Next Article in Journal
The Improved Antineoplastic Activity of Thermophilic L-Asparaginase Tli10209 via Site-Directed Mutagenesis
Previous Article in Journal
Integration of Proteomic and Metabolomic Data Reveals the Lipid Metabolism Disorder in the Liver of Rats Exposed to Simulated Microgravity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

pH Homeodynamics and Male Fertility: A Coordinated Regulation of Acid-Based Balance during Sperm Journey to Fertilization

Institute of Reproductive Medicine, Medical School, Nantong University, Nantong 226019, China
*
Authors to whom correspondence should be addressed.
Biomolecules 2024, 14(6), 685; https://doi.org/10.3390/biom14060685
Submission received: 5 May 2024 / Revised: 3 June 2024 / Accepted: 10 June 2024 / Published: 12 June 2024

Abstract

:
pH homeostasis is crucial for spermatogenesis, sperm maturation, sperm physiological function, and fertilization in mammals. HCO3 and H+ are the most significant factors involved in regulating pH homeostasis in the male reproductive system. Multiple pH-regulating transporters and ion channels localize in the testis, epididymis, and spermatozoa, such as HCO3 transporters (solute carrier family 4 and solute carrier family 26 transporters), carbonic anhydrases, and H+-transport channels and enzymes (e.g., Na+-H+ exchangers, monocarboxylate transporters, H+-ATPases, and voltage-gated proton channels). Hormone-mediated signals impose an influence on the production of some HCO3 or H+ transporters, such as NBCe1, SLC4A2, MCT4, etc. Additionally, ion channels including sperm-specific cationic channels for Ca2+ (CatSper) and K+ (SLO3) are directly or indirectly regulated by pH, exerting specific actions on spermatozoa. The slightly alkaline testicular pH is conducive to spermatogenesis, whereas the epididymis’s low HCO3 concentration and acidic lumen are favorable for sperm maturation and storage. Spermatozoa pH increases substantially after being fused with seminal fluid to enhance motility. In the female reproductive tract, sperm are subjected to increasing concentrations of HCO3 in the uterine and fallopian tube, causing a rise in the intracellular pH (pHi) of spermatozoa, leading to hyperpolarization of sperm plasma membranes, capacitation, hyperactivation, acrosome reaction, and ultimately fertilization. The physiological regulation initiated by SLC26A3, SLC26A8, NHA1, sNHE, and CFTR localized in sperm is proven for certain to be involved in male fertility. This review intends to present the key factors and characteristics of pHi regulation in the testes, efferent duct, epididymis, seminal fluid, and female reproductive tract, as well as the associated mechanisms during the sperm journey to fertilization, proposing insights into outstanding subjects and future research trends.

1. Introduction

Infertility affects approximately 15% of couples worldwide [1], with male infertility accounting for approximately 50% of these cases [2,3]. Several factors contribute to male infertility, including idiopathic abnormalities of sperm quality [4], infection and inflammation of the reproductive tract [5], genetic factors [6], and sperm ion channel dysfunction [7]. Spermatogenesis, sperm maturation, storage, and fertilization require stringent regulation of the internal microenvironment [8,9]. First, spermatogenesis is a complex process of cellular transformation in which spermatogonia undergo two meiotic divisions to form spermatozoa. In the epididymis, spermatozoa become motile and gain the ability to fertilize. Mature sperm then migrate to the vas deferens and its ampulla for storage until ejaculation [10,11]. Throughout these processes, the components of the luminal fluid, e.g., H+, Cl, Na+, and HCO3, keep changing [12], and dynamic adjustments of these ions and pH maintenance are critical for male fertility. An acidic pH (6.6–6.8) in the male reproductive tract reduces sperm motility and facilitates their maturation and storage [13]; after being deposited in the vagina, they come in contact with cervical mucus before proceeding to enter the uterus. The contraction of the uterine muscles facilitates the movement of spermatozoa to the fallopian tubes. Spermatozoa move toward the periphery of the oocyte via thermotropism and chemotropism, before sperm-egg fusion [14]. The ions present notable alterations throughout this process, leading to significant variations in the extracellular pH encountered by spermatozoa. The pH in the female reproductive tract forms a gradient, gradually increasing from the vagina (approximately pH 4.3) to the cervix (pH 6.5–7.5) and further increasing in the fallopian tubes [15]. In addition, the alkaline seminal fluid can neutralize the acidic pH of the vagina, increasing the vaginal pH from 4.3 to 7.2 [16] (Figure 1). Notably, alkalinization of the intracellular pH (pHi) of spermatozoa is required for sperm capacitation in mammals as well as for hyperactivation of sperm motility, chemotaxis, the acrosome reaction (AR), and fertilization [17,18]. HCO3 and H+ transport are critical processes in pHi regulation [19]. pH homeostasis in the male reproductive tract is managed primarily by the transporters solute carrier 4 (SLC4) and solute carrier 26 (SLC26), Na+-H+ exchangers (NHEs), carbonic anhydrases (CAs), monocarboxylate transporters (MCTs), and hydrogen voltage-gated channels (HVCN1), as well as by other channels and enzymes associated with pH regulation, including Ca2+ channels (CatSper) [20], K+ channels (SLO3) [21], and soluble adenylate cyclase (sAC) [22]. A few pH regulators in the testis and epididymis are under the control of the endocrine system [23,24,25]. Moreover, Sertoli cells (SCs) are the primary targets of hormonal transduction signals, contributing to the well-established testicular luminal microenvironment, and lactate transportation to germ cells that is metabolized from glucose in SCs [26]. Evidence suggests NBCe1, SLC4A2 [25], MCT4 [27], and MCT2 [24] levels in SCs are correlated strongly with estrogen, testosterone, follicle-stimulating hormone (FSH), and insulin. Importantly, the actions of pH regulation in male fertility are supported by multiple clinical infertility symptoms, such as asthenozoospermia, teratozoospermia, and by the aberrant sperm physiological functions of rodent sperm. A few pH regulators that are localized in the male reproductive system with determined phenotype or clinical symptoms are summarized in Table 1. This review provides a comprehensive overview of critical molecules that regulate intracellular and extracellular pH before fertilization and investigates the mechanisms responsible for pH homeostasis within microenvironments.

2. pH Homeostasis on Spermatogenesis

Spermatogenesis initiates with the differentiation of spermatogonia into spermatocytes, which undergo meiosis to form round spermatids, ultimately developing into spermatozoa by spermiogenesis [57]. Spermatogenesis is an orchestrating process that requires a highly specialized testicular microenvironment. In particular, HCO3 and H+ play a vital role in testicular pHi homeostasis [58]. SLC4 and MCTs are primarily covered in the regulation of pHi in the testis.
The SLC4 family participates in HCO3 transport and consists of 10 members, namely the Cl/HCO3 transporters (SLC4A1-AE1, SLC4A2-AE2, SLC4A3-AE3), Na+/HCO3 transporters (SLC4A4-NBCe1, SLC4A5-NBCe2, SLC4A7-NBCn1, SLC4A10-NBCn2, SLC4A8-NDCBE), SLC4A9 (AE4), and SLC4A11 (BTR1) [59]. SCs play a vital role in pHi regulation in the testes [60]. Multiple HCO3 transporters are expressed in SCs, including NDCBE, NBCe1, NBCn1, and SLC4A2, which regulate pHi in the testis [25,61]. NDCBE is localized in the apical plasma membrane of SCs, whereas NBCe1 and NBCn1 are found in the basal plasma membrane. NDCBE transports Na+ and HCO3 from the apical lumen into the cell, exchanges them with intracellular Cl, and promotes cellular alkalinization. Simultaneously, NBCe1 and NBCn1 transport Na+ and HCO3 from the basal region into the cell and induce cellular alkalinization [62].
Elevated estrogen has been shown to disturb the content of NBCe1, NBCn1, and NDCBE in human SCs, impairing spermatogenesis and male fertility [62]. Estrogen receptor α (ERα) deficiency leads to decreased levels of NBCe1, CA XIV, and NHE3 (a kind of Na+-H+ exchanger) in the epididymal ducts, and increased luminal pH [23]. SLC4A2 is localized in SCs [25], spermatogonia, and mature spermatozoa [28], and its mutation blocks HCO3-controlled cAMP signaling, disrupting spermatogenesis and causing male sterility [29]. Estrogen increases SLC4A2 levels in SCs and affects HCO3 transport in rats [25]. In contrast, AE1 and AE3 are not expressed in most male reproductive tract tissues. AE1 is expressed only in the Leydig cells of the testis, whereas AE3 is expressed in spermatogenic cells and seminal vesicle cells [63]. The physiological functions of AE1 and AE3 in spermatogenesis remain to be elucidated.
The SLC16 family consists of 14 members, commonly termed the MCT family, which play indispensable roles in nutrient transport, cellular metabolism, and pH regulation [64]. MCT1 is expressed in spermatogonia, spermatocytes, vas deferens, epididymis, and the tails of spermatozoa [30,31,65], whereas MCT2 is expressed in elongated spermatocytes and the tails of mature spermatozoa [31]. MCT1 deficiency results in altered morphology of the seminiferous tubules and SCs, decreased serum 17β-estradiol levels, and the failure of spermatogenesis [32]. In the testes, lactate and pyruvate produced by SCs are the principal energy sources for spermatogenic cells. MCT1-MCT4 transports lactate from SCs to spermatocytes and spermatozoa [66,67]. High-intensity exercise leads to impaired spermatogenesis, decreased viability of SCs and reduced MCT-4, inhibiting lactate export from SCs [68]. In mouse spermatozoa, MCT1 and MCT2 interact with CD147 to mediate the L-lactate transport to spermatogenic cells while decreasing pHi [65]. In addition, MCT1 and MCT4 interact specifically with CD147, and may contribute to MCT expression on the cell surface [67]. MCT5 and MCT8 are also expressed in the mouse testis, epididymis, and epididymal spermatozoa, but the mechanism of action remains to be investigated [65]. Hormones can influence lactate transport from SCs to spermatogenic cells by modulating MCT levels, affecting energy metabolism in the testes. The exposure of SCs to high levels of testosterone and its metabolite 5α-dihydrotestosterone leads to a considerable reduction in the production of MCT4 and lactate [69]. High levels of testosterone and follicle-stimulating hormone (FSH) have been found to decrease MCT2 content [24]. In addition, MCT4 and lactate production are notably reduced in the SCs of patients with diabetes under insulin-deprivation conditions [27].
In the testes, SLC4 and MCT transport HCO3, H+, Na+, and lactate, stabilizing intracellular and extracellular pH within a narrow range, which is essential for maintaining spermatogenesis (Figure 2). In addition to their direct involvement in pHi regulation, SCs supply lactate to germ cells [70]. Therefore, further research on lactate production and ion transport in SCs is essential to elucidate the mechanisms underlying spermatogenesis.

3. pH Involvement in Sperm Maturation

Sperm matures in the epididymis and acquires motility and the ability to fertilize. The epididymis is divided into four parts based on histological, structural, and functional characteristics: the initial segment, caput, corpus, and cauda [57]. As the fluid passes through the rete testis to the epididymal lumen, dramatic changes in composition, including Na+, Cl, and HCO3 reabsorption, H+ secretion, and luminal acidification are observed [12]. The acidic microenvironment of the epididymal luminal fluid and the low concentration of HCO3 are foundational for sperm maturation [71].
The acidic lumen of the epididymis is achieved primarily through HCO3 reabsorption and H+ efflux [13,72]. In the epididymis, NHE3 [41], SLC4A2 [73], NBCe1 [74], NBCn1 [75], CAs [76], and H+-ATPases [77] act synergistically to maintain an acidic microenvironment. NHE3 is expressed throughout the epididymis of rats and is essential for epididymal fluid reabsorption and pH regulation [38]. NHE3 knockout mice exhibited an accumulation of epididymal fluid and abnormal morphological differentiation of epithelial cells, ultimately leading to male sterility [40]. CAs catalyze the reversible hydration of CO2 to H2CO3, thus contributing to the maintenance of intracellular HCO3 concentrations, which in turn influence the viability of spermatozoa. They are vital enzymes that maintain the acid-base balance of fluids [78]. The CA inhibitor acetazolamide significantly inhibits the acidification of epididymal luminal fluid [79]. The significant processes of HCO3 reabsorption in the epididymis include the secretion of H+ into the lumen by the NHE3 transporter and the simultaneous reabsorption of Na+ from the lumen into the principal cells. H+ ions in the lumen react with HCO3 to form H2CO3, while CA IV and CA XIV catalyze to form CO2 and water. CO2 enters the principal cells and is hydrated by CA III in the cytoplasm to form H2CO3, rapidly breaking down into H+ and HCO3. H+ is then recycled into the lumen via NHE3 (Figure 2). HCO3 diffuses passively across the basolateral side of the basement membrane via NBCe1 or SLC4A2 [40,80,81]. ERα regulates NHE3 and CAs in the epididymis. ERα knockout mice have reduced or absent NHE3 and CAs, resulting in increased epididymis luminal pH, decreased sperm viability, and male sterility [23,40]. In addition to CA III, CA IV and CA XIV, CA II, CA IX, and CA XII are also expressed in the epididymis, regulating the pH of the epididymal luminal fluid synergetically [80,81,82]. However, the underlying mechanisms are waiting to be revealed. Moreover, CA IX and CA XII are expressed on the basolateral plasma membrane of efferent duct epithelial cells and mediate transmembrane transport of Cl, Na+ and water reabsorption in the lumen [82]. CA-mediated transmembrane ion transport is coupled to SLC-regulated HCO3 transport, implying that CAs and SLC cooperatively regulate pH homeostasis in the luminal fluid of the efferent ducts. In conclusion, CAs are potential male reproductive pH homeostasis regulators, which further promote sperm maturation in conjunction with the HCO3 transporter.
H+-ATPase is expressed in the initial segment of the epididymis, in the narrow cells in the caput, and the clear cells in the cauda [51,52,53]; it is involved dominantly in the secretion of H+ into the lumen of the epididymis and maintenance of the acidic pH of the epididymal luminal fluid (Figure 2) [52,83]. H+-ATPase exists as two subunits, B1 and B2. Mice with a knockout of the B1 subunit (Atp6v1b1) produce significantly more alkaline urine and are less capable of acid metabolism [84]. The FOXI1 mediates the synthesis of ATP6V1B1 in the mouse epididymis. Knockout of Foxi1 blocks epididymal sperm maturation, disturbing fertilization due to insufficient sperm in the female reproductive tract [56]. Cadmium, a heavy metal environmental pollutant, has been shown to accumulate in the epididymis and kidneys [85], affecting fertility by inhibiting H+-ATPase, leading to alkalinization of the epididymal tubular lumen [86,87]. In addition, functional coordination between principal cells and clear cells in H+ secretion has been determined. HCO3 is secreted by the principal cells into the lumen of the epididymis, enters the clear cells under the regulation of NBCn1, which in turn triggers sAC activation and increased cAMP levels, and finally accelerates the rate of clear cell H+-ATPase-mediated H+ secretion [88]. Angiotensin II induces the production of nitric oxide (NO), which activates soluble guanylate cyclase (sGC) in the clear cells, elevating cGMP, which in turn leads to the aggregation of H+-ATPase in the microvilli of the clear cells. These results indicate that epididymal H+-ATPase is essential to the maintenance of the acidity of the epididymal lumen and plays a crucial role in sperm maturation.
H+ channels and HCO3 transporters located in epithelial cells of the epididymis, and the collaborative actions among various epithelial cell types, are essential for luminal pH regulation. Therefore, investigation of new transporters and H+ secretion channels and elucidation of their synergistic roles are indispensable for understanding the physiological mechanisms of sperm maturation.

4. pH and Regulation of Sperm Function

4.1. pH-Regulated Ion Channels in Spermatozoa

As mature spermatozoa are specialized cells that are transcriptional and translational silencing, the regulation of sperm function is dependent to a large extent on ion channels, second messengers, and protein modifications [89]. Among these, the regulation of intracellular and extracellular ion concentrations by ion channels is crucial for the maintenance of sperm motility and the regulation of sperm function. pH is an indispensable regulator in the activation of ion channels. A few ion channels are present in the sperm plasma membrane, including CatSper, SLO3, voltage-gated Ca2+ channels, and transient receptor potential cation channel subfamily V member 4, which respond to the surroundings in the female reproductive tract [90]. Alkalinization of sperm pHi activates the CatSper and SLO3 channels substantially [91,92]. CatSper is a tetrameric channel composed of four independent pore-forming α-subunits (CatSper1–4) and at least nine auxiliary subunits; it is activated by membrane depolarization and alkalinization of pHi, and mediates Ca2+ influx [93,94]. A comparative genomic assay found that nearly all animals possessing at least four CatSper pore-forming subunit genes also possessed both sAC and at least one sperm-specific Na+/H+ exchanger (SLC9C) gene [95], showing that CatSper activation is possibly associated with NHE activity. The external conditions appear markedly changed once spermatozoa enter the vagina. In particular, the increased pH in the fallopian tube fluid activates the CatSper channel, inducing hyperactivated motility. Male mice with knockout of any CatSper channel subunits are sterile [96,97,98]. The pH sensitivity of CatSper was initially attributed to a conserved histidine-rich region at the N-terminus of CatSper1 in mammals [91,99]. Further studies have revealed that the mechanism by which alkalinization activates CatSper is dependent on the testis-specific protein EF-hand calcium-binding domain protein 9 (EFCAB9), which directly interacts with CATSPERζ in a Ca2+-dependent manner and dissociates at elevated pH. EFCAB9 is significant for pH-dependent and Ca2+-sensitive CatSper channel activation, as CatSper is less sensitive to intracellular Ca2+ changes and pHi alkalinization without EFCAB9 [100]. In human spermatozoa, the H+ channel HVCN1 co-localizes with CatSper in the principal piece of the flagellum. H+ efflux occurs through the HVCN1 channel, which induces pH alkalinization in the flagellum and activates the CatSper ion channel (Figure 3) [101].
SLO3 channels are pHi- and voltage-sensitive K+, which primarily mediate outward K+ currents (IKSper) (Figure 3). IKSper is the only detectable hyperpolarizing current in spermatozoa and significantly regulates their resting membrane potential. Slo3 mutant male mice are sterile, and their sperm have a meager fertilization rate in in vitro fertilization [21]. The voltage sensitivity of heterologously expressed SLO3 channels differed from that of IKSper, implying additional regulatory elements in the spermatozoa [92,102]. The biophysical properties of SLO3 in mice were only similar to those of natural mouse IKSper when coexpressed with the auxiliary subunit LRRC52. Subsequently, LRRC52 was demonstrated to be the auxiliary subunit required to activate the SLO3 channel at physiological voltages and pH values [103,104]. Mice with Lrrc52 gene deficiency produce spermatozoa with severely reduced IKSper voltage sensitivity and have low male fertility. The standard fertilization potential of mouse spermatozoa depends on the co-expression of SLO3/LRRC52 [105]. Intracellular alkalinization significantly enhances IKSper, but this effect can be inhibited by the NHE inhibitor EIPA [92], indicating that IKSper is regulated by H+. In addition, when extracellular pH was at a level of 7.4 and there was no pH buffer in the pipette solution, dimethyl amiloride (DMA), a broad-spectrum NHE inhibitor, inhibited KSper current and membrane depolarization, and subsequently led to a decrease in CatSper current and Ca2+ influx [106], further demonstrating the pH-sensitive regulation of SLO3.
Possible links in functional coupling between pH-sensitive CatSper and SLO3 have been found; however, species differences have led to contradictory results among studies. Evidence from mouse spermatozoa suggests that SLO3 controls Ca2+ influx through CatSper [107]. High HCO3 concentration causes changes in pHi that activate SLO3 channels [108]. The membrane hyperpolarization further raises pHi via NHE or HCO3 transporters, and this pHi alkalinization activates CatSper channels, leading to a rapid increase in [Ca2+]i [107]. In human spermatozoa, KSper is regulated by Ca2+ but is not sensitive to pH [109]. SLO3 channels can inhibit Ca2+ influx via CatSper, proving it may be localized downstream of CatSper [110]; this suggestion is enhanced by the investigation in which some ICSI patients exhibited impaired K+ conductance and abnormal resting membrane potentials but normal resting [Ca2+]i and P4-induced [Ca2+]i responses that were similar to those of normal spermatozoa, suggesting that CatSper function was not altered [111]. Additionally, inhibitor cross-sensitization between CatSper and SLO3 has been revealed. Therefore, further studies are required to investigate the potential interactions between these two channels, with a particular emphasis on elucidating the role of pH in the functional coupling between them, and identifying the similarities and differences between mouse and human spermatozoa. In addition, the emphasis should be on whether other ion channels present in spermatozoa, such as Cl, Na+, and other Ca2+ channels, are modulated by pH and affect sperm function.

4.2. pH and Sperm Motility

The initiation of sperm motility in fish, birds, and mammals has shown that although pH affects sperm motility to slightly differing extents or in subtly different ways, it plays a crucial regulatory role in initiating or maintaining sperm motility, and alkalinization is generally considered to facilitate these processes [112]. After maturation, spermatozoa are stored in the epididymis in an immotile form. In addition to being constrained by high osmotic pressure and the need for various proteins in the spermatozoa to be functionally integrated and separated, the lower pH may block their motility. For example, the motor proteins that drive the movement of the sperm flagellum are pH-dependent, and a lower pH inhibits dynein ATPase activity [113]. This was also observed in an assay concentrating on sea urchin spermatozoa [20]. After ejaculation, sperm motility is initiated by combining factors such as hypotonicity, extracellular Na+, K+ levels, and pH [114,115].
The extracellular pH of the spermatozoa increases instantaneously to approximately 7.5 in seminal fluid. pH value increase contributes to sperm motility, and activation of sperm motility by alkalinization, and was even once believed to be conserved from corals to humans [116]. Alkalinization stimulates cAMP production by sAC, which leads to protein kinase A (PKA)-dependent phosphorylation of flagellin and opening of CatSper channels, promoting sperm motility [117,118]. The elevated pH is critical for activating sperm motility, and this process depends on extracellular Na+ mediated by the activation of the Na+/H+ exchanger in Pacific oysters [119]. However, intracellular alkalinization does not participate in sperm motility in marine fishes such as the European eel [120], and pHi remains unchanged in the sperm motility of rainbow trout [121], implying that species should be taken into consideration in uncovering the pH-mediated process.
In addition to direct regulation of sperm pHi by intracellular factors, the external pH can regulate sperm motility by affecting pHi. The motility was significantly higher in alkaline culture conditions (pH 7.2 and 8.2) compared to that in acidic conditions (pH 5.2 and 6.2) in human spermatozoa, which may be related to the reduced concentration of intracellular Ca2+ in the spermatozoa and the inhibition of Na+/K+-ATPase activity under acidic conditions [122]. Clinical investigations have also found that the pH of seminal plasma in some cases with oligozoospermia and asthenozoospermia is significantly lower than that of the general population (pH 7.2) and that pH is positively correlated with sperm motility [123]. Therefore, alkalinity is essential for sperm motility in vitro. Sperm motility in vivo is significantly activated when they arrive at the glucose-rich uterus from the lactate-rich vagina, demonstrating that glucose is a pH-dependent regulator of sperm motility and that the production of HCO3 from glycolysis is responsible for elevated pHi [124]. Sperm viability decreases during fluid preservation of livestock spermatozoa, accompanied by reduced pH. Artificial pH stabilization is essential for maintaining sperm viability during fluid conservation [125]. Cryopreserved spermatozoa are also associated with decreased pH after resuscitation, which is detrimental to their survival. Therefore, the development of non-toxic pH buffers or diluents that contribute to the recovery of spermatozoa after resuscitation is necessary. In addition, sperm frozen at a higher pH (8.0) have significantly enhanced viability after resuscitation [126], which can be applied to improve the assisted reproduction rate. These data suggest that acidification of the culture medium or seminal fluid after activation of sperm motility is detrimental to normal sperm motility and survival.
Taking advantage of the role of pH in regulating sperm motility, researchers have attempted to selectively screen or enrich sperm with different sex chromosomes (X or Y) through pH manipulation to generate sex-specific offspring in animal husbandry and veterinary medicine [127]. Studies employing human and bovine spermatozoa showed no difference in the response of X- and Y-chromosome spermatozoa to changes in pH. However, spermatozoa from dairy goats in diluents with different pH levels had different proportions of upper-layer X/Y-chromosome spermatozoa using a swim-up technique. At a sperm diluent pH of 6.2, the proportion of upper layer X-chromosome spermatozoa was 67.24% ± 2.61. At a sperm diluent pH of 7.4, the proportion of upper layer X-chromosome spermatozoa was 30.45% ± 1.03, which differed significantly from a proportion of 52.35% ± 1.72 in the control group (pH 6.8) [127]. A higher proportion of X-chromosome spermatozoa (85.57% ± 3.27) was observed in the alkaline diluent (pH 7.4) upon the co-administration of resiquimod [128]. pH affects mitochondrial activity and glucose uptake capacity primarily through phosphorylation of NFκB and GSK3α/β. The viability of X-chromosome spermatozoa is inhibited under acidic conditions and significantly enhanced under alkaline conditions [129]. The sex ratio of offspring can be influenced by adjusting the pH of the semen diluent in pigs, with female offspring more likely to be obtained from semen stored under acidic conditions for short periods (<1 d) [130,131].

4.3. pH Promotion in Sperm Capacitation and Hyperactivation

After entering the female reproductive tract, the progressively more alkaline pH and high HCO3 concentration induce capacitation of the sperm cells. During capacitation, physiological-biochemical variation is observed in spermatozoa, including pHi alkalinization, increased intracellular Ca2+ ([Ca2+]i) levels, plasma membrane hyperpolarization, and tyrosine phosphorylation. These processes are critical for sperm capacitation and hyperactivated motility, further facilitating AR and fertilization [132]. Hyperactivated motility accompanies the onset of capacitation and manifests as high-amplitude, asymmetric oscillation of the sperm flagellum, which enables the sperm’s arrival at the fallopian tube. Alkaline pH plays a double role in hyperactivation. First, intracellular alkalinization promotes sperm hyperactivated motility by activating CatSper channels [133]. In mice, pHi is elevated during sperm capacitation, which activates CatSper channels and triggers asymmetric bending of the flagellum to stimulate hyperactivated motility [134]. Treatment with the alkalizer NH4Cl increases pHi, stimulates elevated sperm [Ca2+]i, and induces hyperactivation [135]. Second, pH alkalinization directly induces hyperactivated motility. The spermatozoa without membrane could be reactivated in solutions containing high concentrations of ATP and 400–1000 nM Ca2+, leading to asymmetric oscillation of their flagella [136,137]. In addition, demembranated spermatozoa were activated at pH 7.0. However, hyperactivated motility occurred only in solutions at pH 7.9–8.5, suggesting that an increase in pH in the axonemal compartment of the spermatozoa is conducive to hyperactivated motility [137].
pHi alkalinization is required for sperm capacitation and largely depends on HCO3 influx or H+ efflux [47,138]. High HCO3 concentration in the female reproductive tract is crucial for the cAMP-PKA signaling pathway, pHi alkalinization, and further membrane hyperpolarization [22,139,140]. HCO3 regulates adenylate cyclase (sAC or ADCY10) activity in the spermatozoa and induces increased cAMP levels and activation of PKA. Male mice with knockout of both sAC genes are sterile, with spermatozoa lacking progressive motility and exhibiting inhibited pHi alkalinization, affecting both sperm capacitation and hyperactivated motility [118,141,142]. In addition, the sAC inhibitor TDI-10229 also significantly inhibits pHi alkalinization, increased cAMP content, and AR induced by sperm capacitation in humans and mice [143].

4.3.1. HCO3 Transport

In spermatozoa, HCO3 transport is mediated through the SLC26 family and CFTR (Figure 3). The central SLC26 family members localized in the male reproductive tract and spermatozoa include SLC26A3, SLC26A6, and SLC26A8 (TAT1) [36,144]. In mouse spermatozoa, SLC26A3 and SLC26A6 are expressed primarily in the midpiece. SLC26A3 is involved in increased intracellular Cl and membrane hyperpolarization induced by dibutyryl cyclic AMP (dbcAMP). Furthermore, HCO3 is transported through SLC26A3, which plays a crucial role in capacitation and induces tyrosine phosphorylation and hyperactivated motility in spermatozoa [33,145]. Mice with Slc26a3 knockout have altered epididymal morphology and reduced sperm counts, which resulted in male sterility [34], whereas Slc26a6 knockout mice remained fertile [146]. A missense mutation in Slc26a3 in human spermatozoa causes decreased male fertility [35]. SLC26A3 and SLC26A6 also interact with CFTR and function together in regulating sperm pHi [33], suggesting that SLC26A3 plays a vital role in male reproduction. SLC26A8 is predominantly localized in the annulus of mature human and mouse spermatids [36,37]. In male mice, deletion of the Slc26a8 gene results in structural anomalies such as defective flagellar differentiation, abnormal flagellar loops, and abnormal sperm capacitation, which ultimately causes male sterility [36]. Notably, SLC26A8 co-localizes with CFTR, another factor that regulates pH homeostasis by transporting HCO3, in the equatorial segment of the sperm head (Figure 3). Together, they form a complex that participates in sperm capacitation and hyperactivated motility. SLC26A8 stimulates CFTR channel activity. Loss of Slc26a8 in mouse spermatozoa leads to decreased intracellular cAMP, which abrogates activation of the sAC/PKA signaling pathway [147].
Mutations in Cftr cause cystic fibrosis, characterized by progressive lung disease, pancreatic insufficiency, and male infertility [49,50,148]. In mice, guinea pigs, and human spermatozoa, CFTR is localized in the equatorial region of the head and the midpiece of the sperm [47]. Homozygous Cftr knockout mice tend to die around puberty; therefore, heterozygous Cftr mutant mice (Cftr+/−) are commonly used to identify the role of CFTR in spermatozoa [48]. Cftr+/− spermatozoa exhibited abnormal capacitation, attenuated response to HCO3-induced membrane hyperpolarization, reduced cAMP production, and decreased motility compared to those of wild-type mice. Additionally, the CFTR inhibitor diphenylamine-2-carboxylic acid (DPC) can affect membrane hyperpolarization and inhibit Cl influx [47,149], whereas the agonist genistein induces membrane hyperpolarization and Cl influx in non-capacitated mouse spermatozoa [149]. Notably, CFTR is in the flagellar annulus of spermatozoa along with both sAC and SLC26A8, promoting local cAMP production; this may be necessary for flagellar PKA pathway activation in spermatozoa [147]. Besides, CFTR can modulate hyperpolarization associated with sperm capacitation in mice by inhibiting epithelial Na+ channels, which activate the cAMP/PKA signaling pathway and regulate membrane potential [150,151].

4.3.2. H+ Transport

H+ efflux is mediated primarily through H+ channels on the sperm plasma membrane, which induces pHi alkalinization and elevated cAMP levels (Figure 3), facilitating sperm capacitation. Sperm pHi alkalinization was abnormal and capacitation was inhibited after H+ efflux blocking [106,152,153]. In rodents, H+ efflux during capacitation is mainly dependent on NHEs. Inhibition of NHE channels significantly reduces sperm viability and progressive motility [106]. NHEs are a class of Na+/H+ exchangers involved in the electroneutral exchange of extracellular Na+ with intracellular H+ to keep pH homeostasis. The NHE family consists of 13 members organized into three distinct subfamilies: (1) SLC9A: SLC9A1(NHE1)–SLC9A9(NHE9); (2) SLC9B: SLC9B1(NHA1) and SLC9B2(NHA2); (3) SLC9C: SLC9C1 (sperm-specific NHE exchanger, sNHE) and SLC9C2 [154]. NHEs regulate the acid-base balance of a wide range of luminal fluids. Several NHEs have been detected in the male reproductive tract or spermatozoa, including NHE1, NHE2, NHE3, NHE5, NHE8, NHA1, NHA2, and sNHE [39].
The localization, molecular weight, and physiological actions of NHE1 vary among species. In rats, NHE1 is localized in the midpiece of the flagellum and cooperates with the a4 isoform of Na+/K+-ATPase to regulate pHi and maintain sperm motility [155]. In sheep and pig spermatozoa, NHE1 is localized in the equatorial segment of the head and flagellum and regulates sperm viability and pH [156]. The absence of NHE1 or NHE5 does not affect fertility in male mice [157,158].
NHE2 is localized in the caput, corpus, and cauda of the epididymis, and in the vas deferens and testes [159,160]; however, limited information is available regarding its function. NHE3 is expressed in the efferent ducts and proximal epididymis in rats and is localized to the efferent ducts in humans and mice, where it is involved in the reabsorption of Na+ and in regulating luminal fluid pH [38,39].
Knockdown of Nhe3 results in abnormal dilation of the rete testis lumen and efferent ducts, leading to obstructive azoospermia [39]. In spermatogenic cells, NHE3 is localized explicitly to developing acrosomal granules, and its absence results in severe defects in the acrosome, indicating that NHE3 is essential for normal acrosome development [40].
Nhe8 knockout mice exhibit impaired spermatogenesis, impaired testicular Leydig cell function, round-headed spermatozoa, lack of acrosomes, and abnormal distribution of mitochondrial sheaths [61]. Mice with SC-specific knockdown of NHE8 produce spermatozoa with normal morphology and exhibit no adverse effects on fertility [161]. In addition, NHE8 is not expressed in Leydig cells [161]. Evidence showed that NHE8 is essential for male fertility; however, its expression and physiological functions in the SCs and Leydig cells, and its working in acrosome formation, remain to be elucidated.
NHA1 and NHA2 are specifically localized to the principal piece of the mouse sperm flagellum (Figure 3) [42]. Nha1 and Nha2 cKO mice exhibit decreased sperm counts and reduced fertility. In addition, the spermatozoa of Nha1 cKO mice have markedly reduced levels of cAMP synthesized by sAC [42]. Nha1/2 double knockout mice showed male sterility [42]. In humans, NHA1 deficiency has been reported in men with teratozoospermia [43].
sNHE is localized in the principal piece of the sperm flagellum [44] and is sensitive to voltage-gating and cAMP (Figure 3) [45]. In mice, sNHE is coupled to the SLC26A3-CFTR complex, facilitating high levels of HCO3 and Cl, thus ensuring elevated cAMP and pHi alkalization [33]. Deficiency or loss of function of sNhe leads to infertility in mice, primarily due to reduced HCO3-sensitive sAC activity, resulting in lower cAMP content and impaired sperm hyperactivation [44,162]. Decreased sNHE was also detected in patients with asthenozoospermia, and the level of sNHE was positively correlated with sperm concentration, total number, and progressive motility [46]. sNHE regulates the pHi of the sperm flagellum, and is involved in sperm motility, functionally coupling with CatSper and Ksper [153]. NHE11 is localized in the acrosomal region of the mature mouse and human sperm [163]; however, its physiological role has not been elucidated.
The HVCN1 is mainly localized in the principal piece of the flagellum as a critical player in the control of pHi of spermatozoa. In human, bovine, and porcine spermatozoa, HVCN1 also mediates H+ efflux besides NHEs (Figure 3) [101]. The activation of HVCN1 depends on membrane potential and the difference between intracellular and extracellular pH. Additionally, fatty acids regulate the HVCN1 gating via the protein kinase C (PKC)-dependent phosphorylation pathway [164,165,166,167]. During sperm capacitation in human or bovine models, HVCN1 induces rapid alkalinization of the principal piece of the flagellum, activating CatSper channels and pH-dependent proteins that contribute to the transition of spermatozoa into motility patterns [168,169]. HVCN1 was found to regulate pHi alkalinization and the hyperactivation of motility in the principal piece of the flagellum in human spermatozoa after inhibition by Cl-GBI [152]. In porcine sperm, HVCN1 is essential for the hyperactivation of motility during capacitation in vitro, but does not influence the luteinizing hormone-induced AR [138]. In addition, eliminating cholesterol and high levels of albumin in the uterine fluid can also enhance the activation of HVCN1 channels [170,171]. Notably, low levels of albumin in the semen (approximately 15 µM) cannot activate HVCN1 channels. In comparison, albumin (500 µM) in the female reproductive tract causes HVCN1 channels to open and induces sperm capacitation, enabling fertilization [170,172]. In contrast, high levels of Zn2+ in human seminal fluid inactivate HVCN1 and reduce sperm viability [173].

4.4. Acrosome Reaction

In mammals, the acrosome is a secretory vesicle located in the apical region of the sperm head, and the acrosome releases its contents by exocytosis upon physiological stimulation, a process known as the AR [174]. The acrosome contains a mixture of hydrolytic enzymes promoting passage of the sperm through the oocyte for sperm-egg fusion [175]. Mouse sperm acrosomal pH (pHa) is maintained at 5.3 but increases to approximately 6.2 during capacitation [176]. Alkalinization during acrosome capacitation activates various enzymes within the acrosome and stimulates spontaneous AR [175,176,177]. Inhibition of lactate metabolism results in an inability to alkalinize pHi, which further affects normal capacitation and the occurrence of AR [178]. In addition, elevated pHa promotes the breakdown of amyloid in the acrosomal matrix amyloid and facilitates the exocytosis of the acrosomal contents [179]. The percentage of AR induction was highest when spermatozoa were at a pH of 7.4, whereas an acidic extracellular condition (pH 6.5) inhibited AR [180,181]. In human spermatozoa, pHa gradually increases when incubated in capacitation fluid, depending on the presence of HCO3 and Ca2+ in the medium, and alkalinization of the acrosome is mediated by a combination of H+-ATPase, HCO3 transporter, and sAC (Figure 3) [182]. A weakly alkaline Ca2+ blocker induces acrosomal alkalinization, triggering AR in mouse and human spermatozoa, even in a calcium-free medium. NH4Cl, a well-known alkalizer, failed to induce AR, suggesting that Ca2+ release from the acrosomes is a critical process in the induction of AR. However, alkalinization alone cannot induce AR [183,184]. Inhibition of H+-ATPase leads to acrosome alkalinization in mice and humans, possibly by maintaining and generating an H+ gradient in the acrosome membrane, which regulates pHa [54]. However, alkalinization caused by H+-ATPase inhibition cannot induce AR, indicating that acrosomal alkalinization and AR may be relatively independent processes or that alkalinization is unnecessary for AR [55]. This may differ from AR in sea urchin spermatozoa, where alkalinization is believed to be required for AR [20].
The subject of how pH regulates AR is largely unknown, and it is currently thought to involve alterations in enzyme activity and the activations of various ion channels associated with H+ and HCO3 transport. SLC4A1, which is expressed in the head and flagellum of human spermatozoa, is essential for the initiation of AR and may be related to the regulation of phosphorylation of two specific protein kinases in the sperm head, namely the tyrosine kinases Syk (which phosphorylates Tyr8, Tyr21, and Tyr904) and Lyn (which phosphorylates Tyr359) [185].
SLC26A8 acts as an anion transporter and may directly regulate AR via pH-dependent activation [36] or functional coupling with CFTR [186]. Inhibition of CFTR with the channel blocker inh-172 also markedly inhibits mouse spermatozoa from AR and affects HCO3-induced alkalinization and membrane hyperpolarization, suppressing the HCO3-dependent increase in cAMP [47]. These results imply that CFTR promotes AR depending on pH variation. In addition, NHE1, another necessary channel regulating pH, was found to be localized in the equatorial segment of the sperm head and the flagellum in porcine spermatozoa [187]. NHE1 is essential for progesterone-induced hyperactivation and AR [187]. Interestingly, reduced sNHE was found in patients with necrospermia, and its expression was also determined to be positively correlated with sperm concentration but not with the rate of AR [46]. However, our data confirmed that sNHE affects progesterone-induced AR by modulating the pH of the sperm flagellum instead of that in the acrosome [153]. HVCN1 is also a considerable channel in AR activation by coupling with cAMP, PKC, or CatSper in bull spermatozoa [168]. Various ion channels are implemented in AR activation, while the associated mechanisms and the possibility of synergistic effects require further determination.

5. Conclusions and Perspectives

Various membrane proteins have been identified as working synergistically to sustain proper intracellular and extracellular pH by transporting HCO3, H+, Na+, etc., contributing to spermatogenesis and sperm maturation. In the female reproductive tract, HCO3 transporters and H+ channels alkalinize sperm pHi, inducing sperm capacitation, hyperactivated motility and AR. The significance of pH homeostasis has been explored in laboratories by effective methods such as the patch-clamp technique, voltage-sensitive and ion-sensitive dyes, bilayer reconstitution, recombinant DNA techniques, single-stranded probe (cRNA) expression in heterologous systems, immunocytochemistry, etc. Further investigations are warranted to uncover the mechanisms of the transporters and ion channels regulating the pH of the luminal fluid of the reproductive tract as well as the pHi of spermatozoa. Our recommendations are as follows: 1. It remains unclear whether SLC4A1 and SLC4A3 (Table 2) are involved in pH regulation during spermatogenesis and sperm maturation. 2. Localized pH-regulating channels other than CatSper and SLO3 need further clarification. 3. Whether pH participates in regulating thermotropism and chemotropism before fertilization needs to be illuminated. 4. NBCe2, NHE2, NHE11, CA II, CA IX, MCT8, etc. (Table 2), are localized in the male reproductive system, and their involvement in male fertility needs to be identified.
Several sperm pH regulators have been shown to be implicated in male fertility during clinical investigations through multiple complicated pathophysiologic processes, including SLC26A3, SLC26A8, NHA1, sNHE, and CFTR. Damaged synergistic action of SLC26A3 and CFTR in HCO3 transport appeared in infertile men [35]. SLC26A8 is localized in the human sperm annulus, the aberrant expression of which caused structural defect of the annulus, further leading to asthenospermia [37]; however, the role of the annulus in sperm motility is far from clear, and the H+/HCO3 transportation activity of SLC26A8 provided a novel clue in the functional exploration of the sperm annulus. Moreover, determining the influence of other SLC26 family members on spermatozoa and their cooperation mechanism with CFTR would contribute to revealing the pathogenesis of male infertility. The expression of NHA1 occurs in a DNA methylation-dependent and independent concerted manner mediated by diverse DNA regulatory elements, and the decrease or absence of NHA1 is relevant to teratozoospermia [43]. Additional studies should focus on the methylation and demethylation patterns in NHA1 CpG islands from teratozoospermic cases, to determine the action of DNA methylation in NHA1 transcription regulation, and its contribution to male reproduction [43]. sNHE deficiency is also responsible for the pathogenicity of asthenospermia, probably due to pHi alkalization disorders [46]. Therefore, further clarifying the mechanisms of pH-regulating transporters involved in male fertility is crucial for targeted treatment in clinical practice. In addition, further screening infertile sperm samples in the clinic to focus on reproductive defects caused by disorders of acid-base balance, and identifying novel HCO3 and H+ transporters as well as their specific functions, will contribute to enhancing our understanding of male infertility, and may lead in turn to the development of novel therapeutic strategies for patients with oligospermia, necrospermia, and idiopathic infertility.

Author Contributions

M.L. and X.Z. (Xiaoning Zhang) designed and conceived the review. P.D. wrote the manuscript. M.Z. and Z.C. performed schematic diagram drawing. X.Z. (Xuhui Zeng) revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Social Development Project of Jiangsu Province (BE2022765), Nantong Social and People’s Livelihood Science and Technology Plan (MS22022087), and Basic Science Research Program of Nantong (JC22022086), and Jiangsu Innovation and Entrepreneurship Talent Plan (JSSCRC2021543).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Macaluso, M.; Wright-Schnapp, T.J.; Chandra, A.; Johnson, R.; Satterwhite, C.L.; Pulver, A.; Berman, S.M.; Wang, R.Y.; Farr, S.L.; Pollack, L.A. A public health focus on infertility prevention, detection, and management. Fertil. Steril. 2010, 93, 16.e1–16.e11. [Google Scholar] [CrossRef] [PubMed]
  2. Datta, J.; Palmer, M.J.; Tanton, C.; Gibson, L.J.; Jones, K.G.; Macdowall, W.; Glasier, A.; Sonnenberg, P.; Field, N.; Mercer, C.H.; et al. Prevalence of infertility and help seeking among 15,000 women and men. Hum. Reprod. 2016, 31, 2108–2118. [Google Scholar] [CrossRef] [PubMed]
  3. Agarwal, A.; Baskaran, S.; Parekh, N.; Cho, C.-L.; Henkel, R.; Vij, S.; Arafa, M.; Panner Selvam, M.K.; Shah, R. Male infertility. Lancet 2021, 397, 319–333. [Google Scholar] [CrossRef] [PubMed]
  4. Cannarella, R.; Condorelli, R.A.; Mongioì, L.M.; La Vignera, S.; Calogero, A.E. Molecular Biology of Spermatogenesis: Novel Targets of Apparently Idiopathic Male Infertility. Int. J. Mol. Sci. 2020, 21, 1728. [Google Scholar] [CrossRef] [PubMed]
  5. Lotti, F.; Maggi, M. Sexual dysfunction and male infertility. Nat. Rev. Urol. 2018, 15, 287–307. [Google Scholar] [CrossRef] [PubMed]
  6. Krausz, C.; Riera-Escamilla, A. Genetics of male infertility. Nat. Rev. Urol. 2018, 15, 369–384. [Google Scholar] [CrossRef] [PubMed]
  7. 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]
  8. Brugh, V.M.; Lipshultz, L.I. Male factor infertility: Evaluation and management. Med. Clin. N. Am. 2004, 88, 367–385. [Google Scholar] [CrossRef] [PubMed]
  9. Lutz, W. Fertility rates and future population trends: Will Europe’s birth rate recover or continue to decline? Int. J. Androl. 2006, 29, 25–33. [Google Scholar] [CrossRef] [PubMed]
  10. Griswold, M.D. Spermatogenesis: The Commitment to Meiosis. Physiol. Rev. 2016, 96, 1–17. [Google Scholar] [CrossRef] [PubMed]
  11. Bergmann, M. Spermatogenesis-physiology and pathophysiology. Der Urologe. Ausg. A 2005, 44, 1131–1138. [Google Scholar] [CrossRef] [PubMed]
  12. Rato, L.; Socorro, S.; Cavaco, J.E.B.; Oliveira, P.F. Tubular fluid secretion in the seminiferous epithelium: Ion transporters and aquaporins in Sertoli cells. J. Membr. Biol. 2010, 236, 215–224. [Google Scholar] [CrossRef] [PubMed]
  13. Hamamah, S.; Gatti, J.L. Role of the ionic environment and internal pH on sperm activity. Hum. Reprod. 1998, 13 (Suppl. 4), 20–30. [Google Scholar] [CrossRef] [PubMed]
  14. Ko, Y.-J.; Maeng, J.-H.; Hwang, S.Y.; Ahn, Y. Design, fabrication, and testing of a microfluidic device for thermotaxis and chemotaxis assays of sperm. Slas Technol. Transl. Life Sci. Innov. 2018, 23, 507–515. [Google Scholar] [CrossRef] [PubMed]
  15. Ng, K.Y.B.; Mingels, R.; Morgan, H.; Macklon, N.; Cheong, Y. In vivo oxygen, temperature and pH dynamics in the female reproductive tract and their importance in human conception: A systematic review. Hum. Reprod. Update 2018, 24, 15–34. [Google Scholar] [CrossRef] [PubMed]
  16. Fox, C.A.; Meldrum, S.J.; Watson, B.W. Continuous measurement by radio-telemetry of vaginal pH during human coitus. J. Reprod. Fertil. 1973, 33, 69–75. [Google Scholar] [CrossRef] [PubMed]
  17. Carr, D.W.; Acott, T.S. Intracellular pH regulates bovine sperm motility and protein phosphorylation. Biol. Reprod. 1989, 41, 907–920. [Google Scholar] [CrossRef] [PubMed]
  18. Suarez, S.S. Control of hyperactivation in sperm. Hum. Reprod. Update 2008, 14, 647–657. [Google Scholar] [CrossRef] [PubMed]
  19. 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]
  20. Nishigaki, T.; José, O.; González-Cota, A.L.; Romero, F.; Treviño, C.L.; Darszon, A. Intracellular pH in sperm physiology. Biochem. Biophys. Res. Commun. 2014, 450, 1149–1158. [Google Scholar] [CrossRef] [PubMed]
  21. 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] [PubMed]
  22. Visconti, P.E.; Krapf, D.; de la Vega-Beltrán, J.L.; Acevedo, J.J.; Darszon, A. Ion channels, phosphorylation and mammalian sperm capacitation. Asian J Androl. 2011, 13, 395–405. [Google Scholar] [CrossRef] [PubMed]
  23. Joseph, A.; Hess, R.A.; Schaeffer, D.J.; Ko, C.; Hudgin-Spivey, S.; Chambon, P.; Shur, B.D. Absence of Estrogen Receptor Alpha Leads to Physiological Alterations in the Mouse Epididymis and Consequent Defects in Sperm Function. Biol. Reprod. 2010, 82, 948–957. [Google Scholar] [CrossRef] [PubMed]
  24. Boussouar, F.; Mauduit, C.; Tabone, E.; Pellerin, L.; Magistretti, P.J.; Benahmed, M. Developmental and hormonal regulation of the monocarboxylate transporter 2 (MCT2) expression in the mouse germ cells. Biol. Reprod. 2003, 69, 1069–1078. [Google Scholar] [CrossRef] [PubMed]
  25. Bernardino, R.L.; Martins, A.D.; Jesus, T.T.; Sá, R.; Sousa, M.; Alves, M.G.; Oliveira, P.F. Estrogenic regulation of bicarbonate transporters from SLC4 family in rat Sertoli cells. Mol. Cell. Biochem. 2015, 408, 47–54. [Google Scholar] [CrossRef] [PubMed]
  26. O’Donnell, L.; Smith, L.B.; Rebourcet, D. Sertoli cells as key drivers of testis function. In Seminars in Cell & Developmental Biology; Academic Press: Cambridge, MA, USA, 2022; pp. 2–9. [Google Scholar]
  27. Oliveira, P.F.; Alves, M.G.; Rato, L.; Laurentino, S.; Silva, J.; Sá, R.; Barros, A.; Sousa, M.; Carvalho, R.A.; Cavaco, J.E.; et al. Effect of insulin deprivation on metabolism and metabolism-associated gene transcript levels of in vitro cultured human Sertoli cells. Biochim. Biophys. Acta 2012, 1820, 84–89. [Google Scholar] [CrossRef]
  28. Holappa, K. Primary structure of a sperm cell anion exchanger and its messenger ribonucleic acid expression during spermatogenesis. Biol. Reprod. 1999, 61, 981–986. [Google Scholar] [CrossRef] [PubMed]
  29. Medina, J.F.; Recalde, S.; Prieto, J.; Lecanda, J.; Saez, E.; Funk, C.D.; Vecino, P.; van Roon, M.A.; Ottenhoff, R.; Bosma, P.J.; et al. Anion exchanger 2 is essential for spermiogenesis in mice. Proc. Natl. Acad. Sci. USA 2003, 100, 15847–15852. [Google Scholar] [CrossRef] [PubMed]
  30. Nakai, M.; Chen, L.; Nowak, R.A. Tissue distribution of basigin and monocarboxylate transporter 1 in the adult male mouse: A study using the wild-type and basigin gene knockout mice. Anat. Record. Part A Discov. Mol. Cell. Evol. Biol. 2006, 288, 527–535. [Google Scholar] [CrossRef] [PubMed]
  31. Kishimoto, A.; Ishiguro-Oonuma, T.; Takahashi, R.; Maekawa, M.; Toshimori, K.; Watanabe, M.; Iwanaga, T. Immunohistochemical localization of GLUT3, MCT1, and MCT2 in the testes of mice and rats: The use of different energy sources in spermatogenesis. Biomed. Res. 2015, 36, 225–234. [Google Scholar] [CrossRef] [PubMed]
  32. Bernardino, R.L.; D’Souza, W.N.; Rato, L.; Rothstein, J.L.; Dias, T.R.; Chui, D.; Wannberg, S.; Alves, M.G.; Oliveira, P.F. Knockout of MCT1 results in total absence of spermatozoa, sex hormones dysregulation, and morphological alterations in the testicular tissue. Cell Tissue Res. 2019, 378, 333–339. [Google Scholar] [CrossRef] [PubMed]
  33. Chávez, J.C.; Hernández-González, E.O.; Wertheimer, E.; Visconti, P.E.; Darszon, 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] [PubMed]
  34. El Khouri, E.; Whitfield, M.; Stouvenel, L.; Kini, A.; Riederer, B.; Lores, P.; Roemermann, D.; di Stefano, G.; Drevet, J.R.; Saez, F.; et al. Slc26a3 deficiency is associated with epididymis dysplasia and impaired sperm fertilization potential in the mouse. Mol. Reprod. Dev. 2018, 85, 682–695. [Google Scholar] [CrossRef] [PubMed]
  35. 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] [PubMed]
  36. Touré, A.; Lhuillier, P.; Gossen, J.A.; Kuil, C.W.; Lhôte, D.; Jégou, B.; Escalier, D.; Gacon, G. The testis anion transporter 1 (Slc26a8) is required for sperm terminal differentiation and male fertility in the mouse. Hum. Mol. Genet. 2007, 16, 1783–1793. [Google Scholar] [CrossRef] [PubMed]
  37. Lhuillier, P.; Rode, B.; Escalier, D.; Lorès, P.; Dirami, T.; Bienvenu, T.; Gacon, G.; Dulioust, E.; Touré, A. Absence of annulus in human asthenozoospermia: Case report. Hum. Reprod. 2009, 24, 1296–1303. [Google Scholar] [CrossRef] [PubMed]
  38. Leung, G.P.; Tse, C.M.; Chew, S.B.; Wong, P.Y. Expression of multiple Na+/H+ exchanger isoforms in cultured epithelial cells from rat efferent duct and cauda epididymidis. Biol. Reprod. 2001, 64, 482–490. [Google Scholar] [CrossRef] [PubMed]
  39. Gardner, C.C.; James, P.F. Na+/H+ Exchangers (NHEs) in Mammalian Sperm: Essential Contributors to Male Fertility. Int. J. Mol. Sci. 2023, 24, 14981. [Google Scholar] [CrossRef] [PubMed]
  40. Zhou, Q.; Clarke, L.; Nie, R.; Carnes, K.; Lai, L.-W.; Lien, Y.-H.H.; Verkman, A.; Lubahn, D.; Fisher, J.S.; Katzenellenbogen, B.S. Estrogen action and male fertility: Roles of the sodium/hydrogen exchanger-3 and fluid reabsorption in reproductive tract function. Proc. Natl. Acad. Sci. USA 2001, 98, 14132–14137. [Google Scholar] [CrossRef] [PubMed]
  41. Bagnis, C.; Marsolais, M.; Biemesderfer, D.; Laprade, R.; Breton, S. Na+/H+-exchange activity and immunolocalization of NHE3 in rat epididymis. Am. J. Physiology. Ren. Physiol. 2001, 280, F426–F436. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, S.-R.; Chen, M.; Deng, S.L.; Hao, X.X.; Wang, X.X.; Liu, Y.X. Sodium-hydrogen exchanger NHA1 and NHA2 control sperm motility and male fertility. Cell Death Dis. 2016, 7, e2152. [Google Scholar] [CrossRef]
  43. Kumar, P.L.; James, P.F. Identification and characterization of methylation-dependent/independent DNA regulatory elements in the human Slc9b1 gene. Gene 2015, 561, 235–248. [Google Scholar] [CrossRef]
  44. Wang, D.; King, S.M.; Quill, T.A.; Doolittle, L.K.; Garbers, D.L. A new sperm-specific Na+/H+ exchanger required for sperm motility and fertility. Nat. Cell Biol. 2003, 5, 1117–1122. [Google Scholar] [CrossRef]
  45. Windler, F.; Bönigk, W.; Körschen, H.G.; Grahn, E.; Strünker, T.; Seifert, R.; Kaupp, U.B. The solute carrier SLC9C1 is a Na+/H+-exchanger gated by an S4-type voltage-sensor and cyclic-nucleotide binding. Nat. Commun. 2018, 9, 2809. [Google Scholar] [CrossRef]
  46. 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]
  47. 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]
  48. Grubb, B.R.; Gabriel, S.E. Intestinal physiology and pathology in gene-targeted mouse models of cystic fibrosis. Am. J. Physiol. 1997, 273, G258–G266. [Google Scholar] [CrossRef]
  49. Ahmed, N.; Corey, M.; Forstner, G.; Zielenski, J.; Tsui, L.C.; Ellis, L.; Tullis, E.; Durie, P. Molecular consequences of cystic fibrosis transmembrane regulator (Cftr) gene mutations in the exocrine pancreas. Gut 2003, 52, 1159–1164. [Google Scholar] [CrossRef]
  50. Grasemann, H.; Ratjen, F. Cystic Fibrosis. N. Engl. J. Med. 2023, 389, 1693–1707. [Google Scholar] [CrossRef]
  51. Brown, D.; Lui, B.; Gluck, S.; Sabolić, I. A plasma membrane proton ATPase in specialized cells of rat epididymis. Am. J. Physiol. 1992, 263, C913–C916. [Google Scholar] [CrossRef]
  52. Breton, S.; Smith, P.J.; Lui, B.; Brown, D. Acidification of the male reproductive tract by a proton pumping (H+)-ATPase. Nat. Med. 1996, 2, 470–472. [Google Scholar] [CrossRef]
  53. Smith, A.N.; Finberg, K.E.; Wagner, C.A.; Lifton, R.P.; Devonald, M.A.; Su, Y.; Karet, F.E. Molecular cloning and characterization of Atp6n1b: A novel fourth murine vacuolar H+-ATPase a-subunit gene. J. Biol. Chem. 2001, 276, 42382–42388. [Google Scholar] [CrossRef]
  54. Sun-Wada, G.-H.; Imai-Senga, Y.; Yamamoto, A.; Murata, Y.; Hirata, T.; Wada, Y.; Futai, M. A proton pump ATPase with testis-specific E1-subunit isoform required for acrosome acidification. J. Biol. Chem. 2002, 277, 18098–18105. [Google Scholar] [CrossRef]
  55. Codelia, V.A.; Cortes, C.J.; Moreno, R.D. Inhibition of the vacuolar H(+)-pump with bafilomycin A1 does not induce acrosome reaction or activate proacrosin in mouse spermatozoa. Biochem. Biophys. Res. Commun. 2005, 337, 1337–1344. [Google Scholar] [CrossRef]
  56. Blomqvist, S.R.; Vidarsson, H.; Söder, O.; Enerbäck, S. Epididymal expression of the forkhead transcription factor Foxi1 is required for male fertility. EMBO J. 2006, 25, 4131–4141. [Google Scholar] [CrossRef]
  57. Dai, P.; Qiao, F.; Chen, Y.; Chan, D.; Yim, H.; Fok, K.; Chen, H. SARS-CoV-2 and male infertility: From short-to long-term impacts. J. Endocrinol. Investig. 2023, 46, 1491–1507. [Google Scholar] [CrossRef]
  58. Bernardino, R.L.; Martins, A.D.; Socorro, S.; Alves, M.G.; Oliveira, P.F. Effect of prediabetes on membrane bicarbonate transporters in testis and epididymis. J. Membr. Biol. 2013, 246, 877–883. [Google Scholar] [CrossRef]
  59. Romero, M.F.; Chen, A.-P.; Parker, M.D.; Boron, W.F. The SLC4 family of bicarbonate (HCO3) transporters. Mol. Asp. Med. 2013, 34, 159–182. [Google Scholar] [CrossRef]
  60. Oliveira, P.F.; Sousa, M.; Barros, A.; Moura, T.; Rebelo da Costa, A. Membrane transporters and cytoplasmatic pH regulation on bovine Sertoli cells. J. Membr. Biol. 2009, 227, 49–55. [Google Scholar] [CrossRef]
  61. Xu, H.; Chen, H.; Li, J.; Zhao, Y.; Ghishan, F.K. Disruption of NHE8 expression impairs Leydig cell function in the testes. Am. J. Physiol. Cell Physiol. 2015, 308, C330–C338. [Google Scholar] [CrossRef]
  62. Bernardino, R.L.; Costa, A.R.; Martins, A.D.; Silva, J.; Barros, A.; Sousa, M.; Sá, R.; Alves, M.G.; Oliveira, P.F. Estradiol modulates Na+-dependent HCO3 transporters altering intracellular pH and ion transport in human Sertoli cells: A role on male fertility? Biol. Cell 2016, 108, 179–188. [Google Scholar] [CrossRef]
  63. Uhlen, M.; Oksvold, P.; Fagerberg, L.; Lundberg, E.; Jonasson, K.; Forsberg, M.; Zwahlen, M.; Kampf, C.; Wester, K.; Hober, S.J.N.b. Towards a knowledge-based human protein atlas. Nat. Biotechnol. 2010, 28, 1248–1250. [Google Scholar] [CrossRef]
  64. Felmlee, M.A.; Jones, R.S.; Rodriguez-Cruz, V.; Follman, K.E.; Morris, M.E. Monocarboxylate transporters (SLC16): Function, regulation, and role in health and disease. Pharmacol. Rev. 2020, 72, 466–485. [Google Scholar] [CrossRef]
  65. Mannowetz, N.; Wandernoth, P.; Wennemuth, G. Basigin interacts with both MCT1 and MCT2 in murine spermatozoa. J. Cell. Physiol. 2012, 227, 2154–2162. [Google Scholar] [CrossRef]
  66. Rato, L.; Alves, M.G.; Socorro, S.; Duarte, A.I.; Cavaco, J.E.; Oliveira, P.F. Metabolic regulation is important for spermatogenesis. Nat. Rev. Urol. 2012, 9, 330–338. [Google Scholar] [CrossRef]
  67. Halestrap, A.P.; PRICE, N.T. The proton-linked monocarboxylate transporter (MCT) family: Structure, function and regulation. Biochem. J. 1999, 343, 281–299. [Google Scholar] [CrossRef]
  68. Soudmand, P.; Tofighi, A.; Azar, J.T.; Razi, M.; Pakdel, F.G. Different continuous exercise training intensities induced effect on sertoli-germ cells metabolic interaction; implication on GLUT-1, GLUT-3 and MCT-4 transporting proteins expression level. Gene 2021, 783, 145553. [Google Scholar] [CrossRef]
  69. Rato, L.; Alves, M.G.; Socorro, S.; Carvalho, R.A.; Cavaco, J.E.; Oliveira, P.F. Metabolic modulation induced by oestradiol and DHT in immature rat Sertoli cells cultured in vitro. Biosci. Rep. 2012, 32, 61–69. [Google Scholar] [CrossRef]
  70. Pereira-Vieira, J.; Azevedo-Silva, J.; Preto, A.; Casal, M.; Queirós, O. MCT1, MCT4 and CD147 expression and 3-bromopyruvate toxicity in colorectal cancer cells are modulated by the extracellular conditions. Biol. Chem. 2019, 400, 787–799. [Google Scholar] [CrossRef]
  71. Okamura, N.; Tajima, Y.; Ishikawa, H.; Yoshii, S.; Koiso, K.; Sugita, Y. Lowered levels of bicarbonate in seminal plasma cause the poor sperm motility in human infertile patients. Fertil. Steril. 1986, 45, 265–272. [Google Scholar] [CrossRef]
  72. Acott, T.S.; Carr, D.W. Inhibition of bovine spermatozoa by caudal epididymal fluid: II. Interaction of pH and a quiescence factor. Biol. Reprod. 1984, 30, 926–935. [Google Scholar] [CrossRef] [PubMed]
  73. Jensen, L.J.; Stuart-Tilley, A.K.; Peters, L.L.; Lux, S.E.; Alper, S.L.; Breton, S. Immunolocalization of AE2 anion exchanger in rat and mouse epididymis. Biol. Reprod. 1999, 61, 973–980. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, Y.; Xu, J.-Y.; Wang, D.-K.; Wang, L.; Chen, L.-M. Cloning and identification of two novel Nbce1 splice variants from mouse reproductive tract tissues: A comparative study of NCBT genes. Genomics 2011, 98, 112–119. [Google Scholar] [CrossRef]
  75. PUSHKIN, A.; CLARK, I.; KWON, T.H.; Nielsen, S.; KURTZ, I. Immunolocalization of NBC3 and NHE3 in the rat epididymis: Colocalization of NBC3 and the vacuolar H+-ATPase. J. Androl. 2000, 21, 708–720. [Google Scholar] [CrossRef] [PubMed]
  76. Ekstedt, E.; Ridderstråle, Y. Histochemical localization of carbonic anhydrase in the testis and epididymis of the rabbit. Acta Anat. 1992, 143, 258–264. [Google Scholar] [CrossRef] [PubMed]
  77. Jensen, L.J.; Schmitt, B.M.; Berger, U.V.; Nsumu, N.N.; Boron, W.F.; Hediger, M.A.; Brown, D.; Breton, S. Localization of sodium bicarbonate cotransporter (NBC) protein and messenger ribonucleic acid in rat epididymis. Biol. Reprod. 1999, 60, 573–579. [Google Scholar] [CrossRef] [PubMed]
  78. Supuran, C.T. Carbonic anhydrases and metabolism. Metabolites 2018, 8, 25. [Google Scholar] [CrossRef] [PubMed]
  79. Caflisch, C.R.; DuBose Jr, T. Direct evaluation of acidification by rat testis and epididymis: Role of carbonic anhydrase. Am. J. Physiol.-Endocrinol. Metab. 1990, 258, E143–E150. [Google Scholar] [CrossRef]
  80. Hermo, L.; Chong, D.L.; Moffatt, P.; Sly, W.S.; Waheed, A.; Smith, C.E. Region-and cell-specific differences in the distribution of carbonic anhydrases II, III, XII, and XIV in the adult rat epididymis. J. Histochem. Cytochem. 2005, 53, 699–713. [Google Scholar] [CrossRef] [PubMed]
  81. Kaunisto, K.; Parkkila, S.; Parkkila, A.-K.; Waheed, A.; Sly, W.S.; Rajaniemi, H. Expression of carbonic anhydrase isoenzymes IV and II in rat epididymal duct. Biol. Reprod. 1995, 52, 1350–1357. [Google Scholar] [CrossRef] [PubMed]
  82. Karhumaa, P.; Kaunisto, K.; Parkkila, S.; Waheed, A.; Pastoreková, S.; Pastorek, J.; Sly, W.S.; Rajaniemi, H. Expression of the transmembrane carbonic anhydrases, CA IX and CA XII, in the human male excurrent ducts. Mol. Hum. Reprod. 2001, 7, 611–616. [Google Scholar] [CrossRef] [PubMed]
  83. Levine, N.; Marsh, D.J. Micropuncture studies of the electrochemical aspects of fluid and electrolyte transport in individual seminiferous tubules, the epididymis and the vas deferens in rats. J. Physiol. 1971, 213, 557–570. [Google Scholar] [CrossRef] [PubMed]
  84. Finberg, K.E.; Wagner, C.A.; Bailey, M.A.; Paunescu, T.G.; Breton, S.; Brown, D.; Giebisch, G.; Geibel, J.P.; Lifton, R.P. The B1-subunit of the H+ATPase is required for maximal urinary acidification. Proc. Natl. Acad. Sci. USA 2005, 102, 13616–13621. [Google Scholar] [CrossRef] [PubMed]
  85. Oldereid, N.B.; Thomassen, Y.; Attramadal, A.; Olaisen, B.; Purvis, K. Concentrations of lead, cadmium and zinc in the tissues of reproductive organs of men. J. Reprod. Fertil. 1993, 99, 421–425. [Google Scholar] [CrossRef]
  86. Caflisch, C.R.; DuBose, T.D. Cadmium-induced changes in luminal fluid pH in testis and epididymis of the rat in vivo. J. Toxicol. Environ. Health 1991, 32, 49–57. [Google Scholar] [CrossRef] [PubMed]
  87. Herak-Kramberger, C.M.; Sabolić, I.; Blanusa, M.; Smith, P.J.; Brown, D.; Breton, S. Cadmium inhibits vacuolar H+ATPase-mediated acidification in the rat epididymis. Biol. Reprod. 2000, 63, 599–606. [Google Scholar] [CrossRef] [PubMed]
  88. Da Silva, N.; Shum, W.W.; Breton, S. Regulation of vacuolar proton pumping ATPase-dependent luminal acidification in the epididymis. Asian J. Androl. 2007, 9, 476–482. [Google Scholar] [CrossRef] [PubMed]
  89. Santiago, J.; Silva, J.V.; Howl, J.; Santos, M.A.; Fardilha, M. All you need to know about sperm RNAs. Hum. Reprod. Update 2022, 28, 67–91. [Google Scholar] [CrossRef]
  90. 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]
  91. Kirichok, Y.; Navarro, B.; Clapham, D.E. Whole-cell patch-clamp measurements of spermatozoa reveal an alkaline-activated Ca2+ channel. Nature 2006, 439, 737–740. [Google Scholar] [CrossRef] [PubMed]
  92. Navarro, B.; Kirichok, Y.; Clapham, D.E. KSper, a pH-sensitive K+ current that controls sperm membrane potential. Proc. Natl. Acad. Sci. USA 2007, 104, 7688–7692. [Google Scholar] [CrossRef] [PubMed]
  93. 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] [PubMed]
  94. Brenker, C.; Goodwin, N.; Weyand, I.; Kashikar, N.D.; Naruse, M.; Krähling, M.; Müller, A.; Kaupp, U.B.; Strünker, T. The CatSper channel: A polymodal chemosensor in human sperm. EMBO J. 2012, 31, 1654–1665. [Google Scholar] [CrossRef] [PubMed]
  95. Romero, F.; Nishigaki, T. Comparative genomic analysis suggests that the sperm-specific sodium/proton exchanger and soluble adenylyl cyclase are key regulators of CatSper among the Metazoa. Zool. Lett. 2019, 5, 25. [Google Scholar] [CrossRef] [PubMed]
  96. Carlson, A.E.; Westenbroek, R.E.; Quill, T.; Ren, D.; Clapham, D.E.; Hille, B.; Garbers, D.L.; Babcock, D.F. CatSper1 required for evoked Ca2+ entry and control of flagellar function in sperm. Proc. Natl. Acad. Sci. USA 2003, 100, 14864–14868. [Google Scholar] [CrossRef] [PubMed]
  97. Quill, T.A.; Sugden, S.A.; Rossi, K.L.; Doolittle, L.K.; Hammer, R.E.; Garbers, D.L. Hyperactivated sperm motility driven by CatSper2 is required for fertilization. Proc. Natl. Acad. Sci. USA 2003, 100, 14869–14874. [Google Scholar] [CrossRef] [PubMed]
  98. Jin, J.; Jin, N.; Zheng, H.; Ro, S.; Tafolla, D.; Sanders, K.M.; Yan, W. Catsper3 and Catsper4 are essential for sperm hyperactivated motility and male fertility in the mouse. Biol. Reprod. 2007, 77, 37–44. [Google Scholar] [CrossRef] [PubMed]
  99. Ren, D.; Navarro, B.; Perez, G.; Jackson, A.C.; Hsu, S.; Shi, Q.; Tilly, J.L.; Clapham, D.E. A sperm ion channel required for sperm motility and male fertility. Nature 2001, 413, 603–609. [Google Scholar] [CrossRef] [PubMed]
  100. 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]
  101. Lishko, P.V.; Kirichok, Y. The role of Hv1 and CatSper channels in sperm activation. J. Physiol. 2010, 588, 4667–4672. [Google Scholar] [CrossRef]
  102. 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]
  103. 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] [PubMed]
  104. 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] [PubMed]
  105. Zeng, X.-H.; Yang, C.; Xia, X.-M.; Liu, M.; Lingle, C.J. SLO3 auxiliary subunit LRRC52 controls gating of sperm KSPER currents and is critical for normal fertility. Proc. Natl. Acad. Sci. USA 2015, 112, 2599–2604. [Google Scholar] [CrossRef] [PubMed]
  106. Kang, H.; Liu, M.; Zhang, W.; Huang, R.-Z.; Zhao, N.; Chen, C.; Zeng, X.-H. Na+/H+ Exchangers Involve in Regulating the pH-Sensitive Ion Channels in Mouse Sperm. Int. J. Mol. Sci. 2021, 22, 1612. [Google Scholar] [CrossRef]
  107. Chávez, J.C.; Ferreira, J.J.; Butler, A.; De La Vega Beltrán, J.L.; Treviño, C.L.; Darszon, A.; Salkoff, L.; Santi, C.M. SLO3 K+ channels control calcium entry through CATSPER channels in sperm. J. Biol. Chem. 2014, 289, 32266–32275. [Google Scholar] [CrossRef] [PubMed]
  108. 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]
  109. Mannowetz, N.; Naidoo, N.M.; Choo, S.-A.S.; Smith, J.F.; Lishko, P.V. SLO1 is the principal potassium channel of human spermatozoa. eLife 2013, 2, e01009. [Google Scholar] [CrossRef] [PubMed]
  110. Kaupp, U.B.; Strunker, T. Signaling in Sperm: More Different than Similar. Trends Cell Biol. 2017, 27, 101–109. [Google Scholar] [CrossRef] [PubMed]
  111. Brown, S.G.; Publicover, S.J.; Mansell, S.A.; Lishko, P.V.; Williams, H.L.; Ramalingam, M.; Wilson, S.M.; Barratt, C.L.R.; Sutton, K.A.; Da Silva, S.M. Depolarization of sperm membrane potential is a common feature of men with subfertility and is associated with low fertilization rate at IVF. Hum. Reprod. 2016, 31, 1147–1157. [Google Scholar] [CrossRef]
  112. Mishra, A.K.; Kumar, A.; Swain, D.K.; Yadav, S.; Nigam, R. Insights into pH regulatory mechanisms in mediating spermatozoa functions. Vet. World 2018, 11, 852–858. [Google Scholar] [CrossRef] [PubMed]
  113. Giroux-Widemann, V.; Jouannet, P.; Pignot-Paintrand, I.; Feneux, D. Effects of pH on the reactivation of human spermatozoa demembranated with Triton X-100. Mol. Reprod. Dev. 1991, 29, 157–162. [Google Scholar] [CrossRef] [PubMed]
  114. Alavi, S.M.; Matsumura, N.; Shiba, K.; Itoh, N.; Takahashi, K.G.; Inaba, K.; Osada, M. Roles of extracellular ions and pH in 5-HT-induced sperm motility in marine bivalve. Reproduction 2014, 147, 331–345. [Google Scholar] [CrossRef] [PubMed]
  115. Lavanya, M.; Selvaraju, S. Microenvironment of the male and female reproductive tracts regulate sperm fertility: Impact of viscosity, pH, and osmolality. Andrology 2022, 10, 92–104. [Google Scholar] [CrossRef] [PubMed]
  116. Speer, K.F.; Allen-Waller, L. Molecular mechanisms of sperm motility are conserved in an early-branching metazoan. Proc. Natl. Acad. Sci. USA 2021, 118, e2109993118. [Google Scholar] [CrossRef] [PubMed]
  117. Vacquier, V.D.; Loza-Huerta, A.; García-Rincón, J.; Darszon, A.; Beltrán, C. Soluble adenylyl cyclase of sea urchin spermatozoa. Biochim. Biophys. Acta 2014, 1842, 2621–2628. [Google Scholar] [CrossRef] [PubMed]
  118. Xie, F.; Garcia, M.A.; Carlson, A.E.; Schuh, S.M.; Babcock, D.F.; Jaiswal, B.S.; Gossen, J.A.; Esposito, G.; van Duin, M.; Conti, M. Soluble adenylyl cyclase (sAC) is indispensable for sperm function and fertilization. Dev. Biol. 2006, 296, 353–362. [Google Scholar] [CrossRef] [PubMed]
  119. Boulais, M.; Suquet, M.; Arsenault-Pernet, E.J. pH controls spermatozoa motility in the Pacific oyster (Crassostrea gigas). Biol. Open 2018, 7, bio031427. [Google Scholar] [CrossRef] [PubMed]
  120. Pérez, L.; Gallego, V.; Asturiano, J.F. Intracellular pH regulation and sperm motility in the European eel. Theriogenology 2020, 145, 48–58. [Google Scholar] [CrossRef] [PubMed]
  121. Boitano, S.; Omoto, C.K. Membrane hyperpolarization activates trout sperm without an increase in intracellular pH. J. Cell Sci. 1991, 98 Pt 3, 343–349. [Google Scholar] [CrossRef]
  122. Zhou, J.; Chen, L.; Li, J.; Li, H.; Hong, Z.; Xie, M.; Chen, S.; Yao, B. The Semen pH Affects Sperm Motility and Capacitation. PLoS ONE 2015, 10, e0132974. [Google Scholar] [CrossRef] [PubMed]
  123. Dhumal, S.S.; Naik, P.; Dakshinamurthy, S.; Sullia, K. Semen pH and its correlation with motility and count—A study in subfertile men. JBRA Assist. Reprod. 2021, 25, 172–175. [Google Scholar] [CrossRef] [PubMed]
  124. Mannowetz, N.; Wandernoth, P.M.; Wennemuth, G. Glucose is a pH-dependent motor for sperm beat frequency during early activation. PLoS ONE 2012, 7, e41030. [Google Scholar] [CrossRef] [PubMed]
  125. Liu, C.H.; Dong, H.B.; Ma, D.L.; Li, Y.W.; Han, D.; Luo, M.J.; Chang, Z.L.; Tan, J.H. Effects of pH during liquid storage of goat semen on sperm viability and fertilizing potential. Anim. Reprod. Sci. 2016, 164, 47–56. [Google Scholar] [CrossRef]
  126. de Mercado, E.; Tomás-Almenar, C.; Gómez-Izquierdo, E. Improvement of the motility of boar sperm after cryopreservation. Anim. Reprod. Sci. 2020, 222, 106610. [Google Scholar] [CrossRef] [PubMed]
  127. He, Q.; Wu, S.; Huang, M.; Wang, Y.; Zhang, K.; Kang, J.; Zhang, Y.; Quan, F. Effects of Diluent pH on Enrichment and Performance of Dairy Goat X/Y Sperm. Front. Cell Dev. Biol. 2021, 9, 747722. [Google Scholar] [CrossRef] [PubMed]
  128. Huang, M.; Cao, X.Y.; He, Q.F.; Yang, H.W.; Chen, Y.Z.; Zhao, J.L.; Ma, H.W.; Kang, J.; Liu, J.; Quang, F.S. Alkaline semen diluent combined with R848 for separation and enrichment of dairy goat X-sperm. J. Dairy Sci. 2022, 105, 10020–10032. [Google Scholar] [CrossRef]
  129. He, Q.; Wu, S.; Gao, F.; Xu, X.; Wang, S.; Xu, Z.; Huang, M.; Zhang, K.; Zhang, Y.; Quan, F. Diluent pH affects sperm motility via GSK3 α/β-hexokinase pathway for the efficient enrichment of X-sperm to increase the female kids rate of dairy goats. Theriogenology 2023, 201, 1–11. [Google Scholar] [CrossRef] [PubMed]
  130. Park, Y.J.; Shin, D.H.; Pang, W.K.; Ryu, D.Y.; Rahman, M.S.; Adegoke, E.O.; Pang, M.G. Short-term storage of semen samples in acidic extender increases the proportion of females in pigs. BMC Vet. Res. 2021, 17, 362. [Google Scholar] [CrossRef] [PubMed]
  131. Park, Y.J.; Kwon, K.J.; Song, W.H.; Pang, W.K.; Ryu, D.Y.; Saidur Rahman, M.; Pang, M.G. New technique of sex preselection for increasing female ratio in boar sperm model. Reprod. Domest. Anim. 2021, 56, 333–341. [Google Scholar] [CrossRef] [PubMed]
  132. Leemans, B.; Stout, T.A.E.; De Schauwer, C.; Heras, S.; Nelis, H.; Hoogewijs, M.; Van Soom, A.; Gadella, B.M. Update on mammalian sperm capacitation: How much does the horse differ from other species? Reproduction 2019, 157, R181–R197. [Google Scholar] [CrossRef] [PubMed]
  133. 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] [PubMed]
  134. Zeng, Y.; Oberdorf, J.A.; Florman, H.M. pH regulation in mouse sperm: Identification of Na+-, Cl, and HCO3-dependent and arylaminobenzoate-dependent regulatory mechanisms and characterization of their roles in sperm capacitation. Dev. Biol. 1996, 173, 510–520. [Google Scholar] [CrossRef] [PubMed]
  135. Marquez, B.; Suarez, S.S. Bovine sperm hyperactivation is promoted by alkaline-stimulated Ca2+ influx. Biol. Reprod. 2007, 76, 660–665. [Google Scholar] [CrossRef] [PubMed]
  136. Ho, H.C.; Suarez, S.S. Hyperactivation of mammalian spermatozoa: Function and regulation. Reproduction 2001, 122, 519–526. [Google Scholar] [CrossRef] [PubMed]
  137. Ho, H.-C.; Granish, K.A.; Suarez, S.S. Hyperactivated motility of bull sperm is triggered at the axoneme by Ca2+ and not cAMP. Dev. Biol. 2002, 250, 208–217. [Google Scholar] [CrossRef] [PubMed]
  138. 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] [PubMed]
  139. Chen, Y.; Cann, M.J.; Litvin, T.N.; Iourgenko, V.; Sinclair, M.L.; Levin, L.R.; Buck, J. Soluble adenylyl cyclase as an evolutionarily conserved bicarbonate sensor. Science 2000, 289, 625–628. [Google Scholar] [CrossRef] [PubMed]
  140. De La Vega-Beltran, J.L.; Sánchez-Cárdenas, C.; Krapf, D.; Hernandez-González, E.O.; Wertheimer, E.; Treviño, C.L.; Visconti, P.E.; Darszon, A. Mouse sperm membrane potential hyperpolarization is necessary and sufficient to prepare sperm for the acrosome reaction. J. Biol. Chem. 2012, 287, 44384–44393. [Google Scholar] [CrossRef] [PubMed]
  141. Esposito, G.; Jaiswal, B.S.; Xie, F.; Krajnc-Franken, M.A.M.; Robben, T.J.A.A.; Strik, A.M.; Kuil, C.; Philipsen, R.L.A.; van Duin, M.; Conti, M.; et al. Mice deficient for soluble adenylyl cyclase are infertile because of a severe sperm-motility defect. Proc. Natl. Acad. Sci. USA 2004, 101, 2993–2998. [Google Scholar] [CrossRef] [PubMed]
  142. 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] [PubMed]
  143. 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] [PubMed]
  144. Pierucci-Alves, F.; Akoyev, V.; Schultz, B.D. Bicarbonate exchangers SLC26A3 and SLC26A6 are localized at the apical membrane of porcine vas deferens epithelium. Physiol. Rep. 2015, 3, e12380. [Google Scholar] [CrossRef]
  145. 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] [CrossRef] [PubMed]
  146. Schweinfest, C.W.; Spyropoulos, D.D.; Henderson, K.W.; Kim, J.-H.; Chapman, J.M.; Barone, S.; Worrell, R.T.; Wang, Z.; Soleimani, M. Slc26a3 (dra)-deficient mice display chloride-losing diarrhea, enhanced colonic proliferation, and distinct up-regulation of ion transporters in the colon. J. Biol. Chem. 2006, 281, 37962–37971. [Google Scholar] [CrossRef] [PubMed]
  147. 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. Genet. 2012, 21, 1287–1298. [Google Scholar] [CrossRef]
  148. Quinton, P.M. Too much salt, too little soda: Cystic fibrosis. Sheng Li Xue Bao 2007, 59, 397–415. [Google Scholar] [PubMed]
  149. Li, C.-Y.; Jiang, L.-Y.; Chen, W.-Y.; Li, K.; Sheng, H.-Q.; Ni, Y.; Lu, J.-X.; Xu, W.-X.; Zhang, S.-Y.; Shi, Q.-X. CFTR is essential for sperm fertilizing capacity and is correlated with sperm quality in humans. Hum. Reprod. 2010, 25, 317–327. [Google Scholar] [CrossRef] [PubMed]
  150. 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]
  151. 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, 9924–9936. [Google Scholar] [CrossRef] [PubMed]
  152. 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]
  153. Liang, M.; Ji, N.; Song, J.; Kang, H. Flagellar pH homeostasis mediated by Na+/H+ exchangers regulates human sperm functions through coupling with CatSper and KSper activation. Hum. Reprod. 2024, 39, 674–688. [Google Scholar] [CrossRef] [PubMed]
  154. Donowitz, M.; Tse, C.M.; Fuster, D. SLC9/NHE gene family, a plasma membrane and organellar family of Na+/H+ exchangers. Mol. Asp. Med. 2013, 34, 236–251. [Google Scholar] [CrossRef] [PubMed]
  155. Woo, A.L.; James, P.F.; Lingrel, J.B. Roles of the Na,K-ATPase alpha4 isoform and the Na+/H+ exchanger in sperm motility. Mol. Reprod. Dev. 2002, 62, 348–356. [Google Scholar] [CrossRef] [PubMed]
  156. 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] [PubMed]
  157. Bell, S.M.; Schreiner, C.M.; Schultheis, P.J.; Miller, M.L.; Evans, R.L.; Vorhees, C.V.; Shull, G.E.; Scott, W.J. Targeted disruption of the murine Nhe1 locus induces ataxia, growth retardation, and seizures. Am. J. Physiol. Cell Physiol. 1999, 276, C788–C795. [Google Scholar] [CrossRef] [PubMed]
  158. Chen, X.; Wang, X.; Tang, L.; Wang, J.; Shen, C.; Liu, J.; Lu, S.; Zhang, H.; Kuang, Y.; Fei, J. Nhe5 deficiency enhances learning and memory via upregulating BDNF/TRKB signaling in mice. Am. J. Med. Genet. Part B Neuropsychiatr. Genet. 2017, 174, 828–838. [Google Scholar] [CrossRef] [PubMed]
  159. Chew, S.B.; Leung, G.P.; Leung, P.Y.; Tse, C.M.; Wong, P.Y. Polarized distribution of NHE1 and NHE2 in the rat epididymis. Biol. Reprod. 2000, 62, 755–758. [Google Scholar] [CrossRef] [PubMed]
  160. Malakooti, J.; Dahdal, R.Y.; Schmidt, L.; Layden, T.J.; Dudeja, P.K.; Ramaswamy, K. Molecular cloning, tissue distribution, and functional expression of the human Na+/H+ exchanger NHE2. Am. J. Physiol. 1999, 277, G383–G390. [Google Scholar] [CrossRef] [PubMed]
  161. Oberheide, K.; Puchkov, D.; Jentsch, T.J. Loss of the Na+/H+ exchanger NHE8 causes male infertility in mice by disrupting acrosome formation. J. Biol. Chem. 2017, 292, 10845–10854. [Google Scholar] [CrossRef] [PubMed]
  162. Miller, M.R.; Mansell, S.A.; Meyers, S.A.; Lishko, P.V. Flagellar ion channels of sperm: Similarities and differences between species. Cell Calcium. 2015, 58, 105–113. [Google Scholar] [CrossRef] [PubMed]
  163. Gardner, C.C.; James, P.F. The Slc9c2 Gene Product (Na+/H+ Exchanger Isoform 11; NHE11) Is a Testis-Specific Protein Localized to the Head of Mature Mammalian Sperm. Int. J. Mol. Sci. 2023, 24, 5329. [Google Scholar] [CrossRef] [PubMed]
  164. 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] [PubMed]
  165. Seredenina, T.; Demaurex, N.; Krause, K.-H. Voltage-Gated Proton Channels as Novel Drug Targets: From NADPH Oxidase Regulation to Sperm Biology. Antioxid. Redox. Signal 2015, 23, 490–513. [Google Scholar] [CrossRef] [PubMed]
  166. Chae, M.R.; Kang, S.J.; Lee, K.P.; Choi, B.R.; Kim, H.K.; Park, J.K.; Kim, C.Y.; Lee, S.W. Onion (Allium cepa L.) peel extract (OPE) regulates human sperm motility via protein kinase C-mediated activation of the human voltage-gated proton channel. Andrology 2017, 5, 979–989. [Google Scholar] [CrossRef] [PubMed]
  167. Musset, B.; Capasso, M.; Cherny, V.V.; Morgan, D.; Bhamrah, M.; Dyer, M.J.S.; DeCoursey, T.E. Identification of Thr29 as a critical phosphorylation site that activates the human proton channel HVCN1 in leukocytes. J. Biol. Chem. 2010, 285, 5117–5121. [Google Scholar] [CrossRef] [PubMed]
  168. 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] [PubMed]
  169. 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] [PubMed]
  170. Zhao, R.; Dai, H.; Arias, R.J.; De Blas, G.A.; Orta, G.; Pavarotti, M.A.; Shen, R.; Perozo, E.; Mayorga, L.S.; Darszon, A.; et al. Direct activation of the proton channel by albumin leads to human sperm capacitation and sustained release of inflammatory mediators by neutrophils. Nat. Commun. 2021, 12, 3855. [Google Scholar] [CrossRef] [PubMed]
  171. Han, S.; Chu, X.-P.; Goodson, R.; Gamel, P.; Peng, S.; Vance, J.; Wang, S. Cholesterol inhibits human voltage-gated proton channel hHv1. Proc. Natl. Acad. Sci. USA 2022, 119, e2205420119. [Google Scholar] [CrossRef] [PubMed]
  172. Casslén, B.; Nilsson, B. Human uterine fluid, examined in undiluted samples for osmolarity and the concentrations of inorganic ions, albumin, glucose, and urea. Am. J. Obs. Gynecol. 1984, 150, 877–881. [Google Scholar] [CrossRef] [PubMed]
  173. Allouche-Fitoussi, D.; Breitbart, H. The Role of Zinc in Male Fertility. Int. J. Mol. Sci. 2020, 21, 7796. [Google Scholar] [CrossRef] [PubMed]
  174. Vigil, P.; Orellana, R.F.; Cortés, M.E. Modulation of spermatozoon acrosome reaction. Biol. Res. 2011, 44, 151–159. [Google Scholar] [CrossRef] [PubMed]
  175. Hirose, M.; Honda, A.; Fulka, H.; Tamura-Nakano, M.; Matoba, S.; Tomishima, T.; Mochida, K.; Hasegawa, A.; Nagashima, K.; Inoue, K.; et al. Acrosin is essential for sperm penetration through the zona pellucida in hamsters. Proc. Natl. Acad. Sci. USA 2020, 117, 2513–2518. [Google Scholar] [CrossRef] [PubMed]
  176. Nakanishi, T.; Ikawa, M.; Yamada, S.; Toshimori, K.; Okabe, M. Alkalinization of acrosome measured by GFP as a pH indicator and its relation to sperm capacitation. Dev. Biol. 2001, 237, 222–231. [Google Scholar] [CrossRef] [PubMed]
  177. Working, P.K.; Meizel, S. Correlation of increased intraacrosomal pH with the hamster sperm acrosome reaction. J. Exp. Zool. 1983, 227, 97–107. [Google Scholar] [CrossRef] [PubMed]
  178. Panneerdoss, S.; Siva, A.B.; Kameshwari, D.B.; Rangaraj, N.; Shivaji, S. Association of lactate, intracellular pH, and intracellular calcium during capacitation and acrosome reaction: Contribution of hamster sperm dihydrolipoamide dehydrogenase, the E3 subunit of pyruvate dehydrogenase complex. J. Androl. 2012, 33, 699–710. [Google Scholar] [CrossRef] [PubMed]
  179. Guyonnet, B.; Egge, N.; Cornwall, G.A. Functional amyloids in the mouse sperm acrosome. Mol. Cell Biol. 2014, 34, 2624–2634. [Google Scholar] [CrossRef] [PubMed]
  180. Mata-Martínez, E.; Darszon, A.; Treviño, C.L. pH-dependent Ca2+ oscillations prevent untimely acrosome reaction in human sperm. Biochem. Biophys. Res. Commun. 2018, 497, 146–152. [Google Scholar] [CrossRef] [PubMed]
  181. Cross, N.L. Effect of pH on the development of acrosomal responsiveness of human sperm. Andrologia 2007, 39, 55–59. [Google Scholar] [CrossRef] [PubMed]
  182. Carrasquel Martínez, G.; Aldana, A.; Carneiro, J.; Treviño, C.L.; Darszon, A. Acrosomal alkalinization occurs during human sperm capacitation. Mol. Hum. Reprod. 2022, 28, gaac005. [Google Scholar] [CrossRef] [PubMed]
  183. Chávez, J.C.; De la Vega-Beltrán, J.L.; José, O.; Torres, P.; Nishigaki, T.; Treviño, C.L.; Darszon, A. Acrosomal alkalization triggers Ca2+ release and acrosome reaction in mammalian spermatozoa. J. Cell. Physiol. 2018, 233, 4735–4747. [Google Scholar] [CrossRef] [PubMed]
  184. Prajapati, P.; Kane, S. Elevated and Sustained Intracellular Calcium Signalling Is Necessary for Efficacious Induction of the Human Sperm Acrosome Reaction. Int. J. Mol. Sci. 2022, 23, 11253. [Google Scholar] [CrossRef] [PubMed]
  185. Donà, G.; Tibaldi, E.; Andrisani, A.; Ambrosini, G.; Sabbadin, C.; Pagano, M.A.; Brunati, A.M.; Armanini, D.; Ragazzi, E.; Bordin, L. Human Sperm Capacitation Involves the Regulation of the Tyr-Phosphorylation Level of the Anion Exchanger 1 (AE1). Int. J. Mol. Sci. 2020, 21, 4063. [Google Scholar] [CrossRef] [PubMed]
  186. El Khouri, E.; Touré, A. Functional interaction of the cystic fibrosis transmembrane conductance regulator with members of the SLC26 family of anion transporters (SLC26A8 and SLC26A9): Physiological and pathophysiological relevance. Int. J. Biochem. Cell Biol. 2014, 52, 58–67. [Google Scholar] [CrossRef] [PubMed]
  187. 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]
  188. Genetet, S.; Desrames, A.; Chouali, Y.; Ripoche, P.; Lopez, C.; Mouro-Chanteloup, I. Stomatin modulates the activity of the Anion Exchanger 1 (AE1, SLC4A1). Sci. Rep. 2017, 7, 46170. [Google Scholar] [CrossRef]
  189. Yannoukakos, D.; Stuart-Tilley, A.; Fernandez, H.A.; Fey, P.; Duyk, G.; Alper, S.L. Molecular cloning, expression, and chromosomal localization of two isoforms of the AE3 anion exchanger from human heart. Circ. Res. 1994, 75, 603–614. [Google Scholar] [CrossRef] [PubMed]
  190. Thorsen, K.; Dam, V.S.; Kjaer-Sorensen, K.; Pedersen, L.N.; Skeberdis, V.A.; Jurevičius, J.; Treinys, R.; Petersen, I.; Nielsen, M.S.; Oxvig, C.; et al. Loss-of-activity-mutation in the cardiac chloride-bicarbonate exchanger AE3 causes short QT syndrome. Nat. Commun. 2017, 8, 1696. [Google Scholar] [CrossRef] [PubMed]
  191. Felder, R.A.; Jose, P.A.; Xu, P.; Gildea, J.J. The renal sodium bicarbonate cotransporter NBCe2: Is it a major contributor to sodium and pH homeostasis? Curr. Hypertens. Rep. 2016, 18, 1–9. [Google Scholar] [CrossRef] [PubMed]
  192. L Bernardino, R.; T Jesus, T.; D Martins, A.; Sousa, M.; Barros, A.; E Cavaco, J.; Socorro, S.; G Alves, M.; F Oliveira, P. Molecular basis of bicarbonate membrane transport in the male reproductive tract. Curr. Med. Chem. 2013, 20, 4037–4049. [Google Scholar] [CrossRef] [PubMed]
  193. Pérez-Carrillo, L.; Aragón-Herrera, A.; Giménez-Escamilla, I.; Delgado-Arija, M.; García-Manzanares, M.; Anido-Varela, L.; Lago, F.; Martínez-Dolz, L.; Portolés, M.; Tarazón, E.J.P. Cardiac sodium/hydrogen exchanger (NHE11) as a novel potential target for SGLT2i in heart failure: A preliminary study. Pharmaceutics 2022, 14, 1996. [Google Scholar] [CrossRef] [PubMed]
  194. Lindskog, S. Structure and mechanism of carbonic anhydrase. Pharmacol. Ther. 1997, 74, 1–20. [Google Scholar] [CrossRef] [PubMed]
  195. Haapasalo, J.; Nordfors, K.; Haapasalo, H.; Parkkila, S. The Expression of Carbonic Anhydrases II, IX and XII in Brain Tumors. Cancers 2020, 12, 1723. [Google Scholar] [CrossRef] [PubMed]
  196. Sly, W.S.; Hewett-Emmett, D.; Whyte, M.P.; Yu, Y.S.; Tashian, R.E. Carbonic anhydrase II deficiency identified as the primary defect in the autosomal recessive syndrome of osteopetrosis with renal tubular acidosis and cerebral calcification. Proc. Natl. Acad. Sci. USA 1983, 80, 2752–2756. [Google Scholar] [CrossRef]
  197. Lönnerholm, G.; Knutson, L.; Wistrand, P.J.; Flemström, G. Carbonic anhydrase in the normal rat stomach and duodenum and after treatment with omeprazole and ranitidine. Acta Physiol. Scand. 1989, 136, 253–262. [Google Scholar] [CrossRef] [PubMed]
  198. Wang, G.; Cheng, Z.; Liu, F.; Zhang, H.; Li, J.; Li, F. CREB is a key negative regulator of carbonic anhydrase IX (CA9) in gastric cancer. Cell Signal. 2015, 27, 1369–1379. [Google Scholar] [CrossRef]
  199. Nortunen, M.; Parkkila, S.; Saarnio, J.; Huhta, H.; Karttunen, T.J. Carbonic Anhydrases II and IX in Non-ampullary Duodenal Adenomas and Adenocarcinoma. J. Histochem. Cytochem. Off. J. Histochem. Soc. 2021, 69, 677–690. [Google Scholar] [CrossRef]
  200. Liao, S.Y.; Ivanov, S.; Ivanova, A.; Ghosh, S.; Cote, M.A.; Keefe, K.; Coca-Prados, M.; Stanbridge, E.J.; Lerman, M.I. Expression of cell surface transmembrane carbonic anhydrase genes Ca9 and Ca12 in the human eye: Overexpression of CA12 (CAXII) in glaucoma. J. Med. Genet. 2003, 40, 257–261. [Google Scholar] [CrossRef] [PubMed]
  201. Wykoff, C.C.; Beasley, N.; Watson, P.H.; Campo, L.; Chia, S.K.; English, R.; Pastorek, J.; Sly, W.S.; Ratcliffe, P.; Harris, A.L. Expression of the hypoxia-inducible and tumor-associated carbonic anhydrases in ductal carcinoma in situ of the breast. Am. J. Pathol. 2001, 158, 1011–1019. [Google Scholar] [CrossRef]
  202. Song, X.; Zhu, S.; Xie, Y.; Liu, J.; Sun, L.; Zeng, D.; Wang, P.; Ma, X.; Kroemer, G.; Bartlett, D.L.; et al. JTC801 Induces pH-dependent Death Specifically in Cancer Cells and Slows Growth of Tumors in Mice. Gastroenterology 2018, 154, 1480–1493. [Google Scholar] [CrossRef] [PubMed]
  203. Lee, M.; Vecchio-Pagán, B.; Sharma, N.; Waheed, A.; Li, X.; Raraigh, K.S.; Robbins, S.; Han, S.T.; Franca, A.L.; Pellicore, M.J.; et al. Loss of carbonic anhydrase XII function in individuals with elevated sweat chloride concentration and pulmonary airway disease. Hum. Mol. Genet. 2016, 25, 1923–1933. [Google Scholar] [CrossRef] [PubMed]
  204. Silagi, E.S.; Schoepflin, Z.R.; Seifert, E.L.; Merceron, C.; Schipani, E.; Shapiro, I.M.; Risbud, M.V. Bicarbonate Recycling by HIF-1-Dependent Carbonic Anhydrase Isoforms 9 and 12 Is Critical in Maintaining Intracellular pH and Viability of Nucleus Pulposus Cells. J. Bone Miner. Res. Off. J. Am. Soc. Bone Miner. Res. 2018, 33, 338–355. [Google Scholar] [CrossRef] [PubMed]
  205. Groeneweg, S.; Kersseboom, S.; van den Berge, A.; Dolcetta-Capuzzo, A.; van Geest, F.S.; van Heerebeek, R.E.A.; Arjona, F.J.; Meima, M.E.; Peeters, R.P.; Visser, W.E.; et al. In Vitro Characterization of Human, Mouse, and Zebrafish MCT8 Orthologues. Thyroid 2019, 29, 1499–1510. [Google Scholar] [CrossRef] [PubMed]
  206. Meng, X.; Yuan, H.; Li, W.; Xiong, Z.; Dong, W.; Xiao, W.; Zhang, X. Solute carrier family 16 member 5 downregulation and its methylation might serve as a prognostic indicator of prostate cancer. IUBMB Life 2021, 73, 1363–1377. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The graphical representation of pH variation in male and female reproductive systems. pH in the testis is determined at 7.2~7.4, which is conducive to maintaining spermatogenesis. From the initial segment to the cauda epididymis, the pH is approximately 6.5~6.8, where the sperm gradually matures and finally enters the vas deferens with a pH of around 7.2~7.4. After ejaculation, from the vagina to the cervix, the pH gradually alkalinizes, increasing from 4.3 to 6.5~7.5; pH 7.0~7.8 is found in the uterus, and in such an alkaline environment, sperm is capacitated and hyperactivated. pH 7.3~7.7 is maintained in the fallopian tube and contributes to AR in sperm to further fertilization.
Figure 1. The graphical representation of pH variation in male and female reproductive systems. pH in the testis is determined at 7.2~7.4, which is conducive to maintaining spermatogenesis. From the initial segment to the cauda epididymis, the pH is approximately 6.5~6.8, where the sperm gradually matures and finally enters the vas deferens with a pH of around 7.2~7.4. After ejaculation, from the vagina to the cervix, the pH gradually alkalinizes, increasing from 4.3 to 6.5~7.5; pH 7.0~7.8 is found in the uterus, and in such an alkaline environment, sperm is capacitated and hyperactivated. pH 7.3~7.7 is maintained in the fallopian tube and contributes to AR in sperm to further fertilization.
Biomolecules 14 00685 g001
Figure 2. The schematic diagram presents the transport of H+ or HCO3 in the testis and epididymis. (A) In the testis, SLC4A2 primarily transports Cl and HCO3 in germ and SCs, and MCT1 transports H+. NDCBE redistributes Na+, Cl and HCO3. NBCn1 and NBCe1 regulate Na+ and HCO3 transportation. (B) In the epididymis, HCO3 reabsorption plays a vital role in pH determination. NHE3, CA III, CA IV, and CA XIV collaborate to manage HCO3 redistribution. In addition, H-ATPase in the caput and cauda epididymis regulates pH by participating in H+ secretion in the epididymal lumen.
Figure 2. The schematic diagram presents the transport of H+ or HCO3 in the testis and epididymis. (A) In the testis, SLC4A2 primarily transports Cl and HCO3 in germ and SCs, and MCT1 transports H+. NDCBE redistributes Na+, Cl and HCO3. NBCn1 and NBCe1 regulate Na+ and HCO3 transportation. (B) In the epididymis, HCO3 reabsorption plays a vital role in pH determination. NHE3, CA III, CA IV, and CA XIV collaborate to manage HCO3 redistribution. In addition, H-ATPase in the caput and cauda epididymis regulates pH by participating in H+ secretion in the epididymal lumen.
Biomolecules 14 00685 g002
Figure 3. The various transporters and ion channels cooperatively regulate the pH in the sperm plasma membrane. In the acrosome of the sperm head, H+-ATPase is present to regulate H+ secretion, and SLC4A1 is involved in HCO3 transport. In equatorial and midpieces, SLC26A3, SLC26A8 and cystic fibrosis transmembrane conductance regulator (CFTR) are found to participate in Cl and HCO3 transport, in which SLC26A8 and CFTR are co-localized. In the principal piece of the flagellum, NHA1, NHA2, sNHE, and HVCN1 are involved in the regulation of H+ expulsion, among which HVCN1 is only present in human, bovine and pig spermatozoa. In addition, two sperm-specific cation channels, CatSper and SLO3, are located in the principal piece and activated by pH alkalization, mediating external calcium inflow and K+ expulsion, respectively. The co-localization of CatSper and HVCN1 is revealed.
Figure 3. The various transporters and ion channels cooperatively regulate the pH in the sperm plasma membrane. In the acrosome of the sperm head, H+-ATPase is present to regulate H+ secretion, and SLC4A1 is involved in HCO3 transport. In equatorial and midpieces, SLC26A3, SLC26A8 and cystic fibrosis transmembrane conductance regulator (CFTR) are found to participate in Cl and HCO3 transport, in which SLC26A8 and CFTR are co-localized. In the principal piece of the flagellum, NHA1, NHA2, sNHE, and HVCN1 are involved in the regulation of H+ expulsion, among which HVCN1 is only present in human, bovine and pig spermatozoa. In addition, two sperm-specific cation channels, CatSper and SLO3, are located in the principal piece and activated by pH alkalization, mediating external calcium inflow and K+ expulsion, respectively. The co-localization of CatSper and HVCN1 is revealed.
Biomolecules 14 00685 g003
Table 1. The silence of pH regulators causes male reproductive deficiency.
Table 1. The silence of pH regulators causes male reproductive deficiency.
Protein Symbol,
Gene ID
LocationFunctionsPhenotype (Knockout in Mouse)Clinical Symptoms/Disease
SLC4A2
(AE2), 6522
Sertoli cells (SCs), spermatozoa [25,28]Sodium-independent anion exchanger, transports Cl/HCO3 in sperm.Spermatogenesis interrupted, male infertility [29]NR
SLC16A1
(MCT1), 6566
Elongated spermatids, spermatozoa tail [30,31]Bidirectional proton-coupled monocarboxylate transporter, transports nutrients, and regulates pH.Seminiferous tubules and SCs morphologically changed, and spermatogenesis failed [32]NR
SLC26A3, 1811Midpiece [33] The chloride/bicarbonate transporter, participates in energy acquisition, induces sperm tyrosine phosphorylation and is hyperactivated.Epididymal morphology altered, and mature sperm decreased [34]Male infertility [35]
SLC26A8
(TAT1), 116369
Annulus of spermatozoa [36,37]Co-localization with CFTR in the equatorial segment of the sperm head, involves sperm capacitation and hyperactivation.Sperm flagellar differentiation is defective, and sperm capacitation fails [36]Asthenozoospermia [37]
SLC9A3
(NHE3), 6550
Efferent duct [38,39], developing acrosomal granule [40], epididymis [41]Reabsorb Na+ in the efferent duct, and regulate lumen fluid pH.The lumen of the rete testis and efferent duct dilated, causing obstructive azoospermia [39]NR
SLC9B1
(NHA1), 150159
SLC9B2
(NHA2), 133308
Principal piece [42]Na+/H+ exchanger, regulates the intracellular pH of spermatozoa, and maintains sperm motility and fertility.Sperm count decreased, causing low fertility (NHA1 or NHA2 cKO mice) and male infertility (double KO mice) [42] Teratospermia, caused by NHA1 deficiency [43]
SLC9C1
(NHE10/sNHE), 285335
Principal piece [44]Sperm-specific sodium/hydrogen exchanger, regulates the intracellular pH of spermatozoa, and maintains sperm hyperactivation.Male infertility and asthenozoospermia [45].Asthenospermia [46]
CFTR, 1080The equatorial region of the sperm head and middle piece [47]The epithelial ion channel, maintains chloride and bicarbonate homeostasis during sperm epididymal maturation and capacitation.Homozygotes for targeted null mutations cause death around puberty. Cftr+/− male mice exhibited abnormal capacitation [48]Cystic fibrosis: progressive lung disorder, pancreatic insufficiency, male infertility [49,50]
B1 subunit of the H+-ATPase (ATP6V1B1), 110935The initial segment of the epididymis, narrow cells in the caput epididymis and bright cells in the cauda epididymis [51,52,53], acrosome [54,55]Participate in the secretion of the H+ in epididymal lumen and maintain the acidic pH of epididymal lumen fluid, and acrosome alkalization.FOXI1 regulates the synthesis of ATP6V1B1. Foxi1−/− mice sperm maturation was blocked, and fertilization failed [56]NR
NR: not reported.
Table 2. pH regulators in the male reproductive system deserve further investigation.
Table 2. pH regulators in the male reproductive system deserve further investigation.
Protein Symbol,
Gene ID
Location in Other TissuesFunctionsLocation in the Male Reproductive System
SLC4A1
(AE1), 6521
Erythroid, kidney [188]Electroneutral anion transporter and structural protein [188] Leydig cells [63]
SLC4A3
(AE3), 6508
brain, heart, and adrenal gland [189,190]Chloride ion exchangers, regulate intracellular pH and cardiac action potential [189,190]Spermatogenic cells and seminal vesicle cells [63]
SLC4A5
(NBCe2), 57835
Liver, spleen, kidney, etc. [191]Sodium-bicarbonate cotransporter [191]Testis, epididymis, prostate, and seminal vesicles [192]
SLC9A2
(NHE2, ID:6549)
Kidney, stomach [160] Plasma membrane Na+-H+ exchanger, maintains intracellular pH [160]Epididymis, vas deferens, testis [159,160]
SLC9C2
(NHE11), 284525
Heart [193]Sodium/hydrogen exchanger, correlates with human cardiac dysfunction, and regulates intracellular pH [193].Spermatozoa head [163]
CA II, 760Stomach, brain, duodenum, etc. [194,195,196,197]Catalyzes reversible hydration, creates and maintains the pH differential in tumor cells, and regulates the pH of duodenal villous epithelial cells [194,195,196,197] Epididymis (narrow cells of the initial segment and principal cells of all regions) [81]
CA IX, 768Brain, duodenum, stomach, eye [195,198,199,200]Catalyzes reversible hydration, participates in necrosis, calcification, acid-base balance, and the formation of aqueous humor, and gastric acid, and maintains pH homeostasis [198,200,201,202]Epididymis (efferent duct) [82]
CA XII, 771Expresses widely in normal tissues [203,204]Reversible hydration of carbon dioxide [203], and maintains intracellular pH [204]Epididymis [81,82]
SLC16A2
(MCT8), 6567
Brain, thyroid, kidney [205]Specific thyroid hormone transmembrane transporter, mediating efflux of thyroid hormone across cell membranes [205]Mouse testis, epididymis, epididymal spermatozoa [65]
SLC16A5
(MCT5), ID: 9121
Liver, kidney, brain, prostate, etc. [206]Proton-linked monocarboxylate transporter [206]Mouse testis, epididymis, epididymal spermatozoa [65]
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

Dai, P.; Zou, M.; Cai, Z.; Zeng, X.; Zhang, X.; Liang, M. pH Homeodynamics and Male Fertility: A Coordinated Regulation of Acid-Based Balance during Sperm Journey to Fertilization. Biomolecules 2024, 14, 685. https://doi.org/10.3390/biom14060685

AMA Style

Dai P, Zou M, Cai Z, Zeng X, Zhang X, Liang M. pH Homeodynamics and Male Fertility: A Coordinated Regulation of Acid-Based Balance during Sperm Journey to Fertilization. Biomolecules. 2024; 14(6):685. https://doi.org/10.3390/biom14060685

Chicago/Turabian Style

Dai, Pengyuan, Meng Zou, Ziyi Cai, Xuhui Zeng, Xiaoning Zhang, and Min Liang. 2024. "pH Homeodynamics and Male Fertility: A Coordinated Regulation of Acid-Based Balance during Sperm Journey to Fertilization" Biomolecules 14, no. 6: 685. https://doi.org/10.3390/biom14060685

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