Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2526 2023-06-13 12:11:44 |
2 layout + 3 word(s) 2529 2023-06-14 05:35:34 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Coppola, M.A.; Tettey-Matey, A.; Imbrici, P.; Gavazzo, P.; Liantonio, A.; Pusch, M. The Gene Family of Voltage-Gated ChLoride Channels. Encyclopedia. Available online: https://encyclopedia.pub/entry/45496 (accessed on 20 May 2024).
Coppola MA, Tettey-Matey A, Imbrici P, Gavazzo P, Liantonio A, Pusch M. The Gene Family of Voltage-Gated ChLoride Channels. Encyclopedia. Available at: https://encyclopedia.pub/entry/45496. Accessed May 20, 2024.
Coppola, Maria Antonietta, Abraham Tettey-Matey, Paola Imbrici, Paola Gavazzo, Antonella Liantonio, Michael Pusch. "The Gene Family of Voltage-Gated ChLoride Channels" Encyclopedia, https://encyclopedia.pub/entry/45496 (accessed May 20, 2024).
Coppola, M.A., Tettey-Matey, A., Imbrici, P., Gavazzo, P., Liantonio, A., & Pusch, M. (2023, June 13). The Gene Family of Voltage-Gated ChLoride Channels. In Encyclopedia. https://encyclopedia.pub/entry/45496
Coppola, Maria Antonietta, et al. "The Gene Family of Voltage-Gated ChLoride Channels." Encyclopedia. Web. 13 June, 2023.
The Gene Family of Voltage-Gated ChLoride Channels
Edit

Endosomes and lysosomes are intracellular vesicular organelles with important roles in cell functions such as protein homeostasis, clearance of extracellular material, and autophagy. Endolysosomes are characterized by an acidic luminal pH that is critical for proper function. Five members of the gene family of voltage-gated ChLoride Channels (CLC proteins) are localized to endolysosomal membranes, carrying out anion/proton exchange activity and thereby regulating pH and chloride concentration. Mutations in these vesicular CLCs cause global developmental delay, intellectual disability, various psychiatric conditions, lysosomal storage diseases, and neurodegeneration, resulting in severe pathologies or even death. 

CLC proteins endosome lysosome chloride transport developmental

1. Introduction—The ChLoride Channel Family

Physiologically, the most abundant anion is chloride. It is an important substrate of many transport proteins, being carried across the membrane as a single anion or coupled with other ions, and is important, for example, for the regulation of the membrane potential, intracellular vesicles acidification and cell volume regulation [1].
In humans, the ChLoride Channel (CLC) family is formed by nine members, which had initially been supposed to be all chloride channels, because of their sequence homology with the founding member, the Torpedo electroplax channel ClC-0 [2]. The discovery that the bacterial Escherichia coli ecClC-1 homologue is not a passive chloride channel but a stoichiometrically coupled secondary active 2 Cl/1 H+ antiporter has dramatically changed the point of view of the entire CLC group [3]. Based on sequence homology, three branches of human CLCs have been distinguished. The first one includes the plasma membrane-localized chloride channels ClC-1, ClC-2 and the two isoforms ClC-Ka and ClC-Kb. The second branch is formed by ClC-3, ClC-4 and ClC-5, while the third branch contains ClC-6 and ClC-7. ClC-3 to -7 are all Cl/H+ exchangers and are localized to the intracellular membranes of endosomes and/or lysosomes [1].
All CLC family members share the same dimeric architecture that is unique to this protein family. Except for ClC-6 and ClC-7 [4], the other CLC proteins can form homo- or hetero-dimers with members of the same branch [1]. Biochemical studies and single-channel analysis on the first cloned Torpedo ClC-0, mutants [2][5][6] and biochemical and low-resolution structural analysis of ecClC-1 [7][8] suggested a homodimeric “double-barreled” architecture, with physically separated anion transport pathways in each protomer. This architecture has been fully confirmed by the determination of ecClC-1 and Salmonella typhimurium stClC crystal structures [9][10]. The structures revealed the presence of distinct anion binding sites, formed by residues that are also highly conserved in human CLCs. The sites are denominated Sext, Scen and Sint, with Sext being occupied by the presumably negatively charged side chain of the “gating glutamate” E148 [9][10]. Each monomer presents 18 α-helices (from A to R) of which 17 are partially embedded in the membrane. The two subunits interact in a tight manner and the architecture follows an inverted and parallel orientation [1]. Two C-terminal tandem cystathionine-β-synthase (CBS) domains are present in most eukaryotic CLC proteins [11][12], but are absent in ecClC-1. The two CBS domains may have a role in the so-called common gating process (that will be discussed in more detail below) and confer unique features to the CLC members [1]. Dutzler and colleagues determined the crystal structures of isolated CBS domains of Torpedo ClC-0 [13], human ClC-5 [14] and human ClC-Ka [15]. CBS domains are present in many different protein families, where they are often implicated in the sensing of adenine nucleotides [11][16]. Structurally, so far, ATP has been found to be bound in the isolated domains of ClC-5 and in the full-length structure of ClC-7 [17], but not in isolated domains of ClC-0 and ClC-Ka and not in full-length structures of bovine ClC-K or human ClC-1 [18][19][20].
Single-channel recordings of the Torpedo ClC-0 channel displayed two kinds of gating mechanisms that regulate the open probability (Po) of the channel: a “fast” or “protopore” gate that acts independently on single pores determines the closing or opening state of each pore of the double-barreled structure [1]. The fast gate is mainly determined by the gating glutamate (E166 in ClC-0), in that its neutralization renders CLC-0 channels voltage independent. Protonation of the gating glutamate and its competition with permeant ions underlie the anion and pH dependent protopore gating of most CLC channels [10][21][22][23]. Conversely, a second mechanism, termed a slow or common gate, operates on both pores simultaneously and is still not well understood [1].
Some CLC proteins require association with a small ancillary subunit for proper function or membrane expression. In particular, the kidney ClC-Ka and ClC-Kb channels require association with the barttin subunit [24]. In glia cells, ClC-2 associates with GlialCAM, a protein with the typical architecture of a cell adhesion molecule, which is mutated in megalencephalic leukoencephalopathy with subcortical cysts (MLC) [25]. It leads to clustering of ClC-2 at glial cell–cell contacts and alters biophysical functions of the ClC-2 channel [25][26]. The complex of ClC-7 with its subunit Ostm1 is mandatory for mutual stabilization [4][27].
The plasma membrane localized chloride channels belonging to the first branch of the CLC family are expressed in a tissue-dependent manner that is different for each member according to their physiological role. All channel CLCs are involved in various human genetic diseases, as reviewed in detail elsewhere [1][28][29].
ClC-3 through ClC-7, which are the focus of this research, function as Cl/H+ exchangers and are localized to intracellular endosomes and/or lysosomes (Figure 1, Table 1). Initially, when the transporter function of the intracellular CLCs was not yet known, it was proposed that they act as charge-shunting chloride channels to assist the luminal acidification of endosomes and lysosomes intracellular organelles [30][31][32][33][34]. Indeed, the maintenance of an acidic pH of the lumen of endo-/lysosomes is required for their proper physiological function. The proton pumping V-ATPase is electrogenic and thus generates an electrical potential difference that would impede acidification if not neutralized by anionic cotransport and/or cationic counter-transport. Somewhat surprisingly and counter-intuitively, model calculations show that a 2 Cl/H+ exchange activity, contributing to a more inside-negative voltage, allows a more acidic steady-state luminal pH compared to a shunting Cl channel [35][36].
Figure 1. Schematic illustration of localization of vesicular CLCs in the endo-/lysosomal pathway.
Among the endo-/lysosomal CLCs, ClC-5 is rather specifically expressed in the kidney with a predominant presence in epithelial cells of the proximal tubule, where it is involved in endocytic uptake [30][31][33]. Indeed, mutations causing impaired ClC-5 transport activity are associated with Dent’s disease, a kidney disorder characterized by the primary symptom of low molecular weight proteinuria, and a series of secondary symptoms including kidney stones and renal failure, caused by defective endocytosis in the proximal tubule [33][37]. ClC-7, together with its subunit Ostm1 [27], is rather ubiquitously expressed in the body and is localized to lysosomes and in the ruffled border of osteoclasts functioning as a 2Cl/H+ antiporter [1]. Accordingly, impaired bone resorption in osteoclast, caused by a functionally defective ClC-7/Ostm1 complex, causes osteopetrosis, a disease characterized by stiff and fragile bones [38].
A large phenotypic spectrum of neuronal diseases is associated with mutations in the genes encoding ClC-3/-4/-6 and ClC-7, as will be described in detail in the following paragraphs (see Table 1).
For all vesicular CLCs, an unsolved question pertains to the direction of exchanger transport. Despite being physiologically localized to endo-/lysosomes, ClC-3 to -6 can reach the plasma membrane when heterologously expressed in HEK293 cells, allowing the investigation of their biophysical properties using the patch clamp technique [45][50][51][52][53]. For ClC-7, the elimination of N-terminal lysosomal targeting motifs leads to plasma membrane expression [4][54]. ClC-3 to -5 all exhibit extreme outward rectification of currents with very little or nonresolvable activation kinetics [50][51][52][53][55]. This current direction corresponds to the transport of luminal Cl out of lysosomes with a parallel influx of cytosolic H+. However, it remains unclear whether the direction of transport is physiologically relevant and whether CLC exchangers work synergistically with V-ATPase, contributing to luminal acidification.
Additionally, ClC-6 and ClC-7 exhibit strongly outwardly rectifying currents, which are, however, characterized by slow activation kinetics and measurable inward “tail” currents [4][45].

2. ClC-3 and ClC-4

The second branch of the CLC family comprises ClC-3, -4 and 5. These three endosomal transporters share high sequence similarity and have similar functional properties [1][29]. Among the human CLCs, they are the most similar to the Escherichia coli ecClC-1 homologue. The renal-specific ClC-5 is found mostly in recycling endosomes and its physiological role will not be discussed in detail [1]. ClC-3 and ClC-4 are localized to sorting endosomes, and ClC-3 is probably localized in late endosomes as well [1][29]. ClC-3 has also been proposed to play a role in synaptic vesicles. This is, however, still controversial and will not be discussed in detail here.

While ClC-4 KO mice have no overt phenotype, in 2013 and 2016, patients (mostly pediatric) with a range of neurodevelopmental and psychiatric complications have been described with X-linked CLCN4 variants [40,42,65] (see Table 1). In heterologous expression, these and some novel variants [66] showed variable loss of function effects. It is important to note that complete loss of ClC-4 protein leads to non-syndromic intellectual disability in males and no disease in heterozygous females. In contrast, de novo and inherited missense variants can lead to severe syndromic neurological disease in males as well as in females, suggesting a dominant effect. In a more recent study, a large number of CLCN4 families was investigated, describing a large spectrum of clinical phenotypes and studying > 50 missense variants in heterologous expression [67]. Novel biophysical mechanisms were discovered for new and already described variants. These included a toxic gain of function characterized by the presence of negative currents at acidic extracellular (luminal) pH, and a shift in the voltage dependence of gating to more positive voltages [67]. Both effects can be expected to exert dominant negative effects in ClC-3/ClC-4 heterodimers.

Almost simultaneously came the discovery of the first variants in CLCN3 that cause global developmental delay, intellectual disability and neurodevelopmental disorders [39] (Table 1). Detailed functional analysis revealed a toxic gain of function for two missense variants, similar to the above-described effects in some CLCN4 variants [39].

3. ClC-6

The third branch of the CLC family comprises ClC-6 and ClC-7, which both function as Cl/H+ exchangers [1][45]. Even though the expression of ClC-6 mRNA appears to be ubiquitous in many tissues [56], biochemical analysis detected native ClC-6 protein predominantly in neurons, where it localizes to late endosomes and partially lysosomes [44]. For a long time, the biophysical profile of ClC-6 remained completely unknown. In the first attempts at heterologous expression, no currents attributable to ClC-6 could be detected [56], possibly caused by the intracellular localization of most of the overexpressed protein [57].
A subtype of lysosomal storage disease, referred to as neuronal ceroid lipofuscinosis (NCL), was observed in ClC-6 knockout mice presenting a mild phenotype with features of reduced pain sensitivity, probably due to strong accumulation of materials in axon initial segments, mild cognitive abnormalities and no impact on their span life [44]. This evidence suggested that CLCN6 variants could be involved in human NCL [44]. Indeed, in a sample of 75 adult-onset variants, including late-onset forms of NLC and Kufs’ disease, two individuals were found to be heterozygous for CLCN6 missense variants (V580M and T628R) [44]. However, no functional analysis had been performed at the time of that study because the transporter had not been successfully functionally characterized.
Preliminary electrophysiological characterization was obtained when the N-terminus of ClC-6 tagged with GFP (GFP-ClC-6) was reported to enhance its cell surface localization [58]. However, the reported currents were small and barely above background levels.

4. ClC-7

Belonging to the third mammalian CLC branch, ClC-7 shares 45% of sequence homology with ClC-6. It was cloned in parallel with ClC-6 in 1995 [56], but could not be functionally analyzed for a long time. Intriguingly, ClC-7 is the only subcellular CLC member to be present almost exclusively in lysosomes [44]. Moreover, it has also been found in the ruffled border of osteoclasts, where it participates in bone resorption [38]. Unlike the other CLC transporters, ClC-7 requires association with a type I transmembrane protein, called Ostm1, for proper function and stability [4][27].
Similarly to ClC-6, no information about electrophysiological ClC-7 characterization has been available for a long time, due to its intracellular localization upon heterologous expression [1]. Ion flux studies with isolated mouse lysosomes showed that ClC-7 is the dominant anion permeation pathway of lysosomal membranes and that it performs 2 Cl/1 H+ antiport activity [59]. A breakthrough was achieved by Stauber and Jentsch, who discovered the sorting motifs that mediate lysosomal targeting [54]. In particular, they found that when four leucine residues localized in the N-terminal portion are changed to alanine, the transporter is at least partially targeted to the plasma membrane [54]. Notably, Ostm1 follows ClC-7 in its expression location. ClC-7PM, the ClC-7 variant in which the two dileucine motifs are mutated to alanine, elicited robust transmembrane, outwardly rectifying voltage-activated currents [4]. Even though some electrophysiological properties of ClC-7 are similar to that of other vesicular CLCs, including the inhibitory effect of acidic pH, ClC-7 differs substantially from ClC-3 to -5. Most importantly, ClC-7PM exhibits very slow activation kinetics in the seconds time range [4]. This slow “gating” phenomenon is strictly linked to conformational changes in the proteins, where the interactions between transmembrane as well as cytoplasmic domains play a key role [60]. In addition to the transport currents, Pusch and Zifarelli discovered that the transporter also exhibits rather large “transient” or “capacitive” currents that reflect charge rearrangements within the protein. These are most likely mediated by movements of the gating glutamate and chloride binding/unbinding events [61]. Similar currents have been observed in ClC-5 and ClC-3 [52][62][63]. The transient currents probably have no physiological role, but represent a biophysical feature that can be useful in deciphering molecular mechanisms of gating and transport. Interestingly, while in ClC-5, neutralization of the so-called proton glutamate completely abolished transport currents, leaving only transient currents [62][63], in ClC-7, residual transport currents were observed in the corresponding E312A mutant [61].
The physiological role of ClC-7 remained unclear for a long time. The first insights were obtained with a mouse KO model that was characterized by severe osteopetrosis [38]. The involvement of ClC-7 in bone resorption was confirmed by the presence of CLCN7 mutations in a human patient with malignant osteopetrosis [38]. Further evidence came from the identification of a spontaneous Ostm1 mutation to be associated with the onset of a severe osteopetrosis in gray lethal mice presenting a fur color defect [64]. In Clcn7−/− mice, even though the number of osteoclasts was normal, their ability to reabsorb calcified bone was impaired [38]. Interestingly, however, no impact on lysosomal acidification was observed, suggesting that the osteoclasts’ ability to acidify intracellular vesicles was preserved in Clcn7−/− mice [38]. The life span of the KO mice was limited to 6–7 weeks.

References

  1. Jentsch, T.J.; Pusch, M. CLC chloride channels and transporters: Structure, function, physiology, and disease. Physiol. Rev. 2018, 98, 1493–1590.
  2. Jentsch, T.J.; Steinmeyer, K.; Schwarz, G. Primary structure of Torpedo marmorata chloride channel isolated by expression cloning in Xenopus oocytes. Nature 1990, 348, 510–514.
  3. Accardi, A.; Miller, C. Secondary active transport mediated by a prokaryotic homologue of ClC Cl− channels. Nature 2004, 427, 803–807.
  4. Leisle, L.; Ludwig, C.F.; Wagner, F.A.; Jentsch, T.J.; Stauber, T. ClC-7 is a slowly voltage-gated 2Cl−/1H+-exchanger and requires Ostm1 for transport activity. EMBO J. 2011, 30, 2140–2152.
  5. Middleton, R.E.; Pheasant, D.J.; Miller, C. Homodimeric architecture of a ClC-type chloride ion channel. Nature 1996, 383, 337–340.
  6. Ludewig, U.; Pusch, M.; Jentsch, T.J. Two physically distinct pores in the dimeric ClC-0 chloride channel. Nature 1996, 383, 340–343.
  7. Maduke, M.; Pheasant, D.J.; Miller, C. High-level expression, functional reconstitution, and quaternary structure of a prokaryotic ClC-type chloride channel. J. Gen. Physiol. 1999, 114, 713–722.
  8. Mindell, J.A.; Maduke, M.; Miller, C.; Grigorieff, N. Projection structure of a ClC-type chloride channel at 6.5 Å resolution. Nature 2001, 409, 219–223.
  9. Dutzler, R.; Campbell, E.B.; Cadene, M.; Chait, B.T.; MacKinnon, R. X-ray structure of a ClC chloride channel at 3.0 Å reveals the molecular basis of anion selectivity. Nature 2002, 415, 287–294.
  10. Dutzler, R.; Campbell, E.B.; MacKinnon, R. Gating the selectivity filter in ClC chloride channels. Science 2003, 300, 108–112.
  11. Ponting, C.P. CBS domains in CIC chloride channels implicated in myotonia and nephrolithiasis (kidney stones). J. Mol. Med. 1997, 75, 160–163.
  12. Estévez, R.; Jentsch, T.J. CLC chloride channels: Correlating structure with function. Curr. Opin. Struct. Biol. 2002, 12, 531–539.
  13. Meyer, S.; Dutzler, R. Crystal structure of the cytoplasmic domain of the chloride channel ClC-0. Structure 2006, 14, 299–307.
  14. Meyer, S.; Savaresi, S.; Forster, I.C.; Dutzler, R. Nucleotide recognition by the cytoplasmic domain of the human chloride transporter ClC-5. Nat. Struct. Mol. Biol. 2007, 14, 60–67.
  15. Markovic, S.; Dutzler, R. The structure of the cytoplasmic domain of the chloride channel ClC-Ka reveals a conserved interaction interface. Structure 2007, 15, 715–725.
  16. Bateman, A. The structure of a domain common to archaebacteria and the homocystinuria disease protein. Trends Biochem. Sci. 1997, 22, 12–13.
  17. Schrecker, M.; Korobenko, J.; Hite, R.K. Cryo-EM structure of the lysosomal chloride-proton exchanger CLC-7 in complex with OSTM1. eLife 2020, 9, e59555.
  18. Park, E.; Campbell, E.B.; MacKinnon, R. Structure of a CLC chloride ion channel by cryo-electron microscopy. Nature 2017, 541, 500–505.
  19. Park, E.; MacKinnon, R. Structure of the CLC-1 chloride channel from Homo sapiens. eLife 2018, 7, e36629.
  20. Wang, K.; Preisler, S.S.; Zhang, L.; Cui, Y.; Missel, J.W.; Gronberg, C.; Gotfryd, K.; Lindahl, E.; Andersson, M.; Calloe, K.; et al. Structure of the human ClC-1 chloride channel. PLoS Biol. 2019, 17, e3000218.
  21. Pusch, M.; Ludewig, U.; Rehfeldt, A.; Jentsch, T.J. Gating of the voltage-dependent chloride channel CIC-0 by the permeant anion. Nature 1995, 373, 527–531.
  22. Traverso, S.; Elia, L.; Pusch, M. Gating competence of constitutively open CLC-0 mutants revealed by the interaction with a small organic Inhibitor. J. Gen. Physiol. 2003, 122, 295–306.
  23. Zifarelli, G.; Murgia, A.R.; Soliani, P.; Pusch, M. Intracellular proton regulation of ClC-0. J. Gen. Physiol. 2008, 132, 185–198.
  24. Estévez, R.; Boettger, T.; Stein, V.; Birkenhäger, R.; Otto, E.; Hildebrandt, F.; Jentsch, T.J. Barttin is a Cl− channel beta-subunit crucial for renal Cl− reabsorption and inner ear K+ secretion. Nature 2001, 414, 558–561.
  25. Jeworutzki, E.; López-Hernández, T.; Capdevila-Nortes, X.; Sirisi, S.; Bengtsson, L.; Montolio, M.; Zifarelli, G.; Arnedo, T.; Müller, C.S.; Schulte, U.; et al. GlialCAM, a protein defective in a leukodystrophy, serves as a ClC-2 Cl− channel auxiliary subunit. Neuron 2012, 73, 951–961.
  26. Jeworutzki, E.; Lagostena, L.; Elorza-Vidal, X.; Lopez-Hernandez, T.; Estevez, R.; Pusch, M. GlialCAM, a CLC-2 Cl− channel subunit, activates the slow gate of CLC chloride channels. Biophys. J. 2014, 107, 1105–1116.
  27. Lange, P.F.; Wartosch, L.; Jentsch, T.J.; Fuhrmann, J.C. ClC-7 requires Ostm1 as a beta-subunit to support bone resorption and lysosomal function. Nature 2006, 440, 220–223.
  28. Stauber, T.; Weinert, S.; Jentsch, T.J. Cell biology and physiology of CLC chloride channels and transporters. Compr. Physiol. 2012, 2, 1701–1744.
  29. Bose, S.; He, H.; Stauber, T. Neurodegeneration upon dysfunction of endosomal/lysosomal CLC chloride transporters. Front. Cell Dev. Biol. 2021, 9, 639231.
  30. Günther, W.; Lüchow, A.; Cluzeaud, F.; Vandewalle, A.; Jentsch, T.J. ClC-5, the chloride channel mutated in Dent’s disease, colocalizes with the proton pump in endocytotically active kidney cells. Proc. Natl. Acad. Sci. USA 1998, 95, 8075–8080.
  31. Günther, W.; Piwon, N.; Jentsch, T.J. The ClC-5 chloride channel knock-out mouse—An animal model for Dent’s disease. Pflügers Arch. 2003, 445, 456–462.
  32. Jentsch, T.J.; Günther, W. Chloride channels: An emerging molecular picture. Bioessays 1997, 19, 117–126.
  33. Piwon, N.; Günther, W.; Schwake, M.; Bösl, M.R.; Jentsch, T.J. ClC-5 Cl−-channel disruption impairs endocytosis in a mouse model for Dent’s disease. Nature 2000, 408, 369–373.
  34. Stobrawa, S.M.; Breiderhoff, T.; Takamori, S.; Engel, D.; Schweizer, M.; Zdebik, A.A.; Bösl, M.R.; Ruether, K.; Jahn, H.; Draguhn, A.; et al. Disruption of ClC-3, a chloride channel expressed on synaptic vesicles, leads to a loss of the hippocampus. Neuron 2001, 29, 185–196.
  35. Weinert, S.; Jabs, S.; Supanchart, C.; Schweizer, M.; Gimber, N.; Richter, M.; Rademann, J.; Stauber, T.; Kornak, U.; Jentsch, T.J. Lysosomal pathology and osteopetrosis upon loss of H+-driven lysosomal Cl− accumulation. Science 2010, 328, 1401–1403.
  36. Ishida, Y.; Nayak, S.; Mindell, J.A.; Grabe, M. A model of lysosomal pH regulation. J. Gen. Physiol. 2013, 141, 705–720.
  37. Lloyd, S.E.; Pearce, S.H.; Fisher, S.E.; Steinmeyer, K.; Schwappach, B.; Scheinman, S.J.; Harding, B.; Bolino, A.; Devoto, M.; Goodyer, P.; et al. A common molecular basis for three inherited kidney stone diseases. Nature 1996, 379, 445–449.
  38. Kornak, U.; Kasper, D.; Bösl, M.R.; Kaiser, E.; Schweizer, M.; Schulz, A.; Friedrich, W.; Delling, G.; Jentsch, T.J. Loss of the ClC-7 chloride channel leads to osteopetrosis in mice and man. Cell 2001, 104, 205–215.
  39. Duncan, A.R.; Polovitskaya, M.M.; Gaitan-Penas, H.; Bertelli, S.; VanNoy, G.E.; Grant, P.E.; O’Donnell-Luria, A.; Valivullah, Z.; Lovgren, A.K.; England, E.M.; et al. Unique variants in CLCN3, encoding an endosomal anion/proton exchanger, underlie a spectrum of neurodevelopmental disorders. Am. J. Hum. Genet. 2021, 108, 1450–1465.
  40. Hu, H.; Haas, S.A.; Chelly, J.; Van Esch, H.; Raynaud, M.; de Brouwer, A.P.; Weinert, S.; Froyen, G.; Frints, S.G.; Laumonnier, F.; et al. X-exome sequencing of 405 unresolved families identifies seven novel intellectual disability genes. Mol. Psychiatry 2016, 21, 133–148.
  41. Palmer, E.E.; Pusch, M.; Picollo, A.; Forwood, C.; Nguyen, M.H.; Suckow, V.; Gibbons, J.; Hoff, A.; Sigfrid, L.; Megarbane, A.; et al. Functional and clinical studies reveal pathophysiological complexity of CLCN4-related neurodevelopmental condition. Mol. Psychiatry 2023, 28, 668–697.
  42. Palmer, E.E.; Stuhlmann, T.; Weinert, S.; Haan, E.; Van Esch, H.; Holvoet, M.; Boyle, J.; Leffler, M.; Raynaud, M.; Moraine, C.; et al. De novo and inherited mutations in the X-linked gene CLCN4 are associated with syndromic intellectual disability and behavior and seizure disorders in males and females. Mol. Psychiatry 2016, 23, 222–230.
  43. Polovitskaya, M.M.; Barbini, C.; Martinelli, D.; Harms, F.L.; Cole, F.S.; Calligari, P.; Bocchinfuso, G.; Stella, L.; Ciolfi, A.; Niceta, M.; et al. A Recurrent Gain-of-Function Mutation in CLCN6, Encoding the ClC-6 Cl−/H+-Exchanger, Causes Early-Onset Neurodegeneration. Am. J. Hum. Genet. 2020, 107, 1062–1077.
  44. Poët, M.; Kornak, U.; Schweizer, M.; Zdebik, A.A.; Scheel, O.; Hoelter, S.; Wurst, W.; Schmitt, A.; Fuhrmann, J.C.; Planells-Cases, R.; et al. Lysosomal storage disease upon disruption of the neuronal chloride transport protein ClC-6. Proc. Natl. Acad. Sci. USA 2006, 103, 13854–13859.
  45. Zifarelli, G.; Pusch, M.; Fong, P. Altered voltage-dependence of slowly activating chloride-proton antiport by late endosomal ClC-6 explains distinct neurological disorders. J. Physiol. 2022, 600, 2147–2164.
  46. He, H.; Cao, X.; Yin, F.; Wu, T.; Stauber, T.; Peng, J. West syndrome caused by a chloride/proton exchange-uncoupling CLCN6 mutation related to autophagic-lysosomal dysfunction. Mol. Neurobiol. 2021, 58, 2990–2999.
  47. Kasper, D.; Planells-Cases, R.; Fuhrmann, J.C.; Scheel, O.; Zeitz, O.; Ruether, K.; Schmitt, A.; Poët, M.; Steinfeld, R.; Schweizer, M.; et al. Loss of the chloride channel ClC-7 leads to lysosomal storage disease and neurodegeneration. EMBO J. 2005, 24, 1079–1091.
  48. Pangrazio, A.; Pusch, M.; Caldana, E.; Frattini, A.; Lanino, E.; Tamhankar, P.M.; Phadke, S.; Lopez, A.G.; Orchard, P.; Mihci, E.; et al. Molecular and clinical heterogeneity in CLCN7-dependent osteopetrosis: Report of 20 novel mutations. Hum. Mutat. 2010, 31, E1071–E1080.
  49. Nicoli, E.R.; Weston, M.R.; Hackbarth, M.; Becerril, A.; Larson, A.; Zein, W.M.; Baker, P.R., 2nd; Burke, J.D.; Dorward, H.; Davids, M.; et al. Lysosomal storage and albinism due to effects of a de novo CLCN7 variant on lysosomal acidification. Am. J. Hum. Genet. 2019, 104, 1127–1138.
  50. Steinmeyer, K.; Schwappach, B.; Bens, M.; Vandewalle, A.; Jentsch, T.J. Cloning and functional expression of rat CLC-5, a chloride channel related to kidney disease. J. Biol. Chem. 1995, 270, 31172–31177.
  51. Friedrich, T.; Breiderhoff, T.; Jentsch, T.J. Mutational analysis demonstrates that ClC-4 and ClC-5 directly mediate plasma membrane currents. J. Biol. Chem. 1999, 274, 896–902.
  52. Guzman, R.E.; Grieschat, M.; Fahlke, C.; Alekov, A.K. ClC-3 is an intracellular chloride/proton exchanger with large voltage-dependent nonlinear capacitance. ACS Chem. Neurosci. 2013, 4, 994–1003.
  53. Guzman, R.E.; Miranda-Laferte, E.; Franzen, A.; Fahlke, C. Neuronal ClC-3 splice variants differ in subcellular localizations, but mediate identical transport functions. J. Biol. Chem. 2015, 290, 25851–25862.
  54. Stauber, T.; Jentsch, T.J. Sorting motifs of the endosomal/lysosomal CLC chloride transporters. J. Biol. Chem. 2010, 285, 34537–34548.
  55. Zifarelli, G.; Pusch, M. Conversion of the 2 Cl−/1 H+ antiporter ClC-5 in a NO3−/H+ antiporter by a single point mutation. Embo J. 2009, 28, 175–182.
  56. Brandt, S.; Jentsch, T.J. ClC-6 and ClC-7 are two novel broadly expressed members of the CLC chloride channel family. FEBS Lett. 1995, 377, 15–20.
  57. Ignoul, S.; Simaels, J.; Hermans, D.; Annaert, W.; Eggermont, J. Human ClC-6 is a late endosomal glycoprotein that associates with detergent-resistant lipid domains. PLoS ONE 2007, 2, e474.
  58. Neagoe, I.; Stauber, T.; Fidzinski, P.; Bergsdorf, E.Y.; Jentsch, T.J. The late endosomal ClC-6 mediates proton/chloride countertransport in heterologous plasma membrane expression. J. Biol. Chem. 2010, 285, 21689–21697.
  59. Graves, A.R.; Curran, P.K.; Smith, C.L.; Mindell, J.A. The Cl−/H+ antiporter ClC-7 is the primary chloride permeation pathway in lysosomes. Nature 2008, 453, 788–792.
  60. Ludwig, C.F.; Ullrich, F.; Leisle, L.; Stauber, T.; Jentsch, T.J. Common gating of both CLC transporter subunits underlies voltage-dependent activation of the 2Cl−/1H+ exchanger ClC-7/Ostm1. J. Biol. Chem. 2013, 288, 28611–28619.
  61. Pusch, M.; Zifarelli, G. Large transient capacitive currents in wild-type lysosomal Cl−/H+ antiporter ClC-7 and residual transport activity in the proton glutamate mutant E312A. J. Gen. Physiol. 2021, 153, e202012583.
  62. Smith, A.J.; Lippiat, J.D. Voltage-dependent charge movement associated with activation of the CLC-5 2Cl−/1H+ exchanger. Faseb J. 2010, 24, 3696–3705.
  63. Zifarelli, G.; De Stefano, S.; Zanardi, I.; Pusch, M. On the mechanism of gating charge movement of ClC-5, a human Cl−/H+ antiporter. Biophys. J. 2012, 102, 2060–2069.
  64. Chalhoub, N.; Benachenhou, N.; Rajapurohitam, V.; Pata, M.; Ferron, M.; Frattini, A.; Villa, A.; Vacher, J. Grey-lethal mutation induces severe malignant autosomal recessive osteopetrosis in mouse and human. Nat. Med. 2003, 9, 399–406.
More
Information
Subjects: Biophysics
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 219
Revisions: 2 times (View History)
Update Date: 14 Jun 2023
1000/1000