikar, N.; Klemm, C.; Krause, E.; Hagen, V.; Kremmer, E.; Strünker, T.; Kaupp, U.B. An atypical CNG channel activated by a single cGMP molecule controls sperm chemotaxis. Sci. Signal. 2009, 2, ra68. [Google Scholar] [CrossRef]
  106. Beltrán, C.; Rodríguez-Miranda, E.; Granados-González, G.; de De la Torre, L.G.; Nishigaki, T.; Darszon, A. Zn2+ induces hyperpolarization by activation of a K+ channel and increases intracellular Ca2+ and pH in sea urchin spermatozoa. Dev. Biol. 2014, 394, 15–23. [Google Scholar] [CrossRef]
  107. Hibino, H.; Inanobe, A.; Furutani, K.; Murakami, S.; Findlay, I.; Kurachi, Y. Inwardly rectifying potassium channels: Their structure, function, and physiological roles. Physiol. Rev. 2010, 90, 291–366. [Google Scholar] [CrossRef]
  108. Munoz-Garay, C.; De la Vega-Beltran, J.L.; Delgado, R.; Labarca, P.; Felix, R.; Darszon, A. Inwardly rectifying K(+) channels in spermatogenic cells: Functional expression and implication in sperm capacitation. Dev. Biol. 2001, 234, 261–274. [Google Scholar] [CrossRef]
  109. Lybaert, P.; Vanbellinghen, A.M.; Quertinmont, E.; Petein, M.; Meuris, S.; Lebrun, P. KATP channel subunits are expressed in the epididymal epithelium in several mammalian species. Biol. Reprod. 2008, 79, 253–261. [Google Scholar] [CrossRef]
  110. Gupta, R.K.; Swain, K.D.; Singh, V.; Anand, M.; Choudhury, S.; Yadav, S.; Saxena, A.; Garg, S.K. Molecular characterization of voltage-gated potassium channel (Kv) and its importance in functional dynamics in bull spermatozoa. Theriogenology 2018, 114, 229–236. [Google Scholar] [CrossRef]
  111. Hernández-González, E.O.; Treviño, C.L.; Castellano, L.E.; de la Vega-Beltrán, J.L.; Ocampo, A.Y.; Wertheimer, E.; Visconti, P.E.; Darszon, A. Involvement of cystic fibrosis transmembrane conductance regulator in mouse sperm capacitation. J. Biol. Chem. 2007, 282, 24397–24406. [Google Scholar] [CrossRef]
  112. Puga Molina, L.C.; Pinto, N.A.; Torres Rodríguez, P.; Romarowski, A.; Vicens Sanchez, A.; Visconti, P.E.; Darszon, A.; Treviño, C.L.; Buffone, M.G. Essential Role of CFTR in PKA-Dependent Phosphorylation, Alkalinization, and Hyperpolarization During Human Sperm Capacitation. J. Cell. Physiol. 2017, 232, 1404–1414. [Google Scholar] [CrossRef]
  113. Shi, Q.X.; Yuan, Y.Y.; Roldan, E.R. gamma-Aminobutyric acid (GABA) induces the acrosome reaction in human spermatozoa. Hum. Reprod. 1997, 3, 677–683. [Google Scholar] [CrossRef] [PubMed]
  114. Ritta, M.N.; Bas, D.E.; Tartaglione, C.M. In vitro effect of gamma-aminobutyric acid on bovine spermatozoa capacitation. Mol. Reprod. Dev. 2004, 67, 478–486. [Google Scholar] [CrossRef] [PubMed]
  115. Ritta, M.N.; Calamera, J.C.; Bas, D.E. Occurrence of GABA and GABA receptors in human spermatozoa. Mol. Hum. Reprod. 1998, 4, 769–773. [Google Scholar] [CrossRef] [PubMed]
  116. Hu, J.H.; He, X.B.; Wu, Q.; Yan, Y.C.; Koide, S.S. Biphasic effect of GABA on rat sperm acrosome reaction: Involvement of GABA(A) and GABA(B) receptors. Arch. Androl. 2002, 48, 369–378. [Google Scholar] [CrossRef]
  117. Shi, Q.X.; Roldan, E.R. Evidence that a GABAA-like receptor is involved in progesterone-induced acrosomal exocytosis in mouse spermatozoa. Biol. Reprod. 1995, 52, 373–381. [Google Scholar] [CrossRef]
  118. Hwang, T.C.; Kirk, K.L. The CFTR ion channel: Gating, regulation, and anion permeation. Cold Spring Harb. Perspect. Med. 2013, 3, a009498. [Google Scholar] [CrossRef]
  119. Gadsby, D.C.; Vergani, P.; Csanády, L. The ABC protein turned chloride channel whose failure causes cystic fibrosis. Nature 2006, 440, 477–483. [Google Scholar] [CrossRef]
  120. Sheppard, D.N.; Welsh, M.J. Structure and function of the CFTR chloride channel. Physiol. Rev. 1999, 79, S23–S45. [Google Scholar] [CrossRef]
  121. Xu, W.M.; Shi, Q.X.; Chen, W.Y.; Zhou, C.X.; Ni, Y.; Rowlands, D.K.; Yi Liu, G.; Zhu, H.; Ma, Z.G.; Wang, X.F.; et al. Cystic fibrosis transmembrane conductance regulator is vital to sperm fertilizing capacity and male fertility. Proc. Natl. Acad. Sci. USA 2007, 104, 9816–9821. [Google Scholar] [CrossRef] [PubMed]
  122. Berdiev, B.K.; Qadri, Y.J.; Benos, D.J. Assessment of the CFTR and ENaC association. Mol. Biosyst. 2009, 5, 123–127. [Google Scholar] [CrossRef]
  123. Konig, J.; Schreiber, R.; Voelcker, T.; Mall, M.; Kunzelmann, K. The cystic fibrosis transmembrane conductance regulator (CFTR) inhibits ENaC through an increase in the intracellular Cl concentration. EMBO Rep. 2001, 2, 1047–1051. [Google Scholar] [CrossRef]
  124. Perez-Cornejo, P.; Arreola, J. Regulation of Ca(2+)-activated chloride channels by cAMP and CFTR in parotid acinar cells. Biophys. Res. Commun. 2004, 316, 612–617. [Google Scholar] [CrossRef]
  125. Cordero-Martínez, J.; Reyes-Miguel, T.; Rodríguez-Páez, L.; Garduño-Siciliano, L.; Maldonado-García, D.; Roa-Espitia, A.L.; Hernández-González, E.O. TMEM16A inhibition impedes capacitation and acquisition of hyperactivated motility in guinea pig sperm. J. Cell. Biochem. 2018, 119, 5944–5959. [Google Scholar] [CrossRef] [PubMed]
  126. Orta, G.; Ferreira, G.; José, O.; Treviño, C.L.; Beltrán, C.; Darszon, A. Human spermatozoa possess a calcium-dependent chloride channel that may participate in the acrosomal reaction. J. Physiol. 2012, 590, 2659–2675. [Google Scholar] [CrossRef]
  127. Kirkman-Brown, J.C.; Smith, D.J. Sperm motility: Is viscosity fundamental to progress? Mol. Hum. Reprod. 2011, 17, 539–544. [Google Scholar] [CrossRef]
  128. Maduke, M.; Miller, C.; Mindell, J.A. A decade of CLC chloride channels: Structure, mechanism, and many unsettled questions. Annu. Rev. Biophys. Biomol. Struct. 2000, 29, 411–438. [Google Scholar] [CrossRef]
  129. Fahlke, C. Ion permeation and selectivity in ClC-type chloride channels. Am. J. Physiol. Physiol. 2001, 280, 748–757. [Google Scholar] [CrossRef]
  130. Kawasaki, M.; Suzuki, M.; Uchida, S.; Sasaki, S.; Marumo, F. Stable and functional expression of the CIC-3 chloride channel in somatic cell lines. Neuron 1995, 14, 1285–1291. [Google Scholar] [CrossRef]
  131. Yeung, C.H.; Barfield, J.P.; Cooper, T.G. Chloride channels in physiological volume regulation of human spermatozoa. Biol. Reprod. 2005, 73, 1057–1063. [Google Scholar] [CrossRef]
  132. Smith, G.D.; Wolf, D.P.; Trautman, K.C.; da Cruz e Silva, E.F.; Greengard, P.; Vijayaraghavan, S. Primate sperm contain protein phosphatase 1, a biochemical mediator of motility. Biol. Reprod. 1996, 54, 719–727. [Google Scholar] [CrossRef] [PubMed]
  133. Blässe, A.K.; Oldenhof, H.; Ekhlasi-Hundrieser, M.; Wolkers, W.F.; Sieme, H.; Bollwein, H. Osmotic tolerance and intracellular ion concentrations of bovine sperm are affected by cryopreservation. Theriogenology 2012, 78, 1312–1320. [Google Scholar] [CrossRef] [PubMed]
  134. Duan, D.; Winter, C.; Cowley, S.; Hume, J.R.; Horowitz, B. Molecular identification of a volume-regulated chloride channel. Nature 1997, 390, 417–421. [Google Scholar] [CrossRef]
  135. Liu, S.W.; Li, Y.; Zou, L.L.; Guan, Y.T.; Peng, S.; Zheng, L.X.; Deng, S.M.; Zhu, L.Y.; Wang, L.W.; Chen, L.X. Chloride channels are involved in sperm motility and are downregulated in spermatozoa from patients with asthenozoospermia. Asian. J. Androl. 2017, 19, 418–424. [Google Scholar] [PubMed]
  136. Russell, J.M. Sodium-potassium-chloride cotransport. Physiol. Rev. 2000, 80, 211–276. [Google Scholar] [CrossRef] [PubMed]
  137. Meizel, S.; Turner, K.O. Chloride efflux during the progesterone-initiated human sperm acrosome reaction is inhibited by lavendustin A, a tyrosine kinase inhibitor. J. Androl. 1996, 17, 327–330. [Google Scholar]
  138. Pace, A.J.; Lee, E.; Athirakul, K.; Coffman, T.M.; O’Brien, D.A.; Koller, B.H. Failure of spermatogenesis in mouse lines deficient in the Na(+)-K(+)-2Cl() cotransporter. J. Clin. Investig. 2000, 105, 441–450. [Google Scholar] [CrossRef]
  139. Hess, K.C.; Jones, B.H.; Marquez, B.; Chen, Y.; Ord, T.S.; Kamenetsky, M.; Miyamoto, C.; Zippin, J.H.; Kopf, G.S.; Suarez, S.S.; et al. The “soluble” adenylyl cyclase in sperm mediates multiple signaling events required for fertilization. Dev. Cell 2005, 9, 249–259. [Google Scholar] [CrossRef]
  140. Okamura, N.; Tajima, Y.; Soejima, A.; Masuda, H.; Sugita, Y. Sodium bicarbonate in seminal plasma stimulates the motility of mammalian spermatozoa through direct activation of adenylate cyclase. J. Biol. Chem. 1985, 260, 9699–9705. [Google Scholar] [CrossRef]
  141. Demarco, I.A.; Espinosa, F.; Edwards, J.; Sosnik, J.; De La Vega-Beltran, J.L.; Hockensmith, J.W.; Kopf, G.S.; Darszon, A.; Visconti, P.E. Involvement of a Na+/HCO3 cotransporter in mouse sperm capacitation. J. Biol. Chem. 2003, 278, 7001–7009. [Google Scholar] [CrossRef]
  142. Törnroth-Horsefield, S.; Hedfalk, K.; Fischer, G.; Lindkvist-Petersson, K.; Neutze, R. Structural insights into eukaryotic aquaporin regulation. FEBS Lett. 2010, 584, 2580–2588. [Google Scholar] [CrossRef]
  143. Delgado-Bermúdez, A.; Ribas-Maynou, J.; Yeste, M. Relevance of Aquaporins for Gamete Function and Cryopreservation. Animals 2022, 12, 573. [Google Scholar] [CrossRef]
  144. Herrera, M.; Garvin, J.L. Aquaporins as gas channels. Pflugers Arch. 2011, 462, 623–630. [Google Scholar] [CrossRef]
  145. Agre, P. The aquaporin water channels. Proc. Am. Thorac. Soc. 2006, 3, 5–13. [Google Scholar] [CrossRef]
  146. Savage, D.F.; Egea, P.F.; Robles-Colmenares, Y.; O’Connell, J.D., 3rd; Stroud, R.M. Architecture and selectivity in aquaporins: 2.5 a X-ray structure of aquaporin Z. PLoS Biol. 2003, 1, e72. [Google Scholar] [CrossRef]
  147. Yasui, M.; Kwon, T.H.; Knepper, M.A.; Nielsen, S.; Agre, P. Aquaporin-6: An intracellular vesicle water channel protein in renal epithelia. Proc. Natl. Acad. Sci. USA 1999, 96, 5808–5813. [Google Scholar] [CrossRef]
  148. Fitzpatrick, K.L.; Reid, R.J. The ever expanding role of aquaglyceroporins: Confirmation of protein-facilitated boron transport. Plant Signal. Behav. 2010, 5, 132–133. [Google Scholar] [CrossRef] [PubMed]
  149. Calamita, G.; Perret, J.; Delporte, C. Aquaglyceroporins: Drug Targets for Metabolic Diseases? Front. Physiol. 2018, 9, 851. [Google Scholar] [CrossRef]
  150. Ishibashi, K.; Tanaka, Y.; Morishita, Y. The role of mammalian superaquaporins inside the cell: An update. Biochim. Biophys. Acta Biomembr. 2021, 1863, 183617. [Google Scholar] [CrossRef]
  151. Zhang, D.; Tan, Y.J.; Qu, F.; Sheng, J.Z.; Huang, H.F. Functions of water channels in male and female reproductive systems. Mol. Aspects Med. 2012, 33, 676–690. [Google Scholar] [CrossRef]
  152. Prieto-Martínez, N.; Morató, R.; Vilagran, I.; Rodríguez-Gil, J.E.; Bonet, S.; Yeste, M. Aquaporins in boar spermatozoa. Part II: Detection and localisation of aquaglyceroporin 3. Reprod. Fertil. Dev. 2015, 29, 703–711. [Google Scholar] [CrossRef]
  153. Prieto-Martinez, N.; Morato, R.; Muino, R.; Hidalgo, C.O.; Rodriguez-Gil, J.E.; Bonet, S.; Yeste, M. Aquaglyceroporins 3 and 7 in bull spermatozoa: Identification, localisation and their relationship with sperm cryotolerance. Reprod. Fertil. Dev. 2017, 29, 1249–1259. [Google Scholar] [CrossRef]
  154. Chen, Q.; Peng, H.; Lei, L.; Zhang, Y.; Kuang, H.; Cao, Y.; Shi, Q.; Ma, T.; Duan, E. Aquaporin 3 is a sperm water channel essential for postcopulatory sperm osmoadaptation and migration. Cell Res. 2011, 21, 922–933. [Google Scholar] [CrossRef]
  155. Alyasin, A.; Momeni, H.R.; Mahdieh, M. Aquaporin3 expression and the potential role of aquaporins in motility and mitochondrial membrane potential in human spermatozoa. Andrologia 2020, 52, e13588. [Google Scholar] [CrossRef]
  156. Vicente-Carrillo, A.; Ekwall, H.; Alvarez-Rodriguez, M.; Rodriguez-Martinez, H.; Álvarez-Rodríguez, M.; Rodríguez-Martínez, H. Membrane Stress During Thawing Elicits Redistribution of Aquaporin 7 But Not of Aquaporin 9 in Boar Spermatozoa. Reprod. Domest. Anim. 2016, 51, 665–679. [Google Scholar] [CrossRef]
  157. Bonilla-Correal, S.; Noto, F.; Garcia-Bonavila, E.; Rodríguez-Gil, J.E.; Yeste, M.; Miró, J. First evidence for the presence of aquaporins in stallion sperm. Reprod. Domest. Anim. 2017, 52, 61–64. [Google Scholar] [CrossRef]
  158. Moretti, E.; Terzuoli, G.; Mazzi, L.; Iacoponi, F.; Collodel, G. Immunolocalization of aquaporin 7 in human sperm and its relationship with semen parameters. Syst. Biol. Reprod. Med. 2012, 58, 129–135. [Google Scholar] [CrossRef]
  159. Yeung, C.H.; Callies, C.; Rojek, A.; Nielsen, S.; Cooper, T.G. Aquaporin Isoforms Involved in Physiological Volume Regulation of Murine Spermatozoa. Biol. Reprod. 2009, 80, 350–357. [Google Scholar] [CrossRef]
  160. Suzuki-Toyota, F.; Ishibashi, K.; Yuasa, S. Immunohistochemical localization of a water channel, aquaporin 7 (AQP7), in the rat testis. Cell Tissue Res. 1999, 295, 279–285. [Google Scholar] [CrossRef]
  161. Calamita, G.; Mazzone, A.; Bizzoca, A.; Svelto, M. Possible Involvement of Aquaporin-7 and -8 in Rat Testis Development and Spermatogenesis. Biochem. Biophys. Res. Commun. 2001, 288, 619–625. [Google Scholar] [CrossRef] [PubMed]
  162. Prieto-Martínez, N.; Vilagran, I.; Morató, R.; Rodríguez-Gil, J.E.; Yeste, M.; Bonet, S.; Prieto-Martinez, N.; Vilagran, I.; Morato, R.; Rodriguez-Gil, J.E.; et al. Aquaporins 7 and 11 in boar spermatozoa: Detection, localisation and relationship with sperm quality. Reprod. Fertil. Dev. 2014, 28, 663–672. [Google Scholar] [CrossRef] [PubMed]
  163. Yeung, C.H.; Cooper, T.G. Aquaporin AQP11 in the testis: Molecular identity and association with the processing of residual cytoplasm of elongated spermatids. Reproduction 2010, 139, 209–216. [Google Scholar] [CrossRef] [PubMed]
  164. Laforenza, U.; Pellavio, G.; Marchetti, A.; Omes, C.; Todaro, F.; Gastaldi, G. Aquaporin-Mediated Water and Hydrogen Peroxide Transport Is Involved in Normal Human Spermatozoa Functioning. Int. J. Mol. Sci. 2017, 18, 66. [Google Scholar] [CrossRef]
  165. Cooper, T.G.; Yeung, C.-H. Acquisition of volume regulatory response of sperm upon maturation in the epididymis and the role of the cytoplasmic droplet. Microsc. Res. Tech. 2003, 61, 28–38. [Google Scholar] [CrossRef]
  166. Drevius, L.-O.; Eriksson, H. Osmotic swelling of mammalian spermatozoa. Exp. Cell Res. 1966, 42, 136–156. [Google Scholar] [CrossRef]
  167. Rossato, M.; Virgilio, F.D.; Foresta, C. Involvement of osmo-sensitive calcium influx in human sperm activation. Mol. Hum. Reprod. 1996, 2, 903–909. [Google Scholar] [CrossRef]
  168. Zanetti, N.; Mayorga, L.S. Acrosomal Swelling and Membrane Docking Are Required for Hybrid Vesicle Formation During the Human Sperm Acrosome Reaction. Biol. Reprod. 2009, 81, 396–405. [Google Scholar] [CrossRef]
  169. Rivlin, J.; Mendel, J.; Rubinstein, S.; Etkovitz, N.; Breitbart, H. Role of Hydrogen Peroxide in Sperm Capacitation and Acrosome Reaction. Biol. Reprod. 2004, 70, 518–522. [Google Scholar] [CrossRef]
  170. Yeung, C.H.; Callies, C.; Tüttelmann, F.; Kliesch, S.; Cooper, T.G. Aquaporins in the human testis and spermatozoa-identification, involvement in sperm volume regulation and clinical relevance. Int. J. Androl. 2010, 33, 629–641. [Google Scholar] [CrossRef]
  171. Ho, H.C. Redistribution of nuclear pores during formation of the redundant nuclear envelope in mouse spermatids. J. Anat. 2010, 216, 525–532. [Google Scholar] [CrossRef] [PubMed]
  172. Brown, S.G.; Publicover, S.J.; Barratt, C.L.R.; Martins da Silva, S.J. Human sperm ion channel (dys)function: Implications for fertilization. Hum. Reprod. Update 2019, 25, 758–776. [Google Scholar] [CrossRef] [PubMed]
  173. Ho, H.C.; Suarez, S.S. Characterization of the intracellular calcium store at the base of the sperm flagellum that regulates hyperactivated motility. Biol. Reprod. 2003, 68, 1590–1596. [Google Scholar] [CrossRef] [PubMed]
  174. Ho, H.C.; Suarez, S.S. An inositol 1,4,5-trisphosphate receptor-gated intracellular Ca(2+) store is involved in regulating sperm hyperactivated motility. Biol. Reprod. 2001, 65, 1606–1615. [Google Scholar] [CrossRef]
  175. Naaby-Hansen, S.; Wolkowicz, M.J.; Klotz, K.; Bush, L.A.; Westbrook, V.A.; Shibahara, H.; Shetty, J.; Coonrod, S.A.; Reddi, P.P.; Shannon, J.; et al. Co-localization of the inositol 1,4,5-trisphosphate receptor and calreticulin in the equatorial segment and in membrane bounded vesicles in the cytoplasmic droplet of human spermatozoa. Mol. Hum. Reprod. 2001, 7, 923–933. [Google Scholar] [CrossRef]
  176. Checchetto, V.; Leanza, L.; De Stefani, D.; Rizzuto, R.; Gulbins, E.; Szabo, I. Mitochondrial K+ channels and their implications for disease mechanisms. Pharmacol. Ther. 2021, 227, 107874. [Google Scholar] [CrossRef]
  177. Szabo, I.; Zoratti, M. Mitochondrial channels: Ion fluxes and more. Physiol. Rev. 2014, 94, 519–608. [Google Scholar] [CrossRef]
  178. Feno, S.; Rizzuto, R.; Raffaello, A.; Vecellio Reane, D. The molecular complexity of the Mitochondrial Calcium Uniporter. Cell Calcium. 2021, 93, 102322. [Google Scholar] [CrossRef]
  179. Heslop, K.A.; Milesi, V.; Maldonado, E.N. VDAC modulation of cancer metabolism: Advances and therapeutic challenges. Front. Physiol. 2021, 12, 742839. [Google Scholar] [CrossRef]
  180. Mannella, C.A. VDAC-A primal perspective. Int. J. Mol. Sci. 2021, 22, 1685. [Google Scholar] [CrossRef]
  181. Triphan, X.; Menzel, V.A.; Petrunkina, A.M.; Cassará, M.C.; Wemheuer, W.; Hinsch, K.D.; Hinsch, E. Localisation and function of voltage-dependent anion channels (VDAC) in bovine spermatozoa. Pflugers Arch. 2008, 455, 677–686. [Google Scholar] [CrossRef] [PubMed]
  182. Kwon, W.S.; Park, Y.J.; Mohamed, E.S.A.; Pang, M.G. Voltage-dependent anion channels are a key factor of male fertility. Fertil. Steril. 2013, 99, 354–361. [Google Scholar] [CrossRef] [PubMed]
  183. Rizzuto, R.; De Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as sensors and regulators of calcium signalling. Nat. Rev. Mol. Cell Biol. 2012, 13, 566–578. [Google Scholar] [CrossRef] [PubMed]
  184. Kirichok, Y.; Krapivinsky, G.; Clapham, D.E. The mitochondrial calcium uniporter is a highly selective ion channel. Nature 2004, 427, 360–364. [Google Scholar] [CrossRef] [PubMed]
  185. Bravo, A.; Treulen, F.; Uribe, P.; Boguen, R.; Felmer, R.; Villegas, J.V. Effect of mitochondrial calcium uniporter blocking on human spermatozoa. Andrologia 2015, 47, 662–668. [Google Scholar] [CrossRef]
  186. Garlid, K.D.; Paucek, P. The mitochondrial potassium cycle. IUBMB Life 2001, 52, 153–158. [Google Scholar] [CrossRef]
  187. Inoue, I.; Nagase, H.; Kishi, K.; Higuti, T. ATP-sensitive K+ channel in the mitochondrial inner membrane. Nature 1991, 352, 244–247. [Google Scholar] [CrossRef]
  188. Tinker, A.; Aziz, Q.; Li, Y.; Specterman, M. ATP-sensitive potassium channels and their physiological and pathophysiological roles. Compr. Physiol. 2018, 8, 1463–1511. [Google Scholar]
  189. Stowe, D.F.; Gadicherla, A.K.; Zhou, Y.; Aldakkak, M.; Cheng, Q.; Kwok, W.M.; Jiang, M.T.; Heisner, J.S.; Yang, M.; Camara, A.K. Protection against cardiac injury by small Ca2+-sensitive K+ channels identified in guinea pig cardiac inner mitochondrial membrane. Biochim. Biophys. Acta 2013, 1828, 427–442. [Google Scholar] [CrossRef]
  190. De Marchi, U.; Sassi, N.; Fioretti, B.; Catacuzzeno, L.; Cereghetti, G.M.; Szabó, I.; Zoratti, M. Intermediate conductance Ca2+- activated potassium channel (KCa3.1) in the inner mitochondrial membrane of human colon cancer cells. Cell Calcium 2009, 45, 509–516. [Google Scholar] [CrossRef]
  191. Balderas, E.; Zhang, J.; Stefani, E.; Toro, L. Mitochondrial BKCa channel. Front. Physiol. 2015, 6, 104–114. [Google Scholar] [CrossRef] [PubMed]
  192. Szabò, I.; Bock, J.; Jekle, A.; Soddemann, M.; Adams, C.; Lang, F.; Zoratti, M.; Gulbins, E. A novel potassium channel in lymphocyte mitochondria. J. Biol. Chem. 2005, 280, 12790–12798. [Google Scholar] [CrossRef] [PubMed]
  193. Kicinska, A.; Augustynek, B.; Kulawiak, B.; Jarmuszkiewicz, W.; Szewczyk, A.; Bednarczyk, P. A largeconductance calcium-regulated K+ channel in human dermal fibroblast mitochondria. Biochem. J. 2016, 473, 4457–4471. [Google Scholar] [CrossRef]
  194. Testai, L.; Barrese, V.; Soldovieri, M.V.; Ambrosino, P.; Martelli, A.; Vinciguerra, I.; Miceli, F.; Greenwood, I.A.; Curtis, M.J.; Breschi, M.C.; et al. Expression and function of Kv7.4 channels in rat cardiac mitochondria: Possible targets for cardioprotection. Cardiovasc. Res. 2016, 110, 40–50. [Google Scholar] [CrossRef] [PubMed]
  195. Peruzzo, R.; Mattarei, A.; Azzolini, M.; Becker-Flegler, K.A.; Romio, M.; Rigoni, G.; Carrer, A.; Biasutto, L.; Parrasia, S.; Kadow, S.; et al. Insight into the mechanism of cytotoxicity of membrane-permeant psoralenic Kv1.3 channel inhibitors by chemical dissection of a novel member of the family. Redox Biol. 2020, 37, 101705. [Google Scholar] [CrossRef]
  196. Brierley, G.P. Energy-linked alteration of mitochondrial permeability to anions. Biochem. Biophys. Res. Commun. 1969, 35, 396–402. [Google Scholar]
  197. Singh, H. Mitochondrial ion channels in cardiac function. Am. J. Physiol. Cell Physiol. 2021, 321, C812–C825. [Google Scholar] [CrossRef]
  198. Ponnalagu, D.; Singh, H. Anion channels of mitochondria. Handb. Exp. Pharmacol. 2017, 240, 71–101. [Google Scholar]
  199. Myers, K.; Somanath, P.R.; Berryman, M.; Vijayaraghavan, S. Identification of chloride intracellular channel proteins in spermatozoa. FEBS Lett. 2004, 566, 136–140. [Google Scholar] [CrossRef]
  200. Bernardi, P.; von Stockum, S. The permeability transition pore as a Ca2+ release channel: New answers to an old question. Cell Calcium 2012, 52, 22–27. [Google Scholar] [CrossRef]
  201. Torrezan-Nitao, E.; Boni, R.; Marques-Santos, L.F. Mitochondrial permeability transition pore (MPTP) desensitization increases sea urchin spermatozoa fertilization rate. Cell Biol. Int. 2016, 40, 1071–1083. [Google Scholar] [CrossRef] [PubMed]
  202. Bernardi, P.; Scorrano, L.; Colonna, R.; Petronilli, V.; Di Lisa, F. Mitochondria and cell death. Mechanistic aspects and methodological issues. Eur. J. Biochem. 1999, 264, 687–701. [Google Scholar] [CrossRef] [PubMed]
  203. Sileikyte, J.; Petronilli, V.; Zulian, A.; Dabbeni-Sala, F.; Tognon, G.; Nikolov, P.; Bernardi, P.; Ricchelli, F. Regulation of the inner membrane mitochondrial permeability transition by the outer membrane translocator protein (peripheral benzodiazepine receptor). J. Biol. Chem. 2011, 286, 1046–1053. [Google Scholar] [CrossRef] [PubMed]
  204. Richardson, A.P.; Halestrap, A.P. Quantification of active mitochondrial permeability transition pores using GNX-4975 inhibitor titrations provides insights into molecular identity. Biochem. J. 2016, 473, 1129–1140. [Google Scholar] [CrossRef]
  205. Petronilli, V.; Cola, C.; Massari, S.; Colonna, R.; Bernardi, P. Physiological effectors modify voltage sensing by the cyclosporin A- sensitive permeability transition pore of mitochondria. J. Biol. Chem. 1993, 268, 21939–21945. [Google Scholar] [CrossRef]
  206. Nicolli, A.; Petronilli, V.; Bernardi, P. Modulation of the mitochondrial cyclosporin A-sensitive permeability transition pore by matrix pH. Evidence that the pore open-closed probability is regulated by reversible histidine protonation. Biochemistry 1993, 32, 4461–4465. [Google Scholar] [CrossRef]
  207. Haworth, R.A.; Hunter, D.R. The Ca2+-induced membrane transition in mitochondria II. Nature of the Ca2+ trigger site. Arch. Biochem. Biophys. 1979, 195, 460–467. [Google Scholar] [CrossRef]
  208. Peng, T.-I.; Jou, M.-J. Oxidative stress caused by mitochondrial calcium overload. Ann. N. Y. Acad. Sci. 2010, 1201, 183–188. [Google Scholar] [CrossRef]
  209. Carafoli, E. Intracellular calcium homeostasis. Annu. Rev. Biochem. 1987, 56, 395–433. [Google Scholar] [CrossRef]
  210. Dan, J.C. Studies on the acrosome. I. Reaction to egg-water and other stimuli. Biol. Bull. 1952, 103, 54–66. [Google Scholar] [CrossRef]
  211. Publicover, S.; Harper, C.V.; Barratt, C. [Ca2+]i signalling in sperm—Making the most of what you’ve got. Nat. Cell Biol. 2007, 9, 235–242. [Google Scholar] [CrossRef] [PubMed]
  212. Wennemuth, G.; Westenbroek, R.E.; Xu, T.; Hille, B.; Babcock, D.F. CaV2.2 and CaV2.3 (N- and R-type) Ca2+ channels in depolarization-evoked entry of Ca2+ into mouse sperm. J. Biol. Chem. 2000, 275, 21210–21217. [Google Scholar] [CrossRef] [PubMed]
  213. Beltrán, J.F.; Belén, L.H.; Lee-Estevez, M.; Figueroa, E.; Dumorné, K.; Farias, J.G. The voltage-gated T-type Ca2+ channel is key to the sperm motility of Atlantic salmon (Salmo salar) Fish. Physiol. Biochem. 2020, 46, 1825–1831. [Google Scholar]
  214. Catterall, W.A. Voltage-gated calcium channels. Cold Spring Harb. Perspect. Biol. 2011, 3, a003947. [Google Scholar]
  215. Clapham, D.E.; Julius, D.; Montell, C.; Schultz, G. International Union of Pharmacology. XLIX. Nomenclature and structure-function relationships of transient receptor potential channels. Pharmacol. Rev. 2005, 57, 427–450. [Google Scholar] [CrossRef]
  216. De Toni, L.; Garolla, A.; Menegazzo, M.; Magagna, S.; Di Nisio, A.; Šabović, I.; Rocca, M.S.; Scattolini, V.; Filippi, A.; Foresta, C. Heat Sensing Receptor TRPV1 Is a Mediator of Thermotaxis in Human Spermatozoa. PLoS ONE. 2016, 11, e0167622. [Google Scholar] [CrossRef]
  217. Chen, Y.; Wang, H.; Wang, F.; Chen, C.; Zhang, P.; Song, D.; Luo, T.; Xu, H.; Zeng, X. Sperm motility modulated by Trpv1 regulates zebrafish fertilization. Theriogenology 2020, 151, 41–51. [Google Scholar] [CrossRef]
  218. Cao, S.; Anishkin, A.; Zinkevich, N.S.; Nishijima, Y.; Korishettar, A.; Wang, Z.; Fang, J.; Wilcox, D.A.; Zhang, D.X. Transient receptor potential vanilloid 4 (TRPV4) activation by arachidonic acid requires protein kinase A-mediated phosphorylation. J. Biol. Chem. 2018, 293, 5307–5322. [Google Scholar] [CrossRef]
  219. Krasznai, Z.; Krasznai, Z.T.; Morisawa, M.; Bazsáné, Z.K.; Hernádi, Z.; Fazekas, Z.; Trón, L.; Goda, K.; Márián, T. Role of the Na+/Ca2+ exchanger in calcium homeostasis and human sperm motility regulation. Cell Motil. Cytoskelet. 2006, 63, 66–76. [Google Scholar] [CrossRef]
  220. Orta, G.; de la Vega-Beltran, J.L.; Martín-Hidalgo, D.; Santi, C.M.; Visconti, P.E.; Darszon, A. CatSper channels are regulated by protein kinase A. J. Biol. Chem. 2018, 293, 16830–16841. [Google Scholar] [CrossRef]
  221. DeCoursey, T.E. Voltage-gated proton channels: Molecular biology, physiology, and pathophysiology of the HV family. Physiol. Rev. 2013, 93, 599–652. [Google Scholar] [CrossRef] [PubMed]
  222. Lishko, P.V.; Botchkina, I.L.; Fedorenko, A.; Kirichok, Y. Acid Extrusion from Human Spermatozoa Is Mediated by Flagellar Voltage-Gated Proton Channel. Cell 2010, 140, 327–337. [Google Scholar] [CrossRef] [PubMed]
  223. Miller, M.R.; Kenny, S.J.; Mannowetz, N.; Mansell, S.A.; Wojcik, M.; Mendoza, S.; Zucker, R.S.; Xu, K.; Lishko, P.V. Asymmetrically Positioned Flagellar Control Units Regulate Human Sperm Rotation. Cell Rep. 2018, 24, 2606–2613. [Google Scholar] [CrossRef] [PubMed]
  224. Chen, B.; Li, S.; Yan, Y.; Duan, Y.; Chang, S.; Wang, H.; Ji, W.; Wu, X.; Si, W. Cryopreservation of cynomolgus macaque (Macaca fascicularis) sperm with glycerol and ethylene glycol, and its effect on sperm-specific ion channels—CatSper and Hv1. Theriogenology 2017, 104, 37–42. [Google Scholar] [CrossRef]
  225. Yeste, M.; Llavanera, M.; Mateo-Otero, Y.; Catalán, J.; Bonet, S.; Pinart, E. HVCN1 Channels Are Relevant for the Maintenance of Sperm Motility During In Vitro Capacitation of Pig Spermatozoa. Int. J. Mol. Sci. 2020, 21, 3255. [Google Scholar] [CrossRef]
  226. Keshtgar, S.; Ghanbari, H.; Ghani, E.; Shid Moosavi, S.M. Effect of CatSper and Hv1 Channel Inhibition on Progesterone Stimulated Human Sperm. J. Reprod. Infertil. 2018, 19, 133–139. [Google Scholar]
  227. Ghanbari, H.; Keshtgar, S.; Zare, H.R.; Gharesi-Fard, B. Inhibition of CatSper and Hv1 Channels and NOX5 Enzyme Affect Progesterone-Induced Increase of Intracellular Calcium Concentration and ROS Generation in Human Sperm. Iran. J. Med. Sci. 2019, 44, 127–134. [Google Scholar] [PubMed]
Figure 1. Ion channels of spermatozoa. Created with BioRender.com (Toronto, ON, Canada, https://app.biorender.com/).
Figure 1. Ion channels of spermatozoa. Created with BioRender.com (Toronto, ON, Canada, https://app.biorender.com/).
Biomedicines 11 02519 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Benko, F.; Urminská, D.; Ďuračka, M.; Tvrdá, E. Signaling Roleplay between Ion Channels during Mammalian Sperm Capacitation. Biomedicines 2023, 11, 2519. https://doi.org/10.3390/biomedicines11092519

AMA Style

Benko F, Urminská D, Ďuračka M, Tvrdá E. Signaling Roleplay between Ion Channels during Mammalian Sperm Capacitation. Biomedicines. 2023; 11(9):2519. https://doi.org/10.3390/biomedicines11092519

Chicago/Turabian Style

Benko, Filip, Dana Urminská, Michal Ďuračka, and Eva Tvrdá. 2023. "Signaling Roleplay between Ion Channels during Mammalian Sperm Capacitation" Biomedicines 11, no. 9: 2519. https://doi.org/10.3390/biomedicines11092519

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop