Skip to main content
The Journal of Biological Chemistry logoLink to The Journal of Biological Chemistry
. 2024 Jul 24;300(9):107599. doi: 10.1016/j.jbc.2024.107599

O-GlcNAc transferase congenital disorder of glycosylation (OGT-CDG): Potential mechanistic targets revealed by evaluating the OGT interactome

Johnathan M Mayfield 1, Naomi L Hitefield 1, Ignacy Czajewski 2, Lotte Vanhye 3, Laura Holden 1, Eva Morava 3, Daan MF van Aalten 2,4,, Lance Wells 1,
PMCID: PMC11381892  PMID: 39059494

Abstract

O-GlcNAc transferase (OGT) is the sole enzyme responsible for the post-translational modification of O-GlcNAc on thousands of target nucleocytoplasmic proteins. To date, nine variants of OGT that segregate with OGT Congenital Disorder of Glycosylation (OGT-CDG) have been reported and characterized. Numerous additional variants have been associated with OGT-CDG, some of which are currently undergoing investigation. This disorder primarily presents with global developmental delay and intellectual disability (ID), alongside other variable neurological features and subtle facial dysmorphisms in patients. Several hypotheses aim to explain the etiology of OGT-CDG, with a prominent hypothesis attributing the pathophysiology of OGT-CDG to mutations segregating with this disorder disrupting the OGT interactome. The OGT interactome consists of thousands of proteins, including substrates as well as interactors that require noncatalytic functions of OGT. A key aim in the field is to identify which interactors and substrates contribute to the primarily neural-specific phenotype of OGT-CDG. In this review, we will discuss the heterogenous phenotypic features of OGT-CDG seen clinically, the variable biochemical effects of mutations associated with OGT-CDG, and the use of animal models to understand this disorder. Furthermore, we will discuss how previously identified OGT interactors causal for ID provide mechanistic targets for investigation that could explain the dysregulated gene expression seen in OGT-CDG models. Identifying shared or unique altered pathways impacted in OGT-CDG patients will provide a better understanding of the disorder as well as potential therapeutic targets.

Keywords: neurodevelopment, O-linked N-acetylglucosamine (O-GlcNAc), O-GlcNAc transferase (OGT), transcription, transcription regulation, histone modification, O-GlcNAcylation, post-translational modification (PTM), protein-protein interaction, intellectual disability


OGT is the sole enzyme responsible for the dynamic posttranslational modification of β-linked N-acetyl-D-glucosamine (O-GlcNAc) onto serine and threonine residues of thousands of nucleocytoplasmic proteins (1, 2, 3, 4, 5). Conversely, O-GlcNAcase (OGA) is responsible for the removal of this modification (6). In humans, OGT resides on the X chromosome (7). Based on the human exome database, OGT is one of the most conserved genes in the human genome and is the most conserved gene encoding a glycosyltransferase (8, 9). OGT shares 73% homology in its amino acid sequence with its Drosophila melanogaster ortholog encoded by the polycomb group (PcG) gene super sex combs (sxc) (7, 10). In mammals, knockout of OGT is embryonic lethal (7). Mammalian cells require OGT, and recent studies have begun to elucidate why OGT is essential (11). Minimum levels of glycosylation are required to maintain cell viability, and noncatalytic functions of OGT are crucial in normal proliferation (8). Specifically, knockout of OGT in mESCs results in increased proteasome activity raising the level of free amino acids, which in turn causes hyperactivation of mTOR, a sensor responsive to amino acid levels, resulting in mitochondrial dysfunction and cell death (12). Beyond glycosyltransferase activity and noncatalytic functions, OGT is also the protease responsible for host cell factor 1 (HCFC1) maturation (13, 14, 15, 16), although this function is not required for cell survival (8). For normal function, the transcriptional regulator HCFC1 must be proteolytically cleaved into N-terminal and C-terminal fragments with OGT being responsible for this in mammals and Taspase 1 being responsible in D. melanogaster (14, 15, 17). Proper HCFC1 maturation is required for regulating cell cycle progression and proliferation as well as gene expression (18, 19).

OGT has 13.5 N-terminal tetratricopeptide repeats (TPR), and this TPR domain is responsible for multiple protein-protein interactions, including substrate selection (Fig. 1A) (20). The C-terminal catalytic domain of OGT is a CAzy glycosyltransferase family 41 domain (21) and has an intervening domain with possible roles in the regulation of OGT and a proposed phosphatidylinositol (3, 4, 5)-trisphosphate (PIP3) binding domain (22, 23, 24, 25). O-GlcNAcylation is involved in numerous biological processes, and increasing evidence suggests a crucial role for OGT in the regulation of multiple aspects of transcription, signaling, and embryonic development (26, 27, 28). This review will summarize our current understanding of variants of OGT causal for the intellectual disability (ID) symptoms seen in patients with OGT-CDG as well as discuss interactors of OGT that are causal for disorders with ID phenotypes. These interactors, often with known O-GlcNAc sites (Fig. 2), provide targets for future investigation that could elucidate molecular mechanisms underpinning the neurodevelopmental phenotype seen in patients with OGT-CDG.

Figure 1.

Figure 1

Model figure: OGT structure, function, and possible mechanisms for OGT-CDG.A, domain map of OGT: OGT structure consists of both a N-terminal Tetratricopeptide repeat (TPR) domain (green) and a C-terminal catalytic domain (orange). The TPR domain of the nucleocytoplasmic isoform of OGT is composed of 13.5 degenerate 34 amino acid repeats and is involved in protein-protein interactions and substrate selection. The catalytic domain is responsible for the catalytic functions of OGT: HCFC1 cleavage and the addition of the O-GlcNAc modification onto substrate proteins. This domain contains both N-terminal and C-terminal catalytic regions linked by an intervening domain of unknown function (yellow). The functions of OGT are shown below the domain map. Depicted along the structure of OGT is a selection of the known variants of the protein, with those bolded and with an asterisk in the process of being characterized. B, the model this review proposes through which OGT is implicated in OGT-CDG: OGT interacts with and glycosylates components of complexes important in embryonic and neural development. This model suggests alterations in these OGT: protein interactions and hypoglycosylation of substrate proteins lead to dysregulation of gene expression and in part OGT-CDG phenotypes. Levels of UDP-GlcNAc in embryonic (high O-GlcNAc) and neural states (low O-GlcNAc) are essential for proper neurodevelopment, and this difference in substrate levels could explain the phenotype specificity of the disorder. Pink indicates proteins that interact with OGT but do not necessarily have known O-GlcNAc sites, termed here as binding partners. BHC, BRAF-HDAC complex; CBP, CREB-binding protein; EP300, Histone acetyltransferase p300; HCFC1, Host Cell Factor 1; Int. domain, intervening domain; KMTs, lysine methyltransferases; (n/np) BAF, neural/neuronal progenitor ATP-dependent BRG1/BRM associated factor; OGT, O-GlcNAc transferase; OGA, O-GlcNAcase; PR-DUB, Polycomb repressive deubiquitinase; TPR, tetratricopeptide repeat; UDP-GlcNAc, uridine diphosphate N-acetylglucosamine.

Figure 2.

Figure 2

Number of O-GlcNAc sites on identified OGT interactors. Shown are OGT interactors identified by Stephen, et al. 2020 (60) that are catalogued by the Online Mendelian Inheritance in Man (OMIM) database to have an intellectual disability (ID) phenotype. The number of mapped O-GlcNAc sites for each interactor was obtained from The O-GlcNAc Database (5) from the Medical College of Wisconsin. Interactors are ordered from least to most mapped O-GlcNAc sites. Zero mapped O-GlcNAc sites do not confirm a protein is not O-GlcNAc modified by OGT, instead that particular protein may have a low O-GlcNAc site occupancy, may be sub-stoichiometrically modified by OGT, may have cell-type-specific modification, or the sites were not mapped at the time of the manuscript preparation.

To date, nine variants of OGT have been established as causal for OGT-CDG. The initial variants characterized were restricted to the TPR domain of OGT (OGTL254F, OGTA259T, OGTR284P, OGTA319T, OGTE339G, and OGTΔ155-177) (29, 30, 31, 32). However, variants in the catalytic domain of OGT have since been characterized and linked to OGT-CDG (OGTN567K, OGTN648Y, and OGTC921Y), confirming causal variants for this disorder are not domain-specific but span the entire protein (33, 34, 35). Characterization of additional variants in both the TPR and catalytic domains of OGT is ongoing (Fig. 1A, bold∗). The phenotypic presentation of the patients, the biochemical characterization of the variants, and the use of animal models will be discussed, highlighting the variable nature of OGT-CDG in all three aspects. Currently, many unanswered questions remain regarding OGT-CDG. As more pathogenic OGT variants are characterized and additional patients are identified, it may become possible to identify a set of common alterations in OGT function as well as a core set of clinical manifestations of OGT-CDG, which would aid in the diagnosis of the disorder.

The OGT-CDG phenotype is variable

OGT-CDG patients present with a common set of core phenotypes but heterogenous secondary characteristics (Table 1). All patients exhibit core phenotypes of ID and developmental delay, often impacting their speech and language development. Common secondary features observed in patients with OGT-CDG include behavioral abnormalities such as Autism Spectrum Disorder (ASD) and Attention-Deficit/Hyperactivity Disorder (ADHD). Patients may present with brain abnormalities including microcephaly, eye abnormalities such as astigmatism, or dysmorphic features. Dysmorphic features vary from patient to patient but include coarse facial features, clinodactyly, or a dolichocephalic head. More variably presenting features include intense drooling, open mouth, hypotonia, and epilepsy or seizures. Finally, some patients affected by OGT-CDG present with more systemic phenotypes including connective tissue disorders, genital abnormalities, or ulcerative colitis. Overall, we see a high penetrance of neurological features; therefore, studies of OGT-CDG have focused on disease models relevant to neuronal development. These variants of OGT may also affect other organs or tissues given the heterogeneous nature of the clinical phenotypes present in patients with OGT-CDG. Additionally, OGTA319T co-segregates with a mutation in the MED12 gene, and OGTL254F patient phenotypes vary within three generations of the same family (29, 31). These results suggest that genetic and environmental background may be influencing the molecular consequences of individual mutations.

Table 1.

Identified OGT interactors linked to ID and their highlighted phenotypes and examples of known ID variants

Gene ID (MIM #) Protein name Alternative names UniProt accession Highlighted features of neurodevelopmental disorder Example ID variants References
BAP1 (619762) Ubiquitin carboxyl-terminal hydrolase BAP1 BRCA1 associated protein 1 Q92560 ID, developmental delay speech delay, motor delay, hypotonia, ASD/ADHD, dysmorphic facial features, eye abnormalities, skeletal malformations, growth failure P12T, P12A, E31K, L49P, C91S, H169R, R718Q (341)
KDM1A (616728) Lysine-specific histone demethylase 1A LSD1, BHC110 O60341 ID, developmental delay, speech delay, growth failure, clinodactyly, craniofacial dysmorphisms, hypotonia, eye abnormalities, thin corpus callosum E403K, D580G, Y785H (95, 115)
SMARCC2 (618362) SWI/SNF complex subunit SMARCC2 BAF170 Q8TAQ2 ID, speech delay, hypotonia, dysmorphic facial features, eye abnormalities, behavioral problems Splice site, W241X, L610P, M896V (342, 343, 344)
TBL1XR1 (616944, 602342) F-box-like/WD repeat-containing protein TBL1XR1 TBLR1 Q9BZK7 ID, delayed psychomotor development, dysmorphic facial features, speech delay, behavioral problems G70D, L282P, Y245C, frameshift (345, 346, 347, 348)
Developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, speech delay, hypotonia Y446C, Y445H, C325Y (349, 350, 351, 352)
ARID1B (135900) AT-rich interactive domain-containing protein 1B BAF250B Q8NFD5 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, delayed psychomotor development, speech delay, seizures, hypotonia, brain abnormalities, behavioral problems Frameshift and nonsense (353, 354, 355, 356, 357)
TRRAP (618364) Transformation/transcription domain-associated protein Tra1 homolog Q9Y4A5 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, fifth finger clinodactyly, hypotonia, seizures, brain abnormalities, behavioral problems R1986Q, L805F, A1043T, W1866R, W1866C (358, 359)
EP300 (618333, 613684) Histone acetyltransferase p300 P300, KAT3B Q09472 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, speech delay, behavioral problems Q1824P, deletion (360)
ID, dysmorphic facial features, ear abnormalities, eye abnormalities, delayed psychomotor development, hypotonia, ASD, behavioral problems Frameshift, termination, F1595V, N1286S (361, 362, 363, 364, 365, 366)
SMARCA4 (614609) Transcription activator BRG1 BRG1, BAF190 P51532 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, delayed psychomotor development, hypotonia, absence of speech, brain abnormalities Deletion, T859M, R885C, L921F, M1011T, R1157G (355)
PRR12 (619539) Proline-rich protein 12 KIAA1205 Q9ULL5 ID, developmental delay, Motor delay, speech delay, ASD/ADHD, Eye abnormalities, hypotonia, growth failure, microcephaly, fifth finger clinodactyly Frameshift, nonsense, splice site, R1169W, L1970P (367, 368, 369)
TNRC6B (619243) Trinucleotide repeat-containing gene 6B protein Q9UPQ9 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, delayed motor skills, speech delay, hypotonia, behavioral problems Termination, frameshift, splice site, V1357E (370, 371)
ASXL2 (617190) Putative polycomb group protein ASXL2 Q76L83 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, hypotonia, delayed speech, delayed psychomotor development, seizures, brain abnormalities, behavioral problems Frameshift (372)
AHDC1 (615829) AT-hook DNA-binding motif-containing protein 1 Gibbin, XIGIS Q5TGY3 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, hypotonia, delayed psychomotor development, brain abnormalities, behavioral problems Frameshift, D1457G (373, 374, 375, 376)
KMT2D (147920) Histone-lysine specific N-methyltransferase 2D MLL2 O14686 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, short fifth finger, seizures, hypotonia Termination (377)
OGT (300997) UDP-N-acetylglucosamine–peptide N-acetylglucosaminyl transferase 110 kDa subunit O15294 ID, developmental delay, speech delay, Behavioral problems, hypotonia, eye abnormalities, ear abnormalities, brain abnormalities (microcephaly, thin corpus callosum), dysmorphic features (clinodactyly), epilepsy, seizures L254F, A259T, R284P, A319T, E339G, N567K, N648Y, Δ155–177, N806S, S93L, A431V (30, 31, 32, 33, 34, 35, 98, 378, 379)
KMT2C (617768) Histone-lysine N-methyltransferase 2C MLL3 Q8NEZ4 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, hypotonia, delayed psychomotor development, speech delay, seizures, behavioral problems Termination, frameshift (380, 381)
PHF21A (618725) PHD finger protein 21A BHC80 Q96BD5 ID, developmental delay, dysmorphic facial features, clinodactyly, impaired motor skills, seizures, hypotonia, behavioral problems Frameshift, termination, G429S (116, 382)
SETD1A (619056) Histone-lysine N-methyltransferase SETD1A KMT2F, SET1A O15047 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, speech delay, seizures, behavioral problems Splice site, Y1499D, frameshift (92, 383)
SIN3A (613406) Paired amphipathic helix protein Sin3a Q96ST3 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, clinodactyly, hypotonia, brain abnormalities, behavioral problems Frameshift, termination (182, 384)
MCM3AP (618124) Germinal-center associated nuclear protein GANP O60318 ID, developmental regression, dysmorphic facial features, hypotonia, speech delay, delayed motor development, developmental regression, seizures E915K, V1272M, frameshift, termination, R878H, S951P, L870S (163, 385, 386, 387, 388)
CREBBP (618332, 180849) CREB-binding protein CBP, KAT3A Q92793 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, fifth finger clinodactyly, speech delay, seizures, behavioral problems C1710R, R1867Q, R1868W, M1872V, E1724K (389, 390)
ID, dysmorphic facial features, ear abnormalities, eye abnormalities, fifth finger clinodactyly, brain abnormalities, speech delay, seizures, hypotonia, behavioral problems Termination, R1378P, splicing (391, 392, 393)
TAF6 (617126) Transcription initiation factor TFIID subunit 6 TAFII70, TAFII80 P49848 ID, dysmorphic facial features, eye abnormalities, hypotonia, delayed psychomotor development, behavioral problems I108T, I71T, R46C (394, 395)
ARID1A (614607) AT-rich interactive domain-containing protein 1A BAF250A O14497 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, delayed psychomotor development, speech delay, hypotonia, seizures, brain abnormalities Frameshift, termination (355, 396, 397)
KMT2A (605130) Histone-lysine N-methyltransferase 2A ALL1, MLL1, HTRX Q03164 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, clinodactyly, hypotonia, delayed psychomotor development, speech delay, behavioral problems Frameshift, termination (398, 399)
ZC3H14 (617125) Zinc finger CCCH domain-containing protein 14 Q6PJT7 ID Termination (306)
KDM3B (618846) Lysine-specific demethylase 3B JMJD1B Q7LBC6 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, motor delay, speech delay, hypotonia, behavioral problems D336G, E1731K, D1032V, termination (119, 120)
CIC (617600) Protein capicua homolog Capicua homolog Q96RK0 ID, delayed psychomotor development, developmental regression, seizures, brain abnormalities, hypotonia, behavioral problems Termination, frameshift (400, 401)
POGZ (616364) Pogo transposable element with ZNF domain Q7Z3K3 ID, dysmorphic facial features, ear abnormalities, eye abnormalities, hypotonia, delayed psychomotor development, behavioral problems Frameshift, termination, splice site (215, 402, 403, 404, 405)
ANKRD17 (619504) Ankryin repeat domain-containing protein 17 GTAR O75179 ID, developmental delay, dysmorphic facial features, eye abnormalities, speech delay, motor delay, seizures Termination, G278V, A377T, L519P, R2434G, S1880P (406)
BPTF (617755) Nucleosome-remodeling factor subunit BPTF FAC1, Fetal Alzheimer Antigen Q12830 ID, dysmorphic facial features, eye abnormalities, clinodactyly, hypotonia, delayed psychomotor development, seizures, brain abnormalities Frameshift, A1924T, M2853R, termination, splicing, in-frame deletion (407, 408)
SPEN (619312) Msx2-interacting protein MINT, SHARP Q96T58 ID, developmental delay, dysmorphic facial features, ear abnormalities, eye abnormalities, hypotonia, brain abnormalities, behavioral problems Termination, frameshift (409)
NUP62 (271930) Nuclear pore glycoprotein p62 P37198 ID, developmental arrest and regression, eye abnormalities, brain abnormalities Q319P (410)
HCFC1 (309541) Host Cell Factor 1 HCF1 P51610 ID, delayed psychomotor development, seizures, hypotonia A864T, G876S, A897V (286, 287, 288, 289)

As shown in Figure 2, listed are OGT interactors catalogued by the Online Mendelian Inheritance in Man (OMIM) database as having an ID phenotype. Identification information for each protein included here are the Gene ID (with associated MIM identification), UniProt accession number, and alternative names by which the protein may be listed. Highlighted Features of Neurodevelopmental Disorder includes phenotypes seen in patients with causal mutations in that gene in relation to those features common to OGT-CDG patients. Example variants listed are variants in a particular interactor which have been found causal for neurodevelopmental disorders. Applicable references are listed. Proteins in bold have been linked to a neurodevelopmental disorder since the publication of the OGT interactome by Stephen et al. 2020 (60). Abbreviations used in this table include: ID (intellectual disability), ASD (autism spectrum disorder), and ADHD (attention-deficit/hyperactivity disorder).

OGT-CDG is an X-linked recessive disorder, and therefore it primarily affects males. In instances where OGT-CDG is inherited, the maternal parent tends to be unaffected by the disorder. X-inactivation, where one chromosome is preferentially inactivated, is often seen in disorders with X-linked intellectual disability (36). This skewing in females toward one X chromosome and the preferential inactivation of the other can be advantageous for survival particularly when there is a genetic defect on the inactivated X chromosome, as it will prevent the expression of the disease in that individual. Male progeny, however, do not share this protective effect and will show disease signs through the inheritance of the disease allele. This is the case with OGT-CDG, where unaffected females transmit the disorder to their sons, but the unaffected females are heavily skewed away from the disease allele, as opposed to the normal 50:50 inactivation of the OGT gene (37). Skewing of the X-chromosome in the maternal parent contributed to the prediction of causality of the OGTL254F mutation in the family analysis (31). Interestingly, OGTN567K represents a unique case where monozygotic female twins share a heterozygous missense mutation in OGT (34). A 98:2 skew in X inactivation was detected in both twins, while the maternal parent was unaffected. However, which X chromosome is inactivated is unclear. Despite sharing the same variant and showing the same skewing in X-inactivation, the severity of the OGT-CDG phenotype varied between the twins suggesting X-inactivation or neonatal environmental factors have a role in the severity of the phenotype in patients. Skewing also suggests that the wildtype allele must have a selective advantage in some fundamental process of early development.

In the clinical setting, there is a need to determine diagnostic techniques for addressing causality for ID. Initial key phenotypic characteristics to observe include cognitive disability and developmental delay, with eye abnormalities and behavioral abnormalities being additional strong indicators of an ID phenotype that could include an OGT-CDG diagnosis. Because of the variability of the dysmorphic features seen, pinpointing specific characteristics is difficult, other than generalizing a dysmorphic feature phenotype for OGT-CDG. However, these phenotypic diagnoses are not sufficient to confirm OGT-CDG. Exome sequencing should be used to identify what particular additional mutations a patient’s genome may harbor, and an exome database allows for ruling out common polymorphisms from unaffected individuals (9). Should a mutation in OGT appear, X-linked skewing of the maternal parent may provide evidence to the causal nature of the variant, particularly if the mother is unaffected with skewing towards the non-affected allele. Finally, patient cells can be assessed for changes relative to healthy controls. Previous work with OGTL254F, OGTR284P, and OGTΔ155-177 indicate OGT and OGA levels decrease in patient cells compared to wildtype controls, but no change in global O-GlcNAc levels is observed (30, 31). In OGTL254F lymphoblastoids, OGA mRNA and OGA promoter-reporter expression both decrease, which provides a potential mechanism by which normal global O-GlcNAcylation is maintained in these cells. Through RNA-sequencing analysis, OGTL254F lymphoblastoids are found to have a subset of disease-relevant differentially regulated genes compared to wild-type controls (31). More research is necessary to identify which genes may be optimal candidates and if this effect is seen for other variants, but this suggests that utilizing RNA-sequencing to search for specific genetic biomarkers may help in assigning causality of OGT-CDG variants.

OGT-CDG variants exhibit variable altered biochemical properties

While the biochemical properties of the OGT-CDG variants studied to date are quite variable (Table 2), there are methods to predict pathogenicity. In silico analyses algorithms, such as Polyphen and SIFT (38, 39), have been used to predict whether OGT-CDG variants are likely deleterious (Table 2) (30, 31, 32). The conservation of amino acid residues can also be used to predict pathogenicity, as the pathogenic variants described to date affect residues that are conserved across vertebrate and invertebrate species (30, 31, 32, 33, 34, 35). Moreover, computational modeling of variants allows for predictions to be made as to the type of effect a variant might have on the enzyme (31, 32, 33, 34, 35). For example, TPR domain variants were predicted to result in steric clashes or loss of secondary structures while catalytic domain variants were predicted to affect substrate binding. Structural studies of OGT-CDG variants have allowed for these predicted changes to be examined. The crystal structures of OGTN567K and OGTN648Y suggest changes in the catalytic domain alter substrate binding (33, 34). Moreover, the crystal structure of OGTL254F shows a small change in TPR 7, and this deviation propagates toward the N terminus showing that it destabilizes the interface between TPRs 6 and 7 (40). In further confirmation of the effects of TPR domain variants on the structure, the spring constants, which measure the elasticity of the TPR domain, were determined for OGTL254F, OGTR284P, and OGTA319T, and all three have alterations that could lead to changes in the ability of these enzymes to adapt to different substrates and complexes (41). Finally, thermal stability assays have also been employed to determine changes in protein stability with varying degrees of change for the TPR domain variants (Table 2) (30, 32, 40). Overall, studies examining the structure of OGT-CDG variants point toward domain-specific changes: TPR domain variants have altered structure which could affect protein-protein interactions, substrate selection, and stability, while catalytic domain variants have changes in the binding site of substrates.

Table 2.

Summary table of OGT-CDG variant characterization

Variant L254F A259T R284P A319T E339G
Domain TPR TPR TPR TPR TPR
Polyphen Score 0.999 1.000 0.967 0.999 1.000
In vitro Glycosyltransferase Activity
  • Normal activity (CKIIα)

  • ↓ activity (de-O-GlcNAcylated HEK293 cell lysate)

  • Normal activity (HEK293 overexpression system)

Normal activity (CKIIα)
  • Normal activity (CKIIα)

  • ↓ activity (de-O-GlcNAcylated HEK293 cell lysate)

Normal activity (CKIIα) Normal activity (CKIIα)
In vitro HCFC1 Glycosyltransferase Activity Normal activity Normal activity Reduced activity Normal activity Normal activity
In vitro HCFC1 Protease Activity Normal activity Normal activity Reduced activity Normal activity Normal activity
Kinetic Analysis
  • Km 1.86-fold change, kcat 0.78-fold change versus WT (CKIIα protein substrate)

  • Km and Vmax unchanged (peptide substrate)

Km 1.29-fold change, kcat 0.78-fold change versus WT (CKIIα protein substrate) Km 1.29-fold change, kcat 0.78-fold change versus WT (CKIIα protein substrate) Km 2.07-fold change, kcat 0.73-fold change versus WT (CKIIα protein substrate) Km 1.71-fold change, kcat 0.89-fold change versus WT (CKIIα protein substrate)
Dimerization Normal dimerization of recombinant TPR domain Normal dimerization of recombinant TPR domain Normal dimerization of recombinant TPR domain Normal dimerization of recombinant TPR domain Normal dimerization of recombinant TPR domain
Thermal Stability
Global O-GlcNAc Levels Unchanged (hESCs) Unchanged (hESCs) Unchanged (hESCs) N/A Unchanged (hESCs)
Global OGT Levels Unchanged (hESCs) Unchanged (hESCs) Unchanged (hESCs) N/A Unchanged (hESCs)
Global OGA Levels Unchanged (hESCs) Unchanged (hESCs) Unchanged (hESCs) N/A Unchanged (hESCs)
Gene Expression
  • Altered (patient derived lymphoblastoids)

  • Altered (hESCs)

Altered (hESCs) Altered (hESCs) N/A Altered (hESCs)
Pluripotency Markers Unchanged POU5F1 (Oct4), SOX2, NANOG (hESCs) Unchanged POU5F1 (Oct4), SOX2, NANOG (hESCs) Unchanged POU5F1 (Oct4), SOX2, NANOG (hESCs) N/A Unchanged POU5F1 (Oct4), SOX2, NANOG (hESCs)
Animal and Cell Models Drosophila:
  • Non-performers in habituation learning

  • ↑ synaptic bouton number, ↑ NMJ length (not significantly), ↑ number of NMJ branches

N/A Drosophila:
  • ↓ climbing speed

  • Non-performers in habituation learning (homozygous)

  • Deficits in habituation learning (heterozygous)

  • ↑ synaptic bouton number, ↑ length of NJM, ↑ perimeter of NMJ, ↑ number of NMJ branches

Drosophila:
  • Deficits in habituation learning (homozygous)

  • ↑ synaptic bouton number, ↑ NMJ length (not significantly), ↑ NMJ length

N/A
References (31, 32, 40, 49) (32) (30, 32, 49) (32, 49) (32)

Variant Δ155–177 N567K N648Y C921Y
Domain TPR Catalytic Catalytic Catalytic
Polyphen Score N/A 0.998 1.000 1.000
In vitro Glycosyltransferase Activity N/A ↓ (TAB1) ↓ (TAB1) ↓ (TAB1)
In vitro HCFC1 Glycosyltransferase Activity N/A ↓ activity N/A ↓ activity
In vitro HCFC1 Protease Activity N/A ↓ activity N/A Normal activity
Kinetic Analysis N/A Km 1.45-fold change, kcat 0.17-fold change, kcat/Km 0.08-fold change versus WT (peptide substrate) Kd 6.88-fold change versus WT (fluorescent probe) Vmax 0.25-fold change, Km 0.18-fold change, kcat/Km unchanged versus WT (peptide substrate)
Dimerization N/A N/A N/A N/A
Thermal Stability ↓ thermal stability N/A Normal thermal stability Normal thermal stability
Global O-GlcNAc Levels
  • Unchanged (patient derived fibroblasts)

  • Unchanged (iPSCs)

  • ↓ (differentiated iPSCs)

  • ↓ (NPCs)

Unchanged (pluripotent mESCs) ↓ (mESCs)
  • ↓ (undifferentiated, mESCs)

  • Unchanged (mESCs, clonogenic conditions)

Global OGT Levels
  • ↓ (patient derived fibroblasts)

  • ↓ (iPSCs)

  • ↓ (differentiated iPSCs)

  • ↓ (NPCs)

Unchanged (pluripotent mESCs) Unchanged (mESCs)
  • ↑ (undifferentiated mESCs)

  • (mESCs, clonogenic conditions)

Global OGA Levels
  • ↓ (patient derived fibroblasts)

  • ↓ (iPSCs)

  • ↓ (differentiated iPSCs)

  • ↓ (NPCs)

  • ↓ (pluripotent mESCs)

  • Unchanged (differentiated mESCs)

↓ (mESCs)
  • Unchanged (undifferentiated mESCs)

Unchanged (mESCs, clonogenic conditions)
Gene Expression
  • Unchanged OGT, OGA mRNA (patient derived fibroblasts)

  • OGT, ↓ OGA mRNA (iPSCs)

  • Unchanged OGT, ↓ OGA mRNA (differentiated iPSCs)

  • NEFH, all other ectoderm markers unchanged (differentiated iPSCs)

  • Unchanged OGT, OGA mRNA (NPCs)

↑ GABPA mRNA (mESCs) N/A Unchanged Ogt, ↓ Oga mRNA (undifferentiated mESCs)
Pluripotency Markers
  • Unchanged POU5F1 (Oct4), SOX2, NANOG mRNA, Unchanged NANOG, OCT4, SOX2 protein (iPSCs)

Unchanged Sox2, Oct4 during first 6 days of differentiation (mESCs) N/A
  • Unchanged Oct4 and Sox2 mRNA, Oct4 and Sox2 protein (undifferentiated mESCs)

↓ Oct4, Sox2 protein (mESCs, clonogenic conditions)
Animal and Cell Models
  • Unchanged pluripotency, differentiation, and morphology (iPSCs)

  • Unchanged neural progenitor markers (NPCs)

  • Fewer and larger rosettes, larger apical lumens (NPCs)

Drosophila:
  • ↓ O-GlcNAc (adult heads)

  • mESCs:

  • ↓ neurite length

  • ↓ proteolytic cleavage of HCFC1

  • Mice:

  • Altered inheritance distribution

  • ↓ Global GlcNAc, ↓ OGA protein, ↓ OGT protein (brain)

  • Oga, ↓ Ogt mRNA, ↓ DIs, ↓ PEs (brain)

  • ↓ body weight, ↓ nose-to-tail length, ↑ pancreas weight, slim appearance

  • ↓ brain weight

Mice:
  • ↓ Global GlcNAc, ↓ OGA protein, ↓ OGT protein (brain)

  • Oga, ↓ Ogt mRNA, ↓ DIs, ↓ PEs (brain)

  • ↓ body weight, ↓ nose-to-tail length, ↑ pancreas weight, slim appearance

  • ↓ brain weight

Drosophila:
  • ↓ O-GlcNAc, ↑ OGT protein (adult heads)

  • ↓ O-GlcNAc, ↑ OGT protein (embryos)

  • ↓ O-GlcNAc, unchanged OGT protein (larvae)

  • ↑ ectopic bristle penetrance

  • ↓ NMJ area, ↓ NJM length, ↓ NMJ bouton numbers

  • ↓ total sleep, ↓ duration and more frequent sleep bouts

  • mESCs:

  • alkaline phosphatase staining (clonogenic conditions)

  • Mice:

  • ↓ Global GlcNAc, ↓ OGA protein, ↓ OGT protein (brain)

  • Oga, ↓ Ogt mRNA, ↓ DIs, ↓ PEs (brain)

  • ↓ body weight, ↓ nose-to-tail length, ↓ body fat mass, ↑ lean body mass, ↓ glycemia, ↑ pancreas weight, slim appearance

  • ↓ skull length, ↓ skull size, ↓ absolute brain weight, rounder and smaller skull

  • Hypothesized anxiety phenotype, compulsive behavior, altered spatial working memory

References (30, 44) (34, 50) (33, 50) (35, 50, 51)

Summarized are characteristics of the six Tetratricopeptide repeat (TPR) domain variants and three catalytic domain variants of OGT discussed in this review. This table shows the results of characterizations of variants in vitro, in cellulo, and in vivo alongside phenotypes observed in vivo. Abbreviations used within include WT (wild-type), TPR (Tetratricopeptide repeat), CKIIα (casein kinase II alpha), TAB1 (TAK-1 binding protein), hESCs (human embryonic stem cells), iPSCs (induced pluripotent stem cells), mESCs (mouse embryonic stem cells), NPCs (neural progenitor cells), NMJ (neuromuscular junction), DI (detained introns), PE (decoy exon), ↑ (upregulated or increased), ↓ (downregulated or decreased), N/A (information unavailable or not determined).

Unlike structural effects, OGT-CDG variants have divergent biochemical properties that do not appear to be domain-specific. All variants to date have been tested for their ability to glycosylate substrates and some for their ability to cleave HCFC1 (Table 2). OGTL254F raises global O-GlcNAc levels in a HEK293 overexpression system to the same degree as wildtype (31), but this same variant has a reduced ability to modify de-O-GlcNAcylated HEK293 lysates, as was also seen for OGTR284P (30, 40). However, all characterized TPR domain variants glycosylate known OGT substrate Casein Kinase II alpha (CKIIα), though with a modest increase in their KM values for the protein substrate (31, 32, 42). This is in contrast to the catalytic domain variants which are greatly impaired in their ability to modify the protein substrate TAK-1 binding protein (TAB1) (33, 34, 35, 43). Kinetics analysis of catalytic domain variants also reveals markedly altered kinetics for peptide substrates (33, 34, 35). In vitro assays investigating the ability of OGT variants to cleave and glycosylate HCFC1 show an inability for OGTN567K to cleave or glycosylate HCFC1, while when examining the TPR domain variants only OGTR284P has a slight decrease in this enzymatic activity (32, 34). Despite the variability of OGT-CDG biochemical properties (Table 2), these results inform future studies. In the future characterization of OGT-CDG variants, HCFC1 glycosylation and cleavage should be probed as this defect may play a role in the mechanism of OGT-CDG for some variants. Furthermore, glycosylation assays show clear defects for catalytic domain variants while TPR domain variants have modest changes in kinetics for a protein substrate. While TPR domain variants have modest changes in glycosylation activity, a reduction in glycosyltransferase activity nonetheless suggests this group of genetic disorders is in fact a disorder of glycosylation. However, the TPR domain variants likely only have a subset of substrates that are differentially O-GlcNAcylated compared to the drastic O-GlcNAcylation defects seen for catalytic domain variants. While these studies provide useful information about the OGT-CDG enzymes, model systems, including relevant cell types and animal models, will help elucidate the intricacies underlying the OGT-CDG phenotype.

Model organisms provide opportunities to elucidate the OGT-CDG mechanisms

Initial experimental work on cell and animal models of OGT-CDG has focused on identifying key phenotypic consequences of disrupting OGT function, which may indicate processes particularly sensitive to OGT dysfunction. OGT is known to be essential for mammalian embryogenesis (27) and many phenotypes seen in OGT-CDG patients indicate defects in early development (i.e. developmental delay, Table 1). Therefore, studies have focused on relevant cell types, namely human and mouse embryonic stem cells (ESCs) (32, 33, 34, 35) and more recently patient cells reprogrammed into induced pluripotent stem cells (iPSCs) (44). Catalytic domain variants have been modeled in mouse ESCs (mESCs), and these models demonstrate heterogeneous alterations in global O-GlcNAcylation, OGT, and OGA levels (Table 2) (33, 34, 35). OGTC921Y and OGTN648Y variants result in lower levels of global O-GlcNAc in mESCs. The OGTN567K variant does not show a decrease in global O-GlcNAc, however, this may be explained by decreased OGA levels in cells modeling this mutation, a phenotype also observed in OGTN648Y mESCs. Furthermore, OGTC921Y shows an increase in OGT levels. Thus, while the changes in the O-GlcNAc cycling enzymes and global O-GlcNAc levels are different among the catalytic domain variants, the results suggest that the cells are attempting to maintain O-GlcNAc homeostasis in response to the variant enzymes. Further heterogeneity of the catalytic domain variants is seen in their ability to process HCFC1. While OGTC921Y is deficient in its ability to glycosylate HCFC1 in vitro, there is no apparent effect on HCFC1 cleavage in vitro or in mESCs (35). This is in contrast to OGTN567K, which shows a marked decrease in the proteolytic fraction of HCFC1 (34). Moreover, OGTN567K mESCs shows a change in expression of a gene regulated by HCFC1, Ets transcription factor GA-binding protein subunit alpha (GABPA) (34). Given that HCFC1 proteolysis is required for normal function in gene expression (13, 14, 15, 16), it is possible that HCFC1 plays a role in the OGT-CDG phenotype for some variants. Furthermore, iPSCs have been generated for a mutation resulting in the deletion of exon 4 of the TPR domain, OGTΔ155-177 (44). While global O-GlcNAc levels are maintained during an undifferentiated state, upon differentiation towards an ectodermal fate global O-GlcNAc levels are altered for the variant in comparison to a corrected iPSC control. OGTΔ155-177 results in a decrease of OGT levels but maintains O-GlcNAc homeostasis by downregulating OGA levels. It appears the homeostatic mechanism regulating the expression of OGT and OGA is insufficient to maintain normal O-GlcNAcylation despite a continued decrease in OGA levels in neural progenitors. This heterogeneity in the homeostasis of O-GlcNAc and its cycling enzymes is not seen when TPR domain variants are modeled in human ESCs (hESCs). Four TPR domain variants (OGTL254F, OGTA259T, OGTR284P, and OGTE339G) show similar levels of global O-GlcNAc, OGT, and OGA (32). The divergence in the models suggests that there may be multiple underlying mechanisms to the OGT-CDG phenotype or differences arising from the cell types used. Some degree of variability seen in patients may arise not only due to differences in genetic background but also due to heterogeneous effects of OGT mutations on the multiple functions fulfilled by OGT.

Beyond alterations to global O-GlcNAc homeostasis, studies have also attempted to uncover downstream effects of OGT-CDG variants on cell and animal models. Some variants have been probed for changes in the pluripotency network, namely SOX2 and OCT4 (Table 2). TPR domain variants modeled in hESCs showed no differences in the expression of SOX2 or OCT4 (32). However, OGTC921Y shows decreased levels of SOX2 and OCT4 protein levels (35), while OGTN567K mESCs and iPSCs of OGTΔ155-177 have normal levels of SOX2 and OCT4 (34, 44). The changes in SOX2 and OCT4 protein levels could explain the altered ability of OGTC921Y mESCs to maintain pluripotency upon withdraw of Leukemia Inhibitory Factor, LIF. Interestingly, upon differentiation of mESCs, OGTN567K shows a decrease in neurite length, while OGTΔ155-177 iPSCs show fewer and larger neural rosettes (34, 44). Alterations in neural rosette formation in vitro further point toward early developmental changes as neural rosette formation is reminiscent of secondary neurulation in vivo (45). However, an important caveat to comparing results from human and mouse cell models is that transcriptional networks maintaining an undifferentiated state differ substantially between these two organisms (46). While variable, these results from models in mESCs and iPSCs provide evidence for multiple downstream targets being affected. Additionally, the transcriptomic profiling of four different TPR domain variants (OGTL254F, OGTA259T, OGTR284P, and OGTE339G) in hESCs shows a consistent phenotype. In this model, several pathways are found to be affected across the four mutations assessed. Notably, genes belonging to “mesoderm development” and “ectoderm development” gene ontology (GO) terms are significantly dysregulated (32). Altered expression of ectoderm and mesoderm genes could explain the ID as well as dysmorphic facial features and clinodactyly often seen in patients (Table 1). An additional pathway of interest downregulated across the four TPR domain variants is the Liver X Receptor/Retinoid X Receptor (LXR/RXR) activation pathway (32). This pathway has been implicated in dopaminergic neuron development, which may be consequential for memory and learning deficits seen in patients (47, 48). Overall, cell models to date have elucidated that gene expression is dysregulated and that the effects of OGT-CDG variants are cell-type specific. Beyond alterations in OGT and OGA mRNA expression, four TPR domain variants have altered gene expression in hESCs, and OGTN567K has dysregulated expression of GABPA in mESCs pointing toward a common downstream impact on gene expression for both TPR domain and catalytic domain variants. Furthermore, the study utilizing iPSCs shows that the cell type is essential for uncovering the effects of OGT-CDG variants and further points toward the neural lineage as sensitive to perturbations in OGT function.

While disease-relevant cell models are invaluable for understanding the role of OGT in the earliest stages of development, many features of the OGT-CDG phenotype (Table 1) can only be recapitulated in animal models. OGT-CDG variants have been modeled in both Drosophila and mice, though only catalytic domain variants have been modeled in the latter (Table 2) (49, 50, 51). Broadly, all catalytic domain mutations modeled in either Drosophila (equivalent to OGTN567K and OGTC921Y) or mice (equivalent to OGTN567K, OGTN648Y, and OGTC921Y)reduce global O-GlcNAcylation in brain tissue to varying degrees, while TPR domain mutations modeled in Drosophila (equivalent to OGTL254F, OGTR284P, and OGTA319T) do not broadly alter O-GlcNAc levels. In mice and consistent with mESC models, the three catalytic domain variants result in decreased OGA protein levels, at least in part through decreased mRNA levels (50). Unexpectedly, OGT protein levels are also significantly decreased in mouse models of OGT-CDG. This decrease in OGT protein levels likely occurs through posttranscriptional or posttranslational mechanisms, as OGT mRNA levels are significantly increased in these mice, unlike in cell models of OGT-CDG (50). These findings further underscore the need to better understand the mechanisms governing O-GlcNAc cycling, including genetic regulation of OGT and OGA, in different cell types and various stages of development and how these are affected by OGT-CDG mutations.

Beyond changes in O-GlcNAc homeostasis, mice and Drosophila models provide opportunities to study anatomical and behavioral effects caused by the OGT-CDG variants (Table 2). Behavioral assays in Drosophila suggest TPR domain variants (equivalent to OGTR284P and OGTA319T) disrupt memory (49). In particular, habituation in these flies is disrupted, which can be rescued by knocking out OGA or by mutagenizing a key catalytic residue. Similarly, knocking out OGA can rescue sleep defects seen in a Drosophila model of catalytic domain variant OGTC921Y (51). These results suggest that behavioral defects in Drosophila models of OGT-CDG may be conveyed through reduced O-GlcNAcylation. Furthermore, the effects of OGT-CDG variants on synaptic development have been assayed through imaging morphological features of the Drosophila larval neuromuscular junction (NMJ), which is the synaptic connection between neurons and muscle cells (49). TPR and catalytic domain variants display inconsistent results regarding the development of this synapse (49, 51). TPR domain mutations increase growth at this axonal terminal, resulting in the formation of more NMJ boutons, which are recognizable structures that contain synaptic vesicles (49). Conversely, the OGTC921Y variant modeled in the fly results in decreased growth at the NMJ, characterized by reduced bouton number and NMJ area. This is phenocopied in larvae homozygous for a catalytically dead variant of OGT and partially rescued by restoring O-GlcNAcylation through knocking out OGA, suggesting this phenotype may be conveyed through impaired O-GlcNAc cycling. In mouse models of catalytic domain variants, low birth weight, microcephaly, and dysmorphic skull features seem to recapitulate the developmental delay, microcephaly, and dysmorphic features seen in patients (Tables 1 and 2) (50). Furthermore, mice with the OGTC921Y variant show impairments in spatial memory, increased anxiety, and hyperactivity, though hyperactivity is not recapitulated in Drosophila (50, 51). Because OGT-CDG patient phenotypes can be reproduced in animal models, including the symptom heterogeneity observed, these models can be used for further study of downstream targets of OGT that may contribute to the mechanisms by which the disorder is conveyed.

The neuroectodermal lineage is sensitive to perturbations in O-GlcNAc homeostasis

The neural-specific phenotype in OGT-CDG patients and the variable results from animal and cell models suggest that there are particular cell types that might be susceptible to alterations in OGT function and O-GlcNAc cycling, especially during early development. In a zebrafish model, ogt transcripts are high throughout the developing embryo pre-gastrulation but have a more localized expression in the brain post-gastrulation (52). In humans, alterations in O-GlcNAc levels in utero caused by gestational diabetes link OGT to neural tube defects (53, 54). In mammals, it has been shown that global O-GlcNAc levels decrease over-development (55). Moreover, multiple studies indicate O-GlcNAc homeostasis is required for normal differentiation of neuroectodermal lineage cells. Perturbations in O-GlcNAc cycling enzymes, and thus altered global O-GlcNAc levels, result in altered neuronal differentiation (56, 57, 58, 59). Furthermore, these alterations in differentiation are accompanied by changes in gene expression of neurogenic transcription factors as well as synaptic proteins. A decrease in global UDP-GlcNAc levels observed during differentiation of embryonic cells into neural lineage cells (56) could serve as a metabolic reprogramming mechanism by which OGT-mediated expression of genes is down-regulated during differentiation (Fig. 1B). The most recent study utilizing iPSCs generated from a patient further confirms the sensitivity of the ectodermal lineage to changes in O-GlcNAc homeostasis (44). Thus, O-GlcNAc homeostasis during development may cause certain cell lineages, such as the neuroectodermal lineages, to be sensitive to alterations in O-GlcNAc cycling and may lead to the OGT-CDG phenotype seen in patients. Furthermore, common alterations in glycosyltransferase activity and changes in gene expression, seen most notably in the TPR domain variant hESCs and in patient lymphoblastoids of OGTL254F, suggest that in part the mechanism behind OGT-CDG lies in hypoglycosylation of key substrates leading to altered gene expression and neurodevelopment. In an attempt to determine the most relevant possible mechanistic targets for OGT-CDG studies, we chose to evaluate a previous interactome generated for OGT’s TPR domain in HeLa cells (60). In the next section, we will discuss how the OGT interactome is enriched for proteins that are causal for an ID phenotype and regulate gene expression at multiple levels.

The OGT interactome is enriched for causal intellectual disability proteins

With the variable biochemical and phenotypic features of causal OGT variants, the next step in understanding the mechanisms underlying the phenotype is to determine relevant targets that could be conveying the phenotype. Since the phenotype seen in patients includes ID and developmental delay regardless of whether the variant lies in the TPR or catalytic domain, there are likely downstream targets of OGT that are hypoglycosylated and are responsible for normal neurodevelopment. This hypoglycosylation is likely a result of reduced catalytic activity from variants of the catalytic domain or as a result of changes in protein-protein interactions resulting from the altered structure of the TPR domain from variants of the TPR domain. Given the key role of OGT in transcriptional regulation and the changes in gene expression seen for the OGT-CDG variants tested, downstream targets are likely responsible for proper spatiotemporal gene expression during embryonic development and downstream targets could be required for normal neuronal functioning.

Previously, a bioinformatics approach using sequence conservation, structural data, and clinical data was used to determine possible targets of OGT that are also linked to neurodevelopmental disorders (61). Many targets with mapped O-GlcNAc sites were identified as potential conveyors of the ID phenotype, and these substrates represent targets worth investigating in the context of OGT-CDG. Here, we expand on this list by discussing interactors of OGT identified in a proximity labeling (BioID) approach for OGT’s TPR domain (60). This basal OGT interactome in HeLa cells contained 24 proteins with OMIM-catalogued neurodevelopmental disorders at the time of initial publication out of 115 total (Table 1). Some of these interactors were also identified using the previously mentioned bioinformatics approach including ARID1A, EP300, and HCFC1(61). This list of ID-linked interactors has since expanded to 32 out of a total of 115, and these data suggest that interactors and substrates of OGT are enriched for proteins implicated in neurodevelopment (Table 1, bold) as has been shown for the Drosophila O-GlcNAcome (49). These interactors represent a diverse array of proteins, but broadly all are regulators of gene expression (Fig. 3). This is in line with other generated OGT interactomes, in which regulators of transcription are enriched (62, 63, 64, 65). Furthermore, the interactors discussed herein are represented in the other interactomes from other cell types as well as from mice. While not all of these interactors have mapped O-GlcNAc sites (Fig. 2), many of these have mapped O-GlcNAc sites in mESCs (i.e. SIN3A, HCFC1, KMT2D, ZC3H14, POGZ, PHF21A, ANKRD17, SPEN, and TAF6) (66). Additionally, SIN3A, HCFC1, KMT2D, ZC3H14, and POGZ all had differentially enriched O-GlcNAc sites upon differentiation with Retinoic Acid (RA) towards an ectodermal fate. While this list of interactors does not represent an exhaustive list of potential targets, they are enriched for neurodevelopmental disorders, represent key regulators of gene expression, and represent both substrates and interactors of OGT which uniquely positions them as critical targets of investigation to understand potential mechanisms of OGT-CDG as well as to better understand the complex role of OGT in regulating gene expression.

Figure 3.

Figure 3

OGT interactors involve multiple aspects of transcription regulation. OGT interactors involve multiple aspects of the regulation of gene expression. All OGT interactors are not only predicted to be causal for disorders with ID phenotypes, like seen in OGT-CDG, but they also all have roles in the regulation of gene expression. Depicted here is each of the 32 interactors discussed in this review and where each plays a role in regulating gene expression: X-inactivation, nucleosome remodeling, histone methylation, histone acetylation, histone ubiquitination, DNA methylation, transcription factors, mRNA processing, nuclear import/export, and mRNA silencing. Ac, acetylation; Me, methylation; RNA Pol II, RNA polymerase II; Ub, ubiquitination.

These interactors consist of many epigenetic regulators of histone modifications including histone methylation, acetylation, and ubiquitination, as well as regulators of mRNA processing, export, and silencing (Fig. 3). In the context of O-GlcNAc biology, some interactors have well-established links to OGT while others are uncharacterized in this regard. Some interactors, like HCFC1 and NUP62, have numerous mapped O-GlcNAc sites (Fig. 2), and only a few interactors, like ARID1B and BAP1, have only one or no mapped O-GlcNAc sites (Fig. 2). Furthermore, many of these proteins have known interactions with each other (Fig. 4) suggesting the existence of OGT-containing multiprotein complexes (Fig. 1B). In support of this concept, several established complexes were recovered in the OGT interactome including many members of the BAF (mSWI/SNF) complex and the Polycomb Repressive Deubiquitinase (PR-DUB) complex. The OGT interactome provides disease-relevant targets that should be investigated to understand possible OGT-CDG mechanisms as well as to better understand the role of OGT in the regulation of gene expression.

Figure 4.

Figure 4

Direct interactions between identified OGT interactors. Using the Biological General Repository for Interaction Datasets (BioGRID) database, we searched which proteins in our group of OGT interactors (Table 1) interacted with other proteins within the subset. If BioGRID contained publications indicating an interaction between two proteins, a blue box is present (example: OGT-HCFC1). A lack of interaction between interactors in BioGRID is denoted by a yellow box (example: PRR12-AHDC1). A yellow box with an asterisk indicates an interaction detected between OGT and another protein that was identified by Stephen et al. 2020 (60) but not in the BioGrid database.

Histone methylation

The OGT interactome is enriched for proteins involved in the regulation of histone methylation including four histone methyltransferases, two histone demethylases, and one reader/scaffolding protein that is all causal for neurodevelopmental disorders (Fig. 3) (60, 67, 68). KMT2A, KMT2C, KMT2D, and SETD1A (KMT2F) are all catalytic subunits of the MLL/SET1 complexes responsible for H3K4me1/2/3 histone methylation and transcriptional activation of genes (Fig. 3) (69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79). Numerous studies have linked dysfunction or insufficiency of these histone methyltransferases to alterations in neurodevelopment and neuronal function (80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95). Moreover, many of the phenotypes observed in models of ID caused by mutations in these methyltransferases can be rescued both in vitro and in vivo (82, 85, 87, 93, 96, 97). In particular, pyruvate supplementation and inhibition of RAS/MAPK, HDACs, and KDM1A have been shown to rescue some or all of the observed phenotypes. Glucosamine supplementation and OGA inhibition have been previously suggested as an intervention for OGT-CDG patients (98), though these supplementations have not been tested in vitro for OGT-CDG variants. KDM1A (LSD1) was also enriched in the OGT interactome, but no studies have mapped O-GlcNAc sites to the protein (Fig. 2). However, the histone methylation reader and scaffold for the CoREST complex (99, 100, 101), PHF21A, has mapped O-GlcNAc sites that could be functionally relevant. Both KDM1A and PHF21A have been linked to neurodevelopmental disorders and have been studied in functionally relevant models (102, 103, 104, 105, 106, 107, 108, 109, 110, 111, 112, 113, 114, 115, 116, 117, 118). Together with other members of the CoREST complex, CoREST, and HDAC1/2, KDM1A functionally silences genes that are normally expressed in neuronal cell lineages. Furthermore, KDM1A has a neuron-specific isoform that is catalytically inactive but is believed to be responsible for the de-repression of neuron-specific genes by inhibiting the function of the ubiquitous KDM1A isoform. Any role of OGT in the regulation of this complex would be worth studying in relevant neural cell lines. KDM3B, another histone demethylase, has recently been reported to be causal for a neurodevelopmental disorder (119, 120), and one study has shown a link between this histone demethylase and neurodevelopmental and neuronal function (121). Lastly, JMJD1C is another interactor identified that has functional roles in histone methylation, and two recent reports have linked it to a neurodevelopmental disorder, but the association needs further confirmation (122, 123). Functionally, perturbing O-GlcNAc levels during neurogenesis results in alterations of histone methylation, and in particular, the levels of H3K4me3 on neurogenic transcription factor promoters increase alongside an increase in KMT2D and a decrease in KDM1A in vitro (58). Moreover, a rat model of maternal hyperglycemia showed an increase in KMT2A levels. Furthermore, EZH2, the catalytic subunit of the Polycomb repressive complex 2 (PRC2) and ID causal protein (MIM#277590), showed changes in phosphorylation accompanied by a change in the repressive H3K27me3 histone modification. While this study used chemical inhibitors to raise O-GlcNAc levels, it is clear that perturbing the O-GlcNAc homeostasis during neurogenesis causes dysregulation of gene expression at least in part by changes in histone methylation (58). The role of OGT in histone methylation is a pertinent question not only for understanding its role in transcription regulation but also for understanding the mechanisms responsible for the OGT-CDG phenotype. When investigating causal OGT-CDG mutants, studies should consider the methylation of histones in relevant models given that OGT interacts with more than seven enzymes responsible for proper histone methylation during development.

Histone acetylation

Further supporting the investigation of epigenetic modifications in the OGT-CDG phenotype is the enrichment of OGT interactors involved in histone acetylation. CBP, EP300, TRRAP, MCM3AP, TBLRXR1, and SIN3A all play a role in histone acetylation or deacetylation either enzymatically or as scaffolds for HDAC-containing complexes (Fig. 3). CBP and EP300 are both histone acetyltransferase as well as modifiers of other proteins (124, 125, 126, 127, 128, 129, 130), and both are causal for neurodevelopmental phenotypes with overlapping presentations (Table 1). Their links to neurodevelopment and neuronal function have been well studied (131, 132, 133, 134, 135, 136, 137, 138, 139, 140, 141), and a possible link to OGT-CDG has been discussed previously (61). Some studies have shown that the O-GlcNAc cycling enzymes cooperate with CBP/EP300 acetylation of substrates (124, 125, 142), suggesting coregulation of transcription between CBP, EP300, and OGT. Both TRRAP and MCM3AP function as scaffolds for histone acetyltransferase complexes such as SAGA, Tip60, and TREX-2 complexes (143, 144, 145, 146, 147, 148, 149, 150, 151, 152, 153, 154, 155, 156, 157, 158). Studies have shown that TRRAP is necessary for normal development including for neural lineage cells (143, 159, 160, 161, 162). However, the role of MCM3AP in neurodevelopment and the functioning of the adult nervous system is less well understood (163). OGT could play a role in stability in the complexes containing TRRAP and MCM3AP either through modification of these scaffolds or by non-catalytic functions (Fig. 1A), but further investigation is required to uncover if either of these is likely. SIN3A and TBLRXR1 both play a scaffolding role as well, but these proteins are involved in the deacetylation of histones and thus transcription repression (164, 165, 166, 167, 168, 169, 170, 171, 172, 173, 174). TBLRXR1 is a part of the NCOR/SMRT complex, which is responsible for gene repression of neuronal genes. TBL1XR1 is causal for a neurodevelopmental phenotype (175, 176), but recent reports also suggest that NCOR1, NCOR2 (SMRT), and HDAC3 are also responsible for an ID phenotype (177), providing a basis for investigation of the relevance of the complex and its members in the OGT-CDG phenotype. The last interactor in this section, SIN3A, has known roles in neuronal function and neurodevelopment (167, 178, 179, 180, 181, 182, 183) and has also been shown to partner with OGT in vivo to regulate gene expression (164, 184). Moreover, SIN3A and TBLRXR1 also tie in another predicted putative conveyor of the OGT-CDG phenotype, MECP2 (61). SIN3A and TBLRXR1 both have links to MECP2 function and Rett syndrome (RTT) in vivo (185, 186, 187, 188, 189). SIN3A and HDAC1/2 levels are altered along with histone acetylation in the RTT mouse cortex, and RTT causal mutations disrupt the binding between TBLRXR1 and MECP2. If OGT-CDG mutations in the TPR domain cause disruptions in normal protein-protein interactions, it is worth investigating how the phenotype could be conveyed through all of these proteins, especially in light of a recent report suggesting that RTT etiological site T203 is a functional O-GlcNAc site with disruptions in normal neuronal function and morphology in vivo (190). Future studies with the aim of site mapping O-GlcNAc sites would determine what sites might differ between wildtype OGT and the OGT-CDG variants in relevant model systems. The enrichment of OGT’s interactome for proteins involved in histone acetylation suggests examining the acetylation status of histones as another potential mechanism by which OGT-CDG phenotype is conveyed.

Histone ubiquitination

The final class of histone modifiers identified in the OGT interactome includes ASXL2, BAP1, and HCFC1 (discussed later) of the PR-DUB complex (Fig. 3) (191). Given that OGT regulates polycomb repression in Drosophila (10), it is possible that the PR-DUB complex is essential to that regulation and possibly to the OGT-CDG mechanism in humans. The PR-DUB complex is responsible for removing ubiquitin added to H2A by the Polycomb Repressive Complex 1 (PRC1) (192, 193). ASXL2 has previously been implicated in the neurodevelopmental phenotype (Table 1), but BAP1 has more recently been found causal for the phenotype (Table 1, bold). The PR-DUB complex and PRC1 regulate gene expression with both complexes being linked to neurodevelopmental phenotypes (194). BAP1 is known to regulate both OGT and HCFC1 levels by catalyzing their deubiquitination resulting in decreased proteasomal degradation (191). OGT might also regulate levels or the integrity of the PR-DUB complex by stabilizing the complex or by targeting the adaptor protein ASXL2, which can be replaced in the complex by paralogs ASXL1 (MIM#:605036) and ASXL3 (MIM#:615485) both also linked to ID. Of note, OGT-CDG causal variants in the TPR domain all exhibited altered LXR pathway gene expression, and ASXL2 has been shown to increase LXRα activity (32, 195). Moreover, a recent study has shown that BAP1 may play a role as an adaptor protein for OGT (65). Knockout of BAP1 resulted in a small change in global O-GlcNAcylation but had major impacts on the O-GlcNAcylation of ASXL3, HCFC1, and ANKRD17 with HCFC1 having clustered increases and decreases. Lastly, a decrease in global O-GlcNAc levels in Drosophila in early development results in redeployment of the PRC, an increase in the H3K27me3 repressive histone modification at genes involved in nervous system development, and a reduction in brain size (196). While this role of O-GlcNAc needs to be confirmed in human systems, this work further supports the notion that OGT and O-GlcNAc are vital for early embryonic development via the regulation of histone modifications. In particular histone methylation, acetylation, and ubiquitination could all possibly link OGT-CDG causal variants to the phenotype seen in patients, and further investigation into these histone modifications could elucidate the common perturbations in gene expression seen in the TPR domain variants.

Nucleosome remodeling

OGT’s interaction with the BAF (mSWI/SNF) complex members SMARCA4, SMARCC2, ARID1A, and ARID1B expands the list of known interactors of OGT (60), and they provide mechanistic targets to understand mechanisms causal in OGT-CDG. The BAF complex, including interactor POGZ, is responsible for remodeling nucleosomes and regulating transcription (Fig. 3) (197, 198, 199, 200, 201, 202, 203, 204, 205, 206). The BAF complex containing SMARCA4 is known to co-regulate nucleosome remodeling (207) as well as the Nucleosome Remodeling Factor (NURF) complex member BPTF (208, 209, 210, 211, 212). POGZ has been studied in relevant models as a high-risk gene for autism, and the altered function of POGZ could explain behavioral problems seen in OGT-CDG patients (197, 213, 214, 215, 216, 217, 218, 219, 220). Furthermore, POGZ is known to co-occupy and coregulate relevant gene expression with ADNP, which was previously identified as a potential conveyor of the OGT-CDG phenotype in a bioinformatic screen (61, 221, 222). Composition and integrity of the BAF complex are essential for early development with the switching of BAF subunits giving rise to embryonic (esBAF), neuroprogenitor (npBAF), and neural (nBAF) specific complexes (Fig. 1B) during development (199, 200, 201, 202, 205, 223, 224, 225, 226, 227, 228, 229, 230, 231, 232, 233, 234, 235, 236, 237). Mutually exclusive subunits ARID1A and ARID1B differ greatly in their number of mapped O-GlcNAc sites (22 versus 1, respectively). Furthermore, ACTL6A was also identified as an interactor and has more recently been reported to be linked to a neurodevelopmental phenotype (238, 239). Given that ACTL6A is also swapped for ACTL6B (MIM#:618470) during similar developmental time periods as the swap from ARID1A to ARID1B, it is worth investigating any role OGT may play in the composition and integrity of the BAF complex in this neural context. The OGT interactor BPTF is also knocn to regulate neurodevelopment, and its interaction with OGT further highlights the importance of studying nucleosome remodeling in the context of O-GlcNAc biology (208, 209, 210, 240). Given that all OGT-CDG variants in the TPR domain showed altered gene expression and OGT interacts with multiple proteins involved in nucleosome remodeling, it is possible that OGT regulates chromatin accessibility. Using techniques like ATAC-seq to probe chromatin accessibility could provide evidence for the involvement of BAF or other chromatin remodelers in the OGT-CDG phenotype and would elucidate the role OGT plays in chromatin accessibility (32, 241).

Transcription factors

Beyond epigenetic regulation, the OGT interactome contains several transcription factors, and it has been well established that O-GlcNAc modification of transcription factors can regulate their function (Fig. 3) (26, 242). TAF6 is a part of the TFIID complex required for basal transcription (243, 244, 245, 246, 247) and has links to the SAGA complex mentioned in the section Histone Acetylation (247, 248, 249). The TFIID complex may represent a neurodevelopmentally important node in transcription as TAF1 (MIM#:300966), TAF2 (MIM#:615599), and TAF13 (MIM#:617432) have already been found causal for neurodevelopmental disorders. The OGT interactor TAF4 (60) has recently been associated with a neurodevelopmental disorder expanding the list of OGT interactors and TFIID members causal for ID (250). CIC, the mammalian homolog of Capicua, is a repressor involved in the MAPK signaling pathway (251, 252, 253, 254, 255, 256, 257, 258, 259, 260, 261, 262, 263, 264). The regulation of MAPK signaling has also been previously suggested as a potential effector of the OGT-CDG phenotype (61). Not only is CIC involved in neurodevelopment and function (265, 266, 267, 268, 269, 270), but it also recruits the SIN3A/HDAC and BAF complexes discussed earlier for repression of gene expression (258, 268). The interplay between interactors CIC and SIN3A is one of many known interactions between the proteins discussed herein (Fig. 4) and highlights the complexity and integrated nature of ID causal proteins that interact with OGT. SPEN (MINT/SHARP) is another OGT interactor involved in transcriptional regulation, and it plays a role in the inactivation of the X chromosome (Fig. 3) (271, 272, 273, 274, 275, 276), Notch signaling (277, 278, 279, 280), and Wnt/β-catenin signaling (281, 282). Furthermore, OGT has been shown to regulate neurogenesis in part through Wnt/β-catenin and Notch signaling (283, 284). With 65 O-GlcNAc sites on SPEN (Fig. 2), this interactor should be investigated within the context of O-GlcNAc biology and OGT-CDG as it represents a transcriptional node connecting different signaling pathways important for early development. In particular, O-GlcNAc sites localized to the nuclear receptor interaction domain of SPEN are likely relevant as these sites may regulate protein-protein interactions (65). The OGT interactor and transcription regulator HCFC1 has been hypothesized to play a role in OGT-CDG previously (61, 98, 285). Of the interactors discussed in this review, HCFC1 by far has the most mapped O-GlcNAc sites (Fig. 2) and is the most well-studied in relation to OGT, which suggests that it plays an important role in O-GlcNAc biology. HCFC1 mutations cause ID, but only mutations in the Kelch domain, which lacks O-GlcNAc sites, result in alterations in cobalamin metabolism (286). Pathogenic HCFC1 variants HCFC1A864T (287), HCFC1G876S (288), and HCFC1A897V (289) are proximal to mapped O-GlcNAc sites, suggesting that these sites, and potentially others, are relevant to the neurodevelopmental phenotype seen in OGT-CDG patients. Furthermore, HCFC1 is cleaved by OGT (13, 14, 15, 16, 290, 291), but this function is separable from the glycosylation of HCFC1 (292). Despite the evidence suggesting that HCFC1 glycosylation or cleavage could be linked to the OGT-CDG phenotype, only OGTN567K of the catalytic domain variants shows alterations in glycosylation or cleavage of an HCFC1 repeat substrate in vivo (34). However, HCFC1 glycosylation and cleavage by TPR domain variants have not been tested in cells, where protein-protein interactions could affect their function. OGT was found to interact with TET2/3, and through this interaction, TET2/3 promotes the glycosylation of HCFC1 by OGT (293). Furthermore, the glycosylation of HCFC1 favors the integrity of the SET1/COMPASS complex containing SETD1A discussed earlier. Given that TET2 was also identified in the OGT interactome (60), it is possible that cell-type specific interactions with OGT could explain the OGT-CDG phenotype via interactors not yet linked to ID. Overall, transcriptional regulators represent important targets of investigation and further add to the evidence that the OGT-CDG mechanism lies in the proper spatiotemporal regulation of gene expression (Fig. 1B).

mRNA processing, export, and silencing

OGT interactors implicated in mRNA processing, export, and silencing further suggest that the OGT-CDG mechanism may involve regulation of gene expression (Fig. 3). With nearly 30 mapped O-GlcNAc sites, ZC3H14 provides a novel target to investigate the role of OGT in mRNA processing, in particular sites S211, S220, and S369 which were found to be enriched in mESCs over that of RA-differentiated mESCs (66). ZC3H14 regulates mRNA processing by controlling poly(A) tail length (294, 295, 296, 297, 298), and its function is important for normal neuronal activity and development (299, 300, 301, 302, 303, 304, 305, 306). NUP62 has been previously identified as a potential target relevant to the OGT-CDG phenotype (285). It is central to the integrity of the Nuclear Pore Complex (NPC), and thus NUP62 plays a role in transport between the nucleus and cytoplasm (307, 308, 309, 310). Without the O-GlcNAc modification on NUP62, it is degraded via the proteasome pathway (311). Therefore, it is possible that OGT-CDG patients have lower levels of NPC, resulting in aberrant transport between the nucleus and cytoplasm. Lastly, TNRC6B represents another protein through which OGT could regulate the expression of genes and another link to neurodevelopmental disorders. TNRC6B is a scaffold for the RNA-induced Silencing Complex (RISC) (312, 313, 314, 315, 316, 317, 318, 319), which silences mRNA transcripts post-transcriptionally and has been implicated in normal neuronal function (320, 321). This complex contains other members linked to neurodevelopmental phenotypes, AGO1 (MIM#:620292) and AGO2 (MIM#:619149), and studies could determine what role OGT plays in the regulation of this complex via its interaction and modification of TNRC6B. Taken together, OGT interactors ZC3H14, NUP62, and TNRC6B implicate post-transcriptional regulation of gene expression as a critical new area within O-GlcNAc biology and as a possible mechanism by which OGT-CDG is conveyed, namely the alterations in gene expression seen.

Other mechanistic targets

In this last section, we will discuss both ID-linked interactors that are currently understudied as well as some other potential targets that might be linked to the OGT-CDG mechanism. ANKRD17, PRR12, and AHDC1 represent understudied proteins in the OGT interactome and provide novel targets for O-GlcNAc research. PRR12 has a role in the microRNA-mediated regulation of gliogenesis (322) and has dynamic expression in the developing brain (323). Additionally, placental reduction of OGT in mice results in altered brain microRNA expression (324), but the mechanisms behind this regulation still need to be determined. ANKRD17 has 40 mapped O-GlcNAc sites (Fig. 2) and may regulate YAP1 and the Hippo pathway (325, 326). OGT has been shown to modulate the Hippo pathway (327, 328), but the role ANKRD17 plays in this modulation has yet to be elucidated. Finally, AHDC1 may play a role in DNA methylation during embryogenesis (329). OGT is known to interact with other proteins involved in DNA methylation and demethylation, including the TET family proteins, so studies investigating the interaction between AHDC1 and OGT would further our understanding of OGT’s role in DNA methylation (330). Studies involving ANKRD17, PRR12, and AHDC1 should consider the role that OGT plays in regulating them to better understand their roles in the cell as well as any possible connections to the OGT-CDG phenotype.

This review has focused on those interactors with cataloged ID phenotypes identified using a TPR-BioID approach (60) (Table 1), but as the list has expanded it is worth noting that the OGT-CDG mechanism may involve interactors or substrates not linked to ID yet or discussed in this review. For example, TET2 was discussed in a previous section as it was related to the modulation of HCFC1 O-GlcNAcylation (293). Beyond being identified as an OGT interactor (60), TET2 has been shown to interact with OGT in mESCs (331), has O-GlcNAc sites enriched in mESCs compared to differentiated mESCs (66), and its modification by OGT competes with known phosphorylation sites (332). While TET2 has not been linked to an ID phenotype, studies have shown that TET2 plays a role in neuronal differentiation (333, 334, 335). Quantification of methylated and hydroxymethylated DNA would provide insight into any role of TET2 in the OGT-CDG mechanism (336). Furthermore, numerous heterogeneous nuclear ribonulceoproteins (HNRNP) were identified as OGT interactors in another study (65), and this particular family of proteins has been identified as a hub for shared neurodevelopmental disorders (337). This family of proteins is responsible for multiple aspects of nucleic acid metabolism including alternative splicing and transcriptional regulation (338), which were discussed in this review as possible mechanisms responsible for OGT-CDG. While they were not discussed in detail, they further highlight the need for the field to investigate the role of O-GlcNAc biology in the regulation of gene expression especially as it relates to possible OGT-CDG mechanisms. While other avenues of investigation will be necessary to fully understand the OGT-CDG phenotype, the large number of OGT interactors and substrates discussed herein highlights the need to probe the role of transcription regulation in the OGT-CDG phenotype.

Future directions

Since the first characterizations of OGT-CDG causal mutations (30, 31), the list of OGT variants has expanded (Table 1) and continues to expand as clinicians test for the genetic causes of ID in patients. In this review, we have discussed the variable biochemical characteristics of the OGT-CDG variants, with defects in glycosylation representing the common alteration though to different degrees depending on the domain. The variable nature of the patient phenotype has been highlighted, and we discussed how animal and cell models recapitulate this variability. We have also provided evidence that, in part, the OGT-CDG mechanism lies in the dysregulation of gene expression as shown by models of TPR domain variants in hESCs and a catalytic domain variant in mESCs (32, 34). This common downstream impact of the variants led to the evaluation of the OGT interactome, which is enriched for gene expression regulators causal for ID phenotypes. Beyond investigating these interactors for links to the OGT-CDG mechanisms, there are other pertinent questions that remain. Could supplementation with glucosamine represent a therapeutic for patients as previously suggested (98)? If models can show that glucosamine supplementation can alleviate the observed effects of OGT-CDG variants, then the next step would be for clinicians to see if supplementation could alleviate some symptoms as has been tested for other CDGs (339). TPR domain variants should also be tested for any alterations in their interactomes in relevant cell types to further narrow the list of likely mechanistic targets. Additionally, the gene expression changes seen in patient lymphoblastoids has not been validated for other variants, but if it is seen in other patients, RNA-seq may represent a diagnostic tool for clinicians to rapidly identify OGT-CDG patients if a set of commonly altered genes is identified. Moreover, models have shown that OGT-CDG variants attempt to maintain O-GlcNAc homeostasis and OGA is essential for normal embryogenesis (340), which elicits the question of whether mutations in OGA might also be causal for an ID phenotype yet to be discovered. Finally, the studies to date show varied effects in models. For example, OGTC921Y shows defects in maintaining the pluripotency network, but other variants do not. Do these varied effects represent differences in study design or do they suggest that OGT-CDG variants may have different mechanistic targets that are affected? Future studies would benefit from the use of multiple OGT-CDG variants to uncover shared or unique affected targets to better address this question. While many questions remain for us to fully understand OGT-CDG, many have been answered since the first characterization of variants, and studies utilizing OGT-CDG variants give us further insight into OGT and O-GlcNAc biology.

Conflict of interest

The authors declare that they have no conflicts of interest with the contents of this article.

Acknowledgments

We would like to thank all of the authors whose work has been cited here. Although we have attempted to be as thorough as possible in our review in the literature, with thousands of papers published in the O-GlcNAc field as well as on the interactors discussed, we may have missed a few relevant manuscripts. Any exclusion is unintentional.

Author contributions

J. M. M., N. L. H., I. C., L. V., E. M., D. M. F. A., and L. W. conceptualization, J. M. M., N. L. H., I. C., and L. V. data curation, J. M. M. and N. L. H. writing–original draft, J. M. M., N. L. H. visualization; I. C., L. V., L. H., E. M., D. M. F. A., and L. W. writing-review & editing; E. M., D. M. F. A., and L. W. supervision; E. M., D. M. F. A., and L. W. funding acquisition.

Funding and additional information

This work was supported in part by grants from NIH/NICHD (R21HD097562 and R01HD110099 to L. W. and F31HD108843 to J. M.), NIH/NINDS (1U54NS115198-04 to E. M.), the National Center for Advancing Translational Sciences (NCATS), and the Rare Disorders Consortium Research Network (RDCRN) (E. M., L. V.). DvA is funded by a Novo Nordisk Fonden Laureate award (NNF21OC0065969) and a Villum Fonden Investigator grant (VIL54496). L. W. is supported by the Georgia Research Alliance. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Reviewed by members of the JBC Editorial Board. Edited by Robert Haltiwanger

Contributor Information

Daan M.F. van Aalten, Email: daan@mbg.au.dk.

Lance Wells, Email: lwells@ccrc.uga.edu.

References

  • 1.Torres C.R., Hart G.W. Topography and polypeptide distribution of terminal N-acetylglucosamine residues on the surfaces of intact lymphocytes. Evidence for O-linked GlcNAc. J. Biol. Chem. 1984;259:3308–3317. [PubMed] [Google Scholar]
  • 2.Haltiwanger R.S., Blomberg M.A., Hart G.W. Glycosylation of nuclear and cytoplasmic proteins. Purification and characterization of a uridine diphospho-N-acetylglucosamine:polypeptide beta-N-acetylglucosaminyltransferase. J. Biol. Chem. 1992;267:9005–9013. [PubMed] [Google Scholar]
  • 3.Kreppel L.K., Blomberg M.A., Hart G.W. Dynamic glycosylation of nuclear and cytosolic proteins. Cloning and characterization of a unique O-GlcNAc transferase with multiple tetratricopeptide repeats. J. Biol. Chem. 1997;272:9308–9315. doi: 10.1074/jbc.272.14.9308. [DOI] [PubMed] [Google Scholar]
  • 4.Lubas W.A., Frank D.W., Krause M., Hanover J.A. O-Linked GlcNAc transferase is a conserved nucleocytoplasmic protein containing tetratricopeptide repeats. J. Biol. Chem. 1997;272:9316–9324. doi: 10.1074/jbc.272.14.9316. [DOI] [PubMed] [Google Scholar]
  • 5.Wulff-Fuentes E., Berendt R.R., Massman L., Danner L., Malard F., Vora J., et al. The human O-GlcNAcome database and meta-analysis. Sci. Data. 2021;8:25. doi: 10.1038/s41597-021-00810-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Gao Y., Wells L., Comer F.I., Parker G.J., Hart G.W. Dynamic O-glycosylation of nuclear and cytosolic proteins: cloning and characterization of a neutral, cytosolic beta-N-acetylglucosaminidase from human brain. J. Biol. Chem. 2001;276:9838–9845. doi: 10.1074/jbc.M010420200. [DOI] [PubMed] [Google Scholar]
  • 7.Shafi R., Iyer S.P., Ellies L.G., O’Donnell N., Marek K.W., Chui D., et al. The O-GlcNAc transferase gene resides on the X chromosome and is essential for embryonic stem cell viability and mouse ontogeny. Proc. Natl. Acad. Sci. U. S. A. 2000;97:5735–5739. doi: 10.1073/pnas.100471497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Levine Z.G., Potter S.C., Joiner C.M., Fei G.Q., Nabet B., Sonnett M., et al. Mammalian cell proliferation requires noncatalytic functions of O-GlcNAc transferase. Proc. Natl. Acad. Sci. U. S. A. 2021;118 doi: 10.1073/pnas.2016778118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Karczewski K.J., Francioli L.C., Tiao G., Cummings B.B., Alföldi J., Wang Q., et al. The mutational constraint spectrum quantified from variation in 141,456 humans. Nature. 2020;581:434–443. doi: 10.1038/s41586-020-2308-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Gambetta M.C., Oktaba K., Müller J. Essential role of the glycosyltransferase sxc/Ogt in polycomb repression. Science. 2009;325:93–96. doi: 10.1126/science.1169727. [DOI] [PubMed] [Google Scholar]
  • 11.Levine Z.G., Walker S. The biochemistry of O-GlcNAc transferase: which functions make it essential in mammalian cells? Annu. Rev. Biochem. 2016;85:631–657. doi: 10.1146/annurev-biochem-060713-035344. [DOI] [PubMed] [Google Scholar]
  • 12.Li X., Yue X., Sepulveda H., Burt R.A., Scott D.A., A Carr S., et al. OGT controls mammalian cell viability by regulating the proteasome/mTOR/mitochondrial axis. Proc. Natl. Acad. Sci. U. S. A. 2023;120 doi: 10.1073/pnas.2218332120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Capotosti F., Guernier S., Lammers F., Waridel P., Cai Y., Jin J., et al. O-GlcNAc transferase catalyzes site-specific proteolysis of HCF-1. Cell. 2011;144:376–388. doi: 10.1016/j.cell.2010.12.030. [DOI] [PubMed] [Google Scholar]
  • 14.Daou S., Mashtalir N., Hammond-Martel I., Pak H., Yu H., Sui G., et al. Crosstalk between O-GlcNAcylation and proteolytic cleavage regulates the host cell factor-1 maturation pathway. Proc. Natl. Acad. Sci. U. S. A. 2011;108:2747–2752. doi: 10.1073/pnas.1013822108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Lazarus M.B., Jiang J., Kapuria V., Bhuiyan T., Janetzko J., Zandberg W.F., et al. HCF-1 is cleaved in the active site of O-GlcNAc transferase. Science. 2013;342:1235–1239. doi: 10.1126/science.1243990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Bhuiyan T., Waridel P., Kapuria V., Zoete V., Herr W. Distinct OGT-binding sites promote HCF-1 cleavage. PLoS One. 2015;10 doi: 10.1371/journal.pone.0136636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Capotosti F., Hsieh J.J.-D., Herr W. Species selectivity of mixed-lineage leukemia/trithorax and HCF proteolytic maturation pathways. Mol. Cell Biol. 2007;27:7063–7072. doi: 10.1128/MCB.00769-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Wysocka J., Herr W. The herpes simplex virus VP16-induced complex: the makings of a regulatory switch. Trends Biochem. Sci. 2003;28:294–304. doi: 10.1016/S0968-0004(03)00088-4. [DOI] [PubMed] [Google Scholar]
  • 19.Zargar Z., Tyagi S. Role of host cell factor-1 in cell cycle regulation. Transcription. 2012;3:187–192. doi: 10.4161/trns.20711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Levine Z.G., Fan C., Melicher M.S., Orman M., Benjamin T., Walker S. O-GlcNAc transferase recognizes protein substrates using an asparagine ladder in the tetratricopeptide repeat (TPR) superhelix. J. Am. Chem. Soc. 2018;140:3510–3513. doi: 10.1021/jacs.7b13546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Drula E., Garron M.-L., Dogan S., Lombard V., Henrissat B., Terrapon N. The carbohydrate-active enzyme database: functions and literature. Nucleic Acids Res. 2022;50:D571–D577. doi: 10.1093/nar/gkab1045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Yang X., Ongusaha P.P., Miles P.D., Havstad J.C., Zhang F., So W.V., et al. Phosphoinositide signalling links O-GlcNAc transferase to insulin resistance. Nature. 2008;451:964–969. doi: 10.1038/nature06668. [DOI] [PubMed] [Google Scholar]
  • 23.Lazarus M.B., Nam Y., Jiang J., Sliz P., Walker S. Structure of human O-GlcNAc transferase and its complex with a peptide substrate. Nature. 2011;469:564–567. doi: 10.1038/nature09638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Blankenship C.M., Xie J., Benz C., Wang A., Ivarsson Y., Jiang J. Motif-dependent binding on the intervening domain regulates O-GlcNAc transferase. Nat. Chem. Biol. 2023;19:1423–1431. doi: 10.1038/s41589-023-01422-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Alteen M.G., Meek R.W., Kolappan S., Busmann J.A., Cao J., O’Gara Z., et al. Phage display uncovers a sequence motif that drives polypeptide binding to a conserved regulatory exosite of O-GlcNAc transferase. Proc. Natl. Acad. Sci. U. S. A. 2023;120 doi: 10.1073/pnas.2303690120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Hart G.W. Nutrient regulation of signaling and transcription. J. Biol. Chem. 2019;294:2211–2231. doi: 10.1074/jbc.AW119.003226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Czajewski I., van Aalten D.M.F. The role of O-GlcNAcylation in development. Development. 2023;150 doi: 10.1242/dev.201370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Sheikh M.A., Emerald B.S., Ansari S.A. Stem cell fate determination through protein O-GlcNAcylation. J. Biol. Chem. 2021;296 doi: 10.1074/jbc.REV120.014915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Bouazzi H., Lesca G., Trujillo C., Alwasiyah M.K., Munnich A. Nonsyndromic X-linked intellectual deficiency in three brothers with a novel MED12 missense mutation [c.5922G>T (p.Glu1974His)] Clin. Case Rep. 2015;3:604–609. doi: 10.1002/ccr3.301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Willems A.P., Gundogdu M., Kempers M.J.E., Giltay J.C., Pfundt R., Elferink M., et al. Mutations in N-acetylglucosamine (O-GlcNAc) transferase in patients with X-linked intellectual disability. J. Biol. Chem. 2017;292:12621–12631. doi: 10.1074/jbc.M117.790097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Vaidyanathan K., Niranjan T., Selvan N., Teo C.F., May M., Patel S., et al. Identification and characterization of a missense mutation in the O-linked β-N-acetylglucosamine (O-GlcNAc) transferase gene that segregates with X-linked intellectual disability. J. Biol. Chem. 2017;292:8948–8963. doi: 10.1074/jbc.M116.771030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Selvan N., George S., Serajee F.J., Shaw M., Hobson L., Kalscheuer V., et al. O-GlcNAc transferase missense mutations linked to X-linked intellectual disability deregulate genes involved in cell fate determination and signaling. J. Biol. Chem. 2018;293:10810–10824. doi: 10.1074/jbc.RA118.002583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Pravata V.M., Gundogdu M., Bartual S.G., Ferenbach A.T., Stavridis M., Õunap K., et al. A missense mutation in the catalytic domain of O-GlcNAc transferase links perturbations in protein O-GlcNAcylation to X-linked intellectual disability. FEBS Lett. 2020;594:717–727. doi: 10.1002/1873-3468.13640. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Pravata V.M., Muha V., Gundogdu M., Ferenbach A.T., Kakade P.S., Vandadi V., et al. Catalytic deficiency of O-GlcNAc transferase leads to X-linked intellectual disability. Proc. Natl. Acad. Sci. U. S. A. 2019;116:14961–14970. doi: 10.1073/pnas.1900065116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Omelková M., Fenger C.D., Murray M., Hammer T.B., Pravata V.M., Bartual S.G., et al. An O-GlcNAc transferase pathogenic variant linked to intellectual disability affects pluripotent stem cell self-renewal. Dis. Model Mech. 2023;16 doi: 10.1242/dmm.049132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Plenge R.M., Stevenson R.A., Lubs H.A., Schwartz C.E., Willard H.F. Skewed X-chromosome inactivation is a common feature of X-linked mental retardation disorders. Am. J. Hum. Genet. 2002;71:168–173. doi: 10.1086/341123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Olivier-Van Stichelen S., Hanover J.A. X-inactivation normalizes O-GlcNAc transferase levels and generates an O-GlcNAc-depleted Barr body. Front. Genet. 2014;5:256. doi: 10.3389/fgene.2014.00256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Adzhubei I.A., Schmidt S., Peshkin L., Ramensky V.E., Gerasimova A., Bork P., et al. A method and server for predicting damaging missense mutations. Nat. Methods. 2010;7:248–249. doi: 10.1038/nmeth0410-248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Ng P.C., Henikoff S. SIFT: predicting amino acid changes that affect protein function. Nucleic Acids Res. 2003;31:3812–3814. doi: 10.1093/nar/gkg509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Gundogdu M., Llabrés S., Gorelik A., Ferenbach A.T., Zachariae U., van Aalten D.M.F. The O-GlcNAc transferase intellectual disability mutation L254F distorts the TPR helix. Cell Chem. Biol. 2018;25:513–518.e4. doi: 10.1016/j.chembiol.2018.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Llabrés S., Tsenkov M.I., MacGowan S.A., Barton G.J., Zachariae U. Disease related single point mutations alter the global dynamics of a tetratricopeptide (TPR) α-solenoid domain. J. Struct. Biol. 2020;209 doi: 10.1016/j.jsb.2019.107405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Tarrant M.K., Rho H.-S., Xie Z., Jiang Y.L., Gross C., Culhane J.C., et al. Regulation of CK2 by phosphorylation and O-GlcNAcylation revealed by semisynthesis. Nat. Chem. Biol. 2012;8:262–269. doi: 10.1038/nchembio.771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Pathak S., Borodkin V.S., Albarbarawi O., Campbell D.G., Ibrahim A., van Aalten D.M. O-GlcNAcylation of TAB1 modulates TAK1-mediated cytokine release. EMBO J. 2012;31:1394–1404. doi: 10.1038/emboj.2012.8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Murray M., Davidson L., Ferenbach A.T., Lefeber D., van Aalten D.M.F. Neuroectoderm phenotypes in a human stem cell model of O-GlcNAc transferase intellectual disability. bioRxiv. 2023 doi: 10.1101/2023.09.18.558285. [preprint] [DOI] [PubMed] [Google Scholar]
  • 45.Fedorova V., Vanova T., Elrefae L., Pospisil J., Petrasova M., Kolajova V., et al. Differentiation of neural rosettes from human pluripotent stem cells in vitro is sequentially regulated on a molecular level and accomplished by the mechanism reminiscent of secondary neurulation. Stem Cell Res. 2019;40 doi: 10.1016/j.scr.2019.101563. [DOI] [PubMed] [Google Scholar]
  • 46.Schnerch A., Cerdan C., Bhatia M. Distinguishing between mouse and human pluripotent stem cell regulation: the best laid plans of mice and men. Stem Cells. 2010;28:419–430. doi: 10.1002/stem.298. [DOI] [PubMed] [Google Scholar]
  • 47.Tsetsenis T., Badyna J.K., Wilson J.A., Zhang X., Krizman E.N., Subramaniyan M., et al. Midbrain dopaminergic innervation of the hippocampus is sufficient to modulate formation of aversive memories. Proc. Natl. Acad. Sci. U. S. A. 2021;118 doi: 10.1073/pnas.2111069118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Courtney R., Landreth G.E. LXR regulation of brain cholesterol: from development to disease. Trends Endocrinol. Metab. 2016;27:404–414. doi: 10.1016/j.tem.2016.03.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Fenckova M., Muha V., Mariappa D., Catinozzi M., Czajewski I., Blok L.E.R., et al. Intellectual disability-associated disruption of O-GlcNAc cycling impairs habituation learning in Drosophila. PLoS Genet. 2022;18 doi: 10.1371/journal.pgen.1010159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Authier F., Ondruskova N., Ferenbach A.T., McNeilly A., van Aalten D.M.F. Neurodevelopmental defects in a mouse model of O-GlcNAc transferase intellectual disability. bioRxiv. 2023 doi: 10.1101/2023.08.23.554427. [preprint] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Czajewski I., McDowall L., Ferenbach A., van Aalten D.M.F. Rescuable sleep and synaptogenesis phenotypes in a Drosophila model of O-GlcNAc transferase intellectual disability. bioRxiv. 2023 doi: 10.1101/2023.06.28.546900. [preprint] [DOI] [Google Scholar]
  • 52.Webster D.M., Teo C.F., Sun Y., Wloga D., Gay S., Klonowski K.D., et al. O-GlcNAc modifications regulate cell survival and epiboly during zebrafish development. BMC Dev. Biol. 2009;9:28. doi: 10.1186/1471-213X-9-28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Cui Y., Cruz M., Palatnik A., Olivier-Van Stichelen S. O-GlcNAc transferase contributes to sex-specific placental deregulation in gestational diabetes. Placenta. 2023;131:1–12. doi: 10.1016/j.placenta.2022.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kim G., Cao L., Reece E.A., Zhao Z. Impact of protein O-GlcNAcylation on neural tube malformation in diabetic embryopathy. Sci. Rep. 2017;7 doi: 10.1038/s41598-017-11655-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Liu Y., Li X., Yu Y., Shi J., Liang Z., Run X., et al. Developmental regulation of protein O-GlcNAcylation, O-GlcNAc transferase, and O-GlcNAcase in mammalian brain. PLoS One. 2012;7 doi: 10.1371/journal.pone.0043724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Andres L.M., Blong I.W., Evans A.C., Rumachik N.G., Yamaguchi T., Pham N.D., et al. Chemical modulation of protein O-GlcNAcylation via OGT inhibition promotes human neural cell differentiation. ACS Chem. Biol. 2017;12:2030–2039. doi: 10.1021/acschembio.7b00232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Parween S., Varghese D.S., Ardah M.T., Prabakaran A.D., Mensah-Brown E., Emerald B.S., et al. Higher O-GlcNAc levels are associated with defects in progenitor proliferation and Premature neuronal differentiation during in-vitro human embryonic cortical neurogenesis. Front. Cell Neurosci. 2017;11:415. doi: 10.3389/fncel.2017.00415. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Parween S., Alawathugoda T.T., Prabakaran A.D., Dheen S.T., Morse R.H., Emerald B.S., et al. Nutrient sensitive protein O-GlcNAcylation modulates the transcriptome through epigenetic mechanisms during embryonic neurogenesis. Life Sci. Allian. 2022;5 doi: 10.26508/lsa.202201385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Liu X., Song S., Chen Z., Gao C., Li Y., Luo Y., et al. Release of O-GlcNAc transferase inhibitor promotes neuronal differentiation of neural stem cells in 3D bioprinted supramolecular hydrogel scaffold for spinal cord injury repair. Acta Biomater. 2022;151:148–162. doi: 10.1016/j.actbio.2022.08.031. [DOI] [PubMed] [Google Scholar]
  • 60.Stephen H.M., Praissman J.L., Wells L. Generation of an interactome for the tetratricopeptide repeat domain of O-GlcNAc transferase Indicates a role for the enzyme in intellectual disability. J. Proteome Res. 2020;20:1229–1242. doi: 10.1021/acs.jproteome.0c00604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Mitchell C.W., Czajewski I., van Aalten D.M.F. Bioinformatic prediction of putative conveyers of O-GlcNAc transferase intellectual disability. J. Biol. Chem. 2022;298 doi: 10.1016/j.jbc.2022.102276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Gao J., Yang Y., Qiu R., Zhang K., Teng X., Liu R., et al. Proteomic analysis of the OGT interactome: novel links to epithelial-mesenchymal transition and metastasis of cervical cancer. Carcinogenesis. 2018;39:1222–1234. doi: 10.1093/carcin/bgy097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Martinez M., Renuse S., Kreimer S., O’Meally R., Natov P., Madugundu A.K., et al. Quantitative proteomics reveals that the OGT interactome is remodeled in response to oxidative stress. Mol. Cell Proteomics. 2021;20 doi: 10.1016/j.mcpro.2021.100069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Ma J., Hou C., Li Y., Chen S., Wu C. OGT protein interaction network (OGT-PIN): a curated database of experimentally identified interaction proteins of OGT. Int. J. Mol. Sci. 2021;22:9620. doi: 10.3390/ijms22179620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Griffin M.E., Thompson J.W., Xiao Y., Sweredoski M.J., Aksenfeld R.B., Jensen E.H., et al. Functional glycoproteomics by integrated network assembly and partitioning. bioRxiv. 2023 doi: 10.1101/2023.06.13.541482. [preprint] [DOI] [Google Scholar]
  • 66.Hao Y., Li X., Qin K., Shi Y., He Y., Zhang C., et al. Chemoproteomic and transcriptomic analysis reveals that O-GlcNAc regulates mouse embryonic stem cell fate through the pluripotency network. Angew. Chem. Int. Ed. Engl. 2023;62 doi: 10.1002/anie.202300500. [DOI] [PubMed] [Google Scholar]
  • 67.Gabriele M., Lopez Tobon A., D’Agostino G., Testa G. The chromatin basis of neurodevelopmental disorders: rethinking dysfunction along the molecular and temporal axes. Prog. Neuropsychopharmacol. Biol. Psychiatry. 2018;84:306–327. doi: 10.1016/j.pnpbp.2017.12.013. [DOI] [PubMed] [Google Scholar]
  • 68.Vallianatos C.N., Iwase S. Disrupted intricacy of histone H3K4 methylation in neurodevelopmental disorders. Epigenomics. 2015;7:503–519. doi: 10.2217/epi.15.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Narayanan A., Ruyechan W.T., Kristie T.M. The coactivator host cell factor-1 mediates Set1 and MLL1 H3K4 trimethylation at herpesvirus immediate early promoters for initiation of infection. Proc. Natl. Acad. Sci. U. S. A. 2007;104:10835–10840. doi: 10.1073/pnas.0704351104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Lee Y.-T., Ayoub A., Park S.-H., Sha L., Xu J., Mao F., et al. Mechanism for DPY30 and ASH2L intrinsically disordered regions to modulate the MLL/SET1 activity on chromatin. Nat. Commun. 2021;12:2953. doi: 10.1038/s41467-021-23268-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Jakovcevski M., Ruan H., Shen E.Y., Dincer A., Javidfar B., Ma Q., et al. Neuronal Kmt2a/Mll1 histone methyltransferase is essential for prefrontal synaptic plasticity and working memory. J. Neurosci. 2015;35:5097–5108. doi: 10.1523/JNEUROSCI.3004-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Lee J.-E., Wang C., Xu S., Cho Y.-W., Wang L., Feng X., et al. H3K4 mono- and di-methyltransferase MLL4 is required for enhancer activation during cell differentiation. Elife. 2013;2 doi: 10.7554/eLife.01503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Vlasevska S., Garcia-Ibanez L., Duval R., Holmes A.B., Jahan R., Cai B., et al. KMT2D acetylation by CREBBP reveals a cooperative functional interaction at enhancers in normal and malignant germinal center B cells. Proc. Natl. Acad. Sci. U. S. A. 2023;120 doi: 10.1073/pnas.2218330120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Chai H., Pan C., Zhang M., Huo H., Shan H., Wu J. Histone methyltransferase SETD1A interacts with notch and promotes notch transactivation to augment ovarian cancer development. BMC Cancer. 2023;23:96. doi: 10.1186/s12885-023-10573-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Imran A., Moyer B.S., Canning A.J., Kalina D., Duncan T.M., Moody K.J., et al. Kinetics of the multitasking high-affinity Win binding site of WDR5 in restricted and unrestricted conditions. Biochem. J. 2021;478:2145–2161. doi: 10.1042/BCJ20210253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Shinsky S.A., Monteith K.E., Viggiano S., Cosgrove M.S. Biochemical reconstitution and phylogenetic comparison of human SET1 family core complexes involved in histone methylation. J. Biol. Chem. 2015;290:6361–6375. doi: 10.1074/jbc.M114.627646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Kim D.-H., Tang Z., Shimada M., Fierz B., Houck-Loomis B., Bar-Dagen M., et al. Histone H3K27 trimethylation inhibits H3 binding and function of SET1-like H3K4 methyltransferase complexes. Mol. Cell Biol. 2013;33:4936–4946. doi: 10.1128/MCB.00601-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.van Nuland R., Smits A.H., Pallaki P., Jansen P.W., Vermeulen M., Timmers H.T. Quantitative dissection and stoichiometry determination of the human SET1/MLL histone methyltransferase complexes. Mol. Cell Biol. 2013;33:2067–2077. doi: 10.1128/MCB.01742-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Dharmarajan V., Lee J.-H., Patel A., Skalnik D.G., Cosgrove M.S. Structural basis for WDR5 interaction (Win) motif recognition in human SET1 family histone methyltransferases. J. Biol. Chem. 2012;287:27275–27289. doi: 10.1074/jbc.M112.364125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Lim D.A., Huang Y.-C., Swigut T., Mirick A.L., Garcia-Verdugo J.M., Wysocka J., et al. Chromatin remodelling factor Mll1 is essential for neurogenesis from postnatal neural stem cells. Nature. 2009;458:529–533. doi: 10.1038/nature07726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Bharadwaj R., Peter C.J., Jiang Y., Roussos P., Vogel-Ciernia A., Shen E.Y., et al. Conserved higher-order chromatin regulates NMDA receptor gene expression and cognition. Neuron. 2014;84:997–1008. doi: 10.1016/j.neuron.2014.10.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Bjornsson H.T., Benjamin J.S., Zhang L., Weissman J., Gerber E.E., Chen Y.-C., et al. Histone deacetylase inhibition rescues structural and functional brain deficits in a mouse model of Kabuki syndrome. Sci. Transl. Med. 2014;6 doi: 10.1126/scitranslmed.3009278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Huang Y.-C., Shih H.-Y., Lin S.-J., Chiu C.-C., Ma T.-L., Yeh T.-H., et al. The epigenetic factor Kmt2a/Mll1 regulates neural progenitor proliferation and neuronal and glial differentiation. Dev. Neurobiol. 2015;75:452–462. doi: 10.1002/dneu.22235. [DOI] [PubMed] [Google Scholar]
  • 84.Van Laarhoven P.M., Neitzel L.R., Quintana A.M., Geiger E.A., Zackai E.H., Clouthier D.E., et al. Kabuki syndrome genes KMT2D and KDM6A: functional analyses demonstrate critical roles in craniofacial, heart and brain development. Hum. Mol. Genet. 2015;24:4443–4453. doi: 10.1093/hmg/ddv180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Tsai I.-C., McKnight K., McKinstry S.U., Maynard A.T., Tan P.L., Golzio C., et al. Small molecule inhibition of RAS/MAPK signaling ameliorates developmental pathologies of Kabuki Syndrome. Sci. Rep. 2018;8 doi: 10.1038/s41598-018-28709-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Carosso G.A., Boukas L., Augustin J.J., Nguyen H.N., Winer B.L., Cannon G.H., et al. Precocious neuronal differentiation and disrupted oxygen responses in Kabuki syndrome. JCI Insight. 2019;4 doi: 10.1172/jci.insight.129375. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Mukai J., Cannavò E., Crabtree G.W., Sun Z., Diamantopoulou A., Thakur P., et al. Recapitulation and reversal of schizophrenia-related phenotypes in Setd1a-deficient mice. Neuron. 2019;104:471–487.e12. doi: 10.1016/j.neuron.2019.09.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Yu X., Yang L., Li J., Li W., Li D., Wang R., et al. De novo and inherited SETD1A variants in early-onset epilepsy. Neurosci. Bull. 2019;35:1045–1057. doi: 10.1007/s12264-019-00400-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Schwenty-Lara J., Nehl D., Borchers A. The histone methyltransferase KMT2D, mutated in Kabuki syndrome patients, is required for neural crest cell formation and migration. Hum. Mol. Genet. 2020;29:305–319. doi: 10.1093/hmg/ddz284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Nagahama K., Sakoori K., Watanabe T., Kishi Y., Kawaji K., Koebis M., et al. Setd1a insufficiency in mice attenuates excitatory synaptic function and recapitulates schizophrenia-related behavioral abnormalities. Cell Rep. 2020;32 doi: 10.1016/j.celrep.2020.108126. [DOI] [PubMed] [Google Scholar]
  • 91.Shpargel K.B., Mangini C.L., Xie G., Ge K., Magnuson T. The KMT2D Kabuki syndrome histone methylase controls neural crest cell differentiation and facial morphology. Development. 2020;147 doi: 10.1242/dev.187997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kummeling J., Stremmelaar D.E., Raun N., Reijnders M.R.F., Willemsen M.H., Ruiterkamp-Versteeg M., et al. Characterization of SETD1A haploinsufficiency in humans and Drosophila defines a novel neurodevelopmental syndrome. Mol. Psychiatry. 2021;26:2013–2024. doi: 10.1038/s41380-020-0725-5. [DOI] [PubMed] [Google Scholar]
  • 93.Wang S., van Rhijn J.-R., Akkouh I., Kogo N., Maas N., Bleeck A., et al. Loss-of-function variants in the schizophrenia risk gene SETD1A alter neuronal network activity in human neurons through the cAMP/PKA pathway. Cell Rep. 2022;39 doi: 10.1016/j.celrep.2022.110790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Clifton N.E., Bosworth M.L., Haan N., Rees E., Holmans P.A., Wilkinson L.S., et al. Developmental disruption to the cortical transcriptome and synaptosome in a model of SETD1A loss-of-function. Hum. Mol. Genet. 2022;31:3095–3106. doi: 10.1093/hmg/ddac105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Chong Z.-S., Khong Z.J., Tay S.H., Ng S.-Y. Metabolic contributions to neuronal deficits caused by genomic disruption of schizophrenia risk gene SETD1A. Schizophrenia (Heidelb) 2022;8:115. doi: 10.1038/s41537-022-00326-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Cao K., Collings C.K., Morgan M.A., Marshall S.A., Rendleman E.J., Ozark P.A., et al. An Mll4/COMPASS-Lsd1 epigenetic axis governs enhancer function and pluripotency transition in embryonic stem cells. Sci. Adv. 2018;4 doi: 10.1126/sciadv.aap8747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Rabadan M.A., De La Cruz E.D., Rao S.B., Chen Y., Gong C., Crabtree G., et al. An in vitro model of neuronal ensembles. Nat. Commun. 2022;13:3340. doi: 10.1038/s41467-022-31073-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Pravata V.M., Omelková M., Stavridis M.P., Desbiens C.M., Stephen H.M., Lefeber D.J., et al. An intellectual disability syndrome with single-nucleotide variants in O-GlcNAc transferase. Eur. J. Hum. Genet. 2020;28:706–714. doi: 10.1038/s41431-020-0589-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Lan F., Collins R.E., De Cegli R., Alpatov R., Horton J.R., Shi X., et al. Recognition of unmethylated histone H3 lysine 4 links BHC80 to LSD1-mediated gene repression. Nature. 2007;448:718–722. doi: 10.1038/nature06034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Garay P.M., Wallner M.A., Iwase S. Yin-yang actions of histone methylation regulatory complexes in the brain. Epigenomics. 2016;8:1689–1708. doi: 10.2217/epi-2016-0090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Lakowski B., Roelens I., Jacob S. CoREST-like complexes regulate chromatin modification and neuronal gene expression. J. Mol. Neurosci. 2006;29:227–239. doi: 10.1385/JMN:29:3:227. [DOI] [PubMed] [Google Scholar]
  • 102.Lopez C.I., Saud K.E., Aguilar R., Berndt F.A., Cánovas J., Montecino M., et al. The chromatin modifying complex CoREST/LSD1 negatively regulates notch pathway during cerebral cortex development. Dev. Neurobiol. 2016;76:1360–1373. doi: 10.1002/dneu.22397. [DOI] [PubMed] [Google Scholar]
  • 103.Swahari V., West A.E. Histone demethylases in neuronal differentiation, plasticity, and disease. Curr. Opin. Neurobiol. 2019;59:9–15. doi: 10.1016/j.conb.2019.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Laurent B., Ruitu L., Murn J., Hempel K., Ferrao R., Xiang Y., et al. A specific LSD1/KDM1A isoform regulates neuronal differentiation through H3K9 demethylation. Mol. Cell. 2015;57:957–970. doi: 10.1016/j.molcel.2015.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Rusconi F., Paganini L., Braida D., Ponzoni L., Toffolo E., Maroli A., et al. LSD1 neurospecific alternative splicing controls neuronal excitability in mouse models of epilepsy. Cereb. Cortex. 2015;25:2729–2740. doi: 10.1093/cercor/bhu070. [DOI] [PubMed] [Google Scholar]
  • 106.Li A., Sun Y., Dou C., Chen J., Zhang J. Lysine-specific demethylase 1 expression in zebrafish during the early stages of neuronal development. Neural Regen Res. 2012;7:2719–2726. doi: 10.3969/j.issn.1673-5374.2012.34.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Christopher M.A., Myrick D.A., Barwick B.G., Engstrom A.K., Porter-Stransky K.A., Boss J.M., et al. LSD1 protects against hippocampal and cortical neurodegeneration. Nat. Commun. 2017;8:805. doi: 10.1038/s41467-017-00922-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Rusconi F., Grillo B., Ponzoni L., Bassani S., Toffolo E., Paganini L., et al. LSD1 modulates stress-evoked transcription of immediate early genes and emotional behavior. Proc. Natl. Acad. Sci. U. S. A. 2016;113:3651–3656. doi: 10.1073/pnas.1511974113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Hirano K., Namihira M. LSD1 mediates neuronal differentiation of human fetal neural stem cells by controlling the expression of a novel target gene, HEYL. Stem Cells. 2016;34:1872–1882. doi: 10.1002/stem.2362. [DOI] [PubMed] [Google Scholar]
  • 110.Sáez J.E., Gómez A.V., Barrios Á.P., Parada G.E., Galdames L., González M., et al. Decreased expression of CoREST1 and CoREST2 together with LSD1 and HDAC1/2 during neuronal differentiation. PLoS One. 2015;10 doi: 10.1371/journal.pone.0131760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Zhang F., Xu D., Yuan L., Sun Y., Xu Z. Epigenetic regulation of Atrophin1 by lysine-specific demethylase 1 is required for cortical progenitor maintenance. Nat. Commun. 2014;5:5815. doi: 10.1038/ncomms6815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Han X., Gui B., Xiong C., Zhao L., Liang J., Sun L., et al. Destabilizing LSD1 by Jade-2 promotes neurogenesis: an antibraking system in neural development. Mol. Cell. 2014;55:482–494. doi: 10.1016/j.molcel.2014.06.006. [DOI] [PubMed] [Google Scholar]
  • 113.Zibetti C., Adamo A., Binda C., Forneris F., Toffolo E., Verpelli C., et al. Alternative splicing of the histone demethylase LSD1/KDM1 contributes to the modulation of neurite morphogenesis in the mammalian nervous system. J. Neurosci. 2010;30:2521–2532. doi: 10.1523/JNEUROSCI.5500-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Ferdous S., Shelton D.A., Getz T.E., Chrenek M.A., L’Hernault N., Sellers J.T., et al. Deletion of histone demethylase Lsd1 (Kdm1a) during retinal development leads to defects in retinal function and structure. Front. Cell Neurosci. 2023;17 doi: 10.3389/fncel.2023.1104592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Pilotto S., Speranzini V., Marabelli C., Rusconi F., Toffolo E., Grillo B., et al. LSD1/KDM1A mutations associated to a newly described form of intellectual disability impair demethylase activity and binding to transcription factors. Hum. Mol. Genet. 2016;25:2578–2587. doi: 10.1093/hmg/ddw120. [DOI] [PubMed] [Google Scholar]
  • 116.Kim H.-G., Rosenfeld J.A., Scott D.A., Bénédicte G., Labonne J.D., Brown J., et al. Disruption of PHF21A causes syndromic intellectual disability with craniofacial anomalies, epilepsy, hypotonia, and neurobehavioral problems including autism. Mol. Autism. 2019;10:35. doi: 10.1186/s13229-019-0286-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Klajn A., Ferrai C., Stucchi L., Prada I., Podini P., Baba T., et al. The rest repression of the neurosecretory phenotype is negatively modulated by BHC80, a protein of the BRAF/HDAC complex. J. Neurosci. 2009;29:6296–6307. doi: 10.1523/JNEUROSCI.5943-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Iwase S., Januma A., Miyamoto K., Shono N., Honda A., Yanagisawa J., et al. Characterization of BHC80 in BRAF-HDAC complex, involved in neuron-specific gene repression. Biochem. Biophys. Res. Commun. 2004;322:601–608. doi: 10.1016/j.bbrc.2004.07.163. [DOI] [PubMed] [Google Scholar]
  • 119.Zhao X., Yu T., Tang J., Yao R.-E., Li N., Wang J. Two patients with KDM3B variants and new presentations of Diets-Jongmans syndrome. Neurogenetics. 2023;24:95–101. doi: 10.1007/s10048-023-00711-1. [DOI] [PubMed] [Google Scholar]
  • 120.Diets I.J., van der Donk R., Baltrunaite K., Waanders E., Reijnders M.R.F., Dingemans A.J.M., et al. De novo and inherited pathogenic variants in KDM3B cause intellectual disability, short stature, and facial dysmorphism. Am. J. Hum. Genet. 2019;104:758–766. doi: 10.1016/j.ajhg.2019.02.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Kim Y.G., Bak M.S., Kim A., Kim Y., Chae Y.-C., Kim Y.L., et al. Kdm3b haploinsufficiency impairs the consolidation of cerebellum-dependent motor memory in mice. Mol. Brain. 2021;14:106. doi: 10.1186/s13041-021-00815-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Slavotinek A., van Hagen J.M., Kalsner L., Pai S., Davis-Keppen L., Ohden L., et al. Jumonji domain containing 1C (JMJD1C) sequence variants in seven patients with autism spectrum disorder, intellectual disability and seizures. Eur. J. Med. Genet. 2020;63 doi: 10.1016/j.ejmg.2020.103850. [DOI] [PubMed] [Google Scholar]
  • 123.Sáez M.A., Fernández-Rodríguez J., Moutinho C., Sanchez-Mut J.V., Gomez A., Vidal E., et al. Mutations in JMJD1C are involved in Rett syndrome and intellectual disability. Genet. Med. 2016;18:378–385. doi: 10.1038/gim.2015.100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Ma Z., Chalkley R.J., Vosseller K. Hyper-O-GlcNAcylation activates nuclear factor κ-light-chain-enhancer of activated B cells (NF-κB) signaling through interplay with phosphorylation and acetylation. J. Biol. Chem. 2017;292:9150–9163. doi: 10.1074/jbc.M116.766568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Allison D.F., Wamsley J.J., Kumar M., Li D., Gray L.G., Hart G.W., et al. Modification of RelA by O-linked N-acetylglucosamine links glucose metabolism to NF-κB acetylation and transcription. Proc. Natl. Acad. Sci. U. S. A. 2012;109:16888–16893. doi: 10.1073/pnas.1208468109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Li X., Chen X., Zhou W., Ji S., Li X., Li G., et al. Effect of melatonin on neuronal differentiation requires CBP/p300-mediated acetylation of histone H3 lysine 14. Neuroscience. 2017;364:45–59. doi: 10.1016/j.neuroscience.2017.07.064. [DOI] [PubMed] [Google Scholar]
  • 127.Bai B., Zhang Q., Wan C., Li D., Zhang T., Li H. CBP/p300 inhibitor C646 prevents high glucose exposure induced neuroepithelial cell proliferation. Birth Defects Res. 2018;110:1118–1128. doi: 10.1002/bdr2.1360. [DOI] [PubMed] [Google Scholar]
  • 128.Wang J., Weaver I.C.G., Gauthier-Fisher A., Wang H., He L., Yeomans J., et al. CBP histone acetyltransferase activity regulates embryonic neural differentiation in the normal and Rubinstein-Taybi syndrome brain. Dev. Cell. 2010;18:114–125. doi: 10.1016/j.devcel.2009.10.023. [DOI] [PubMed] [Google Scholar]
  • 129.Alarcón J.M., Malleret G., Touzani K., Vronskaya S., Ishii S., Kandel E.R., et al. Chromatin acetylation, memory, and LTP are impaired in CBP+/- mice: a model for the cognitive deficit in Rubinstein-Taybi syndrome and its amelioration. Neuron. 2004;42:947–959. doi: 10.1016/j.neuron.2004.05.021. [DOI] [PubMed] [Google Scholar]
  • 130.Hung J.-J., Wang Y.-T., Chang W.-C. Sp1 deacetylation induced by phorbol ester recruits p300 to activate 12(S)-lipoxygenase gene transcription. Mol. Cell Biol. 2006;26:1770–1785. doi: 10.1128/MCB.26.5.1770-1785.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Shimizu Y., Kawasaki T. Histone acetyltransferase EP300 regulates the proliferation and differentiation of neural stem cells during adult neurogenesis and regenerative neurogenesis in the zebrafish optic tectum. Neurosci. Lett. 2021;756 doi: 10.1016/j.neulet.2021.135978. [DOI] [PubMed] [Google Scholar]
  • 132.Li Y., Stockton M.E., Eisinger B.E., Zhao Y., Miller J.L., Bhuiyan I., et al. Reducing histone acetylation rescues cognitive deficits in a mouse model of Fragile X syndrome. Nat. Commun. 2018;9:2494. doi: 10.1038/s41467-018-04869-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Bhattacherjee V., Horn K.H., Singh S., Webb C.L., Pisano M.M., Greene R.M. CBP/p300 and associated transcriptional co-activators exhibit distinct expression patterns during murine craniofacial and neural tube development. Int. J. Dev. Biol. 2009;53:1097–1104. doi: 10.1387/ijdb.072489vb. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Nakashima K., Yanagisawa M., Arakawa H., Kimura N., Hisatsune T., Kawabata M., et al. Synergistic signaling in fetal brain by STAT3-Smad1 complex bridged by p300. Science. 1999;284:479–482. doi: 10.1126/science.284.5413.479. [DOI] [PubMed] [Google Scholar]
  • 135.Babu A., Kamaraj M., Basu M., Mukherjee D., Kapoor S., Ranjan S., et al. Chemical and genetic rescue of an ep300 knockdown model for Rubinstein Taybi Syndrome in zebrafish. Biochim. Biophys. Acta Mol. Basis Dis. 2018;1864:1203–1215. doi: 10.1016/j.bbadis.2018.01.029. [DOI] [PubMed] [Google Scholar]
  • 136.Schoof M., Launspach M., Holdhof D., Nguyen L., Engel V., Filser S., et al. The transcriptional coactivator and histone acetyltransferase CBP regulates neural precursor cell development and migration. Acta Neuropathol. Commun. 2019;7:199. doi: 10.1186/s40478-019-0849-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Vieira P.A., Korzus E. CBP-Dependent memory consolidation in the prefrontal cortex supports object-location learning. Hippocampus. 2015;25:1532–1540. doi: 10.1002/hipo.22473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Impey S., Fong A.L., Wang Y., Cardinaux J.R., Fass D.M., Obrietan K., et al. Phosphorylation of CBP mediates transcriptional activation by neural activity and CaM kinase IV. Neuron. 2002;34:235–244. doi: 10.1016/s0896-6273(02)00654-2. [DOI] [PubMed] [Google Scholar]
  • 139.Tanaka Y., Naruse I., Hongo T., Xu M., Nakahata T., Maekawa T., et al. Extensive brain hemorrhage and embryonic lethality in a mouse null mutant of CREB-binding protein. Mech. Dev. 2000;95:133–145. doi: 10.1016/s0925-4773(00)00360-9. [DOI] [PubMed] [Google Scholar]
  • 140.Partanen A., Motoyama J., Hui C.C. Developmentally regulated expression of the transcriptional cofactors/histone acetyltransferases CBP and p300 during mouse embryogenesis. Int. J. Dev. Biol. 1999;43:487–494. [PubMed] [Google Scholar]
  • 141.Malik A.N., Vierbuchen T., Hemberg M., Rubin A.A., Ling E., Couch C.H., et al. Genome-wide identification and characterization of functional neuronal activity-dependent enhancers. Nat. Neurosci. 2014;17:1330–1339. doi: 10.1038/nn.3808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Hayakawa K., Hirosawa M., Tabei Y., Arai D., Tanaka S., Murakami N., et al. Epigenetic switching by the metabolism-sensing factors in the generation of orexin neurons from mouse embryonic stem cells. J. Biol. Chem. 2013;288:17099–17110. doi: 10.1074/jbc.M113.455899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Tapias A., Zhou Z.-W., Shi Y., Chong Z., Wang P., Groth M., et al. Trrap-dependent histone acetylation specifically regulates cell-cycle gene transcription to control neural progenitor fate decisions. Cell Stem Cell. 2014;14:632–643. doi: 10.1016/j.stem.2014.04.001. [DOI] [PubMed] [Google Scholar]
  • 144.Rajagopalan D., Pandey A.K., Xiuzhen M.C., Lee K.K., Hora S., Zhang Y., et al. TIP60 represses telomerase expression by inhibiting Sp1 binding to the TERT promoter. PLoS Pathog. 2017;13 doi: 10.1371/journal.ppat.1006681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Chen Y.-J.C., Dent S.Y.R. Conservation and diversity of the eukaryotic SAGA coactivator complex across kingdoms. Epigenetics Chromatin. 2021;14:26. doi: 10.1186/s13072-021-00402-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Yin B.-K., Wang Z.-Q. Beyond HAT adaptor: TRRAP liaisons with Sp1-mediated transcription. Int. J. Mol. Sci. 2021;22 doi: 10.3390/ijms222212445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Herbst D.A., Esbin M.N., Louder R.K., Dugast-Darzacq C., Dailey G.M., Fang Q., et al. Structure of the human SAGA coactivator complex. Nat. Struct. Mol. Biol. 2021;28:989–996. doi: 10.1038/s41594-021-00682-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Elías-Villalobos A., Fort P., Helmlinger D. New insights into the evolutionary conservation of the sole PIKK pseudokinase Tra1/TRRAP. Biochem. Soc. Trans. 2019;47:1597–1608. doi: 10.1042/BST20180496. [DOI] [PubMed] [Google Scholar]
  • 149.Giaimo B.D., Ferrante F., Vallejo D.M., Hein K., Gutierrez-Perez I., Nist A., et al. Histone variant H2A.Z deposition and acetylation directs the canonical Notch signaling response. Nucleic Acids Res. 2018;46:8197–8215. doi: 10.1093/nar/gky551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Sawan C., Hernandez-Vargas H., Murr R., Lopez F., Vaissière T., Ghantous A.Y., et al. Histone acetyltransferase cofactor Trrap maintains self-renewal and restricts differentiation of embryonic stem cells. Stem Cells. 2013;31:979–991. doi: 10.1002/stem.1341. [DOI] [PubMed] [Google Scholar]
  • 151.Murr R., Vaissière T., Sawan C., Shukla V., Herceg Z. Orchestration of chromatin-based processes: mind the TRRAP. Oncogene. 2007;26:5358–5372. doi: 10.1038/sj.onc.1210605. [DOI] [PubMed] [Google Scholar]
  • 152.Gause M., Eissenberg J.C., Macrae A.F., Dorsett M., Misulovin Z., Dorsett D. Nipped-A, the Tra1/TRRAP subunit of the Drosophila SAGA and Tip60 complexes, has multiple roles in Notch signaling during wing development. Mol. Cell Biol. 2006;26:2347–2359. doi: 10.1128/MCB.26.6.2347-2359.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Evangelista F.M., Maglott-Roth A., Stierle M., Brino L., Soutoglou E., Tora L. Transcription and mRNA export machineries SAGA and TREX-2 maintain monoubiquitinated H2B balance required for DNA repair. J. Cell Biol. 2018;217:3382–3397. doi: 10.1083/jcb.201803074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Singh S.K., Maeda K., Eid M.M.A., Almofty S.A., Ono M., Pham P., et al. GANP regulates recruitment of AID to immunoglobulin variable regions by modulating transcription and nucleosome occupancy. Nat. Commun. 2013;4:1830. doi: 10.1038/ncomms2823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Umlauf D., Bonnet J., Waharte F., Fournier M., Stierle M., Fischer B., et al. The human TREX-2 complex is stably associated with the nuclear pore basket. J. Cell Sci. 2013;126:2656–2667. doi: 10.1242/jcs.118000. [DOI] [PubMed] [Google Scholar]
  • 156.Wickramasinghe V.O., Andrews R., Ellis P., Langford C., Gurdon J.B., Stewart M., et al. Selective nuclear export of specific classes of mRNA from mammalian nuclei is promoted by GANP. Nucleic Acids Res. 2014;42:5059–5071. doi: 10.1093/nar/gku095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Jani D., Lutz S., Hurt E., Laskey R.A., Stewart M., Wickramasinghe V.O. Functional and structural characterization of the mammalian TREX-2 complex that links transcription with nuclear messenger RNA export. Nucleic Acids Res. 2012;40:4562–4573. doi: 10.1093/nar/gks059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Takei Y., Swietlik M., Tanoue A., Tsujimoto G., Kouzarides T., Laskey R. MCM3AP, a novel acetyltransferase that acetylates replication protein MCM3. EMBO Rep. 2001;2:119–123. doi: 10.1093/embo-reports/kve026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Suzuki T., Hirai Y., Uehara T., Ohga R., Kosaki K., Kawahara A. Involvement of the zebrafish trrap gene in craniofacial development. Sci. Rep. 2021;11 doi: 10.1038/s41598-021-03123-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Xia W., Hu J., Ma J., Huang J., Wang X., Jiang N., et al. Novel TRRAP mutation causes autosomal dominant non-syndromic hearing loss. Clin. Genet. 2019;96:300–308. doi: 10.1111/cge.13590. [DOI] [PubMed] [Google Scholar]
  • 161.Yin B.-K., Lázaro D., Wang Z.-Q. TRRAP-mediated acetylation on Sp1 regulates adult neurogenesis. Comput. Struct. Biotechnol. J. 2023;21:472–484. doi: 10.1016/j.csbj.2022.12.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Tapias A., Lázaro D., Yin B.-K., Rasa S.M.M., Krepelova A., Kelmer Sacramento E., et al. HAT cofactor TRRAP modulates microtubule dynamics via SP1 signaling to prevent neurodegeneration. Elife. 2021;10 doi: 10.7554/eLife.61531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Woldegebriel R., Kvist J., Andersson N., Õunap K., Reinson K., Wojcik M.H., et al. Distinct effects on mRNA export factor GANP underlie neurological disease phenotypes and alter gene expression depending on intron content. Hum. Mol. Genet. 2020;29:1426–1439. doi: 10.1093/hmg/ddaa051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Yang X., Zhang F., Kudlow J.E. Recruitment of O-GlcNAc transferase to promoters by corepressor mSin3A: coupling protein O-GlcNAcylation to transcriptional repression. Cell. 2002;110:69–80. doi: 10.1016/s0092-8674(02)00810-3. [DOI] [PubMed] [Google Scholar]
  • 165.Streubel G., Fitzpatrick D.J., Oliviero G., Scelfo A., Moran B., Das S., et al. Fam60a defines a variant Sin3a-Hdac complex in embryonic stem cells required for self-renewal. EMBO J. 2017;36:2216–2232. doi: 10.15252/embj.201696307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Grimes J.A., Nielsen S.J., Battaglioli E., Miska E.A., Speh J.C., Berry D.L., et al. The co-repressor mSin3A is a functional component of the REST-CoREST repressor complex. J. Biol. Chem. 2000;275:9461–9467. doi: 10.1074/jbc.275.13.9461. [DOI] [PubMed] [Google Scholar]
  • 167.Bridi M., Schoch H., Florian C., Poplawski S.G., Banerjee A., Hawk J.D., et al. Transcriptional corepressor SIN3A regulates hippocampal synaptic plasticity via Homer1/mGluR5 signaling. JCI Insight. 2020;5 doi: 10.1172/jci.insight.92385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Roopra A., Sharling L., Wood I.C., Briggs T., Bachfischer U., Paquette A.J., et al. Transcriptional repression by neuron-restrictive silencer factor is mediated via the Sin3-histone deacetylase complex. Mol. Cell Biol. 2000;20:2147–2157. doi: 10.1128/mcb.20.6.2147-2157.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Huang Y., Myers S.J., Dingledine R. Transcriptional repression by REST: recruitment of Sin3A and histone deacetylase to neuronal genes. Nat. Neurosci. 1999;2:867–872. doi: 10.1038/13165. [DOI] [PubMed] [Google Scholar]
  • 170.Seki M., Masaki H., Arauchi T., Nakauchi H., Sugano S., Suzuki Y. A comparison of the rest complex binding patterns in embryonic stem cells and epiblast stem cells. PLoS One. 2014;9 doi: 10.1371/journal.pone.0095374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Yoon H.-G., Chan D.W., Huang Z.-Q., Li J., Fondell J.D., Qin J., et al. Purification and functional characterization of the human N-CoR complex: the roles of HDAC3, TBL1 and TBLR1. EMBO J. 2003;22:1336–1346. doi: 10.1093/emboj/cdg120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Yoon H.-G., Choi Y., Cole P.A., Wong J. Reading and function of a histone code involved in targeting corepressor complexes for repression. Mol. Cell Biol. 2005;25:324–335. doi: 10.1128/MCB.25.1.324-335.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Tomita A., Buchholz D.R., Shi Y.-B. Recruitment of N-CoR/SMRT-TBLR1 corepressor complex by unliganded thyroid hormone receptor for gene repression during frog development. Mol. Cell Biol. 2004;24:3337–3346. doi: 10.1128/MCB.24.8.3337-3346.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Perissi V., Aggarwal A., Glass C.K., Rose D.W., Rosenfeld M.G. A corepressor/coactivator exchange complex required for transcriptional activation by nuclear receptors and other regulated transcription factors. Cell. 2004;116:511–526. doi: 10.1016/s0092-8674(04)00133-3. [DOI] [PubMed] [Google Scholar]
  • 175.Hu Y., Lauffer P., Stewart M., Codner G., Mayerl S., Heuer H., et al. An animal model for Pierpont syndrome: a mouse bearing the Tbl1xr1Y446C/Y446C mutation. Hum. Mol. Genet. 2022;31:2951–2963. doi: 10.1093/hmg/ddac086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Mastrototaro G., Zaghi M., Massimino L., Moneta M., Mohammadi N., Banfi F., et al. TBL1XR1 ensures balanced neural development through NCOR complex-mediated regulation of the MAPK pathway. Front. Cell Dev. Biol. 2021;9 doi: 10.3389/fcell.2021.641410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Zhou W., He Y., Rehman A.U., Kong Y., Hong S., Ding G., et al. Loss of function of NCOR1 and NCOR2 impairs memory through a novel GABAergic hypothalamus-CA3 projection. Nat. Neurosci. 2019;22:205–217. doi: 10.1038/s41593-018-0311-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Strobl-Mazzulla P.H., Bronner M.E. A PHD12-Snail2 repressive complex epigenetically mediates neural crest epithelial-to-mesenchymal transition. J. Cell Biol. 2012;198:999–1010. doi: 10.1083/jcb.201203098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Schuff M., Rössner A., Wacker S.A., Donow C., Gessert S., Knöchel W. FoxN3 is required for craniofacial and eye development of Xenopus laevis. Dev. Dyn. 2007;236:226–239. doi: 10.1002/dvdy.21007. [DOI] [PubMed] [Google Scholar]
  • 180.Zhang B., Su W., Hu J., Xu J., Askar P., Bao S., et al. Transcriptome analysis of schwann cells at various stages of myelination implicates chromatin regulator Sin3A in control of myelination identity. Neurosci. Bull. 2022;38:720–740. doi: 10.1007/s12264-022-00850-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Halder D., Lee C.-H., Hyun J.Y., Chang G.-E., Cheong E., Shin I. Suppression of Sin3A activity promotes differentiation of pluripotent cells into functional neurons. Sci. Rep. 2017;7 doi: 10.1038/srep44818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Witteveen J.S., Willemsen M.H., Dombroski T.C.D., van Bakel N.H.M., Nillesen W.M., van Hulten J.A., et al. Haploinsufficiency of MeCP2-interacting transcriptional co-repressor SIN3A causes mild intellectual disability by affecting the development of cortical integrity. Nat. Genet. 2016;48:877–887. doi: 10.1038/ng.3619. [DOI] [PubMed] [Google Scholar]
  • 183.Ching G.Y., Liem R.K.H. RE1 silencing transcription factor is involved in regulating neuron-specific expression of alpha-internexin and neurofilament genes. J. Neurochem. 2009;109:1610–1623. doi: 10.1111/j.1471-4159.2009.06052.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Cox E.J., Marsh S.A. Exercise and diabetes have opposite effects on the assembly and O-GlcNAc modification of the mSin3A/HDAC1/2 complex in the heart. Cardiovasc. Diabetol. 2013;12:101. doi: 10.1186/1475-2840-12-101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Koerner M.V., FitzPatrick L., Selfridge J., Guy J., De Sousa D., Tillotson R., et al. Toxicity of overexpressed MeCP2 is independent of HDAC3 activity. Genes Dev. 2018;32:1514–1524. doi: 10.1101/gad.320325.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Kruusvee V., Lyst M.J., Taylor C., Tarnauskaitė Ž., Bird A.P., Cook A.G. Structure of the MeCP2-TBLR1 complex reveals a molecular basis for Rett syndrome and related disorders. Proc. Natl. Acad. Sci. U. S. A. 2017;114:E3243–E3250. doi: 10.1073/pnas.1700731114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Cheng P.-Y., Lin Y.-P., Chen Y.-L., Lee Y.-C., Tai C.-C., Wang Y.-T., et al. Interplay between SIN3A and STAT3 mediates chromatin conformational changes and GFAP expression during cellular differentiation. PLoS One. 2011;6 doi: 10.1371/journal.pone.0022018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Martinowich K., Hattori D., Wu H., Fouse S., He F., Hu Y., et al. DNA methylation-related chromatin remodeling in activity-dependent BDNF gene regulation. Science. 2003;302:890–893. doi: 10.1126/science.1090842. [DOI] [PubMed] [Google Scholar]
  • 189.Napoletani G., Vigli D., Cosentino L., Grieco M., Talamo M.C., Lacivita E., et al. Stimulation of the serotonin receptor 7 restores brain histone H3 acetylation and MeCP2 corepressor protein levels in a female mouse model of Rett syndrome. J. Neuropathol. Exp. Neurol. 2021;80:265–273. doi: 10.1093/jnen/nlaa158. [DOI] [PubMed] [Google Scholar]
  • 190.Cheng J., Zhao Z., Chen L., Li Y., Du R., Wu Y., et al. Loss of O-GlcNAcylation on MeCP2 at threonine 203 leads to neurodevelopmental disorders. Neurosci. Bull. 2022;38:113–134. doi: 10.1007/s12264-021-00784-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Dey A., Seshasayee D., Noubade R., French D.M., Liu J., Chaurushiya M.S., et al. Loss of the tumor suppressor BAP1 causes myeloid transformation. Science. 2012;337:1541–1546. doi: 10.1126/science.1221711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Sahtoe D.D., van Dijk W.J., Ekkebus R., Ovaa H., Sixma T.K. BAP1/ASXL1 recruitment and activation for H2A deubiquitination. Nat. Commun. 2016;7 doi: 10.1038/ncomms10292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Campagne A., Lee M.-K., Zielinski D., Michaud A., Le Corre S., Dingli F., et al. BAP1 complex promotes transcription by opposing PRC1-mediated H2A ubiquitylation. Nat. Commun. 2019;10:348. doi: 10.1038/s41467-018-08255-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Doyle L.A., Unlu Bektas F., Chatzantonaki E., Repton C., Derrien A., Illingworth R.S. RINGs, DUBs and abnormal brain growth-histone H2A ubiquitination in brain development and disease. Epigenomes. 2022;6:42. doi: 10.3390/epigenomes6040042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Park U.-H., Seong M., Kim E.-J., Hur W., Kim S.W., Yoon S.K., et al. Reciprocal regulation of LXRα activity by ASXL1 and ASXL2 in lipogenesis. Biochem. Biophys. Res. Commun. 2014;443:489–494. doi: 10.1016/j.bbrc.2013.11.124. [DOI] [PubMed] [Google Scholar]
  • 196.Zhang Y., Yu H., Wang D., Lei X., Meng Y., Zhang N., et al. Protein O-GlcNAcylation homeostasis regulates facultative heterochromatin to fine-tune sog-Dpp signaling during Drosophila early embryogenesis. J. Genet. Genomics. 2023;50:948–959. doi: 10.1016/j.jgg.2023.05.014. [DOI] [PubMed] [Google Scholar]
  • 197.Sun X., Cheng L., Sun Y. Autism-associated protein POGZ controls ESCs and ESC neural induction by association with esBAF. Mol. Autism. 2022;13:24. doi: 10.1186/s13229-022-00502-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Bi-Lin K.W., Seshachalam P.V., Tuoc T., Stoykova A., Ghosh S., Singh M.K. Critical role of the BAF chromatin remodeling complex during murine neural crest development. PLoS Genet. 2021;17 doi: 10.1371/journal.pgen.1009446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Nguyen H., Sokpor G., Pham L., Rosenbusch J., Stoykova A., Staiger J.F., et al. Epigenetic regulation by BAF (mSWI/SNF) chromatin remodeling complexes is indispensable for embryonic development. Cell Cycle. 2016;15:1317–1324. doi: 10.1080/15384101.2016.1160984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Narayanan R., Pirouz M., Kerimoglu C., Pham L., Wagener R.J., Kiszka K.A., et al. Loss of BAF (mSWI/SNF) complexes causes global transcriptional and chromatin State changes in forebrain development. Cell Rep. 2015;13:1842–1854. doi: 10.1016/j.celrep.2015.10.046. [DOI] [PubMed] [Google Scholar]
  • 201.Zhang X., Li B., Li W., Ma L., Zheng D., Li L., et al. Transcriptional repression by the BRG1-SWI/SNF complex affects the pluripotency of human embryonic stem cells. Stem Cell Rep. 2014;3:460–474. doi: 10.1016/j.stemcr.2014.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Ho L., Ronan J.L., Wu J., Staahl B.T., Chen L., Kuo A., et al. An embryonic stem cell chromatin remodeling complex, esBAF, is essential for embryonic stem cell self-renewal and pluripotency. Proc. Natl. Acad. Sci. U. S. A. 2009;106:5181–5186. doi: 10.1073/pnas.0812889106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Rada-Iglesias A., Bajpai R., Swigut T., Brugmann S.A., Flynn R.A., Wysocka J. A unique chromatin signature uncovers early developmental enhancers in humans. Nature. 2011;470:279–283. doi: 10.1038/nature09692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Lei I., West J., Yan Z., Gao X., Fang P., Dennis J.H., et al. BAF250a protein regulates nucleosome occupancy and histone modifications in priming embryonic stem cell differentiation. J. Biol. Chem. 2015;290:19343–19352. doi: 10.1074/jbc.M115.637389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Flores-Alcantar A., Gonzalez-Sandoval A., Escalante-Alcalde D., Lomelí H. Dynamics of expression of ARID1A and ARID1B subunits in mouse embryos and in cells during the cell cycle. Cell Tissue Res. 2011;345:137–148. doi: 10.1007/s00441-011-1182-x. [DOI] [PubMed] [Google Scholar]
  • 206.Li X.S., Trojer P., Matsumura T., Treisman J.E., Tanese N. Mammalian SWI/SNF--a subunit BAF250/ARID1 is an E3 ubiquitin ligase that targets histone H2B. Mol. Cell Biol. 2010;30:1673–1688. doi: 10.1128/MCB.00540-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Laurette P., Coassolo S., Davidson G., Michel I., Gambi G., Yao W., et al. Chromatin remodellers Brg1 and Bptf are required for normal gene expression and progression of oncogenic Braf-driven mouse melanoma. Cell Death Differ. 2020;27:29–43. doi: 10.1038/s41418-019-0333-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Zapata G., Yan K., Picketts D.J. Generation of a mouse model of the neurodevelopmental disorder with dysmorphic facies and distal limb anomalies syndrome. Hum. Mol. Genet. 2022;31:3405–3421. doi: 10.1093/hmg/ddac119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Landry J., Sharov A.A., Piao Y., Sharova L.V., Xiao H., Southon E., et al. Essential role of chromatin remodeling protein Bptf in early mouse embryos and embryonic stem cells. PLoS Genet. 2008;4 doi: 10.1371/journal.pgen.1000241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Barak O., Lazzaro M.A., Lane W.S., Speicher D.W., Picketts D.J., Shiekhattar R. Isolation of human NURF: a regulator of Engrailed gene expression. EMBO J. 2003;22:6089–6100. doi: 10.1093/emboj/cdg582. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Li H., Ilin S., Wang W., Duncan E.M., Wysocka J., Allis C.D., et al. Molecular basis for site-specific read-out of histone H3K4me3 by the BPTF PHD finger of NURF. Nature. 2006;442:91–95. doi: 10.1038/nature04802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Wysocka J., Swigut T., Xiao H., Milne T.A., Kwon S.Y., Landry J., et al. A PHD finger of NURF couples histone H3 lysine 4 trimethylation with chromatin remodelling. Nature. 2006;442:86–90. doi: 10.1038/nature04815. [DOI] [PubMed] [Google Scholar]
  • 213.Kitagawa T., Kobayashi D., Baron B., Okita H., Miyamoto T., Takai R., et al. AT-hook DNA-binding motif-containing protein one knockdown downregulates EWS-FLI1 transcriptional activity in Ewing’s sarcoma cells. PLoS One. 2022;17 doi: 10.1371/journal.pone.0269077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Suliman-Lavie R., Title B., Cohen Y., Hamada N., Tal M., Tal N., et al. Pogz deficiency leads to transcription dysregulation and impaired cerebellar activity underlying autism-like behavior in mice. Nat Commun. 2020;11:5836. doi: 10.1038/s41467-020-19577-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Stessman H.A.F., Willemsen M.H., Fenckova M., Penn O., Hoischen A., Xiong B., et al. Disruption of POGZ is associated with intellectual disability and autism spectrum disorders. Am. J. Hum. Genet. 2016;98:541–552. doi: 10.1016/j.ajhg.2016.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Matsumura K., Seiriki K., Okada S., Nagase M., Ayabe S., Yamada I., et al. Pathogenic POGZ mutation causes impaired cortical development and reversible autism-like phenotypes. Nat. Commun. 2020;11:859. doi: 10.1038/s41467-020-14697-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Deng L., Mojica-Perez S.P., Azaria R.D., Schultz M., Parent J.M., Niu W. Loss of POGZ alters neural differentiation of human embryonic stem cells. Mol. Cell Neurosci. 2022;120 doi: 10.1016/j.mcn.2022.103727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Zhao W., Tan J., Zhu T., Ou J., Li Y., Shen L., et al. Rare inherited missense variants of POGZ associate with autism risk and disrupt neuronal development. J. Genet. Genomics. 2019;46:247–257. doi: 10.1016/j.jgg.2019.04.002. [DOI] [PubMed] [Google Scholar]
  • 219.Cunniff M.M., Markenscoff-Papadimitriou E., Ostrowski J., Rubenstein J.L., Sohal V.S. Altered hippocampal-prefrontal communication during anxiety-related avoidance in mice deficient for the autism-associated gene Pogz. Elife. 2020;9 doi: 10.7554/eLife.54835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Ibaraki K., Hamada N., Iwamoto I., Ito H., Kawamura N., Morishita R., et al. Expression analyses of POGZ, A responsible gene for neurodevelopmental disorders, during mouse brain development. Dev. Neurosci. 2019;41:139–148. doi: 10.1159/000502128. [DOI] [PubMed] [Google Scholar]
  • 221.Markenscoff-Papadimitriou E., Binyameen F., Whalen S., Price J., Lim K., Ypsilanti A.R., et al. Autism risk gene POGZ promotes chromatin accessibility and expression of clustered synaptic genes. Cell Rep. 2021;37 doi: 10.1016/j.celrep.2021.110089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Conrow-Graham M., Williams J.B., Martin J., Zhong P., Cao Q., Rein B., et al. A convergent mechanism of high risk factors ADNP and POGZ in neurodevelopmental disorders. Brain. 2022;145:3250–3263. doi: 10.1093/brain/awac152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Staahl B.T., Crabtree G.R. Creating a neural specific chromatin landscape by npBAF and nBAF complexes. Curr. Opin. Neurobiol. 2013;23:903–913. doi: 10.1016/j.conb.2013.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Yan Z., Wang Z., Sharova L., Sharov A.A., Ling C., Piao Y., et al. BAF250B-associated SWI/SNF chromatin-remodeling complex is required to maintain undifferentiated mouse embryonic stem cells. Stem Cells. 2008;26:1155–1165. doi: 10.1634/stemcells.2007-0846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Tuoc T.C., Narayanan R., Stoykova A. BAF chromatin remodeling complex: cortical size regulation and beyond. Cell Cycle. 2013;12:2953–2959. doi: 10.4161/cc.25999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Tuoc T., Dere E., Radyushkin K., Pham L., Nguyen H., Tonchev A.B., et al. Ablation of BAF170 in developing and postnatal dentate gyrus affects neural stem cell proliferation, differentiation, and learning. Mol. Neurobiol. 2017;54:4618–4635. doi: 10.1007/s12035-016-9948-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Tuoc T.C., Boretius S., Sansom S.N., Pitulescu M.-E., Frahm J., Livesey F.J., et al. Chromatin regulation by BAF170 controls cerebral cortical size and thickness. Dev. Cell. 2013;25:256–269. doi: 10.1016/j.devcel.2013.04.005. [DOI] [PubMed] [Google Scholar]
  • 228.Nguyen H., Sokpor G., Parichha A., Pham L., Saikhedkar N., Xie Y., et al. BAF (mSWI/SNF) complex regulates mediolateral cortical patterning in the developing forebrain. Front. Cell Dev. Biol. 2022;10 doi: 10.3389/fcell.2022.1011109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Holdhof D., Schoof M., Al-Kershi S., Spohn M., Kresbach C., Göbel C., et al. Brahma-related gene 1 has time-specific roles during brain and eye development. Development. 2021;148 doi: 10.1242/dev.196147. [DOI] [PubMed] [Google Scholar]
  • 230.Jin Y., Gao X., Lu M., Chen G., Yang X., Ren N., et al. Loss of BAF (mSWI/SNF) chromatin-remodeling ATPase Brg1 causes multiple malformations of cortical development in mice. Hum. Mol. Genet. 2022;31:3504–3520. doi: 10.1093/hmg/ddac127. [DOI] [PubMed] [Google Scholar]
  • 231.Chandler R.L., Magnuson T. The SWI/SNF BAF-A complex is essential for neural crest development. Dev. Biol. 2016;411:15–24. doi: 10.1016/j.ydbio.2016.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Deng L., Li G., Rao B., Li H. Central nervous system-specific knockout of Brg1 causes growth retardation and neuronal degeneration. Brain Res. 2015;1622:186–195. doi: 10.1016/j.brainres.2015.06.027. [DOI] [PubMed] [Google Scholar]
  • 233.Eroglu B., Wang G., Tu N., Sun X., Mivechi N.F. Critical role of Brg1 member of the SWI/SNF chromatin remodeling complex during neurogenesis and neural crest induction in zebrafish. Dev. Dyn. 2006;235:2722–2735. doi: 10.1002/dvdy.20911. [DOI] [PubMed] [Google Scholar]
  • 234.Liu P.-P., Dai S.-K., Mi T.-W., Tang G.-B., Wang Z., Wang H., et al. Acetate supplementation restores cognitive deficits caused by ARID1A haploinsufficiency in excitatory neurons. EMBO Mol. Med. 2022;14 doi: 10.15252/emmm.202215795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Pagliaroli L., Porazzi P., Curtis A.T., Scopa C., Mikkers H.M.M., Freund C., et al. Inability to switch from ARID1A-BAF to ARID1B-BAF impairs exit from pluripotency and commitment towards neural crest formation in ARID1B-related neurodevelopmental disorders. Nat. Commun. 2021;12:6469. doi: 10.1038/s41467-021-26810-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Moffat J.J., Jung E.-M., Ka M., Jeon B.T., Lee H., Kim W.-Y. Differential roles of ARID1B in excitatory and inhibitory neural progenitors in the developing cortex. Sci. Rep. 2021;11:3856. doi: 10.1038/s41598-021-82974-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Celen C., Chuang J.-C., Luo X., Nijem N., Walker A.K., Chen F., et al. Arid1b haploinsufficient mice reveal neuropsychiatric phenotypes and reversible causes of growth impairment. Elife. 2017;6 doi: 10.7554/eLife.25730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Marom R., Jain M., Burrage L.C., Song I.-W., Graham B.H., Brown C.W., et al. Heterozygous variants in ACTL6A, encoding a component of the BAF complex, are associated with intellectual disability. Hum. Mutat. 2017;38:1365–1371. doi: 10.1002/humu.23282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Yuan Z.-Z., Xie X.-H., Gu H., Zhang W.-Z., Hu Y.-Q., Yang Y.-F., et al. Case Report: BAF-Opathies/SSRIDDs due to a de novo ACTL6A variant, previously considered to be heart-hand syndrome. Front. Cardiovasc. Med. 2021;8 doi: 10.3389/fcvm.2021.708033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Ma Y., Liu X., Liu Z., Wei S., Shang H., Xue Y., et al. The chromatin remodeling protein Bptf promotes posterior neuroectodermal fate by enhancing Smad2-activated wnt8a expression. J. Neurosci. 2015;35:8493–8506. doi: 10.1523/JNEUROSCI.0377-15.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Grandi F.C., Modi H., Kampman L., Corces M.R. Chromatin accessibility profiling by ATAC-seq. Nat. Protoc. 2022;17:1518–1552. doi: 10.1038/s41596-022-00692-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Brimble S., Wollaston-Hayden E.E., Teo C.F., Morris A.C., Wells L. The role of the O-GlcNAc modification in regulating eukaryotic gene expression. Curr. Signal. Transduct. Ther. 2010;5:12–24. doi: 10.2174/157436210790226465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Antonova S.V., Haffke M., Corradini E., Mikuciunas M., Low T.Y., Signor L., et al. Chaperonin CCT checkpoint function in basal transcription factor TFIID assembly. Nat. Struct. Mol. Biol. 2018;25:1119–1127. doi: 10.1038/s41594-018-0156-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Bieniossek C., Papai G., Schaffitzel C., Garzoni F., Chaillet M., Scheer E., et al. The architecture of human general transcription factor TFIID core complex. Nature. 2013;493:699–702. doi: 10.1038/nature11791. [DOI] [PubMed] [Google Scholar]
  • 245.Weake V.M., Swanson S.K., Mushegian A., Florens L., Washburn M.P., Abmayr S.M., et al. A novel histone fold domain-containing protein that replaces TAF6 in Drosophila SAGA is required for SAGA-dependent gene expression. Genes Dev. 2009;23:2818–2823. doi: 10.1101/gad.1846409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Aoyagi N., Wassarman D.A. Developmental and transcriptional consequences of mutations in Drosophila TAF(II)60. Mol. Cell Biol. 2001;21:6808–6819. doi: 10.1128/MCB.21.20.6808-6819.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Lee T.I., Causton H.C., Holstege F.C., Shen W.C., Hannett N., Jennings E.G., et al. Redundant roles for the TFIID and SAGA complexes in global transcription. Nature. 2000;405:701–704. doi: 10.1038/35015104. [DOI] [PubMed] [Google Scholar]
  • 248.Wang H., Dienemann C., Stützer A., Urlaub H., Cheung A.C.M., Cramer P. Structure of the transcription coactivator SAGA. Nature. 2020;577:717–720. doi: 10.1038/s41586-020-1933-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Dahiya R., Natarajan K. Mutational analysis of TAF6 revealed the essential requirement of the histone-fold domain and the HEAT repeat domain for transcriptional activation. FEBS J. 2018;285:1491–1510. doi: 10.1111/febs.14423. [DOI] [PubMed] [Google Scholar]
  • 250.Janssen B.D.E., van den Boogaard M.-J.H., Lichtenbelt K., Seaby E.G., Stals K., Ellard S., et al. De novo putative loss-of-function variants in TAF4 are associated with a neuro-developmental disorder. Hum. Mutat. 2022;43:1844–1851. doi: 10.1002/humu.24444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Park J., Park G.-Y., Lee J., Park J., Kim S., Kim E., et al. ERK phosphorylation disrupts the intramolecular interaction of capicua to promote cytoplasmic translocation of capicua and tumor growth. Front. Mol. Biosci. 2022;9 doi: 10.3389/fmolb.2022.1030725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Kim J.W., Luck C., Wu W., Ponce R.K., Lin Y.K., Gupta N., et al. Capicua suppresses YAP1 to limit tumorigenesis and maintain drug sensitivity in human cancer. Cell Rep. 2022;41 doi: 10.1016/j.celrep.2022.111443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Patel A.L., Zhang L., Keenan S.E., Rushlow C.A., Fradin C., Shvartsman S.Y. Capicua is a fast-acting transcriptional brake. Curr. Biol. 2021;31:3639–3647.e5. doi: 10.1016/j.cub.2021.05.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Keenan S.E., Blythe S.A., Marmion R.A., Djabrayan N.J.-V., Wieschaus E.F., Shvartsman S.Y. Rapid dynamics of Signal-dependent transcriptional repression by capicua. Dev. Cell. 2020;52:794–801.e4. doi: 10.1016/j.devcel.2020.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Bunda S., Heir P., Li A.S.C., Mamatjan Y., Zadeh G., Aldape K. c-Src Phosphorylates and inhibits the function of the CIC tumor suppressor protein. Mol. Cancer Res. 2020;18:774–786. doi: 10.1158/1541-7786.MCR-18-1370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 256.Paul S., Yang L., Mattingly H., Goyal Y., Shvartsman S.Y., Veraksa A. Activation-induced substrate engagement in ERK signaling. Mol. Biol. Cell. 2020;31:235–243. doi: 10.1091/mbc.E19-07-0355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Astigarraga S., Grossman R., Díaz-Delfín J., Caelles C., Paroush Z., Jiménez G. A MAPK docking site is critical for downregulation of Capicua by Torso and EGFR RTK signaling. EMBO J. 2007;26:668–677. doi: 10.1038/sj.emboj.7601532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Weissmann S., Cloos P.A., Sidoli S., Jensen O.N., Pollard S., Helin K. The tumor suppressor CIC directly regulates MAPK pathway genes via histone deacetylation. Cancer Res. 2018;78:4114–4125. doi: 10.1158/0008-5472.CAN-18-0342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Simón-Carrasco L., Jiménez G., Barbacid M., Drosten M. The Capicua tumor suppressor: a gatekeeper of Ras signaling in development and cancer. Cell Cycle. 2018;17:702–711. doi: 10.1080/15384101.2018.1450029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.LeBlanc V.G., Firme M., Song J., Chan S.Y., Lee M.H., Yip S., et al. Comparative transcriptome analysis of isogenic cell line models and primary cancers links capicua (CIC) loss to activation of the MAPK signalling cascade. J. Pathol. 2017;242:206–220. doi: 10.1002/path.4894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Lim B., Samper N., Lu H., Rushlow C., Jiménez G., Shvartsman S.Y. Kinetics of gene derepression by ERK signaling. Proc. Natl. Acad. Sci. U. S. A. 2013;110:10330–10335. doi: 10.1073/pnas.1303635110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Grimm O., Sanchez Zini V., Kim Y., Casanova J., Shvartsman S.Y., Wieschaus E. Torso RTK controls Capicua degradation by changing its subcellular localization. Development. 2012;139:3962–3968. doi: 10.1242/dev.084327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Ajuria L., Nieva C., Winkler C., Kuo D., Samper N., Andreu M.J., et al. Capicua DNA-binding sites are general response elements for RTK signaling in Drosophila. Development. 2011;138:915–924. doi: 10.1242/dev.057729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Tseng A.-S.K., Tapon N., Kanda H., Cigizoglu S., Edelmann L., Pellock B., et al. Capicua regulates cell proliferation downstream of the receptor tyrosine kinase/ras signaling pathway. Curr. Biol. 2007;17:728–733. doi: 10.1016/j.cub.2007.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Lee S.D., Song J., LeBlanc V.G., Marra M.A. Integrative multi-omic analysis reveals neurodevelopmental gene dysregulation in CIC-knockout and IDH1-mutant cells. J. Pathol. 2022;256:297–309. doi: 10.1002/path.5835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Li K., Shao S., Ji T., Liu M., Wang L., Pang Y., et al. Capicua regulates dendritic morphogenesis through Ets in hippocampal neurons in vitro. Front. Neuroanat. 2021;15 doi: 10.3389/fnana.2021.669310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Hourigan B., Balay S.D., Yee G., Sharma S., Tan Q. Capicua regulates the development of adult-born neurons in the hippocampus. Sci. Rep. 2021;11 doi: 10.1038/s41598-021-91168-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Hwang I., Pan H., Yao J., Elemento O., Zheng H., Paik J. CIC is a critical regulator of neuronal differentiation. JCI Insight. 2020;5 doi: 10.1172/jci.insight.135826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Ahmad S.T., Rogers A.D., Chen M.J., Dixit R., Adnani L., Frankiw L.S., et al. Capicua regulates neural stem cell proliferation and lineage specification through control of Ets factors. Nat. Commun. 2019;10:2000. doi: 10.1038/s41467-019-09949-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Lee C.-J., Chan W.-I., Scotting P.J. CIC, a gene involved in cerebellar development and ErbB signaling, is significantly expressed in medulloblastomas. J. Neurooncol. 2005;73:101–108. doi: 10.1007/s11060-004-4598-2. [DOI] [PubMed] [Google Scholar]
  • 271.Dossin F., Pinheiro I., Żylicz J.J., Roensch J., Collombet S., Le Saux A., et al. SPEN integrates transcriptional and epigenetic control of X-inactivation. Nature. 2020;578:455–460. doi: 10.1038/s41586-020-1974-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Robert-Finestra T., Tan B.F., Mira-Bontenbal H., Timmers E., Gontan C., Merzouk S., et al. SPEN is required for Xist upregulation during initiation of X chromosome inactivation. Nat. Commun. 2021;12:7000. doi: 10.1038/s41467-021-27294-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Markaki Y., Gan Chong J., Wang Y., Jacobson E.C., Luong C., Tan S.Y.X., et al. Xist nucleates local protein gradients to propagate silencing across the X chromosome. Cell. 2021;184:6174–6192.e32. doi: 10.1016/j.cell.2021.10.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Moindrot B., Cerase A., Coker H., Masui O., Grijzenhout A., Pintacuda G., et al. A Pooled shRNA screen identifies Rbm15, Spen, and Wtap as factors required for Xist RNA-mediated silencing. Cell Rep. 2015;12:562–572. doi: 10.1016/j.celrep.2015.06.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Monfort A., Di Minin G., Postlmayr A., Freimann R., Arieti F., Thore S., et al. Identification of Spen as a crucial factor for Xist function through forward genetic screening in haploid embryonic stem cells. Cell Rep. 2015;12:554–561. doi: 10.1016/j.celrep.2015.06.067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.McHugh C.A., Chen C.-K., Chow A., Surka C.F., Tran C., McDonel P., et al. The Xist lncRNA interacts directly with SHARP to silence transcription through HDAC3. Nature. 2015;521:232–236. doi: 10.1038/nature14443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Yabe D., Fukuda H., Aoki M., Yamada S., Takebayashi S., Shinkura R., et al. Generation of a conditional knockout allele for mammalian Spen protein Mint/SHARP. Genesis. 2007;45:300–306. doi: 10.1002/dvg.20296. [DOI] [PubMed] [Google Scholar]
  • 278.Pan L., Hoffmeister P., Turkiewicz A., Huynh N.N.D., Große-Berkenbusch A., Knippschild U., et al. Transcription factor RBPJL is able to repress notch target gene expression but is non-responsive to notch activation. Cancers (Basel) 2021;13:5027. doi: 10.3390/cancers13195027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Doroquez D.B., Orr-Weaver T.L., Rebay I. Split ends antagonizes the Notch and potentiates the EGFR signaling pathways during Drosophila eye development. Mech. Dev. 2007;124:792–806. doi: 10.1016/j.mod.2007.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Kuroda K., Han H., Tani S., Tanigaki K., Tun T., Furukawa T., et al. Regulation of marginal zone B cell development by MINT, a suppressor of Notch/RBP-J signaling pathway. Immunity. 2003;18:301–312. doi: 10.1016/s1074-7613(03)00029-3. [DOI] [PubMed] [Google Scholar]
  • 281.Feng Y., Bommer G.T., Zhai Y., Akyol A., Hinoi T., Winer I., et al. Drosophila split ends homologue SHARP functions as a positive regulator of Wnt/beta-catenin/T-cell factor signaling in neoplastic transformation. Cancer Res. 2007;67:482–491. doi: 10.1158/0008-5472.CAN-06-2314. [DOI] [PubMed] [Google Scholar]
  • 282.Lin H.V., Doroquez D.B., Cho S., Chen F., Rebay I., Cadigan K.M. Splits ends is a tissue/promoter specific regulator of Wingless signaling. Development. 2003;130:3125–3135. doi: 10.1242/dev.00527. [DOI] [PubMed] [Google Scholar]
  • 283.Shen H., Zhao X., Chen J., Qu W., Huang X., Wang M., et al. O-GlcNAc transferase Ogt regulates embryonic neuronal development through modulating Wnt/β-catenin signaling. Hum. Mol. Genet. 2021;31:57–68. doi: 10.1093/hmg/ddab223. [DOI] [PubMed] [Google Scholar]
  • 284.Chen J., Dong X., Cheng X., Zhu Q., Zhang J., Li Q., et al. Ogt controls neural stem/progenitor cell pool and adult neurogenesis through modulating Notch signaling. Cell Rep. 2021;34 doi: 10.1016/j.celrep.2021.108905. [DOI] [PubMed] [Google Scholar]
  • 285.Muha V., Fenckova M., Ferenbach A.T., Catinozzi M., Eidhof I., Storkebaum E., et al. O-GlcNAcase contributes to cognitive function in Drosophila. J. Biol. Chem. 2020;295:8636–8646. doi: 10.1074/jbc.RA119.010312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Castro V.L., Quintana A.M. The role of HCFC1 in syndromic and non-syndromic intellectual disability. Med. Res. Arch. 2020;8 doi: 10.18103/mra.v8i6.2122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Tarpey P.S., Smith R., Pleasance E., Whibley A., Edkins S., Hardy C., et al. A systematic, large-scale resequencing screen of X-chromosome coding exons in mental retardation. Nat Genet. 2009;41:535–543. doi: 10.1038/ng.367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Huang L., Jolly L.A., Willis-Owen S., Gardner A., Kumar R., Douglas E., et al. A noncoding, regulatory mutation implicates HCFC1 in nonsyndromic intellectual disability. Am. J. Hum. Genet. 2012;91:694–702. doi: 10.1016/j.ajhg.2012.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.Koufaris C., Alexandrou A., Tanteles G.A., Anastasiadou V., Sismani C. A novel HCFC1 variant in male siblings with intellectual disability and microcephaly in the absence of cobalamin disorder. Biomed. Rep. 2016;4:215–218. doi: 10.3892/br.2015.559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Kapuria V., Röhrig U.F., Waridel P., Lammers F., Borodkin V.S., van Aalten D.M.F., et al. The conserved threonine-rich region of the HCF-1PRO repeat activates promiscuous OGT:UDP-GlcNAc glycosylation and proteolysis activities. J. Biol. Chem. 2018;293:17754–17768. doi: 10.1074/jbc.RA118.004185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Janetzko J., Trauger S.A., Lazarus M.B., Walker S. How the glycosyltransferase OGT catalyzes amide bond cleavage. Nat. Chem. Biol. 2016;12:899–901. doi: 10.1038/nchembio.2173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Kapuria V., Röhrig U.F., Bhuiyan T., Borodkin V.S., van Aalten D.M.F., Zoete V., et al. Proteolysis of HCF-1 by Ser/Thr glycosylation-incompetent O-GlcNAc transferase:UDP-GlcNAc complexes. Genes Dev. 2016;30:960–972. doi: 10.1101/gad.275925.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Deplus R., Delatte B., Schwinn M.K., Defrance M., Méndez J., Murphy N., et al. TET2 and TET3 regulate GlcNAcylation and H3K4 methylation through OGT and SET1/COMPASS. EMBO J. 2013;32:645–655. doi: 10.1038/emboj.2012.357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Fasken M.B., Corbett A.H., Stewart M. Structure-function relationships in the Nab2 polyadenosine-RNA binding Zn finger protein family. Protein Sci. 2019;28:513–523. doi: 10.1002/pro.3565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Morris K.J., Corbett A.H. The polyadenosine RNA-binding protein ZC3H14 interacts with the THO complex and coordinately regulates the processing of neuronal transcripts. Nucleic Acids Res. 2018;46:6561–6575. doi: 10.1093/nar/gky446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Soucek S., Zeng Y., Bellur D.L., Bergkessel M., Morris K.J., Deng Q., et al. The evolutionarily-conserved polyadenosine RNA binding protein, Nab2, cooperates with splicing machinery to regulate the fate of pre-mRNA. Mol. Cell Biol. 2016;36:2697–2714. doi: 10.1128/MCB.00402-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 297.Grenier St-Sauveur V., Soucek S., Corbett A.H., Bachand F. Poly(A) tail-mediated gene regulation by opposing roles of Nab2 and Pab2 nuclear poly(A)-binding proteins in pre-mRNA decay. Mol. Cell Biol. 2013;33:4718–4731. doi: 10.1128/MCB.00887-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Kelly S., Pak C., Garshasbi M., Kuss A., Corbett A.H., Moberg K. New kid on the ID block: neural functions of the Nab2/ZC3H14 class of Cys₃His tandem zinc-finger polyadenosine RNA binding proteins. RNA Biol. 2012;9:555–562. doi: 10.4161/rna.20187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Corgiat E.B., List S.M., Rounds J.C., Yu D., Chen P., Corbett A.H., et al. The Nab2 RNA-binding protein patterns dendritic and axonal projections through a planar cell polarity-sensitive mechanism. G3 (Bethesda) 2022;12 doi: 10.1093/g3journal/jkac100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Corgiat E.B., List S.M., Rounds J.C., Corbett A.H., Moberg K.H. The RNA-binding protein Nab2 regulates the proteome of the developing Drosophila brain. J. Biol. Chem. 2021;297 doi: 10.1016/j.jbc.2021.100877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Rounds J.C., Corgiat E.B., Ye C., Behnke J.A., Kelly S.M., Corbett A.H., et al. The disease-associated proteins Drosophila Nab2 and Ataxin-2 interact with shared RNAs and coregulate neuronal morphology. Genetics. 2022;220 doi: 10.1093/genetics/iyab175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 302.Bienkowski R.S., Banerjee A., Rounds J.C., Rha J., Omotade O.F., Gross C., et al. The conserved, disease-associated RNA binding protein dNab2 interacts with the Fragile X protein ortholog in Drosophila neurons. Cell Rep. 2017;20:1372–1384. doi: 10.1016/j.celrep.2017.07.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Rha J., Jones S.K., Fidler J., Banerjee A., Leung S.W., Morris K.J., et al. The RNA-binding protein, ZC3H14, is required for proper poly(A) tail length control, expression of synaptic proteins, and brain function in mice. Hum. Mol. Genet. 2017;26:3663–3681. doi: 10.1093/hmg/ddx248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Fasken M.B., Corbett A.H. Links between mRNA splicing, mRNA quality control, and intellectual disability. RNA Dis. 2016;3 [PMC free article] [PubMed] [Google Scholar]
  • 305.Kelly S.M., Bienkowski R., Banerjee A., Melicharek D.J., Brewer Z.A., Marenda D.R., et al. The Drosophila ortholog of the Zc3h14 RNA binding protein acts within neurons to pattern axon projection in the developing brain. Dev. Neurobiol. 2016;76:93–106. doi: 10.1002/dneu.22301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 306.Pak C., Garshasbi M., Kahrizi K., Gross C., Apponi L.H., Noto J.J., et al. Mutation of the conserved polyadenosine RNA binding protein, ZC3H14/dNab2, impairs neural function in Drosophila and humans. Proc. Natl. Acad. Sci. U. S. A. 2011;108:12390–12395. doi: 10.1073/pnas.1107103108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Singh U., Bindra D., Samaiya A., Mishra R.K. Overexpressed Nup88 stabilized through interaction with Nup62 promotes NF-κB dependent pathways in cancer. Front. Oncol. 2023;13 doi: 10.3389/fonc.2023.1095046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Madheshiya P.K., Shukla E., Singh J., Bawaria S., Ansari M.Y., Chauhan R. Insights into the role of Nup62 and Nup93 in assembling cytoplasmic ring and central transport channel of the nuclear pore complex. Mol. Biol. Cell. 2022;33:ar139. doi: 10.1091/mbc.E22-01-0027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 309.Fontana P., Dong Y., Pi X., Tong A.B., Hecksel C.W., Wang L., et al. Structure of cytoplasmic ring of nuclear pore complex by integrative cryo-EM and AlphaFold. Science. 2022;376 doi: 10.1126/science.abm9326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Huang G., Zhan X., Zeng C., Liang K., Zhu X., Zhao Y., et al. Cryo-EM structure of the inner ring from the Xenopus laevis nuclear pore complex. Cell Res. 2022;32:451–460. doi: 10.1038/s41422-022-00633-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311.Zhu Y., Liu T.-W., Madden Z., Yuzwa S.A., Murray K., Cecioni S., et al. Post-translational O-GlcNAcylation is essential for nuclear pore integrity and maintenance of the pore selectivity filter. J. Mol. Cell Biol. 2016;8:2–16. doi: 10.1093/jmcb/mjv033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Meister G., Landthaler M., Peters L., Chen P.Y., Urlaub H., Lührmann R., et al. Identification of novel argonaute-associated proteins. Curr. Biol. 2005;15:2149–2155. doi: 10.1016/j.cub.2005.10.048. [DOI] [PubMed] [Google Scholar]
  • 313.Johnson S.T., Chu Y., Liu J., Corey D.R. Impact of scaffolding protein TNRC6 paralogs on gene expression and splicing. RNA. 2021;27:1004–1016. doi: 10.1261/rna.078709.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 314.Liu Z., Johnson S.T., Zhang Z., Corey D.R. Expression of TNRC6 (GW182) proteins is not necessary for gene silencing by fully complementary RNA duplexes. Nucleic Acid Ther. 2019;29:323–334. doi: 10.1089/nat.2019.0815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Liu J., Liu Z., Corey D.R. The requirement for GW182 scaffolding protein depends on whether argonaute is mediating translation, transcription, or splicing. Biochemistry. 2018;57:5247–5256. doi: 10.1021/acs.biochem.8b00602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Wakiyama M., Ogami K., Iwaoka R., Aoki K., Hoshino S.-I. MicroRNP-mediated translational activation of nonadenylated mRNAs in a mammalian cell-free system. Genes Cells. 2018;23:332–344. doi: 10.1111/gtc.12580. [DOI] [PubMed] [Google Scholar]
  • 317.Baillat D., Shiekhattar R. Functional dissection of the human TNRC6 (GW182-related) family of proteins. Mol. Cell Biol. 2009;29:4144–4155. doi: 10.1128/MCB.00380-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 318.Lazzaretti D., Tournier I., Izaurralde E. The C-terminal domains of human TNRC6A, TNRC6B, and TNRC6C silence bound transcripts independently of Argonaute proteins. RNA. 2009;15:1059–1066. doi: 10.1261/rna.1606309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Zipprich J.T., Bhattacharyya S., Mathys H., Filipowicz W. Importance of the C-terminal domain of the human GW182 protein TNRC6C for translational repression. RNA. 2009;15:781–793. doi: 10.1261/rna.1448009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Rybak A., Fuchs H., Hadian K., Smirnova L., Wulczyn E.A., Michel G., et al. The let-7 target gene mouse lin-41 is a stem cell specific E3 ubiquitin ligase for the miRNA pathway protein Ago2. Nat. Cell Biol. 2009;11:1411–1420. doi: 10.1038/ncb1987. [DOI] [PubMed] [Google Scholar]
  • 321.Lessel D., Zeitler D.M., Reijnders M.R.F., Kazantsev A., Hassani Nia F., Bartholomäus A., et al. Germline AGO2 mutations impair RNA interference and human neurological development. Nat. Commun. 2020;11:5797. doi: 10.1038/s41467-020-19572-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Mao Y., Shen G., Su Z., Du J., Xu F., Yu Y. RAD21 inhibited transcription of tumor suppressor MIR4697HG and led to glioma tumorigenesis. Biomed. Pharmacother. 2020;123 doi: 10.1016/j.biopha.2019.109759. [DOI] [PubMed] [Google Scholar]
  • 323.Córdova-Fletes C., Domínguez M.G., Delint-Ramirez I., Martínez-Rodríguez H.G., Rivas-Estilla A.M., Barros-Núñez P., et al. A de novo t(10;19)(q22.3;q13.33) leads to ZMIZ1/PRR12 reciprocal fusion transcripts in a girl with intellectual disability and neuropsychiatric alterations. Neurogenetics. 2015;16:287–298. doi: 10.1007/s10048-015-0452-2. [DOI] [PubMed] [Google Scholar]
  • 324.Howerton C.L., Morgan C.P., Fischer D.B., Bale T.L. O-GlcNAc transferase (OGT) as a placental biomarker of maternal stress and reprogramming of CNS gene transcription in development. Proc. Natl. Acad. Sci. U. S. A. 2013;110:5169–5174. doi: 10.1073/pnas.1300065110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 325.Sidor C., Borreguero-Munoz N., Fletcher G.C., Elbediwy A., Guillermin O., Thompson B.J. Mask family proteins ANKHD1 and ANKRD17 regulate YAP nuclear import and stability. Elife. 2019;8 doi: 10.7554/eLife.48601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Dong L., Lin F., Wu W., Huang W., Cai Z. Transcriptional cofactor Mask2 is required for YAP-induced cell growth and migration in bladder cancer cell. J. Cancer. 2016;7:2132–2138. doi: 10.7150/jca.16438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Zhai L., Yang X., Dong J., Qian L., Gao Y., Lv Y., et al. O-GlcNAcylation mediates endometrial cancer progression by regulating the Hippo-YAP pathway. Int. J. Oncol. 2023;63:90. doi: 10.3892/ijo.2023.5538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Peng C., Zhu Y., Zhang W., Liao Q., Chen Y., Zhao X., et al. Regulation of the Hippo-YAP pathway by glucose Sensor O-GlcNAcylation. Mol Cell. 2017;68:591–604.e5. doi: 10.1016/j.molcel.2017.10.010. [DOI] [PubMed] [Google Scholar]
  • 329.Collier A., Liu A., Torkelson J., Pattison J., Gaddam S., Zhen H., et al. Gibbin mesodermal regulation patterns epithelial development. Nature. 2022;606:188–196. doi: 10.1038/s41586-022-04727-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Konzman D., Abramowitz L.K., Steenackers A., Mukherjee M.M., Na H.-J., Hanover J.A. O-GlcNAc: regulator of signaling and epigenetics linked to X-linked intellectual disability. Front. Genet. 2020;11 doi: 10.3389/fgene.2020.605263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Vella P., Scelfo A., Jammula S., Chiacchiera F., Williams K., Cuomo A., et al. Tet proteins connect the O-linked N-acetylglucosamine transferase Ogt to chromatin in embryonic stem cells. Mol. Cell. 2013;49:645–656. doi: 10.1016/j.molcel.2012.12.019. [DOI] [PubMed] [Google Scholar]
  • 332.Bauer C., Göbel K., Nagaraj N., Colantuoni C., Wang M., Müller U., et al. Phosphorylation of TET proteins is regulated via O-GlcNAcylation by the O-linked N-acetylglucosamine transferase (OGT) J. Biol. Chem. 2015;290:4801–4812. doi: 10.1074/jbc.M114.605881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Hahn M.A., Jin S.-G., Li A.X., Liu J., Huang Z., Wu X., et al. Reprogramming of DNA methylation at NEUROD2-bound sequences during cortical neuron differentiation. Sci. Adv. 2019;5 doi: 10.1126/sciadv.aax0080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Hahn M.A., Qiu R., Wu X., Li A.X., Zhang H., Wang J., et al. Dynamics of 5-hydroxymethylcytosine and chromatin marks in Mammalian neurogenesis. Cell Rep. 2013;3:291–300. doi: 10.1016/j.celrep.2013.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.He F., Wu H., Zhou L., Lin Q., Cheng Y., Sun Y.E. Tet2-mediated epigenetic drive for astrocyte differentiation from embryonic neural stem cells. Cell Death Discov. 2020;6:30. doi: 10.1038/s41420-020-0264-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Kurdyukov S., Bullock M. DNA methylation analysis: choosing the right method. Biology (Basel). 2016;5:3. doi: 10.3390/biology5010003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Gillentine M.A., Wang T., Hoekzema K., Rosenfeld J., Liu P., Guo H., et al. Rare deleterious mutations of HNRNP genes result in shared neurodevelopmental disorders. Genome Med. 2021;13:63. doi: 10.1186/s13073-021-00870-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 338.Geuens T., Bouhy D., Timmerman V. The hnRNP family: insights into their role in health and disease. Hum. Genet. 2016;135:851–867. doi: 10.1007/s00439-016-1683-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 339.Brasil S., Pascoal C., Francisco R., Marques-da-Silva D., Andreotti G., Videira P.A., et al. CDG therapies: from bench to bedside. Int. J. Mol. Sci. 2018;19:1304. doi: 10.3390/ijms19051304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 340.Muha V., Authier F., Szoke-Kovacs Z., Johnson S., Gallagher J., McNeilly A., et al. Loss of O-GlcNAcase catalytic activity leads to defects in mouse embryogenesis. J. Biol. Chem. 2021;296 doi: 10.1016/j.jbc.2021.100439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 341.Küry S., Ebstein F., Mollé A., Besnard T., Lee M.-K., Vignard V., et al. Rare germline heterozygous missense variants in BRCA1-associated protein 1, BAP1, cause a syndromic neurodevelopmental disorder. Am. J. Hum. Genet. 2022;109:361–372. doi: 10.1016/j.ajhg.2021.12.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Zhu X., Petrovski S., Xie P., Ruzzo E.K., Lu Y.-F., McSweeney K.M., et al. Whole-exome sequencing in undiagnosed genetic diseases: interpreting 119 trios. Genet. Med. 2015;17:774–781. doi: 10.1038/gim.2014.191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 343.Machol K., Rousseau J., Ehresmann S., Garcia T., Nguyen T.T.M., Spillmann R.C., et al. Expanding the spectrum of BAF-related disorders: de novo variants in SMARCC2 cause a syndrome with intellectual disability and developmental delay. Am. J. Hum. Genet. 2019;104:164–178. doi: 10.1016/j.ajhg.2018.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 344.Martínez F., Caro-Llopis A., Roselló M., Oltra S., Mayo S., Monfort S., et al. High diagnostic yield of syndromic intellectual disability by targeted next-generation sequencing. J. Med. Genet. 2017;54:87–92. doi: 10.1136/jmedgenet-2016-103964. [DOI] [PubMed] [Google Scholar]
  • 345.Saitsu H., Tohyama J., Walsh T., Kato M., Kobayashi Y., Lee M., et al. A girl with West syndrome and autistic features harboring a de novo TBL1XR1 mutation. J. Hum. Genet. 2014;59:581–583. doi: 10.1038/jhg.2014.71. [DOI] [PubMed] [Google Scholar]
  • 346.O’Roak B.J., Vives L., Girirajan S., Karakoc E., Krumm N., Coe B.P., et al. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature. 2012;485:246–250. doi: 10.1038/nature10989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 347.O’Roak B.J., Vives L., Fu W., Egertson J.D., Stanaway I.B., Phelps I.G., et al. Multiplex targeted sequencing identifies recurrently mutated genes in autism spectrum disorders. Science. 2012;338:1619–1622. doi: 10.1126/science.1227764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 348.Armour C.M., Smith A., Hartley T., Chardon J.W., Sawyer S., Schwartzentruber J., et al. Syndrome disintegration: exome sequencing reveals that Fitzsimmons syndrome is a co-occurrence of multiple events. Am. J. Med. Genet. A. 2016;170:1820–1825. doi: 10.1002/ajmg.a.37684. [DOI] [PubMed] [Google Scholar]
  • 349.Heinen C.A., Jongejan A., Watson P.J., Redeker B., Boelen A., Boudzovitch-Surovtseva O., et al. A specific mutation in TBL1XR1 causes Pierpont syndrome. J. Med. Genet. 2016;53:330–337. doi: 10.1136/jmedgenet-2015-103233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 350.Kahlert A.-K., Weidensee S., Mackenroth L., Porrmann J., Rump A., Di Donato N., et al. Pierpont syndrome: report of a new patient. Clin. Dysmorphol. 2017;26:205–208. doi: 10.1097/MCD.0000000000000184. [DOI] [PubMed] [Google Scholar]
  • 351.Tesarova M., Baxova A., Hansikova H., Lambert L., Vondrackova A., Leiska A., et al. Pierpont syndrome due to mutation c.1337A>G in TBL1XR1 gene. Clin. Dysmorphol. 2022;31:145–148. doi: 10.1097/MCD.0000000000000416. [DOI] [PubMed] [Google Scholar]
  • 352.Lemattre C., Thevenon J., Duffourd Y., Nambot S., Haquet E., Vuadelle B., et al. TBL1XR1 mutations in Pierpont syndrome are not restricted to the recurrent p.Tyr446Cys mutation. Am. J. Med. Genet. A. 2018;176:2813–2818. doi: 10.1002/ajmg.a.40510. [DOI] [PubMed] [Google Scholar]
  • 353.van der Sluijs P.J., Jansen S., Vergano S.A., Adachi-Fukuda M., Alanay Y., AlKindy A., et al. The ARID1B spectrum in 143 patients: from nonsyndromic intellectual disability to Coffin–Siris syndrome. Genet. Med. 2019;21:1295–1307. doi: 10.1038/s41436-018-0330-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 354.Hoyer J., Ekici A.B., Endele S., Popp B., Zweier C., Wiesener A., et al. Haploinsufficiency of ARID1B, a member of the SWI/SNF-a chromatin-remodeling complex, is a frequent cause of intellectual disability. Am. J. Hum. Genet. 2012;90:565–572. doi: 10.1016/j.ajhg.2012.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 355.Tsurusaki Y., Okamoto N., Ohashi H., Kosho T., Imai Y., Hibi-Ko Y., et al. Mutations affecting components of the SWI/SNF complex cause Coffin-Siris syndrome. Nat. Genet. 2012;44:376–378. doi: 10.1038/ng.2219. [DOI] [PubMed] [Google Scholar]
  • 356.Santen G.W.E., Aten E., Sun Y., Almomani R., Gilissen C., Nielsen M., et al. Mutations in SWI/SNF chromatin remodeling complex gene ARID1B cause Coffin-Siris syndrome. Nat. Genet. 2012;44:379–380. doi: 10.1038/ng.2217. [DOI] [PubMed] [Google Scholar]
  • 357.Zweier M., Peippo M.M., Pöyhönen M., Kääriäinen H., Begemann A., Joset P., et al. The HHID syndrome of hypertrichosis, hyperkeratosis, abnormal corpus callosum, intellectual disability, and minor anomalies is caused by mutations in ARID1B. Am. J. Med. Genet. A. 2017;173:1440–1443. doi: 10.1002/ajmg.a.38143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 358.Mavros C.F., Brownstein C.A., Thyagrajan R., Genetti C.A., Tembulkar S., Graber K., et al. De novo variant of TRRAP in a patient with very early onset psychosis in the context of non-verbal learning disability and obsessive-compulsive disorder: a case report. BMC Med. Genet. 2018;19:197. doi: 10.1186/s12881-018-0711-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 359.Cogné B., Ehresmann S., Beauregard-Lacroix E., Rousseau J., Besnard T., Garcia T., et al. Missense variants in the histone acetyltransferase complex component gene TRRAP cause autism and syndromic intellectual disability. Am. J. Hum. Genet. 2019;104:530–541. doi: 10.1016/j.ajhg.2019.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360.Menke L.A., DDD study. Gardeitchik T., Hammond P., Heimdal K.R., Houge G., et al. Further delineation of an entity caused by CREBBP and EP300 mutations but not resembling Rubinstein-Taybi syndrome. Am. J. Med. Genet. A. 2018;176:862–876. doi: 10.1002/ajmg.a.38626. [DOI] [PubMed] [Google Scholar]
  • 361.Roelfsema J.H., White S.J., Ariyürek Y., Bartholdi D., Niedrist D., Papadia F., et al. Genetic heterogeneity in Rubinstein-Taybi syndrome: mutations in both the CBP and EP300 genes cause disease. Am. J. Hum. Genet. 2005;76:572–580. doi: 10.1086/429130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 362.Zimmermann N., Acosta A.M.B.F., Kohlhase J., Bartsch O. Confirmation of EP300 gene mutations as a rare cause of Rubinstein-Taybi syndrome. Eur. J. Hum. Genet. 2007;15:837–842. doi: 10.1038/sj.ejhg.5201791. [DOI] [PubMed] [Google Scholar]
  • 363.Foley P., Bunyan D., Stratton J., Dillon M., Lynch S.A. Further case of Rubinstein-Taybi syndrome due to a deletion in EP300. Am. J. Med. Genet. A. 2009;149A:997–1000. doi: 10.1002/ajmg.a.32771. [DOI] [PubMed] [Google Scholar]
  • 364.Bartsch O., Labonté J., Albrecht B., Wieczorek D., Lechno S., Zechner U., et al. Two patients with EP300 mutations and facial dysmorphism different from the classic Rubinstein-Taybi syndrome. Am. J. Med. Genet. A. 2010;152A:181–184. doi: 10.1002/ajmg.a.33153. [DOI] [PubMed] [Google Scholar]
  • 365.Woods S.A., Robinson H.B., Kohler L.J., Agamanolis D., Sterbenz G., Khalifa M. Exome sequencing identifies a novel EP300 frame shift mutation in a patient with features that overlap Cornelia de Lange syndrome. Am. J. Med. Genet. A. 2014;164A:251–258. doi: 10.1002/ajmg.a.36237. [DOI] [PubMed] [Google Scholar]
  • 366.Hamilton M.J., Newbury-Ecob R., Holder-Espinasse M., Yau S., Lillis S., Hurst J.A., et al. Rubinstein-Taybi syndrome type 2: report of nine new cases that extend the phenotypic and genotypic spectrum. Clin. Dysmorphol. 2016;25:135–145. doi: 10.1097/MCD.0000000000000143. [DOI] [PubMed] [Google Scholar]
  • 367.Chowdhury F., Wang L., Al-Raqad M., Amor D.J., Baxová A., Bendová Š., et al. Haploinsufficiency of PRR12 causes a spectrum of neurodevelopmental, eye, and multisystem abnormalities. Genet. Med. 2021;23:1234–1245. doi: 10.1038/s41436-021-01129-6. [DOI] [PubMed] [Google Scholar]
  • 368.Reis L.M., Costakos D., Wheeler P.G., Bardakjian T., Schneider A., Fung S.S.M., et al. Dominant variants in PRR12 result in unilateral or bilateral complex microphthalmia. Clin. Genet. 2021;99:437–442. doi: 10.1111/cge.13897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 369.Leduc M.S., Mcguire M., Madan-Khetarpal S., Ortiz D., Hayflick S., Keller K., et al. De novo apparent loss-of-function mutations in PRR12 in three patients with intellectual disability and iris abnormalities. Hum. Genet. 2018;137:257–264. doi: 10.1007/s00439-018-1877-0. [DOI] [PubMed] [Google Scholar]
  • 370.Eising E., Carrion-Castillo A., Vino A., Strand E.A., Jakielski K.J., Scerri T.S., et al. A set of regulatory genes co-expressed in embryonic human brain is implicated in disrupted speech development. Mol. Psychiatry. 2019;24:1065–1078. doi: 10.1038/s41380-018-0020-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 371.Granadillo J.L., P A Stegmann A., Guo H., Xia K., Angle B., Bontempo K., et al. Pathogenic variants in TNRC6B cause a genetic disorder characterised by developmental delay/intellectual disability and a spectrum of neurobehavioural phenotypes including autism and ADHD. J. Med. Genet. 2020;57:717–724. doi: 10.1136/jmedgenet-2019-106470. [DOI] [PubMed] [Google Scholar]
  • 372.Shashi V., Pena L.D.M., Kim K., Burton B., Hempel M., Schoch K., et al. De novo truncating variants in ASXL2 are associated with a unique and recognizable clinical phenotype. Am. J. Hum. Genet. 2016;99:991–999. doi: 10.1016/j.ajhg.2016.08.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 373.Xia F., Bainbridge M.N., Tan T.Y., Wangler M.F., Scheuerle A.E., Zackai E.H., et al. De novo truncating mutations in AHDC1 in individuals with syndromic expressive language delay, hypotonia, and sleep apnea. Am. J. Hum. Genet. 2014;94:784–789. doi: 10.1016/j.ajhg.2014.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 374.Yang H., Douglas G., Monaghan K.G., Retterer K., Cho M.T., Escobar L.F., et al. De novo truncating variants in the AHDC1 gene encoding the AT-hook DNA-binding motif-containing protein 1 are associated with intellectual disability and developmental delay. Cold Spring Harb. Mol. Case Stud. 2015;1 doi: 10.1101/mcs.a000562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 375.García-Acero M., Acosta J. Whole-Exome Sequencing Identifies a de novo AHDC1 Mutation in a Colombian Patient with Xia-Gibbs Syndrome. Mol. Syndromol. 2017;8:308–312. doi: 10.1159/000479357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 376.Gumus E. Extending the phenotype of Xia-Gibbs syndrome in a two-year-old patient with craniosynostosis with a novel de novo AHDC1 missense mutation. Eur. J. Med. Genet. 2020;63 doi: 10.1016/j.ejmg.2019.03.001. [DOI] [PubMed] [Google Scholar]
  • 377.Daly T., Roberts A., Yang E., Mochida G.H., Bodamer O. Holoprosencephaly in Kabuki syndrome. Am. J. Med. Genet. A. 2020;182:441–445. doi: 10.1002/ajmg.a.61454. [DOI] [PubMed] [Google Scholar]
  • 378.Kim S.Y., Jang S.S., Kim H., Hwang H., Choi J.E., Chae J.-H., et al. Genetic diagnosis of infantile-onset epilepsy in the clinic: application of whole-exome sequencing following epilepsy gene panel testing. Clin. Genet. 2021;99:418–424. doi: 10.1111/cge.13903. [DOI] [PubMed] [Google Scholar]
  • 379.Verheijen J., Vanhye L., Niaz Shaikh A., Radenkovic S., Ranatunga W., Pichurin P., et al. P341: idenfication of novel variants and phenotypic expansion in OGT-CDG. Genet. Med. Open. 2023;1:100369. [Google Scholar]
  • 380.Kleefstra T., Kramer J.M., Neveling K., Willemsen M.H., Koemans T.S., Vissers L.E.L.M., et al. Disruption of an EHMT1-associated chromatin-modification module causes intellectual disability. Am. J. Hum. Genet. 2012;91:73–82. doi: 10.1016/j.ajhg.2012.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.Koemans T.S., Kleefstra T., Chubak M.C., Stone M.H., Reijnders M.R.F., de Munnik S., et al. Functional convergence of histone methyltransferases EHMT1 and KMT2C involved in intellectual disability and autism spectrum disorder. PLoS Genet. 2017;13 doi: 10.1371/journal.pgen.1006864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382.Hamanaka K., Sugawara Y., Shimoji T., Nordtveit T.I., Kato M., Nakashima M., et al. De novo truncating variants in PHF21A cause intellectual disability and craniofacial anomalies. Eur. J. Hum. Genet. 2019;27:378–383. doi: 10.1038/s41431-018-0289-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Singh T., Kurki M.I., Curtis D., Purcell S.M., Crooks L., McRae J., et al. Rare loss-of-function variants in SETD1A are associated with schizophrenia and developmental disorders. Nat. Neurosci. 2016;19:571–577. doi: 10.1038/nn.4267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 384.Narumi-Kishimoto Y., Araki N., Migita O., Kawai T., Okamura K., Nakabayashi K., et al. Novel SIN3A mutation identified in a Japanese patient with Witteveen-Kolk syndrome. Eur J Med Genet. 2019;62 doi: 10.1016/j.ejmg.2018.09.014. [DOI] [PubMed] [Google Scholar]
  • 385.Schuurs-Hoeijmakers J.H.M., Vulto-van Silfhout A.T., Vissers L.E.L.M., van de Vondervoort I.I.G.M., van Bon B.W.M., de Ligt J., et al. Identification of pathogenic gene variants in small families with intellectually disabled siblings by exome sequencing. J. Med. Genet. 2013;50:802–811. doi: 10.1136/jmedgenet-2013-101644. [DOI] [PubMed] [Google Scholar]
  • 386.Ylikallio E., Woldegebriel R., Tumiati M., Isohanni P., Ryan M.M., Stark Z., et al. MCM3AP in recessive Charcot-Marie-Tooth neuropathy and mild intellectual disability. Brain. 2017;140:2093–2103. doi: 10.1093/brain/awx138. [DOI] [PubMed] [Google Scholar]
  • 387.Karakaya M., Mazaheri N., Polat I., Bharucha-Goebel D., Donkervoort S., Maroofian R., et al. Biallelic MCM3AP mutations cause Charcot-Marie-Tooth neuropathy with variable clinical presentation. Brain. 2017;140:e65. doi: 10.1093/brain/awx222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 388.Kennerson M.L., Corbett A.C., Ellis M., Perez-Siles G., Nicholson G.A. A novel MCM3AP mutation in a Lebanese family with recessive Charcot-Marie-Tooth neuropathy. Brain. 2018;141:e66. doi: 10.1093/brain/awy184. [DOI] [PubMed] [Google Scholar]
  • 389.Menke L.A., van Belzen M.J., Alders M., Cristofoli F., DDD Study. Ehmke N., et al. CREBBP mutations in individuals without Rubinstein-Taybi syndrome phenotype. Am. J. Med. Genet. A. 2016;170:2681–2693. doi: 10.1002/ajmg.a.37800. [DOI] [PubMed] [Google Scholar]
  • 390.Angius A., Uva P., Oppo M., Persico I., Onano S., Olla S., et al. Confirmation of a new phenotype in an individual with a variant in the last part of exon 30 of CREBBP. Am. J. Med. Genet. A. 2019;179:634–638. doi: 10.1002/ajmg.a.61052. [DOI] [PubMed] [Google Scholar]
  • 391.Petrij F., Giles R.H., Dauwerse H.G., Saris J.J., Hennekam R.C., Masuno M., et al. Rubinstein-Taybi syndrome caused by mutations in the transcriptional co-activator CBP. Nature. 1995;376:348–351. doi: 10.1038/376348a0. [DOI] [PubMed] [Google Scholar]
  • 392.Murata T., Kurokawa R., Krones A., Tatsumi K., Ishii M., Taki T., et al. Defect of histone acetyltransferase activity of the nuclear transcriptional coactivator CBP in Rubinstein-Taybi syndrome. Hum. Mol. Genet. 2001;10:1071–1076. doi: 10.1093/hmg/10.10.1071. [DOI] [PubMed] [Google Scholar]
  • 393.Kalkhoven E., Roelfsema J.H., Teunissen H., den Boer A., Ariyurek Y., Zantema A., et al. Loss of CBP acetyltransferase activity by PHD finger mutations in Rubinstein-Taybi syndrome. Hum. Mol. Genet. 2003;12:441–450. doi: 10.1093/hmg/ddg039. [DOI] [PubMed] [Google Scholar]
  • 394.Yuan B., Pehlivan D., Karaca E., Patel N., Charng W.-L., Gambin T., et al. Global transcriptional disturbances underlie Cornelia de Lange syndrome and related phenotypes. J. Clin. Invest. 2015;125:636–651. doi: 10.1172/JCI77435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 395.Alazami A.M., Patel N., Shamseldin H.E., Anazi S., Al-Dosari M.S., Alzahrani F., et al. Accelerating novel candidate gene discovery in neurogenetic disorders via whole-exome sequencing of prescreened multiplex consanguineous families. Cell Rep. 2015;10:148–161. doi: 10.1016/j.celrep.2014.12.015. [DOI] [PubMed] [Google Scholar]
  • 396.Santen G.W.E., Aten E., Vulto-van Silfhout A.T., Pottinger C., van Bon B.W., van Minderhout I.J., et al. Coffin-Siris syndrome and the BAF complex: genotype-phenotype study in 63 patients. Hum. Mutat. 2013;34:1519–1528. doi: 10.1002/humu.22394. [DOI] [PubMed] [Google Scholar]
  • 397.Wieczorek D., Bögershausen N., Beleggia F., Steiner-Haldenstätt S., Pohl E., Li Y., et al. A comprehensive molecular study on Coffin-Siris and Nicolaides-Baraitser syndromes identifies a broad molecular and clinical spectrum converging on altered chromatin remodeling. Hum. Mol. Genet. 2013;22:5121–5135. doi: 10.1093/hmg/ddt366. [DOI] [PubMed] [Google Scholar]
  • 398.Miyake N., Tsurusaki Y., Koshimizu E., Okamoto N., Kosho T., Brown N.J., et al. Delineation of clinical features in Wiedemann-Steiner syndrome caused by KMT2A mutations. Clin. Genet. 2016;89:115–119. doi: 10.1111/cge.12586. [DOI] [PubMed] [Google Scholar]
  • 399.Jones W.D., Dafou D., McEntagart M., Woollard W.J., Elmslie F.V., Holder-Espinasse M., et al. De novo mutations in MLL cause Wiedemann-Steiner syndrome. Am. J. Hum. Genet. 2012;91:358–364. doi: 10.1016/j.ajhg.2012.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 400.Vissers L.E., de Ligt J., Gilissen C., Janssen I., Steehouwer M., de Vries P., et al. A de novo paradigm for mental retardation. Nat Genet. 2010;42:1109–1112. doi: 10.1038/ng.712. [DOI] [PubMed] [Google Scholar]
  • 401.Lu H.-C., Tan Q., Rousseaux M.W.C., Wang W., Kim J.-Y., Richman R., et al. Disruption of the ATXN1-CIC complex causes a spectrum of neurobehavioral phenotypes in mice and humans. Nat. Genet. 2017;49:527–536. doi: 10.1038/ng.3808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 402.Deciphering Developmental Disorders Study Large-scale discovery of novel genetic causes of developmental disorders. Nature. 2015;519:223–228. doi: 10.1038/nature14135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 403.White J., Beck C.R., Harel T., Posey J.E., Jhangiani S.N., Tang S., et al. POGZ truncating alleles cause syndromic intellectual disability. Genome Med. 2016;8:3. doi: 10.1186/s13073-015-0253-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 404.Tan B., Zou Y., Zhang Y., Zhang R., Ou J., Shen Y., et al. A novel de novo POGZ mutation in a patient with intellectual disability. J. Hum. Genet. 2016;61:357–359. doi: 10.1038/jhg.2015.156. [DOI] [PubMed] [Google Scholar]
  • 405.Assia Batzir N., Posey J.E., Song X., Akdemir Z.C., Rosenfeld J.A., Brown C.W., et al. Phenotypic expansion of POGZ-related intellectual disability syndrome (White-Sutton syndrome) Am. J. Med. Genet. A. 2020;182:38–52. doi: 10.1002/ajmg.a.61380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 406.Chopra M., McEntagart M., Clayton-Smith J., Platzer K., Shukla A., Girisha K.M., et al. Heterozygous ANKRD17 loss-of-function variants cause a syndrome with intellectual disability, speech delay, and dysmorphism. Am. J. Hum. Genet. 2021;108:1138–1150. doi: 10.1016/j.ajhg.2021.04.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 407.Glinton K.E., Hurst A.C.E., Bowling K.M., Cristian I., Haynes D., Adstamongkonkul D., et al. Phenotypic expansion of the BPTF-related neurodevelopmental disorder with dysmorphic facies and distal limb anomalies. Am. J. Med. Genet. A. 2021;185:1366–1378. doi: 10.1002/ajmg.a.62102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 408.Stankiewicz P., Khan T.N., Szafranski P., Slattery L., Streff H., Vetrini F., et al. Haploinsufficiency of the chromatin remodeler BPTF causes syndromic developmental and speech delay, postnatal microcephaly, and dysmorphic features. Am. J. Hum. Genet. 2017;101:503–515. doi: 10.1016/j.ajhg.2017.08.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 409.Radio F.C., Pang K., Ciolfi A., Levy M.A., Hernández-García A., Pedace L., et al. SPEN haploinsufficiency causes a neurodevelopmental disorder overlapping proximal 1p36 deletion syndrome with an episignature of X chromosomes in females. Am. J. Hum. Genet. 2021;108:502–516. doi: 10.1016/j.ajhg.2021.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 410.Basel-Vanagaite L., Muncher L., Straussberg R., Pasmanik-Chor M., Yahav M., Rainshtein L., et al. Mutated nup62 causes autosomal recessive infantile bilateral striatal necrosis. Ann. Neurol. 2006;60:214–222. doi: 10.1002/ana.20902. [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Biological Chemistry are provided here courtesy of American Society for Biochemistry and Molecular Biology

RESOURCES