Abstract
The inositol 1,4,5-trisphosphate (IP3) receptors (IP3Rs), which form tetrameric channels, play pivotal roles in regulating the spatiotemporal patterns of intracellular calcium signals. Mutations in IP3Rs have been increasingly associated with many debilitating human diseases such as ataxia, Gillespie syndrome, and generalized anhidrosis. However, how these mutations affect IP3R function, and how the perturbation of as-sociated calcium signals contribute to the pathogenesis and severity of these diseases remains largely uncharacterized. Moreover, many of these diseases occur as the result of autosomal dominant inheritance, suggesting that WT and mutant subunits associate in heterotetrameric channels. How the in-corporation of different numbers of mutant subunits within the tetrameric channels affects its activities and results in different disease phenotypes is also unclear. In this report, we investigated representative disease-associated missense mutations to determine their effects on IP3R channel activity. Additionally, we designed concatenated IP3R constructs to create tetrameric channels with a predefined subunit composition to explore the functionality of heteromeric channels. Using calcium imaging techniques to assess IP3R channel function, we observed that all the mutations studied resulted in severely attenuated Ca2+ release when expressed as homotetramers. However, some heterotetramers retained varied degrees of function dependent on the composition of the tetramer. Our findings suggest that the effect of mutations depends on the location of the mutation in the IP3R structure, as well as on the stoichiometry of mutant subunits assembled within the tetrameric channel. These studies provide insight into the pathogenesis and penetrance of these devastating human diseases.
Keywords: calcium intracellular release; calcium imaging; signal transduction; calcium signaling; inositol 1,4,5-trisphosphate receptor (IP3R); spinocerebellar ataxia (SCA); Gillespie syndrome (GS); anhidrosis; calcium-intracellular release; calcium channel; inositol 1,4,5-trisphosphate (IP3); inositol trisphosphate receptor (InsP3R); imaging; spinocerebellar ataxia
Inositol 1,4,5-trisphosphate (IP3) receptor (IP3R)-mediated calcium (Ca2+) release regulates many cellular activities including proliferation, secretion, division, contraction, and even apoptosis (1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11). There are three IP3R genes in the human genome (ITPR1, ITPR2, and ITPR3). The resulting monomeric isoforms of IP3R (IP3R1, IP3R2, and IP3R3) share ∼60–70% sequence homology and form homo- and heterotetrameric Ca2+ channels predominantly located in membranes of intracellular Ca2+ stores (3, 12, 13, 14, 15, 16, 17, 18). In response to various extracellular stimuli that induce phospholipase C-mediated hydrolysis of phosphatidylinositol 4,5-bisphosphate, IP3 is generated and diffuses through the cytosol to bind and subsequently activate IP3R channels resulting in movement of Ca2+ down its concentration gradient into the cytosol (19, 20). Each monomeric IP3R isoform is a ∼300 kDa protein and is generally divided into four functional domains (Fig. 1). The N-terminal suppressor domain (SD; 1-225 aa)1 (21) and ligand-binding domain (LBD; 226-578 aa) (22), the central regulatory and coupling domain (579-2232 aa) (23), and the C-terminal channel domain containing six membrane spanning regions and a cytosolic tail (2233-2710 aa) (24, 25, 26). It is believed that IP3 binding to the N terminus of IP3R triggers conformational changes that are transmitted intramolecularly or intermolecularly to open the channel pore located at the C terminus (27, 28, 29, 30); however, the precise mechanism of this gating is uncertain (31).
IP3R genes are ubiquitously expressed, although various tissues and cell types often preferentially express one subtype or a combination of subtypes (32, 33, 34, 35). Despite the overlapping expression patterns, these receptors can mediate distinct functions depending on the cell's physiological context. IP3R1 is highly enriched in the nervous system, whereas IP3R2 and IP3R3 are predominately expressed in the periphery including the digestive system and secretory cells of the exocrine system (36, 37, 38). Studies have shown that IP3R1 is essential for normal nervous system development as mice lacking IP3R1 often die in utero or suffer from severe ataxia and seizure prior to death very early in postnatal life (39, 40, 41). Mice lacking either IP3R2 or IP3R3 appear to develop only subtle phenotypes. However, mice lacking both IP3R2 and IP3R3 suffer from severe salivary and pancreatic insufficiencies and die after weaning due to inability to assimilate macronutrients from dry food (42).
Many devastating human diseases have been associated with abnormal Ca2+ homeostasis. Indeed, there is a growing body of literature illustrating that diseases, such as Huntington's disease, schizophrenia, ALS, Parkinson's disease, Alzheimer's disease, and spinocerebellar ataxia (SCA), are all accompanied by dysregulated Ca2+ signals (43, 44, 45, 46). Although the underlying etiologies of these diseases are heterogeneous, many studies have shown that perturbed IP3R signaling is among the contributing or causal factors. Mutations in IP3R genes have been associated with several human diseases, including SCA (47, 48, 49, 50, 51, 52, 53), Gillespie syndrome (GS) (52, 54), and anhidrosis (55). These mutations have generally been identified through whole exome sequencing and linkage analysis in a clinical setting. Nevertheless, the molecular and cellular consequences of many of these mutations, as well as the mechanism whereby individual mutations lead to the clinical manifestations, remain largely unknown. As mutations have been identified in all four functional domains of the IP3R, it is conceivable that mutations can interfere with various aspects of IP3R biology including IP3 binding, allosteric regulation by various cellular factors, ion permeation, protein folding, stability, and subcellular localization. It is likely that in patients, tetrameric IP3R channels are assembled from a combination of both WT and mutant subunits and given the autosomal dominant inheritance of the majority of cases, the incorporation of mutant subunits likely results in dominant-negative activity. Nevertheless, how the specific stoichiometry of the resultant heterotetrameric channels might affect function and lead to different clinical manifestations and disease penetrance have not been explored.
In this study, we have examined the consequences of a representative set of missense mutations on IP3R channel function that are illustrative of mutations in distinct functional domains of the IP3R and of differing modes of inheritance. Missense mutations of residues Arg-269 (R269W) (50, 51, 52, 53) and Asn-602 (N602D) (47, 48, 52) both lead to multiple cases of autosomal dominant, early-onset, nonprogressive ataxia, even though they are located in the LBD and regulatory/coupling domain, respectively (Fig. 1, A and B). In contrast, the G2498S mutation, located in the C-terminal selectivity filter of the IP3R2 channel pore (Fig. 1C), is inherited in an autosomal recessive fashion such that individuals homozygous for the mutation suffered from generalized anhidrosis, whereas their family members heterozygous for the mutation were asymptomatic for the disease (55). We utilized two heterologous expression systems, which are unambiguously IP3R null (56, 57) to directly investigate how expression of these mutant channels affect channel expression and localization, tetramer formation, IP3 binding, channel gating, Ca2+ store content, and Ca2+ release activities. To investigate how incorporation of WT and disease-associated mutant subunits into the assembled receptor alters channel function, we also used a concatenated subunit approach that allows for a defined subunit stoichiometry. Our data indicate that the impact of missense mutations is domain dependent and markedly impacted by the stoichiometry of incorporation into the functional heterotetrameric channel. These studies provide a framework to understand the molecular basis of human IP3R pathologies and offer novel insight into roles of key residues in IP3R channel activities.
Results
IP3R1 is highly expressed in the nervous system (35) and mutations result in a broad range of disease phenotypes: ataxia alone or ataxia with cognitive difficulties, iris hypoplasia and/or cerebellar hypotrophy. Indeed, there is a growing list of missense (47, 48, 49, 50, 51, 52, 53, 54, 55), nonsense, insertion/deletion, and splice mutations (58, 59) scattered throughout the IP3R1 gene (Fig. 1). Nevertheless, how IP3R1 mutations affect channel activity and how these mutations lead to pathogenesis remains to be elucidated. Here we investigated in detail, the functional consequences of three representative mutations located in different functional domains of IP3R following expression of homotetrameric or defined heterotetrameric mutant channels in an IP3R null background.
Mutation of R269W attenuates IP3R activity by reducing IP3 binding
R269W has been identified in patients diagnosed with autosomal dominant, nonprogressive congenital ataxia (NPCA) (50), spinocerebellar ataxia 29 (SCA29) (52), ataxic cerebral palsy (51), and early-onset ataxia (EOA) (53). These patients exhibit nonprogressive early-onset ataxia, hypotonia, and developmental delays (Table S1). Residue Arg-269 is located in the ligand-binding core of the IP3R1 (Table S2) that also constitutes the IP3R-binding protein released with inositol 1,4,5-trisphosphate (IRBIT)-binding domain (60, 61). This mutation (R269W) substitutes an evolutionarily conserved (Fig. 2A), positively charged, arginine residue, with a large hydrophobic, aromatic, tryptophan residue (Fig. 2B). This residue has been previously identified as one of the 12 positively charged amino acids absolutely required for IP3 binding and early studies have demonstrated that mutation of arginine to glutamine at this site (R269Q) resulted in an 80% loss of IP3 binding (22, 62). We therefore investigated the consequences of the R269W mutation on IP3 binding. WT, rat IP3R1 (WT rIP3R1), and rat cDNA engineered to express the R269W mutation (rIP3R1RW) were stably transfected into DT40-3KO cells (56) (Fig. 2C). [3H]IP3 binding was measured by competitive binding assays. When exogenously expressed in DT40-3KO, rIP3R1RW IP3-binding activity was severely compromised compared with WT rIP3R1 (Fig. 2D). Total specific binding was reduced by 60%, whereas the apparent EC50 for binding was not markedly altered under these conditions. This latter observation is surprising but may reflect severe alterations in the binding kinetics (either an alteration of on- or off-rate) such that equilibrium binding is not reached for the rIP3R1RW mutant under the same conditions as the WT rIP3R1.
Next, we assessed the ability of rIP3R1RW to support IP3-induced Ca2+ release when stably expressed in DT40-3KO cells. In single cell imaging experiments, stimulation with trypsin, which activates the Gαq-coupled protease-activated receptor 2 (PAR2), generates IP3 and results in Ca2+ release in cells expressing rIP3R1 but not in IP3R null, DT40-3KO (Fig. 2E) (56, 57). Similarly, whereas stimulation of cells expressing rIP3R1 receptors mediated robust Ca2+ signals, cells expressing rIP3R1RW did not exhibit any measurable response to a maximal concentration of trypsin (Fig. 2E). The pooled data illustrates that despite retaining modest residual binding activity, in terms of the magnitude of signal (Fig. 2F), and the number of cells that respond above a defined threshold (Fig. 2G), rIP3R1RW is refractory to stimulation in DT40-3KO cells. Additionally, multiple, distinct clonal rIP3R1RW cell lines generated in DT40-3KO cells exhibited similar de-creases in amplitude (data not shown). This lack of response may be related to the requirement of each IP3R monomer to be occupied by IP3 for activation (57), which given the reduced binding, would be predicted to limit this occurrence.
In support of the idea that assemblies of mutant (binding deficient) IP3R1 and WT IP3R1 are generated in cells, we performed co-immunoprecipitation experiments in cells transfected with a mCherry-tagged rIP3R1 and IP3R1RW. The IP3R1RW mutant was readily detected in samples immunoprecipitated with a mCherry antibody (Fig. S1A) and thus heterotetrameric channels containing both WT and mutant subunits were formed.
A prediction from our previous data (57) is that heterotetramers consisting of rIP3R1 and rIP3R1RW would also fail to respond to stimulation. To test this idea, we generated cDNA encoding concatenated dimers consisting of rIP3R1 and rIP3R1RW expressed as a single polypeptide chain of IP3R1-IP3R1RW (R1/R1RW) or IP3R1RW-IP3R1 (R1RW/R1) and stably expressed these constructs in DT40-3KO cells (Fig. 3A). Dimers are assembled by the cell to produce tetrameric channels consisting of two IP3R1 and two IP3R1RW subunits. Cells expressing concatenated dimers encoding only rIP3R1 (R1/R1) responded to PAR2 stimulation with robust Ca2+ signals, as previously reported (38, 57), whereas cells expressing concatenated dimers encoding rIP3R1RW (R1RW/R1RW) failed to respond (Fig. 3B). Additionally, cells expressing R1/R1RW or R1RW/R1 dimers also failed to respond to trypsin stimulation (Fig. 3B). The pooled data illustrates that despite some R1/R1RW or R1RW/R1 dimers exhibiting a very small increase in the magnitude of the signal (Fig. 3C), the number of cells that respond above to a defined threshold remained not significantly different from the null DT40-3KO cell line (Fig. 3D). Additional R1/R1RW, R1RW/R1, and R1RW/R1RW dimeric clonal cell lines generated in DT40-3KO cells exhibited similar decreases in amplitude and percentage of cells responding above a specific threshold (data not shown). Consistent with these observations, IP3 also failed to evoke Ca2+ release in permeabilized DT40-3KO cells stably expressing channels containing two or more rIP3R1RW subunits (Fig. 3E). These data indicate that tetramers containing mutant subunits likely are poorly functional. Thus, the implication is that patients with heterozygous expression of the mutant protein, only tetramers consisting of nonmutant IP3R1 subunits would support significant function.
Although the DT40-3KO chicken lymphocyte cell line has been used for last two decades to probe IP3R biology, we have recently used CRISPR/Cas9 technology to disrupt all three IP3R genes in human embryonic kidney (HEK293) cells to generate IP3R null HEK293 cells (HEK-3KO) (57). Tangible advantages of these cells include their mammalian origin, with presumably more relevant cellular milieu in terms of potential regulatory factors, and that HEK293 cells can be easily maintained and transfected. Therefore, we next asked if the R269W mutant behaves similarly in HEK-3KO compared with DT40 cells. Stable HEK-3KO cell lines were generated expressing WT rIP3R1, rIP3R1RW, or hIP3R1RW (Fig. 4A). Stable expression resulted in an obvious reticular pattern of expression, not obviously different from the localization of WT IP3R1 (Fig. 4B). Expression of rIP3R1RW did not result in alteration of the apparent content of the ER Ca2+ store, as assessed by the Ca2+ released following exposure to a SERCA pump inhibitor in a Ca2+-free extracellular buffer (Fig. S2, A and B). See supplemental experimental procedures. In single cell assays, stimulation of HEK-3KO expressing WT rIP3R1 with 5, 50, or 500 nm trypsin generated robust Ca2+ release (Fig. 4, C and D) with nearly 100% of cells responding (Fig. 4E). Notably, stimulation of cells expressing rIP3R1RW also resulted in measurable Ca2+ increases (Fig. 4, C and D), albeit, at much lower maximal amplitudes and with a lower percentage of responding cells responding (Fig. 4E). A small increase in amplitude was likewise seen in other clonal cell lines expressing rIP3R1RW and hIP3RRW in HEK-3KO cells (data not shown). These findings were recapitulated in populations of cells where [Ca2+] changes were monitored in response to varying [trypsin] using a Flexstation3 plate reader (Fig. 4F). Additionally, the corresponding mutation in the human IP3R1 (hIP3R1RW) stably expressed in HEK-3KO cells displayed a similar marked loss of function when compared with trypsin-induced release evoked in HEK-3KO cells stably expressing hIP3R1 (Fig. 4F). The seemingly disparate findings obtained upon expression of R269W mutations in DT40-3KO and HEK-3KO could reflect differences in stable expression level of the mutant IP3R in HEK-3KO cells compared with DT40-3KOs (Fig. S3A). In total, these findings show in two expression systems that R269W exhibits compromised IP3 binding that results in severely reduced channel activity.
Mutation of N602D attenuates IP3R1 activity without altering IP3 binding
The N602D mutation in IP3R1 has been identified in two unrelated families and is associated with diagnoses of ataxic cerebral palsy without cerebral atrophy (48) and autosomal dominant congenital nonprogressive spinocerebellar ataxia (47, 52) (Table S1). Asn-602 is conserved among the three IP3R subtypes and across many species (Fig. 5A). The residue is located just outside of the LBD, but formally within the IRBIT-binding domain (Table S3) and serves to connect two α-helical regions of the receptor (Fig. 5B). The mutation introduces a negatively charged residue at this site, replacing an asparagine residue with a polar neutral side chain with an aspartate residue harboring a polar acidic side chain. DT40-3KO cells stably expressing WT rIP3R1 or rIP3R1 expressing N602D (rIP3R1ND) mutations were generated (Fig. 5C). To establish the effects of this mutation on channel function, we first assessed IP3 binding to the mutant proteins. In a competitive binding assay, rIP3R1ND exhibited binding comparable with that of WT rIP3R1, suggesting that the mutation does not interfere with IP3 binding despite the proximity of the mutation to the LBD (Fig. 5D). To examine the consequences of the mutation on Ca2+ release activities, cells expressing WT rIP3R1 or rIP3R1ND were stimulated with the PAR agonist, trypsin. Trypsin did not elicit any measurable [Ca2+] increase in cells expressing rIP3R1ND (Fig. 5E). The pooled data demonstrates that in terms of the magnitude of signal (Fig. 5F), and the number of cells that respond above a defined threshold (Fig. 5G), rIP3R1ND is refractory to stimulation in DT40-3KO cells.
Given, that the rIP3R1RW mutation exhibited more activity when stably expressed in HEK-3KO cells when compared with DT40-3KOs, we next investigated whether rIP3R1ND was similarly more active when stably expressed in this mammalian cell line (Fig. 6A). Expression of rIP3R1ND in HEK-3KO cells did not result in alteration of the apparent content of the ER Ca2+ store (Fig. S2C) and resulted in an obvious reticular pattern of expression (Fig. S4). Consistent, with data obtained in cells expressing rIP3R1RW, stimulation of single HEK-3KO stably expressing rIP3R1ND with various [trypsin] resulted in modest Ca2+ release (Fig. 6B) manifested as reduced maximal initial release (Fig. 6C) and a lower percentage of cells responding at all [trypsin] (Fig. 6D). Once again, differences observed when the mutant is stably expressed in HEK-3KO cells compared with DT40-3KOs may result from differences in mutant protein expression in the two cell lines (Fig. S3B). A mutation in a neighboring residue in the same region, Thr-594 (T594I) (Fig. S5A), which is associated with infantile onset SCA (49), exhibited a similar decreased Ca2+ release (Fig. S5D) and percentage of cells responding (Fig. S5E) when stably expressed in HEK-3KO cells (Fig. S5B) and stimulated with a maximal concentration of trypsin (Fig. S5C). Reflecting the behavior at the single cell level, in population assays, the concentration versus response relationship for trypsin-induced [Ca2+] release reflected a decrease in both efficacy and sensitivity of stimulation (Fig. 6E and Fig. S5F). Thus, although competent to bind IP3, homomeric IP3R1 harboring mutations in this region are largely nonfunctional.
We next investigated the impact of the presence of variable numbers of mutant rIP3R1ND within a heterotetrametric assembly of subunits, which likely represents the situation in patients heterozygous for the mutation. Once again, the IP3R1 protein was readily detected in samples immunoprecipitated with a mCherry antibody suggesting heterotetrameric channels containing both WT and mutant channels are formed (Fig. S1A). Concatenated dimers were generated containing one WT rIP3R1 and one rIP3R1ND (rIP3R1-rIP3R1ND: R1/R1ND or rIP3R1ND-rIP3R1: R1/R1ND) and stably expressed in DT40-3KO cells (Fig. 7A). As shown previously, tetramers expressed from either monomeric rIP3R1 (WT rIP3R1) or dimeric rIP3R1 (R1/R1) were capable of supporting robust PAR2-mediated Ca2+ signaling, however, tetramers assembled from R1ND/R1 or R1/R1ND were completely refractory to stimulation at maximal [trypsin] (Fig. 7, B–D). These data indicate that tetramers harboring two mutant IP3R1ND subunits are likely severely compromised even given the modest activity of this mutant in HEK cells. To investigate if tetramers harboring only a single mutant rIP3R1ND retain any activity, concatenated tetramers were generated with three WT rIP3R1 and one rIP3R1ND (rIP3R1-rIP3R1-rIP3R1-rIP3R1ND: R1/R1/R1/R1ND) and stably expressed in DT40-3KO cells (Fig. 7E). Notably, some cells expressing R1/R1/R1/R1ND tetramers retained a comparable ability to support Ca2+ release to a concatenated tetramer assembled from WT rIP3R1 in terms of both trypsin (Fig. 7, F–H) and IP3-evoked Ca2+ release (Fig. 7I). In total these data demonstrate that rIP3RND retains modest activity and that tetramers harboring this gating mutant are not necessarily completely disabled.
Mutation of G2498S in the channel pore attenuates IP3R2 activity
We next investigated a missense mutation (G2498S) in IP3R2, identified in a family with generalized anhidrosis (55). IP3R2 is the predominant IP3R isoform expressed in eccrine sweat glands. Homozygous individuals were reported to have morphologically normal eccrine sweat glands but suffer from severe hyperthermia resulting from an absence of perspiration in response to physiological or pharmacological stimuli (Table S1). Notably, however, heterozygous siblings did not exhibit any symptoms of disease and thus this mutation represents a rare example of autosomal recessive inheritance of disease symptoms associated with IP3R mutations. The Gly-2498 residue in IP3R2 is located in a motif (GGGXG) conserved in all IP3R subtypes (Fig. 8A) as well as the superfamily of cation channels and forms the selectivity filter of the ion conducting pathway (Fig. 8B). This ion permeability pathway is a hot spot of other disease-associated mutations in multiple IP3R types (Table S4). This mutation (G2498S; designated mIP3R2GS) substitutes an evolutionarily conserved, nonpolar, hydrophobic, glycine residue, with a polar, hydrophillic serine residue and thus might reasonably be expected to disrupt the pore of the channel. Consistent with previous studies, when stably expressed in DT40-3KO cells (Fig. 8C), IP3 binding to mIP3R2GS was comparable with WT mIP3R2 (Fig. 8D). Nevertheless, trypsin stimulation failed to elicit any Ca2+ release in these cells (Fig. 8, E–G). Again, stable expression of mIP3R2GS in HEK-3KO cells did not alter the apparent store content of the ER (Fig. S2D) and resulted in an obvious reticular pattern of expression (Fig. S4B). However, notably, in contrast to the previous mutations investigated, when expressed in HEK-3KO cells (Fig. 9A) mIP3R2GS was unable to support any measurable Ca2+ release (Fig. 9, B–E). Similar results were seen in other clonal lines expressing the G2498S mutations (data not shown), as well as HEK-3KO cells stably expressing a mutation in hIP3R1 G2506R (Fig. S6, A–F), corresponding to the neighboring glycine in the GGGXG motif. This mutation is associated with spinocerebellar ataxia and Gillespie syndrome. Again, differences in expression in HEK-3KO cells versus DT40-3KO cells may in part explain these results (Fig. S3C). Overall, these data confirm that homozygous expression of mIP3R2GS results in a complete abrogation of IP3R activity and is consistent with the severe phenotype in individuals homozygous for the mutation.
As the heterozygous siblings of the affected individuals were seemingly not markedly impacted by harboring the mutation, we next investigated the behavior of tetrameric channels assembled from WT and mIP3R2GS. Immunoprecipitation once again confirmed that mutated mIP3R2GS monomers could interact with other monomeric subunits (Fig. S1, B and C). The current structural model based on the cryo-EM structure of IP3R1 (PDB codes 6MU1 and 6MU2) indicates that the Ser substitution for Gly-2547 (Gly-2498 in IP3R2) may form H bonds with the neighboring Arg-2544 residue (Fig. 10A), potentially resulting in constrained movement and restricted opening of the pore in a homotetramer harboring this mutant. Concatenated cDNA encoding either homodimeric mIP3R2-mIP3R2 (R2/R2) and mIP3R2GS-mIP3R2GS (R2GS/R2GS) or heterodimeric IP3R2GS-IP3R2 (R2GS/R2 and R2/R2GS) were generated and stably expressed in DT40-3KO cells (Fig. 10B). As shown previously (38), trypsin stimulation of cells expressing both mIP3R2 and R2/R2 elicited vigorous Ca2+ release (Fig. 10C). Although cells expressing R2GS/R2GS were refractory to stimulation, surprisingly, stimulation of cells expressing R2GS/R2 or R2/R2GS exhibited robust Ca2+ release (Fig. 10C and pooled data in Fig. 10, D and F). This partial channel functionality was confirmed through stimulation of additional cell lines expressing R2GS/R2 or R2/R2GS (data not shown). Similarly, exposure to IP3 in permeabilized cells also evoked robust Ca2+ release in WT homotetramers, intermediate Ca2+ release in heterotetramers, and no Ca2+ release in mutant homotetramers (Fig. 10F). These studies therefore reveal that heterotetrameric channels composed of two WT and two G2498S mutant subunits retain significant function as IP3-sensitive channels. Our data suggest that substitution of two or fewer Gly-Ser results in less restriction of pore movement required for gating, such that significant Ca2+ flux can occur that likely supports activity sufficient to render heterozygous individuals asymptomatic.
Discussion
Mutations have been identified throughout human IP3R genes and are associated with numerous diseases (58, 59). How these mutations affect channel function and contribute to the pathogenesis of the disease remain, in most cases, largely unexplored. Here we showed that three different mutations in three distinct domains of the IP3R result in decreased channel function. The extent and penetrance of disease varies with these mutations, and our data suggest that this depends on the mechanism of channel disruption and upon the pattern of inheritance. Both factors combine to alter the stoichiometry of functional subunits in the IP3R tetramer and ultimately the extent of Ca2+ release supported by the assembled channel.
Expression of each mutation in DT40-3KO cells resulted in nonfunctional IP3R Ca2+ channels, consistent with previous studies (22, 87). In each case, the complete loss in function was not due to a difference in protein localization and expression or ER store Ca2+ content. These data indicate that it is unlikely that these mutations result in mislocalization, altered degradation, or a constitutively leaky channel pore. Instead, the altered functionality occurred by mutation specific mechanisms. Patients harboring a mutation of Arg-269 to both a larger tryptophan and smaller glycine present with early-onset, nonprogressive SCA. Our data show that the R269W mutation resulted in a severe disruption of IP3 binding. Mechanistically this is consistent with the substitution of the positively charged arginine residue with the neutral tryptophan residue impeding the coordination of negatively charged phosphate moieties of IP3 in the binding pocket. In addition, it is likely that the bulky rings of the tryptophan residue may also sterically hinder IP3 binding to other critical residues in the binding pocket core. Both factors likely severely compromise the conformational change in the binding pocket necessary for communication of the gating information required for channel opening.
The N602D mutant, in contrast to the R269W mutant, did not alter IP3 binding. These results correlate with data showing that this region of the receptor is outside of the IP3-binding core (22). This mutation, together with the T594I mutation is located between two α-helices in the ARM1 domain, adjacent to the LBD. Although it is possible that these mutations result in the alteration of IP3R activity by channel modulators, it is more likely that these mutations result in an alteration of the native IP3R conformation, which abrogates the communication of information between the binding core and the adjacent ARM1 domain. Thus, these mutations are likely best termed “gating mutants.”
Binding of IP3 was also not altered by mutation of G2498S. This residue is integral to the selectivity filter in the conduction pathway of the channel pore. It is likely that the mutation from a small glycine residue to a larger serine residue results in a “dead pore.” Consistent with this observation, mutation of a neighboring selectivity filter residue in IP3R1, Gly-2506, to both arginine (Fig. S6), and alanine also produces a similarly nonfunctional, pore mutant (26).
In contrast to expression of the mutant constructs in DT40-3KO, stable expression the R269W and N602D mutations in HEK-3KO cells resulted in channels with varying degrees of limited functionality in terms of the ability of Gαq-coupled agonists to elicit Ca2+ release in a subset of cells. Increased channel activity of these mutants observed in the HEK-3KO cells could be due in part to increased IP3R expression in HEK-3KO cells compared with DT40-3KO cells (Fig. S3). Conceivably, this higher expression level could facilitate receptor clustering, which would be expected to amplify small, unitary Ca2+ signals. Additionally, it is possible that HEK cells express so-called “licensing” factors, which are reported to render refractory IP3R, activatable, thus revealing a degree of functionality to mutant IP3R modulated in this fashion (63). Other factors could also contribute to increased functionality in HEK-3KOs including differences in accessory proteins, regulation, and the intracellular environment present in avian versus mammalian cells. In contrast to R269W and N602D rIP3R1, stable expression of G2498S mIP3R2 or G2506R hIP3R1 in HEK-3KO cells failed to yield functional channels, reflecting the severe disruption of the channel pore.
Experiments documenting the activity in both DT40-3KO and HEK-3KO of the mutants stably expressed as homotetramers provides some insight into the mechanisms underlying the altered activity. However, both the R269W and N602D are inherited in an autosomal dominant manner, as are the vast majority of documented diseases associated with IP3R mutations (Tables S2–S4). Given that IP3Rs are well documented to form heterotetramers and this is the dominant IP3R form when multiple species are expressed (64), it would be predicted that these patients express a variety of tetramers comprised of different ratios of WT IP3R1 and mutant monomers. Indeed, our data confirms that heterotetramers of WT and WT channels are readily assembled. How incorporation of mutant monomers with altered activity influences the function of the assembled channel is therefore an important question. We address this issue using concatenated IP3R constructs as a powerful tool to precisely control the stoichiometry of WT versus mutant subunits in the tetramer.
For example, heterotetramers consisting of two WT subunits and two R269W subunits are predominantly nonfunctional. Our previous studies have indicated that because of the requirement of each monomer to be liganded by IP3, binding deficient mutants effectively act as dominant-negative modulators if incorporated into the heterotetramer (57). These results are also consistent with previous work from our laboratory showing other IP3R1 mutations inherited in an autosomal dominant manner that result in Gillespie syndrome are dominant-negative (65). Moreover, these results suggest that in patients with heterozygous expression of these mutant proteins, only WT homotetramers will be able to contribute a significant amount of Ca2+ release. An extension of this idea, if it is assumed that tetramers form without bias, is that given a normal distribution of mutant and WT subunits in the tetramer, only a small percentage (∼9%) would be predicted to be homotetramers of WT subunits, likely severely attenuating Ca2+ release of the compliment of IP3R1 expressed.
Tetramers assembled from dimers consisting of a WT and a N602D mutant, thus expressing two mutant and two WT sub-units, respectively, were similarly largely nonfunctional. Notably, however, expression of a concatenated tetramer with three WT subunits and one mutated N602D subunit resulted in a significant increase in channel function compared with cells expressing equal numbers of mutant and WT subtypes, indicating that the N602D subunit does not function as a true “dominant-negative.” Although the channel activity of this mutant tetramer is still significantly decreased compared with tetramers with four WT IP3R1 subunits (WT homotetramers), these data would predict that heterozygous individuals would express a greater percentage of IP3R tetramers, which retain function when compared with individuals expressing the R269W mutation.
In contrast to the R269W and N602D mutants, anhidrosis caused by the G2498S mutation is one of the few documented IP3R disorders that is inherited in an autosomal recessive manner (55). Although individuals homozygous for the mutation exhibit a profound heat intolerance through the inability to sweat, those heterozygous for the mutation are asymptomatic. Symptomatic individuals are therefore homozygous for the expression of the mutant proteins, whereas heterozygous individuals would be predicted to express a combination of mutant and WT subunits. Our data using concatenated dimers of mutant and WT subunits provides insight into why heterozygous individuals are asymptomatic. We demonstrate that heterotetramers comprised of two WT IP3R2 subunits and two G2498S mutant subunits retain a significant amount of function, such that Gαq-coupled stimulation results in Ca2+ release activity intermediate between that of the WT homotetramers and the G2498S mutant homotetramers. Thus, these data indicate that in individuals heterozygous for the mutation, potentially a larger fraction of IP3R2 channels are functional and thus presumably provide a sufficient Ca2+ signal to sustain sweat secretion. Notably, numerous cell types in the periphery, including hepatocytes and exocrine cells, express IP3R2, usually in combination with other isoforms. These data likely also explain the highly specific phenotype observed with this mutation; whereas human sweat glands almost exclusively express IP3R2, there appeared to be no obvious consequences of expression of the mutant in other tissues likely because the G2498S mutation does not act in a dominant-negative fashion. Thus, heterotetramers of mutant and other WT IP3R subtypes would not be predicted to be substantially disabled and thus impact cellular function. In total, studies presented here using concatenated subunits provide unprecedented insight into how mutant IP3R may function in heterozygous individuals and help demystify these challenging channelopathies.
Experimental procedures
Reagents
Antibodies generated by Pocono Rabbit Farms and Laboratories included rabbit polyclonal antibodies for IP3R2 against amino acids 320-338 (NT2) and amino acids 2686-2701 in mouse IP3R2 (CT2), as well as a rabbit polyclonal antibody against the C-terminal 19 aa of IP3R1 (CT1) (66). Reagents used for SDS-PAGE were obtained from Bio-Rad. Protein A/G-agarose beads were from Santa Cruz. Alexa Fluor 488 and Dylight™ 800CW secondary antibodies were from Thermo Scientific. Restriction enzymes and DNA T4 ligase were purchased from New England Biolabs. β-Mercaptoethanol, chicken serum, penicillin/streptomycin, RPMI 1640 media, and G418 sulfate were obtained from Life Technologies. Fetal bovine serum and goat serum were from Gibco. Fura2-AM was from TEFLABS. All other chemicals were obtained from Sigma unless otherwise indicated.
Plasmid construction
Mutagenesis and all DNA modifications were carried out using Pfu Ultra II Hotstart 2× Master Mix (Agilent). Primers used in this study were synthesized by Integrated DNA Technologies (IDT) (Table S5). A QuikChange mutagenesis protocol was used to introduce R269W in cDNAs encoding the rat IP3R1 (rIP3R1; NM_001270596) in pDNA3.1 using mutagenic primers (primers 1 and 2). These primers encode R269W, in addition to silently introducing the indicated restriction site (Table S5). Similarly, two mutagenesis primers were used to introduce N602D into rIP3R1 (primers 5 and 6), whereas mutagenesis primers were used to introduce R269W (primers 3 and 4), T494I (primers 7 and 8), and G2506R (primers 9 and 10 for G > A and primers 11 and 12 for G > C) into human IP3R1 (hIP3R1; NM_001099952) in pCDNA3.1 (obtained from Annetta Wronska at Columbia University). Finally, primers 13 and 14 introduced G2498S in mouse IP3R2 (mIP3R2; NM_019923) in pcDNA3.1. The coding regions for all constructs were confirmed by sequencing.
IP3R1 and IP3R2 concatenated constructs were generated as previously described in detail (38, 64, 67). Briefly, the cDNAs of the corresponding IP3R subunits, designated as head and tail, in pcDNA3.1 vector was modified and subcloned in a tandem fashion, between the two arms of the pJAZZ mamm linear vector (Lucigen). The resultant construct encodes one ORF consisting of two IP3R subunits connected with a 14-amino acid linker. Similarly, tetrameric IP3R1 constructs were generated using four modified IP3R1 subunits (numbered I-IV). The resulting cDNA forms one reading frame encoding four IP3R1 subunits with 14 amino acid linkers separating the subunits. To create concatenated constructs harboring one or more mutant subunits, the corresponding mutations were introduced in the individual subunits before assembling the concatenated constructs.
Cell culture and transfection
DT40-3KO, chicken B lymphocyte cells engineered through homologous recombination for the deletion of the three endogenous IP3R isoforms (56), were grown at 39 °C with 5% CO2 in RPMI 1640 media supplemented with 1% chicken serum, 10% fetal bovine serum, 100 units/ml of penicillin, 100 µg/ml of streptomycin. DT40-3KO transfection was performed as described before (66). In brief, 5 million cells were washed once with PBS and electroporated with 5-10 µg of DNA using Amaxa cell nucleofector kit T (Lonza laboratories). Cells were allowed to recover for 24 h at 39˚C with 5% CO2, and then passed into 96-well–plates containing media supplemented with 2 mg/ml of G418. Clones expressing the desired constructs were identified by immunoblots 10-14 days after transfection. HEK-3KO, HEK293 cells engineered in our laboratory though CRISPR/Cas9 for the deletion of the three endogenous IP3R isoforms were grown at 37 °C with 5% CO2 in Dulbecco's modified Eagle's medium supplemented with 10% fetal bovine serum, 100 units/ml of penicillin, 100 µg/ml of streptomycin.
HEK-3KO transfection was also performed similarly to as previously described (57). In brief, 1 million cells were washed once with PBS and electroporated with 5-10 µg of DNA using Amaxa cell nucleofector kit T (Lonza laboratories). Cells recovered for 48 h at 37 °C with 5% CO2, and then were passed into new 10-cm plates containing media supplemented with 2 mg/ml of G418. After 7 days of selection, immunoblot was performed to assess expression of the desired constructs in these stable HEK cell pools. Following confirmation of expression, cells were either picked and transferred to new 24-well– plates or diluted at various concentrations into 96-well–plates or containing media supplemented with 2 mg/ml of G418. Monoclonal lines were identified 7 days after dilution and clones expressing the desired constructs were identified by immunoblots 10-14 days after identification.
Cell lysis, co-immunoprecipitation, and SDS-PAGE analyses
Following treatment, cells were harvested by centrifugation, washed once with PBS, and solubilized in cell lysis buffer containing 10 mm Tris-HCl, 10 mm NaCl, 1 mm EGTA, 1 mm EDTA, 1 mm NaF, 20 mm Na4P2O7, 2 mm Na3 VO4, 1% Triton X-100 (v/v), 0.5% sodium deoxycholate (w/v), and 10% glycerol supplemented with protease inhibitors. Lysates were incubated for 30 min on ice and cleared by centrifugation at 16,000 × g for 10 min at 4 °C. Protein concentrations in cleared lysates were determined using Dc protein assay kit (Bio-Rad). Cleared supernatants were rocked with the indicated antibodies and 30 μl of protein A/G slurry overnight. Immunocomplexes were washed five times and resuspended in gel loading buffer. Proteins were fractionated on SDS-PAGE gel and transferred to a nitrocellulose membrane, which was probed with the indicated primary antibodies and corresponding secondary antibodies. Membranes were imaged with an Odyssey IR imaging system (LICOR Biosciences).
Immunocytochemistry and confocal microscopy
Stable HEK cell lines were plated on poly-d-lysine–coated coverslips. When roughly 50% confluent cells were fixed using ice-cold methanol at −20 °C or paraformaldehyde at room temperature for 10 min. Subsequently, coverslips were washed with PBS and cells were blocked in goat serum or BSA for 1.5 h. Following blocking, cells were incubated in primary antibody (IP3R1 or IP3R2) and DAPI overnight at 4 °C. The following day, primary antibody was removed, and coverslips were washed 3 times with PBS for 10 min with gentle rocking. Subsequently, the secondary antibody conjugated to Alexa Fluor 488 or 594 was incubated for 1 h at room temperature with gentle rocking. After incubation, coverslips were washed with PBS and mounted on slides. After allowing slides to dry, coverslips were sealed onto slides and imaged using confocal microscopy.
Measurement of cytosolic Ca2+ in intact cells
Single cell Ca2+ imaging was performed as described previously (66). DT40-3KO and HEK-3KO cells expressing IP3R constructs were washed once with imaging buffer (10 mm HEPES, 1.26 mm Ca2+, 137 mm NaCl, 4.7 mm KCl, 5.5 mm glucose, 1 mm Na2HPO4, 0.56 mm MgCl2, pH 7.4). Cells were then resuspended in imaging buffer containing 2 μm Fura2-AM and placed on a coverslip for 20 min at room temperature to allow for fluorescent dye loading and attachment of cells to the glass coverslip. Cells were then perfused with imaging buffer and stimulated with the desired agonist. Ca2+ imaging was performed using an inverted epifluorescence Nikon microscope with a ×40 oil immersion objective. Cells were alternately excited at 340 and 380 nm, and emission was monitored at 505 nm. Images were captured every second with an exposure of 10 ms and 4 × 4 binning using a digital camera driven by TILL Photonics software. Image acquisition was performed using TILLvisION software and data were exported to Microsoft excel. Experiments were repeated at least three times.
Population-based Ca2+ imaging was performed in intact HEK cells by loading cells with 4 μm Fura-2 AM in culture medium. After 1 h, cells were harvested and subsequently washed, resuspended in imaging buffer, and dispensed into a black-walled flat-bottom 96-well–plate (∼300,000 cells/well). The plate was spun at 200 × g for 2 min to plate cells to the bottom of each well. The plate rested for 30 min prior to commencing the assay. Fluorescence imaging was carried out using FlexStation 3 from Molecular Devices (excitation alternated between 340 and 380 nm and emission 510 nm) and analyzed by using SoftMax© Pro Microplate Data Acquisition and Analysis software to determine peak of the response. Data were averaged from at least 3 individual plates and curves fitting was done using a logistic dose-response equation in the OriginPro 6.1 software.
Single cell-permeabilized DT40 cell IP3-induced Ca2+ release (IICR) assays
DT40-3KO cells stably expressing defined IP3R constructs were loaded with 20 μm MgFura-2 AM for 50–60 min and subsequently permeabilized by superfusion of 40 μm β-escin in imaging buffer as previously described (57, 64). The duration of permeabilization depended on the flow rate, and care was taken to prevent excessive (internal membrane) permeabilization by careful monitoring of the fluorescence. Permeabilized cells were then washed in ICM without β-escin for 15 min to facilitate removal of cytosolic dye. The internal stores of permeabilized cells were loaded by activating SERCA through superfusion with ICM containing 1.4 mm MgCl2, 3 mm NaATP, and 0.65 mm CaCl2 (free [Ca2+] of 200 nm (MaxChelator freeware)). Upon stabilization of fluorescence, MgCl2 was removed from the superfused solution to disable SERCA. Varying concentrated IP3 were applied through superfusion to induce unidirectional Ca2+ release from internals stores. After washing out IP3, stores were repeatedly refilled and released to allow for repeated stimulations. Each experiment contained between 30 and 60 cells in a field of view and performed a minimum of three times. The initial rates of Ca2+ release were determined by fitting the release curves of each individual cell to a single exponential function (OriginPro 6.1). Concentration-response curves were plotted using the determined rates. All imaging was performed on an inverted epifluorescence Nikon microscope with a ×40 oil immersion objective. Cells were alternately excited at 340 and 380 nm, and emission was monitored at 505 nm. Images were captured every 5 s during permeabilization, loading, and disabling and every second during release. This was done with an exposure of 10 ms and 4 × 4 binning using a digital camera driven by TILL Photonics software.
Competitive IP3-binding assay
DT40-3KO cells stably expressing the indicated IP3R constructs were washed once with PBS and lysed in Igepal lysis buffer containing (120 mm NaCl, 50 mm Tris-HCl, 0.5% Igepal (v/v), 1 mm EDTA) supplemented with a mixture of protease inhibitors. Lysates were incubated on ice for 30 min and cleared by centrifugation at 16,000 × g for 10 min at 4 °C. IP3R proteins were immunoprecipitated overnight using anti-IP3R and protein A/G PLUS-agarose beads. Immunocomplexes were washed three times with lysis buffer and twice with binding buffer (50 mm Tris-base, 1 mm EDTA, pH 8, plus 1 mm β-mercaptoethanol) and equally distributed into 1.5-ml microcentrifuge tubes. One pair of tubes was immunoblotted using SDS-PAGE to ensure equal immunoprecipitation of WT and mutant protein, whereas all others underwent the binding reaction consisting of a 100-μl volume containing equal amounts of immunoprecipitated proteins, 2.5 nm tritiated IP3 ([3H]IP3) in the presence of different concentrations of unlabeled IP3. Tubes were incubated for 1 h at 4 °C with mixing every 10 min. Finally, tubes were centrifuged at 16,000 × g, supernatants were removed, and 500 µl of 1% of SDS was added to each tube. The contents of the tubes were transferred to vials containing scintillation liquid after 12 h and bound radioactivity were measured using liquid scintillation counter. Nonspecific binding was calculated as the amount of bound radioactivity in the presence of 50 μm unlabeled IP3. Specific binding is determined by subtracting nonspecific binding from the CPM values obtained in other conditions. Total specific binding was determined as the binding observed in the absence of unlabeled IP3. Normalized specific binding from 3 experiments were averaged and curves were fit using a logistic dose-response equation using the OriginPro 6.1 software.
Data analysis and statistical analysis
Data obtained were exported into Microsoft Excel where appropriate ratios, normalizations, and averages were calculated. Statistical analysis was performed in GraphPad Prism for experiments comparing two cell lines (unpaired t test) and three or more cell lines (one-way ANOVA with Tukey's test). Logistic fits, as well as IC50 were calculated using OriginPro 6.1.
Data availability
All constructs used are available freely on request. All data needed to evaluate the conclusions are presented in the paper or Supporting material.
Acknowledgments
The thank Dr. Irina Serysheva and Dr. Mariah Baker (University of Texas, Health Science Center at Houston) for helpful insights into the structural consequences of these missense mutations. We also thank members of the lab, especially Sydney Stremick for helpful and constructive comments.
NIH NIDCR (DE014756) to David I. Yule
Edited by Roger J. Colbran
Footnotes
This article contains supporting information.
Author contributions—L. E. T. formal analysis; L. E. T., K. J. A., A. M. W., and S. M. investigation; L. E. T. writing-original draft; K. J. A. and D. I. Y. conceptualization; K. J. A. and A. M. W. methodology; SM and D. I. Y. data curation; D. I. Y. funding acquisition; D. I. Y. project administration; D. I. Y. writing-review and editing.
Funding and additional information—This work was supported by National Institutes of Health Grant R01 DE014756 (to D. I. Y.). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.
Conflict of interest—The authors declare no conflicts of interest in regards to the content of this manuscript.
Present address for Kamil J. Alzayady: Physician Assistant Department, University of Maryland Eastern Shore, Princess Anne, Maryland, USA
- IP3
- inositol 1,4,5-trisphosphate
- IP3R
- inositol 1,4,5-trisphosphate receptor
- LBD
- ligand-binding domain
- SCA
- spinocerebellar ataxia
- aa
- amino acid(s)
- GS
- Gillespie syndrome
- PAR
- protease-activated receptor
- HEK
- human embryonic kidney
- SERCA
- sarcoplasmic reticulum Ca2+-ATPase
- DAPI
- 4′,6-diamidino-2-phenylindole
- PDB
- Protein Data Bank
- ANOVA
- analysis of variance
- GAPDH
- glyceraldehyde-3-phosphate dehydrogenase.
The nucleotide sequences, and thus amino acid numbering, used throughout this paper are: rIP3R1, NM_001007235; hIP3R1, NM_001099952; mIP3R2, NM_019923. See Table S6 for alignment guide of these sequences with other commonly used sequences.
Supplementary Material
References
- 1.Chin K.V., Cade C., Brostrom C.O., Galuska E.M., Brostrom M.A. Calcium-dependent regulation of protein synthesis at translational initiation in eukaryotic cells. J. Biol. Chem. 1987;262:16509–16514. 3680263. [PubMed] [Google Scholar]
- 2.Nicotera P., Orrenius S. Ca2+ and cell death. Ann. N. Y. Acad. Sci. 1992;648:17–27. doi: 10.1111/j.1749-6632.1992.tb24520.x. 1637044. [DOI] [PubMed] [Google Scholar]
- 3.Berridge M.J. Inositol trisphosphate and calcium signalling. Nature. 1993;361:315–325. doi: 10.1038/361315a0. 8381210. [DOI] [PubMed] [Google Scholar]
- 4.Berridge M.J., Bootman M.D., Lipp P. Calcium: a life and death signal. Nature. 1998;395:645–648. doi: 10.1038/27094. 9790183. [DOI] [PubMed] [Google Scholar]
- 5.Berridge M.J., Lipp P., Bootman M.D. The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 2000;1:11–21. doi: 10.1038/35036035. 11413485. [DOI] [PubMed] [Google Scholar]
- 6.Mattson M.P., Chan S.L. Calcium orchestrates apoptosis. Nat. Cell Biol. 2003;5:1041–1043. doi: 10.1038/ncb1203-1041. 14647298. [DOI] [PubMed] [Google Scholar]
- 7.Limbäck-Stokin K., Korzus E., Nagaoka-Yasuda R., Mayford M. Nuclear calcium/calmodulin regulates memory consolidation. J. Neurosci. 2004;24:10858–10867. doi: 10.1523/JNEUROSCI.1022-04.2004. 15574736. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Harr M.W., Distelhorst C.W. Apoptosis and autophagy: decoding calcium signals that mediate life or death. Cold Spring Harbor Perspect. Biol. 2010;2 doi: 10.1101/cshperspect.a005579. 20826549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Iino M. Spatiotemporal dynamics of Ca2+ signaling and its physiological roles. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2010;86:244–256. doi: 10.2183/pjab.86.244. 20228624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Decuypere J.P., Monaco G., Bultynck G., Missiaen L., De Smedt H., Parys J.B. The IP(3) receptor-mitochondria connection in apoptosis and autophagy. Biochim. Biophys. Acta. 2011;1813:1003–1013. doi: 10.4061/2011/920178. 21146562. [DOI] [PubMed] [Google Scholar]
- 11.Goonasekera S.A., Molkentin J.D. Unraveling the secrets of a double life: contractile versus signaling Ca2+ in a cardiac myocyte. J. Mol. Cell. Cardiol. 2012;52:317–322. doi: 10.1016/j.yjmcc.2011.05.001. 21600216. [DOI] [PubMed] [Google Scholar]
- 12.Mignery G.A., Südhof T.C., Takei K., De Camilli P. Putative receptor for inositol 1,4,5-trisphosphate similar to ryanodine receptor. Nature. 1989;342:192–195. doi: 10.1038/342192a0. 2554146. [DOI] [PubMed] [Google Scholar]
- 13.Joseph S.K., Lin C., Pierson S., Thomas A.P., Maranto A.R. Heteroligomers of type-I and type-III inositol trisphosphate receptors in WB rat liver epithelial cells. J. Biol. Chem. 1995;270:23310–23316. doi: 10.1074/jbc.270.40.23310. 7559486. [DOI] [PubMed] [Google Scholar]
- 14.Joseph S.K., Boehning D., Pierson S., Nicchitta C.V. Membrane insertion, glycosylation, and oligomerization of inositol trisphosphate receptors in a cell-free translation system. J. Biol. Chem. 1997;272:1579–1588. doi: 10.1074/jbc.272.3.1579. 8999831. [DOI] [PubMed] [Google Scholar]
- 15.Patel S., Joseph S.K., Thomas A.P. Molecular properties of inositol 1,4,5-trisphosphate receptors. Cell Calcium. 1999;25:247–264. doi: 10.1054/ceca.1999.0021. 10378086. [DOI] [PubMed] [Google Scholar]
- 16.Onoue H., Tanaka H., Tanaka K., Doira N., Ito Y. Heterooligomer of type 1 and type 2 inositol 1,4,5-trisphosphate receptor expressed in rat liver membrane fraction exists as tetrameric complex. Biochem. Biophys. Res. Commun. 2000;267:928–933. doi: 10.1006/bbrc.1999.2065. 10673393. [DOI] [PubMed] [Google Scholar]
- 17.Joseph S.K., Bokkala S., Boehning D., Zeigler S. Factors determining the composition of inositol trisphosphate receptor hetero-oligomers expressed in COS cells. J. Biol. Chem. 2000;275:16084–16090. doi: 10.1074/jbc.M000506200. 10747920. [DOI] [PubMed] [Google Scholar]
- 18.Foskett J.K., White C., Cheung K.H., Mak D.O. Inositol trisphosphate receptor Ca2+ release channels. Physiol. Rev. 2007;87:593–658. doi: 10.1152/physrev.00035.2006. 17429043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Streb H., Irvine R.F., Berridge M.J., Schulz I. Release of Ca2+ from a nonmitochondrial intracellular store in pancreatic acinar cells by inositol-1,4,5-trisphosphate. Nature. 1983;306:67–69. doi: 10.1038/306067a0. 6605482. [DOI] [PubMed] [Google Scholar]
- 20.Streb H., Bayerdörffer E., Haase W., Irvine R.F., Schulz I. Effect of inositol-1,4,5-trisphosphate on isolated subcellular fractions of rat pancreas. J. Membr. Biol. 1984;81:241–253. doi: 10.1007/BF01868717. 6334162. [DOI] [PubMed] [Google Scholar]
- 21.Iwai M., Michikawa T., Bosanac I., Ikura M., Mikoshiba K. Molecular basis of the isoform-specific ligand-binding affinity of inositol 1,4,5-trisphosphate receptors. J. Biol. Chem. 2007;282:12755–12764. doi: 10.1074/jbc.M609833200. 17327232. [DOI] [PubMed] [Google Scholar]
- 22.Yoshikawa F., Morita M., Monkawa T., Michikawa T., Furuichi T., Mikoshiba K. Mutational analysis of the ligand binding site of the inositol 1,4,5-trisphosphate receptor. J. Biol. Chem. 1996;271:18277–18284. doi: 10.1074/jbc.271.30.18277. 8663526. [DOI] [PubMed] [Google Scholar]
- 23.Yule D.I., Betzenhauser M.J., Joseph S.K. Linking structure to function: Recent lessons from inositol 1,4,5-trisphosphate receptor mutagenesis. Cell Calcium. 2010;47:469–479. doi: 10.1016/j.ceca.2010.04.005. 20510450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Ramos-Franco J., Galvan D., Mignery G.A., Fill M. Location of the permeation pathway in the recombinant type 1 inositol 1,4,5-trisphosphate receptor. J. Gen. Physiol. 1999;114:243–250. doi: 10.1085/jgp.114.2.243. 10436000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Boehning D., Mak D.O., Foskett J.K., Joseph S.K. Molecular determinants of ion permeation and selectivity in inositol 1,4,5-trisphosphate receptor Ca2+ channels. J. Biol. Chem. 2001;276:13509–13512. doi: 10.1074/jbc.C100094200. 11278266. [DOI] [PubMed] [Google Scholar]
- 26.Schug Z.T., da Fonseca P.C.A., Bhanumathy C.D., Wagner L., 2nd, Zhang X., Bailey B., Morris E.P., Yule D.I., Joseph S.K. Molecular characterization of the inositol 1,4,5-trisphosphate receptor pore-forming segment. J. Biol. Chem. 2008;283:2939–2948. doi: 10.1074/jbc.M706645200. 18025085. [DOI] [PubMed] [Google Scholar]
- 27.Fan G., Baker M.L., Wang Z., Baker M.R., Sinyagovskiy P.A., Chiu W., Ludtke S.J., Serysheva I.I. Gating machinery of InsP3R channels revealed by electron cryomicroscopy. Nature. 2015;527:336–341. doi: 10.1038/nature15249. 26458101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Baker M.R., Fan G., Serysheva I.I. Structure of IP3R channel: high-resolution insights from cryo-EM. Curr. Opin. Struct. Biol. 2017;46:38–47. doi: 10.1016/j.sbi.2017.05.014. 28618351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Hamada K., Miyatake H., Terauchi A., Mikoshiba K. IP3-mediated gating mechanism of the IP3 receptor revealed by mutagenesis and X-ray crystallography. Proc. Natl. Acad. Sci. U.S.A. 2017;114:4661–4666. doi: 10.1073/pnas.1701420114. 28416699. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Fan G., Baker M.R., Wang Z., Seryshev A.B., Ludtke S.J., Baker M.L., Serysheva I.I. Cryo-EM reveals ligand induced allostery underlying InsP3R channel gating. Cell Res. 2018;28:1158–1170. doi: 10.1038/s41422-018-0108-5. 30470765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Taylor C.W., Tovey S.C. IP(3) receptors: toward understanding their activation. Cold Spring Harbor Perspect. Biol. 2010;2 doi: 10.1101/cshperspect.a004010. 20980441. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Maranto A.R. Primary structure, ligand binding, and localization of the human type 3 inositol 1,4,5-trisphosphate receptor expressed in intestinal epithelium. J. Biol. Chem. 1994;269:1222–1230. 8288584. [PubMed] [Google Scholar]
- 33.Newton C.L., Mignery G.A., Sudhof T.C. Co-expression in vertebrate tissues and cell lines of multiple inositol 1,4,5-trisphosphate (InsP3) receptors with distinct affinities for InsP3. J. Biol. Chem. 1994;269:28613–28619. 7961809. [PubMed] [Google Scholar]
- 34.De Smedt H., Missiaen L., Parys J.B., Henning R.H., Sienaert I., Vanlingen S., Gijsens A., Himpens B., Casteels R. Isoform diversity of the inositol trisphosphate receptor in cell types of mouse origin. Biochem. J. 1997;322:575–583. doi: 10.1042/bj3220575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Taylor C.W., Genazzani A.A., Morris S.A. Expression of inositol trisphosphate receptors. Cell Calcium. 1999;26:237–251. doi: 10.1054/ceca.1999.0090. 10668562. [DOI] [PubMed] [Google Scholar]
- 36.Furuichi T., Yoshikawa S., Miyawaki A., Wada K., Maeda N., Mikoshiba K. Primary structure and functional expression of the inositol 1,4,5-trisphosphate-binding protein P400. Nature. 1989;342:32–38. doi: 10.1038/342032a0. 2554142. [DOI] [PubMed] [Google Scholar]
- 37.Wojcikiewicz R.J. Type I, II, and III inositol 1,4,5-trisphosphate receptors are unequally susceptible to down-regulation and are expressed in markedly different proportions in different cell types. J. Biol. Chem. 1995;270:11678–11683. doi: 10.1074/jbc.270.19.11678. 7744807. [DOI] [PubMed] [Google Scholar]
- 38.Alzayady K.J., Wagner L.E., 2nd, Chandrasekhar R., Monteagudo A., Godiska R., Tall G.G., Joseph S.K., Yule D.I. Functional inositol 1,4,5-trisphosphate receptors assembled from concatenated homo- and heteromeric subunits. J. Biol. Chem. 2013;288:29772–29784. doi: 10.1074/jbc.M113.502203. 23955339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Matsumoto M., Nakagawa T., Inoue T., Nagata E., Tanaka K., Takano H., Minowa O., Kuno J., Sakakibara S., Yamada M., Yoneshima H., Miyawaki A., Fukuuchi Y., Furuichi T., Okano H. Ataxia and epileptic seizures in mice lacking type 1 inositol 1,4,5-trisphosphate receptor. Nature. 1996;379:168–171. doi: 10.1038/379168a0. 8538767. [DOI] [PubMed] [Google Scholar]
- 40.Street V.A., Bosma M.M., Demas V.P., Regan M.R., Lin D.D., Robinson L.C., Agnew W.S., Tempel B.L. The type 1 inositol 1,4,5-trisphosphate receptor gene is altered in the opisthotonos mouse. J. Neurosci. 1997;17:635–645. doi: 10.1523/JNEUROSCI.17-02-00635.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.van de Leemput J., Chandran J., Knight M.A., Holtzclaw L.A., Scholz S., Cookson M.R., Houlden H., Gwinn-Hardy K., Fung H.C., Lin X., Hernandez D., Simon-Sanchez J., Wood N.W., Giunti P., Rafferty I. Deletion at ITPR1 underlies ataxia in mice and spinocerebellar ataxia 15 in humans. PLoS Genet. 2007;3:e108. doi: 10.1371/journal.pgen.0030108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Futatsugi A., Nakamura T., Yamada M.K., Ebisui E., Nakamura K., Uchida K., Kitaguchi T., Takahashi-Iwanaga H., Noda T., Aruga J., Mikoshiba K. IP3 receptor types 2 and 3 mediate exocrine secretion underlying energy metabolism. Science. 2005;309:2232–2234. doi: 10.1126/science.1114110. 16195467. [DOI] [PubMed] [Google Scholar]
- 43.Bezprozvanny I., Hayden M.R. Deranged neuronal calcium signaling and Huntington disease. Biochem. Biophys. Res. Commun. 2004;322:1310–1317. doi: 10.1016/j.bbrc.2004.08.035. 15336977. [DOI] [PubMed] [Google Scholar]
- 44.Foskett J.K. Inositol trisphosphate receptor Ca2+ release channels in neurological diseases. Pflugers Arch. 2010;460:481–494. doi: 10.1007/s00424-010-0826-0. 20383523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Bezprozvanny I. Role of inositol 1,4,5-trisphosphate receptors in pathogenesis of Huntington's disease and spinocerebellar ataxias. Neurochem. Res. 2011;36:1186–1197. doi: 10.1007/s11064-010-0393-y. 21210219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Hisatsune C., Mikoshiba K. IP3 receptor mutations and brain diseases in human and rodents. J. Neurochem. 2017;141:790–807. doi: 10.1111/jnc.13991. 28211945. [DOI] [PubMed] [Google Scholar]
- 47.Huang L., Chardon J.W., Carter M.T., Friend K.L., Dudding T.E., Schwartzentruber J., Zou R., Schofield P.W., Douglas S., Bulman D.E., Boycott K.M. Missense mutations in ITPR1 cause autosomal dominant congenital nonprogressive spinocerebellar ataxia. Orphanet. J. Rare Dis. 2012;7:67. doi: 10.1186/1750-1172-7-67. 22986007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Parolin Schnekenberg R., Perkins E.M., Miller J.W., Davies W.I., D'Adamo M.C., Pessia M., Fawcett K.A., Sims D., Gillard E., Hudspith K., Skehel P., Williams J., O'Regan M., Jayawant S., Jefferson R. De novo point mutations in patients diagnosed with ataxic cerebral palsy. Brain. 2015;138:1817–1832. doi: 10.1093/brain/awv117. 25981959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Sasaki M., Ohba C., Iai M., Hirabayashi S., Osaka H., Hiraide T., Saitsu H., Matsumoto N. Sporadic infantile-onset spinocerebellar ataxia caused by missense mutations of the inositol 1,4,5-triphosphate receptor type 1 gene. J. Neurol. 2015;262:1278–1284. doi: 10.1007/s00415-015-7705-8. 25794864. [DOI] [PubMed] [Google Scholar]
- 50.Barresi S., Niceta M., Alfieri P., Brankovic V., Piccini G., Bruselles A., Barone M.R., Cusmai R., Tartaglia M., Bertini E., Zanni G. Mutations in the IRBIT domain of ITPR1 are a frequent cause of autosomal dominant nonprogressive congenital ataxia. Clin. Genet. 2017;91:86–91. doi: 10.1111/cge.12783. 27062503. [DOI] [PubMed] [Google Scholar]
- 51.Das J., Lilleker J., Shereef H., Ealing J. Missense mutation in the ITPR1 gene presenting with ataxic cerebral palsy: Description of an affected family and literature review. Neurol. Neurochir. Pol. 2017;51:497–500. doi: 10.1016/j.pjnns.2017.06.012. 28826917. [DOI] [PubMed] [Google Scholar]
- 52.Zambonin J.L., Bellomo A., Ben-Pazi H., Everman D.B., Frazer L.M., Geraghty M.T., Harper A.D., Jones J.R., Kamien B., Kernohan K., Koenig M.K., Lines M., Palmer E.E., Richardson R., Segel R. Spinocerebellar ataxia type 29 due to mutations in ITPR1: a case series and review of this emerging congenital ataxia. Orphanet. J. Rare Dis. 2017;12:121. doi: 10.1186/s13023-017-0672-7. 28659154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Synofzik M., Helbig K.L., Harmuth F., Deconinck T., Tanpaiboon P., Sun B., Guo W., Wang R., Palmaer E., Tang S., Schaefer G.B., Gburek-Augustat J., Züchner S., Krägeloh-Mann I., Baets J. De novo ITPR1 variants are a recurrent cause of early-onset ataxia, acting via loss of channel function. Eur. J. Hum. Genet. 2018;26:1623–1634. doi: 10.1038/s41431-018-0206-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.McEntagart M., Williamson K.A., Rainger J.K., Wheeler A., Seawright A., De Baere E., Verdin H., Bergendahl L.T., Quigley A., Rainger J., Dixit A., Sarkar A., López Laso E., Sanchez-Carpintero R., Barrio J. A restricted repertoire of de novo mutations in ITPR1 cause Gillespie syndrome with evidence for dominant-negative effect. Am. J. Hum. Genet. 2016;98:981–992. doi: 10.1016/j.ajhg.2016.03.018. 27108798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Klar J., Hisatsune C., Baig S.M., Tariq M., Johansson A.C., Rasool M., Malik N.A., Ameur A., Sugiura K., Feuk L., Mikoshiba K., Dahl N. Abolished InsP3R2 function inhibits sweat secretion in both humans and mice. J. Clin. Invest. 2014;124:4773–4780. doi: 10.1172/JCI70720. 25329695. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Sugawara H., Kurosaki M., Takata M., Kurosaki T. Genetic evidence for involvement of type 1, type 2 and type 3 inositol 1,4,5-trisphosphate receptors in signal transduction through the B-cell antigen receptor. EMBO J. 1997;16:3078–3088. doi: 10.1093/emboj/16.11.3078. 9214625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Alzayady K.J., Wang L., Chandrasekhar R., Wagner L.E., 2nd, Van Petegem F., Yule D.I. Defining the stoichiometry of inositol 1,4,5-trisphosphate binding required to initiate Ca2+ release. Sci. Signal. 2016;9:ra35. doi: 10.1126/scisignal.aad6281. 27048566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Terry L.E., Alzayady K.J., Furati E., Yule D.I. Inositol 1,4,5-trisphosphate receptor mutations associated with human disease. Messenger. 2018;6:29–44. doi: 10.1166/msr.2018.1075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kerkhofs M., Seitaj B., Ivanova H., Monaco G., Bultynck G., Parys J.B. Pathophysiological consequences of isoform-specific IP3 receptor mutations. Biochim. Biophys. Acta Mol. Cell Res. 2018;1865:1707–1717. doi: 10.1016/j.bbamcr.2018.06.004. 29906486. [DOI] [PubMed] [Google Scholar]
- 60.Ando H., Mizutani A., Matsu-Ura T., Mikoshiba K. IRBIT, a novel inositol 1,4,5-trisphosphate (IP3) receptor-binding protein, is released from the IP3 receptor upon IP3 binding to the receptor. J. Biol. Chem. 2003;278:10602–10612. doi: 10.1074/jbc.M210119200. 12525476. [DOI] [PubMed] [Google Scholar]
- 61.Ando H., Mizutani A., Kiefer H., Tsuzurugi D., Michikawa T., Mikoshiba K. IRBIT suppresses IP3 receptor activity by competing with IP3 for the common binding site on the IP3 receptor. Mol. Cell. 2006;22:795–806. doi: 10.1016/j.molcel.2006.05.017. 16793548. [DOI] [PubMed] [Google Scholar]
- 62.Bosanac I., Alattia J.R., Mal T.K., Chan J., Talarico S., Tong F.K., Tong K.I., Yoshikawa F., Furuichi T., Iwai M., Michikawa T., Mikoshiba K., Ikura M. Structure of the inositol 1,4,5-trisphosphate receptor binding core in complex with its ligand. Nature. 2002;420:696–700. doi: 10.1038/nature01268. 12442173. [DOI] [PubMed] [Google Scholar]
- 63.Taylor C.W., Machaca K. IP3 receptors and store-operated Ca2+ entry: a license to fill. Curr. Opin. Cell Biol. 2019;57:1–7. doi: 10.1016/j.ceb.2018.10.001. 30368032. [DOI] [PubMed] [Google Scholar]
- 64.Chandrasekhar R., Alzayady K.J., Wagner L.E., 2nd, Yule D.I. Unique regulatory properties of heterotetrameric inositol 1,4,5-trisphosphate receptors revealed by studying concatenated receptor constructs. J. Biol. Chem. 2016;291:4846–4860. doi: 10.1074/jbc.M115.705301. 26755721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Gerber S., Alzayady K.J., Burglen L., Bremond-Gignac D., Marchesin V., Roche O., Rio M., Funalot B., Calmon R., Durr A., Gil-da-Silva-Lopes V.L., Ribeiro Bittar M.F., Orssaud C., Héron B., Ayoub E. Recessive and dominant de novo ITPR1 mutations cause Gillespie syndrome. Am. J. Hum. Genet. 2016;98:971–980. doi: 10.1016/j.ajhg.2016.03.004. 27108797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Alzayady K.J., Chandrasekhar R., Yule D.I. Fragmented inositol 1,4,5-trisphosphate receptors retain tetrameric architecture and form functional Ca2+ release channels. J. Biol. Chem. 2013;288:11122–11134. doi: 10.1074/jbc.M113.453241. 23479737. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Chandrasekhar R., Alzayady K.J., Yule D.I. Using concatenated subunits to investigate the functional consequences of heterotetrameric inositol 1,4,5-trisphosphate receptors. Biochem. Soc. Trans. 2015;43:364–370. doi: 10.1042/BST20140287. 26009177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Ganesamoorthy D., Bruno D.L., Schoumans J., Storey E., Delatycki M.B., Zhu D., Wei M.K., Nicholson G.A., McKinlay Gardner R.J., Slater H.R. Development of a multiplex ligation-dependent probe amplification assay for diagnosis and estimation of the frequency of spinocerebellar ataxia type 15. Clin. Chem. 2009;55:1415–1418. doi: 10.1373/clinchem.2009.124958. 19423733. [DOI] [PubMed] [Google Scholar]
- 69.Ohba C., Osaka H., Iai M., Yamashita S., Suzuki Y., Aida N., Shimozawa N., Takamura A., Doi H., Tomita-Katsumoto A., Nishiyama K., Tsurusaki Y., Nakashima M., Miyake N., Eto Y. Diagnostic utility of whole exome sequencing in patients showing cerebellar and/or vermis atrophy in childhood. Neurogenetics. 2013;14:225–232. doi: 10.1007/s10048-013-0375-8. 24091540. [DOI] [PubMed] [Google Scholar]
- 70.Fogel B.L., Lee H., Deignan J.L., Strom S.P., Kantarci S., Wang X., Quintero-Rivera F., Vilain E., Grody W.W., Perlman S., Geschwind D.H., Nelson S.F. Exome sequencing in the clinical diagnosis of sporadic or familial cerebellar ataxia. JAMA Neurol. 2014;71:1237–1246. doi: 10.1001/jamaneurol.2014.1944. 25133958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Wang L., Hao Y., Yu P., Cao Z., Zhang J., Zhang X., Chen Y., Zhang H., Gu W. Identification of a splicing mutation in ITPR1 via WES in a Chinese early-onset spinocerebellar ataxia family. Cerebellum. 2018;17:294–299. doi: 10.1007/s12311-017-0896-z. 29196976. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Dudding T.E., Friend K., Schofield P.W., Lee S., Wilkinson I.A., Richards R.I. Autosomal dominant congenital non-progressive ataxia overlaps with the SCA15 locus. Neurology. 2004;63:2288–2292. doi: 10.1212/01.wnl.0000147299.80872.d1. 15623688. [DOI] [PubMed] [Google Scholar]
- 73.Hara K., Shiga A., Nozaki H., Mitsui J., Takahashi Y., Ishiguro H., Yomono H., Kurisaki H., Goto J., Ikeuchi T., Tsuji S., Nishizawa M., Onodera O. Total deletion and a missense mutation of ITPR1 in Japanese SCA15 families. Neurology. 2008;71:547–551. doi: 10.1212/01.wnl.0000311277.71046.a0. 18579805. [DOI] [PubMed] [Google Scholar]
- 74.Schabhüttl M., Wieland T., Senderek J., Baets J., Timmerman V., De Jonghe P., Reilly M.M., Stieglbauer K., Laich E., Windhager R., Erwa W., Trajanoski S., Strom T.M., Auer-Grumbach M. Whole-exome sequencing in patients with inherited neuropathies: outcome and challenges. J. Neurol. 2014;261:970–982. doi: 10.1007/s00415-014-7289-8. 24627108. [DOI] [PubMed] [Google Scholar]
- 75.Laššuthová P., Šafka Brožková D., Krůtová M., Neupauerová J., Haberlová J., Mazanec R., Dřímal P., Seeman P. Improving diagnosis of inherited peripheral neuropathies through gene panel analysis. Orphanet. J. Rare Dis. 2016;11:118. doi: 10.1186/s13023-016-0500-5. 27549087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Shadrina M.I., Shulskaya M.V., Klyushnikov S.A., Nikopensius T., Nelis M., Kivistik P.A., Komar A.A., Limborska S.A., Illarioshkin S.N., Slominsky P.A. ITPR1 gene p.Val1553Met mutation in Russian family with mild spinocerebellar ataxia. Cerebellum Ataxias. 2016;3:2. doi: 10.1186/s40673-016-0040-8. 26770814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Klar J., Ali Z., Farooq M., Khan K., Wikstrom J., Iqbal M., Zulfiqar S., Faryal S., Baig S.M., Dahl N. A missense variant in ITPR1 provides evidence for autosomal recessive SCA29 with asymptomatic cerebellar hypoplasia in carriers. Eur. J. Hum. Genet. 2017;25:848–853. doi: 10.1038/ejhg.2017.54. 28488678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Carvalho D.R., Medeiros J.E.G., Ribeiro D.S.M., Martins B., Sobreira N.L.M. Additional features of Gillespie syndrome in two Brazilian siblings with a novel ITPR1 homozygous pathogenic variant. Eur. J. Med. Genet. 2018;61:134–138. doi: 10.1016/j.ejmg.2017.11.005. 29169895. [DOI] [PubMed] [Google Scholar]
- 79.Elert-Dobkowska E., Stepniak I., Krysa W., Ziora-Jakutowicz K., Rakowicz M., Sobanska A., Pilch J., Antczak-Marach D., Zaremba J., Sulek A. Next-generation sequencing study reveals the broader variant spectrum of hereditary spastic paraplegia and related phenotypes. Neurogenetics. 2019;20:27–38. doi: 10.1007/s10048-019-00565-6. 30778698. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Gorski M.M., Lecchi A., Femia E.A., La Marca S., Cairo A., Pappalardo E., Lotta L.A., Artoni A., Peyvandi F. Complications of whole-exome sequencing for causal gene discovery in primary platelet secretion defects. Haematologica. 2019;104:2084–2090. doi: 10.3324/haematol.2018.204990. 30819905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Stendel C., Wagner M., Rudolph G., Klopstock T. Gillespie's syndrome with minor cerebellar involvement and no intellectual disability associated with a novel ITPR1 mutation: report of a case and literature review. Neuropediatrics. 2019;50:382–386. doi: 10.1055/s-0039-1693150. 31340402. [DOI] [PubMed] [Google Scholar]
- 82.Cetani F., Pardi E., Aretini P., Saponaro F., Borsari S., Mazoni L., Apicella M., Civita P., La Ferla M., Caligo M.A., Lessi F., Mazzanti C.M., Torregossa L., Oppo A., Marcocci C. Whole exome sequencing in familial isolated primary hyperparathyroidism. J. Endocrinol. Invest. 2020;43:231–245. doi: 10.1007/s40618-019-01107-5. 31486992. [DOI] [PubMed] [Google Scholar]
- 83.Gonzaga-Jauregui C., Harel T., Gambin T., Kousi M., Griffin L.B., Francescatto L., Ozes B., Karaca E., Jhangiani S.N., Bainbridge M.N., Lawson K.S., Pehlivan D., Okamoto Y., Withers M., Mancias P. Exome sequence analysis suggests that genetic burden contributes to phenotypic variability and complex neuropathy. Cell Rep. 2015;12:1169–1183. doi: 10.1016/j.celrep.2015.07.023. 26257172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Prasad A., Rabionet R., Espinet B., Zapata L., Puiggros A., Melero C., Puig A., Sarria-Trujillo Y., Ossowski S., Garcia-Muret M.P., Estrach T., Servitje O., Lopez-Lerma I., Gallardo F., Pujol R.M. Identification of gene mutations and fusion genes in patients with Sezary syndrome. J. Invest. Dermatol. 2016;136:1490–1499. doi: 10.1016/j.jid.2016.03.024. 27039262. [DOI] [PubMed] [Google Scholar]
- 85.Dentici M.L., Barresi S., Nardella M., Bellacchio E., Alfieri P., Bruselles A., Pantaleoni F., Danieli A., Iarossi G., Cappa M., Bertini E., Tartaglia M., Zanni G. Identification of novel and hotspot mutations in the channel domain of ITPR1 in two patients with Gillespie syndrome. Gene. 2017;628:141–145. doi: 10.1016/j.gene.2017.07.017. 28698159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Hayashi S., Uehara D.T., Tanimoto K., Mizuno S., Chinen Y., Fukumura S., Takanashi J.I., Osaka H., Okamoto N., Inazawa J. Comprehensive investigation of CASK mutations and other genetic etiologies in 41 patients with intellectual disability and microcephaly with pontine and cerebellar hypoplasia (MICPCH) PLoS ONE. 2017;12 doi: 10.1371/journal.pone.0181791. 28783747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Hsiao C.T., Liu Y.T., Liao Y.C., Hsu T.Y., Lee Y.C., Soong B.W. Mutational analysis of ITPR1 in a Taiwanese cohort with cerebellar ataxias. PLoS ONE. 2017;12 doi: 10.1371/journal.pone.0187503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.van Dijk T., Barth P., Reneman L., Appelhof B., Baas F., Poll-The B.T. A de novo missense mutation in the inositol 1,4,5-triphosphate receptor type 1 gene causing severe pontine and cerebellar hypoplasia: expanding the phenotype of ITPR1-related spinocerebellar ataxia's. Am. J. Med. Genet. A. 2017;173:207–212. doi: 10.1002/ajmg.a.37962. 27862915. [DOI] [PubMed] [Google Scholar]
- 89.De Silva D., Williamson K.A., Dayasiri K.C., Suraweera N., Quinters V., Abeysekara H., Wanigasinghe J., De Silva D., De Silva H. Gillespie syndrome in a South Asian child: a case report with confirmation of a heterozygous mutation of the ITPR1 gene and review of the clinical and molecular features. BMC Pediatr. 2018;18:308. doi: 10.1186/s12887-018-1286-5. 30249237. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All constructs used are available freely on request. All data needed to evaluate the conclusions are presented in the paper or Supporting material.