Mol. Cells 2021; 44(10): 758-769
Published online October 29, 2021
https://doi.org/10.14348/molcells.2021.0131
© The Korean Society for Molecular and Cellular Biology
Correspondence to : physiolksj@gmail.com
This is an open-access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 Unported License. To view a copy of this license, visit http://creativecommons.org/licenses/by-nc-sa/3.0/.
Calcium homeostasis modulator 1 (CALHM1) is a membrane protein with four transmembrane helices that form an octameric ion channel with voltage-dependent activation. There are four conserved cysteine (Cys) residues in the extracellular domain that form two intramolecular disulfide bonds. We investigated the roles of C42-C127 and C44-C161 in human CALHM1 channel biogenesis and the ionic current (ICALHM1). Replacing Cys with Ser or Ala abolished the membrane trafficking as well as ICALHM1. Immunoblotting analysis revealed dithiothreitol-sensitive multimeric CALHM1, which was markedly reduced in C44S and C161S, but preserved in C42S and C127S. The mixed expression of C42S and wild-type did not show a dominant-negative effect. While the heteromeric assembly of CALHM1 and CALHM3 formed active ion channels, the co-expression of C42S and CALHM3 did not produce functional channels. Despite the critical structural role of the extracellular cysteine residues, a treatment with the membrane-impermeable reducing agent tris(2-carboxyethyl) phosphine (TCEP, 2 mM) did not affect ICALHM1 for up to 30 min. Interestingly, incubation with TCEP (2 mM) for 2-6 h reduced both ICALHM1 and the surface expression of CALHM1 in a time-dependent manner. We propose that the intramolecular disulfide bonds are essential for folding, oligomerization, trafficking and maintenance of CALHM1 in the plasma membrane, but dispensable for the voltage-dependent activation once expressed on the plasma membrane.
Keywords CALHM1, disulfide bond, membrane trafficking, oligomerization, reducing agent
Calcium homeostasis modulator 1 (CALHM1) is a newly identified voltage-gated nonselective ion channel, and its functional expression has been reported in cerebral cortical neuronal cells and taste bud cells (Dreses-Werringloer et al., 2008; Taruno et al., 2013). In cerebral neurons, CALHM1 is reportedly involved in the action potential frequency-dependent regulation of excitability (Vingtdeux et al., 2016). In type II taste bud cells, CALHM1 interacts with CALHM3 as a pore-forming subunit, and the CALHM1/3 hetero-multimeric channel functions as an endogenous ATP-release channel signaling to the purinoceptors in the sensory nerve ending (Kashio et al., 2019; Ma et al., 2018).
In whole-cell patch clamp studies, CALHM1-expressing cells showed slowly activating outward currents under strong depolarized conditions, and the voltage-dependence was shifted to the left by lowering the extracellular Ca2+ concentration ([Ca2+]e) (Ma et al., 2012). We have recently reported that both voltage dependence and activation speed were markedly potentiated by raising the temperature to physiological ranges (Jeon et al., 2021). Also, CALHM1 current activation shows tendency of sensitization by repetitive stimuli as like the voltage-dependent activation of Anoctamin 6/TMEM16F (Roh et al., 2021).
All six paralogues of the CALHM family proteins have four transmembrane domains with both N-and C-termini located on the cytoplasmic side. Recent cryo-EM studies have revealed octameric (CALHM1) and undecameric (CALHM2 and 4) assemblies with wide pore diameters consistent with the unselective permeability to various ion sizes, including ATP4- (Choi et al., 2019; Drozdzyk et al., 2020; Foskett, 2020; Syrjanen et al., 2020). There are four strictly conserved cysteine residues in the extracellular domains of human CALHM1, CALHM2, and CALHM3 (Figs. 1A and 1B). Based on the cryo-EM study, the conserved C42, C44, C127, and C161 residues in CALHM1 are linked by disulfide bonds (C42-C127 and C44-C161) (Demura et al., 2020; Ren et al., 2020) (Figs. 1C-1E). The supplementary data from a study of
The roles of intramolecular (between protein domains) and intermolecular (units of multimeric assemblage) disulfide bonds vary depending on the biogenesis, trafficking, and gating mechanisms (see Discussion section). Therefore, it is critical to elucidate the structural and functional roles of disulfide bonds in newly identified ion channels such as CALHM1. In the present study, we investigated the role of two disulfide bridges in the extracellular domains of human CALHM1 overexpressed in CHO cells. All four Cys mutants completely lost their channel activity and membrane expression. Although the disulfide bond breakage using the reducing agents dithiothreitol (DTT) and tris(2-carboxyethyl) phosphine (TCEP) did not acutely change the channel activity, sustained treatment with TCEP (≥2 h) significantly reduced
Chinese hamster ovary (CHO) and human embryonic kidney cell line (HEK293) cells were purchased from ATCC (USA) and incubated in Dulbecco’s modified Eagle’s medium (DMEM; Gibco, USA) supplemented with 10% fetal bovine serum (FBS; Gibco) and 1% penicillin-streptomycin (Gibco). The cells were incubated at 37°C and 5% CO2. The cells were sub-cultured by centrifuging at 160 ×
To insert the coding sequence (CDS), the cDNA of human CALHM1 and pEGFP-N1 vector (Takara Bio, USA) were truncated with the restriction enzymes
The transfected CHO cells were transferred to a bath mounted on the stage of an inverted microscope (
CALHM1-overexpressed cells were imaged on an 18 mm coverslip with an A1 confocal microscope (Nikon) one day after transfection. Cell-attached coverslips were fixed with 4% paraformaldehyde and washed three times with ice-cold phosphate-buffered saline (PBS). The membranes and nucleus of cells were stained in the dark using 5 µg/ml wheat germ agglutinin (WGA) and 1 µg/ml 4′,6-diamidino-2-phenylindole (DAPI), respectively. Images were scanned with a 100× immersion objective lens at 512 × 512 pixels using a digital zoom. All confocal images were processed and transferred using NIS software (Nikon) and ImageJ software (NIH, USA).
Proteins were extracted from overexpressed CHO cells using lysis buffer containing 0.5 M EDTA, 25 mM Tris-HCl, 150 mM NaCl, and 1% Triton X-100 with a phosphatase and protease inhibitor cocktail (Roche, Germany), pH 7.4. After scraping the cells and treating them with the lysis buffer, the lysates were homogenized 10 times with a sterile 26-gauge needle. The protein was quantified using the Bradford assay (Bio-Rad, USA). After incubation at 95°C with an appropriate concentration of DTT (indicated in the results), lysates were fractionated using SDS-PAGE on 3%-8% gradient gels (Thermo Fisher Scientific) and transferred onto a PVDF membrane in 25 mM Tris, 192 mM glycine, and 20% methanol. Membranes were blocked with 5% skim milk containing 1% TBS and Tween-20 (TBS-T) for 4 h at room temperature, with gentle rocking. Membranes were then incubated overnight at 4°C with anti-GFP (Invitrogen, USA) and anti-GAPDH (USA) primary antibodies, and this was followed by incubation with secondary antibodies after washing three times with TBS-T. Blots were developed using ECL Plus western blotting detection reagents (Merck, Germany) for western blot analysis.
To detect CALHM1 in the surface membrane fraction, CHO cells were washed with ice-cold PBS and incubated in 0.5 mg/well Sulfo-NHS-SS-Biotin (Thermo Fisher Scientific) for 1 h at 4°C. Unreacted biotin was quenched using 50 mM Tris (pH 7.4) and washed with ice-cold PBS. Biotinylated proteins were then extracted as described above. Avidin slurry (100 µl; Thermo Fisher Scientific) was added to 500 µl of cell lysate (500 µg of biotinylated protein) in spin columns and incubated, with rotation, for 1 h at room temperature. After washing the beads three times with TBS-T, proteins were eluted with elution buffer (62.5 mM Tris, 1% SDS, 10% glycerol, and 50 mM DTT). The eluted proteins were fractionated using SDS-PAGE on 4%-12% gradient gels and immunoblotted as described above.
To explore the biochemical interaction of WT CALHM1 and C42S CALHM1, after co-transfection for 24 h, cells were washed with ice-cold PBS and harvested in lysis buffer (120 mM NaCl, 50 mM HEPES, 2 mM EDTA, 2 mM MgCl2, and 0.5% Triton X-100 with a phosphatase and protease inhibitor cocktail, pH 7.4). The lysates were passed 17 times through a sterile 26-gauge needle and incubated on a rotator for 30 min at 4°C. After centrifugation, the protein concentration in the supernatants were measured and 200 µg of lysates (500 µl) were incubated with 30 µl of Protein G-agarose beads (Thermo Fisher Scientific) coupled to 1 µg of anti-FLAG M2 antibody (Sigma-Aldrich) or anti-GFP antibody (Thermo Fisher Scientific) on a rotator at 4°C overnight. After beads were washed twice with wash buffer and once with ice-cold PBS, the precipitates were eluted with 25 µl of 2× sample buffer (120 mM Tris-Cl, 4% SDS, 0.02% bromophenol blue, 20% glycerol, 10 mM DTT) at 60°C for 5 min. The lysates for input (10 µg) were incubated in sample buffer at 95°C for 5 min. All samples were fractionated using SDS-PAGE on 4%-12% gradient gels and immunoblotted as described above.
Statistical analysis was performed using Origin Pro 2021 (OriginLab Corporation, USA) and Prism 9.2.0 (GraphPad Software, USA). Data are presented as the mean ± SD. Student’s unpaired
The schematic figure of barrels and connecting lines demonstrates the topology and domains of CALHM1 (Fig. 1A). Amino acid sequences of the three human CALHM paralogues (CALHM1-3) and the calhm1 sequences of mice and killifish were compared (Fig. 1B). Four cysteine residues (C42, C44, C127, and C161 in human CALHM1) were conserved across all five proteins. The densities of the two disulfide bonds between the conserved cysteines (C42-C127 and C44-C161) in the killifish calhm1 cryo-EM structure were determined (Figs. 1C-1E). All these residues were located in the extracellular space, and the two disulfide bonds were located relatively close to each other (Figs. 1C and 1D, red color). Since the cryo-EM structure of human CALHM1 has not been reported before, we used the homologous model to confirm whether the less-conserved cysteine molecules also formed disulfide bonds. There were eight additional cysteine residues in cryo-EM structure, but they were too far from each other to form disulfide bonds (Fig. 1D, orange color). Taken together, we confirmed that only the four conserved cysteine residues formed the two disulfide bonds identified (Fig. 1E).
We compared the electrophysiological properties of cysteine mutants and wild-type (WT) CALHM1 in CHO cells. In the whole-cell patch clamp configuration, step-like depolarization was applied from –40 mV of holding voltage to 80 mV (5 s) with various levels, followed by repolarization to –40 mV. At room temperature, slowly activating outward currents (
Replacement of any of the four cysteines residues with serine or alanine residues by site-directed mutagenesis completely abolished
To investigate CALHM1 trafficking to the plasma membrane, we conducted confocal microscopic analysis of CHO cells expressing CALHM1 with an EGFP tag (Figs. 3A and 3B). The cells were co-stained with WGA, which binds to the glycoproteins in the cell membrane (Fig. 3A, red color). Wild-type CALHM1 was localized to the plasma membrane, along with the puncta form of intracellular expression (Fig. 3A, uppermost panel). Notably, a considerable amount of the CALHM1 protein was located in the perinuclear region. A pattern of relatively larger cytosolic expression of WT CALHM1 has also been reported (Dreses-Werringloer et al., 2008). In contrast to WT, none of the Cys mutants showed plasma membrane localization, while the perinuclear expression pattern was evident (Figs. 3A and 3B). We also compared the plasma membrane expressions of WT with the expression of cysteine mutants using a surface biotinylation technique. While WT CALHM1 was found in the plasma membrane fraction, none of the four cysteine mutants were detected (Fig. 3C). The multi-bands of WT and Cys mutants in the whole-cell lysates indicate variations of
We then evaluated whether intramolecular disulfide bonds are necessary for multimerization in the biogenesis of the CALHM1 channel. Immunoblotting using 3%-8% polyacrylamide gels was performed with or without a reducing agent (DTT, 0, 0.1, 1, and 10 mM). In the control condition, CALHM1 proteins of various sizes were detected, pointing to the presence of monomers, dimers, and tetramers (Fig. 4A). The calculated molecular weight of CALHM1-GFP is ~65.9 kDa, and the predicted molecular masses for dimeric and tetrameric forms are ~131.8 kDa and ~263.6 kDa, respectively. The multimeric forms of CALHM1 decreased with increasing concentrations of DTT, while the monomeric form increased. When either C42 or C127 was replaced with serine, multimeric bands were detected in the non-reducing condition (0 mM DTT), similar to WT (Fig. 4B). In contrast, the multimeric sizes of C44S and C161S decreased significantly, even under non-reducing conditions. We compared the intensity of monomer and tetramer bands in four immunoblots with increasing concentrations of DTT. Whereas the relative expression ratios of monomer gradually increased to 3.24 at 10 mM DTT in WT, those of tetramer decreased to 0.22, when normalized to signal intensity of 0 mM DTT (Fig. 4C). The ratios of monomer and tetramer bands in C42S and C127S were similar to WT (monomers, 3.12 and 2.97; tetramers, 0.28 and 0.14, respectively, Fig. 4D), while not significantly altered in C44S and C161S (monomers, 1.53 and 1.44; tetramers, 0.55 and 0.67, respectively).
Co-expression of WT with mutated channel units often showed dominant-negative effects, and hetero-multimerized channels showed impaired electrophysiological function (Bannister et al., 1999; Cho et al., 2000). Since the functional CALHM1 channel has an octameric structure, a putative hetero-multimeric assembly might disrupt the function of the channel. When WT and C42S cDNAs were co-transfected using constant quantities of WT and increasing quantities of C42S (10:1, 10:2, 10:5, and 10:10), the
To determine whether WT and C42S CALHM1 form hetero-multimerized channels, we performed co-expression and immunoprecipitation of FLAG tagged WT CALHM1 and GFP tagged C42S CALHM1 cDNAs with a positive control expressed with different (FLAG or GFP) tagged WT CALHM1 cDNAs. While CALHM1 co-immunoprecipitates with itself as a homo-multimer, GFP tagged C42S CALHM1 was not pulled down with FLAG tagged WT CALHM1, indicating C42S CALHM1 could not interact with WT CALHM1 (Fig. 5B).
The intrinsically expressed ATP-releasing channel in taste buds is composed of CALHM1 and CALHM3 (Kashio et al., 2019; Ma et al., 2018). When expressed with CALHM3 alone, CHO cells showed negligible voltage-dependent currents (Fig. 5C). However, when co-expressed with CALHM1, the voltage-dependent outward
Finally, we investigated whether reducing agents used to break the cysteine bridges could affect CALHM1 activity. Step depolarization from –40 to 60 mV (5 s) was repetitively applied to activate CHO cells expressing WT CALHM1 every 30 s. After confirming the constant amplitudes of
The structures of various transmembrane proteins, including ion channels, are stabilized by intra- or intermolecular disulfide bonds. In the CALHM family, four cysteine residues were highly conserved across and within species (Fig. 1B). Recent cryo-EM studies have confirmed the intramolecular disulfide bridges investigated here (Choi et al., 2019; Demura et al., 2020; Ren et al., 2020; Yang et al., 2020). Although the cryo-EM structure of human CALHM1 has not been reported, we indirectly confirmed the existence of two disulfide bonds between four conserved cysteine residues using homology modeling of CALHM1 based on human CALHM2 structure. These disulfide bonds are located in the extracellular space and occur approximately 7 Å apart, when measured using the homology models and killifish CALHM1 cryo-EM structure. Interestingly, eight non-conserved cysteine residues were also located in CALHM1 cryo-EM structure, but the distances between them were too large to form intermolecular or intramolecular disulfide bridges (Fig. 1D). This implies that only the four conserved cysteine residues could form two intramolecular disulfide bonds (Figs. 1C-1E).
The nascent channel proteins are targeted to the endoplasmic reticulum (ER) and are folded and multimerized within the ER prior to membrane expression (Deutsch, 2003; Isacoff et al., 2013). In the ER, oxidative protein folding, a process of disulfide bond formation, and a conformational folding reaction are generated (Narayan, 2012). It is very important that only properly folded and assembled channels are trafficked to the plasma membrane as dysfunctional channels can have serious effects. Starting with ER quality control, maturation in Golgi shuttled channel complexes is also important to proceed in forward traffic to reach the cell membrane.
Our study showed the critical roles of the intramolecular disulfide bonds (C42-C127 and C44-C161) in the biogenesis and membrane trafficking of CALHM1 ion channels. Specifically, we found that the cysteine bridge between C44 and C161 (C44-C161) is essential, while that between C42 and C127 (C42-C127) is dispensable for multimerization in the ER (Fig. 4). Nevertheless, the patch clamp and confocal microscopy demonstrated the necessity of both disulfide bonds as all mutants of the four individual cysteine residues showed impaired trafficking to the membrane, with no electrophysiological function (Figs. 2 and 3). We cautiously interpret that each disulfide bridge is critical to pass the quality check during CALHM1 localization to the plasma membrane.
While WT CALHM1 forms a homo-multimer, the single C42 mutation (C42S) lacking the C42-C127 bridge did not assemble into WT CALHM1 when it was co-expressed with WT CALHM1 (Figs. 5A and 5B). In addition, although CALHM1 and CALHM3 form a heteromeric ion channel (CALHM1/3) with higher activity than CALHM1 alone (Ma et al., 2018), the triple co-expression of WT CALHM1, CALHM3 and C42S CALHM1 did not affect the electrophysiological function (Figs. 5C and 5D). These results suggest that C42S might be unable to assemble with CALHM1/3 hetero-multimeric proteins.
Finally, we investigated the sensitivity of CALHM1 to DTT and TCEP. Surprisingly, despite the critical role of extracellular disulfide bridges in biogenesis and trafficking,
Because TCEP is impermeable to the plasma membrane, we investigated the effect of sustained treatment on ion channel activity and membrane expression of CALHM1. We observed a decrease in
The time gap between the nonsignificant acute effect on
A number of studies have reported that intermolecular or inter subunit disulfide bridges are important for the assembly of ion channels; shaker-type voltage-gated K+ channels (Schulteis et al., 1995; 1996), voltage-gated Na+ channels (Chen et al., 2012; Yereddi et al., 2013), voltage-gated Ca2+ channels (Calderon-Rivera et al., 2012), voltage-gated H+ channels (Fujiwara et al., 2013), and acid-sensing ion channels (Zha et al., 2009). Among the two-pore domain (K2P) K+ channel family, TWIK1 and TREK1 form a heterodimeric channel via a disulfide bridge in the extracellular cap region (Hwang et al., 2014). However, the role of disulfide bridges in the cap structure formed by the extracellular M1-P1 linkers of each K2P monomer remains controversial (Zuniga and Zuniga, 2016).
The critical role of intramolecular disulfide bridges in functional channel expression has been reported in other types of ion channels; inward rectifier K+ channels (Kir2.1 and 2.3) (Bannister et al., 1999; Cho et al., 2000), transient receptor potential channels (TRPC4 and 5) (Duan et al., 2018; 2019), transient receptor potential ankyrin channel (TRPA1) (Wang et al., 2012), TIM23 complex of mitochondria (Ramesh et al., 2016), chloride intracellular ion channel (CLIC1) (Al Khamici et al., 2016), and CALHM1, identified in this study. In contrast to the role of CALHM1 in membrane trafficking, the extracellular cysteine residues of Kir2.1 and 2.3 are required for their electrophysiological function but not for expression in the plasma membrane. Furthermore, the disulfide bond on the extracellular side of the pore and the preceding small loop of TRPC5 seem to be critical for the function and pharmacological activation of the agonist, Englerin A. The functional modulation of TRPC5 that results from changing the state of disulfide bonds using reducing agents has been reported (Xu et al., 2008). In contrast to TRPC5, acute treatment with a reducing agent did not affect
As discussed above, the roles of disulfide bridges in channel complexes are widely variable and include biogenesis, gating process, and pharmacological modulation. Our study confirmed the essential role of highly conserved intramolecular disulfide bridges in the proper assembly and trafficking of CALHM1. Although TCEP treatment did not acutely change the channel activity, prolonged exposure to the reducing environment could significantly lower CALHM1 expression, and this might have pathophysiological implications in the cells and tissues where the functions of CALHM1 are being revealed.
This work was supported by grants from the National Research Foundation of Korea (NRF-2018R1A5A2025964), EDISON (EDucation-research Integration through Simulation On the Net) Program (NRF-2016M3C1A6936605) and the Korea Health Technology R&D Project, through the Korea Health Industry Development Institute (KHIDI), funded by the Ministry of Health & Welfare, Republic of Korea (grant No. HP20C0199). This work was also supported by a grant from the M.D., Ph.D./Medical Scientist Training Programs through KHIDI to Y.K.J. We thank Prof. Chansik Hong for technical advices and helpful discussions.
J.W.K. and Y.K.J. conceived and performed the experiments. J.K. gave technical support. S.J.K. (Sang Jeong Kim) provided expertise and feedback. S.J.K. (Sung Joon Kim) wrote the manuscript and supervised the study.
The authors have no potential conflicts of interest to disclose.
Mol. Cells 2021; 44(10): 758-769
Published online October 31, 2021 https://doi.org/10.14348/molcells.2021.0131
Copyright © The Korean Society for Molecular and Cellular Biology.
Jae Won Kwon1,2,4 , Young Keul Jeon1,2,4
, Jinsung Kim1,2
, Sang Jeong Kim1,2
, and Sung Joon Kim1,2,3,*
1Department of Physiology, Seoul National University College of Medicine, Seoul 03080, Korea, 2Department of Biomedical Sciences, Seoul National University College of Medicine, Seoul 03080, Korea, 3Ischemic/Hypoxic Disease Institute, Seoul National University College of Medicine, Seoul 03080, Korea, 4These authors contributed equally to this work.
Correspondence to:physiolksj@gmail.com
This is an open-access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 Unported License. To view a copy of this license, visit http://creativecommons.org/licenses/by-nc-sa/3.0/.
Calcium homeostasis modulator 1 (CALHM1) is a membrane protein with four transmembrane helices that form an octameric ion channel with voltage-dependent activation. There are four conserved cysteine (Cys) residues in the extracellular domain that form two intramolecular disulfide bonds. We investigated the roles of C42-C127 and C44-C161 in human CALHM1 channel biogenesis and the ionic current (ICALHM1). Replacing Cys with Ser or Ala abolished the membrane trafficking as well as ICALHM1. Immunoblotting analysis revealed dithiothreitol-sensitive multimeric CALHM1, which was markedly reduced in C44S and C161S, but preserved in C42S and C127S. The mixed expression of C42S and wild-type did not show a dominant-negative effect. While the heteromeric assembly of CALHM1 and CALHM3 formed active ion channels, the co-expression of C42S and CALHM3 did not produce functional channels. Despite the critical structural role of the extracellular cysteine residues, a treatment with the membrane-impermeable reducing agent tris(2-carboxyethyl) phosphine (TCEP, 2 mM) did not affect ICALHM1 for up to 30 min. Interestingly, incubation with TCEP (2 mM) for 2-6 h reduced both ICALHM1 and the surface expression of CALHM1 in a time-dependent manner. We propose that the intramolecular disulfide bonds are essential for folding, oligomerization, trafficking and maintenance of CALHM1 in the plasma membrane, but dispensable for the voltage-dependent activation once expressed on the plasma membrane.
Keywords: CALHM1, disulfide bond, membrane trafficking, oligomerization, reducing agent
Calcium homeostasis modulator 1 (CALHM1) is a newly identified voltage-gated nonselective ion channel, and its functional expression has been reported in cerebral cortical neuronal cells and taste bud cells (Dreses-Werringloer et al., 2008; Taruno et al., 2013). In cerebral neurons, CALHM1 is reportedly involved in the action potential frequency-dependent regulation of excitability (Vingtdeux et al., 2016). In type II taste bud cells, CALHM1 interacts with CALHM3 as a pore-forming subunit, and the CALHM1/3 hetero-multimeric channel functions as an endogenous ATP-release channel signaling to the purinoceptors in the sensory nerve ending (Kashio et al., 2019; Ma et al., 2018).
In whole-cell patch clamp studies, CALHM1-expressing cells showed slowly activating outward currents under strong depolarized conditions, and the voltage-dependence was shifted to the left by lowering the extracellular Ca2+ concentration ([Ca2+]e) (Ma et al., 2012). We have recently reported that both voltage dependence and activation speed were markedly potentiated by raising the temperature to physiological ranges (Jeon et al., 2021). Also, CALHM1 current activation shows tendency of sensitization by repetitive stimuli as like the voltage-dependent activation of Anoctamin 6/TMEM16F (Roh et al., 2021).
All six paralogues of the CALHM family proteins have four transmembrane domains with both N-and C-termini located on the cytoplasmic side. Recent cryo-EM studies have revealed octameric (CALHM1) and undecameric (CALHM2 and 4) assemblies with wide pore diameters consistent with the unselective permeability to various ion sizes, including ATP4- (Choi et al., 2019; Drozdzyk et al., 2020; Foskett, 2020; Syrjanen et al., 2020). There are four strictly conserved cysteine residues in the extracellular domains of human CALHM1, CALHM2, and CALHM3 (Figs. 1A and 1B). Based on the cryo-EM study, the conserved C42, C44, C127, and C161 residues in CALHM1 are linked by disulfide bonds (C42-C127 and C44-C161) (Demura et al., 2020; Ren et al., 2020) (Figs. 1C-1E). The supplementary data from a study of
The roles of intramolecular (between protein domains) and intermolecular (units of multimeric assemblage) disulfide bonds vary depending on the biogenesis, trafficking, and gating mechanisms (see Discussion section). Therefore, it is critical to elucidate the structural and functional roles of disulfide bonds in newly identified ion channels such as CALHM1. In the present study, we investigated the role of two disulfide bridges in the extracellular domains of human CALHM1 overexpressed in CHO cells. All four Cys mutants completely lost their channel activity and membrane expression. Although the disulfide bond breakage using the reducing agents dithiothreitol (DTT) and tris(2-carboxyethyl) phosphine (TCEP) did not acutely change the channel activity, sustained treatment with TCEP (≥2 h) significantly reduced
Chinese hamster ovary (CHO) and human embryonic kidney cell line (HEK293) cells were purchased from ATCC (USA) and incubated in Dulbecco’s modified Eagle’s medium (DMEM; Gibco, USA) supplemented with 10% fetal bovine serum (FBS; Gibco) and 1% penicillin-streptomycin (Gibco). The cells were incubated at 37°C and 5% CO2. The cells were sub-cultured by centrifuging at 160 ×
To insert the coding sequence (CDS), the cDNA of human CALHM1 and pEGFP-N1 vector (Takara Bio, USA) were truncated with the restriction enzymes
The transfected CHO cells were transferred to a bath mounted on the stage of an inverted microscope (
CALHM1-overexpressed cells were imaged on an 18 mm coverslip with an A1 confocal microscope (Nikon) one day after transfection. Cell-attached coverslips were fixed with 4% paraformaldehyde and washed three times with ice-cold phosphate-buffered saline (PBS). The membranes and nucleus of cells were stained in the dark using 5 µg/ml wheat germ agglutinin (WGA) and 1 µg/ml 4′,6-diamidino-2-phenylindole (DAPI), respectively. Images were scanned with a 100× immersion objective lens at 512 × 512 pixels using a digital zoom. All confocal images were processed and transferred using NIS software (Nikon) and ImageJ software (NIH, USA).
Proteins were extracted from overexpressed CHO cells using lysis buffer containing 0.5 M EDTA, 25 mM Tris-HCl, 150 mM NaCl, and 1% Triton X-100 with a phosphatase and protease inhibitor cocktail (Roche, Germany), pH 7.4. After scraping the cells and treating them with the lysis buffer, the lysates were homogenized 10 times with a sterile 26-gauge needle. The protein was quantified using the Bradford assay (Bio-Rad, USA). After incubation at 95°C with an appropriate concentration of DTT (indicated in the results), lysates were fractionated using SDS-PAGE on 3%-8% gradient gels (Thermo Fisher Scientific) and transferred onto a PVDF membrane in 25 mM Tris, 192 mM glycine, and 20% methanol. Membranes were blocked with 5% skim milk containing 1% TBS and Tween-20 (TBS-T) for 4 h at room temperature, with gentle rocking. Membranes were then incubated overnight at 4°C with anti-GFP (Invitrogen, USA) and anti-GAPDH (USA) primary antibodies, and this was followed by incubation with secondary antibodies after washing three times with TBS-T. Blots were developed using ECL Plus western blotting detection reagents (Merck, Germany) for western blot analysis.
To detect CALHM1 in the surface membrane fraction, CHO cells were washed with ice-cold PBS and incubated in 0.5 mg/well Sulfo-NHS-SS-Biotin (Thermo Fisher Scientific) for 1 h at 4°C. Unreacted biotin was quenched using 50 mM Tris (pH 7.4) and washed with ice-cold PBS. Biotinylated proteins were then extracted as described above. Avidin slurry (100 µl; Thermo Fisher Scientific) was added to 500 µl of cell lysate (500 µg of biotinylated protein) in spin columns and incubated, with rotation, for 1 h at room temperature. After washing the beads three times with TBS-T, proteins were eluted with elution buffer (62.5 mM Tris, 1% SDS, 10% glycerol, and 50 mM DTT). The eluted proteins were fractionated using SDS-PAGE on 4%-12% gradient gels and immunoblotted as described above.
To explore the biochemical interaction of WT CALHM1 and C42S CALHM1, after co-transfection for 24 h, cells were washed with ice-cold PBS and harvested in lysis buffer (120 mM NaCl, 50 mM HEPES, 2 mM EDTA, 2 mM MgCl2, and 0.5% Triton X-100 with a phosphatase and protease inhibitor cocktail, pH 7.4). The lysates were passed 17 times through a sterile 26-gauge needle and incubated on a rotator for 30 min at 4°C. After centrifugation, the protein concentration in the supernatants were measured and 200 µg of lysates (500 µl) were incubated with 30 µl of Protein G-agarose beads (Thermo Fisher Scientific) coupled to 1 µg of anti-FLAG M2 antibody (Sigma-Aldrich) or anti-GFP antibody (Thermo Fisher Scientific) on a rotator at 4°C overnight. After beads were washed twice with wash buffer and once with ice-cold PBS, the precipitates were eluted with 25 µl of 2× sample buffer (120 mM Tris-Cl, 4% SDS, 0.02% bromophenol blue, 20% glycerol, 10 mM DTT) at 60°C for 5 min. The lysates for input (10 µg) were incubated in sample buffer at 95°C for 5 min. All samples were fractionated using SDS-PAGE on 4%-12% gradient gels and immunoblotted as described above.
Statistical analysis was performed using Origin Pro 2021 (OriginLab Corporation, USA) and Prism 9.2.0 (GraphPad Software, USA). Data are presented as the mean ± SD. Student’s unpaired
The schematic figure of barrels and connecting lines demonstrates the topology and domains of CALHM1 (Fig. 1A). Amino acid sequences of the three human CALHM paralogues (CALHM1-3) and the calhm1 sequences of mice and killifish were compared (Fig. 1B). Four cysteine residues (C42, C44, C127, and C161 in human CALHM1) were conserved across all five proteins. The densities of the two disulfide bonds between the conserved cysteines (C42-C127 and C44-C161) in the killifish calhm1 cryo-EM structure were determined (Figs. 1C-1E). All these residues were located in the extracellular space, and the two disulfide bonds were located relatively close to each other (Figs. 1C and 1D, red color). Since the cryo-EM structure of human CALHM1 has not been reported before, we used the homologous model to confirm whether the less-conserved cysteine molecules also formed disulfide bonds. There were eight additional cysteine residues in cryo-EM structure, but they were too far from each other to form disulfide bonds (Fig. 1D, orange color). Taken together, we confirmed that only the four conserved cysteine residues formed the two disulfide bonds identified (Fig. 1E).
We compared the electrophysiological properties of cysteine mutants and wild-type (WT) CALHM1 in CHO cells. In the whole-cell patch clamp configuration, step-like depolarization was applied from –40 mV of holding voltage to 80 mV (5 s) with various levels, followed by repolarization to –40 mV. At room temperature, slowly activating outward currents (
Replacement of any of the four cysteines residues with serine or alanine residues by site-directed mutagenesis completely abolished
To investigate CALHM1 trafficking to the plasma membrane, we conducted confocal microscopic analysis of CHO cells expressing CALHM1 with an EGFP tag (Figs. 3A and 3B). The cells were co-stained with WGA, which binds to the glycoproteins in the cell membrane (Fig. 3A, red color). Wild-type CALHM1 was localized to the plasma membrane, along with the puncta form of intracellular expression (Fig. 3A, uppermost panel). Notably, a considerable amount of the CALHM1 protein was located in the perinuclear region. A pattern of relatively larger cytosolic expression of WT CALHM1 has also been reported (Dreses-Werringloer et al., 2008). In contrast to WT, none of the Cys mutants showed plasma membrane localization, while the perinuclear expression pattern was evident (Figs. 3A and 3B). We also compared the plasma membrane expressions of WT with the expression of cysteine mutants using a surface biotinylation technique. While WT CALHM1 was found in the plasma membrane fraction, none of the four cysteine mutants were detected (Fig. 3C). The multi-bands of WT and Cys mutants in the whole-cell lysates indicate variations of
We then evaluated whether intramolecular disulfide bonds are necessary for multimerization in the biogenesis of the CALHM1 channel. Immunoblotting using 3%-8% polyacrylamide gels was performed with or without a reducing agent (DTT, 0, 0.1, 1, and 10 mM). In the control condition, CALHM1 proteins of various sizes were detected, pointing to the presence of monomers, dimers, and tetramers (Fig. 4A). The calculated molecular weight of CALHM1-GFP is ~65.9 kDa, and the predicted molecular masses for dimeric and tetrameric forms are ~131.8 kDa and ~263.6 kDa, respectively. The multimeric forms of CALHM1 decreased with increasing concentrations of DTT, while the monomeric form increased. When either C42 or C127 was replaced with serine, multimeric bands were detected in the non-reducing condition (0 mM DTT), similar to WT (Fig. 4B). In contrast, the multimeric sizes of C44S and C161S decreased significantly, even under non-reducing conditions. We compared the intensity of monomer and tetramer bands in four immunoblots with increasing concentrations of DTT. Whereas the relative expression ratios of monomer gradually increased to 3.24 at 10 mM DTT in WT, those of tetramer decreased to 0.22, when normalized to signal intensity of 0 mM DTT (Fig. 4C). The ratios of monomer and tetramer bands in C42S and C127S were similar to WT (monomers, 3.12 and 2.97; tetramers, 0.28 and 0.14, respectively, Fig. 4D), while not significantly altered in C44S and C161S (monomers, 1.53 and 1.44; tetramers, 0.55 and 0.67, respectively).
Co-expression of WT with mutated channel units often showed dominant-negative effects, and hetero-multimerized channels showed impaired electrophysiological function (Bannister et al., 1999; Cho et al., 2000). Since the functional CALHM1 channel has an octameric structure, a putative hetero-multimeric assembly might disrupt the function of the channel. When WT and C42S cDNAs were co-transfected using constant quantities of WT and increasing quantities of C42S (10:1, 10:2, 10:5, and 10:10), the
To determine whether WT and C42S CALHM1 form hetero-multimerized channels, we performed co-expression and immunoprecipitation of FLAG tagged WT CALHM1 and GFP tagged C42S CALHM1 cDNAs with a positive control expressed with different (FLAG or GFP) tagged WT CALHM1 cDNAs. While CALHM1 co-immunoprecipitates with itself as a homo-multimer, GFP tagged C42S CALHM1 was not pulled down with FLAG tagged WT CALHM1, indicating C42S CALHM1 could not interact with WT CALHM1 (Fig. 5B).
The intrinsically expressed ATP-releasing channel in taste buds is composed of CALHM1 and CALHM3 (Kashio et al., 2019; Ma et al., 2018). When expressed with CALHM3 alone, CHO cells showed negligible voltage-dependent currents (Fig. 5C). However, when co-expressed with CALHM1, the voltage-dependent outward
Finally, we investigated whether reducing agents used to break the cysteine bridges could affect CALHM1 activity. Step depolarization from –40 to 60 mV (5 s) was repetitively applied to activate CHO cells expressing WT CALHM1 every 30 s. After confirming the constant amplitudes of
The structures of various transmembrane proteins, including ion channels, are stabilized by intra- or intermolecular disulfide bonds. In the CALHM family, four cysteine residues were highly conserved across and within species (Fig. 1B). Recent cryo-EM studies have confirmed the intramolecular disulfide bridges investigated here (Choi et al., 2019; Demura et al., 2020; Ren et al., 2020; Yang et al., 2020). Although the cryo-EM structure of human CALHM1 has not been reported, we indirectly confirmed the existence of two disulfide bonds between four conserved cysteine residues using homology modeling of CALHM1 based on human CALHM2 structure. These disulfide bonds are located in the extracellular space and occur approximately 7 Å apart, when measured using the homology models and killifish CALHM1 cryo-EM structure. Interestingly, eight non-conserved cysteine residues were also located in CALHM1 cryo-EM structure, but the distances between them were too large to form intermolecular or intramolecular disulfide bridges (Fig. 1D). This implies that only the four conserved cysteine residues could form two intramolecular disulfide bonds (Figs. 1C-1E).
The nascent channel proteins are targeted to the endoplasmic reticulum (ER) and are folded and multimerized within the ER prior to membrane expression (Deutsch, 2003; Isacoff et al., 2013). In the ER, oxidative protein folding, a process of disulfide bond formation, and a conformational folding reaction are generated (Narayan, 2012). It is very important that only properly folded and assembled channels are trafficked to the plasma membrane as dysfunctional channels can have serious effects. Starting with ER quality control, maturation in Golgi shuttled channel complexes is also important to proceed in forward traffic to reach the cell membrane.
Our study showed the critical roles of the intramolecular disulfide bonds (C42-C127 and C44-C161) in the biogenesis and membrane trafficking of CALHM1 ion channels. Specifically, we found that the cysteine bridge between C44 and C161 (C44-C161) is essential, while that between C42 and C127 (C42-C127) is dispensable for multimerization in the ER (Fig. 4). Nevertheless, the patch clamp and confocal microscopy demonstrated the necessity of both disulfide bonds as all mutants of the four individual cysteine residues showed impaired trafficking to the membrane, with no electrophysiological function (Figs. 2 and 3). We cautiously interpret that each disulfide bridge is critical to pass the quality check during CALHM1 localization to the plasma membrane.
While WT CALHM1 forms a homo-multimer, the single C42 mutation (C42S) lacking the C42-C127 bridge did not assemble into WT CALHM1 when it was co-expressed with WT CALHM1 (Figs. 5A and 5B). In addition, although CALHM1 and CALHM3 form a heteromeric ion channel (CALHM1/3) with higher activity than CALHM1 alone (Ma et al., 2018), the triple co-expression of WT CALHM1, CALHM3 and C42S CALHM1 did not affect the electrophysiological function (Figs. 5C and 5D). These results suggest that C42S might be unable to assemble with CALHM1/3 hetero-multimeric proteins.
Finally, we investigated the sensitivity of CALHM1 to DTT and TCEP. Surprisingly, despite the critical role of extracellular disulfide bridges in biogenesis and trafficking,
Because TCEP is impermeable to the plasma membrane, we investigated the effect of sustained treatment on ion channel activity and membrane expression of CALHM1. We observed a decrease in
The time gap between the nonsignificant acute effect on
A number of studies have reported that intermolecular or inter subunit disulfide bridges are important for the assembly of ion channels; shaker-type voltage-gated K+ channels (Schulteis et al., 1995; 1996), voltage-gated Na+ channels (Chen et al., 2012; Yereddi et al., 2013), voltage-gated Ca2+ channels (Calderon-Rivera et al., 2012), voltage-gated H+ channels (Fujiwara et al., 2013), and acid-sensing ion channels (Zha et al., 2009). Among the two-pore domain (K2P) K+ channel family, TWIK1 and TREK1 form a heterodimeric channel via a disulfide bridge in the extracellular cap region (Hwang et al., 2014). However, the role of disulfide bridges in the cap structure formed by the extracellular M1-P1 linkers of each K2P monomer remains controversial (Zuniga and Zuniga, 2016).
The critical role of intramolecular disulfide bridges in functional channel expression has been reported in other types of ion channels; inward rectifier K+ channels (Kir2.1 and 2.3) (Bannister et al., 1999; Cho et al., 2000), transient receptor potential channels (TRPC4 and 5) (Duan et al., 2018; 2019), transient receptor potential ankyrin channel (TRPA1) (Wang et al., 2012), TIM23 complex of mitochondria (Ramesh et al., 2016), chloride intracellular ion channel (CLIC1) (Al Khamici et al., 2016), and CALHM1, identified in this study. In contrast to the role of CALHM1 in membrane trafficking, the extracellular cysteine residues of Kir2.1 and 2.3 are required for their electrophysiological function but not for expression in the plasma membrane. Furthermore, the disulfide bond on the extracellular side of the pore and the preceding small loop of TRPC5 seem to be critical for the function and pharmacological activation of the agonist, Englerin A. The functional modulation of TRPC5 that results from changing the state of disulfide bonds using reducing agents has been reported (Xu et al., 2008). In contrast to TRPC5, acute treatment with a reducing agent did not affect
As discussed above, the roles of disulfide bridges in channel complexes are widely variable and include biogenesis, gating process, and pharmacological modulation. Our study confirmed the essential role of highly conserved intramolecular disulfide bridges in the proper assembly and trafficking of CALHM1. Although TCEP treatment did not acutely change the channel activity, prolonged exposure to the reducing environment could significantly lower CALHM1 expression, and this might have pathophysiological implications in the cells and tissues where the functions of CALHM1 are being revealed.
This work was supported by grants from the National Research Foundation of Korea (NRF-2018R1A5A2025964), EDISON (EDucation-research Integration through Simulation On the Net) Program (NRF-2016M3C1A6936605) and the Korea Health Technology R&D Project, through the Korea Health Industry Development Institute (KHIDI), funded by the Ministry of Health & Welfare, Republic of Korea (grant No. HP20C0199). This work was also supported by a grant from the M.D., Ph.D./Medical Scientist Training Programs through KHIDI to Y.K.J. We thank Prof. Chansik Hong for technical advices and helpful discussions.
J.W.K. and Y.K.J. conceived and performed the experiments. J.K. gave technical support. S.J.K. (Sang Jeong Kim) provided expertise and feedback. S.J.K. (Sung Joon Kim) wrote the manuscript and supervised the study.
The authors have no potential conflicts of interest to disclose.
Tiffany W. Todd, and Janghoo Lim
Mol. Cells 2013; 36(3): 185-194 https://doi.org/10.1007/s10059-013-0167-xChang Ho Kang, Byeong Cheol Moon, Hyeong Cheol Park, Sung Cheol Koo, Yong Hun Chi, Yong Hwa Cheong, Byung-Dae Yoon, Sang Yeol Lee, and Cha Young Kim
Mol. Cells 2013; 35(5): 381-387 https://doi.org/10.1007/s10059-013-2185-0Jae Yoon Shin, Jae Il Shin, Jun Seob Kim, Yoo Soo Yang, Yeon-Kyun Shin, Kyeong Kyu Kim,
Sangho Lee, and Dae-Hyuk Kweon