Next Article in Journal
Protein Disulfide Isomerase A3 (PDIA3): A Pharmacological Target in Glioblastoma?
Next Article in Special Issue
SUMOylation Inhibition Enhances Protein Transcription under CMV Promoter: A Lesson from a Study with the F508del-CFTR Mutant
Previous Article in Journal
Decreased Hippocampal Neurogenesis in Aged Male Wistar Rats Is Not Associated with Memory Acquisition in a Water Maze
Previous Article in Special Issue
Personalized CFTR Modulator Therapy for G85E and N1303K Homozygous Patients with Cystic Fibrosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pathogenic Relationships in Cystic Fibrosis and Renal Diseases: CFTR, SLC26A9 and Anoctamins

1
Physiological Institute, University of Regensburg, University Street 31, 93053 Regensburg, Germany
2
Department of Nephrology and Hypertension, Friedrich Alexander University Erlangen Nuremberg, 91054 Erlangen, Germany
3
Department of Pediatrics, University Hospital Münster, 48149 Münster, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(17), 13278; https://doi.org/10.3390/ijms241713278
Submission received: 28 June 2023 / Revised: 31 July 2023 / Accepted: 22 August 2023 / Published: 26 August 2023
(This article belongs to the Special Issue Cystic Fibrosis and CFTR Interactions 2.0)

Abstract

:
The Cl-transporting proteins CFTR, SLC26A9, and anoctamin (ANO1; ANO6) appear to have more in common than initially suspected, as they all participate in the pathogenic process and clinical outcomes of airway and renal diseases. In the present review, we will therefore concentrate on recent findings concerning electrolyte transport in the airways and kidneys, and the role of CFTR, SLC26A9, and the anoctamins ANO1 and ANO6. Special emphasis will be placed on cystic fibrosis and asthma, as well as renal alkalosis and polycystic kidney disease. In essence, we will summarize recent evidence indicating that CFTR is the only relevant secretory Cl channel in airways under basal (nonstimulated) conditions and after stimulation by secretagogues. Information is provided on the expressions of ANO1 and ANO6, which are important for the correct expression and function of CFTR. In addition, there is evidence that the Cl transporter SLC26A9 expressed in the airways may have a reabsorptive rather than a Cl-secretory function. In the renal collecting ducts, bicarbonate secretion occurs through a synergistic action of CFTR and the Cl/HCO3 transporter SLC26A4 (pendrin), which is probably supported by ANO1. Finally, in autosomal dominant polycystic kidney disease (ADPKD), the secretory function of CFTR in renal cyst formation may have been overestimated, whereas ANO1 and ANO6 have now been shown to be crucial in ADPKD and therefore represent new pharmacological targets for the treatment of polycystic kidney disease.

1. Introduction

Cystic fibrosis (CF) is a genetic disorder caused by variants in the cystic fibrosis transmembrane conductance regulator (CFTR) gene that affects approximately 90,000 people worldwide. The absence or impaired function of the CFTR protein is associated with dysfunction in several organs, particularly the respiratory and gastrointestinal tracts. Autosomal dominant polycystic kidney disease (ADPKD) is the most commonly inherited nephropathy. It is characterized by the development and enlargement of renal cysts due to increased cell proliferation, extracellular matrix abnormalities, and increased CFTR-mediated transepithelial fluid secretion. CFTR is also required for the excretion of bicarbonate in the renal collecting ducts. A defect in renal bicarbonate excretion can lead to systemic alkalosis.
A large number of reports describe various functional changes induced either by the knockout or knockdown or the overexpression of CFTR. Although CFTR is the essential Cl channel for transepithelial Cl secretion, many of the morphological and functional changes that occur when CFTR expression is altered are thought to be caused indirectly by the up- or downregulation of intracellular signaling pathways or by metabolic changes that affect the function of other, independent proteins. In this context, physical protein–protein interactions leading to functional coupling with partner proteins (e.g., the epithelial Na+ channel ENaC or the Cl/HCO3 exchanger pendrin) are reported to control ion transport and other cell and tissue functions. Large “interactomes” for CFTR now exist, leading to the somewhat provocative question of whether CFTR “interacts with everything” [1,2,3,4,5,6]. Indeed, CFTR is a true hub for kinases and the crosstalk of cAMP and Ca2+ [7,8]. In this brief review, we focus specifically on the interplay of CFTR with the Cl transporter SLC26A9, the Ca2+-activated Cl channel anoctamin 1 (ANO1), and the phospholipid scramblase anoctamin 6 (ANO6) in the airways and kidneys.

2. CFTR Causes Constitutive Basal Cl Secretion in the Airways

The airways show basal Cl secretion in the absence of secretagogues, i.e., cAMP- or Ca2+-dependent stimulation, raising the question which of the above Cl channels and transporters are actually responsible for basal secretion. It should be noted that no spontaneous basal CFTR activity was observed in patch clamp recordings in the absence of PKA- or PKC-dependent stimulation. Previous studies provided conflicting data on the origin of basal Cl secretion. While some studies suggested spontaneous activity of CFTR as the cause of basal Cl secretion [9,10,11,12], other studies proposed SLC26A9 as the responsible transporter [13,14,15,16,17]. Until now, separating the two pathways (CFTR and SLC26A9) has been difficult because (i) there have been no specific inhibitors for SLC26A9, (ii) CFTR inhibitors do not target specifically [18], and (iii) intracellular transport and activity of SLC26A9 depend on the expression and function of CFTR [16,19,20,21,22].
In the study by Jo et al., the SLC26A9 inhibitor S9-A13 had no inhibitory effect on airway Cl transport either in vitro or ex vivo, whereas CFTRinh172 inhibited both basal and cAMP-induced Cl secretion [23]. Interestingly, neither inhibition of adenosine receptors nor inhibition of adenylate cyclase blocked basal Cl secretion, raising questions about additional mechanisms for the activation of CFTR. Thus, it is possible that increased protein kinase C (PKC) activity, e.g., through ATP release and binding to purinergic receptors, keeps part of the CFTR active at basal cAMP levels and PKA phosphorylation [24,25,26,27]. The Hanrahan lab has also recently demonstrated expression of SLC26A4 (Pendrin) in primary nasal and bronchial ciliated epithelial cells, which enhances Cl secretion through the stimulation of CFTR [28]. The molecular mechanism of STAS/R domain interaction has been previously shown for the activation of CFTR by SLC26A6 [29]. As discussed in the next section, the Ca2+-activated Cl channel ANO1 may contribute to the maintenance of basal CFTR activity through a Ca2+-dependent mechanism.

3. Relationship between CFTR and Anoctamins

Early studies showed cAMP/PKA and increases in intracellular Ca2+ as two independent second messenger pathways that lead to epithelial Cl secretion [30]. The pharmacological tools used to discriminate between both Cl conductances, however, are rather non-specific, and thus our team could not clearly keep these conductances apart [31,32]. We and others also reported that CFTR seemingly “inhibits” endogenous Ca2+-activated Cl currents (CaCC) in Xenopus oocytes, bovine pulmonary artery endothelium cells and isolated parotid acinar cells [33,34,35,36,37]. After molecular identification of the Ca2+-activated Cl channel (CACC) as anoctamin 1 (ANO1), it was found that CFTR does not inhibit ANO1 and that ANO1 currents and CFTR currents are not additive, i.e., they do not add up to the sum of both currents [38,39].
In fact, Ca2+-enhancing agonists such as purinergic or muscarinic ligands mostly activate CFTR-dependent secretion in the airways [24,40,41], while ANO1 currents rapidly inactivate it due to the mechanisms outlined in previous reports [42,43,44,45]. Therefore, after inhibition or in the absence of CFTR, CACC is very short-lived, and in the intestine, there is not even an apical CACC (ANO1) [45,46,47,48]. The rapid inactivation of ANO1 is reasonably well understood [42,49] and raises the question as to whether direct pharmacological activation of ANO1 by the synthetical compound ETX001 (called ETD-002 in the clinical trial) can be successful in restoring Cl secretion in the airways of CF patients [50,51]. It should also be noted that ANO1 is expressed at only very low levels in the airways [52].
The activation of ANO1 in CF may even be counterproductive, as ANO1 is a proinflammatory factor which enhances mucus production and mucus secretion (at least in our hands) and supports pain sensation [53,54,55,56,57]. Moreover, during inflammatory airway diseases such as asthma and CF, ANO1 is upregulated in pulmonary arterial vessels, where it supports airway constriction [52,58,59,60,61,62,63,64,65]. Finally, ANO1 supports the release of inflammatory cytokines such as IL-8 and the accumulation of pulmonary CD-45 positive cells [53]. A phase 1 clinical trial with the ANO1-activator ETD-002 finished more than a year ago, but so far, no outcome has been reported. Because of the partially conflicting studies, further studies are required in animals with clear pulmonary inflammation and increased expression of ANO1. While pure cell culture studies are insufficient, the F508del-piglet model might be very helpful in determining the true role of ANO1 in human airways, particularly under inflammatory conditions.

4. Crosstalk between CFTR and ANO1

The previous section provokes the question whether there exists a crosstalk between CFTR and ANO1. Studies reported attenuated expression of ANO1 in the apical membrane of airway epithelial cells, when coexpressed with F508del-CFTR [31,66]. We found evidence for an interaction between ANO1 and CFTR through PSD-95/Dlg/ZO-1 (PDZ) domain proteins, as described for SLC26A9 [16]. The functional interaction between ANO1 and CFTR is based on the crosstalk of intracellular Ca2+ and the intracellular cAMP signaling pathway. Crosstalk is facilitated by exchange proteins directly activated by cAMP (EPAC1) and Ca2+-sensitive adenylate cyclase type 1 (ADCY1). The assembly of such a local signalosome also depends on the presence of G-protein coupled receptors (GPCRs) [32,67,68]. In the next chapter, we will show that a functional interaction of CFTR and ANO1 also exists at the level of the membrane expression of CFTR.

5. Reduced Plasma Membrane Expression of CFTR in the Absence of ANO1

Cell-specific knockout of ANO1 in ciliated airway epithelial cells abolished Ca2+-activated Cl currents and largely reduced Ca2+-dependent Cl secretion in mouse airways. Moreover, Ca2+-dependent Cl transport was abolished in intestinal epithelial cells from epithelial-specific ANO1-knockout mice [31]. However, we reported the surprising observation that in parallel to the loss of ANO1-dependent transport, CFTR-dependent Cl transport was also lost in these ANO1-knockout animals [31] (Figure 1). In both the airways and the intestine, we found that the expression of CFTR in the apical membrane was largely attenuated, if not abolished. It should be noted that the expression of ANO1 in mouse airways is very low, while clear expression of ANO1 is detected in colonic epithelial cells, mainly located in the basolateral membrane [69,70].
How are these findings explained? From earlier studies, we know that ANO1 tethers the endoplasmic reticulum (ER) near the plasma membrane (PM) via binding to the inositol trisphosphate receptor (IP3R). Due to this, IP3-mediated Ca2+-release from the ER and store-operated Ca2+ influx are strongly improved in airway sub-apical or colonic sub-basolateral membrane compartments (please note that in the large intestine, ANO1 is located primarily near or in the basolateral membrane [71,72,73]). Along this line, it is of note that extended synaptotagmin-1 (ESYT1), another ER-PM tether, was found to further enhance PM expression of ANO1 and Ca2+ signaling [74]. PM expression of CFTR requires exocytosis, which is enhanced by the higher local sub-membranous Ca2+ levels facilitated by ANO1 [75]. Moreover, exocytosis also depends on ANO6 [75,76]. In knockout mice for ANO1 and ANO6 and in a number of human cell lines, we and others showed that both ANO1 and ANO6 are important for the PM insertion and activation of CFTR [52,53,75,76,77,78,79,80]. In this context, Ca2+-dependent activation of PKC could play a role [24,81,82]. Enhanced sub-membranous Ca2+ may further support CFTR activity via Ca2+- activated adenylate cyclases and EPAC. Finally, both anoctamins are equally important for mucus secretion by goblet cells and the release of lysozyme and other antimicrobial factors by Paneth cells. We speculate that ANO1 facilitates local Ca2+ signaling and not, or at least not primarily, Cl secretion. Evidence for this will be provided in the next chapter outlining data obtained from the first two patients that lacked expression of functional ANO1.

6. A Loss of Function Mutation of ANO1 in Patients Also Abolished CFTR-Mediated Cl Transport

The first two patients expressing the ANO1-variant c.897 + 3_897 + 6delAAGT were reported recently. These patients expressed a dysfunctional ANO1 and lack of Ca2+-activated Cl currents [83]. The two reported siblings presented in early infancy with reduced intestinal peristalsis and recurrent episodes of hemorrhagic diarrhea. Analysis of isolated primary airway epithelial cells obtained from one of the patients reproduced the results obtained earlier in tissue-specific ANO1-knockout mice [31]. Apart from the absence of Ca2+-activated Cl transport, CFTR Cl currents were also completely absent, possibly due to a lack of expression of CFTR in the apical membrane. Moreover, analysis of cells obtained from a heterozygous sibling showed reduced Cl secretion [83]. Rather surprisingly, the patients did not show a CF-like lung phenotype, although sweat tests were positive, indicating defective CFTR Cl conductance. This is even more surprising given the fact that both Ca2+-activated ANO1 and cAMP-activated CFTR Cl conductances were absent. Cytokine levels measured in sputum samples obtained from one of the ANO1 patients were largely reduced when compared to the cytokine levels measured in samples from two CF patients (Figure 2). While this may provide further evidence for the pro-inflammatory role of ANO1 [57], it also raises questions regarding the true contribution of apical Cl conductance for CF pathology [54].

7. Contribution of CFTR and ANO1/ANO6 to Regulated Cell Death

In previous chapters, we reported the functional relationship between CFTR and ANO1. Here, we will elaborate on the relationship between CFTR and ANO6, a Ca2+-activated phospholipid scramblase that is also permeable to ions [85]. Prior to the era of CFTR, the so-called intermediate conductance outwardly rectifying Cl channel ICOR (ORCC, ORDIC) was shown in many reports to be the essential epithelial secretory Cl channel that is absent in cystic fibrosis [86,87,88,89,90,91,92], while others identified the ICOR as a patch clamp artifact occurring during membrane excision [93]. After its identification in 1989 [94,95,96], CFTR was considered to be the real secretory Cl channel [81], and later it became clear that there was a regulatory relationship between CFTR and ICOR/ORCC [97,98]. ORCC has since been described as an apoptosis-related Cl channel, but its molecular identity still remains unclear [99]. After the identification of the anoctamin family, we reported CFTR as being an activator of ANO6, and we demonstrated that ANO6 is a core component of ICOR involved in apoptotic cell death [85,100]. Thus, the phospholipid scramblase ANO6 found a role in CFTR-dependent cell death and a place within the pathogenic relationships of CFTR with other proteins [101].
CFTR has been proposed to release glutathione (GSH) from airway epithelial cells to be enriched in the apical airway surface liquid, which will neutralize reactive oxygen species (ROS) [102,103,104]. Apparently, GSH efflux does not change cytosolic GSH content [105], and we were therefore unable to detect different ROS levels depending on the expression of CFTR [106]. However, we observed an enhanced activity of ANO6 in the presence of wtCFTR. As in most other cell types, ANO6 is also expressed in airway epithelial cells, where it can scramble plasma membrane phospholipids, which leads to cell death [85]. The importance of ANO6 for regulated cell death is also demonstrated in ANO6 knockout mice. In these animals, the number of apoptotic cells within the intestinal epithelium was drastically reduced [106].
In vivo inoculation with P. aeruginosa or Staphylococcus aureus induced lipid peroxidation in the lungs of CFTR-knockout mice and in wild-type animals. Exposure of human airway epithelial cells to P. aeruginosa induced an increase in reactive oxygen species (ROS) and caused lipid peroxidation and cell death. P. aeruginosa-induced cell death was independent of expression of wt-CFTR or F508del-CFTR [107]. In contrast, knockout of ANO1 clearly reduced cell death, possibly because ANO1 supports Ca2+-dependent activation of ANO6 and thus phospholipid scrambling [108]. Notwithstanding these results, good evidence exists for enhanced oxidative stress in the lungs of people with CF, while ROS have been shown to directly activate ANO1 and ANO6, providing another functional link between a lack of CFTR function, anoctamins and cell death [44,109,110].

8. Airway Secretion of Bicarbonate (HCO3) by CFTR

Disruption of CFTR in mice causes organ diseases typical of cystic fibrosis (CF), such as meconium ileus, distal intestinal obstructions with mucus accumulation, blockage of pancreatic ducts and lacrimal gland dilatation, along with some developmental defects [111]. These initial studies were confirmed in a number of subsequent transgenic models for cystic fibrosis [112,113]. However, a central aspect of CF pathology, namely the chronic inflammatory airway disease, was hardly detectable in CF mice, and pH values in the airway surface liquid were not different [113]. However, in a number of studies with a CF pig model, the human CF pathology could be nicely reproduced [114,115]. In contrast to transgenic F508del-cftr mice, CF pigs demonstrated reduced airway surface pH, impaired bacterial killing and adhesive mucus that disrupts mucociliary transport [116,117,118,119]. It was concluded that dysfunctional CFTR leads to a lack of HCO3 secretion, thus causing acidification of the airway surface liquid (ASL), followed by mucus abnormalities, attenuation of airway defenses, inflammation and a typical CF lung phenotype [118,119,120]. However, in another porcine CFTR-knockout model, acidic ASL pH could not be detected [12,121]. In this study, micro pH-electrode measurements were used to assess ASL pH directly in the small airways of lung sections from acutely sacrificed newborn piglets [12]. Pathological changes in these CFTR−/− lungs were not detected. Along this line, another study reported mucus accumulation preceding pulmonary infection in children with CF [122]. Moreover, using a novel luminescent technology integrated with fiberoptic probes, an acidic airway surface liquid pH could not be detected in children with cystic fibrosis [123]. Taken together, these findings may provoke the question as to whether HCO3 does also use another secretory pathway that is different to CFTR.

9. SLC26A9 Is Expressed in the Apical Membrane of Airways from CFTR-Knockout Piglets, but Not in Airways Expressing CFTR-F508del

SLC26A9 is one out of eleven proteins of the SLC26A family of anion transporters. It is expressed in the gastrointestinal tract, the respiratory system, male tissues and skin, and may have different functions depending on the organ in which it is expressed. Variants of SLC26A9 were associated with an increased incidence of meconium ileus and diabetes in patients with cystic fibrosis (CF) [124,125,126,127,128,129]. While there is good evidence for the role of SLC26A9 in gastrointestinal transport, it remains unclear whether it affects CF lung disease severity and airway responses to CFTR therapeutics [130,131]. Recently, the expression of the Cl/HCO3 exchanger SLC26A9 was found to be absent in the apical membrane of human F508del-CFTR/F508del-CFTR airways [132], which corresponds to the well-known inhibitory effect of F508del-CFTR on the membrane expression of SLC26A9 [16,21]. In contrast, SLC26A9 was found to be well expressed in the apical membrane of airway epithelial cells in non-CF lungs and in lungs from CFTR-knockout piglets [132]. Plasma membrane expression of SLC26A9 in the absence of CFTR was also shown in cell cultures, whereas coexpression with F508del-CFTR abrogates the biosynthesis, trafficking and function of SLC26A9 [14,16,21]. We therefore speculate that the normal ASL pH measured in airways of CFTR−/− piglets is due to the normal location and function of SLC26A9, which suggests that HCO3 can be secreted by SLC26A9 to the luminal side of the airways.
The transport of HCO3 by SLC26A9 has been proposed in some studies [133,134,135], but was not found by other laboratories [13,136,137,138]. Additional species-specific tissue factors like epithelial polarization or coexpression with additional proteins like CFTR may affect SLC26A9 transport function. Along this line, the vast majority of SLC26A9 expressed in non-polarized cells remains in the cytosol, while it is nicely expressed in the apical membrane of polarized cells [23]. Coexpression with wtCFTR or the complete absence of CFTR (CFTR knockout piglets) allows the proper plasma membrane location of SLC26A9, in contrast to the airways of CF-patients expressing a F508del-CFTR allele [132]. Like other SLC26A proteins (SLC26A3,4,6,8), SLC26A9 may also interact physically with CFTR via R (regulatory) and STAS (Sulphate Transporter and AntiSigma factor antagonist) domains, and probably through PDZ-domain interaction [13,29,139,140,141,142,143]. A recent study showed a contribution of SLC26A9 to airway bicarbonate secretion using the novel SLC26A9 inhibitor S9-A13. Online recordings of ASL pH in primary human nasal epithelial cells under thin film conditions indicated a sustained decrease in ASL pH caused by S9-A13, while subsequent activation of CFTR was unable to re-alkalinize ASL pH [23]. These initial results require confirmation by additional studies in vivo to clearly define the role of SLC26A9-dependent bicarbonate transport in airways (Figure 3). It will also be interesting to learn to what extent SLC26A9 contributes to airway HCO3 secretion when compared to SLC26A4 (pendrin), which probably secretes the most HCO3, particularly during inflammation [28]. In conclusion, the main task of SLC26A9 in the airways and particularly in the alveoli could actually be the reabsorption of Cl rather than Cl secretion (reviewed in [144]).

10. Bicarbonate Is Secreted in Renal Collecting Ducts, Which Requires CFTR, Pendrin and Possibly ANO1

CFTR is also expressed in the tubular epithelial cells of the human kidneys, where it affects different transport functions. Early studies suggested a role of CFTR for renal bicarbonate (HCO3) transport [145,146], which was later confirmed for many other epithelial organs [147,148,149]. HCO3 excretion was found to be largely reduced in people with CF, particularly when patients were challenged with the hormone secretin, which binds to its receptor and increases intracellular cAMP. A defect in renal bicarbonate excretion can lead to metabolic alkalosis occasionally observed in CF patients. Detailed studies in mice lacking expression of CFTR or the HCO3 transporter SLC26A4 (pendrin) finally uncovered the molecular mechanism [150,151]. Physical interaction of CFTR with pendrin and/or Cl recycling via CFTR drives the tubular release of HCO3 through apical pendrin and urinary excretion. This process takes place in ß-intercalated cells of the renal collecting duct, which coexpress CFTR, pendrin and receptors for secretin [150]. ANO1 is colocalized together with pendrin (and CFTR) in the apical membrane of renal ß-intercalated cells and may support the activity of CFTR [150] (Figure 4).

11. CFTR and ANO1 in Polycystic Kidney Disease: Which One Counts?

While CF and CF-associated metabolic alkalosis are rare, autosomal dominant polycystic kidney disease (ADPKD) is the most common monogenic kidney disease, affecting approximately 1 in 1000 individuals, often resulting in end-stage renal disease [152]. Mutations in either the PKD1 (~78%) or PKD2 (~15%) gene [153] cause the formation of multiple renal cysts which originate from renal tubule epithelial cells, predominantly the principal cells of the collecting duct [154,155]. The cysts grow continuously over years and cause compression of the adjacent intact nephrons, resulting in a decline of renal function [156]. Two key features are identified for cyst growth: a change from an absorptive to a secretory epithelium and the abnormal proliferation of cyst epithelial cells [157]. It is assumed that the major secretory force for cyst fluid secretion is apical cAMP-dependent Cl secretion, and several studies have suggested CFTR as the essential Cl channel [158,159,160].
However, recently, Cabrita et al. demonstrated that cyst growth in ADPKD is prevented by pharmacological and genetic inhibition of the calcium-activated chloride channel ANO1 [161]. Loss of PKD1 increased the expression of ANO1 and CFTR and induced Cl secretion in murine kidneys. Importantly, upregulated ANO1 enhanced intracellular Ca2+ signaling and the proliferation of PKD1-deficient renal epithelial cells. In contrast, increases in Ca2+ signaling, cell proliferation and CFTR expression were not observed in PKD1/ANO1 double knockout mice. In a sophisticated renal collecting duct M1 cell organoid model and in primary renal epithelial cells, cell proliferation and Cl secretion were also dependent on enhanced expression of ANO1 [162,163]. Knockdown of PKD1 or PKD2 increased basal intracellular Ca2+ levels and enhanced purinergic Ca2+ release from the endoplasmic reticulum. Ca2+ signals, proliferation, and Cl secretion were largely reduced via the knockdown or blockade of ANO1. ANO1 is therefore central to enhanced Ca2+ release from IP3-sensitive ER Ca2+ stores, and is a central player in ADPKD caused by mutations in PKD1 and PKD2. The data strongly suggest that pharmacological inhibition of ANO1 slows down the progression of ADPKD.
Concerning disease progression, male gender is a major risk factor [164,165]. Talbi et al. found that kidneys from PKD1 knockout mice had a more pronounced phenotype in males compared to females. The proliferation of cells from the cyst epithelium was enhanced in male when compared to female kidneys. This was paralleled by higher basal intracellular Ca2+ concentrations in cells isolated from PKD1 knockout males. These results again suggest enhanced intracellular Ca2+ levels contributing to enhanced proliferation and cyst development in male kidneys. Notably, the incubation of renal cells with dihydrotestosterone enhanced basal Ca2+ levels and ATP-stimulated ANO1 currents [166]. Similar results were obtained in a mouse model for autosomal recessive polycystic kidney disease (ARPKD) [157,167]. Finally, polycystic kidneys are under constant oxidative stress, which causes lipid peroxidation and the activation of ANO1 and ANO6 [44,110,168]. Therefore, inhibition of anoctamins may be a new avenue for therapeutic intervention in ADPKD.

12. Targeting ANO1 or CFTR in ADPKD?

Inhibitors of Cl currents such as diphenylamine-2-carboxylate and knockdown of CFTR by antisense oligo-nucleotides inhibited cAMP-activated Cl currents in cyst cells [169]; 8807590). CFTRinh-172 or Ph-GlyH-101 reduced the cyst growth of renal MDCK cells in a metanephric mouse kidney model and a rapidly progressive neonatal Pkd1 knockout mouse model [159,160]. In three CF patients with concomitant ADPKD, disease progression was delayed when compared to their siblings without CF [170,171]. However, the ADPKD-protective effect provided by CF was not confirmed in a subsequent report [172], and CFTR expression in isolated ADPKD cyst cells was shown to be very heterogeneous [158,169,173]. It is therefore not entirely clear whether inhibiting CFTR slows down cyst progression. Moreover, both CFTRinh-172 or Ph-GlyH-101 have pronounced off-target effects and affect intracellular Ca2+ signals which actually inhibit ANO1 [18].
Initial studies showed that Ca2+-activated ANO1 Cl currents contribute to cyst growth [174]. ATP is released by cyst cells, accumulates in the cyst lumen and activates ANO1 via the stimulation of purinergic receptors [155,174,175]. By contrast, the scavenging of ATP by apyrase, the P2Y2 receptor antagonist suramin, and the knockdown of P2Y2 inhibited cyst growth [155]. These studies and the subsequent work outlined above [161,162] suggested ANO1 as the relevant pharmacological target to inhibit in ADPKD.
Taken together, in mouse studies, ANO1 is a dominant driver of secretion-dependent cyst enlargement, while we also found that knockout of CFTR had no significant impact on cyst growth [176] (Figure 5). Nevertheless, it is important to keep in mind that the physiological contribution of ANO1 in mice is probably greater while the contribution of CFTR is lower than in humans. In mice, CFTR shows lower activity in the airways but has a more pronounced contribution to intestinal transport [177]. In the kidneys of healthy mice, CFTR is only clearly expressed in ß-intercalated cells, where it controls HCO3 secretion [150]. Interestingly, a recent report showed that the application of the CFTR-corrector VX-809 (Lumacaftor) in a Pkd1 knockout and the Pkd1RC/RC mouse model reduced cyst growth [178,179]. These findings were explained by a cellular translocalization of CFTR and the Na+/H+ exchanger 3. A clinical phase 2 placebo-controlled randomized trial investigated the efficacy and safety of the CFTR corrector GLPG2737 in ADPKD patients (NCT04578548) [180]. More studies are required to analyse the contribution of ANO1 to cyst formation in human tissue. The central aspects of ANO1 include its obvious pro-proliferative and de-differentiating properties [101], which after all may have a larger impact on cyst progression than fluid secretion.

13. Inhibitors of ANO1

Given the promising results obtained through the genetic and pharmacological inhibition of ANO1, ANO1 qualifies as a potential target for the treatment of ADPKD. ANO1 function can be addressed by drugs that have already been approved for other indications, like niclosamide or benzbromarone [161]. Niclosamide is an essential oral anthelminthic drug used for decades to treat parasitic infections, but it is meant for short-term use [181]. Benzbromarone is a uricosuric drug that has been used in the treatment of gout over the last 30 years. Although withdrawn by Sanofi for safety reasons after reports of hepatotoxicity, it is still marketed in several countries by other drug companies (drugs.com/international/benzbromarone.html (accessed on 27 June 2023)). Hepatotoxicity is rare and occurs in 1 in 17,000 patients [182]. For comparison, the only drug approved for treatment of ADPKD, the vasopressin-2-receptor antagonist tolvaptan, has a hepatotoxic risk of 1 in 3000. Therefore, many experts in the field have questioned the withdrawal of benzbromarone [182].

14. Conclusions

This review summarizes recent findings on CFTR, SLC26A9, and anoctamin 1 and 6 in the airways and kidneys. It becomes clear that these ion channels and transports do not stand alone, but rather operate in a functional and metabolic network and should therefore be analyzed in the context of their molecular and functional interactions. However, data obtained in different animal and cell culture models often cause confusion due to the differential expression of proteins in the different species, tissues or cell lines. This is particularly evident for the differences in CFTR expression in human, piglets and mice, and even more so for ANO1, which is almost absent in native airways but abundant in cultured airway cells. The authors believe that progress in understanding cystic fibrosis and polycystic kidney disease can only be achieved by considering data from animal studies.
The current trend towards over-regulated and lengthy animal welfare applications should be corrected. Pure cell culture studies, even when performed on primary cells in differentiated culture, carry the risk of misinterpretation. This is particularly important in the development of new pharmacological strategies. Because of the increasingly complex pathogenic relationships, new therapeutic strategies must first be thoroughly evaluated in animal studies before they can be applied in humans.

Authors Contributions

Conceptualization, K.K., J.O., A.K., R.S. and B.B.; methodology., J.O., A.K., J.H.P., T.M., R.S. and B.B.; formal analysis, J.O., A.K. and B.B.; writing—original draft preparation, K.K., R.S. and B.B.; writing—review and editing, J.O., A.K., J.H.P., T.M., R.S. and B.B. All authors have read and agreed to the published version of the manuscript.

Funding

Supported by DFG Transregio-SFB, Project-ID 509149993, TRR 374 (Project A3).

Institutional Review Board Statement

All animal experiments complied with the general guidelines for animal research, in accordance with the United Kingdom Animals Act, 1986, and associated guidelines, and EU Directive 2010/63/EU for animal experiments. All animal experiments were approved by the local Ethics Committee of the Government of Unterfranken/Wurzburg/Germany (AZ: 55.2-2532-2-677) and were conducted according to the guidelines of the American Physiologic Society and German Law for the Welfare of Animals.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kunzelmann, K. CFTR: Interacting with everything? News Physiol. Sci. 2001, 17, 167–170. [Google Scholar] [CrossRef]
  2. Wang, X.; Venable, J.; LaPointe, P.; Hutt, D.M.; Koulov, A.V.; Coppinger, J.; Gurkan, C.; Kellner, W.; Matteson, J.; Plutner, H.; et al. Hsp90 cochaperone Aha1 downregulation rescues misfolding of CFTR in cystic fibrosis. Cell 2006, 127, 803–815. [Google Scholar] [CrossRef]
  3. Pankow, S.; Bamberger, C.; Calzolari, D.; Martínez-Bartolomé, S.; Lavallée-Adam, M.; Balch, W.E.; Yates, J.R., 3rd. ∆F508 CFTR interactome remodelling promotes rescue of cystic fibrosis. Nature 2015, 528, 510–516. [Google Scholar] [CrossRef] [PubMed]
  4. Lim, S.H.; Snider, J.; Birimberg-Schwartz, L.; Ip, W.; Serralha, J.C.; Botelho, H.M.; Lopes-Pacheco, M.; Pinto, M.C.; Moutaoufik, M.T.; Zilocchi, M.; et al. CFTR interactome mapping using the mammalian membrane two-hybrid high-throughput screening system. Mol. Syst. Biol. 2022, 18, e10629. [Google Scholar] [CrossRef]
  5. Li, C.; Naren, A.P. Macromolecular complexes of cystic fibrosis transmembrane conductance regulator and its interacting partners. Pharmacol. Ther. 2005, 108, 208–223. [Google Scholar] [CrossRef]
  6. Roxo-Rosa, M.; Davezac, N.; Bensalem, N.; Majumder, M.; Heda, G.D.; Simas, A.; Penque, D.; Amaral, M.D.; Lukacs, G.L.; Edelman, A. Proteomics techniques for cystic fibrosis research. J. Cyst. Fibros. 2004, 3 (Suppl. S2), 85–89. [Google Scholar] [CrossRef]
  7. Kunzelmann, K.; Mehta, A. CFTR: A hub for kinases and cross-talk of cAMP and Ca. FEBS J. 2013, 280, 4417–4429. [Google Scholar] [CrossRef] [PubMed]
  8. Guggino, W.B.; Banks-Schlegel, S.P. Macromolecular Interactions and Ion Transport in Cystic Fibrosis. Am. J. Respir. Crit. Care Med. 2004, 170, 815–820. [Google Scholar] [CrossRef]
  9. Ishibashi, K.; Okamura, K.; Yamazaki, J. Involvement of apical P2Y2 receptor-regulated CFTR activity in muscarinic stimulation of Cl(−) reabsorption in rat submandibular gland. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2008, 294, R1729–R1736. [Google Scholar] [CrossRef] [PubMed]
  10. Wang, X.; Lytle, C.; Quinton, P.M. Predominant constitutive CFTR conductance in small airways. Respir. Res. 2005, 6, 7. [Google Scholar] [CrossRef] [PubMed]
  11. Song, Y.; Salinas, D.; Nielson, D.W.; Verkman, A.S. Hyperacidity of secreted fluid from submucosal glands in early cystic fibrosis. Am. J. Physiol. Cell Physiol. 2006, 290, C741–C749. [Google Scholar] [CrossRef] [PubMed]
  12. Benedetto, R.; Centeio, R.; Ousingsawat, J.; Schreiber, R.; Janda, M.; Kunzelmann, K. Transport properties in CFTR−/− knockout piglets suggest normal airway surface liquid pH and enhanced amiloride-sensitive Na(+) absorption. Pflug. Arch. 2020, 472, 1507–1519. [Google Scholar] [CrossRef] [PubMed]
  13. Bertrand, C.A.; Zhang, R.; Pilewski, J.M.; Frizzell, R.A. SLC26A9 is a constitutively active, CFTR-regulated anion conductance in human bronchial epithelia. J. Gen. Physiol. 2009, 133, 421–438. [Google Scholar] [CrossRef] [PubMed]
  14. Ousingsawat, J.; Schreiber, R.; Kunzelmann, K. Differential contribution of SLC26A9 to Cl(−) conductance in polarized and non-polarized epithelial cells. J. Cell Physiol. 2011, 227, 2323–2329. [Google Scholar] [CrossRef]
  15. Salomon, J.J.; Spahn, S.; Wang, X.; Fullekrug, J.; Bertrand, C.A.; Mall, M.A. Generation and functional characterization of epithelial cells with stable expression of SLC26A9 Cl channels. Am. J. Physiol. Lung Cell Mol. Physiol. 2016, 310, L593–L602. [Google Scholar] [CrossRef]
  16. Bertrand, C.A.; Mitra, S.; Mishra, S.K.; Wang, X.; Zhao, Y.; Pilewski, J.M.; Madden, D.R.; Frizzell, R.A. The CFTR trafficking mutation F508del inhibits the constitutive activity of SLC26A9. Am. J. Physiol. Lung Cell Mol. Physiol. 2017, 312, L912–L925. [Google Scholar] [CrossRef]
  17. Larsen, M.B.; Choi, J.J.; Wang, X.; Myerburg, M.M.; Frizzell, R.A.; Bertrand, C.A. Separating the contributions of SLC26A9 and CFTR to anion secretion in primary human bronchial epithelia (HBE). Am. J. Physiol. Lung Cell Mol. Physiol. 2021, 321, L1147–L1160. [Google Scholar] [CrossRef] [PubMed]
  18. Lin, J.; Gettings, S.M.; Talbi, K.; Schreiber, R.; Taggart, M.J.; Preller, M.; Kunzelmann, K.; Althaus, M.; Gray, M.A. Pharmacological inhibitors of the cystic fibrosis transmembrane conductance regulator exert off-target effects on epithelial cation channels. Pflug. Arch. 2023, 475, 167–179. [Google Scholar] [CrossRef] [PubMed]
  19. Sato, Y.; Thomas, D.Y.; Hanrahan, J.W. The anion transporter SLC26A9 localizes to tight junctions and is degraded by the proteasome when co-expressed with F508del-CFTR. J. Biol. Chem. 2019, 294, 18269–18284. [Google Scholar] [CrossRef]
  20. Pinto, M.C.; Quaresma, M.C.; Silva, I.A.L.; Railean, V.; Ramalho, S.S.; Amaral, M.D. Synergy in Cystic Fibrosis Therapies: Targeting SLC26A9. Int. J. Mol. Sci. 2021, 22, 13064. [Google Scholar] [CrossRef]
  21. Needham, P.G.; Goeckeler-Fried, J.L.; Zhang, C.; Sun, Z.; Wetzel, A.R.; Bertrand, C.A.; Brodsky, J.L. SLC26A9 is selected for endoplasmic reticulum associated degradation (ERAD) via Hsp70-dependent targeting of the soluble STAS domain. Biochem. J. 2021, 478, 4203–4220. [Google Scholar] [CrossRef]
  22. Avella, M.; Loriol, C.; Boulukos, K.; Borgese, F.; Ehrenfeld, J. SLC26A9 stimulates CFTR expression and function in human bronchial cell lines. J. Cell Physiol. 2011, 226, 212–223. [Google Scholar] [CrossRef] [PubMed]
  23. Jo, S.; Centeio, R.; Park, J.; Ousingsawat, J.; Jeon, D.K.; Talbi, K.; Schreiber, R.; Ryu, K.; Kahlenberg, K.; Somoza, V.; et al. The SLC26A9 inhibitor S9-A13 provides no evidence for a role of SLC26A9 in airway chloride secretion but suggests a contribution to regulation of ASL pH and gastric proton secretion. FASEB J. 2022, 36, e22534. [Google Scholar] [CrossRef]
  24. Palmer, M.L.; Lee, S.Y.; Carlson, D.; Fahrenkrug, S.; O’Grady, S.M. Stable knockdown of CFTR establishes a role for the channel in P2Y receptor-stimulated anion secretion. J. Cell Physiol. 2006, 206, 759–770. [Google Scholar] [CrossRef] [PubMed]
  25. Chappe, V.; Hinkson, D.A.; Howell, L.D.; Evagelidis, A.; Liao, J.; Chang, X.B.; Riordan, J.R.; Hanrahan, J.W. Stimulatory and inhibitory protein kinase C consensus sequences regulate the cystic fibrosis transmembrane conductance regulator. Proc. Natl. Acad. Sci. USA 2004, 101, 390–395. [Google Scholar] [CrossRef]
  26. Jia, Y.; Mathews, C.J.; Hanrahan, J.W. Phosphorylation by protein kinase C is required for acute activation of cystic fibrosis transmembrane conductance regulator by protein kinase A. J. Biol. Chem. 1997, 272, 4978–4984. [Google Scholar] [CrossRef]
  27. Winpenny, J.P.; McAlroy, H.L.; Gray, M.A.; Argent, B.E. Protein kinase C regulates the magnitude and stability of CFTR currents in pancreatic duct cells. Am. J. Physiol. 1995, 268, C823–C828. [Google Scholar] [CrossRef] [PubMed]
  28. Kim, D.; Huang, J.; Billet, A.; Abu-Arish, A.; Goepp, J.; Matthes, E.; Tewfik, M.A.; Frenkiel, S.; Hanrahan, J.W. Pendrin Mediates Bicarbonate Secretion and Enhances Cystic Fibrosis Transmembrane Conductance Regulator Function in Airway Surface Epithelia. Am. J. Respir. Cell Mol. Biol. 2019, 60, 705–716. [Google Scholar] [CrossRef] [PubMed]
  29. Ko, S.B.; Zeng, W.; Dorwart, M.R.; Luo, X.; Kim, K.H.; Millen, L.; Goto, H.; Naruse, S.; Soyombo, A.; Thomas, P.J.; et al. Gating of CFTR by the STAS domain of SLC26 transporters. Nat. Cell Biol. 2004, 6, 343–350. [Google Scholar] [CrossRef] [PubMed]
  30. Anderson, M.P.; Welsh, M.J. Calcium and cAMP activate different chloride channels in the apical membrane of normal and cystic fibrosis epithelia. Proc. Natl. Acad. Sci. USA 1991, 88, 6003–6007. [Google Scholar] [CrossRef]
  31. Benedetto, R.; Ousingsawat, J.; Wanitchakool, P.; Zhang, Y.; Holtzman, M.J.; Amaral, M.; Rock, J.R.; Schreiber, R.; Kunzelmann, K. Epithelial Chloride Transport by CFTR Requires TMEM16A. Sci. Rep. 2017, 7, 12397. [Google Scholar] [CrossRef] [PubMed]
  32. Lerias, J.; Pinto, M.; Benedetto, R.; Schreiber, R.; Amaral, M.; Aureli, M.; Kunzelmann, K. Compartmentalized crosstalk of CFTR and TMEM16A (ANO1) through EPAC1 and ADCY1. Cell Signal 2018, 44, 10–19. [Google Scholar] [CrossRef]
  33. Kunzelmann, K.; Mall, M.; Briel, M.; Hipper, A.; Nitschke, R.; Ricken, S.; Greger, R. The cystic fibrosis transmembrane conductance regulator attenuates the endogenous Ca2+ activated Cl conductance in Xenopus ooyctes. Pflüg. Arch. 1997, 434, 178–181. [Google Scholar] [CrossRef]
  34. Wei, L.; Vankeerberghen, A.; Cuppens, H.; Eggermont, J.; Cassiman, J.J.; Droogmans, G.; Nilius, B. Interaction between calcium-activated chloride channels and the cystic fibrosis transmembrane conductance regulator. Pflug. Arch. 1999, 438, 635–641. [Google Scholar] [CrossRef]
  35. Wei, L.; Vankeerberghen, A.; Cuppens, H.; Cassiman, J.J.; Droogmans, G.; Nilius, B. The C-terminal part of the R-domain, but not the PDZ binding motif, of CFTR is involved in interaction with Ca2+ -activated Cl channels. Pflüg. Arch. 2001, 442, 280–285. [Google Scholar] [CrossRef]
  36. Perez-Cornejo, P.; Arreola, J. Regulation of Ca(2+)-activated chloride channels by cAMP and CFTR in parotid acinar cells. Biochem. Biophys. Res. Commun. 2004, 316, 612–617. [Google Scholar] [CrossRef]
  37. Ousingsawat, J.; Kongsuphol, P.; Schreiber, R.; Kunzelmann, K. CFTR and TMEM16A are Separate but Functionally Related Cl Channels. Cell Physiol. Biochem. 2011, 28, 715–724. [Google Scholar] [CrossRef]
  38. Kunzelmann, K.; Tian, Y.; Martins, J.R.; Faria, D.; Kongsuphol, P.; Ousingsawat, J.; Wolf, L.; Schreiber, R. Cells in focus: Airway epithelial cells-Functional links between CFTR and anoctamin dependent Cl() secretion. Int. J. Biochem. Cell Biol. 2012, 44, 1897–1900. [Google Scholar] [CrossRef] [PubMed]
  39. Bozoky, Z.; Ahmadi, S.; Milman, T.; Kim, T.H.; Du, K.; Di Paola, M.; Pasyk, S.; Pekhletski, R.; Keller, J.P.; Bear, C.E.; et al. Synergy of cAMP and calcium signaling pathways in CFTR regulation. Proc. Natl. Acad. Sci. USA 2017, 114, E2086–E2095. [Google Scholar] [CrossRef]
  40. Namkung, W.; Finkbeiner, W.E.; Verkman, A.S. CFTR-Adenylyl Cyclase I Association Is Responsible for UTP Activation of CFTR in Well-Differentiated Primary Human Bronchial Cell Cultures. Mol. Biol. Cell 2010, 21, 2639–2648. [Google Scholar] [CrossRef]
  41. Billet, A.; Hanrahan, J.W. The secret life of CFTR as a calcium-activated chloride channel. J. Physiol. 2013, 591, 5273–5278. [Google Scholar] [CrossRef]
  42. Le, S.C.; Jia, Z.; Chen, J.; Yang, H. Molecular basis of PIP2-dependent regulation of the Ca(2+)-activated chloride channel TMEM16A. Nat. Commun. 2019, 10, 3769. [Google Scholar] [CrossRef]
  43. Ko, W.; Jung, S.R.; Kim, K.W.; Yeon, J.H.; Park, C.G.; Nam, J.H.; Hille, B.; Suh, B.C. Allosteric modulation of alternatively spliced Ca(2+)-activated Cl() channels TMEM16A by PI(4,5)P(2) and CaMKII. Proc. Natl. Acad. Sci. USA 2020, 117, 30787–30798. [Google Scholar] [CrossRef]
  44. Schreiber, R.; Ousingsawat, J.; Wanitchakool, P.; Sirianant, L.; Benedetto, R.; Reiss, K.; Kunzelmann, K. Regulation of TMEM16A/ANO1 and TMEM16F/ANO6 ion currents and phospholipid scrambling by Ca2+ and plasma membrane lipid. J. Physiol. 2018, 596, 217–229. [Google Scholar] [CrossRef]
  45. Talbi, K.; Ousingsawat, J.; Centeio, R.; Schreiber, R.; Kunzelmann, K. Calmodulin-Dependent Regulation of Overexpressed but Not Endogenous TMEM16A Expressed in Airway Epithelial Cells. Membranes 2021, 11, 723. [Google Scholar] [CrossRef]
  46. Mall, M.; Gonska, T.; Thomas, J.; Schreiber, R.; Seydewitz, H.H.; Kuehr, J.; Brandis, M.; Kunzelmann, K. Modulation of Ca2+ activated Cl secretion by basolateral K+ channels in human normal and cystic fibrosis airway epithelia. Pediatr. Res. 2003, 53, 608–618. [Google Scholar] [CrossRef]
  47. Mall, M.; Bleich, M.; Greger, R.; Schürlein, M.; Kühr, J.; Seydewitz, H.H.; Brandis, M.; Kunzelmann, K. Cholinergic ion secretion in human colon requires co-activation by cAMP. Am. J. Physiol. 1998, 275, G1274–G1281. [Google Scholar]
  48. Faria, D.; Schreiber, R.; Kunzelmann, K. CFTR is activated through stimulation of purinergic P2Y2 receptors. Pflüg. Arch. 2009, 457, 1373–1380. [Google Scholar] [CrossRef]
  49. Tembo, M.; Wozniak, K.L.; Bainbridge, R.E.; Carlson, A.E. Phosphatidylinositol 4,5-bisphosphate (PIP2) and Ca(2+) are both required to open the Cl() channel TMEM16A. J. Biol. Chem. 2019, 294, 12556–12564. [Google Scholar] [CrossRef]
  50. Danahay, H.L.; Lilley, S.; Fox, R.; Charlton, H.; Sabater, J.; Button, B.; McCarthy, C.; Collingwood, S.P.; Gosling, M. TMEM16A Potentiation: A Novel Therapeutic Approach for the Treatment of Cystic Fibrosis. Am. J. Respir. Crit. Care Med. 2020, 201, 946–954. [Google Scholar] [CrossRef] [PubMed]
  51. Danahay, H.; Gosling, M. TMEM16A: An Alternative Approach to Restoring Airway Anion Secretion in Cystic Fibrosis? Int. J. Mol. Sci. 2020, 21, 2386. [Google Scholar] [CrossRef]
  52. Centeio, R.; Ousingsawat, J.; Cabrita, I.; Schreiber, R.; Talbi, K.; Benedetto, R.; Doušová, T.; Verbeken, E.K.; De Boeck, K.; Cohen, I.; et al. Mucus Release and Airway Constriction by TMEM16A May Worsen Pathology in Inflammatory Lung Disease. Int. J. Mol. Sci. 2021, 22, 7852. [Google Scholar] [CrossRef] [PubMed]
  53. Benedetto, R.; Cabrita, I.; Schreiber, R.; Kunzelmann, K. TMEM16A is indispensable for basal mucus secretion in airways and intestine. FASEB J. 2019, 33, 4502–4512. [Google Scholar] [CrossRef]
  54. Kunzelmann, K.; Ousingsawat, J.; Cabrita, I.; Doušová, T.; Bähr, A.; Janda, M.; Schreiber, R.; Benedetto, R. TMEM16A in Cystic Fibrosis: Activating or Inhibiting? Front. Pharmacol. 2019, 10, 3. [Google Scholar] [CrossRef]
  55. Bai, W.; Liu, M.; Xiao, Q. The diverse roles of TMEM16A Ca(2+)-activated Cl() channels in inflammation. J. Adv. Res. 2021, 33, 53–68. [Google Scholar] [CrossRef]
  56. Liu, B.; Linley, J.E.; Du, X.; Zhang, X.; Ooi, L.; Zhang, H.; Gamper, N. The acute nociceptive signals induced by bradykinin in rat sensory neurons are mediated by inhibition of M-type K+ channels and activation of Ca2+-activated Cl channels. J. Clin. Investig. 2010, 120, 1240–1252. [Google Scholar] [CrossRef]
  57. Lee, B.; Cho, H.; Jung, J.; Yang, Y.D.; Yang, D.J.; Oh, U. Anoctamin 1 contributes to inflammatory and nerve-injury induced hypersensitivity. Mol. Pain. 2014, 10, 1744–8069. [Google Scholar] [CrossRef]
  58. Leblanc, N.; Forrest, A.S.; Ayon, R.J.; Wiwchar, M.; Angermann, J.E.; Pritchard, H.A.; Singer, C.A.; Valencik, M.L.; Britton, F.; Greenwood, I.A. Molecular and functional significance of Ca(2+)-activated Cl(−) channels in pulmonary arterial smooth muscle. Pulm. Circ. 2015, 5, 244–268. [Google Scholar] [CrossRef]
  59. Zhang, C.H.; Li, Y.; Zhao, W.; Lifshitz, L.M.; Li, H.; Harfe, B.D.; Zhu, M.S.; ZhuGe, R. The transmembrane protein 16A Ca(2+)-activated Cl- channel in airway smooth muscle contributes to airway hyperresponsiveness. Am. J. Respir. Crit. Care Med. 2013, 187, 374–381. [Google Scholar] [CrossRef]
  60. Gallos, G.; Remy, K.E.; Danielsson, J.; Funayama, H.; Fu, X.W.; Chang, H.-Y.S.; Xu, D.; Emala, C.W.; Bradley, E.; Large, R.J.; et al. Functional expression of the TMEM16 family of calcium-activated chloride channels in airway smooth muscle. Am. J. Physiol. Cell. Mol. Physiol. 2013, 305, L625–L634. [Google Scholar] [CrossRef] [PubMed]
  61. Danielsson, J.; Perez-Zoghbi, J.; Bernstein, K.; Barajas, M.B.; Zhang, Y.; Kumar, S.; Sharma, P.K.; Gallos, G.; Emala, C.W. Antagonists of the TMEM16A Calcium-activated Chloride Channel Modulate Airway Smooth Muscle Tone and Intracellular Calcium. Anesthesiology 2015, 123, 569–581. [Google Scholar] [CrossRef] [PubMed]
  62. Kondo, M.; Tsuji, M.; Hara, K.; Arimura, K.; Yagi, O.; Tagaya, E.; Takeyama, K.; Tamaoki, J. Chloride ion transport and overexpression of TMEM16A in a guinea-pig asthma model. Clin. Exp. Allergy 2017, 47, 795–804. [Google Scholar] [CrossRef]
  63. Huang, F.; Zhang, H.; Wu, M.; Yang, H.; Kudo, M.; Peters, C.J.; Woodruff, P.G.; Solberg, O.D.; Donne, M.L.; Huang, X.; et al. Calcium-activated chloride channel TMEM16A modulates mucin secretion and airway smooth muscle contraction. Proc. Natl. Acad. Sci. USA 2012, 109, 16354–16359. [Google Scholar] [CrossRef]
  64. Hao, A.; Guo, S.; Shi, S.; Wang, X.; Zhan, Y.; Chen, Y.; An, H. Emerging Modulators of TMEM16A and Their Therapeutic Potential. J. Membr. Biol. 2021, 254, 353–365. [Google Scholar] [CrossRef] [PubMed]
  65. Miner, K.; Labitzke, K.; Liu, B.; Elliot, R.; Wang, P.; Henckels, K.; Gaida, K.; Elliot, R.; Chen, J.J.; Liu, L.; et al. Drug Repurposing: The Anthelmintics Niclosamide and Nitazoxanide Are Potent TMEM16A Antagonists That Fully Bronchodilate Airways. Front. Pharmacol. 2019, 10, 51. [Google Scholar] [CrossRef] [PubMed]
  66. Ruffin, M.; Voland, M.; Marie, S.; Bonora, M.; Blanchard, E.; Blouquit-Laye, S.; Naline, E.; Puyo, P.; Le Rouzic, P.; Guillot, L.; et al. Anoctamin 1 Dysregulation Alters Bronchial Epithelial Repair in Cystic Fibrosis. Biochim. Biophys. Acta 2013, 1832, 2340–2351. [Google Scholar] [CrossRef] [PubMed]
  67. Lobo, M.J.; Amaral, M.D.; Zaccolo, M.; Farinha, C.M. EPAC1 activation by cAMP stabilizes CFTR at the membrane by promoting its interaction with NHERF1. J. Cell Sci. 2016, 129, 2599–2612. [Google Scholar] [CrossRef]
  68. Murabito, A.; Bhatt, J.; Ghigo, A. It Takes Two to Tango! Protein–Protein Interactions behind cAMP-Mediated CFTR Regulation. Int. J. Mol. Sci. 2023, 24, 10538. [Google Scholar] [CrossRef]
  69. Schreiber, R.; Faria, D.; Skryabin, B.V.; Rock, J.R.; Kunzelmann, K. Anoctamins support calcium-dependent chloride secretion by facilitating calcium signaling in adult mouse intestine. Pflüg. Arch. 2015, 467, 1203–1213. [Google Scholar] [CrossRef]
  70. He, Q.; Halm, S.T.; Zhang, J.; Halm, D.R. Activation of the basolateral membrane Cl conductance essential for electrogenic K secretion suppresses electrogenic Cl secretion. Exp. Physiol. 2011, 96, 305–316. [Google Scholar] [CrossRef] [PubMed]
  71. Cabrita, I.; Benedetto, R.; Fonseca, A.; Wanitchakool, P.; Sirianant, L.; Skryabin, B.V.; Schenk, L.K.; Pavenstadt, H.; Schreiber, R.; Kunzelmann, K. Differential effects of anoctamins on intracellular calcium signals. FASEB J. 2017, 31, 2123–2134. [Google Scholar] [CrossRef]
  72. Kunzelmann, K.; Cabrita, I.; Wanitchakool, P.; Ousingsawat, J.; Sirianant, L.; Benedetto, R.; Schreiber, R. Modulating Ca2+ signals: A common theme for TMEM16, Ist2, and TMC. Pflüg. Arch. 2016, 468, 475–490. [Google Scholar] [CrossRef]
  73. Jin, X.; Shah, S.; Du, X.; Zhang, H.; Gamper, N. Activation of Ca2+-activated Cl channel ANO1 by localized Ca2+ signals. J. Physiol. 2016, 594, 19–30. [Google Scholar] [CrossRef]
  74. Lerias, J.R.; Pinto, M.C.; Botelho, H.M.; Awatade, N.T.; Quaresma, M.C.; Silva, I.A.L.; Wanitchakool, P.; Schreiber, R.; Pepperkok, R.; Kunzelmann, K.; et al. A novel microscopy-based assay identifies extended synaptotagmin-1 (ESYT1) as a positive regulator of anoctamin 1 traffic. Biochim. Biophys. Acta 2018, 1865, 421–431. [Google Scholar] [CrossRef]
  75. Centeio, R.; Cabrita, I.; Schreiber, R.; Kunzelmann, K. TMEM16A/F support exocytosis but do not inhibit Notch-mediated goblet cell metaplasia of BCi-NS1.1 human airway epithelium. Front. Physiol. 2023, 14, 1157704. [Google Scholar] [CrossRef] [PubMed]
  76. Bricogne, C.; Fine, M.; Pereira, P.M.; Sung, J.; Tijani, M.; Wang, Y.; Henriques, R.; Collins, M.K.; Hilgemann, D. TMEM16F activation by Ca(2+) triggers plasma membrane expansion and directs PD-1 trafficking. Sci. Rep. 2019, 9, 619. [Google Scholar] [CrossRef]
  77. Benedetto, R.; Ousingsawat, J.; Cabrita, I.; Pinto, M.; Lerias, J.; Wanitchakool, P.; Schreiber, R.; Kunzelmann, K. Plasma membrane localized TMEM16 Proteins are Indispensable for expression of CFTR. J. Mol. Med. 2019, 97, 711–722. [Google Scholar] [CrossRef]
  78. Cabrita, I.; Benedetto, R.; Schreiber, R.; Kunzelmann, K. Niclosamide repurposed for the treatment of inflammatory airway disease. JCI Insight 2019, 8, 128414. [Google Scholar] [CrossRef]
  79. Yimnual, C.; Satitsri, S.; Ningsih, B.N.S.; Rukachaisirikul, V.; Muanprasat, C. A fungus-derived purpactin A as an inhibitor of TMEM16A chloride channels and mucin secretion in airway epithelial cells. Biomed. Pharmacother. Biomed. Pharmacother. 2021, 139, 111583. [Google Scholar] [CrossRef] [PubMed]
  80. Centeio, R.; Ousingsawat, J.; Schreiber, R.; Kunzelmann, K. CLCA1 Regulates Airway Mucus Production and Ion Secretion through TMEM16A. Int. J. Mol. Sci. 2021, 22, 5133. [Google Scholar] [CrossRef] [PubMed]
  81. Chappe, V.; Hinkson, D.A.; Zhu, T.; Chang, X.B.; Riordan, J.R.; Hanrahan, J.W. Phosphorylation of protein kinase C sites in NBD1 and the R domain control CFTR channel activation by PKA. J. Physiol. 2003, 548, 39–52. [Google Scholar] [CrossRef]
  82. Okuda, K.; Shaffer, K.M.; Ehre, C. Mucins and CFTR: Their Close Relationship. Int. J. Mol. Sci. 2022, 23, 10232. [Google Scholar] [CrossRef]
  83. Park, J.H.; Ousingsawat, J.; Cabrita, I.; Bettels, R.E.; Große-Onnebrink, J.; Schmalstieg, C.; Biskup, S.; Reunert, J.; Rust, S.; Schreiber, R.; et al. TMEM16A deficiency: A potentially fatal neonatal disease resulting from impaired chloride currents. J. Med. Genet. 2020, 58, 247–253. [Google Scholar] [CrossRef] [PubMed]
  84. Sagel, S.D.; Chmiel, J.F.; Konstan, M.W. Sputum biomarkers of inflammation in cystic fibrosis lung disease. Proc. Am. Thorac. Soc. 2007, 4, 406–417. [Google Scholar] [CrossRef]
  85. Ousingsawat, J.; Wanitchakool, P.; Kmit, A.; Romao, A.M.; Jantarajit, W.; Schreiber, S.; Kunzelmann, K. Anoctamin 6 mediates effects essential for innate immunity downstream of P2X7-receptors in macrophages. Nat. Commun. 2015, 6, 6245. [Google Scholar] [CrossRef]
  86. Frizzell, R.A.; Rechkemmer, G.R.; Shoemaker, R.L. Altered regulation of airway epithelial cell chloride channels in cystic fibrosis. Science 1986, 233, 558–560. [Google Scholar] [CrossRef] [PubMed]
  87. Li, M.; McCann, J.D.; Liedtke, C.M.; Nairn, A.C.; Greengard, P.; Welsh, M.J. Cyclic AMP-dependent protein kinase opens chloride channels in normal but not cystic fibrosis airway epithelium. Nature 1988, 331, 358–360. [Google Scholar] [CrossRef]
  88. Schoumacher, R.A.; Shoemaker, R.L.; Halm, D.R.; Tallant, E.A.; Wallace, R.W.; Frizzell, R.A. Phosphorylation fails to activate chloride channels from cystic fibrosis airway cells. Nature 1987, 330, 752–754. [Google Scholar] [CrossRef] [PubMed]
  89. Hryciw, D.H.; Guggino, W.B. Cystic Fibrosis Transmembrane Conductance Regulator and the Outwardly Rectifying Chloride Channel: A Relationship between Two Chloride Channels Expressed In Epithelial Cells. Clin. Exp. Pharmacol. Physiol. 2000, 27, 892–895. [Google Scholar] [CrossRef]
  90. Schwiebert, E.M.; Benos, D.J.; Egan, M.E.; Stutts, M.J.; Guggino, W.B.; Bertrand, C.A.; Mitra, S.; Mishra, S.K.; Wang, X.; Zhao, Y.; et al. CFTR Is a Conductance Regulator as well as a Chloride Channel. Physiol. Rev. 1999, 79, S145–S166. [Google Scholar] [CrossRef]
  91. Egan, M.E.; Schwiebert, E.M.; Guggino, W.B. Differential expression of ORCC and CFTR induced by low temperature in CF airway epithelial cells. Am. J. Physiol. Physiol. 1995, 268, C243–C251. [Google Scholar] [CrossRef] [PubMed]
  92. Hayslett, J.P.; Gögelein, H.; Kunzelmann, K.; Greger, R. Characteristics of apical chloride channels in human colon cells (HT29). Pflug. Arch. 1987, 410, 487–494. [Google Scholar] [CrossRef]
  93. Kunzelmann, K.; Pavenstädt, H.; Greger, R. Properties and regulation of chloride channels in cystic fibrosis and normal airway epithelial cells. Pflüg. Arch. 1989, 415, 172–182. [Google Scholar] [CrossRef]
  94. Riordan, J.R.; Rommens, J.M.; Kerem, B.-S.; Alon, N.; Rozmahel, R.; Grzelczak, Z.; Zielenski, J.; Plavsic, S.L.N.; Chou, J.-L.; Drumm, M.L.; et al. Identification of the cystic fibrosis gene: Cloning and characterization of complementary DNA. Science 1989, 245, 1066–1072. [Google Scholar] [CrossRef] [PubMed]
  95. Rommens, J.M.; Iannuzzi, M.C.; Kerem, B.; Drumm, M.L.; Melmer, G.; Dean, M.; Rozmahel, R.; Cole, J.L.; Kennedy, D.; Hidaka, N.; et al. Identification of the cystic fibrosis gene: Chromosome walking and jumping. Science 1989, 245, 1059–1065. [Google Scholar] [CrossRef] [PubMed]
  96. Kerem, B.-S.; Rommens, M.; Buchanan, J.; Markiewicz, D.; Cox, T.; Aravinda, C.; Buchwald, M.; Tsui, L.-C. Identification of the cystic fibrosis gene: Genetic analysis. Science 1989, 245, 1073–1080. [Google Scholar] [CrossRef] [PubMed]
  97. Gabriel, S.E.; Clarke, L.L.; Boucher, R.C.; Stutts, M.J. CFTR and outward rectifying chloride channels are distinct proteins with a regulatory relationship. Nature 1993, 363, 263–268. [Google Scholar] [CrossRef] [PubMed]
  98. Schwiebert, E.M.; Flotte, T.; Cutting, G.R.; Guggino, W.B. Both CFTR and outwardly rectifying chloride channels contribute to cAMP-stimulated whole cell chloride currents. Am. J. Physiol. 1994, 266, C1464–C1477. [Google Scholar] [CrossRef] [PubMed]
  99. Szabo, I.; Lepple-Wienhues, A.; Kaba, K.N.; Zoratti, M.; Gulbins, E.; Lang, F. Tyrosine kinase-dependent activation of a chloride channel in CD95-induced apoptosis in T lymphocytes. Proc. Natl. Acad. Sci. USA 1998, 95, 6169–6174. [Google Scholar] [CrossRef]
  100. Martins, J.R.; Faria, D.; Kongsuphol, P.; Reisch, B.; Schreiber, R.; Kunzelmann, K. Anoctamin 6 is an essential component of the outwardly rectifying chloride channel. Proc. Natl. Acad. Sci. USA 2011, 108, 18168–18172. [Google Scholar] [CrossRef]
  101. Kunzelmann, K.; Ousingsawat, J.; Benedetto, R.; Cabrita, I.; Schreiber, R. Contribution of Anoctamins to Cell Survival and Cell Death. Cancers 2019, 19, E382. [Google Scholar] [CrossRef]
  102. Cantin, A.M.; North, S.L.; Hubbard, R.C.; Crystal, R.G. Normal alveolar epithelial lining fluid contains high levels of glutathione. J. Appl. Physiol. 1987, 63, 152–157. [Google Scholar] [CrossRef]
  103. Rahman, I.; MacNee, W. Oxidative stress and regulation of glutathione in lung inflammation. Eur. Respir. J. 2000, 16, 534–554. [Google Scholar] [CrossRef]
  104. Kogan, I.; Ramjeesingh, M.; Li, C.; Kidd, J.F.; Wang, Y.; Leslie, E.M.; Cole, S.P.; Bear, C.E. CFTR directly mediates nucleotide-regulated glutathione flux. EMBO J. 2003, 22, 1981–1989. [Google Scholar] [CrossRef]
  105. Gao, L.; Kim, K.J.; Yankaskas, J.R.; Forman, H.J. Abnormal glutathione transport in cystic fibrosis airway epithelia. Am. J. Physiol. 1999, 277, L113–L118. [Google Scholar] [CrossRef]
  106. Simoes, F.; Ousingsawat, J.; Wanitchakool, P.; Fonseca, A.; Cabrita, I.; Benedetto, R.; Schreiber, R.; Kunzelmann, K. CFTR supports cell death through ROS-dependent activation of TMEM16F (anoctamin 6). Pflug. Arch. 2018, 470, 305–314. [Google Scholar] [CrossRef]
  107. Ousingsawat, J.; Schreiber, R.; Gulbins, E.; Kamler, M.; Kunzelmann, K.P. aeruginosa Induced Lipid Peroxidation Causes Ferroptotic Cell Death in Airways. Cell Physiol. Biochem. 2021, 55, 590–604. [Google Scholar] [CrossRef]
  108. Ousingsawat, J.; Schreiber, R.; Kunzelmann, K. TMEM16F/Anoctamin 6 in Ferroptotic Cell Death. Cancers 2019, 11, 625. [Google Scholar] [CrossRef] [PubMed]
  109. Moliteo, E.; Sciacca, M.; Palmeri, A.; Papale, M.; Manti, S.; Parisi, G.F.; Leonardi, S. Cystic Fibrosis and Oxidative Stress: The Role of CFTR. Molecules 2022, 27, 5324. [Google Scholar] [CrossRef] [PubMed]
  110. Schreiber, R.; Buchholz, B.; Kraus, A.; Schley, G.; Scholz, J.; Ousingsawat, J.; Kunzelmann, K. Lipid peroxidation drives renal cyst growth in vitro through activation of TMEM16A. J. Am. Soc. Nephrol. 2019, 30, 228–242. [Google Scholar] [CrossRef]
  111. Ratcliff, R.; Evans, M.J.; Cuthbert, A.W.; MacVinish, L.J.; Foster, D.; Anderson, J.R.; Colledge, W.H. Production of a severe cystic fibrosis mutation in mice by gene targeting. Nat. Genet. 1993, 4, 35–41. [Google Scholar] [CrossRef]
  112. Grubb, B.R.; Boucher, R.C. Pathophysiology of gene-targeted mouse models for cystic fibrosis. Physiol. Rev. 1999, 79, S193–S214. [Google Scholar] [CrossRef]
  113. Wilke, M.; Buijs-Offerman, R.M.; Aarbiou, J.; Colledge, W.H.; Sheppard, D.N.; Touqui, L.; Bot, A.; Jorna, H.; De Jonge, H.R.; Scholte, B.J. Mouse models of cystic fibrosis: Phenotypic analysis and research applications. J. Cyst. Fibros. 2011, 10 (Suppl. S2), 9. [Google Scholar] [CrossRef]
  114. Chen, J.H.; Stoltz, D.A.; Karp, P.H.; Ernst, S.E.; Pezzulo, A.A.; Moninger, T.O.; Rector, M.V.; Reznikov, L.R.; Launspach, J.L.; Chaloner, K.; et al. Loss of anion transport without increased sodium absorption characterizes newborn porcine cystic fibrosis airway epithelia. Cell 2010, 143, 911–923. [Google Scholar] [CrossRef] [PubMed]
  115. Ostedgaard, L.S.; Meyerholz, D.K.; Chen, J.H.; Pezzulo, A.A.; Karp, P.H.; Rokhlina, T.; Ernst, S.E.; Hanfland, R.A.; Reznikov, L.R.; Ludwig, P.S.; et al. The {Delta}F508 Mutation Causes CFTR Misprocessing and Cystic Fibrosis-Like Disease in Pigs. Sci. Transl. Med. 2011, 3, 74ra24. [Google Scholar] [CrossRef] [PubMed]
  116. Pedersen, K.A.; Schroder, R.L.; Skaaning-Jensen, B.; Strobaek, D.; Olesen, S.P.; Christophersen, P. Activation of the human intermediate-conductance Ca(2+)-activated K(+) channel by 1-ethyl-2-benzimidazolinone is strongly Ca(2+)-dependent. Biochim. Biophys. Acta 1999, 1420, 231–240. [Google Scholar] [CrossRef]
  117. Stoltz, D.A.; Meyerholz, D.K.; Pezzulo, A.A.; Ramachandran, S.; Rogan, M.P.; Davis, G.J.; Hanfland, R.A.; Wohlford-Lenane, C.; Dohrn, C.L.; Bartlett, J.A.; et al. Cystic fibrosis pigs develop lung disease and exhibit defective bacterial eradication at birth. Sci. Transl. Med. 2010, 2, 29ra31. [Google Scholar] [CrossRef] [PubMed]
  118. Pezzulo, A.A.; Tang, X.X.; Hoegger, M.J.; Alaiwa, M.H.; Ramachandran, S.; Moninger, T.O.; Karp, P.H.; Wohlford-Lenane, C.L.; Haagsman, H.P.; van Eijk, M.; et al. Reduced airway surface pH impairs bacterial killing in the porcine cystic fibrosis lung. Nature 2012, 487, 109–113. [Google Scholar] [CrossRef]
  119. Hoegger, M.J.; Fischer, A.J.; McMenimen, J.D.; Ostedgaard, L.S.; Tucker, A.J.; Awadalla, M.A.; Moninger, T.O.; Michalski, A.S.; Hoffman, E.A.; Zabner, J.; et al. Impaired mucus detachment disrupts mucociliary transport in a piglet model of cystic fibrosis. Science 2014, 345, 818–822. [Google Scholar] [CrossRef]
  120. Shah, V.S.; Ernst, S.; Tang, X.X.; Karp, P.H.; Parker, C.P.; Ostedgaard, L.S.; Welsh, M.J. Relationships among CFTR expression, HCO3- secretion, and host defense may inform gene- and cell-based cystic fibrosis therapies. Proc. Natl. Acad. Sci. USA 2016, 113, 5382–5387. [Google Scholar] [CrossRef]
  121. Klymiuk, N.; Mundhenk, L.; Kraehe, K.; Wuensch, A.; Plog, S.; Emrich, D.; Langenmayer, M.C.; Stehr, M.; Holzinger, A.; Kroner, C.; et al. Sequential targeting of CFTR by BAC vectors generates a novel pig model of cystic fibrosis. J. Mol. Med. 2012, 90, 597–608. [Google Scholar] [CrossRef]
  122. Esther, C.R., Jr.; Muhlebach, M.S.; Ehre, C.; Hill, D.B.; Wolfgang, M.C.; Kesimer, M.; Ramsey, K.A.; Markovetz, M.R.; Garbarine, I.C.; Forest, M.G.; et al. Mucus accumulation in the lungs precedes structural changes and infection in children with cystic fibrosis. Sci. Transl. Med. 2019, 11, eaav3488. [Google Scholar] [CrossRef]
  123. Schultz, A.; Puvvadi, R.; Borisov, S.M.; Shaw, N.C.; Klimant, I.; Berry, L.J.; Montgomery, S.T.; Nguyen, T.; Kreda, S.M.; Kicic, A.; et al. Airway surface liquid pH is not acidic in children with cystic fibrosis. Nat. Commun. 2017, 8, 1409. [Google Scholar] [CrossRef]
  124. Chen, A.P.; Chang, M.H.; Romero, M.F. Functional analysis of nonsynonymous single nucleotide polymorphisms in human SLC26A9. Hum. Mutat. 2012, 33, 1275–1284. [Google Scholar] [CrossRef]
  125. Sun, L.; Rommens, J.M.; Corvol, H.; Li, W.; Li, X.; Chiang, T.A.; Lin, F.; Dorfman, R.; Busson, P.F.; Parekh, R.V.; et al. Multiple apical plasma membrane constituents are associated with susceptibility to meconium ileus in individuals with cystic fibrosis. Nat. Genet. 2012, 44, 562–569. [Google Scholar] [CrossRef]
  126. Miller, M.R.; Soave, D.; Li, W.; Gong, J.; Pace, R.G.; Boelle, P.Y.; Cutting, G.R.; Drumm, M.L.; Knowles, M.R.; Sun, L.; et al. Variants in Solute Carrier SLC26A9 Modify Prenatal Exocrine Pancreatic Damage in Cystic Fibrosis. J. Pediatr. 2015, 166, 1152–1157.e1156. [Google Scholar] [CrossRef]
  127. Lam, A.N.; Aksit, M.A.; Vecchio-Pagan, B.; Shelton, C.A.; Osorio, D.L.; Anzmann, A.F.; Goff, L.A.; Whitcomb, D.C.; Blackman, S.M.; Cutting, G.R. Increased expression of anion transporter SLC26A9 delays diabetes onset in cystic fibrosis. J. Clin. Investig. 2020, 130, 272–286. [Google Scholar] [CrossRef]
  128. Li, T.; di Stefano, G.; Raza, G.S.; Sommerer, I.; Riederer, B.; Römermann, D.; Tan, X.; Tan, Q.; Pallagi, P.; Hollenbach, M.; et al. Hydrokinetic pancreatic function and insulin secretion are moduled by Cl() uniporter Slc26a9 in mice. Acta Physiol. 2022, 234, e13729. [Google Scholar] [CrossRef]
  129. Aksit, M.A.; Ling, H.; Pace, R.G.; Raraigh, K.S.; Onchiri, F.; Faino, A.V.; Pagel, K.; Pugh, E.; Stilp, A.M.; Sun, Q.; et al. Pleiotropic modifiers of age-related diabetes and neonatal intestinal obstruction in cystic fibrosis. Am. J. Hum. Genet. 2022, 109, 1894–1908. [Google Scholar] [CrossRef]
  130. Strug, L.J.; Gonska, T.; He, G.; Keenan, K.; Ip, W.; Boelle, P.Y.; Lin, F.; Panjwani, N.; Gong, J.; Li, W.; et al. Cystic fibrosis gene modifier SLC26A9 modulates airway response to CFTR-directed therapeutics. Hum. Mol. Genet. 2016, 25, 4590–4600. [Google Scholar] [CrossRef]
  131. Eastman, A.C.; Pace, R.G.; Dang, H.; Aksit, M.A.; Vecchio-Pagán, B.; Lam, A.N.; O’Neal, W.K.; Blackman, S.M.; Knowles, M.R.; Cutting, G.R. SLC26A9 SNP rs7512462 is not associated with lung disease severity or lung function response to ivacaftor in cystic fibrosis patients with G551D-CFTR. J. Cyst. Fibros. Off. J. Eur. Cyst. Fibros. Soc. 2021, 20, 851–856. [Google Scholar] [CrossRef]
  132. Ousingsawat, J.; Centeio, R.; Schreiber, R.; Kunzelmann, K. Expression of SLC26A9 in Airways and Its Potential Role in Asthma. Int. J. Mol. Sci. 2022, 23, 2998. [Google Scholar] [CrossRef]
  133. Chang, M.H.; Plata, C.; Zandi-Nejad, K.; Sindic, A.; Sussman, C.R.; Mercado, A.; Broumand, V.; Raghuram, V.; Mount, D.B.; Romero, M.F. Slc26a9-Anion Exchanger, Channel and Na(+) Transporter. J. Membr. Biol. 2009, 228, 125–140. [Google Scholar] [CrossRef] [PubMed]
  134. Xu, J.; Henriksnäs, J.; Barone, S.; Witte, D.; Shull, G.E.; Forte, J.G.; Holm, L.; Soleimani, M. SLC26A9 is expressed in gastric surface epithelial cells, mediates Cl/HCO3 exchange, and is inhibited by NH4+. Am. J. Physiol. Cell Physiol. 2005, 289, C493–C505. [Google Scholar] [CrossRef]
  135. Demitrack, E.S.; Soleimani, M.; Montrose, M.H. Damage to the gastric epithelium activates cellular bicarbonate secretion via SLC26A9 Cl()/HCO(3)(). Am. J. Physiol. Gastrointest. Liver Physiol. 2010, 299, G255–G264. [Google Scholar] [CrossRef]
  136. Dorwart, M.R.; Shcheynikov, N.; Wang, Y.; Stippec, S.; Muallem, S. SLC26A9 is a Cl() channel regulated by the WNK kinases. J. Physiol. 2007, 584, 333–345. [Google Scholar] [CrossRef]
  137. Loriol, C.; Dulong, S.; Avella, M.; Gabillat, N.; Boulukos, K.; Borgese, F.; Ehrenfeld, J. Characterization of SLC26A9, facilitation of Cl() transport by bicarbonate. Cell Physiol. Biochem. 2008, 22, 15–30. [Google Scholar] [CrossRef]
  138. Walter, J.D.; Sawicka, M.; Dutzler, R. Cryo-EM structures and functional characterization of murine Slc26a9 reveal mechanism of uncoupled chloride transport. eLife 2019, 8, e46986. [Google Scholar] [CrossRef]
  139. Ko, S.B.; Shcheynikov, N.; Choi, J.Y.; Luo, X.; Ishibashi, K.; Thomas, P.J.; Kim, J.Y.; Kim, K.H.; Lee, M.G.; Naruse, S.; et al. A molecular mechanism for aberrant CFTR-dependent HCO(3)() transport in cystic fibrosis. EMBO J. 2002, 21, 5662–5672. [Google Scholar] [CrossRef] [PubMed]
  140. Rakonczay, Z., Jr.; Hegyi, P.; Hasegawa, M.; Inoue, M.; You, J.; Iida, A.; Ignáth, I.; Alton, E.W.; Griesenbach, U.; Ovári, G.; et al. CFTR gene transfer to human cystic fibrosis pancreatic duct cells using a Sendai virus vector. J. Cell Physiol. 2008, 214, 442–455. [Google Scholar] [CrossRef]
  141. 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]
  142. El Khouri, E.; Toure, 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]
  143. Bozoky, Z.; Krzeminski, M.; Muhandiram, R.; Birtley, J.R.; Al-Zahrani, A.; Thomas, P.J.; Frizzell, R.A.; Ford, R.C.; Forman-Kay, J.D. Regulatory R region of the CFTR chloride channel is a dynamic integrator of phospho-dependent intra- and intermolecular interactions. Proc. Natl. Acad. Sci. USA 2013, 110, E4427–E4436. [Google Scholar] [CrossRef]
  144. Kunzelmann, K.; Centeio, R.; Ousingsawat, J.; Talbi, K.; Seidler, U.; Schreiber, R. SLC26A9 in airways and intestine: Secretion or absorption? Channels 2023, 17, 2186434. [Google Scholar] [CrossRef] [PubMed]
  145. Bretscher, D.; Schneider, A.; Hagmann, R.; Hadorn, B.; Howald, B.; Lüthy, C.; Oetliker, O. Response of renal handling of sodium and bicarbonate to secretin in normals and in patients with cystic fibrosis. Pediatr. Res. 1974, 8, 899. [Google Scholar] [CrossRef]
  146. Windstetter, D.; Schaefer, F.; Scharer, K.; Reiter, K.; Eife, R.; Harms, H.K.; Bertele-Harms, R.; Fiedler, F.; Tsui, L.C.; Reitmeir, P.; et al. Renal function and renotropic effects of secretin in cystic fibrosis. Eur. J. Med. Res. 1997, 2, 431–436. [Google Scholar]
  147. Johansen, P.G.; Hadorn, B.; Anderson, C.M. Effect of Secretin on Human Rectal Mucosa in vivo. Nature 1968, 217, 468. [Google Scholar] [CrossRef]
  148. Hadorn, B.; Johansen, P.G.; Anderson, C.M. Pancreozymin secretin test of exocrine pancreatic funtion in cystic fribrosis and the significance of the result for the pathogenesis of the disease. Can. Med. Assoc. J. 1968, 98, 377–385. [Google Scholar]
  149. Berg, P.; Sorensen, M.V.; Rousing, A.Q.; Olesen, H.V.; Jensen-Fangel, S.; Jeppesen, M.; Leipziger, J. Challenged Urine Bicarbonate Excretion as a Measure of Cystic Fibrosis Transmembrane Conductance Regulator Function in Cystic Fibrosis. Ann. Intern. Med. 2022, 175, 1543–1551. [Google Scholar] [CrossRef] [PubMed]
  150. Berg, P.; Svendsen, S.L.; Sorensen, M.V.; Larsen, C.K.; Andersen, J.F.; Jensen-Fangel, S.; Jeppesen, M.; Schreiber, R.; Cabrita, I.; Kunzelmann, K.; et al. Impaired Renal HCO(3)(−) Excretion in Cystic Fibrosis. J. Am. Soc. Nephrol. 2020, 31, 1711–1727. [Google Scholar] [CrossRef]
  151. Berg, P.; Svendsen, S.L.; Sorensen, M.V.; Schreiber, R.; Kunzelmann, K.; Leipziger, J. The molecular mechanism of CFTR- and secretin-dependent renal bicarbonate excretion. J. Physiol. 2021, 599, 3003–3011. [Google Scholar] [CrossRef]
  152. Torres, V.E.; Harris, P.C.; Pirson, Y. Autosomal dominant polycystic kidney disease. Lancet 2007, 369, 1287–1301. [Google Scholar] [CrossRef]
  153. Senum, S.R.; Li, Y.S.M.; Benson, K.A.; Joli, G.; Olinger, E.; Lavu, S.; Madsen, C.D.; Gregory, A.V.; Neatu, R.; Kline, T.L.; et al. Monoallelic IFT140 pathogenic variants are an important cause of the autosomal dominant polycystic kidney-spectrum phenotype. Am. J. Hum. Genet. 2022, 109, 136–156. [Google Scholar] [CrossRef]
  154. Raphael, K.L.; Strait, K.A.; Stricklett, P.K.; Miller, R.L.; Nelson, R.D.; Piontek, K.B.; Germino, G.G.; Kohan, D.E. Inactivation of Pkd1 in principal cells causes a more severe cystic kidney disease than in intercalated cells. Kidney Int. 2009, 75, 626–633. [Google Scholar] [CrossRef]
  155. Buchholz, B.; Teschemacher, B.; Schley, G.; Schillers, H.; Eckardt, K.U. Formation of cysts by principal-like MDCK cells depends on the synergy of cAMP- and ATP-mediated fluid secretion. J. Mol. Med. 2011, 89, 251–261. [Google Scholar] [CrossRef]
  156. Grantham, J.J.; Mulamalla, S.; Swenson-Fields, K.I. Why kidneys fail in autosomal dominant polycystic kidney disease. Nat. Rev. Nephrol. 2011, 7, 556–566. [Google Scholar] [CrossRef]
  157. Terryn, S.; Ho, A.; Beauwens, R.; Devuyst, O. Fluid transport and cystogenesis in autosomal dominant polycystic kidney disease. Biochim. Biophys. Acta 2011, 1812, 1314–1321. [Google Scholar] [CrossRef]
  158. Davidow, C.J.; Maser, R.L.; Rome, L.A.; Calvet, J.P.; Grantham, J.J. The cystic fibrosis transmembrane conductance regulator mediates transepithelial fluid secretion by human autosomal dominant polycystic kidney disease epithelium in vitro. Kidney Int. 1996, 50, 208–218. [Google Scholar] [CrossRef]
  159. Magenheimer, B.S.; St John, P.L.; Isom, K.S.; Abrahamson, D.R.; De Lisle, R.C.; Wallace, D.P.; Maser, R.L.; Grantham, J.J.; Calvet, J.P. Early embryonic renal tubules of wild-type and polycystic kidney disease kidneys respond to cAMP stimulation with cystic fibrosis transmembrane conductance regulator/Na(+),K(+),2Cl(−) Co-transporter-dependent cystic dilation. J. Am. Soc. Nephrol. 2006, 17, 3424–3437. [Google Scholar] [CrossRef]
  160. Yang, B.; Sonawane, N.D.; Zhao, D.; Somlo, S.; Verkman, A.S. Small-Molecule CFTR Inhibitors Slow Cyst Growth in Polycystic Kidney Disease. J. Am. Soc. Nephrol. 2008, 19, 1300–1310. [Google Scholar] [CrossRef]
  161. Cabrita, I.; Kraus, A.; Scholz, J.K.; Skoczynski, K.; Schreiber, R.; Kunzelmann, K.; Buchholz, B. Cyst growth in ADPKD is prevented by pharmacological and genetic inhibition of TMEM16A in vivo. Nat. Commun. 2020, 11, 4320. [Google Scholar] [CrossRef]
  162. Cabrita, I.; Buchholz, B.; Schreiber, R.; Kunzelmann, K. TMEM16A drives renal cyst growth by augmenting Ca(2+) signaling in M1 cells. J. Mol. Med. 2020, 98, 659–671. [Google Scholar] [CrossRef]
  163. Cabrita, I.; Talbi, K.; Kunzelmann, K.; Schreiber, R. Loss of PKD1 and PKD2 share common effects on intracellular Ca2+ signaling. Cell Calcium 2021, 97, 102413. [Google Scholar] [CrossRef] [PubMed]
  164. Gabow, P.A.; Johnson, A.M.; Kaehny, W.D.; Kimberling, W.J.; Lezotte, D.C.; Duley, I.T.; Jones, R.H. Factors affecting the progression of renal disease in autosomal-dominant polycystic kidney disease. Kidney Int. 1992, 41, 1311–1319. [Google Scholar] [CrossRef]
  165. Stewart, J.H. End-stage renal failure appears earlier in men than in women with polycystic kidney disease. Am. J. Kidney Dis. 1994, 24, 181–183. [Google Scholar] [CrossRef]
  166. Talbi, K.; Cabrita, I.; Schreiber, R.; Kunzelmann, K. Gender-Dependent Phenotype in Polycystic Kidney Disease Is Determined by Differential Intracellular Ca2+ Signals. Int. J. Mol. Sci. 2021, 22, 6019. [Google Scholar] [CrossRef] [PubMed]
  167. Nakanishi, K.; Sweeney, W.E., Jr.; Macrae, D.K.; Cotton, C.U.; Avner, E.D. Role of CFTR in autosomal recessive polycystic kidney disease. J. Am. Soc. Nephrol. 2001, 12, 719–725. [Google Scholar] [CrossRef]
  168. Menon, V.; Rudym, D.; Chandra, P.; Miskulin, D.; Perrone, R.; Sarnak, M. Inflammation, oxidative stress, and insulin resistance in polycystic kidney disease. Clin. J. Am. Soc. Nephrol. CJASN 2011, 6, 7–13. [Google Scholar] [CrossRef]
  169. Hanaoka, K.; Devuyst, O.; Schwiebert, E.M.; Wilson, P.D.; Guggino, W.B. A role for CFTR in human autosomal dominant polycystic kidney disease. Am. J. Physiol. 1996, 270, C389–C399. [Google Scholar] [CrossRef]
  170. O’Sullivan, D.A.; Torres, V.E.; Gabow, P.A.; Thibodeau, S.N.; King, B.F.; Bergstralh, E.J. Cystic fibrosis and the phenotypic expression of autosomal dominant polycystic kidney disease. Am. J. Kidney Dis. 1998, 32, 976–983. [Google Scholar] [CrossRef]
  171. Xu, N.; Glockner, J.F.; Rossetti, S.; Babovich-Vuksanovic, D.; Harris, P.C.; Torres, V.E. Autosomal dominant polycystic kidney disease coexisting with cystic fibrosis. J. Nephrol. 2006, 19, 529–534. [Google Scholar]
  172. Persu, A.; Devuyst, O.; Lannoy, N.; Materne, R.; Brosnahan, G.; Gabow, P.A.; Pirson, Y.; Verellen-Dumoulin, C. CF gene and cystic fibrosis transmembrane conductance regulator expression in autosomal dominant polycystic kidney disease. J. Am. Soc. Nephrol. 2000, 11, 2285–2296. [Google Scholar] [CrossRef] [PubMed]
  173. Brill, S.R.; Ross, K.E.; Davidow, C.J.; Ye, M.; Grantham, J.J.; Caplan, M.J. Immunolocalization of ion transport proteins in human autosomal dominant polycystic kidney epithelial cells. Proc. Natl. Acad. Sci. USA 1996, 93, 10206–10211. [Google Scholar] [CrossRef]
  174. Buchholz, B.; Faria, D.; Schley, G.; Schreiber, R.; Eckardt, K.U.; Kunzelmann, K. Anoctamin 1 induces calcium-activated chloride secretion and tissue proliferation in polycystic kidney disease. Kidney Int. 2014, 85, 1058–1067. [Google Scholar] [CrossRef]
  175. Schwiebert, E.M.; Wallace, D.P.; Braunstein, G.M.; King, S.R.; Peti-Peterdi, J.; Hanaoka, K.; Guggino, W.B.; Guay-Woodford, L.M.; Bell, P.D.; Sullivan, L.P.; et al. Autocrine extracellular purinergic signaling in epithelial cells derived from polycystic kidneys. Am. J. Physiol. Ren. Physiol. 2002, 282, F763–F775. [Google Scholar] [CrossRef] [PubMed]
  176. Talbi, K.; Cabrita, I.; Kraus, A.; Hofmann, S.; Skoczynski, K.; Kunzelmann, K.; Buchholz, B.; Schreiber, R. The chloride channel CFTR is not required for cyst growth in an ADPKD mouse model. FASEB J. 2021, 35, e21897. [Google Scholar] [CrossRef] [PubMed]
  177. Guilbault, C.; Saeed, Z.; Downey, G.P.; Radzioch, D. Cystic fibrosis mouse models. Am. J. Respir. Cell Mol. Biol. 2007, 36, 1–7. [Google Scholar] [CrossRef] [PubMed]
  178. Yanda, M.K.; Cha, B.; Cebotaru, C.V.; Cebotaru, L. Pharmacological reversal of renal cysts from secretion to absorption suggests a potential therapeutic strategy for managing autosomal dominant polycystic kidney disease. J. Biol. Chem. 2019, 294, 17090–17104. [Google Scholar] [CrossRef]
  179. Yanda, M.K.; Cebotaru, L. VX-809 mitigates disease in a mouse model of autosomal dominant polycystic kidney disease bearing the R3277C human mutation. FASEB J. 2021, 35, e21987. [Google Scholar] [CrossRef]
  180. van Koningsbruggen-Rietschel, S.; Conrath, K.; Fischer, R.; Sutharsan, S.; Kempa, A.; Gleiber, W.; Schwarz, C.; Hector, A.; Van Osselaer, N.; Pano, A.; et al. GLPG2737 in lumacaftor/ivacaftor-treated CF subjects homozygous for the F508del mutation: A randomized phase 2A trial (PELICAN). J. Cyst. Fibros. Off. J. Eur. Cyst. Fibros. Soc. 2020, 19, 292–298. [Google Scholar] [CrossRef]
  181. Chen, W.; Mook, R.A., Jr.; Premont, R.T.; Wang, J. Niclosamide: Beyond an antihelminthic drug. Cell Signal 2018, 41, 89–96. [Google Scholar] [CrossRef] [PubMed]
  182. Lee, M.H.; Graham, G.G.; Williams, K.M.; Day, R.O. A benefit-risk assessment of benzbromarone in the treatment of gout. Was its withdrawal from the market in the best interest of patients? Drug Saf. 2008, 31, 643–665. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Attenuated CFTR-dependent Cl secretion in mice with intestinal epithelial knockout of Ano1. (A) Original Ussing chamber recordings obtained from colonic epithelia under open circuit conditions, as described in [31]. Stimulation of colonic epithelia with IBMX and forskolin (I/F; 100 µM/2 µM) induced a pronounced voltage deflection in normal mouse colonic epithelium (Ano1+/+), indicating pronounced cAMP-activated secretion. The voltage deflection, i.e., Cl secretion, was strongly attenuated in a colonic tissue obtained from a mouse lacking expression of Ano1 (Ano1−/−). (B) Calculated equivalent short circuit currents (Isc) indicate strongly attenuated CFTR-dependent (I/F-stimulated) as well as Ca2+-dependent (100 µM ATP-stimulated) Cl secretion in mouse colon lacking epithelial expression of Ano1 (−/−), when compared to wild-type colons (−/−). Mean ± SEM (number of experiments for +/+ and −/−). § Significantly reduced when compared to basal Isc in +/+ tissues (ANOVA). # Significantly reduced when compared to stimulated Isc in +/+ tissues (ANOVA). For methods, see [31].
Figure 1. Attenuated CFTR-dependent Cl secretion in mice with intestinal epithelial knockout of Ano1. (A) Original Ussing chamber recordings obtained from colonic epithelia under open circuit conditions, as described in [31]. Stimulation of colonic epithelia with IBMX and forskolin (I/F; 100 µM/2 µM) induced a pronounced voltage deflection in normal mouse colonic epithelium (Ano1+/+), indicating pronounced cAMP-activated secretion. The voltage deflection, i.e., Cl secretion, was strongly attenuated in a colonic tissue obtained from a mouse lacking expression of Ano1 (Ano1−/−). (B) Calculated equivalent short circuit currents (Isc) indicate strongly attenuated CFTR-dependent (I/F-stimulated) as well as Ca2+-dependent (100 µM ATP-stimulated) Cl secretion in mouse colon lacking epithelial expression of Ano1 (−/−), when compared to wild-type colons (−/−). Mean ± SEM (number of experiments for +/+ and −/−). § Significantly reduced when compared to basal Isc in +/+ tissues (ANOVA). # Significantly reduced when compared to stimulated Isc in +/+ tissues (ANOVA). For methods, see [31].
Ijms 24 13278 g001
Figure 2. Cytokines in a sputum sample from a patient carrying the ANO1 loss-of-function variant c.897 + 3_897 + 6delAAGT are strongly reduced when compared to samples from CF patients. Sputum samples were obtained from a healthy volunteer, the patient carrying the ANO1 variant c.897 + 3_897 + 6delAAGT, and from two CF patients, and concentrations of the cytokines IL-8 and IL-1ß were determined. Although the ANO1 patient lacks CFTR function in addition to the defect in ANO1 Ca2+-dependent Cl secretion, cytokines were strongly reduced when compared to CF patients carrying known CFTR mutations. Mean ± SEM (number of measurements). # Significantly enhanced when compared to the healthy volunteer (ANOVA). $ Significantly enhanced when compared to the ANO1 patients (ANOVA). For methods see [84].
Figure 2. Cytokines in a sputum sample from a patient carrying the ANO1 loss-of-function variant c.897 + 3_897 + 6delAAGT are strongly reduced when compared to samples from CF patients. Sputum samples were obtained from a healthy volunteer, the patient carrying the ANO1 variant c.897 + 3_897 + 6delAAGT, and from two CF patients, and concentrations of the cytokines IL-8 and IL-1ß were determined. Although the ANO1 patient lacks CFTR function in addition to the defect in ANO1 Ca2+-dependent Cl secretion, cytokines were strongly reduced when compared to CF patients carrying known CFTR mutations. Mean ± SEM (number of measurements). # Significantly enhanced when compared to the healthy volunteer (ANOVA). $ Significantly enhanced when compared to the ANO1 patients (ANOVA). For methods see [84].
Ijms 24 13278 g002
Figure 3. Proposed modified scheme for Cl and HCO3 secretion in airways. Cl and HCO3 ions are taken up on the basolateral site of airway epithelial cells by the Na+/K+/2Cl cotransporter NKCC1, NBCe1, and AE2, respectively. The Na+/K+-ATPase and the basolateral K+ channels KCNQ1/KCNE3 and KCNN4 provide the driving force for luminal anion secretion. Luminal secretion of Cl occurs essentially through CFTR, with only a negligible contribution of ANO1. ANO1 supports Cl secretion through tethering of the endoplasmic reticulum near the apical compartment and an increase in the subapical Ca2+ concentration that facilitates activation and membrane expression of CFTR. CFTR also provides a recycling pathway for Cl that allows the secretion of HCO3 by SLC26A4 and probably SLC26A9.
Figure 3. Proposed modified scheme for Cl and HCO3 secretion in airways. Cl and HCO3 ions are taken up on the basolateral site of airway epithelial cells by the Na+/K+/2Cl cotransporter NKCC1, NBCe1, and AE2, respectively. The Na+/K+-ATPase and the basolateral K+ channels KCNQ1/KCNE3 and KCNN4 provide the driving force for luminal anion secretion. Luminal secretion of Cl occurs essentially through CFTR, with only a negligible contribution of ANO1. ANO1 supports Cl secretion through tethering of the endoplasmic reticulum near the apical compartment and an increase in the subapical Ca2+ concentration that facilitates activation and membrane expression of CFTR. CFTR also provides a recycling pathway for Cl that allows the secretion of HCO3 by SLC26A4 and probably SLC26A9.
Ijms 24 13278 g003
Figure 4. ANO1 is colocalized with pendrin in the apical membrane of ß-intercalated cells. (A) Immunocytochemistry demonstrating colocalization of ANO1 and pendrin in the apical membrane of ß-intercalated cells of mouse collecting ducts. For methods, see [150]. (B) Model showing the molecular mechanisms for HCO3 excretion by collecting duct ß-intercalated cells. Blood HCO3 is taken up into ß-intercalated cells and is transported by pendrin into the collecting duct lumen in exchange with Cl, which is recycled via colocalized CFTR. In addition, CFTR may directly interact with SLC26A4. Colocalized ANO1 tethers the endoplasmic reticulum (ER) to the apical membrane and facilitates efficient Ca2+ signaling in the apical compartment, which supports insertion of CFTR into the apical membrane and its activation. An increase in blood secretin leads to the activation of basolateral secretin receptors (SCTR), which further activates CFTR and HCO3 excretion. For methods, see [150].
Figure 4. ANO1 is colocalized with pendrin in the apical membrane of ß-intercalated cells. (A) Immunocytochemistry demonstrating colocalization of ANO1 and pendrin in the apical membrane of ß-intercalated cells of mouse collecting ducts. For methods, see [150]. (B) Model showing the molecular mechanisms for HCO3 excretion by collecting duct ß-intercalated cells. Blood HCO3 is taken up into ß-intercalated cells and is transported by pendrin into the collecting duct lumen in exchange with Cl, which is recycled via colocalized CFTR. In addition, CFTR may directly interact with SLC26A4. Colocalized ANO1 tethers the endoplasmic reticulum (ER) to the apical membrane and facilitates efficient Ca2+ signaling in the apical compartment, which supports insertion of CFTR into the apical membrane and its activation. An increase in blood secretin leads to the activation of basolateral secretin receptors (SCTR), which further activates CFTR and HCO3 excretion. For methods, see [150].
Ijms 24 13278 g004
Figure 5. Knockout of Cftr does not affect cyst growth in an ADPKD mouse model. KspCreERT2; Pkd1lox;lox mice (Pkd1−/−; n = 10) and KspCreERT2; Pkd1lox;lox/Cftrlox;lox mice (Pkd1−/−/Cftr−/−; n = 15) received daily intraperitoneal injections of tamoxifen (2 mg/kg body weight dissolved in 5% ethanol and 95% neutral oil at postnatal days 20–22) to induce tubule-specific deletion of Pkd1 or co-deletion of Pkd1 and Cftr. Non-induced KspCreERT2; Pkd1lox;lox mice (Pkd1+/+; n = 8) served as controls. Analyses were performed 10 weeks after induction with tamoxifen. (A) Representative kidney sections at the end of the experiment. For methods, see [161]. (B) Analysis of the cystic indices defined as the ratio of the cortical cystic area divided by the whole cortex area. (C) Two-kidney weight per body weight ratio. No significant effect of CFTR-knockout was found [161]. Bars show means ± SEM, and dots indicate individual values. * Significant increase compared to Pkd1+/+ (p < 0.05; one-way ANOVA). For methods, see [176].
Figure 5. Knockout of Cftr does not affect cyst growth in an ADPKD mouse model. KspCreERT2; Pkd1lox;lox mice (Pkd1−/−; n = 10) and KspCreERT2; Pkd1lox;lox/Cftrlox;lox mice (Pkd1−/−/Cftr−/−; n = 15) received daily intraperitoneal injections of tamoxifen (2 mg/kg body weight dissolved in 5% ethanol and 95% neutral oil at postnatal days 20–22) to induce tubule-specific deletion of Pkd1 or co-deletion of Pkd1 and Cftr. Non-induced KspCreERT2; Pkd1lox;lox mice (Pkd1+/+; n = 8) served as controls. Analyses were performed 10 weeks after induction with tamoxifen. (A) Representative kidney sections at the end of the experiment. For methods, see [161]. (B) Analysis of the cystic indices defined as the ratio of the cortical cystic area divided by the whole cortex area. (C) Two-kidney weight per body weight ratio. No significant effect of CFTR-knockout was found [161]. Bars show means ± SEM, and dots indicate individual values. * Significant increase compared to Pkd1+/+ (p < 0.05; one-way ANOVA). For methods, see [176].
Ijms 24 13278 g005
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

Kunzelmann, K.; Ousingsawat, J.; Kraus, A.; Park, J.H.; Marquardt, T.; Schreiber, R.; Buchholz, B. Pathogenic Relationships in Cystic Fibrosis and Renal Diseases: CFTR, SLC26A9 and Anoctamins. Int. J. Mol. Sci. 2023, 24, 13278. https://doi.org/10.3390/ijms241713278

AMA Style

Kunzelmann K, Ousingsawat J, Kraus A, Park JH, Marquardt T, Schreiber R, Buchholz B. Pathogenic Relationships in Cystic Fibrosis and Renal Diseases: CFTR, SLC26A9 and Anoctamins. International Journal of Molecular Sciences. 2023; 24(17):13278. https://doi.org/10.3390/ijms241713278

Chicago/Turabian Style

Kunzelmann, Karl, Jiraporn Ousingsawat, Andre Kraus, Julien H. Park, Thorsten Marquardt, Rainer Schreiber, and Björn Buchholz. 2023. "Pathogenic Relationships in Cystic Fibrosis and Renal Diseases: CFTR, SLC26A9 and Anoctamins" International Journal of Molecular Sciences 24, no. 17: 13278. https://doi.org/10.3390/ijms241713278

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