Next Article in Journal
Curcumin for Treating Breast Cancer: A Review of Molecular Mechanisms, Combinations with Anticancer Drugs, and Nanosystems
Next Article in Special Issue
Impact of Drug-Mediated Inhibition of Intestinal Transporters on Nutrient and Endogenous Substrate Disposition…an Afterthought?
Previous Article in Journal
Customizable Three-Dimensional Printed Earring Tap for Treating Affections Caused by Aesthetic Perforations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multiple Regulatory Signals and Components in the Modulation of Bicarbonate Transporters

by
Hyeong Jae Kim
and
Jeong Hee Hong
*
Department of Physiology, Lee Gil Ya Cancer and Diabetes Institute, College of Medicine, Gachon University, 155 Getbeolro, Yeonsu-gu, Incheon 21999, Republic of Korea
*
Author to whom correspondence should be addressed.
Pharmaceutics 2024, 16(1), 78; https://doi.org/10.3390/pharmaceutics16010078
Submission received: 13 December 2023 / Revised: 1 January 2024 / Accepted: 3 January 2024 / Published: 5 January 2024
(This article belongs to the Special Issue New Insights into Transporters in Drug Development)

Abstract

:
Bicarbonate transporters are responsible for the appropriate flux of bicarbonate across the plasma membrane to perform various fundamental cellular functions. The functions of bicarbonate transporters, including pH regulation, cell migration, and inflammation, are highlighted in various cellular systems, encompassing their participation in both physiological and pathological processes. In this review, we focused on recently identified modulatory signaling components that regulate the expression and activity of bicarbonate transporters. Moreover, we addressed recent advances in our understanding of cooperative systems of bicarbonate transporters and channelopathies. This current review aims to provide a new, in-depth understanding of numerous human diseases associated with the dysfunction of bicarbonate transporters.

1. Homeostatic Role of Bicarbonate Transporters

Bicarbonate is a fundamental ion in cellular pH regulation and crosses the membrane via various bicarbonate transporters, such as the chloride/bicarbonate exchangers (CBE; SLC26A family and anion exchanger (AE) family), sodium/bicarbonate cotransporter (NBC), sodium-dependent chloride–bicarbonate exchanger (NDCBE), and the cystic fibrosis (CF) transmembrane conductance regulator (CFTR) channel [1,2,3,4,5]. It has been known that the human adult bicarbonate level is maintained between 23 to 29 mEq/L in serum and 10 mEq/L in the cytosol [6]. Since bicarbonate transporters were first identified, their biological and physiological roles have been revealed in various cellular systems, including pathological processes. Defects in bicarbonate transporters result in disease-state signaling [7]. Although acid-base cellular pH regulation and basic physiological concepts of bicarbonate transporters are highlighted in various reviews [8,9,10,11], the regulatory signals or components that act on the modulation of bicarbonate transporters are relatively unknown. This review summarizes recent advances in signaling components and the bicarbonate-associated enzymes in the modulation of bicarbonate transporter function and the understanding of bicarbonate transporters to develop drug targets against bicarbonate transporter-associated diseases.

2. Modulatory Signals in the Regulation of Bicarbonate Transporters

2.1. Inflammatory Disease-Associated Cytokines

Inflammatory cytokines increase bicarbonate secretion in epithelia [12,13]. The pH of airway epithelial surface liquid, for example, is a critical factor in respiratory defense against inhaled pathogens [14]. Alkalization of the cell surface fluid through bicarbonate secretion is associated with host defense mechanisms. Thus, the roles of bicarbonate transporters in inflammatory settings have been investigated in several studies of airway and intestinal systems, and these findings are discussed in this section.
Among the chloride–bicarbonate exchangers, downregulated in adenoma (DRA, also termed SLC26A3) is found in the gastrointestinal system and is mainly implicated in inflammatory diseases. For instance, the interleukin (IL)-10-deleted colitis mouse model and patients with ulcerative colitis revealed reduced DRA expression [15]. The downregulation of DRA leads to a defect in bicarbonate secretion and may affect the protective function against overt diarrhea [16].
In addition, inflammatory diseases, such as bowel disease, reveal the enhanced tumor necrosis factor (TNF)-α levels in the mucosa and lamina propria [17,18]. Several studies have verified the relationship between bicarbonate transporters and TNF-α. Treatment with TNF-α reduces mRNA and protein expression of DRA in colorectal adenocarcinoma HT-29 cells and Caco-2 3D-cysts model [19]. Moreover, injection of TNF-α decreases mRNA and protein expression levels of DRA compared to controls in crypt-derived mouse ileal enteroids on matrigel [19]. Overexpression of nuclear factor-kappa B (NF-κB) subunits also inhibits mRNA and protein expression levels of DRA [19]. Conversely, mRNA expression of DRA is increased by the treatment of NF-κB inhibitor caffeic acid phenethyl ester in HT-29 cells and ileal enteroids [19]. In ileal enteroid-derived mouse crypts, TNF-α decreases DRA mRNA expression, while having no effect on another CBE, Slc26a6 (mouse PAT-1) [19]. DRA knockout mice, not PAT-1 knockout, reveal dysregulated electrolyte balance; thus DRA, not PAT-1, is important in the regulation of intestinal chloride and fluid volume [20]. In ulcerative colitis (UC) patient samples, DRA expression levels are decreased, whereas serum TNF-α levels are upregulated compared to a control group [21]. Consistent with these findings, in a dextran sulfate sodium (DSS)-induced colitis mouse model, expression levels of DRA mRNA and protein are downregulated, whereas TNF-α is upregulated [21]. In DRA-overexpressed Caco2-brush border of enterocyte (Caco2/BBE) cells, TNF-α treatment induces the downregulation of DRA [21]. Knockdown of TNF-α with siRNA significantly upregulates mRNA and protein expression of DRA [21]. Conversely, knockdown of DRA with siRNA induces the upregulated mRNA and protein expression of TNF-α compared to control [21]. Thus, stability of DRA might be affected by inflammatory cytokines and DRA activity is dominantly involved in the maintenance of electrolyte homeostasis.
Additionally, it is known that anion exchanger 2 (AE2) secretes bicarbonate to protect biliary mucosa in the digestive system from related diseases such as primary biliary cholangitis (PBC) [22]. Pro-inflammatory cytokines such as IL-8, IL-12, IL-17, IL-18, and TNF-α may contribute to the reduced expression and activity of AE2 by enhancing miR-506 expression in H69 cholangiocytes [23]. Dysregulated AE2 may impair the bicarbonate umbrella against bile salt-associated cellular apoptosis [23,24].
In addition to DRA and AE2, pendrin, also termed SLC26A4, has been associated with the airway system. IL-17 treatment time-dependently enhances mRNA and protein expression of pendrin in human bronchial epithelial (HBE) cells [25]. DRA mRNA is decreased with IL-17 treatment in HBE cells [25]. Regarding DRA expression in tissues, DRA is not detected in the respiratory system [26]. In the presence of the NF-κB inhibitor JSH-23, IL-17 treatment does not increase pendrin mRNA and protein expression in HBE cells [25]. In cystic fibrosis (CF) HBE cells, pendrin mRNA expression and pHi are increased, whereas treatment with siRNA-pendrin decreases chloride/bicarbonate exchanger activity [25].
The combined administration of TNF-α and IL-17 time-dependently elevates the pH of the airway epithelial surface liquid [14]. Theoretically, the pH of airway epithelial liquid may increase as a consequence of increased bicarbonate secretion, decreased H+ secretion, or both [14]. These results indicate that TNF-α acidifies the pH of airway epithelial liquid by enhancement of H+ secretion, whereas IL-17 has no effect on H+ transport. Additionally, co-treatment with both TNF-α and IL-17 increases the pH of airway epithelial liquid through enhanced bicarbonate secretion, not by reduced H+ secretion, in human airway epithelia [14].
The combined treatment of IL-17 and TNF-α increases the mRNA expression of a gene subset of bicarbonate transporters, including CFTR, members of the SLC26A family, and the SLC4 family (NBCs) and its associated enzymes, carbonic anhydrases (CAs), in human airway epithelia [14]. We included CFTR in this review because the chloride channel CFTR is considered as the bicarbonate efflux transporter [27,28]. Concomitant addition of IL-17 and TNF-α with CFTR (inh)-172, a CFTR inhibitor, induces decreased pH of airway epithelial liquid [29]. Co-treatment of IL-17 and TNF-α with CFTR modulators (VX-445, VX-661, and VX-770) causes an increase in the pH of airway epithelial surface liquid [29]. Additionally, treatment with pendrin siRNA in the co-presence of IL-17 and TNF-α induces reduced pH of airway epithelial liquid [14]. Thus, combined treatment with IL-17 and TNF-α induces bicarbonate secretion through the involvement of CFTR and pendrin [29], alkalinizing the pH of airway epithelial liquid and providing an important barrier defense mechanism against airway inflammation (Figure 1). Understanding the effects of cytokines on the pH regulation of airway surface fluid through bicarbonate transporters would support the development of therapeutic strategies against airway inflammation.

2.2. Angiotensin II-Mediated Modulation of Bicarbonate Transport

Angiotensin II (ANG II) regulates systemic and renal circulation through the release of aldosterone in the adrenal cortex. It is known that ANG II is involved in bicarbonate reabsorption in the proximal tubule [30]. ANG II-mediated bicarbonate absorption occurs through the involvement of the bicarbonate transporter NBC in proximal tubules [30,31,32]. Wang et al. [33] demonstrated that the addition of ANG II increases fluid and bicarbonate absorption rates in the early distal tubules. However, ANG II fails to stimulate bicarbonate absorption in the late distal tubules. In an in vivo study of separated tubule segments, treatment with ANG II increased fluid absorption, while its antagonist Sarilesin decreased fluid absorption in both early and late distal tubules [33]. Thus, they found that ANG II stimulates bicarbonate and Na+ reabsorption in the proximal tubule and showed that ANG II receptors are located in both the lumen and basement membrane of proximal tubules [33]. In addition to ANG II-mediated bicarbonate absorption by NBC, ANG II stimulates sodium–hydrogen exchanger (NHE) activity in proximal and early distal tubules and cardiac muscles [30,33,34]. Although we highlighted the effect of ANG II on NBC and NHE in the renal system, the fundamental mechanism of acid-base transport and bicarbonate transporters in body system is described in several reviews [35,36].
ANG II receptors are categorized into two major subtypes: type 1 (AT1) and type 2 (AT2) receptors. ANG II-mediated calcium responses in the renal system are investigated using AT1 and AT2 antagonists, valsartan, and PD123319, respectively [37]. ANG II-mediated calcium responses are inhibited by treatment with valsartan but not PD123319 in isolated proximal tubules of wild-type (WT) mice. These results suggest that ANG II-evoked calcium responses are mediated by AT1 [37]. Additionally, isolated tubules of type A-AT1-knockout (KO) mice lose biphasic regulation of ANG II-mediated NBC activity [37]. ANG II stimulates NBC activity in adult ventricular myocytes from rats [38] and cats [39]. The hypertensive rat heart revealed impaired electrogenic NBC1 (NBCe1) activity and electroneutral type of NBC (NBCn1)-dependent pH recovery, suggesting that impaired NBCe1 activity by ANG II may be compensated by upregulated expression of NBCn1 in ventricular myocytes [40]. ANG II also mediates the production of reactive oxygen species (ROS), which activates NBC activity through the extracellular signal-regulated kinase (ERK) pathway in cat ventricular myocytes [39]. The ERK inhibitor U0126 treatment inhibits ANG II-induced equivalent acid efflux, indicating pH recovery through the involvement of NBC [39]. Furthermore, ANG II treatment inhibits the activity of NBCe1 through p38 kinase, not through an ERK1/2 pathway in cat cardiomyocytes [41]. Additionally, ANG II downregulates NBCe1 activation via p38-kinase activation, whereas NBCn1 activity is upregulated by ERK1/2- and ROS-dependent pathways in cat ventricular myocytes [40,41]. ANG II-activated AT1 receptors mediate increased calcium, alteration of the redox balance of cardiomyocytes, and subsequent cardiac hypertrophy through the activation of phospholipase C [42].
ANG II-mediated intracellular pH modulation is also involved in AE activity. ANG II treatment facilitates the reduction of myocardial intracellular pH by activating AE under conditions of an alkali load induced by trimethylamine hydrochloride (TMA), a protein kinase C (PKC) activator, treatment in cat papillary muscles [43]. Among the AE family members AE1 to AE3, full-length AE3 is involved in ANG II-mediated activation of PKC in the myocardium [44]. Additionally, the protein expression of AE1 is upregulated by a combination of the aldosterone analogue deoxycorticosterone acetate and NaHCO3 to induce alkalosis in the mouse kidney medulla, helping to maintain acid–base balance [45].
In addition, although sodium-chloride cotransporter (NCC) is not included in the bicarbonate transporter family, the NCC in plasma membrane is enhanced by ANG II treatment, and co-stimulation of aldosterone enhances pendrin expression, along with the expression of phosphorylated NCC, in adrenalectomized mice [46]. This implies that the upregulation of NCCs and pendrin by ANG II is required for the maintenance of blood pressure via the renin–ANG–aldosterone system. Furthermore, treatment with ANG II decreases CFTR mRNA and protein expression in the basilar arteries of the brain [47]. We illustrated the differential role of ANG II in Figure 2. We emphasized that CFTR is understood to be a regulator of vasoconstriction. Thus, the effect of ANG II on the modulation of bicarbonate transporters is complex and requires careful consideration, taking into account the different functions of bicarbonate transporters such as pendrin and CFTR in each organ.

2.3. Regulation of Bicarbonate Transport by Endogenous Peptides

2.3.1. Neuropeptide Vasoactive Intestinal Peptide

Vasoactive intestinal polypeptide (VIP) is a 28-amino-acid neuropeptide produced by neurons, T cells, B cells, and endocrine cells and is found in most organs [48,49]. VIP is released in response to various stimuli, including serotonin, acetylcholine, substance P, and others [49]. VIP stimulation contributes to the regulation of inflammation and smooth muscle relaxation [50]. It also has been reported that VIP activates bicarbonate transport by involving CFTR in mice with disrupted CFTR genes, established as a colony model for cystic fibrosis [51].
VIP is a physiological activator of CFTR, contributing to mucus hydration and local nonspecific immune system activity in mammalian airways [52,53]. VIP enhances CFTR activity through both protein kinase A (PKA)- and PKC-dependent molecular mechanisms [54,55,56]. Stimulation of VIP receptors (VPAC1 and VPAC2) enhances cAMP and calcium production through the G protein-coupled VPACs and subsequently activates CFTR function [54,55,57]. Accordingly, VIP is a regulator of CFTR-mediated secretion in respiratory submucosal glands [52,58]. In Calu-3 lung epithelial cells, treatment with VIP time-dependently increases CFTR expression in the apical membrane [59]. VIP-induced CFTR expression is downregulated via the treatment of PKC inhibitors bisindolylmaleimide X or chelerythrine chloride [59].
Oxidative damage caused by ozone exposure or cigarette smoke leads to an upregulation of glutathione and a downregulation of chloride secretion by reducing CFTR expression levels at the apical side in Calu-3 and T84 cells [60,61]. Ozone stress induces the downregulation of CFTR expression and activity through the signal transducer and activator of transcription 1 (STAT1) signal pathway in HBE cells [62]. VIP treatment has been shown to recover ozone stress-mediated CFTR dysfunction in HBE cells through the PKA and PKC signaling pathway [56]. Furthermore, treatment with the PKCε inhibitor peptide, Epsilon-V1-2, reduces the surface expression of CFTR in the presence of VIP in Calu-3 cells [63], suggesting that PKCε is involved in VIP-modulated CFTR membrane stability.
Sodium–hydrogen exchange regulatory factor-1 (NHERF1) is a PDZ domain-containing scaffolding protein that interacts with CFTR [64,65]. VIP exposure has been shown to enhance NHERF1 expression and co-localization with CFTR [63]. However, VIP treatment in the presence of siRNA-NHERF1 does not enhance the surface expression of CFTR compared to VIP treatment alone [63]. These results suggest that NHERF1 participates in the VIP-CFTR signaling cascade. Although VIP-mediated bicarbonate or chloride regulation is well-defined for CFTR function [66,67,68,69], the verification of regulatory interactions between VIP and other bicarbonate transporters should be investigated in the future.

2.3.2. Neuropeptide Y

Neuropeptide Y (NPY) is a 36-amino acid peptide hormone extensively expressed in submucosal neurons and the myenteric plexus throughout the gastrointestinal tract of humans [70], mice [71], and rats [72]. The NPY family includes NPY, peptide YY, and pancreatic polypeptide [73] (sequences are illustrated in Figure 3), sharing significant biological homology [74,75,76,77]. Among them, NPY is known to be a vasoconstriction peptide [78]. Since NPY is secreted simultaneously with norepinephrine (NE) in sympathetic responses, NE-induced vasoconstriction is accelerated by NPY-mediated NPY Y1 receptor activation [78].
Sakena et al. investigated the short-term regulation of NPY on CBE to understand NPY-mediated chloride and sodium transport in the gastrointestinal tract [79]. NPY treatment upregulates CBE activity in Caco2 cells [79]. Furthermore, selective agonists for NPY/Y1 or Y2 receptors increase CBE activity, while the ERK1/2 inhibitor U0126 downregulates NPY-mediated CBE activity in Caco2 cells [79]. NPY treatment also enhances DRA expression in lipid raft fractions [79]. It is generally believed that the association of membrane transporters in lipid rafts is required for optimal activity of transporters [80], and cholesterol content is crucial for the integrity of lipid rafts [81]. Treatment with the cholesterol-sequestering agent methyl-β-cyclodextrin (MβCD) reduces DRA expression in Caco2 lipid rafts, and, conversely, reduced CBE activity is restored by cholesterol addition [79].
NPY and VIP exhibit opposing functions in CFTR-associated inflammatory reactions, where NPY stimulates cytokine release while VIP reduces cytokine secretion [67]. NPY treatment downregulates VIP-mediated acidification and cellular shrinkage through the inhibition of cAMP in primary serous cells [67]. The activation of cAMP is blocked by the VIP receptor antagonist VIP (6-28), and this effect is restored by VIP treatment [67]. These findings confirm that NPY treatment inhibits CFTR-mediated chloride, bicarbonate, and fluid secretion by reducing cAMP signaling. Consequently, elevated NPY levels may contribute to reduced fluid secretion and subsequently induce airway inflammation. Although it has been reported that NPY and VIP levels are increased in human nasal polyps, inflammatory cells of ovalbumin-induced chronic asthma mice, and inflammatory bowel disease (IBD) patients, both peptides are considered immune modulators in airway diseases [82,83,84] and their sophisticated regulation and disease-specific roles should be further investigated in the future.

2.4. Intracellular Calcium-Associated Modulation of Bicarbonate Transport

Intracellular calcium homeostasis is a critical process for maintaining various physiological functions, including cardiac and neural function [85,86], and is regulated by numerous signaling proteins in a highly coordinated manner.
For instance, in a myocardial infarction mouse model, levels of NBCe1 mRNA and protein are upregulated in ventricular myocytes [87]. In cardiac-specific transgenic mice overexpressing human NBCe1, the mortality rate and infarct size are higher, and left ventricular (LV) systolic function and contractility are decreased compared to WT mice, suggesting that overexpression of NBCe1 results in dysfunction of the LV [87]. NBCe1 overexpression increases hypoxia-induced sodium and calcium signaling changes in adult mouse ventricular myocytes [87]. However, treatment with the NBC inhibitor S0859 reduces the concentration of these two ions in the setting of NBCe1 overexpression and hypoxia [87]. Overexpressed NBCe1 may be involved in the excessive sodium load and subsequent calcium overload through the sodium–calcium exchanger, which induces sodium efflux and calcium influx in cardiomyocytes [88]. Thus, sodium and calcium overload induce contractile dysfunction. Modulation of cardiac NBCe1 could be a potential strategy against cardiac contractile dysfunction and cardiac sudden death.
In addition, drugs that modulate calcium signaling play a role in the regulation of bicarbonate transporters. Treatment with the calcium ionophore 4Br-A23187 increases intracellular calcium, while it inhibits DRA activity in DRA-transfected human embryonic kidney (HEK) cells [89].
Regarding calcium signaling-associated proteins, as a calcium-associated protein, IRBIT (IP3R coupling protein released with inositol 1,4,5-trisphosphate) has been identified as a molecule that competes with IP3 for binding to the IP3 receptor [90,91] and is now recognized as a multifactorial regulator due to its wide range of target proteins such as NBCe1-B [92] and calcium/calmodulin kinase IIα [93]. IRBIT binds to NBCe1-B, one variant of NBCe1, and regulates NBC activity, playing a critical role in epithelial secretion [92,94]. The activity of NBCe1-B exhibits basal properties due to the auto-inhibitory domain (AID), which consists of residues 1-85 [95,96]. The AID of NBCe1-B is released by the binding of IRBIT through electrostatic interaction, resulting in the activation of NBCe1-B [96,97,98]. In addition, AE2 is identified as one of the binding targets of both IRBIT and long (L)-IRBIT in HEK293T cells [99]. L-IRBIT KO via CRISPR/Cas9 induces downregulated AE2 activity and expression in B16F10 melanoma cells [99]. IRBIT/L-IRBIT double KO reduces AE2 expression, whereas IRBIT overexpression restores AE2 activity in IRBIT/L-IRBIT double KO B16-F10 cells [99]. In L-IRBIT KO B16F10 cells, the lysosomal degradation inhibitor bafilomycin upregulates AE2 protein expression and recovers L-IRBIT KO-induced AE2 downregulation [99]. IRBIT and L-IRBIT possess the AHCY domain, which forms multimer. The AE2 binds to IRBIT multimer, which contains homo-multimer or hetero-multimer in HEK293T cells [99]. Although only IRBIT homo-multimer is involved in the regulation of AE2 expression, the binding of IRBIT family modulates the protein stability of AE2 through the modulation of endosome/lysosome-dependent degradation pathway.
Additionally, IRBIT antagonizes kinases such as Ste20-related proline/alanine-rich kinase (SPAK) or with-no-lysine (WNK) kinase, which destabilizes membrane expression of NBCe1-B [92]. Hwang et al. reported that IRBIT stabilizes NBCn1 expression in the plasma membrane to support cellular migration in A549 lung cancer cells [100]. Knockdown of IRBIT destabilizes plasma NBCn1 expression and reduces cellular migration, suggesting that IRBIT mediates the maintenance of a stable migratory module [100]. Although the calcium-associated protein IRBIT is related to several bicarbonate transporters such as AE2 and NBCs, identifying the regulatory relationships between IRBIT or IRBIT family and other bicarbonate transporters will be challenging in the future.

3. Host Defense Mechanisms

Bacterial prostatitis results from a bacterial infection in prostate glands, and is one of the inflammation to verify the host defense mechanism. Although we discussed the effect of inflammatory cytokines on bicarbonate transporters in Section 2.1, this section will discuss the bacterial-associated inflammation and its bicarbonate transporter regulation in detail. A common feature of prostatitis is an increase in pH, termed as an alkaline shift (>pH 8.0) [101]. Inflammatory cytokines, particularly TNF-α and IL-10, are elevated in prostatitis, and their presence has been detected in the urine of prostatitis patients [102]. These elevated inflammatory mediators induce an upregulation in CFTR expression [103]. In response to bacterial infection, prostatitis exhibits enhanced bicarbonate secretion involving CFTR, resulting in an elevation of pH [104]. The alkaline shift in pH, facilitated by bicarbonate secretion, serves as a defense mechanism against bacterial infection and inflammation. Enhanced pH of airway surface liquid plays a crucial role in respiratory host defense mechanisms, including mucociliary movement and anti-bacterial activity [105,106]. In addition, the infection of Helicobacter pylori leads to downregulated expressions and activities of duodenal CFTR and PAT-1 through transforming growth factor (TGF)-β-mediated MAPK pathway, which participates in the development of duodenal ulcers in non-tumorigenic duodenal epithelial cells from human patients [107].
Since bicarbonate has been shown to be involved in antibacterial activity, the reinforced CFTR-mediated bicarbonate secretion in prostatitis can function as a host defense mechanism [108,109,110,111]. For instance, E. coli-derived LPS treatment upregulates the expression of mRNA and protein for various inflammatory cytokines such as IL-1β, TNF-α, and IL-6 as well as the bicarbonate–associated CFTR and CA2 in prostate epithelial cells isolated from rats [104]. Inhibition of CA by acetazolamide or CFTR inhibition by CFTR (inh)-172 enhance bacterial activity in E. coli–injected prostate epithelial (PE) cells [104]. Consistently, mRNA and protein levels of CA2, CFTR, and pro-inflammatory cytokines are upregulated in E. coli–injected PE cells [104].
In IBD, decreased NaCl absorption is observed and results in IBD-associated diarrhea [112]. The absorption of NaCl involves the activity of NHE such as NHE3 and CBE such as DRA [112].
There are several studies related to anion exchange systems which involve chloride-dependent bicarbonate secretion and pH modulation. Mutations or dysregulation of DRA leads to diarrheal disorders including UC [113]. Free chloride from bicarbonate solution (containing 10 mM glucose, 25 mM NaHCO3, 115 mM Na+-gluconate, 1 mM Mg2+-gluconate, 6 mM Ca2+-gluconate, and K+-sulfate) induces alkaline intracellular pH in colonic crypts isolated from patients [113]. Additionally, DRA activity is enhanced in surface regions of distal colon crypts compared to the middle and bases [113]. However, DRA activity and mRNA expression level are downregulated, and subsequent chloride absorption is decreased in surface crypts of UC patients [112,113]. Additionally, a similar decrease in DRA mRNA expression is also observed in IL-10 KO mice after developing colonic inflammation [15]. Inflammatory diseases are attenuated via the inhibition of p-STAT signaling. For instance, all-trans retinoic acid (ATRA) treatment upregulates DRA activity and protein expression in interferon (IFN)-γ-treated Caco-2 cells [114]. Moreover, ATRA treatment downregulates IFN-γ-induced enhanced p-STAT1 expression, reducing the severity of inflammation in DSS-induced colitis mice and its isolated primary colon tissues [114].

4. Bicarbonate-Associated Carbonic Anhydrases

Carbonic anhydrases (CAs) are enzymes responsible for various biological processes such as cellular pH homeostasis, ion transport, cell respiration, and gluconeogenesis, and they function to catalyze the reversible reaction: CO2 + H2O ⇄ HCO3 + H+ [115,116]. The human CAs consist of 15 isoforms and exist in various cellular locations: there are cytosolic isoforms (CA1, CA2, CA3, CA7, and CA8), mitochondrial isoforms (CA5-A and CA5-B), a secretory isoform (CA6), and membrane-coupled isoforms (CA4, CA9, CA12, CA14, and CA15) [116,117,118,119,120,121]. CAs are an enzymatic source of bicarbonate and thus are closely related to the activity of bicarbonate transporters. CA2, for example, induces an upregulation of NBCe1 activity in Xenopus oocytes as determined by elevations in intracellular sodium concentration, and increases membrane currents and conductance [122]. The membrane current is downregulated via treatment with CA inhibitor, ethoxyzolamide, in CA1-NBCe1 and CA3-NBCe1 co-expressing oocytes [123]. Furthermore, CA1 protein injection increases NBCe1 activity and its catalytic activity in oocytes [123].
CA4 is a membrane-attached enzyme with a glycosyl-phosphatidyl-inositol anchor and is involved in bicarbonate transport in brain, lung, kidney, and muscle [118,124]. In particular, CA4 interacts with bicarbonate transporters, such as AE2 and NBCe1, to regulate intracellular pH [125]. Mechanically, CA4 catalyzes the reversible reaction to produce carbon dioxide. The carbon dioxide diffuses into the cells, or bicarbonate serves as a source of bicarbonate influx mediated by NBCe1-B, thereby participating in the bicarbonate metabolon that regulates intracellular pH [125]. The CA4 protein is widely expressed in hypoxic tumors and CA4 overexpression upregulates cell migration in Madin–Darby canine kidney (MDCK) cells with or without hepatocyte growth factor (HGF), an epithelial-mesenchymal transition inducer [126]. In addition, hypoxia induces co-localization of CA4 with AE2 and NBCe1 at the leading-edge regions of migrating SiHa and A549 cells [126], suggesting that bicarbonate transporters provide mechanical machinery for movement of cells.
CA12 is a plasma membrane-associated enzyme which can be described as an oncogenic marker such as CA9 [127,128,129]. CA12 co-localizes with NBCe1 in the cellular membrane of MDCK and HEK cells [130]. However, a mutation in CA12 (E143K) results in mis-localization of the protein to the ER in transfected MDCK or HEK cells [130]. The mutated CA12 (E143K) downregulates the activities of NBCe1 and AE2 in HeLa cells [130]. This CA12 mutation is found in patients presenting with symptoms of dry mouth and dry tongues which occur due to reduced fluid secretion [130], suggesting that downregulation of NBCe1 and AE2 via CA12 mutation are involved in reduced fluid secretion and salivation. We summarized the biological effect of CAs on the modulation of bicarbonate transporters (Table 1).
It also has been found that the conversion ratio of bicarbonate to carbon dioxide regulates the formation of neutrophil extracellular trap around inflammation foci [131]. Involvement of CA activity and bicarbonate transporter activity with conversion ability may regulate the conversion ratio through the modulation of bicarbonate content. Thus, modulation of CA or bicarbonate transporter activity on a neutrophil extracellular trap might be considered as potential strategies to attenuate inflammation foci.

5. Acidic Microenvironment and Proliferation

Hypoxic condition and production of acidic metabolites such as lactate, carbon dioxide, and proton ion are essential processes of cancer progression [5,132]. It has been known that progression of breast cancer, lung cancer, pancreatic cancer, and melanoma is associated with acidic stress and the microenvironment [100,133,134,135,136]. Although alkaline-pH-shift mechanism in urine against bacterial infection should be distinguished, acidic stress stimulates cancer cell survival and migration through the involvement of bicarbonate transporters such as NDCBE and NBCn1 [5]. NDCBE expression was also enhanced by acid loads in the rat brain [100,137]. In addition, an acidic environment in the inflamed synovium in rheumatoid arthritis (RA) triggers membrane recruitment of NBCn1 in fibroblast-like synoviocytes (FLS), mediating the aggressiveness of RA-FLS [138]. More recently, the stimulation of insulin and growth factor enhances NBCn1 protein expression in HeLa cells [139]. The mechanistic target of rapamycin complex 1 (mTORC1)-S6K1 axis stimulates the transcription of SLC4A7 mRNA [139]. Depleted Slc4a7-mice reveal reduced tumor growth in CAL-51 breast cancer cell line-implanted xenograft mice and thus recruited NBCn1 protein into the plasma membrane, and subsequent enhancement of bicarbonate influx is involved in de novo nucleotide synthesis and the cell growth of the tumor [139]. Although it has been found that mTORC1 signaling modulates NBCn1 expression in proliferating cells, future studies on the mechanism remain to be carried out for other pathophysiological states of cells and specific tissues.

6. Conclusions and Perspectives

Regulatory signals or proteins that act on the modulation of bicarbonate transporters are discussed in physiological and pathological conditions (Table 2). Inflammatory cytokines, peptides, and enzymes, mentioned in this review, are involved in the modulation of bicarbonate flux through specific bicarbonate transporters, thereby regulating cellular pH and various pathophysiological roles. Although not all members of the bicarbonate transporter family are discussed here, the large number of known isotypes provide potential research themes that need investigating in physiological or pathological conditions. In addition, although structural homology of bicarbonate transporters must be overcome, development of a specific activator or inhibitor on bicarbonate transporters will be an attractive research area. Thus, we are challenged to gain greater understanding of activating mechanisms and to identify new activators and inhibitors to modulate specific bicarbonate transporters. Thus, future additional verification of mechanisms of the modulation of bicarbonate transporters will be beneficial to obtain new paradigms in bicarbonate transporter-associated whole-body systems and diseases, such as immune, cancer, respiratory, tooth development, digestive, cardiac, and reproductive systems.

Author Contributions

H.J.K. and J.H.H. conceptualized and designed the study and acquired and interpreted information from PubMed; H.J.K. drew all figures; H.J.K. and J.H.H. drafted the manuscript; J.H.H. revised the manuscript critically for important intellectual content; J.H.H. contributed to the funding acquisition and final approval of the published version and is responsible for all aspects of the work as regards the accuracy and integrity of the review. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT; 2022R1A2C1003890: J.H.H.).

Acknowledgments

All figures were developed by the authors. The authors did not use generative AI and AI-assisted technology.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Borowitz, D. CFTR, bicarbonate, and the pathophysiology of cystic fibrosis. Pediatr. Pulmonol. 2015, 50 (Suppl. S40), S24–S30. [Google Scholar] [CrossRef] [PubMed]
  2. Boron, W.F.; Fong, P.; Hediger, M.A.; Boulpaep, E.L.; Romero, M.F. The electrogenic Na/HCO3 cotransporter. Wien. Klin. Wochenschr. 1997, 109, 445–456. [Google Scholar] [CrossRef] [PubMed]
  3. Kopito, R.R. Molecular biology of the anion exchanger gene family. Int. Rev. Cytol. 1990, 123, 177–199. [Google Scholar] [CrossRef] [PubMed]
  4. Alper, S.L.; Sharma, A.K. The SLC26 gene family of anion transporters and channels. Mol. Aspects Med. 2013, 34, 494–515. [Google Scholar] [CrossRef]
  5. Yang, O.C.Y.; Loh, S.H. Acidic Stress Triggers Sodium-Coupled Bicarbonate Transport and Promotes Survival in A375 Human Melanoma Cells. Sci. Rep. 2019, 9, 6858. [Google Scholar] [CrossRef]
  6. Hall, J.E.; Guyton, A.C. Guyton and Hall Textbook of Medical Physiology, 12th ed.; Saunders/Elsevier: Philadelphia, PA, USA, 2011; 1091p. [Google Scholar]
  7. Cordat, E.; Casey, J.R. Bicarbonate transport in cell physiology and disease. Biochem. J. 2009, 417, 423–439. [Google Scholar] [CrossRef]
  8. Hamm, L.L.; Nakhoul, N.; Hering-Smith, K.S. Acid-Base Homeostasis. Clin. J. Am. Soc. Nephrol. 2015, 10, 2232–2242. [Google Scholar] [CrossRef]
  9. Wang, H.; An, J.X.; Jin, H.; He, S.Y.; Liao, C.C.; Wang, J.; Tuo, B.G. Roles of Cl/HCO3 anion exchanger 2 in the physiology and pathophysiology of the digestive system (Review). Mol. Med. Rep. 2021, 24, 491. [Google Scholar] [CrossRef]
  10. Martinez-Crespo, L.; Valkenier, H. Transmembrane Transport of Bicarbonate by Anion Receptors. Chempluschem 2022, 87, e202200266. [Google Scholar] [CrossRef]
  11. Boedtkjer, E. Ion Channels, Transporters, and Sensors Interact with the Acidic Tumor Microenvironment to Modify Cancer Progression. Rev. Physiol. Biochem. Pharmacol. 2022, 182, 39–84. [Google Scholar] [CrossRef]
  12. Kreindler, J.L.; Bertrand, C.A.; Lee, R.J.; Karasic, T.; Aujla, S.; Pilewski, J.M.; Frizzell, R.A.; Kolls, J.K. Interleukin-17A induces bicarbonate secretion in normal human bronchial epithelial cells. Am. J. Physiol. Lung Cell Mol. Physiol. 2009, 296, L257–L266. [Google Scholar] [CrossRef] [PubMed]
  13. Gorrieri, G.; Scudieri, P.; Caci, E.; Schiavon, M.; Tomati, V.; Sirci, F.; Napolitano, F.; Carrella, D.; Gianotti, A.; Musante, I.; et al. Goblet Cell Hyperplasia Requires High Bicarbonate Transport To Support Mucin Release. Sci. Rep. 2016, 6, 36016. [Google Scholar] [CrossRef] [PubMed]
  14. Rehman, T.; Thornell, I.M.; Pezzulo, A.A.; Thurman, A.L.; Ibarra, G.S.R.; Karp, P.H.; Tan, P.; Duffey, M.E.; Welsh, M.J. TNF alpha and IL-17 alkalinize airway surface liquid through CFTR and pendrin. Am. J. Physiol.-Cell Physiol. 2020, 319, C331–C344. [Google Scholar] [CrossRef] [PubMed]
  15. Yang, H.; Jiang, W.; Furth, E.E.; Wen, X.; Katz, J.P.; Sellon, R.K.; Silberg, D.G.; Antalis, T.M.; Schweinfest, C.W.; Wu, G.D. Intestinal inflammation reduces expression of DRA, a transporter responsible for congenital chloride diarrhea. Am. J. Physiol. 1998, 275, G1445–G1453. [Google Scholar] [CrossRef] [PubMed]
  16. Xiao, F.; Juric, M.; Li, J.; Riederer, B.; Yeruva, S.; Singh, A.K.; Zheng, L.; Glage, S.; Kollias, G.; Dudeja, P.; et al. Loss of downregulated in adenoma (DRA) impairs mucosal HCO3 secretion in murine ileocolonic inflammation. Inflamm. Bowel Dis. 2012, 18, 101–111. [Google Scholar] [CrossRef] [PubMed]
  17. Friedrich, M.; Pohin, M.; Powrie, F. Cytokine Networks in the Pathophysiology of Inflammatory Bowel Disease. Immunity 2019, 50, 992–1006. [Google Scholar] [CrossRef] [PubMed]
  18. Geremia, A.; Biancheri, P.; Allan, P.; Corazza, G.R.; Di Sabatino, A. Innate and adaptive immunity in inflammatory bowel disease. Autoimmun. Rev. 2014, 13, 3–10. [Google Scholar] [CrossRef]
  19. Kumar, A.; Chatterjee, I.; Gujral, T.; Alakkam, A.; Coffing, H.; Anbazhagan, A.N.; Borthakur, A.; Saksena, S.; Gill, R.K.; Alrefai, W.A.; et al. Activation of Nuclear Factor-κB by Tumor Necrosis Factor in Intestinal Epithelial Cells and Mouse Intestinal Epithelia Reduces Expression of the Chloride Transporter SLC26A3. Gastroenterology 2017, 153, 1338–1350.e3. [Google Scholar] [CrossRef]
  20. Schweinfest, C.W.; Spyropoulos, D.D.; Henderson, K.W.; Kim, J.H.; Chapman, J.M.; Barone, S.; Worrell, R.T.; Wang, Z.; Soleimani, M. slc26a3 (dra)-deficient mice display chloride-losing diarrhea, enhanced colonic proliferation, and distinct up-regulation of ion transporters in the colon. J. Biol. Chem. 2006, 281, 37962–37971. [Google Scholar] [CrossRef]
  21. Ding, X.M.; Li, D.X.; Li, M.K.; Tian, D.; Yu, H.B.; Yu, Q. Tumor necrosis factor-α acts reciprocally with solute carrier family 26, member 3, (downregulated-in-adenoma) and reduces its expression, leading to intestinal inflammation. Int. J. Mol. Med. 2018, 41, 1224–1232. [Google Scholar] [CrossRef]
  22. Sasaki, M.; Sato, Y.; Nakanuma, Y. An impaired biliary bicarbonate umbrella may be involved in dysregulated autophagy in primary biliary cholangitis. Lab. Investig. 2018, 98, 745–754. [Google Scholar] [CrossRef] [PubMed]
  23. Erice, O.; Munoz-Garrido, P.; Vaquero, J.; Perugorria, M.J.; Fernandez-Barrena, M.G.; Saez, E.; Santos-Laso, A.; Arbelaiz, A.; Jimenez-Aguero, R.; Fernandez-Irigoyen, J.; et al. MicroRNA-506 promotes primary biliary cholangitis-like features in cholangiocytes and immune activation. Hepatology 2018, 67, 1420–1440. [Google Scholar] [CrossRef] [PubMed]
  24. Chang, J.C.; Go, S.; de Waart, D.R.; Munoz-Garrido, P.; Beuers, U.; Paulusma, C.C.; Oude Elferink, R. Soluble Adenylyl Cyclase Regulates Bile Salt-Induced Apoptosis in Human Cholangiocytes. Hepatology 2016, 64, 522–534. [Google Scholar] [CrossRef] [PubMed]
  25. Adams, K.M.; Abraham, V.; Spielman, D.; Kolls, J.K.; Rubenstein, R.C.; Conner, G.E.; Cohen, N.A.; Kreindler, J.L. IL-17A induces Pendrin expression and chloride-bicarbonate exchange in human bronchial epithelial cells. PLoS ONE 2014, 9, e103263. [Google Scholar] [CrossRef] [PubMed]
  26. Wheat, V.J.; Shumaker, H.; Burnham, C.; Shull, G.E.; Yankaskas, J.R.; Soleimani, M. CFTR induces the expression of DRA along with Cl/HCO3 exchange activity in tracheal epithelial cells. Am. J. Physiol. Cell Physiol. 2000, 279, C62–C71. [Google Scholar] [CrossRef]
  27. Poulsen, J.H.; Fischer, H.; Illek, B.; Machen, T.E. Bicarbonate conductance and pH regulatory capability of cystic fibrosis transmembrane conductance regulator. Proc. Natl. Acad. Sci. USA 1994, 91, 5340–5344. [Google Scholar] [CrossRef]
  28. Garnett, J.P.; Hickman, E.; Burrows, R.; Hegyi, P.; Tiszlavicz, L.; Cuthbert, A.W.; Fong, P.; Gray, M.A. Novel role for pendrin in orchestrating bicarbonate secretion in cystic fibrosis transmembrane conductance regulator (CFTR)-expressing airway serous cells. J. Biol. Chem. 2011, 286, 41069–41082. [Google Scholar] [CrossRef]
  29. Rehman, T.; Karp, P.H.; Tan, P.; Goodell, B.J.; Pezzulo, A.A.; Thurman, A.L.; Thornell, I.M.; Durfey, S.L.; Duffey, M.E.; Stoltz, D.A.; et al. Inflammatory cytokines TNF-α and IL-17 enhance the efficacy of cystic fibrosis transmembrane conductance regulator modulators. J. Clin. Investig. 2021, 131, e150398. [Google Scholar] [CrossRef]
  30. Geibel, J.; Giebisch, G.; Boron, W.F. Angiotensin II stimulates both Na(+)-H+ exchange and Na+/HCO3 cotransport in the rabbit proximal tubule. Proc. Natl. Acad. Sci. USA 1990, 87, 7917–7920. [Google Scholar] [CrossRef]
  31. Romero, M.F.; Hopfer, U.; Madhun, Z.T.; Zhou, W.; Douglas, J.G. Angiotensin II actions in the rabbit proximal tubule. Angiotensin II mediated signaling mechanisms and electrolyte transport in the rabbit proximal tubule. Ren. Physiol. Biochem. 1991, 14, 199–207. [Google Scholar] [CrossRef]
  32. Coppola, S.; Fromter, E. An electrophysiological study of angiotensin II regulation of Na-HCO3 cotransport and K conductance in renal proximal tubules. I. Effect of picomolar concentrations. Pflugers Arch. 1994, 427, 143–150. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, T.; Giebisch, G. Effects of angiotensin II on electrolyte transport in the early and late distal tubule in rat kidney. Am. J. Physiol. 1996, 271, F143–F149. [Google Scholar] [CrossRef] [PubMed]
  34. Kohout, T.A.; Rogers, T.B. Angiotensin-Ii Activates the Na+/HCO3- Symport through a Phosphoinositide-Independent Mechanism in Cardiac-Cells. J. Biol. Chem. 1995, 270, 20432–20438. [Google Scholar] [CrossRef] [PubMed]
  35. Skelton, L.A.; Boron, W.F.; Zhou, Y. Acid-base transport by the renal proximal tubule. J. Nephrol. 2010, 23 (Suppl. S16), S4–S18. [Google Scholar] [PubMed]
  36. Romero, M.F.; Chen, A.P.; Parker, M.D.; Boron, W.F. The SLC4 family of bicarbonate (HCO3) transporters. Mol. Aspects Med. 2013, 34, 159–182. [Google Scholar] [CrossRef]
  37. Horita, S.; Zheng, Y.; Hara, C.; Yamada, H.; Kunimi, M.; Taniguchi, S.; Uwatoko, S.; Sugaya, T.; Goto, A.; Fujita, T.; et al. Biphasic regulation of Na+-HCO3 cotransporter by angiotensin II type 1A receptor. Hypertension 2002, 40, 707–712. [Google Scholar] [CrossRef]
  38. Baetz, D.; Haworth, R.S.; Avkiran, M.; Feuvray, D. The ERK pathway regulates Na+-HCO3 cotransport activity in adult rat cardiomyocytes. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H2102–H2109. [Google Scholar] [CrossRef]
  39. De Giusti, V.C.; Garciarena, C.D.; Aiello, E.A. Role of reactive oxygen species (ROS) in angiotensin II-induced stimulation of the cardiac Na+/HCO3 cotransport. J. Mol. Cell. Cardiol. 2009, 47, 716–722. [Google Scholar] [CrossRef]
  40. Orlowski, A.; Ciancio, M.C.; Caldiz, C.I.; De Giusti, V.C.; Aiello, E.A. Reduced sarcolemmal expression and function of the NBCe1 isoform of the NaHCO3 cotransporter in hypertrophied cardiomyocytes of spontaneously hypertensive rats: Role of the reninangiotensin system. Cardiovasc. Res. 2014, 101, 211–219. [Google Scholar] [CrossRef]
  41. De Giusti, V.C.; Orlowski, A.; Aiello, E.A. Angiotensin II inhibits the electrogenic Na+/HCO3 cotransport of cat cardiac myocytes. J. Mol. Cell. Cardiol. 2010, 49, 812–818. [Google Scholar] [CrossRef]
  42. Bhullar, S.K.; Dhalla, N.S. Angiotensin II-Induced Signal Transduction Mechanisms for Cardiac Hypertrophy. Cells 2022, 11, 3336. [Google Scholar] [CrossRef] [PubMed]
  43. Camilion de Hurtado, M.C.; Alvarez, B.V.; Perez, N.G.; Ennis, I.L.; Cingolani, H.E. Angiotensin II activates Na+-independent Cl-HCO3 exchange in ventricular myocardium. Circ. Res. 1998, 82, 473–481. [Google Scholar] [CrossRef] [PubMed]
  44. Alvarez, B.V.; Fujinaga, J.; Casey, J.R. Molecular basis for angiotensin II-induced increase of chloride/bicarbonate exchange in the myocardium. Circ. Res. 2001, 89, 1246–1253. [Google Scholar] [CrossRef] [PubMed]
  45. Mohebbi, N.; Perna, A.; van der Wijst, J.; Becker, H.M.; Capasso, G.; Wagner, C.A. Regulation of two renal chloride transporters, AE1 and pendrin, by electrolytes and aldosterone. PLoS ONE 2013, 8, e55286. [Google Scholar] [CrossRef]
  46. Hirohama, D.; Ayuzawa, N.; Ueda, K.; Nishimoto, M.; Kawarazaki, W.; Watanabe, A.; Shimosawa, T.; Marumo, T.; Shibata, S.; Fujita, T. Aldosterone Is Essential for Angiotensin II-Induced Upregulation of Pendrin. J. Am. Soc. Nephrol. 2018, 29, 57–68. [Google Scholar] [CrossRef]
  47. Zhao, L.; Yuan, F.; Pan, N.; Yu, Y.; Yang, H.; Liu, Y.; Wang, R.; Zhang, B.; Wang, G. CFTR deficiency aggravates Ang II induced vasoconstriction and hypertension by regulating Ca2+ influx and RhoA/Rock pathway in VSMCs. Front. Biosci. 2021, 26, 1396–1410. [Google Scholar] [CrossRef]
  48. Henning, R.J.; Sawmiller, D.R. Vasoactive intestinal peptide: Cardiovascular effects. Cardiovasc. Res. 2001, 49, 27–37. [Google Scholar] [CrossRef]
  49. Iwasaki, M.; Akiba, Y.; Kaunitz, J.D. Recent advances in vasoactive intestinal peptide physiology and pathophysiology: Focus on the gastrointestinal system. F1000Research 2019, 8, 1629. [Google Scholar] [CrossRef]
  50. Leceta, J.; Gomariz, R.P.; Martinez, C.; Carrion, M.; Arranz, A.; Juarranz, Y. Vasoactive intestinal peptide regulates Th17 function in autoimmune inflammation. Neuroimmunomodulation 2007, 14, 134–138. [Google Scholar] [CrossRef]
  51. Hogan, D.L.; Crombie, D.L.; Isenberg, J.I.; Svendsen, P.; Schaffalitzky de Muckadell, O.B.; Ainsworth, M.A. Acid-stimulated duodenal bicarbonate secretion involves a CFTR-mediated transport pathway in mice. Gastroenterology 1997, 113, 533–541. [Google Scholar] [CrossRef]
  52. Wine, J.J.; Joo, N.S. Submucosal glands and airway defense. Proc. Am. Thorac. Soc. 2004, 1, 47–53. [Google Scholar] [CrossRef] [PubMed]
  53. Maggi, C.A.; Giachetti, A.; Dey, R.D.; Said, S.I. Neuropeptides as Regulators of Airway Function—Vasoactive-Intestinal-Peptide and the Tachykinins. Physiol. Rev. 1995, 75, 277–322. [Google Scholar] [CrossRef] [PubMed]
  54. Derand, R.; Montoni, A.; Bulteau-Pignoux, L.; Janet, T.; Moreau, B.; Muller, J.M.; Becq, F. Activation of VPAC1 receptors by VIP and PACAP-27 in human bronchial epithelial cells induces CFTR-dependent chloride secretion. Br. J. Pharmacol. 2004, 141, 698–708. [Google Scholar] [CrossRef] [PubMed]
  55. Laburthe, M.; Couvineau, A.; Tan, V. Class II G protein-coupled receptors for VIP and PACAP: Structure, models of activation and pharmacology. Peptides 2007, 28, 1631–1639. [Google Scholar] [CrossRef] [PubMed]
  56. Qu, F.; Liu, H.J.; Xiang, Y.; Tan, Y.R.; Liu, C.; Zhu, X.L.; Qin, X.Q. Activation of CFTR trafficking and gating by vasoactive intestinal peptide in human bronchial epithelial cells. J. Cell. Biochem. 2011, 112, 902–908. [Google Scholar] [CrossRef] [PubMed]
  57. Rumalla, K.; Yarbrough, C.K.; Pugely, A.J.; Koester, L.; Dorward, I.G. Spinal fusion for pediatric neuromuscular scoliosis: National trends, complications, and in-hospital outcomes. J. Neurosurg. Spine 2016, 25, 500–508. [Google Scholar] [CrossRef] [PubMed]
  58. Ianowski, J.P.; Choi, J.Y.; Wine, J.J.; Hanrahan, J.W. Mucus secretion by single tracheal submucosal glands from normal and cystic fibrosis transmembrane conductance regulator knockout mice. J. Physiol. 2007, 580, 301–314. [Google Scholar] [CrossRef]
  59. Chappe, F.; Loewen, M.E.; Hanrahan, J.W.; Chappe, V. Vasoactive intestinal peptide increases cystic fibrosis transmembrane conductance regulator levels in the apical membrane of Calu-3 cells through a protein kinase C-dependent mechanism. J. Pharmacol. Exp. Ther. 2008, 327, 226–238. [Google Scholar] [CrossRef]
  60. Cantin, A.M.; Bilodeau, G.; Ouellet, C.; Liao, J.; Hanrahan, J.W. Oxidant stress suppresses CFTR expression. Am. J. Physiol. Cell Physiol. 2006, 290, C262–C270. [Google Scholar] [CrossRef]
  61. Cantin, A.M.; Hanrahan, J.W.; Bilodeau, G.; Ellis, L.; Dupuis, A.; Liao, J.; Zielenski, J.; Durie, P. Cystic fibrosis transmembrane conductance regulator function is suppressed in cigarette smokers. Am. J. Respir. Crit. Care 2006, 173, 1139–1144. [Google Scholar] [CrossRef]
  62. Qu, F.; Qin, X.Q.; Cui, Y.R.; Xiang, Y.; Tan, Y.R.; Liu, H.J.; Peng, L.H.; Zhou, X.Y.; Liu, C.; Zhu, X.L. Ozone stress down-regulates the expression of cystic fibrosis transmembrane conductance regulator in human bronchial epithelial cells. Chem. Biol. Interact. 2009, 179, 219–226. [Google Scholar] [CrossRef] [PubMed]
  63. Alshafie, W.; Chappe, F.G.; Li, M.S.; Anini, Y.; Chappe, V.M. VIP regulates CFTR membrane expression and function in Calu-3 cells by increasing its interaction with NHERF1 and P-ERM in a VPAC1-and PKC ε-dependent manner. Am. J. Physiol.-Cell Physiol. 2014, 307, C107–C119. [Google Scholar] [CrossRef] [PubMed]
  64. Short, D.B.; Trotter, K.W.; Reczek, D.; Kreda, S.M.; Bretscher, A.; Boucher, R.C.; Stutts, M.J.; Milgram, S.L. An apical PDZ protein anchors the cystic fibrosis transmembrane conductance regulator to the cytoskeleton. J. Biol. Chem. 1998, 273, 19797–19801. [Google Scholar] [CrossRef] [PubMed]
  65. Cheng, J.; Moyer, B.D.; Milewski, M.; Loffing, J.; Ikeda, M.; Mickle, J.E.; Cutting, G.R.; Li, M.; Stanton, B.A.; Guggino, W.B. A Golgi-associated PDZ domain protein modulates cystic fibrosis transmembrane regulator plasma membrane expression. J. Biol. Chem. 2002, 277, 3520–3529. [Google Scholar] [CrossRef] [PubMed]
  66. Yu, B.; Zhu, X.; Yang, X.; Jin, L.; Xu, J.; Ma, T.; Yang, H. Plumbagin Prevents Secretory Diarrhea by Inhibiting CaCC and CFTR Channel Activities. Front. Pharmacol. 2019, 10, 1181. [Google Scholar] [CrossRef] [PubMed]
  67. McMahon, D.B.; Carey, R.M.; Kohanski, M.A.; Tong, C.C.L.; Papagiannopoulos, P.; Adappa, N.D.; Palmer, J.N.; Lee, R.J. Neuropeptide regulation of secretion and inflammation in human airway gland serous cells. Eur. Respir. J. 2020, 55, 1901386. [Google Scholar] [CrossRef] [PubMed]
  68. Oak, A.A.; Chhetri, P.D.; Rivera, A.A.; Verkman, A.S.; Cil, O. Repurposing calcium-sensing receptor agonist cinacalcet for treatment of CFTR-mediated secretory diarrheas. JCI Insight 2021, 6, e146823. [Google Scholar] [CrossRef]
  69. Rodrat, M.; Wongdee, K.; Teerapornpuntakit, J.; Thongbunchoo, J.; Tanramluk, D.; Aeimlapa, R.; Thammayon, N.; Thonapan, N.; Wattano, P.; Charoenphandhu, N. Vasoactive intestinal peptide and cystic fibrosis transmembrane conductance regulator contribute to the transepithelial calcium transport across intestinal epithelium-like Caco-2 monolayer. PLoS ONE 2022, 17, e0277096. [Google Scholar] [CrossRef]
  70. Nichols, K.; Staines, W.; Krantis, A. Neural sites of the human colon colocalize nitric oxide synthase-related NADPH diaphorase activity and neuropeptide Y. Gastroenterology 1994, 107, 968–975. [Google Scholar] [CrossRef]
  71. Sandgren, K.; Larsson, L.T.; Ekblad, E. Widespread changes in neurotransmitter expression and number of enteric neurons and interstitial cells of Cajal in lethal spotted mice—An explanation for persisting dysmotility after operation for Hirschsprung’s disease? Dig. Dis. Sci. 2002, 47, 1049–1064. [Google Scholar] [CrossRef]
  72. Ekblad, E.; Ekman, R.; Hakanson, R.; Sundler, F. Projections of peptide-containing neurons in rat colon. Neuroscience 1988, 27, 655–674. [Google Scholar] [CrossRef] [PubMed]
  73. Holzer, P.; Reichmann, F.; Farzi, A. Neuropeptide Y, peptide YY and pancreatic polypeptide in the gut-brain axis. Neuropeptides 2012, 46, 261–274. [Google Scholar] [CrossRef] [PubMed]
  74. Savage, A.P.; Adrian, T.E.; Carolan, G.; Chatterjee, V.K.; Bloom, S.R. Effects of peptide YY (PYY) on mouth to caecum intestinal transit time and on the rate of gastric emptying in healthy volunteers. Gut 1987, 28, 166–170. [Google Scholar] [CrossRef] [PubMed]
  75. Floyd, J.C., Jr.; Fajans, S.S.; Pek, S.; Chance, R.E. A newly recognized pancreatic polypeptide; plasma levels in health and disease. Recent. Prog. Horm. Res. 1976, 33, 519–570. [Google Scholar] [CrossRef] [PubMed]
  76. Tatemoto, K. Neuropeptide Y: Complete amino acid sequence of the brain peptide. Proc. Natl. Acad. Sci. USA 1982, 79, 5485–5489. [Google Scholar] [CrossRef] [PubMed]
  77. Tatemoto, K. Isolation and characterization of peptide YY (PYY), a candidate gut hormone that inhibits pancreatic exocrine secretion. Proc. Natl. Acad. Sci. USA 1982, 79, 2514–2518. [Google Scholar] [CrossRef] [PubMed]
  78. Tan, C.M.J.; Green, P.; Tapoulal, N.; Lewandowski, A.J.; Leeson, P.; Herring, N. The Role of Neuropeptide Y in Cardiovascular Health and Disease. Front. Physiol. 2018, 9, 1281. [Google Scholar] [CrossRef] [PubMed]
  79. Saksena, S.; Tyagi, S.; Goyal, S.; Gill, R.K.; Alrefai, W.A.; Ramaswamy, K.; Dudeja, P.K. Stimulation of apical Cl/HCO3(OH) exchanger, SLC26A3 by neuropeptide Y is lipid raft dependent. Am. J. Physiol.-Gastrointest. Liver Physiol. 2010, 299, G1334–G1343. [Google Scholar] [CrossRef]
  80. Magnani, F.; Tate, C.G.; Wynne, S.; Williams, C.; Haase, J. Partitioning of the serotonin transporter into lipid microdomains modulates transport of serotonin. J. Biol. Chem. 2004, 279, 38770–38778. [Google Scholar] [CrossRef]
  81. Simons, K.; Ikonen, E. Cell biology—How cells handle cholesterol. Science 2000, 290, 1721–1726. [Google Scholar] [CrossRef]
  82. Fang, S.Y.; Shen, C.L.; Ohyama, M. Presence of neuropeptides in human nasal polyps. Acta Otolaryngol. 1994, 114, 324–328. [Google Scholar] [CrossRef] [PubMed]
  83. Makinde, T.O.; Steininger, R.; Agrawal, D.K. NPY and NPY receptors in airway structural and inflammatory cells in allergic asthma. Exp. Mol. Pathol. 2013, 94, 45–50. [Google Scholar] [CrossRef] [PubMed]
  84. Chandrasekharan, B.; Nezami, B.G.; Srinivasan, S. Emerging neuropeptide targets in inflammation: NPY and VIP. Am. J. Physiol. Gastrointest. Liver Physiol. 2013, 304, G949–G957. [Google Scholar] [CrossRef] [PubMed]
  85. Barry, W.H.; Bridge, J.H. Intracellular calcium homeostasis in cardiac myocytes. Circulation 1993, 87, 1806–1815. [Google Scholar] [CrossRef] [PubMed]
  86. Verkhratsky, A.; Orkand, R.K.; Kettenmann, H. Glial calcium: Homeostasis and signaling function. Physiol. Rev. 1998, 78, 99–141. [Google Scholar] [CrossRef]
  87. Chen, Z.H.; Chen, L.; Chen, K.T.; Lin, H.R.; Shen, M.J.; Chen, L.; Zhu, H.L.; Zhu, Y.Q.; Wang, Q.C.; Xi, F.; et al. Overexpression of Na+-HCO3- cotransporter contributes to the exacerbation of cardiac remodeling in mice with myocardial infarction by increasing intracellular calcium overload. BBA-Mol. Basis Dis. 2020, 1866, 165623. [Google Scholar] [CrossRef]
  88. Ji, M.J.; Son, K.H.; Hong, J.H. Addition of oh8dG to Cardioplegia Attenuated Myocardial Oxidative Injury through the Inhibition of Sodium Bicarbonate Cotransporter Activity. Antioxidants 2022, 11, 1641. [Google Scholar] [CrossRef]
  89. Lamprecht, G.; Hsieh, C.J.; Lissner, S.; Nold, L.; Heil, A.; Gaco, V.; Schafer, J.; Turner, J.R.; Gregor, M. Intestinal Anion Exchanger Down-regulated in Adenoma (DRA) Is Inhibited by Intracellular Calcium. J. Biol. Chem. 2009, 284, 19744–19753. [Google Scholar] [CrossRef]
  90. Ando, H.; Mizutani, A.; Matsu-ura, T.; Mikoshiba, K. IRBIT, a novel inositol 1,4,5-trisphosphate (IP3) receptor-binding protein, is released from the IP3 receptor upon IP3 binding to the receptor. J. Biol. Chem. 2003, 278, 10602–10612. [Google Scholar] [CrossRef]
  91. Ando, H.; Mizutani, A.; Kiefer, H.; Tsuzurugi, D.; Michikawa, T.; Mikoshiba, K. IRBIT suppresses IP3 receptor activity by competing with IP3 for the common binding site on the IP3 receptor. Mol. Cell 2006, 22, 795–806. [Google Scholar] [CrossRef]
  92. Yang, D.; Li, Q.; So, I.; Huang, C.L.; Ando, H.; Mizutani, A.; Seki, G.; Mikoshiba, K.; Thomas, P.J.; Muallem, S. IRBIT governs epithelial secretion in mice by antagonizing the WNK/SPAK kinase pathway. J. Clin. Investig. 2011, 121, 956–965. [Google Scholar] [CrossRef] [PubMed]
  93. Kawaai, K.; Mizutani, A.; Shoji, H.; Ogawa, N.; Ebisui, E.; Kuroda, Y.; Wakana, S.; Miyakawa, T.; Hisatsune, C.; Mikoshiba, K. IRBIT regulates CaMKII α activity and contributes to catecholamine homeostasis through tyrosine hydroxylase phosphorylation. Proc. Natl. Acad. Sci. USA 2015, 112, 5515–5520. [Google Scholar] [CrossRef] [PubMed]
  94. Shirakabe, K.; Priori, G.; Yamada, H.; Ando, H.; Horita, S.; Fujita, T.; Fujimoto, I.; Mizutani, A.; Seki, G.; Mikoshiba, K. IRBIT, an inositol 1,4,5-trisphosphate receptor-binding protein, specifically binds to and activates pancreas-type Na+/HCO3 cotransporter 1 (pNBC1). Proc. Natl. Acad. Sci. USA 2006, 103, 9542–9547. [Google Scholar] [CrossRef] [PubMed]
  95. Lee, S.K.; Boron, W.F.; Parker, M.D. Relief of autoinhibition of the electrogenic Na-HCO3 cotransporter NBCe1-B: Role of IRBIT vs.amino-terminal truncation. Am. J. Physiol. Cell Physiol. 2012, 302, C518–C526. [Google Scholar] [CrossRef] [PubMed]
  96. Hong, J.H.; Yang, D.; Shcheynikov, N.; Ohana, E.; Shin, D.M.; Muallem, S. Convergence of IRBIT, phosphatidylinositol (4,5) bisphosphate, and WNK/SPAK kinases in regulation of the Na+-HCO3 cotransporters family. Proc. Natl. Acad. Sci. USA 2013, 110, 4105–4110. [Google Scholar] [CrossRef]
  97. Su, P.; Wu, H.; Wang, M.; Cai, L.; Liu, Y.; Chen, L.M. IRBIT activates NBCe1-B by releasing the auto-inhibition module from the transmembrane domain. J. Physiol. 2021, 599, 1151–1172. [Google Scholar] [CrossRef]
  98. Wang, M.; Wu, H.; Liu, Y.; Chen, L.M. Activation of mouse NBCe1-B by Xenopus laevis and mouse IRBITs: Role of the variable Nt appendage of IRBITs. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183240. [Google Scholar] [CrossRef]
  99. Itoh, R.; Hatano, N.; Murakami, M.; Mitsumori, K.; Kawasaki, S.; Wakagi, T.; Kanzaki, Y.; Kojima, H.; Kawaai, K.; Mikoshiba, K.; et al. Both IRBIT and long-IRBIT bind to and coordinately regulate Cl/HCO3 exchanger AE2 activity through modulating the lysosomal degradation of AE2. Sci. Rep. 2021, 11, 5990. [Google Scholar] [CrossRef]
  100. Hwang, S.; Shin, D.M.; Hong, J.H. Protective Role of IRBIT on Sodium Bicarbonate Cotransporter-n1 for Migratory Cancer Cells. Pharmaceutics 2020, 12, 816. [Google Scholar] [CrossRef]
  101. Blacklock, N.J.; Beavis, J.P. The response of prostatic fluid pH in inflammation. Br. J. Urol. 1974, 46, 537–542. [Google Scholar] [CrossRef]
  102. He, L.; Wang, Y.; Long, Z.; Jiang, C. Clinical significance of IL-2, IL-10, and TNF-α in prostatic secretion of patients with chronic prostatitis. Urology 2010, 75, 654–657. [Google Scholar] [CrossRef] [PubMed]
  103. Cafferata, E.G.; Gonzalez-Guerrico, A.M.; Giordano, L.; Pivetta, O.H.; Santa-Coloma, T.A. Interleukin-1β regulates CFTR expression in human intestinal T84 cells. Biochim. Biophys. Acta 2000, 1500, 241–248. [Google Scholar] [CrossRef] [PubMed]
  104. Xie, C.; Tang, X.; Xu, W.; Diao, R.; Cai, Z.; Chan, H.C. A host defense mechanism involving CFTR-mediated bicarbonate secretion in bacterial prostatitis. PLoS ONE 2010, 5, e15255. [Google Scholar] [CrossRef] [PubMed]
  105. Tang, X.X.; Ostedgaard, L.S.; Hoegger, M.J.; Moninger, T.O.; Karp, P.H.; McMenimen, J.D.; Choudhury, B.; Varki, A.; Stoltz, D.A.; Welsh, M.J. Acidic pH increases airway surface liquid viscosity in cystic fibrosis. J. Clin. Investig. 2016, 126, 879–891. [Google Scholar] [CrossRef] [PubMed]
  106. Simonin, J.; Bille, E.; Crambert, G.; Noel, S.; Dreano, E.; Edwards, A.; Hatton, A.; Pranke, I.; Villeret, B.; Cottart, C.H.; et al. Airway surface liquid acidification initiates host defense abnormalities in Cystic Fibrosis. Sci. Rep. 2019, 9, 6516. [Google Scholar] [CrossRef] [PubMed]
  107. Wen, G.R.; Deng, S.L.; Song, W.F.; Jin, H.; Xu, J.Y.; Liu, X.; Xie, R.M.; Song, P.H.; Tuo, B.G. Helicobacter pylori infection downregulates duodenal CFTR and SLC26A6 expressions through TGF beta signaling pathway. Bmc Microbiol. 2018, 18, 87. [Google Scholar] [CrossRef]
  108. Thompson, K.D.; Welykyj, S.; Massa, M.C. Antibacterial activity of lidocaine in combination with a bicarbonate buffer. J. Dermatol. Surg. Oncol. 1993, 19, 216–220. [Google Scholar] [CrossRef]
  109. Drake, D.R.; Vargas, K.; Cardenzana, A.; Srikantha, R. Enhanced bactericidal activity of Arm and Hammer Dental Care. Am. J. Dent. 1995, 8, 308–312. [Google Scholar]
  110. Craig, S.B.; Concannon, M.J.; McDonald, G.A.; Puckett, C.L. The antibacterial effects of tumescent liposuction fluid. Plast. Reconstr. Surg. 1999, 103, 666–670. [Google Scholar] [CrossRef]
  111. Jarvis, G.N.; Fields, M.W.; Adamovich, D.A.; Arthurs, C.E.; Russell, J.B. The mechanism of carbonate killing of Escherichia coli. Lett. Appl. Microbiol. 2001, 33, 196–200. [Google Scholar] [CrossRef]
  112. Priyamvada, S.; Gomes, R.; Gill, R.K.; Saksena, S.; Alrefai, W.A.; Dudeja, P.K. Mechanisms Underlying Dysregulation of Electrolyte Absorption in Inflammatory Bowel Disease-Associated Diarrhea. Inflamm. Bowel Dis. 2015, 21, 2926–2935. [Google Scholar] [CrossRef] [PubMed]
  113. Farkas, K.; Yeruva, S.; Rakonczay, Z., Jr.; Ludolph, L.; Molnar, T.; Nagy, F.; Szepes, Z.; Schnur, A.; Wittmann, T.; Hubricht, J.; et al. New therapeutic targets in ulcerative colitis: The importance of ion transporters in the human colon. Inflamm. Bowel Dis. 2011, 17, 884–898. [Google Scholar] [CrossRef] [PubMed]
  114. Priyamvada, S.; Anbazhagan, A.N.; Kumar, A.; Chatterjee, I.; Borthakur, A.; Saksena, S.; Gill, R.K.; Alrefai, W.A.; Dudeja, P.K. All-trans Retinoic Acid Counteracts Diarrhea and Inhibition of Downregulated in Adenoma Expression in Gut Inflammation. Inflamm. Bowel Dis. 2020, 26, 534–545. [Google Scholar] [CrossRef] [PubMed]
  115. Supuran, C.T. Carbonic anhydrases--an overview. Curr. Pharm. Des. 2008, 14, 603–614. [Google Scholar] [CrossRef] [PubMed]
  116. Tashian, R.E. The carbonic anhydrases: Widening perspectives on their evolution, expression and function. Bioessays 1989, 10, 186–192. [Google Scholar] [CrossRef] [PubMed]
  117. Sly, W.S.; Hu, P.Y. Human carbonic anhydrases and carbonic anhydrase deficiencies. Annu. Rev. Biochem. 1995, 64, 375–401. [Google Scholar] [CrossRef] [PubMed]
  118. Supuran, C.T. Carbonic anhydrases: Novel therapeutic applications for inhibitors and activators. Nat. Rev. Drug Discov. 2008, 7, 168–181. [Google Scholar] [CrossRef]
  119. Supuran, C.T.; Scozzafava, A.; Conway, J. Carbonic Anhydrase: Its Inhibitors and Activators; CRC Press: Boca Raton, FL, USA, 2004; pp. 363–364. [Google Scholar]
  120. Imtaiyaz Hassan, M.; Shajee, B.; Waheed, A.; Ahmad, F.; Sly, W.S. Structure, function and applications of carbonic anhydrase isozymes. Bioorg Med. Chem. 2013, 21, 1570–1582. [Google Scholar] [CrossRef]
  121. Waheed, A.; Zhu, X.L.; Sly, W.S. Membrane-associated carbonic anhydrase from rat lung. Purification, characterization, tissue distribution, and comparison with carbonic anhydrase IVs of other mammals. J. Biol. Chem. 1992, 267, 3308–3311. [Google Scholar] [CrossRef]
  122. Becker, H.M.; Deitmer, J.W. Carbonic anhydrase II increases the activity of the human electrogenic Na+/HCO3 cotransporter. J. Biol. Chem. 2007, 282, 13508–13521. [Google Scholar] [CrossRef]
  123. Schueler, C.; Becker, H.M.; McKenna, R.; Deitmer, J.W. Transport activity of the sodium bicarbonate cotransporter NBCe1 is enhanced by different isoforms of carbonic anhydrase. PLoS ONE 2011, 6, e27167. [Google Scholar] [CrossRef] [PubMed]
  124. Waheed, A.; Sly, W.S. Membrane associated carbonic anhydrase IV (CA IV): A personal and historical perspective. Subcell. Biochem. 2014, 75, 157–179. [Google Scholar] [CrossRef] [PubMed]
  125. Morgan, P.E.; Pastorekova, S.; Stuart-Tilley, A.K.; Alper, S.L.; Casey, J.R. Interactions of transmembrane carbonic anhydrase, CAIX, with bicarbonate transporters. Am. J. Physiol. Cell Physiol. 2007, 293, C738–C748. [Google Scholar] [CrossRef]
  126. Svastova, E.; Witarski, W.; Csaderova, L.; Kosik, I.; Skvarkova, L.; Hulikova, A.; Zatovicova, M.; Barathova, M.; Kopacek, J.; Pastorek, J.; et al. Carbonic anhydrase IX interacts with bicarbonate transporters in lamellipodia and increases cell migration via its catalytic domain. J. Biol. Chem. 2012, 287, 3392–3402. [Google Scholar] [CrossRef] [PubMed]
  127. Hynninen, P.; Vaskivuo, L.; Saarnio, J.; Haapasalo, H.; Kivela, J.; Pastorekova, S.; Pastorek, J.; Waheed, A.; Sly, W.S.; Puistola, U.; et al. Expression of transmembrane carbonic anhydrases IX and XII in ovarian tumours. Histopathology 2006, 49, 594–602. [Google Scholar] [CrossRef] [PubMed]
  128. Kopecka, J.; Campia, I.; Jacobs, A.; Frei, A.P.; Ghigo, D.; Wollscheid, B.; Riganti, C. Carbonic anhydrase XII is a new therapeutic target to overcome chemoresistance in cancer cells. Oncotarget 2015, 6, 6776–6793. [Google Scholar] [CrossRef] [PubMed]
  129. Waheed, A.; Sly, W.S. Carbonic anhydrase XII functions in health and disease. Gene 2017, 623, 33–40. [Google Scholar] [CrossRef] [PubMed]
  130. Hong, J.H.; Muhammad, E.; Zheng, C.; Hershkovitz, E.; Alkrinawi, S.; Loewenthal, N.; Parvari, R.; Muallem, S. Essential role of carbonic anhydrase XII in secretory gland fluid and HCO3 secretion revealed by disease causing human mutation. J. Physiol. 2015, 593, 5299–5312. [Google Scholar] [CrossRef]
  131. Maueroder, C.; Mahajan, A.; Paulus, S.; Gosswein, S.; Hahn, J.; Kienhofer, D.; Biermann, M.H.; Tripal, P.; Friedrich, R.P.; Munoz, L.E.; et al. Menage-a-Trois: The Ratio of Bicarbonate to CO2 and the pH Regulate the Capacity of Neutrophils to Form NETs. Front. Immunol. 2016, 7, 583. [Google Scholar] [CrossRef]
  132. Ady, J.W.; Desir, S.; Thayanithy, V.; Vogel, R.I.; Moreira, A.L.; Downey, R.J.; Fong, Y.; Manova-Todorova, K.; Moore, M.A.; Lou, E. Intercellular communication in malignant pleural mesothelioma: Properties of tunneling nanotubes. Front. Physiol. 2014, 5, 400. [Google Scholar] [CrossRef]
  133. Stock, C.; Gassner, B.; Hauck, C.R.; Arnold, H.; Mally, S.; Eble, J.A.; Dieterich, P.; Schwab, A. Migration of human melanoma cells depends on extracellular pH and Na+/H+ exchange. J. Physiol. 2005, 567, 225–238. [Google Scholar] [CrossRef] [PubMed]
  134. Chen, W.; Zhong, R.; Ming, J.; Zou, L.; Zhu, B.B.; Lu, X.Z.; Ke, J.T.; Zhang, Y.; Liu, L.; Miao, X.; et al. The SLC4A7 variant rs4973768 is associated with breast cancer risk: Evidence from a case-control study and a meta-analysis. Breast Cancer Res. Treat. 2012, 136, 847–857. [Google Scholar] [CrossRef] [PubMed]
  135. Lee, S.; Axelsen, T.V.; Jesse, N.; Pedersen, S.F.; Vahi, P.; Boedtkjer, E. Na+,HCO3-cotransporter NBCn1 (Slc4a7) accelerates ErbB2-induced breast cancer development and tumor growth in mice. Oncogene 2018, 37, 5569–5584. [Google Scholar] [CrossRef] [PubMed]
  136. Cappellesso, F.; Orban, M.P.; Shirgaonkar, N.; Berardi, E.; Serneels, J.; Neveu, M.A.; Di Molfetta, D.; Piccapane, F.; Caroppo, R.; Debellis, L.; et al. Targeting the bicarbonate transporter SLC4A4 overcomes immunosuppression and immunotherapy resistance in pancreatic cancer. Nat. Cancer 2022, 3, 1464–1483. [Google Scholar] [CrossRef] [PubMed]
  137. Lee, H.J.; Park, H.J.; Lee, S.; Kim, Y.H.; Choi, I. The sodium-driven chloride/bicarbonate exchanger NDCBE in rat brain is upregulated by chronic metabolic acidosis. Brain Res. 2011, 1377, 13–20. [Google Scholar] [CrossRef]
  138. Ji, M.; Ryu, H.J.; Baek, H.M.; Shin, D.M.; Hong, J.H. Dynamic synovial fibroblasts are modulated by NBCn1 as a potential target in rheumatoid arthritis. Exp. Mol. Med. 2022, 54, 503–517. [Google Scholar] [CrossRef]
  139. Ali, E.S.; Liponska, A.; O’Hara, B.P.; Amici, D.R.; Torno, M.D.; Gao, P.; Asara, J.M.; Yap, M.F.; Mendillo, M.L.; Ben-Sahra, I. The mTORC1-SLC4A7 axis stimulates bicarbonate import to enhance de novo nucleotide synthesis. Mol. Cell. 2022, 82, 3284–3298.e3287. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of inflammatory cytokine-mediated ion transporters in airway epithelia. Bicarbonate secretion by airway epithelial transporters plays a crucial role in the barrier defense mechanism against inflammation by enhancing the pH of airway epithelial fluid. The presence of TNF and IL-17 in airway epithelial fluid enhances the efficacy of CFTR and pendrin function [29]. CFTR: cystic fibrosis transmembrane conductance regulator, CAs: carbonic anhydrases (enhanced membrane-associated CA9 and CA12) [14].
Figure 1. Schematic illustration of inflammatory cytokine-mediated ion transporters in airway epithelia. Bicarbonate secretion by airway epithelial transporters plays a crucial role in the barrier defense mechanism against inflammation by enhancing the pH of airway epithelial fluid. The presence of TNF and IL-17 in airway epithelial fluid enhances the efficacy of CFTR and pendrin function [29]. CFTR: cystic fibrosis transmembrane conductance regulator, CAs: carbonic anhydrases (enhanced membrane-associated CA9 and CA12) [14].
Pharmaceutics 16 00078 g001
Figure 2. Schematic illustration of angiotensin II-mediated bicarbonate transporter regulation. ANG II represents differential roles in NBCs in ventricular myocytes [40,41]. There are differential effects of ANG II on pendrin/NCC and CFTR in the kidney and basilar arteries of the brain, respectively [46,47]. Black and red arrow indicates down- or up-regulated protein expression, respectively. ANG II: angiotensin II, ERK1/2: extracellular signal-regulated kinase1/2, ROS: reactive oxygen species, N: nucleus, NCCs: sodium chloride cotransporter, CFTR: cystic fibrosis transmembrane conductance regulator.
Figure 2. Schematic illustration of angiotensin II-mediated bicarbonate transporter regulation. ANG II represents differential roles in NBCs in ventricular myocytes [40,41]. There are differential effects of ANG II on pendrin/NCC and CFTR in the kidney and basilar arteries of the brain, respectively [46,47]. Black and red arrow indicates down- or up-regulated protein expression, respectively. ANG II: angiotensin II, ERK1/2: extracellular signal-regulated kinase1/2, ROS: reactive oxygen species, N: nucleus, NCCs: sodium chloride cotransporter, CFTR: cystic fibrosis transmembrane conductance regulator.
Pharmaceutics 16 00078 g002
Figure 3. Amino acid sequence of NPY family members (NPY, peptide YY, and pancreatic polypeptide). The NPY family members share significant biological homology (red color) [75,76,77]. The multiple sequence alignment was performed with a CLUSTALW.
Figure 3. Amino acid sequence of NPY family members (NPY, peptide YY, and pancreatic polypeptide). The NPY family members share significant biological homology (red color) [75,76,77]. The multiple sequence alignment was performed with a CLUSTALW.
Pharmaceutics 16 00078 g003
Table 1. Biological effects of CAs on bicarbonate transporters.
Table 1. Biological effects of CAs on bicarbonate transporters.
Isoform TypeTransportersRelated MechanismsSpeciesRef
CA1
(Microinjection)
NBCe1Increased NBC activity and catalytic activityOocytes[123]
CA2
(Microinjection)
NBCe1Elevated NBC activityOocytes[122]
CA2
(Transfection)
AE2
NBCe1
Upregulated AE2 activities
No effect on NBC activity
HeLa cells[130]
CA4
(Transfection)
AE2, NBCe1Regulation of intracellular pHHEK293T[125]
CA12
(Transfection)
NBCe1, AE2Upregulated NBC and AE2 activitiesHeLa cells[130]
CA12 (E143K)
(Transfection)
NBCe1, AE2Downregulated NBC and AE2 activities, reduced fluid secretionHeLa and primary parotid gland cells
Table 2. Various stimulants of bicarbonate transporters.
Table 2. Various stimulants of bicarbonate transporters.
TypeBTIonsExpression TissuesRelated DiseasesStimulantsRef
CBE; SLC26A familyDRA
(SLC26A3)
Cl−,
HCO3
Colon, ileal, and intestineBowel diseases, InflammatoryInflammatory cytokines (IL-17 and TNF-α), Caffeic acid phenethyl ester, NPY, MβCD, 4Br-A23187, and ATRA[19,21,25,79,89,114]
Pendrin
(SLC26A4)
Respiratory tract, kidney Inflammatory, Dysregulation of airway surface liquid pH TNF-α, Ang II, IL-17 [14,25,46]
PAT-1DuodenalH. pylori-associated
duodenal ulcer
H. pylori[107]
CBE; Anion exchanger familyAE1Cl,
HCO3
KidneyMaintenance of electrolyte and acid-base disturbancesAcidic stress, alkaloid stress[45]
AE2Skin, kidney, and liverMelanoma, Primary biliary cholangitis.L-IRBIT K/O, IRBIT/L-IRBIT double K/O, and pro-inflammatory cytokines (IL-8, 12, 17, 18, and TNF-α)[23,99]
AE3HeartCardiac hypertrophyAng II [44]
NBCNBCe1Na+,
HCO3
Heart, airway, pancreatic ducts, parotid ducts, and oocytesHypertrophy, heart failureIL-17 + TNF-α, ANG II, IRBIT,
CA1, and CA2
[14,41,92,122,123]
NBCn1Airway, heart, lung, and cervixHypertrophy, heart failure, carcinoma, adenocarcinomaIL-17 + TNF- α, ANG II, IRBIT,
Insulin, and growth factor
[14,41,100,139]
NDCBECl,
HCO3
Brain, skinChronic metabolic acidosis, melanomaAcidic stress[5,137]
ATP-binding cassette transporterCFTRCl,
HCO3
Duodenal, airway, brain, lung, and prostateH. pylori-associated duodenal ulcer, hypertension, adenocarcinomaH. pylori, TNF-a + IL17, ANG II, VIP, LPS, E. coli[14,47,59,104,107]
Abbreviations: CBE, chloride–bicarbonate exchanger; SLC26A, solute carrier family 26; DRA, downregulated in adenoma; PAT-1, putative anion transporter 1; IL, interleukin; TNF-α, tumor necrosis factor α; NPY, neuropeptide Y; MβCD, methyl-β-cyclodextrin; ATRA, all-trans retinoic acid; Ang II, angiotensin II; AE, anion exchanger; L-IRBIT, long-IP3R coupling protein released with inositol 1,4,5-trisphosphate; IRBIT, IP3R coupling protein released with inositol 1,4,5-trisphosphate; NBC, sodium-bicarbonate cotransporter; NBCe1, electrogenic sodium bicarbonate cotransporter1; NBCn1, electroneutral sodium bicarbonate cotransporter1; NDCBE, sodium-dependent chloride-bicarbonate exchanger; CA, carbonic anhydrase; CFTR, cystic fibrosis transmembrane conductance; VIP, vasoactive intestinal peptide; LPS, lipopolysaccharide; H. pylori, Helicobacter pylori; E. coli, Escherichia coli.
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

Kim, H.J.; Hong, J.H. Multiple Regulatory Signals and Components in the Modulation of Bicarbonate Transporters. Pharmaceutics 2024, 16, 78. https://doi.org/10.3390/pharmaceutics16010078

AMA Style

Kim HJ, Hong JH. Multiple Regulatory Signals and Components in the Modulation of Bicarbonate Transporters. Pharmaceutics. 2024; 16(1):78. https://doi.org/10.3390/pharmaceutics16010078

Chicago/Turabian Style

Kim, Hyeong Jae, and Jeong Hee Hong. 2024. "Multiple Regulatory Signals and Components in the Modulation of Bicarbonate Transporters" Pharmaceutics 16, no. 1: 78. https://doi.org/10.3390/pharmaceutics16010078

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