Next Article in Journal
Therapeutic Delivery of Soluble Fractalkine Ameliorates Vascular Dysfunction in the Diabetic Retina
Next Article in Special Issue
K+-Driven Cl/HCO3 Exchange Mediated by Slc4a8 and Slc4a10
Previous Article in Journal
Metabolic Profiling to Assess Response to Targeted and Immune Therapy in Melanoma
Previous Article in Special Issue
A Comparative Study on the Lysosomal Cation Channel TMEM175 Using Automated Whole-Cell Patch-Clamp, Lysosomal Patch-Clamp, and Solid Supported Membrane-Based Electrophysiology: Functional Characterization and High-Throughput Screening Assay Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

The Role of Ion-Transporting Proteins in Human Disease

by
Yoshinori Marunaka
1,2,3
1
Medical Research Institute, Kyoto Industrial Health Association, 67 Kitatsuboi-cho, Nishinokyo, Nakagyo-ku, Kyoto 604-8472, Japan
2
Research Organization of Science and Technology, Ritsumeikan University, Kusatsu 525-8577, Japan
3
Graduate School of Medical Science, Kyoto Prefectural University of Medicine, Kyoto 602-8566, Japan
Int. J. Mol. Sci. 2024, 25(3), 1726; https://doi.org/10.3390/ijms25031726
Submission received: 16 January 2024 / Accepted: 22 January 2024 / Published: 31 January 2024
(This article belongs to the Special Issue The Role of Ion-Transporting Proteins in Human Disease)
This Special Issue focuses on the significance of ion-transporting proteins, such as ion channels and transporters, providing evidence for their significant contribution to bodily and cellular functions via the regulation of signal transduction and ionic environments [1,2,3,4,5]. In the late 1980s, the cystic fibrosis transmembrane regulator (CFTR) gene and its most common cystic fibrosis (CF)-causing mutation (ΔF508) were discovered [6]. Subsequently, many other genes which encode proteins from ion-transporting channels and transporters have been found to be associated with human disease. Even though the discovery of genes which are associated with human diseases marks a major advance within the field, the development of therapies to treat such channelopathies and transportopathies has not yet been fully accomplished.
CF patients are well known for showing deficient water secretion in the lungs, the liver, the sinuses, the small and large intestines, pancreatic and hepatobiliary ducts and male reproductive tracts [7,8,9,10,11,12]. One of the most serious problems in patients with CF is death from pneumonia caused by impaired fluid secretion into the airway cavity, which leads to dry lungs and an increased susceptibility to infection [8]. A gene responsible for one of the most serious problems in CF patients, impaired fluid secretion into the airway cavity, has been cloned by a research group in Toronto, Canada [6], and the gene has been called the cystic fibrosis transmembrane regulator (CFTR) [6]. This discovery has clearly pointed out that CF is a genetic disease, and various types of CFTR mutations have been classified from a functional perspective [13,14,15,16]: e.g., G542, ΔF508, G551D, R117H, A445E [13] (Figure 1), 3849 + 10kbC > T and C1400X [14]. The most common CFTR mutation is ΔF508 (loss of phenylalanine at position 508), which leads to a deficiency in the amount of protein being trafficked to the cell membrane [14,15,16]. Thus, ΔF508 CFTR has no role in the secretion of Cl into the airway lumen [14,15,16]. Deficient Cl secretion into the airway lumen, resulting in impaired water secretion, leads to dryness in the airway lumen and causes bacterial and viral infections. Similar CFTR dysfunction is not only observed in the lung but is also present in other organs such as the liver, the sinuses, the intestine, the colon, hepatobiliary and pancreatic ducts and male reproductive tracts [8,9,10,11]. CFTR modulator therapy is currently implemented by enhancing CFTR function through two mechanisms [17]. Potentiators such as ivacaftor increase the probability of the channel being open, allowing Cl and HCO3 to pass across the cell membrane and through the channel more easily. Correctors such as lumacaftor, tezacaftor and elexacaftor improve the number of CFTR Cl channels on the cell surface by helping proteins to fold properly, allowing the translocation of CFTR Cl channels to the cell membrane surface. Severe variants such as ΔF508 required both potentiators and correctors to ameliorate the quantity and function of channels at the membrane surface of cells. Currently, four modulators have been approved by drug regulatory agencies in Europe and the USA, and the indication for each therapeutic agent depends on the specific CFTR genetic variants present [18]. Further, Cl channels play a variety of important roles in bodily and cellular functions, such as regulation of cell volume [19,20] (Figure 2) and cell migration [21] (Figure 3).
Epithelial Na+ reabsorption in the renal collecting ducts contributes to blood pressure regulation by controlling the volume of bodily fluids [28]. Amiloride-sensitive epithelial Na+ channels in the renal epithelia contribute to Na+ reabsorption, and the mutation of amiloride epithelial Na+ channels causes disorders which affect blood pressure control. The amiloride-sensitive epithelial Na+ channel (ENaC) has been cloned by a research group in Lausanne, Switzerland, as it has been recognized as a causative gene for Liddle syndrome [29,30]. Each ENaC subunit consists of two transmembrane domains with intracellular N and C termini, and a large extracellular domain. The ion selectivity filter can specifically discriminate Na+, and the filter is located in the middle of the transmembrane domains [31]. The extracellular domain contains protease cleavage sites that enable its inhibitory effects on ENaC to be eliminated; it plays a key role in regulating ENaC activation [32,33]. The N-terminal ubiquitylation of the α and γ subunits has been implicated in the endocytosis and degradation of ENaC [34]; both the HGxxR sequence in the N-terminal and the PPPxY sequence in the C-terminal participate in regulating the ENaC. Mutations in the HGxxR sequence or the PPPxY sequence lead to the ENaC functioning abnormally, and are associated with the occurrence of Liddle syndrome [35] and pseudohypoaldosteronism (PHA) [36], i.e., the truncation of β or γ ENaC causes Liddle syndrome by elevating the amount of Na+ reabsorption in the kidney by increasing the number of functional ENaC located at the plasma membrane [36,37,38,39,40,41].
In addition to hypertension, ENaC plays important roles in human diseases [28] (Figure 4). In recent years, the ENaC has been found to contribute to immune cell activation, cystic fibrosis, endothelial cell dysfunction, pseudohypoaldosteronism (PHA), aggravated inflammation involved in high salt-induced hypertension, tumors and taste dysfunction [42]. ENaC hyperfunction elevates the concentration of intracellular Na+ ([Na+]i), leading to an intracellular-Ca2+ overload due to the activation of the Na+/Ca2+ exchanger; the overloaded intracellular Ca2+ is a key factor in ENaC-related inflammation [28].
The pH of interstitial fluid also plays an important role in various bodily and cellular functions [43,44,45,46,47]; lowering the pH of the interstitial fluid causes insulin resistance, and increases the accumulation of amyloid β, a candidate causative factor in Alzheimer’s Disease. The interstitial fluid’s pH is controlled by various types of ion transporters [48] (Figure 5), e.g., anion exchanger (AE), monocarboxylate transporter (MCT), Na+-HCO3 cotransporter (NBC), Na+-driven Cl/HCO3 exchanger (NDCBE), Na+/H+ exchanger (NHE), H+-ATPase (H+ pump). Variations in the interstitial fluid’s pH are primarily affected by glucose metabolism performed via the anaerobic process mediated through the glycolytic pathway followed by the aerobic one through the Krebs (TCA) cycle [49,50,51]. The interstitial fluid’s pH-buffer capacity is very low compared with that of blood and/or the intracellular space [52,53,54,55,56,57,58,59,60] (Figure 5A). Therefore, the interstitial fluid’s pH changes more easily compared with the pH of blood and/or the intracellular space under metabolic disordered conditions [52,53,54,55,56,57,58,59,60] (Figure 5B). The pH of the fluid around enzymes and their substrates (proteins) can affect the binding affinity of enzymes to their substrates (proteins), influencing the activity of the enzyme by altering the protein tertiary structure [61]. For example, a lowered interstitial fluid pH diminishes the insulin’s binding affinity to its receptor, leading to insulin resistance [62] (Figure 5B).
The lowered interstitial fluid pH enhances the accumulation of amyloid-β which is observed in patients suffering with Alzheimer’s disease, and leads to hyperphosphorylation of the tau protein causing neural loss, neural inflammation and synaptic impairment, which are associated with behavioral abnormalities and cognitive decline [63,64,65,66,67,68,69,70]. Patients with type 2 diabetes are at high risk of developing Alzheimer’s disease [71,72,73,74,75,76]. Lowered interstitial fluid pH, an acidic condition, activates the β- and γ-secretases involved in the formation of amyloid-β from the amyloid precursor protein [77,78,79,80,81,82,83,84]. Thus, the acidic conditions which occur in type 2 diabetes patients increases the production of amyloid-β by activating the β- and γ-secretases [68,69,70,77]. Therefore, the interstitial fluid’s pH should be maintained within the normal range by various ion transporting proteins and elevating the pH-buffer capacity to ensure the maintenance of healthy body conditions [48,85,86] (Figure 5).
Thus, ion-transporting proteins’ dysfunction play important roles in human diseases. This Special Issue aims to provide insights into recent advances in the function and structure of ion-transporting proteins as they relate to human disease and the molecular mechanisms that cause ion-transporting-protein dysfunction. Finally, I would like to conclude my editorial by presenting the essence of the content of the articles in this Special Issue. Characteristics of the lysosomal cation channel TMEM175, a Parkinson’s disease-related protein and a promising drug target, is reported using a modified whole-cell patch clamp technique applied to lysosome [87]. The advantages of this new technique are described in this article [87]. NHE plays an important role in extrusion of H+ produced in the intracellular space [48] (Figure 5). Salari et al. [88] have provided a detailed report of the expression sites of NHE in the colon, indicating the significance of NHE expression. Connexins, unique hemichannels, are transmembrane proteins which form gap junctions in vertebrates, and allow cell–cell or/and paracrine communication by releasing ATP, glutamate and NAD+, thus regulating processes such as synaptic transmission and cell migration [89]. The Na+,K+-ATPase (Na+,K+-pump) maintains critical cellular functions by creating fundamental Na+ and K+ environments in the intracellular space. For example, the Na+,K+-ATPase creates ionic environments for the maintenance of voltage-dependent excitatory membrane function in nerve and muscle cells, the secondary active transport of Na+-coupled glucose and amino acids into epithelial cells and the intracellular concentration of Ca2+ in heart myocardia [90]. The review by Baloglu [90] focuses on the regulation of Na+,K+-ATPases in ischemic heart disease and discusses the regulation of Na+,K+-ATPases under conditions of myocardial stress and their therapeutic potential based on the perspective of hypoxia-inducible factors. A review about ENaC is also published in this Special Issue [91]. The significance of ENaC has been described above in the early sections of this editorial. This review [91] focuses on discussing some recent developments in the search for novel therapeutic agents. I hope the articles published in this Special Issue provide researchers with new insights into the roles of ion-transporting proteins in human diseases and their potential as therapeutics for human diseases.

Funding

This research was funded by Grants-in-Aid for Scientific Research (B) from the Japan Society of the Promotion of Science (JSPS KAKENHI Grant Number JP18H03182 and JP21H03368) to Yoshinori Marunaka.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Levring, J.; Terry, D.S.; Kilic, Z.; Fitzgerald, G.; Blanchard, S.C.; Chen, J. CFTR function, pathology and pharmacology at single-molecule resolution. Nature 2023, 616, 606–614. [Google Scholar] [CrossRef]
  2. Oka, Y.; Butnaru, M.; von Buchholtz, L.; Ryba, N.J.; Zuker, C.S. High salt recruits aversive taste pathways. Nature 2013, 494, 472–475. [Google Scholar] [CrossRef]
  3. Tang, T.; Jian, B.; Liu, Z. Transmembrane protein 175, a lysosomal ion channel related to Parkinson’s disease. Biomolecules 2023, 13, 802. [Google Scholar] [CrossRef]
  4. Kato, H.E.; Inoue, K.; Abe-Yoshizumi, R.; Kato, Y.; Ono, H.; Konno, M.; Hososhima, S.; Ishizuka, T.; Hoque, M.R.; Kunitomo, H.; et al. Structural basis for Na+ transport mechanism by a light-driven Na+ pump. Nature 2015, 521, 48–53. [Google Scholar] [CrossRef]
  5. Bernardazzi, C.; Sheikh, I.A.; Xu, H.; Ghishan, F.K. The physiological function and potential role of the ubiquitous Na+/H+ exchanger Isoform 8 (NHE8): An overview Data. Int. J. Mol. Sci. 2022, 23, 10857. [Google Scholar] [CrossRef]
  6. Riordan, J.R.; Rommens, J.M.; Kerem, B.; Alon, N.; Rozmahel, R.; Grzelczak, Z.; Zielenski, J.; Lok, S.; Plavsic, N.; Chou, J.L.; et al. Identification of the cystic fibrosis gene: Cloning and characterization of complementary DNA. Science 1989, 245, 1066–1073. [Google Scholar] [CrossRef]
  7. Welsh, M.J.; Smith, A.E. Molecular mechanisms of CFTR chloride channel dysfunction in cystic fibrosis. Cell 1993, 73, 1251–1254. [Google Scholar] [CrossRef]
  8. Cabrini, G.; Rimessi, A.; Borgatti, M.; Pinton, P.; Gambari, R. Overview of CF lung pathophysiology. Curr. Opin. Pharmacol. 2022, 64, 102214. [Google Scholar] [CrossRef]
  9. Putman, M.S.; Norris, A.W.; Hull, R.L.; Rickels, M.R.; Sussel, L.; Blackman, S.M.; Chan, C.L.; Ode, K.L.; Daley, T.; Stecenko, A.A.; et al. Cystic fibrosis-related diabetes workshop: Research priorities spanning disease pathophysiology, diagnosis, and outcomes. Diabetes 2023, 72, 677–689. [Google Scholar] [CrossRef]
  10. Costaguta, G.; Patey, N.; Álvarez, F. Cystic fibrosis liver disease in children—A review of our current understanding. Arch. Argent Pediatr. 2023, 121, e202202905. [Google Scholar]
  11. Raymond, C.M.; Gaul, S.P.; Han, S.; Huang, G.; Dong, J. Variability of clinical presentation in patients heterozygous for the F508del cystic fibrosis variant: A series of three cases and a review of the literature. Cureus 2023, 15, e40185. [Google Scholar] [CrossRef] [PubMed]
  12. Mall, M.A.; Mayer-Hamblett, N.; Rowe, S.M. Cystic fibrosis: Emergence of highly effective targeted therapeutics and potential clinical implications. Am. J. Respir. Crit. Care Med. 2020, 201, 1193–1208. [Google Scholar] [CrossRef] [PubMed]
  13. Koivula, F.N.; McClenaghan, N.H.; Harper, A.G.; Kelly, C. Islet-intrinsic effects of CFTR mutation. Diabetologia 2016, 59, 1350–1355. [Google Scholar] [CrossRef]
  14. Ong, T.; Ramsey, B.W. Cystic Fibrosis: A review. JAMA 2023, 329, 1859–1871. [Google Scholar] [CrossRef] [PubMed]
  15. Fontes, G.; Ghislain, J.; Benterki, I.; Zarrouki, B.; Trudel, D.; Berthiaume, Y.; Poitout, V. The ΔF508 mutation in the cystic fibrosis transmembrane conductance regulator is associated with progressive insulin resistance and decreased functional beta-cell mass in mice. Diabetes 2015, 64, 4112–4122. [Google Scholar] [CrossRef] [PubMed]
  16. Marunaka, Y. The mechanistic links between insulin and cystic fibrosis transmembrane conductance regulator (CFTR) Cl channel. Int. J. Mol. Sci. 2017, 18, 1767. [Google Scholar] [CrossRef]
  17. Veit, G.; Avramescu, R.G.; Chiang, A.N.; Houck, S.A.; Cai, Z.; Peters, K.W.; Hong, J.S.; Pollard, H.B.; Guggino, W.B.; Balch, W.E.; et al. From CFTR biology toward combinatorial pharmacotherapy: Expanded classification of cystic fibrosis mutations. Mol. Biol. Cell 2016, 27, 424–433. [Google Scholar] [CrossRef]
  18. Ramsey, B.W.; Davies, J.; McElvaney, N.G.; Tullis, E.; Bell, S.C.; Dřevínek, P.; Griese, M.; McKone, E.F.; Wainwright, C.E.; Konstan, M.W.; et al. A CFTR potentiator in patients with cystic fibrosis and the G551D mutation. N. Engl. J. Med. 2011, 365, 1663–1672. [Google Scholar] [CrossRef]
  19. Sinha, M.; Zabini, D.; Guntur, D.; Nagaraj, C.; Enyedi, P.; Olschewski, H.; Kuebler, W.M.; Olschewski, A. Chloride channels in the lung: Challenges and perspectives for viral infections, pulmonary arterial hypertension, and cystic fibrosis. Pharmacol. Ther. 2022, 237, 108249. [Google Scholar] [CrossRef]
  20. Okada, Y.; Okada, T.; Sato-Numata, K.; Islam, M.R.; Ando-Akatsuka, Y.; Numata, T.; Kubo, M.; Shimizu, T.; Kurbannazarova, R.S.; Marunaka, Y.; et al. Cell volume-activated and volume-correlated anion channels in mammalian cells: Their biophysical, molecular, and pharmacological properties. Pharmacol. Rev. 2019, 71, 49–88. [Google Scholar] [CrossRef]
  21. Marunaka, Y. Physiological roles of chloride ions in bodily and cellular functions. J. Physiol. Sci. 2023, 73, 31. [Google Scholar] [CrossRef] [PubMed]
  22. Shekarabi, M.; Zhang, J.; Khanna, A.R.; Ellison, D.H.; Delpire, E.; Kahle, K.T. WNK kinase signaling in ion homeostasis and human disease. Cell Metab. 2017, 25, 285–299. [Google Scholar] [CrossRef] [PubMed]
  23. Köchl, R.; Thelen, F.; Vanes, L.; Brazão, T.F.; Fountain, K.; Xie, J.; Huang, C.L.; Lyck, R.; Stein, J.V.; Tybulewicz, V.L. WNK1 kinase balances T cell adhesion versus migration in vivo. Nat. Immunol. 2016, 17, 1075–1083. [Google Scholar] [CrossRef] [PubMed]
  24. Garzon-Muvdi, T.; Schiapparelli, P.; ap Rhys, C.; Guerrero-Cazares, H.; Smith, C.; Kim, D.H.; Kone, L.; Farber, H.; Lee, D.Y.; An, S.S.; et al. Regulation of brain tumor dispersal by NKCC1 through a novel role in focal adhesion regulation. PLoS Biol. 2012, 10, e1001320. [Google Scholar] [CrossRef]
  25. Haas, B.R.; Cuddapah, V.A.; Watkins, S.; Rohn, K.J.; Dy, T.E.; Sontheimer, H. With-No-Lysine Kinase 3 (WNK3) stimulates glioma invasion by regulating cell volume. Am. J. Physiol. Cell Physiol. 2011, 301, C1150–C1160. [Google Scholar] [CrossRef]
  26. Schwab, A.; Fabian, A.; Hanley, P.J.; Stock, C. Role of ion channels and transporters in cell migration. Physiol. Rev. 2012, 92, 1865–1913. [Google Scholar] [CrossRef] [PubMed]
  27. Carmona, B.; Marinho, H.S.; Matos, C.L.; Nolasco, S.; Soares, H. Tubulin post-translational modifications: The elusive roles of acetylation. Biology 2023, 12, 561. [Google Scholar] [CrossRef]
  28. Chen, Y.; Yu, X.; Yan, Z.; Zhang, S.; Zhang, J.; Guo, W. Role of epithelial sodium channel-related inflammation in human diseases. Front. Immunol. 2023, 14, 1178410. [Google Scholar] [CrossRef]
  29. Canessa, C.M.; Horisberger, J.D.; Rossier, B.C. Epithelial sodium channel related to proteins involved in neurodegeneration. Nature 1993, 361, 467–470. [Google Scholar] [CrossRef]
  30. Canessa, C.M.; Schild, L.; Buell, G.; Thorens, B.; Gautschi, I.; Horisberger, J.D.; Rossier, B.C. Amiloride-sensitive epithelial Na+ channel is made of three homologous subunits. Nature 1994, 367, 463–467. [Google Scholar] [CrossRef]
  31. Baconguis, I.; Bohlen, C.J.; Goehring, A.; Julius, D.; Gouaux, E. X-ray structure of acid-sensing ion channel 1-snake toxin complex reveals open state of a Na+-selective channel. Cell 2014, 156, 717–729. [Google Scholar] [CrossRef]
  32. Kleyman, T.R.; Kashlan, O.B.; Hughey, R.P. Epithelial Na+ channel Rregulation by extracellular and intracellular factors. Annu. Rev. Physiol. 2018, 80, 263–281. [Google Scholar] [CrossRef]
  33. Zhang, L.; Wang, X.; Chen, J.; Kleyman, T.R.; Sheng, S. Accessibility of ENaC extracellular domain central core residues. J. Biol. Chem. 2022, 298, 101860. [Google Scholar] [CrossRef]
  34. Staub, O.; Gautschi, I.; Ishikawa, T.; Breitschopf, K.; Ciechanover, A.; Schild, L.; Rotin, D. Regulation of stability and function of the epithelial Na+ channel (ENaC) by ubiquitination. EMBO J. 1997, 16, 6325–6336. [Google Scholar] [CrossRef]
  35. Shimkets, R.A.; Warnock, D.G.; Bositis, C.M.; Nelson-Williams, C.; Hansson, J.H.; Schambelan, M.; Gill, J.R., Jr.; Ulick, S.; Milora, R.V.; Findling, J.W.; et al. Liddle’s syndrome: Heritable human hypertension caused by mutations in the beta subunit of the epithelial sodium channel. Cell 1994, 79, 407–414. [Google Scholar] [CrossRef]
  36. Pradervand, S.; Barker, P.M.; Wang, Q.; Ernst, S.A.; Beermann, F.; Grubb, B.R.; Burnier, M.; Schmidt, A.; Bindels, R.J.; Gatzy, J.T.; et al. Salt restriction induces pseudohypoaldosteronism type 1 in mice expressing low levels of the beta-subunit of the amiloride-sensitive epithelial sodium channel. Proc. Natl. Acad. Sci. USA 1999, 96, 1732–1737. [Google Scholar] [CrossRef] [PubMed]
  37. Hansson, J.H.; Nelson-Williams, C.; Suzuki, H.; Schild, L.; Shimkets, R.; Lu, Y.; Canessa, C.; Iwasaki, T.; Rossier, B.; Lifton, R.P. Hypertension caused by a truncated epithelial sodium channel gamma subunit: Genetic heterogeneity of Liddle syndrome. Nat. Genet. 1995, 11, 76–82. [Google Scholar] [CrossRef] [PubMed]
  38. Hansson, J.H.; Schild, L.; Lu, Y.; Wilson, T.A.; Gautschi, I.; Shimkets, R.; Nelson-Williams, C.; Rossier, B.C.; Lifton, R.P. A de novo missense mutation of the beta subunit of the epithelial sodium channel causes hypertension and Liddle syndrome, identifying a proline-rich segment critical for regulation of channel activity. Proc. Natl. Acad. Sci. USA 1995, 92, 11495–11499. [Google Scholar] [CrossRef] [PubMed]
  39. Tamura, H.; Schild, L.; Enomoto, N.; Matsui, N.; Marumo, F.; Rossier, B.C. Liddle disease caused by a missense mutation of beta subunit of the epithelial sodium channel gene. J. Clin. Investig. 1996, 97, 1780–1784. [Google Scholar] [CrossRef] [PubMed]
  40. Staub, O.; Dho, S.; Henry, P.; Correa, J.; Ishikawa, T.; McGlade, J.; Rotin, D. WW domains of Nedd4 bind to the proline-rich PY motifs in the epithelial Na+ channel deleted in Liddle’s syndrome. EMBO J. 1996, 15, 2371–2380. [Google Scholar] [CrossRef] [PubMed]
  41. Tabibzadeh, N.; Crambert, G. Mechanistic insights into the primary and secondary alterations of renal ion and water transport in the distal nephron. J. Intern. Med. 2023, 293, 4–22. [Google Scholar] [CrossRef]
  42. Taruno, A.; Gordon, M.D. Molecular and cellular mechanisms of salt taste. Annu. Rev. Physiol. 2023, 85, 25–45. [Google Scholar] [CrossRef]
  43. Marunaka, Y. Roles of interstitial fluid pH and weak organic acids in development and amelioration of insulin resistance. Biochem. Soc. Trans. 2021, 49, 715–726. [Google Scholar] [CrossRef]
  44. Pillai, S.R.; Damaghi, M.; Marunaka, Y.; Spugnini, E.P.; Fais, S.; Gillies, R.J. Causes, consequences, and therapy of tumors acidosis. Cancer Metastasis Rev. 2019, 38, 205–222. [Google Scholar] [CrossRef]
  45. Gillies, R.J.; Pilot, C.; Marunaka, Y.; Fais, S. Targeting acidity in cancer and diabetes. Biochim. Biophys. Acta. Rev. Cancer 2019, 1871, 273–280. [Google Scholar] [CrossRef]
  46. Fais, S.; Marunaka, Y. The acidic microenvironment: Is it a phenotype of all cancers? A focus on multiple myeloma and some analogies with diabetes mellitus. Cancers 2020, 12, 3226. [Google Scholar] [CrossRef] [PubMed]
  47. Puppulin, L.; Hosogi, S.; Sun, H.; Matsuo, K.; Inui, T.; Kumamoto, Y.; Suzaki, T.; Tanaka, H.; Marunaka, Y. Bioconjugation strategy for cell surface labelling with gold nanostructures designed for highly localized pH measurement. Nat. Commun. 2018, 9, 5278. [Google Scholar] [CrossRef] [PubMed]
  48. Marunaka, Y. Molecular mechanisms of obesity-induced development of insulin resistance and promotion of amyloid-β accumulation: Dietary therapy using weak organic acids via improvement of lowered interstitial fluid pH. Biomolecules 2023, 13, 779. [Google Scholar] [CrossRef] [PubMed]
  49. Kamariah, N.; Ragunathan, P.; Shin, J.; Saw, W.G.; Wong, C.F.; Dick, T.; Grüber, G. Unique structural and mechanistic properties of mycobacterial F-ATP synthases: Implications for drug design. Prog. Biophys. Mol. Biol. 2020, 152, 64–73. [Google Scholar] [CrossRef]
  50. Patel, H.; Kerndt, C.C.; Bhardwaj, A. Physiology, Respiratory Quotient. In StatPearls; StatPearls Publishing LLC: Treasure Island, FL, USA, 2020. [Google Scholar]
  51. Pessin, J.E.; Bell, G.I. Mammalian facilitative glucose transporter family: Structure and molecular regulation. Annu. Rev. Physiol. 1992, 54, 911–930. [Google Scholar] [CrossRef]
  52. Diaz-Vegas, A.; Sanchez-Aguilera, P.; Krycer, J.R.; Morales, P.E.; Monsalves-Alvarez, M.; Cifuentes, M.; Rothermel, B.A.; Lavandero, S. Is mitochondrial dysfunction a common root of noncommunicable chronic diseases? Endocr. Rev. 2020, 41, 491–517. [Google Scholar] [CrossRef]
  53. Michailidis, M.; Moraitou, D.; Tata, D.A.; Kalinderi, K.; Papamitsou, T.; Papaliagkas, V. Alzheimer’s disease as type 3 Diabetes: Common pathophysiological mechanisms between Alzheimer’s disease and type 2 diabetes. Int. J. Mol. Sci. 2022, 23, 2687. [Google Scholar] [CrossRef]
  54. Mthembu, S.X.H.; Mazibuko-Mbeje, S.E.; Ziqubu, K.; Nyawo, T.A.; Obonye, N.; Nyambuya, T.M.; Nkambule, B.B.; Silvestri, S.; Tiano, L.; Muller, C.J.F.; et al. Impact of physical exercise and caloric restriction in patients with type 2 diabetes: Skeletal muscle insulin resistance and mitochondrial dysfunction as ideal therapeutic targets. Life Sci. 2022, 297, 120467. [Google Scholar] [CrossRef]
  55. Shane, M.A.; Nofziger, C.; Blazer-Yost, B.L. Hormonal regulation of the epithelial Na+ channel: From amphibians to mammals. Gen. Comp. Endocrinol. 2006, 147, 85–92. [Google Scholar] [CrossRef]
  56. Lynn, J.; Park, M.; Ogunwale, C.; Acquaah-Mensah, G.K. A tale of two diseases: Exploring mechanisms linking diabetes mellitus with Alzheimer’s disease. J. Alzheimers Dis. 2022, 85, 485–501. [Google Scholar] [CrossRef] [PubMed]
  57. Potenza, M.A.; Sgarra, L.; Desantis, V.; Nacci, C.; Montagnani, M. Diabetes and Alzheimer’s disease: Might mitochondrial dysfunction help deciphering the common path? Antioxidants 2021, 10, 1257. [Google Scholar] [CrossRef] [PubMed]
  58. Paul, S.; Saha, D.; Bk, B. Mitochondrial dysfunction and mitophagy closely cooperate in neurological deficits associated with Alzheimer’s disease and type 2 diabetes. Mol. Neurobiol. 2021, 58, 3677–3691. [Google Scholar] [CrossRef] [PubMed]
  59. Woo, C.Y.; Jang, J.E.; Lee, S.E.; Koh, E.H.; Lee, K.U. Mitochondrial dysfunction in adipocytes as a primary cause of adipose tissue inflammation. Diabetes Metab. J. 2019, 43, 247–256. [Google Scholar] [CrossRef] [PubMed]
  60. Takano, C.; Ogawa, E.; Hayakawa, S. Insulin resistance in mitochondrial diabetes. Biomolecules 2023, 13, 126. [Google Scholar] [CrossRef] [PubMed]
  61. Petrenko, A.G.; Zozulya, S.A.; Deyev, I.E.; Eladari, D. Insulin receptor-related receptor as an extracellular pH sensor involved in the regulation of acid-base balance. Biochim. Biophys. Acta 2013, 1834, 2170–2175. [Google Scholar] [CrossRef]
  62. Hayata, H.; Miyazaki, H.; Niisato, N.; Yokoyama, N.; Marunaka, Y. Lowered extracellular pH is involved in the pathogenesis of skeletal muscle insulin resistance. Biochem. Biophys. Res. Commun. 2014, 445, 170–174. [Google Scholar] [CrossRef]
  63. Hampel, H.; Vassar, R.; De Strooper, B.; Hardy, J.; Willem, M.; Singh, N.; Zhou, J.; Yan, R.; Vanmechelen, E.; De Vos, A.; et al. The β-secretase BACE1 in Alzheimer’s disease. Biol. Psychiatry 2021, 89, 745–756. [Google Scholar] [CrossRef]
  64. Ju, Y.; Tam, K.Y. Pathological mechanisms and therapeutic strategies for Alzheimer’s disease. Neural Regen. Res. 2022, 17, 543–549. [Google Scholar]
  65. Kamble, S.; Barale, S.; Dhanavade, M.; Sonawane, K. Structural significance of neprylysin from Streptococcus suis GZ1 in the degradation of Aβ peptides, a causative agent in Alzheimer’s disease. Comput. Biol. Med. 2021, 136, 104691. [Google Scholar] [CrossRef]
  66. Kim, N.; Lee, H.J. Redox-active metal ions and amyloid-degrading enzymes in Alzheimer’s disease. Int. J. Mol. Sci. 2021, 22, 7697. [Google Scholar] [CrossRef]
  67. Bell, S.M.; Barnes, K.; De Marco, M.; Shaw, P.J.; Ferraiuolo, L.; Blackburn, D.J.; Venneri, A.; Mortiboys, H. Mitochondrial dysfunction in Alzheimer’s disease: A biomarker of the future? Biomedicines 2021, 9, 63. [Google Scholar] [CrossRef] [PubMed]
  68. Roda, A.R.; Serra-Mir, G.; Montoliu-Gaya, L.; Tiessler, L.; Villegas, S. Amyloid-beta peptide and tau protein crosstalk in Alzheimer’s disease. Neural Regen. Res. 2022, 17, 1666–1674. [Google Scholar] [PubMed]
  69. Cheng, L.; Li, W.; Chen, Y.; Lin, Y.; Wang, B.; Guo, Q.; Miao, Y. Plasma Aβ as a biomarker for predicting Aβ-PET status in Alzheimer’s disease: A systematic review with meta-analysis. J. Neurol. Neurosurg. Psychiatry 2022, 93, 513–520. [Google Scholar] [CrossRef] [PubMed]
  70. Kumar, N.; Kumar, V.; Anand, P.; Kumar, V.; Ranjan Dwivedi, A.; Kumar, V. Advancements in the development of multi-target directed ligands for the treatment of Alzheimer’s disease. Bioorg. Med. Chem. 2022, 61, 116742. [Google Scholar] [CrossRef] [PubMed]
  71. Peng, X.; Fan, R.; Xie, L.; Shi, X.; Dong, K.; Zhang, S.; Tao, J.; Xu, W.; Ma, D.; Chen, J.; et al. A growing link between circadian rhythms, type 2 diabetes mellitus and Alzheimer’s disease. Int. J. Mol. Sci. 2022, 23, 504. [Google Scholar] [CrossRef] [PubMed]
  72. Pakdin, M.; Toutounchian, S.; Namazi, S.; Arabpour, Z.; Pouladi, A.; Afsahi, S.; Poudineh, M.; Nasab, M.M.M.; Yaghoobpoor, S.; Deravi, N. Type 2 diabetes mellitus and Alzheimer disease: A review of the potential links. Curr. Diabetes Rev. 2022, 18, e051121197760. [Google Scholar]
  73. Burillo, J.; Marqués, P.; Jiménez, B.; González-Blanco, C.; Benito, M.; Guillén, C. Insulin resistance and diabetes mellitus in Alzheimer’s disease. Cells 2021, 10, 1236. [Google Scholar] [CrossRef]
  74. Humpel, C. Intranasal neprilysin rapidly eliminates amyloid-beta plaques, but causes plaque compensations: The explanation why the amyloid-beta cascade may fail? Neural. Regen. Res. 2022, 17, 1881–1884. [Google Scholar] [CrossRef] [PubMed]
  75. Nagai, N.; Ito, Y.; Tanino, T. Effect of high glucose levels on amyloid β production in retinas of spontaneous diabetes mellitus Otsuka Long-Evans Tokushima fatty rats. Biol. Pharm. Bull. 2015, 38, 601–610. [Google Scholar] [CrossRef]
  76. Lai, M.C.; Liu, W.Y.; Liou, S.S.; Liu, I.M. The citrus flavonoid hesperetin encounters diabetes-mediated Alzheimer-type neuropathologic changes through relieving advanced glycation end-products Inducing endoplasmic reticulum stress. Nutrients 2022, 14, 745. [Google Scholar] [CrossRef] [PubMed]
  77. Hur, J.-Y. γ-Secretase in Alzheimer’s disease. Exp. Mol. Med. 2022, 54, 433–446. [Google Scholar] [CrossRef]
  78. Sasaguri, H.; Hashimoto, S.; Watamura, N.; Sato, K.; Takamura, R.; Nagata, K.; Tsubuki, S.; Ohshima, T.; Yoshiki, A.; Sato, K.; et al. Recent advances in the modeling of Alzheimer’s disease. Front. Neurosci. 2022, 16, 807473. [Google Scholar] [CrossRef]
  79. Haass, C.; Hung, A.Y.; Schlossmacher, M.G.; Teplow, D.B.; Selkoe, D.J. beta-Amyloid peptide and a 3-kDa fragment are derived by distinct cellular mechanisms. J. Biol. Chem. 1993, 268, 3021–3024. [Google Scholar] [CrossRef]
  80. Knops, J.; Suomensaari, S.; Lee, M.; McConlogue, L.; Seubert, P.; Sinha, S. Cell-type and amyloid precursor protein-type specific inhibition of A beta release by bafilomycin A1, a selective inhibitor of vacuolar ATPases. J. Biol. Chem. 1995, 270, 2419–2422. [Google Scholar] [CrossRef] [PubMed]
  81. Ren, W.; Lou, H.; Ren, X.; Wen, G.; Wu, X.; Xia, X.; Wang, S.; Yu, X.; Yan, L.; Zhang, G.; et al. Ketamine promotes the amyloidogenic pathway by regulating endosomal pH. Toxicology 2022, 471, 153163. [Google Scholar] [CrossRef]
  82. Ellis, C.R.; Shen, J. pH-Dependent Population Shift Regulates BACE1 Activity and Inhibition. J. Am. Chem. Soc. 2015, 137, 9543–9546. [Google Scholar] [CrossRef]
  83. Maesako, M.; Houser, M.C.Q.; Turchyna, Y.; Wolfe, M.S.; Berezovska, O. Presenilin/γ-secretase activity is located in acidic compartments of live neurons. J. Neurosci. 2022, 42, 145–154. [Google Scholar] [CrossRef]
  84. Cai, T.; Hatano, A.; Kanatsu, K.; Tomita, T. Histidine 131 in presenilin 1 is the pH-sensitive residue that causes the increase in Aβ42 level in acidic pH. J. Biochem. 2020, 167, 463–471. [Google Scholar] [CrossRef]
  85. Marunaka, Y. The proposal of molecular mechanisms of weak organic acids intake-induced improvement of insulin resistance in diabetes mellitus via elevation of interstitial fluid pH. Int. J. Mol. Sci. 2018, 19, 3244. [Google Scholar] [CrossRef]
  86. Hosogi, S.; Kuwahara, A.; Kuwahara, Y.; Tanaka, S.; Shimamoto, C.; Tagawa, N.; Kato, I.; Yoshimoto, K.; Aoi, W.; Takata, K.; et al. Mumefural prevents insulin resistance and amyloid-beta accumulation in the brain by improving lowered interstitial fluid pH in type 2 diabetes mellitus. Biomed. Res. 2023, 44, 17–29. [Google Scholar] [CrossRef]
  87. Bazzone, A.; Barthmes, M.; George, C.; Brinkwirth, N.; Zerlotti, R.; Prinz, V.; Cole, K.; Friis, S.; Dickson, A.; Rice, S.; et al. A comparative study on the lysosomal cation channel TMEM175 using automated whole-cell patch-clamp, lysosomal patch-clamp, and solid supported membrane-based electrophysiology: Functional characterization and high-throughput screening assay development. Int. J. Mol. Sci. 2023, 24, 12788. [Google Scholar] [CrossRef] [PubMed]
  88. Salari, A.; Zhou, K.; Nikolovska, K.; Seidler, U.; Amiri, M. Human colonoid-myofibroblast coculture for study of apical Na+/H+ exchangers of the lower cryptal nNeck region. Int. J. Mol. Sci. 2023, 24, 4266. [Google Scholar] [CrossRef] [PubMed]
  89. Tovar, L.M.; Burgos, C.F.; Yévenes, G.E.; Moraga-Cid, G.; Fuentealba, J.; Coddou, C.; Bascunan-Godoy, L.; Catrupay, C.; Torres, A.; Castro, P.A. Understanding the role of ATP release through connexins hemichannels during neurulation. Int. J. Mol. Sci. 2023, 24, 2159. [Google Scholar] [CrossRef] [PubMed]
  90. Baloglu, E. Hypoxic stress-dependent regulation of Na,K-ATPase in ischemic heart disease. Int. J. Mol. Sci. 2023, 24, 7855. [Google Scholar] [CrossRef] [PubMed]
  91. Lemmens-Gruber, R.; Tzotzos, S. The epithelial sodium channel—An underestimated drug target. Int. J. Mol. Sci. 2023, 24, 7775. [Google Scholar] [CrossRef]
Figure 1. Classification of CFTR mutations. This figure has been published in an article by Koivula et al. [13], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Figure 1. Classification of CFTR mutations. This figure has been published in an article by Koivula et al. [13], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Ijms 25 01726 g001
Figure 2. Six families of volume-regulated Cl channels. (i) TMEM16 family: calcium spark activates TMEM16A/Ano1, thereby leading to Cl-efflux. (ii) Bestrophins: Bestrophins like Best1, are activated due to intracellular calcium spark or cell swelling and result in Cl efflux. (iii) Maxi-Cl: can be activated by ATP and/or cell swelling, resulting in the secretion of Cl in bulk amounts. (iv) CLC family: CLC-2 on the plasma membrane is activated by cell swelling, resulting in Cl efflux. CLC-3 in the ER is involved in the secretion of Cl in exchange for hydrogen ions, similarly to CLC-4 and 5 in the endosomes. CLC-6 and 7 are involved in the secretion of Cl from the lysosomes. (v) Ligand-gated Cl channels comprising glycine- and GABAa-gated Cl channels: when the ligands bind to the membrane receptor, the receptors act as Cl channels. (vi) CFTR: cytoplasmic cAMP activates the CFTR channel and thereby causes Cl efflux. ER: endoplasmic reticulum, RyR: ryanodine receptor, IP3R: Inositol-triphosphate receptor. Ion denotation: Red: Ca2+; Green: Cl; Red: hydrogen ion; Blue: GABA ligand; Green: Glycine ligand. This figure has been published in an article by Sinha et al. [19], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Figure 2. Six families of volume-regulated Cl channels. (i) TMEM16 family: calcium spark activates TMEM16A/Ano1, thereby leading to Cl-efflux. (ii) Bestrophins: Bestrophins like Best1, are activated due to intracellular calcium spark or cell swelling and result in Cl efflux. (iii) Maxi-Cl: can be activated by ATP and/or cell swelling, resulting in the secretion of Cl in bulk amounts. (iv) CLC family: CLC-2 on the plasma membrane is activated by cell swelling, resulting in Cl efflux. CLC-3 in the ER is involved in the secretion of Cl in exchange for hydrogen ions, similarly to CLC-4 and 5 in the endosomes. CLC-6 and 7 are involved in the secretion of Cl from the lysosomes. (v) Ligand-gated Cl channels comprising glycine- and GABAa-gated Cl channels: when the ligands bind to the membrane receptor, the receptors act as Cl channels. (vi) CFTR: cytoplasmic cAMP activates the CFTR channel and thereby causes Cl efflux. ER: endoplasmic reticulum, RyR: ryanodine receptor, IP3R: Inositol-triphosphate receptor. Ion denotation: Red: Ca2+; Green: Cl; Red: hydrogen ion; Blue: GABA ligand; Green: Glycine ligand. This figure has been published in an article by Sinha et al. [19], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Ijms 25 01726 g002
Figure 3. Changes to cell shape required during cell migration. Na+-K+-2Cl cotransporter 1 (NKCC1) and aquaporin (AQP) are expressed on the migrating-side membrane. NKCC1 is involved in Cl uptake into the cytosolic space with Na+ and K+ [22,23,24,25,26]. The uptake of these ions results in an influx of water into the cytosolic space via AQP due to an increase in cytosolic osmolarity [22]. The movement of Cl, Na+, K+ and water causes an increase in cell volume accompanied by [Cl]c elevation, which promotes tubulin polymerization (elongation) [27] by inhibiting GTPase activity [21]. Like tubulin polymerization, actin monomers are enhanced to be polymerized. Then, cells migrate via these processes. On the one hand, K+-Cl cotransporter (KCC), volume-regulated anion channel (VRAC), Ca2+-activated K+ channel (K+Ca3.1) and AQP are expressed on the tail-end membrane during cell migration and Cl with K+ are excreted to the extracellular space via KCC, VRAC and K+Ca3.1 [22,26]. AQP-mediated water efflux to the extracellular space is caused by diminution in cytosolic osmolarity due to excretion of these ions. The movement of Cl, K+ and water results in a decrease in cell volume accompanied by [Cl]c diminution, which leads to tubulin depolymerization (shortening) during the tail-end of the cell migration [27] by activating GTPase [21]. WNK activated by lowered [Cl]c induces phosphorylation (activation) of OSR1/SPACK, which increases NKCC1 activity by phosphorylating NKCC1 [23]. Thus, WNK is crucial within cell migration [23]. This figure has been published in an article by Marunaka [21] and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Figure 3. Changes to cell shape required during cell migration. Na+-K+-2Cl cotransporter 1 (NKCC1) and aquaporin (AQP) are expressed on the migrating-side membrane. NKCC1 is involved in Cl uptake into the cytosolic space with Na+ and K+ [22,23,24,25,26]. The uptake of these ions results in an influx of water into the cytosolic space via AQP due to an increase in cytosolic osmolarity [22]. The movement of Cl, Na+, K+ and water causes an increase in cell volume accompanied by [Cl]c elevation, which promotes tubulin polymerization (elongation) [27] by inhibiting GTPase activity [21]. Like tubulin polymerization, actin monomers are enhanced to be polymerized. Then, cells migrate via these processes. On the one hand, K+-Cl cotransporter (KCC), volume-regulated anion channel (VRAC), Ca2+-activated K+ channel (K+Ca3.1) and AQP are expressed on the tail-end membrane during cell migration and Cl with K+ are excreted to the extracellular space via KCC, VRAC and K+Ca3.1 [22,26]. AQP-mediated water efflux to the extracellular space is caused by diminution in cytosolic osmolarity due to excretion of these ions. The movement of Cl, K+ and water results in a decrease in cell volume accompanied by [Cl]c diminution, which leads to tubulin depolymerization (shortening) during the tail-end of the cell migration [27] by activating GTPase [21]. WNK activated by lowered [Cl]c induces phosphorylation (activation) of OSR1/SPACK, which increases NKCC1 activity by phosphorylating NKCC1 [23]. Thus, WNK is crucial within cell migration [23]. This figure has been published in an article by Marunaka [21] and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Ijms 25 01726 g003
Figure 4. Distribution of ENaC in organs and related diseases. This figure has been published in an article by Chen et al. [28], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Figure 4. Distribution of ENaC in organs and related diseases. This figure has been published in an article by Chen et al. [28], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Ijms 25 01726 g004
Figure 5. Mitochondrial dysfunction-induced insulin resistance via lowering interstitial fluid pH. (A) Metabolic cells with normal mitochondrial function. (B) Metabolic cells with mitochondrial dysfunction. AE, anion exchanger; CA, carbonic anhydrase; MCT, monocarboxylate transporter; NBC, Na+-HCO3 cotransporter; NDCBE, Na+-driven Cl/HCO3 exchanger; NHE, Na+/H+ exchanger. This figure has been published in an article by Marunaka [48], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Figure 5. Mitochondrial dysfunction-induced insulin resistance via lowering interstitial fluid pH. (A) Metabolic cells with normal mitochondrial function. (B) Metabolic cells with mitochondrial dysfunction. AE, anion exchanger; CA, carbonic anhydrase; MCT, monocarboxylate transporter; NBC, Na+-HCO3 cotransporter; NDCBE, Na+-driven Cl/HCO3 exchanger; NHE, Na+/H+ exchanger. This figure has been published in an article by Marunaka [48], and is described under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/) on 16 January 2024.
Ijms 25 01726 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

Marunaka, Y. The Role of Ion-Transporting Proteins in Human Disease. Int. J. Mol. Sci. 2024, 25, 1726. https://doi.org/10.3390/ijms25031726

AMA Style

Marunaka Y. The Role of Ion-Transporting Proteins in Human Disease. International Journal of Molecular Sciences. 2024; 25(3):1726. https://doi.org/10.3390/ijms25031726

Chicago/Turabian Style

Marunaka, Yoshinori. 2024. "The Role of Ion-Transporting Proteins in Human Disease" International Journal of Molecular Sciences 25, no. 3: 1726. https://doi.org/10.3390/ijms25031726

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