Skip to main content
Science Advances logoLink to Science Advances
. 2022 Jan 14;8(2):eabm2059. doi: 10.1126/sciadv.abm2059

How can same-gene mutations promote both cancer and developmental disorders?

Ruth Nussinov 1,2,*, Chung-Jung Tsai 1, Hyunbum Jang 1
PMCID: PMC8759737  PMID: 35030014

This review offers why same-gene mutations can lead to cancer and neurodevelopmental disorders and why a high risk of cancer.

Abstract

The question of how same-gene mutations can drive both cancer and neurodevelopmental disorders has been puzzling. It has also been puzzling why those with neurodevelopmental disorders have a high risk of cancer. Ras, MEK, PI3K, PTEN, and SHP2 are among the oncogenic proteins that can harbor mutations that encode diseases other than cancer. Understanding why some of their mutations can promote cancer, whereas others promote neurodevelopmental diseases, and why even the same mutations may promote both phenotypes, has important clinical ramifications. Here, we review the literature and address these tantalizing questions. We propose that cell type–specific expression of the mutant protein, and of other proteins in the respective pathway, timing of activation (during embryonic development or sporadic emergence), and the absolute number of molecules that the mutations activate, alone or in combination, are pivotal in determining the pathological phenotypes—cancer and (or) developmental disorders.

INTRODUCTION: THE PUZZLES

Among the puzzling questions confronting basic and translational research is how certain oncogenic mutations can promote cancer, while other mutations of the same proteins, or even the same mutations, can provoke other syndromes. Among the syndromes are RASopathies (1), a group of genetic neurodevelopmental disorders arising from mutations in mitogen-activated protein kinase (MAPK) pathway–related proteins (2), recently shown to share traits with autism spectrum disorder (ASD) (3), cerebral palsy (4), and more. The question of why patients with neurodevelopmental disorders are more prone to cancer has also been enigmatic. Here, we suggest that these seemingly puzzling observations can be understood based on four key considerations: (i) the level of expression of the protein in the type-specific cell, (ii) the potency of the single or combined mutations, (iii) the time window (expression levels in a specific cell type vary, e.g., during embryonic development or in the adult differentiated state), and (iv) the expression levels of the nodes in the pathway that the signal flows through. Often, the mutations are considered independently of the mRNA levels. However, the number of activated molecules depends on protein expression, and signal propagation requires that the expression levels of all proteins in the pathway not be too low. Notably, even without the mutations, very high expression levels can drive cancer. Increased levels of signaling can also drive neurodevelopmental disorders (5).

Recently, there has been a rapid hike in the number of reports relating to the linkage between neurodevelopmental disorders and cancer. These include many excellent reviews on the various aspects of neurodevelopmental disorders, such as RASopathy including, e.g., neurofibromatosis type 1 (NF1), Noonan syndrome (NS), Costello syndrome (CS) [e.g., (623)], PIK3CA-related overgrowth spectrum (PROS) [e.g., (2426)], ASD [e.g., (2733)], cerebral palsy, and more [e.g., (3437)], and one uniquely also covering genetically engineered mouse models to study cancer or congenital disorders (38). The linkage between developmental signaling cascades and aggressive central nervous system (CNS) tumors has also been reviewed [e.g., (37, 39)]. Much progress has been made and vast amounts of experimental and clinical data have accumulated. Yet, some key enigmatic questions are still waiting to be deciphered.

NEURODEVELOPMENTAL DISORDERS

Most of the disorders arise from changes in the development of the nervous system (40). Transcriptomics of the developing brain revealed that autism risk genes are highly expressed in prefrontal and primary motor cortices, striatum, cerebellum, and medial dorsal nucleus of the thalamus, particularly in early and midfetal development. Connective tissue dysplasia is common in NF1 microdeletions with a potential for associated cardiac manifestation. NF1 patients have higher chances of CNS tumors (gliomas and astrocytomas), neurofibrosarcomas, and leukemias than the normal population, suggesting dysregulated expression in these tissues (41). Cerebral palsy is associated with connective tissue in the gastrocnemius muscle (plural gastrocnemii), a superficial two-headed muscle that is in the back part of the lower leg (42). Germline mutations in Ras signaling pathways that relate to different familial, developmental syndromes frequently share phenotypic cardiofaciocutaneous features. Proteins involved in neurodevelopmental disorders relate to common pathways, including synaptic plasticity/function, chromatin remodelers, and the phosphoinositide 3-kinase (PI3K)/protein kinase B (AKT)/mechanistic target of rapamycin (mTOR) (43) and MAPK pathways, revealing RNA enrichment of genes involved in neuronal development, chromatin dynamics, and cell cycle regulation. RASopathies include NF1, NS, NSML (NS with multiple lentigines, formerly known as LEOPARD syndrome), CS, LS (Legius syndrome), CFC (cardiofaciocutaneous) syndrome, CM-AVM (capillary malformation–arteriovenous malformation) syndrome, and autosomal dominant intellectual disability type 5. They also present an increased risk of cancer (44). The mutations occur in Ras proteins, their effectors, e.g., B-Raf, regulators, e.g., guanine nucleotide exchange factors (GEFs), guanosine triphosphatase (GTPase)–activating proteins (GAPs), associated scaffolding proteins, phosphatases, e.g., Src homology 2 (SH2) domain–containing phosphatase 2 (SHP2), ubiquitin ligases, and pathway inhibitors. They are also observed in epidermal growth factor receptor (EGFR) (45), extracellular signal–regulated kinase (ERK) (46), and other members of the Ras superfamily, e.g., Rho in cerebral palsy (4749), Rac3 (50), RalA (51), and Rab (52), and their regulators (53), e.g., RhoGAP in schizophrenia risk (5456). The disorders emerge from germline mutations during embryonic development, with their timing and location critically influencing the resulting disease phenotype. Mutations associated with neurodevelopmental disorders also include oncogenic proteins in other pathways, such as overgrowth disorder involving PI3Kα (57), autism, a neurological developmental disorder involving phosphatase and tensin homolog (PTEN) (58), and recently observed in cyclin-dependent kinases (CDKs) and their activating cyclin subunits in the cell cycle (59). Structural studies of the proteins suggest that the conformational and dynamic changes promoted by the mutations, including at the protein-protein interaction sites, can be relatively minor and apparently inconsequential as in case of Ras, or substantial as in EGFR, where the extracellular domain becomes unstable, negatively affecting EGF binding (45). The patterns of direct interactions of the mutated protein with other proteins may, as in SHP2 (Fig. 1) (60, 61), or may not be affected by the mutations, leaving the seemingly baffling question of same protein with—mostly different but also the same—mutations, different disease (1) unanswered. Why patients with neurodevelopmental disorders have higher chances of developing cancer has also been enigmatic. Here, we focus on these questions.

Fig. 1. SHP2-associated Noonan syndrome in RASopathy.

Fig. 1.

(Top) Domain structure of SHP2 and its activation by disease-associated mutations. SHP2 contains nSH2 and cSH2 domains in the N-terminal region, and a phosphatase [protein tyrosine phosphatase (PTP)] and an unstructured tail in the C-terminal region. The C-terminal tail contains two tyrosine phosphorylation sites (Tyr542 and Tyr580) and a proline-rich domain with the PxxP motif. Upon upstream stimulation, the phosphorylated tyrosine (pY) motif of receptor tyrosine kinase (RTK) targets the SH2 domains of SHP2, recruiting it to the plasma membrane and releasing the autoinhibition of SHP2. Adapter proteins, such as Grb2-associated binding protein 1 (GAB1), also provide the pY motif, leading to recruitment of SHP2. Inactive SHP2 exhibits low basal activity owing to the autoinhibitory interaction between the nSH2 and PTP domains. Mutations in SHP2 causing human diseases, such as cancers or neurodevelopmental disorders, lead to gain of function. (Middle) These mutations mainly cluster in the nSH2-PTP interface, abolishing the inhibitory interaction. The locations and types of the germline mutations related to the NS and LEOPARD (LPRD) syndrome (also known as NS with multiple lentigines) are not the same as those responsible for the cancers, juvenile myelomonocytic leukemia (JMML), and acute myeloid leukemia (AML), although the same residues are mutated. (Bottom left) Inactive SHP2 does not contribute to the MAPK pathway. (Bottom right) Active SHP2 mutants with open conformation can bind to RTK or GAB1, promoting catalytic activity. SHP2 dephosphorylates the Ras GTPase-activating protein (RasGAP) binding sites in RTK, preventing RasGAP recruitment to the plasma membrane (133, 134). This negative regulation of RasGAP leads to increase of Ras activity, and thus MAPK signaling. NS and leukemia-associated SHP2 mutants increase MAPK signaling, while LPRD syndrome–associated mutants decrease PTP activity, with both enzymatically activating and inactivating mutations in PTPN11 unexpectedly resulting in neurodevelopmental disorders with overlapping clinical phenotypes (127). PDB, Protein Data Bank.

CANCER EVOLUTION AND NEURODEVELOPMENTAL PATHOLOGIES

Cancer is expressed by uncontrolled cell proliferation. Cancer evolution requires (i) that the cell type–specific expression level of the associated protein be above a certain threshold, and the additive contributions of the oncogenic mutations be potent, together leading to a high population of activated mutant molecules, and (ii) that the activation signal propagates downstream. This requires that the cell type–specific expression levels of all nodes (proteins) in the pathway should not be low. If low, signaling in that pathway would not go through and another protein from a parallel pathway might be recruited (62), leading to network rewiring. Neurodevelopmental pathologies require (i) a high expression level of the protein in the specific embryonic brain cell type (63) and potent mutations, single or in combination, and (ii) propagation of the signal downstream (Fig. 2). For this, all nodes in the pathway should be sufficiently highly expressed in that cell. The expression levels of the proteins are driven by chromatin accessibility and signaling in the specific cell (64). Pathological and nonpathological brain cell type–specific profiles of the mRNA are critical (65), particularly in the striatum, an essential relay of motor, cognitive, and limbic processes, suggesting highly specific roles in cell subpopulations (66). This scenario is supported by analysis of the mRNA of brain cells, which observed gene expression changes linked to ASD (67, 68) and other neurodevelopmental disorders, such as bipolar disorders (69). The phenotypes were associated with neurons in the upper layers of the cortex and microglial activation genes, and the developmental transcription factor genes were most affected (70). It is also supported by SHP2-associated NS in RASopathy (Fig. 1), where binding partners mediating Ras signaling are abundant in excitatory but not inhibitory neurons, also pointing to cell type–specific changes in gene expression and the signaling network (61). Transcriptional dysregulation of trophic signaling pathways was documented in cerebral palsy patient-derived cell lines from an unselected cohort of 182 affected individuals (4). High expression levels require chromatin accessibility of the associated genes (64). The involvement of chromatin remodeling is supported by improved viability by histone deacetylase inhibitors in most RASopathy cell lines as compared to several MAPK inhibitors (71). The mutations should be sufficiently strong to obtain a high population of the activated proteins. Mutation strength relates to its mechanism of protein activation (7275). In cancer, hyperactivation leads to cell proliferation; extremely vigorous activation signal can lead to oncogene-induced senescence (OIS). In development, it can lead to premature senescence (Fig. 2) (76). Cells with aborted cell cycle can promote developmental disorders.

Fig. 2. The cellular network and cell cycle.

Fig. 2.

Schematic diagrams of the networks and cell cycles of a wild-type cell, a cancer cell, and a mutant embryonic brain cell. Constitutive potent activation of the proteins is via single or combination of mutations (red circle and explosion shape). Mutations in proteins associated with the transcription machinery and chromatin remodeling can lead to high expression levels of these (red circle) proteins in the type-specific cell. The level of expression of the protein, coupled with the potency of activation by the mutations, results in a large number of active molecules, leading to cancer or neurodevelopmental disorder phenotypes, respectively. The time window—germline (in neurodevelopmental disorders) or somatic (in cancer)—of the mutation and cell type (embryonic brain cell or dedifferentiated cell) distinguish between them. In addition, the expression levels of the nodes in the pathway that the signal flows through, controlled by the chromatin remodeling proteins, should not be low to permit the signal to transmit. In cancer, the resulting hyperactivated signaling can lead to OIS. In development, they can lead to premature senescence. In embryonic cells, cell cycle arrest due to the premature senescence alters phenotypic traits toward the developmental disorders, including RASopathies. Germline embryonic neurodevelopmental disorder mutations can increase the risk of cancer by coupling with emerging somatic mutations. G1, S, G2, and M are stages in the cell cycle. In both cancer and neurodevelopmental disorders, signaling levels are critical in determining the disease consequences, and these levels are the outcome of (i) mutation strengths, (ii) expression levels, (iii) cell type, and (iv) timing window.

Multiple proteins are involved in cell cycle progress. Perturbations of their expression levels in specific cell types coupled with the signal strength, as can be measured by the absolute numbers of molecules (products) that the respective mutant protein(s) activate and propagation of the signal in the pathway, make them challenging to understand and especially to forecast. The different proteins in cancer and OIS, and in premature senescence in neurodevelopmental disorders such as RASopathy, suggest rewired signaling.

OIS AND PREMATURE SENESCENCE

Precursor cells divide until they reach a fully differentiated state; differentiation arrests proliferation, permanently exiting from the division cycle (77). Strong activity of CDK-cyclin in the G1 cell cycle phase supports the undifferentiated state; however, hyperactivated OIS can emerge as protection against genomic instability and DNA damage, blocking cell division cycles (78, 79). PI3K lipid kinase, which phosphorylates phosphatidylinositol 4,5-bisphosphate (PIP2) to phosphatidylinositol 3,4,5-trisphosphate (PIP3), a signaling lipid required for signal transduction via the PI3K/AKT/mTOR pathway, can provide an example (Fig. 3). The number of PIP3 molecules produced by PI3K reflects its expression level and the potency of the mutations. A very high number of PIP3 and a consequent vigorous signaling can lead to cell proliferation. However, too high numbers by hyperactivated PI3K—generated by multiple PI3K mutations, or with collaborating mutant proteins—can lead to short signaling duration due to OIS. Mutations in both PI3K and PTEN, a tumor suppressor that dephosphorylates PIP3 to PIP2, can result in extremely high levels of PIP3 and such an outcome.

Fig. 3. Mutations and expression levels of PI3K and PTEN in different cell lines.

Fig. 3.

PI3K and PTEN signaling in cancer (top) and PI3K (middle) and PTEN (bottom) signaling in neurodevelopmental disorders. Oncogenic and neurodevelopmental driver mutations powerfully activate PI3K, increasing the production of signaling lipid PIP3. PTEN dephosphorylates PIP3 back to PIP2, impairing its tumor suppressor activity and leading to the same outcome. High expression levels of PI3K with embryonic postzygotic mutations in specific brain (nerve) cells can lead to PROS, which encompasses a range of unique clinical entities, but with a continuum and overlapping diagnoses (135). High expression levels of PTEN with embryonic germline mutations in different brain cell types can lead to ASD. The tumor suppressor mutants can present the PTEN hamartoma tumor syndrome (PHTS), with variable phenotypes, such as those diagnosed with autism or cancer.

OIS is a robust physiological antitumor response. Developmental senescence commonly takes place in aging and can also be recruited as a contributor to neurodevelopmental disorders (44, 76, 80, 81). A disorder regime requires high expression level of the mutant protein in the specific neuron cell and that the expression of nodes in the signaling pathway not be too low. If too low, or the level of a node in a parallel pathway is high, cell type–specific pathway branching or rewiring can take place (62). An activated ERK can regulate targets in the cytosol or translocate to the nucleus to phosphorylate a range of transcription factors altering gene expression. SHP2 branching of the Ras signaling pathway involves proteins that are populated in excitatory but not inhibitory neurons, rewiring the pathway that can result in the NS phenotype (Fig. 1) (61).

Recruited downstream factors can rewire the network. In contrast to replicative senescence (RS) and OIS, developmentally programmed senescence occurs with undetectable γH2AX (phosphorylated histone H2AX) and p53 signaling. OIS is mostly mediated by the higher expression of p16INK4A (CDK inhibitor 2A), p19ARF (ARF tumor suppressor, mice; p14ARF, human), p21WAF1 (CDK inhibitor 1 or CDK-interacting protein 1, p21CIP1), and p53 (76). Senescence during embryonic development does not appear to be dependent on p16INK4A, but on p21WAF1, p15INK4B (CDK4 inhibitor B), and TGF-β (transforming growth factor–β)/SMAD (mothers against decapentaplegic homolog) and PI3K/FOXO (forkhead box protein O) pathways, suggesting modified expression levels and resultant rewiring (76, 8284). The altered patterns of expression associated with senescence are linked to changes in the genome architecture. In OIS, the heterochromatin undergoes extensive three-dimensional (3D) genome reorganization, resulting in active genes that are positioned in regions bordering senescence-associated domains to be in spatial proximity, promoting their expression (85). In neural stem cells (NSCs), all genes linked to cancer can affect proliferation; however, only those that are expressed during a certain time window may elicit neuronal developmental disorders (86). Overexpression of cyclin D1 and CDK4 in NSCs can shorten the G1 phase and delay neurogenesis (77, 87, 88). Changes in expression levels and consequent rewiring are mediated by expression levels of cell type–specific transcription factor isoforms, generated by, e.g., alternative splicing (89), or by different factors. Data support the involvement of transcription factors in rewiring [e.g., (90, 91)].

Recent data have provided mechanistic level insight into the role of p21 in developmental senescence (92). The inhibition of NSC senescence underlies corticogenesis—that is, the formation of the cerebral cortex, which is the outer layer of the brain, during the development of the nervous system—and neurogenesis, the formation of new neurons that later differentiate into specific types of neurons. The receptor for activated C kinase (RACK1) is abundant in NSCs. Depletion of RACK1 as the cerebral cortex is forming during the development of the nervous system can lead to developmental senescence in NSCs; its presence prevents senescence during corticogenesis. p21WAF1 and TGF-β/SMAD are among the critical mediators of developmental senescence during embryonic development (84). RACK1 interacts with SMAD, blocking the TGF-β/SMAD pathway, which suppresses p21 expression and cellular senescence.

p19ARF and p16INK4A are critical mediators of cellular senescence (93, 94). Recent data suggest that p19ARF can also be among the proteins taking part in senescence during embryonic development, albeit through a different scenario (95). Valproic acid induces p19ARF-mediated cellular senescence causing neurodevelopmental defects including cognitive defects and autism. Valproic acid has been used to treat epilepsy and bipolar disorder, but not in pregnancy, as it can elicit birth defects, cognitive impairment, and autism. p19ARF-mediated cellular senescence and microcephaly was detected in valproic acid neurogenesis-defective neuroepithelial tissues. Considering that p21WAF1 acts in both OIS (96) and developmental senescence, the expression levels and consequent network wiring of senescence mediators may control both developmental senescence and OIS.

CHROMATIN REMODELING IN CANCER AND IN NEURODEVELOPMENTAL DISORDERS

Proteins with mutations associated with both cancer and neurodevelopmental disorders occur broadly, commonly in those associated with expression. Leucine-zipper transcription regulator 1 (LZTR1), a highly mutated tumor suppressor gene involved in several cancer types and developmental disorders, is one example (97, 98). Cohesin, a chromosome-associated multi-subunit protein complex (99), is another, possibly due to chromatin association. Chromatin regulators appear to be among the most frequently mutated genes in individuals with autism, and single-cell transcriptomics supports a role of CHD8, a DNA helicase that acts as a chromatin remodeling factor and regulates transcription in autism (100, 101). Genome-wide analyses pointed to candidate genes encoding nuclear factors implicated in chromatin remodeling. These include histone demethylation, and the recognition of DNA methylation (102, 103) and promoter architecture reorganization during neuronal cell differentiation implicates target genes for neurodevelopmental disorders (104). Remodeling can affect regulatory regions of distal genes via long-range chromatin interactions, in a cell type–, developmental stage–, and disease-specific manner (105). Variants of mediator MED12 subunit, with a central role in RNA polymerase II transcription and regulation of cell growth, development, and differentiation, have also been linked to neurodevelopmental disorders (106). Last, Pol III mutations linked to neurodevelopmental disorder defects were identified and up-regulation of Pol III transcription was observed in cancer (107). Cancer and neurodevelopmental disorders are linked, with an increased cancer risk among patients with certain developmental syndromes, and cancer-driving genes making up more than a third of the risk genes for developmental disorders (108). As in cancer, unless the expression level is very high, single mutations are unlikely to be sufficient for neurodevelopmental disorders. A single cancer driver confers a selective growth advantage during cancer development but is insufficient for cancer initiation or maintenance (57, 109). This may be the case for developmental disorders as well (57, 110). Thus, preexisting germline mutations may couple with emerging sporadic mutations and high expression, resulting in higher incidence of cancer in neurodevelopmental disorders. Chromatin may also be involved in the apparent absence (e.g., in PIK3CA) or scarcity (e.g., in Ras) of germline driver hotspots in humans. The relative sparseness of highly transforming driver mutations in the germline has been attributed to embryonic lethality resulting from severe phenotype that includes cardiomegaly and abnormal brain development (38). However, single mutations may be insufficient for emergence of cancer. These mutant alleles could couple either with mutations in chromatin remodelers that are likely to be rare in embryonic cells or, more likely, with high expression levels during transient exposure of the corresponding genes taking place during chromatin reorganization in the course of embryo development. Here, we focus on proteins in the major Ras signaling pathways, MAPK and PI3K/AKT/mTOR.

DEVELOPMENTAL DISORDERS MAY OR MAY NOT INVOLVE THE SAME MUTATIONS AS CANCER

Exactly how to link same-gene mutations to neurodevelopmental disorders versus cancer has been unclear. Some publications focused on developmental syndromes and the cell cycle, whereas others explored the mutational properties and performed network analysis (111, 112). A decade-old attractive hypothesis argued that the primary feature distinguishing between the two types of mutations is signaling strength and duration (113). If strong and short, it can elicit developmental syndromes; if weak and sustained, the likely outcome is cancer. This hypothesis has been tested by measuring signal strength, for example, phosphorylation levels for MAPK-associated RASopathies. However, the linkage between a mutation and signal strength and duration, and the differential clinical outcomes have been elusive. Signal strength and duration are also unlikely to be independent parameters. Strength may tune duration via senescence.

Mutations in RAS and related pathway genes such as NF1 can lead to cancer and developmental disorders (1). Apart from CS, RASopathy-associated mutations involve germline variants, whereas cancer emerges mostly from somatic mutations. KRas and NRas germline RASopathy variants are rarely at G12, G13, and Q61, which are frequent in cancers. KRas mutations are especially frequent at G12, particularly in pancreatic ductal adenocarcinoma (PDAC), colorectal adenocarcinoma (CRC), and non–small cell lung cancer (NSCLC) (114, 115). G13X, A146T, and K117N are common in CRC (116). G12/13 block GAP-catalyzed and intrinsic hydrolysis. A146T enhances nucleotide exchange, with a 12-fold higher guanosine diphosphate (GDP) dissociation rate than wild-type KRas. K117N increases guanosine triphosphate (GTP) binding up to five- to sixfold (117). On the other hand, germline KRas mutations were identified at, e.g., K5, V14, Q22, P34, I36, T58, G60, V152, D153, and F156 (Fig. 4). They are associated with NS and CFC syndrome (1). Most increase the levels of active KRas, but with a milder phenotype than the oncogenic mutations (116). Only a moderate increase in the phosphorylation levels of downstream signaling proteins was observed.

Fig. 4. Ras mutations and their associated diseases.

Fig. 4.

(Left) Locations of the somatic mutations of Ras proteins in cancer. The somatic mutations can be observed in all Ras isoforms including HRas, NRas, and KRas, but expressed with different strengths. For example, G12X is populated in KRas and HRas, while Q61X is populated in NRas. (Right) Locations of the embryonic mutations of KRas in RASopathy. The white cartoons represent the catalytic domain of Ras protein. The locations and types of the mutations with associated diseases are highlighted in the boxes.

In a B-Raf example, phosphoproteomic analyses showed mutations in septo-optic dysplasia (SOD) including hypopituitarism and CFC syndrome. Activation of the MAPK pathway by B-RafV600E, or B-RafQ241R, leads to abnormal cell lineage determination and terminal differentiation of hormone-producing cells, causing hypopituitarism. Expression of the B-RafV600E in embryonic pituitary progenitors promotes expression of cell cycle inhibitors, cell growth arrest, and apoptosis, but not cancer (118). B-RafV600E is also a common strong driver of cancer. Thus, there appears to be no rule: Different diseases may arise from different or the same mutations. The mutations are not differentiated by their strengths. Neither type seems to be consistently stronger than the other. Computations observed a trend that energy changes upon residue substitutions are higher for cancer compared to RASopathy mutations (111).

PIK3CA, encoding the catalytic subunit of PI3Kα lipid kinase in the PI3K/AKT/mTOR pathway, provides an example for another developmental disorder (Fig. 3). It is the second (third) most highly mutated protein in cancer. Its mutations are also observed in benign PROS (57). Although PROS are not inherited germline line mutations, they emerge during embryonic development, with timing and location (skin, vasculature, bones, fat, and brain tissue) depending on the specific disease and determining the disease phenotype. Germline mutations in PIK3CD can lead to the activated PI3Kδ syndrome (APDS), with neurodevelopmental abnormalities such as speech and developmental retardation (119). APDS, particularly APDS2, also increases the risk of Hodgkin’s and non-Hodgkin’s lymphomas. PI3Kδ isoform mostly expresses in lymphocytes. APDS can also promote T cell senescence and a rare primary immunodeficiency disorder (120). PIK3CD mutations elevate basal activity through enhanced membrane binding of p110δ.

Somatic mutations in PTEN occur in many cancers, most frequently in prostate, and endometrial cancers, as well as melanoma. They are also common in brain glioblastomas and astrocytomas. Preexisting germline autism-associated mutations could couple with somatic mutations contributing to the evolution of brain cancers (112).

As observed in RASopathies, mutations can express unique developmental syndromes in specific cell types; however, there can also be certain overlapping phenotypic features, including craniofacial abnormalities, congenital heart defects, and neurocognitive delay (46). These may arise from contributions from other cell types if the expression levels exceed the threshold and expression of downstream nodes is at sufficiently high levels for the potent signaling to go through.

THE MECHANISM OF ACTIVATION AND SIGNALING STRENGTH

Despite the absence of a clear pattern of relative strength, the mutation should be sufficiently strong for a high percentage of the molecules to be activated. Molecular activation mechanisms can be a key factor in deciding signal strength and disease outcome. One example is KRasG12D predominating in pancreatic cancer but not KRasA146T, which is almost exclusively in cancers of the intestinal tract particularly colon, and blood (121). The overall frequency of the G12D mutation is 23.4%, and that of the G12V mutation is 21.1% (122); that of A146T is approximately 4% in colorectal cancers (123). The statistics correlates with the strength. Prognosis of KRasG12V mutant tumors is poorer than that of other subtypes (124), and in metastatic pancreatic cancer, it is correlated with poor overall survival (125). Tissue specificity and varied strengths point to differences in signaling downstream (121). Ras is activated by exchange of GDP by GTP; it is deactivated by GTP hydrolysis. KRasG12X mutations interfere with GTP hydrolysis, retaining KRas in an active, GTP-bound state. On the other hand, KRasA146T promotes an extension of Switch I away from the KRas catalytic domain and the nucleotide binding site, and nucleotide exchange. The differences in the strengths of the mutations derive from the location of the mutations and their consequent activation mechanism. GAP-induced hydrolysis is impaired by Gly12 but not by A146T mutations, which act by accelerating nucleotide exchange. On the other hand, mutations in tumor suppressor PTEN abolish its activation mechanism at the membrane (75).

Mutations in kinases and phosphatases may lead to divergent phenotypes by phosphorylating/dephosphorylating different substrates in different tissue types (126). Recent data suggest that HRasG12V is involved in hypotonia in CS. A mouse model showed an inverse correlation of Ras/MAPK and p38, which is essential for skeletal muscle growth and homeostasis through the dual-specificity phosphatase (DUSP). DUSP6 dephosphorylates ERK1/2, as well as p38 in the presence of pERK2, with DUSP6, pERK2, and p-p38 forming a stable complex.

Recently, SHP2 liquid-liquid phase separation (LLPS) helped address the perplexing question of why both enzymatically activating and inactivating mutations in PTPN11 result in developmental disorder with overlapping clinical manifestations (127). The authors showed that the inactivating mutation of SHP2 activates wild-type SHP2 through inducing LLPS, thereby accounting for the activation mechanism of wild-type SHP2. The activation of SHP2 is required for some receptor tyrosine kinase (RTK) signaling pathways (128).

In addition to the location, the biophysical and biochemical attributes of the mutations in cancer and neurodevelopmental disorders matter. They can alter the stabilities, conformational distributions, dynamics, and interactions of the respective protein. SHP2 is an example of the potential therapeutic consequences of the conformational properties of the mutations in specific proteins (127); Ras is an example of the mutational properties and activation mechanism (74). At the same time, the observations that even the same mutations in the same proteins can lead to different phenotypic disease manifestations, and the lack of consistency of the mutational strengths in cancer versus neurodevelopmental disorders, argue that mutational properties per se are unable to resolve the enigma of their different clinical outcomes.

CONCLUDING REMARKS: CANCER, NEURODEVELOPMENTAL DISORDERS, PHARMACOLOGY, AND THE ULTIMATE AIM OF MEASUREMENTS

Cancer depends on the number of the corresponding activated proteins, which is determined by their expression levels and oncogenic mutations. Neurodevelopmental disorders embody the same principles. The mutations can occur in the same proteins and may or may not overlap. No uniform pattern is observed in their relative strengths. In both, propagation of a potent signal requires large-enough numbers of available active molecules in each node in the respective pathways. These numbers depend on protein expression levels, which rest on the 3D structures of the genome and chromatin accessibility of the corresponding genes, which vary across cell subtypes and timing (embryonic development or differentiated) windows. If the number of molecules in a specific node is too low, pathway rewiring through involvement of transcription factor isoforms (or other factors) can take place, resulting in premature developmental senescence. One way for this blueprint to be tested is by the mRNA levels of the nodes in the specific cell (67).

From the pharmacological standpoint, deciphering the same gene mutations but different disease enigma is vastly important. Recent evidence indicates that PI3K pathway inhibitors undergoing trials in cancer can provide a therapy for PROS. Since both mutation types are in the same protein, this can be understood. However, considering the dependence of signaling on expression of every single node in the pathway and rewiring, combinatorial drug regimens adopted in cancer, such as canonical Ras pathway inhibitors, may or may not work in neurodevelopmental disorders. Network analysis suggested that RASopathy proteins participate in 33 pathways (111). Inclusion of drugs acting on chromatin modifiers, such as statins and histone deacetylase inhibitors, may have a better chance (71); however, contending with toxicity remains a challenge (129). Last, with the advent of the impact of artificial intelligence (AI), the question has been raised as to whether AI can offer a breakthrough in predicting risks and prognosis for neurodevelopmental disorders (130) and drug repurposing (131). Within this framework, the challenging aim is an integrative systems biology platform that merges in vitro and tissue-specific in vivo models, allelic series, transcriptomics and proteomics, and patient samples to start to measure the differences and gauge the effects of these on neurodevelopmental disorders and cancer phenotypes and on therapeutics.

Within this ambitious framework, the first hurdle is measurement of the mutational activation strength and signal potency. We consider it in terms of signaling by the numbers, that is, aiming to measure the absolute number of activated molecules, and asking what thresholds should be considered for a signal to propagate through the pathway, reach the cell cycle, and activate (repress) a transcription factor. In line with our discussion here, the strength needs to account for the mutation potency and cell type–specific transcription level. The sparse measurements of the single-cell mRNA levels, especially across time windows during development and consequent likely dropout events, transcription bursts, and noisy data, are big hurdles to overcome. Deep learning, low-dimensional representation for scRNA-seq (single-cell RNA sequencing) data visualization and analysis is expected to assist (132).

The premise of measurements of signals induced by driver hotspots (and drug resistance mutations in cancer), leading to signal transduction from an upstream to a downstream node, to activate (or repress) function in a specific cell at a specific developmental state, with respect to an established threshold set by experiments, and exploited in deep learning predictions, is a powerful aim. If (or when) attained, it could provide a compelling starting point for biologists to investigate.

Acknowledgments

Funding: This project has been funded in whole or in part with federal funds from the National Cancer Institute, NIH, under contract no. HHSN261201500003I. The content of this publication does not necessarily reflect the views or policies of the Department of Health and Human Services, nor does mention of trade names, commercial products, or organizations imply endorsement by the U.S. government. This research was supported (in part) by the Intramural Research Program of the NIH, National Cancer Institute, Center for Cancer Research.

Author contributions: Conceptualization: R.N. and H.J. Investigation: R.N. C.-J.T., and H.J. Visualization: R.N. and H.J. Supervision: R.N. Writing—original draft: R.N. Writing—review and editing: All authors.

Competing interests: The authors declare that they have no competing interests.

References and Notes

  • 1.Dunnett-Kane V., Burkitt-Wright E., Blackhall F. H., Malliri A., Evans D. G., Lindsay C. R., Germline and sporadic cancers driven by the RAS pathway: Parallels and contrasts. Ann. Oncol. 31, 873–883 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Tidyman W. E., Rauen K. A., Pathogenetics of the RASopathies. Hum. Mol. Genet. 25, R123–R132 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Adviento B., Corbin I. L., Widjaja F., Desachy G., Enrique N., Rosser T., Risi S., Marco E. J., Hendren R. L., Bearden C. E., Rauen K. A., Weiss L. A., Autism traits in the RASopathies. J. Med. Genet. 51, 10–20 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.van Eyk C. L., Corbett M. A., Gardner A., van Bon B. W., Broadbent J. L., Harper K., MacLennan A. H., Gecz J., Analysis of 182 cerebral palsy transcriptomes points to dysregulation of trophic signalling pathways and overlap with autism. Transl. Psychiatry 8, 88 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Goyal Y., Jindal G. A., Pelliccia J. L., Yamaya K., Yeung E., Futran A. S., Burdine R. D., Schupbach T., Shvartsman S. Y., Divergent effects of intrinsically active MEK variants on developmental Ras signaling. Nat. Genet. 49, 465–469 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Inoue D., Polaski J. T., Taylor J., Castel P., Chen S., Kobayashi S., Hogg S. J., Hayashi Y., Pineda J. M. B., El Marabti E., Erickson C., Knorr K., Fukumoto M., Yamazaki H., Tanaka A., Fukui C., Lu S. X., Durham B. H., Liu B., Wang E., Mehta S., Zakheim D., Garippa R., Penson A., Chew G. L., McCormick F., Bradley R. K., Abdel-Wahab O., Minor intron retention drives clonal hematopoietic disorders and diverse cancer predisposition. Nat. Genet. 53, 707–718 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Christou E. E., Zafeiropoulos P., Rallis D., Baltogianni M., Asproudis C., Stefaniotou M., Giapros V., Asproudis I., A narrative review of the ocular manifestations in Noonan syndrome. Semin. Ophthalmol. (2021). [DOI] [PubMed] [Google Scholar]
  • 8.Van R., Cuevas-Navarro A., Castel P., McCormick F., The molecular functions of RIT1 and its contribution to human disease. Biochem. J. 477, 2755–2770 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Gross A. M., Frone M., Gripp K. W., Gelb B. D., Schoyer L., Schill L., Stronach B., Biesecker L. G., Esposito D., Hernandez E. R., Legius E., Loh M. L., Martin S., Morrison D. K., Rauen K. A., Wolters P. L., Zand D., McCormick F., Savage S. A., Stewart D. R., Widemann B. C., Yohe M. E., Advancing RAS/RASopathy therapies: An NCI-sponsored intramural and extramural collaboration for the study of RASopathies. Am. J. Med. Genet. A 182, 866–876 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Baltanas F. C., Garcia-Navas R., Santos E., SOS2 comes to the fore: Differential functionalities in physiology and pathology. Int. J. Mol. Sci. 22, (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Riller Q., Rieux-Laucat F., RASopathies: From germline mutations to somatic and multigenic diseases. Biomed. J. 44, 422–432 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Weber S. M., Carroll S. L., The role of R-Ras proteins in normal and pathologic migration and morphologic change. Am. J. Pathol. 191, 1499–1510 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Prior I., Ras variant biology and contributions to human disease. Methods Mol. Biol. 2262, 3–18 (2021). [DOI] [PubMed] [Google Scholar]
  • 14.Gripp K. W., Schill L., Schoyer L., Stronach B., Bennett A. M., Blaser S., Brown A., Burdine R., Burkitt-Wright E., Castel P., Darilek S., Dias A., Dyer T., Ellis M., Erickson G., Gelb B. D., Green T., Gross A., Ho A., Holder J. L. Jr., Inoue S. I., Jelin A. C., Kennedy A., Klein R., Kontaridis M. I., Magoulas P., McConnell D. B., McCormick F., Neel B. G., Prada C. E., Rauen K. A., Roberts A., Rodriguez-Viciana P., Rosen N., Rumbaugh G., Sablina A., Solman M., Tartaglia M., Thomas A., Timmer W. C., Venkatachalam K., Walsh K. S., Wolters P. L., Yi J. S., Zenker M., Ratner N., The sixth international RASopathies symposium: Precision medicine-from promise to practice. Am. J. Med. Genet. A 182, 597–606 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Kessler D., Gerlach D., Kraut N., McConnell D. B., Targeting son of sevenless 1: The pacemaker of KRAS. Curr. Opin. Chem. Biol. 62, 109–118 (2021). [DOI] [PubMed] [Google Scholar]
  • 16.Klomp J. E., Klomp J. A., Der C. J., The ERK mitogen-activated protein kinase signaling network: The final frontier in RAS signal transduction. Biochem. Soc. Trans. 49, 253–267 (2021). [DOI] [PubMed] [Google Scholar]
  • 17.Leclerc J., Vermaut C., Buisine M. P., Diagnosis of lynch syndrome and strategies to distinguish lynch-related tumors from sporadic MSI/dMMR tumors. Cancers 13, 467 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Mitri F., Bersano A., Herve D., Kraemer M., Cutaneous manifestations in moyamoya angiopathy: A review. Eur. J. Neurol. 28, 1784–1793 (2021). [DOI] [PubMed] [Google Scholar]
  • 19.Bergqvist C., Wolkenstein P., MEK inhibitors in RASopathies. Curr. Opin. Oncol. 33, 110–119 (2021). [DOI] [PubMed] [Google Scholar]
  • 20.Rauen K. A., Maeda Y., Egense A., Tidyman W. E., Familial cardio-facio-cutaneous syndrome: Vertical transmission of the BRAF p.G464R pathogenic variant and review of the literature. Am. J. Med. Genet. A 185, 469–475 (2021). [DOI] [PubMed] [Google Scholar]
  • 21.Lee Y., Choi Y., Seo G. H., Kim G. H., Choi I. H., Keum C., Ko J. M., Cheon C. K., Jeon J., Choi J. H., Yoo H. W., Lee B. H., Clinical and molecular spectra of BRAF-associated RASopathy. J. Hum. Genet. 66, 389–399 (2021). [DOI] [PubMed] [Google Scholar]
  • 22.Baltanas F. C., Zarich N., Rojas-Cabaneros J. M., Santos E., SOS GEFs in health and disease. Biochim. Biophys. Acta Rev. Cancer 1874, 188445 (2020). [DOI] [PubMed] [Google Scholar]
  • 23.Malaquias A. C., Jorge A. A. L., Activation of the MAPK pathway (RASopathies) and partial growth hormone insensitivity. Mol. Cell. Endocrinol. 519, 111040 (2021). [DOI] [PubMed] [Google Scholar]
  • 24.Martinez-Lopez A., Blasco-Morente G., Perez-Lopez I., Herrera-Garcia J. D., Luque-Valenzuela M., Sanchez-Cano D., Lopez-Gutierrez J. C., Ruiz-Villaverde R., Tercedor-Sanchez J., CLOVES syndrome: Review of a PIK3CA-related overgrowth spectrum (PROS). Clin. Genet. 91, 14–21 (2017). [DOI] [PubMed] [Google Scholar]
  • 25.Venot Q., Canaud G., PIK3CA-related overgrowth syndrome (PROS). Nephrol. Ther. 13( suppl. 1), S155–S156 (2017). [DOI] [PubMed] [Google Scholar]
  • 26.Venot Q., Blanc T., Rabia S. H., Berteloot L., Ladraa S., Duong J. P., Blanc E., Johnson S. C., Hoguin C., Boccara O., Sarnacki S., Boddaert N., Pannier S., Martinez F., Magassa S., Yamaguchi J., Knebelmann B., Merville P., Grenier N., Joly D., Cormier-Daire V., Michot C., Bole-Feysot C., Picard A., Soupre V., Lyonnet S., Sadoine J., Slimani L., Chaussain C., Laroche-Raynaud C., Guibaud L., Broissand C., Amiel J., Legendre C., Terzi F., Canaud G., Targeted therapy in patients with PIK3CA-related overgrowth syndrome. Nature 558, 540–546 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Tiwari A., Rahi S., Mehan S., Elucidation of abnormal extracellular regulated kinase (ERK) signaling and associations with syndromic and non-syndromic autism. Curr. Drug Targets 22, 1071–1086 (2021). [DOI] [PubMed] [Google Scholar]
  • 28.Yehia L., Keel E., Eng C., The clinical spectrum of PTEN mutations. Annu. Rev. Med. 71, 103–116 (2020). [DOI] [PubMed] [Google Scholar]
  • 29.Yehia L., Ngeow J., Eng C., PTEN-opathies: From biological insights to evidence-based precision medicine. J. Clin. Invest. 129, 452–464 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Spina Nagy G., Kawamoto E. M., Bridi J. C., The role of PTEN signaling in synaptic function: Implications in autism spectrum disorder. Neurosci. Lett. 759, 136015 (2021). [DOI] [PubMed] [Google Scholar]
  • 31.Liu K. X., Sethi R. V., Pulsifer M. B., D’Gama A. M., LaVally B., Ebb D. H., Tarbell N. J., Yock T. I., MacDonald S. M., Clinical outcomes of pediatric patients with autism spectrum disorder and other neurodevelopmental disorders and intracranial germ cell tumors. Pediatr. Blood Cancer 68, e28935 (2021). [DOI] [PubMed] [Google Scholar]
  • 32.Valentine A. Z., Brown B. J., Groom M. J., Young E., Hollis C., Hall C. L., A systematic review evaluating the implementation of technologies to assess, monitor and treat neurodevelopmental disorders: A map of the current evidence. Clin. Psychol. Rev. 80, 101870 (2020). [DOI] [PubMed] [Google Scholar]
  • 33.Hanly C., Shah H., Au P. Y. B., Murias K., Description of neurodevelopmental phenotypes associated with 10 genetic neurodevelopmental disorders: A scoping review. Clin. Genet. 99, 335–346 (2021). [DOI] [PubMed] [Google Scholar]
  • 34.Bartels E. M., Korbo L., Harrison A. P., Novel insights into cerebral palsy. J. Muscle Res. Cell Motil. 41, 265–267 (2020). [DOI] [PubMed] [Google Scholar]
  • 35.Fahey M. C., Maclennan A. H., Kretzschmar D., Gecz J., Kruer M. C., The genetic basis of cerebral palsy. Dev. Med. Child Neurol. 59, 462–469 (2017). [DOI] [PubMed] [Google Scholar]
  • 36.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 84, 306–327 (2018). [DOI] [PubMed] [Google Scholar]
  • 37.Curry R. N., Glasgow S. M., The Role of neurodevelopmental pathways in brain tumors. Front. Cell Dev. Biol. 9, 659055 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Castel P., Rauen K. A., McCormick F., The duality of human oncoproteins: Drivers of cancer and congenital disorders. Nat. Rev. Cancer 20, 383–397 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Ostrom Q. T., Gittleman H., Xu J., Kromer C., Wolinsky Y., Kruchko C., Barnholtz-Sloan J. S., CBTRUS statistical report: Primary brain and other central nervous system tumors diagnosed in the United States in 2009-2013. Neuro Oncol. 18, v1–v75 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Briscoe J., Marin O., Looking at neurodevelopment through a big data lens. Science 369, eaaz8627 (2020). [DOI] [PubMed] [Google Scholar]
  • 41.Mensink K. A., Ketterling R. P., Flynn H. C., Knudson R. A., Lindor N. M., Heese B. A., Spinner R. J., Babovic-Vuksanovic D., Connective tissue dysplasia in five new patients with NF1 microdeletions: Further expansion of phenotype and review of the literature. J. Med. Genet. 43, e8 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Pingel J., Kampmann M. L., Andersen J. D., Wong C., Dossing S., Borsting C., Nielsen J. B., Gene expressions in cerebral palsy subjects reveal structural and functional changes in the gastrocnemius muscle that are closely associated with passive muscle stiffness. Cell Tissue Res. 384, 513–526 (2021). [DOI] [PubMed] [Google Scholar]
  • 43.Parenti I., Rabaneda L. G., Schoen H., Novarino G., Neurodevelopmental disorders: From genetics to functional pathways. Trends Neurosci. 43, 608–621 (2020). [DOI] [PubMed] [Google Scholar]
  • 44.Rauen K. A., The RASopathies. Annu. Rev. Genomics Hum. Genet. 14, 355–369 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ganetzky R., Finn E., Bagchi A., Zollo O., Conlin L., Deardorff M., Harr M., Simpson M. A., McGrath J. A., Zackai E., Lemmon M. A., Sondheimer N., EGFR mutations cause a lethal syndrome of epithelial dysfunction with progeroid features. Mol. Genet. Genomic Med. 3, 452–458 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Patel A. L., Shvartsman S. Y., Outstanding questions in developmental ERK signaling. Development 145, dev143818 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Jin S. C., Lewis S. A., Bakhtiari S., Zeng X., Sierant M. C., Shetty S., Nordlie S. M., Elie A., Corbett M. A., Norton B. Y., van Eyk C. L., Haider S., Guida B. S., Magee H., Liu J., Pastore S., Vincent J. B., Brunstrom-Hernandez J., Papavasileiou A., Fahey M. C., Berry J. G., Harper K., Zhou C., Zhang J., Li B., Zhao H., Heim J., Webber D. L., Frank M. S. B., Xia L., Xu Y., Zhu D., Zhang B., Sheth A. H., Knight J. R., Castaldi C., Tikhonova I. R., Lopez-Giraldez F., Keren B., Whalen S., Buratti J., Doummar D., Cho M., Retterer K., Millan F., Wang Y., Waugh J. L., Rodan L., Cohen J. S., Fatemi A., Lin A. E., Phillips J. P., Feyma T., MacLennan S. C., Vaughan S., Crompton K. E., Reid S. M., Reddihough D. S., Shang Q., Gao C., Novak I., Badawi N., Wilson Y. A., McIntyre S. J., Mane S. M., Wang X., Amor D. J., Zarnescu D. C., Lu Q., Xing Q., Zhu C., Bilguvar K., Padilla-Lopez S., Lifton R. P., Gecz J., MacLennan A. H., Kruer M. C., Mutations disrupting neuritogenesis genes confer risk for cerebral palsy. Nat. Genet. 52, 1046–1056 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Zamboni V., Jones R., Umbach A., Ammoni A., Passafaro M., Hirsch E., Merlo G. R., Rho GTPases in intellectual disability: From genetics to therapeutic opportunities. Int. J. Mol. Sci. 19, 1821 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. News. Genet. Med. 22, 1924 (2020). [Google Scholar]
  • 50.de Curtis I., The Rac3 GTPase in neuronal development, neurodevelopmental disorders, and cancer. Cells 8, 1063 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hiatt S. M., Neu M. B., Ramaker R. C., Hardigan A. A., Prokop J. W., Hancarova M., Prchalova D., Havlovicova M., Prchal J., Stranecky V., Yim D. K. C., Powis Z., Keren B., Nava C., Mignot C., Rio M., Revah-Politi A., Hemati P., Stong N., Iglesias A. D., Suchy S. F., Willaert R., Wentzensen I. M., Wheeler P. G., Brick L., Kozenko M., Hurst A. C. E., Wheless J. W., Lacassie Y., Myers R. M., Barsh G. S., Sedlacek Z., Cooper G. M., De novo mutations in the GTP/GDP-binding region of RALA, a RAS-like small GTPase, cause intellectual disability and developmental delay. PLOS Genet. 14, e1007671 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Banworth M. J., Li G., Consequences of Rab GTPase dysfunction in genetic or acquired human diseases. Small GTPases 9, 158–181 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Qu L., Pan C., He S. M., Lang B., Gao G. D., Wang X. L., Wang Y., The Ras superfamily of small GTPases in non-neoplastic cerebral diseases. Front. Mol. Neurosci. 12, 121 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Sekiguchi M., Sobue A., Kushima I., Wang C., Arioka Y., Kato H., Kodama A., Kubo H., Ito N., Sawahata M., Hada K., Ikeda R., Shinno M., Mizukoshi C., Tsujimura K., Yoshimi A., Ishizuka K., Takasaki Y., Kimura H., Xing J., Yu Y., Yamamoto M., Okada T., Shishido E., Inada T., Nakatochi M., Takano T., Kuroda K., Amano M., Aleksic B., Yamomoto T., Sakuma T., Aida T., Tanaka K., Hashimoto R., Arai M., Ikeda M., Iwata N., Shimamura T., Nagai T., Nabeshima T., Kaibuchi K., Yamada K., Mori D., Ozaki N., ARHGAP10, which encodes Rho GTPase-activating protein 10, is a novel gene for schizophrenia risk. Transl. Psychiatry 10, 247 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Teng J. P., Yang Z. Y., Zhu Y. M., Ni D., Zhu Z. J., Li X. Q., The roles of ARHGAP10 in the proliferation, migration and invasion of lung cancer cells. Oncol. Lett. 14, 4613–4618 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Hada K., Wulaer B., Nagai T., Itoh N., Sawahata M., Sobue A., Mizoguchi H., Mori D., Kushima I., Nabeshima T., Ozaki N., Yamada K., Mice carrying a schizophrenia-associated mutation of the Arhgap10 gene are vulnerable to the effects of methamphetamine treatment on cognitive function: Association with morphological abnormalities in striatal neurons. Mol. Brain 14, 21 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Madsen R. R., Vanhaesebroeck B., Semple R. K., Cancer-associated PIK3CA mutations in overgrowth disorders. Trends Mol. Med. 24, 856–870 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Morris S. M., Acosta M. T., Garg S., Green J., Huson S., Legius E., North K. N., Payne J. M., Plasschaert E., Frazier T. W., Weiss L. A., Zhang Y., Gutmann D. H., Constantino J. N., Disease burden and symptom structure of autism in neurofibromatosis type 1: A study of the International NF1-ASD Consortium Team (INFACT). JAMA Psychiat. 73, 1276–1284 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Colas P., Cyclin-dependent kinases and rare developmental disorders. Orphanet J. Rare Dis. 15, 203 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Ryu H. H., Kim T., Kim J. W., Kang M., Park P., Kim Y. G., Kim H., Ha J., Choi J. E., Lee J., Lim C. S., Kim C. H., Kim S. J., Silva A. J., Kaang B. K., Lee Y. S., Excitatory neuron-specific SHP2-ERK signaling network regulates synaptic plasticity and memory. Sci. Signal. 12, eaau5755 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.L. K. Ferrarelli, in Science (American Association for the Advancement of Science, 2019), vol. 363, chap. 1052, pp. 1052–1054. [Google Scholar]
  • 62.Nussinov R., Tsai C. J., Jang H., Are parallel proliferation pathways redundant? Trends Biochem. Sci. 45, 554–563 (2020). [DOI] [PubMed] [Google Scholar]
  • 63.Kleijer K. T. E., Huguet G., Tastet J., Bourgeron T., Burbach J. P. H., Anatomy and cell biology of autism spectrum disorder: Lessons from human genetics. Adv. Anat. Embryol. Cell Biol. 224, 1–25 (2017). [DOI] [PubMed] [Google Scholar]
  • 64.Nussinov R., Jang H., Nir G., Tsai C. J., Cheng F., A new precision medicine initiative at the dawn of exascale computing. Signal Transduct. Target. Ther. 6, 3 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Merienne N., Meunier C., Schneider A., Seguin J., Nair S. S., Rocher A. B., Le Gras S., Keime C., Faull R., Pellerin L., Chatton J. Y., Neri C., Merienne K., Deglon N., Cell-type-specific gene expression profiling in adult mouse brain reveals normal and disease-state signatures. Cell Rep. 26, 2477–2493.e9 (2019). [DOI] [PubMed] [Google Scholar]
  • 66.Silberberg G., Bolam J. P., Local and afferent synaptic pathways in the striatal microcircuitry. Curr. Opin. Neurobiol. 33, 182–187 (2015). [DOI] [PubMed] [Google Scholar]
  • 67.O’Connor S. A., Feldman H. M., Arora S., Hoellerbauer P., Toledo C. M., Corrin P., Carter L., Kufeld M., Bolouri H., Basom R., Delrow J., McFaline-Figueroa J. L., Trapnell C., Pollard S. M., Patel A., Paddison P. J., Plaisier C. L., Neural G0: A quiescent-like state found in neuroepithelial-derived cells and glioma. Mol. Syst. Biol. 17, e9522 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Rahman M. R., Petralia M. C., Ciurleo R., Bramanti A., Fagone P., Shahjaman M., Wu L., Sun Y., Turanli B., Arga K. Y., Islam M. R., Islam T., Nicoletti F., Comprehensive analysis of RNA-seq gene expression profiling of brain transcriptomes reveals novel genes, regulators, and pathways in autism spectrum disorder. Brain Sci. 10, 747 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Mullins N., Forstner A. J., O’Connell K. S., Coombes B., Coleman J. R. I., Qiao Z., Als T. D., Bigdeli T. B., Borte S., Bryois J., Charney A. W., Drange O. K., Gandal M. J., Hagenaars S. P., Ikeda M., Kamitaki N., Kim M., Krebs K., Panagiotaropoulou G., Schilder B. M., Sloofman L. G., Steinberg S., Trubetskoy V., Winsvold B. S., Won H. H., Abramova L., Adorjan K., Agerbo E., Al Eissa M., Albani D., Alliey-Rodriguez N., Anjorin A., Antilla V., Antoniou A., Awasthi S., Baek J. H., Baekvad-Hansen M., Bass N., Bauer M., Beins E. C., Bergen S. E., Birner A., Pedersen C. B., Boen E., Boks M. P., Bosch R., Brum M., Brumpton B. M., Brunkhorst-Kanaan N., Budde M., Bybjerg-Grauholm J., Byerley W., Cairns M., Casas M., Cervantes P., Clarke T. K., Cruceanu C., Cuellar-Barboza A., Cunningham J., Curtis D., Czerski P. M., Dale A. M., Dalkner N., David F. S., Degenhardt F., Djurovic S., Dobbyn A. L., Douzenis A., Elvsashagen T., Escott-Price V., Ferrier I. N., Fiorentino A., Foroud T. M., Forty L., Frank J., Frei O., Freimer N. B., Frisen L., Gade K., Garnham J., Gelernter J., Pedersen M. G., Gizer I. R., Gordon S. D., Gordon-Smith K., Greenwood T. A., Grove J., Guzman-Parra J., Ha K., Haraldsson M., Hautzinger M., Heilbronner U., Hellgren D., Herms S., Hoffmann P., Holmans P. A., Huckins L., Jamain S., Johnson J. S., Kalman J. L., Kamatani Y., Kennedy J. L., Kittel-Schneider S., Knowles J. A., Kogevinas M., Koromina M., Kranz T. M., Kranzler H. R., Kubo M., Kupka R., Kushner S. A., Lavebratt C., Lawrence J., Leber M., Lee H. J., Lee P. H., Levy S. E., Lewis C., Liao C., Lucae S., Lundberg M., MacIntyre D. J., Magnusson S. H., Maier W., Maihofer A., Malaspina D., Maratou E., Martinsson L., Mattheisen M., McCarroll S. A., McGregor N. W., McGuffin P., McKay J. D., Medeiros H., Medland S. E., Millischer V., Montgomery G. W., Moran J. L., Morris D. W., Muhleisen T. W., O’Brien N., O’Donovan C., Loohuis L. M. O., Oruc L., Papiol S., Pardinas A. F., Perry A., Pfennig A., Porichi E., Potash J. B., Quested D., Raj T., Rapaport M. H., DePaulo J. R., Regeer E. J., Rice J. P., Rivas F., Rivera M., Roth J., Roussos P., Ruderfer D. M., Sanchez-Mora C., Schulte E. C., Senner F., Sharp S., Shilling P. D., Sigurdsson E., Sirignano L., Slaney C., Smeland O. B., Smith D. J., Sobell J. L., Hansen C. S., Artigas M. S., Spijker A. T., Stein D. J., Strauss J. S., Swiatkowska B., Terao C., Thorgeirsson T. E., Toma C., Tooney P., Tsermpini E. E., Vawter M. P., Vedder H., Walters J. T. R., Witt S. H., Xi S., Xu W., Yang J. M. K., Young A. H., Young H., Zandi P. P., Zhou H., Zillich L., Psychiatry H. A.-I., Adolfsson R., Agartz I., Alda M., Alfredsson L., Babadjanova G., Backlund L., Baune B. T., Bellivier F., Bengesser S., Berrettini W. H., Blackwood D. H. R., Boehnke M., Borglum A. D., Breen G., Carr V. J., Catts S., Corvin A., Craddock N., Dannlowski U., Dikeos D., Esko T., Etain B., Ferentinos P., Frye M., Fullerton J. M., Gawlik M., Gershon E. S., Goes F. S., Green M. J., Grigoroiu-Serbanescu M., Hauser J., Henskens F., Hillert J., Hong K. S., Hougaard D. M., Hultman C. M., Hveem K., Iwata N., Jablensky A. V., Jones I., Jones L. A., Kahn R. S., Kelsoe J. R., Kirov G., Landen M., Leboyer M., Lewis C. M., Li Q. S., Lissowska J., Lochner C., Loughland C., Martin N. G., Mathews C. A., Mayoral F., McElroy S. L., McIntosh A. M., McMahon F. J., Melle I., Michie P., Milani L., Mitchell P. B., Morken G., Mors O., Mortensen P. B., Mowry B., Muller-Myhsok B., Myers R. M., Neale B. M., Nievergelt C. M., Nordentoft M., Nothen M. M., O’Donovan M. C., Oedegaard K. J., Olsson T., Owen M. J., Paciga S. A., Pantelis C., Pato C., Pato M. T., Patrinos G. P., Perlis R. H., Posthuma D., Ramos-Quiroga J. A., Reif A., Reininghaus E. Z., Ribases M., Rietschel M., Ripke S., Rouleau G. A., Saito T., Schall U., Schalling M., Schofield P. R., Schulze T. G., Scott L. J., Scott R. J., Serretti A., Shannon Weickert C., Smoller J. W., Stefansson H., Stefansson K., Stordal E., Streit F., Sullivan P. F., Turecki G., Vaaler A. E., Vieta E., Vincent J. B., Waldman I. D., Weickert T. W., Werge T., Wray N. R., Zwart J. A., Biernacka J. M., Nurnberger J. I., Cichon S., Edenberg H. J., Stahl E. A., McQuillin A., Di Florio A., Ophoff R. A., Andreassen O. A., Genome-wide association study of more than 40,000 bipolar disorder cases provides new insights into the underlying biology. Nat. Genet. 53, 817–829 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Velmeshev D., Schirmer L., Jung D., Haeussler M., Perez Y., Mayer S., Bhaduri A., Goyal N., Rowitch D. H., Kriegstein A. R., Single-cell genomics identifies cell type-specific molecular changes in autism. Science 364, 685–689 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Das T. K., Gatto J., Mirmira R., Hourizadeh E., Kaufman D., Gelb B. D., Cagan R., Drosophila RASopathy models identify disease subtype differences and biomarkers of drug efficacy. iScience 24, 102306 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Zhang M., Jang H., Nussinov R., PI3K driver mutations: A biophysical membrane-centric perspective. Cancer Res. 81, 237–247 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Zhang M., Jang H., Nussinov R., The mechanism of PI3Kα activation at the atomic level. Chem. Sci. 10, 3671–3680 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Nussinov R., Jang H., Gursoy A., Keskin O., Gaponenko V., Inhibition of nonfunctional Ras. Cell Chem. Biol. 28, 121–133 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Jang H., Smith I. N., Eng C., Nussinov R., The mechanism of full activation of tumor suppressor PTEN at the phosphoinositide-enriched membrane. iScience 24, 102438 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Engler M., Fidan M., Nandi S., Cirstea I. C., Senescence in RASopathies, a possible novel contributor to a complex pathophenoype. Mech. Ageing Dev. 194, 111411 (2021). [DOI] [PubMed] [Google Scholar]
  • 77.Ruijtenberg S., van den Heuvel S., Coordinating cell proliferation and differentiation: Antagonism between cell cycle regulators and cell type-specific gene expression. Cell Cycle 15, 196–212 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Cichowski K., Hahn W. C., Unexpected pieces to the senescence puzzle. Cell 133, 958–961 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Kilbey A., Terry A., Cameron E. R., Neil J. C., Oncogene-induced senescence: An essential role for Runx. Cell Cycle 7, 2333–2340 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Tajan M., Paccoud R., Branka S., Edouard T., Yart A., The RASopathy family: Consequences of germline activation of the RAS/MAPK pathway. Endocr. Rev. 39, 676–700 (2018). [DOI] [PubMed] [Google Scholar]
  • 81.Santoriello C., Deflorian G., Pezzimenti F., Kawakami K., Lanfrancone L., d’Adda di Fagagna F., Mione M., Expression of H-RASV12 in a zebrafish model of Costello syndrome causes cellular senescence in adult proliferating cells. Dis. Model. Mech. 2, 56–67 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Davaapil H., Brockes J. P., Yun M. H., Conserved and novel functions of programmed cellular senescence during vertebrate development. Development 144, 106–114 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Munoz-Espin D., Canamero M., Maraver A., Gomez-Lopez G., Contreras J., Murillo-Cuesta S., Rodriguez-Baeza A., Varela-Nieto I., Ruberte J., Collado M., Serrano M., Programmed cell senescence during mammalian embryonic development. Cell 155, 1104–1118 (2013). [DOI] [PubMed] [Google Scholar]
  • 84.Storer M., Mas A., Robert-Moreno A., Pecoraro M., Ortells M. C., Di Giacomo V., Yosef R., Pilpel N., Krizhanovsky V., Sharpe J., Keyes W. M., Senescence is a developmental mechanism that contributes to embryonic growth and patterning. Cell 155, 1119–1130 (2013). [DOI] [PubMed] [Google Scholar]
  • 85.Sati S., Bonev B., Szabo Q., Jost D., Bensadoun P., Serra F., Loubiere V., Papadopoulos G. L., Rivera-Mulia J. C., Fritsch L., Bouret P., Castillo D., Gelpi J. L., Orozco M., Vaillant C., Pellestor F., Bantignies F., Marti-Renom M. A., Gilbert D. M., Lemaitre J. M., Cavalli G., 4D genome rewiring during oncogene-induced and replicative senescence. Mol. Cell 78, 522–538.e9 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Ernst C., Proliferation and differentiation deficits are a major convergence point for neurodevelopmental disorders. Trends Neurosci. 39, 290–299 (2016). [DOI] [PubMed] [Google Scholar]
  • 87.Lange C., Huttner W. B., Calegari F., Cdk4/cyclinD1 overexpression in neural stem cells shortens G1, delays neurogenesis, and promotes the generation and expansion of basal progenitors. Cell Stem Cell 5, 320–331 (2009). [DOI] [PubMed] [Google Scholar]
  • 88.Calegari F., Huttner W. B., An inhibition of cyclin-dependent kinases that lengthens, but does not arrest, neuroepithelial cell cycle induces premature neurogenesis. J. Cell Sci. 116, 4947–4955 (2003). [DOI] [PubMed] [Google Scholar]
  • 89.Zhang Y., Li S., Xiang Y., Qiu L., Zhao H., Hayes J. D., The selective post-translational processing of transcription factor Nrf1 yields distinct isoforms that dictate its ability to differentially regulate gene expression. Sci. Rep. 5, 12983 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Poramba-Liyanage D. W., Korthout T., Cucinotta C. E., van Kruijsbergen I., van Welsem T., El Atmioui D., Ovaa H., Tsukiyama T., van Leeuwen F., Inhibition of transcription leads to rewiring of locus-specific chromatin proteomes. Genome Res. 30, 635–646 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Krivtsova O., Makarova A., Lazarevich N., Aberrant expression of alternative isoforms of transcription factors in hepatocellular carcinoma. World J. Hepatol. 10, 645–661 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Zhu Q., Chen L., Li Y., Huang M., Shao J., Li S., Cheng J., Yang H., Wu Y., Zhang J., Feng J., Fan M., Wu H., Rack1 is essential for corticogenesis by preventing p21-dependent senescence in neural stem cells. Cell Rep. 36, 109639 (2021). [DOI] [PubMed] [Google Scholar]
  • 93.Yetil A., Anchang B., Gouw A. M., Adam S. J., Zabuawala T., Parameswaran R., van Riggelen J., Plevritis S., Felsher D. W., p19ARF is a critical mediator of both cellular senescence and an innate immune response associated with MYC inactivation in mouse model of acute leukemia. Oncotarget 6, 3563–3577 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Randle D. H., Zindy F., Sherr C. J., Roussel M. F., Differential effects of p19(Arf) and p16(Ink4a) loss on senescence of murine bone marrow-derived preB cells and macrophages. Proc. Natl. Acad. Sci. U.S.A. 98, 9654–9659 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.M. Rhinn, I. Zapata-Bodalo, A. Klein, J.-L. Plassat, T. Knauer-Meyer, W. M. Keyes, Valproic acid induces p19Arf-mediated cellular senescence to cause neurodevelopmental defects. bioRxiv 2021.05.28.446099 [Preprint]. 29 May 2021. 10.1101/2021.05.28.446099. [DOI]
  • 96.Zakut R., Givol D., The tumor suppression function of p21Waf is contained in its N-terminal half (‘half-WAF’). Oncogene 11, 393–395 (1995). [PubMed] [Google Scholar]
  • 97.Piotrowski A., Xie J., Liu Y. F., Poplawski A. B., Gomes A. R., Madanecki P., Fu C., Crowley M. R., Crossman D. K., Armstrong L., Babovic-Vuksanovic D., Bergner A., Blakeley J. O., Blumenthal A. L., Daniels M. S., Feit H., Gardner K., Hurst S., Kobelka C., Lee C., Nagy R., Rauen K. A., Slopis J. M., Suwannarat P., Westman J. A., Zanko A., Korf B. R., Messiaen L. M., Germline loss-of-function mutations in LZTR1 predispose to an inherited disorder of multiple schwannomas. Nat. Genet. 46, 182–187 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Yamamoto G. L., Aguena M., Gos M., Hung C., Pilch J., Fahiminiya S., Abramowicz A., Cristian I., Buscarilli M., Naslavsky M. S., Malaquias A. C., Zatz M., Bodamer O., Majewski J., Jorge A. A., Pereira A. C., Kim C. A., Passos-Bueno M. R., Bertola D. R., Rare variants in SOS2 and LZTR1 are associated with Noonan syndrome. J. Med. Genet. 52, 413–421 (2015). [DOI] [PubMed] [Google Scholar]
  • 99.Carico Z. M., Stefan H. C., Justice M., Yimit A., Dowen J. M., A cohesin cancer mutation reveals a role for the hinge domain in genome organization and gene expression. PLOS Genet. 17, e1009435 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Hoffmann A., Spengler D., Single-cell transcriptomics supports a role of CHD8 in autism. Int. J. Mol. Sci. 22, 3261 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Hurley S., Mohan C., Suetterlin P., Ellingford R., Riegman K. L. H., Ellegood J., Caruso A., Michetti C., Brock O., Evans R., Rudari F., Delogu A., Scattoni M. L., Lerch J. P., Fernandes C., Basson M. A., Distinct, dosage-sensitive requirements for the autism-associated factor CHD8 during cortical development. Mol. Autism 12, 16 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Lasalle J. M., Autism genes keep turning up chromatin. OA Autism 1, 14 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Cotney J., Muhle R. A., Sanders S. J., Liu L., Willsey A. J., Niu W., Liu W., Klei L., Lei J., Yin J., Reilly S. K., Tebbenkamp A. T., Bichsel C., Pletikos M., Sestan N., Roeder K., State M. W., Devlin B., Noonan J. P., The autism-associated chromatin modifier CHD8 regulates other autism risk genes during human neurodevelopment. Nat. Commun. 6, 6404 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Su C., Argenziano M., Lu S., Pippin J. A., Pahl M. C., Leonard M. E., Cousminer D. L., Johnson M. E., Lasconi C., Wells A. D., Chesi A., Grant S. F. A., 3D promoter architecture re-organization during iPSC-derived neuronal cell differentiation implicates target genes for neurodevelopmental disorders. Prog. Neurobiol. 201, 102000 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Harabula I., Pombo A., The dynamics of chromatin architecture in brain development and function. Curr. Opin. Genet. Dev. 67, 84–93 (2021). [DOI] [PubMed] [Google Scholar]
  • 106.Polla D. L., Bhoj E. J., Verheij J., Wassink-Ruiter J. S. K., Reis A., Deshpande C., Gregor A., Hill-Karfe K., Silfhout A. T. V., Pfundt R., Bongers E., Hakonarson H., Berland S., Gradek G., Banka S., Chandler K., Gompertz L., Huffels S. C., Stumpel C., Wennekes R., Stegmann A. P. A., Reardon W., Leenders E., de Vries B. B. A., Li D., Zackai E., Ragge N., Lynch S. A., Cuddapah S., van Bokhoven H., Zweier C., de Brouwer A. P. M., De novo variants in MED12 cause X-linked syndromic neurodevelopmental disorders in 18 females. Genet. Med. 23, 645–652 (2021). [DOI] [PubMed] [Google Scholar]
  • 107.Ramsay E. P., Abascal-Palacios G., Daiss J. L., King H., Gouge J., Pilsl M., Beuron F., Morris E., Gunkel P., Engel C., Vannini A., Structure of human RNA polymerase III. Nat. Commun. 11, 6409 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Qi H., Dong C., Chung W. K., Wang K., Shen Y., Deep genetic connection between cancer and developmental disorders. Hum. Mutat. 37, 1042–1050 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Stratton M. R., Campbell P. J., Futreal P. A., The cancer genome. Nature 458, 719–724 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Halaban R., Krauthammer M., RASopathy gene mutations in melanoma. J. Invest. Dermatol. 136, 1755–1759 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Kiel C., Serrano L., Structure-energy-based predictions and network modelling of RASopathy and cancer missense mutations. Mol. Syst. Biol. 10, 727 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Smith I. N., Thacker S., Seyfi M., Cheng F., Eng C., Conformational dynamics and allosteric regulation landscapes of germline PTEN mutations associated with autism compared to those associated with cancer. Am. J. Hum. Genet. 104, 861–878 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Agell N., Bachs O., Rocamora N., Villalonga P., Modulation of the Ras/Raf/MEK/ERK pathway by Ca2+, and calmodulin. Cell. Signal. 14, 649–654 (2002). [DOI] [PubMed] [Google Scholar]
  • 114.Cerami E., Gao J., Dogrusoz U., Gross B. E., Sumer S. O., Aksoy B. A., Jacobsen A., Byrne C. J., Heuer M. L., Larsson E., Antipin Y., Reva B., Goldberg A. P., Sander C., Schultz N., The cBio cancer genomics portal: An open platform for exploring multidimensional cancer genomics data. Cancer Discov. 2, 401–404 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Gao J., Aksoy B. A., Dogrusoz U., Dresdner G., Gross B., Sumer S. O., Sun Y., Jacobsen A., Sinha R., Larsson E., Cerami E., Sander C., Schultz N., Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci. Signal. 6, pl1 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Gremer L., Merbitz-Zahradnik T., Dvorsky R., Cirstea I. C., Kratz C. P., Zenker M., Wittinghofer A., Ahmadian M. R., Germline KRAS mutations cause aberrant biochemical and physical properties leading to developmental disorders. Hum. Mutat. 32, 33–43 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Munoz-Maldonado C., Zimmer Y., Medova M., A comparative analysis of individual RAS mutations in cancer biology. Front. Oncol. 9, 1088 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Gualtieri A., Kyprianou N., Gregory L. C., Vignola M. L., Nicholson J. G., Tan R., Inoue S. I., Scagliotti V., Casado P., Blackburn J., Abollo-Jimenez F., Marinelli E., Besser R. E. J., Hogler W., Karen Temple I., Davies J. H., Gagunashvili A., Robinson I., Camper S. A., Davis S. W., Cutillas P. R., Gevers E. F., Aoki Y., Dattani M. T., Gaston-Massuet C., Activating mutations in BRAF disrupt the hypothalamo-pituitary axis leading to hypopituitarism in mice and humans. Nat. Commun. 12, 2028 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Singh A., Joshi V., Jindal A. K., Mathew B., Rawat A., An updated review on activated PI3 kinase delta syndrome (APDS). Genes Dis. 7, 67–74 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Moreno-Corona N., Chentout L., Poggi L., Thouenon R., Masson C., Parisot M., Mouel L. L., Picard C., Andre I., Cavazzana M., Perrin L., Durandy A., Azarnoush S., Kracker S., Two monogenetic disorders, activated PI3-Kinase-δ syndrome 2 and smith-magenis syndrome, in one patient: Case report and a literature review of neurodevelopmental impact in primary immunodeficiencies associated with disturbed PI3K signaling. Front. Pediatr. 9, 688022 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Poulin E. J., Bera A. K., Lu J., Lin Y. J., Strasser S. D., Paulo J. A., Huang T. Q., Morales C., Yan W., Cook J., Nowak J. A., Brubaker D. K., Joughin B. A., Johnson C. W., DeStefanis R. A., Ghazi P. C., Gondi S., Wales T. E., Iacob R. E., Bogdanova L., Gierut J. J., Li Y., Engen J. R., Perez-Mancera P. A., Braun B. S., Gygi S. P., Lauffenburger D. A., Westover K. D., Haigis K. M., Tissue-specific oncogenic activity of KRASA146T. Cancer Discov. 9, 738–755 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Jia Y., Jiang T., Li X., Zhao C., Zhang L., Zhao S., Liu X., Qiao M., Luo J., Shi J., Yang H., Wang Y., Xi L., Zhang S., Gao G., Su C., Ren S., Zhou C., Characterization of distinct types of KRAS mutation and its impact on first-line platinum-based chemotherapy in Chinese patients with advanced non-small cell lung cancer. Oncol. Lett. 14, 6525–6532 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Edkins S., O’Meara S., Parker A., Stevens C., Reis M., Jones S., Greenman C., Davies H., Dalgliesh G., Forbes S., Hunter C., Smith R., Stephens P., Goldstraw P., Nicholson A., Chan T. L., Velculescu V. E., Yuen S. T., Leung S. Y., Stratton M. R., Futreal P. A., Recurrent KRAS codon 146 mutations in human colorectal cancer. Cancer Biol. Ther. 5, 928–932 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Ako S., Nouso K., Kinugasa H., Dohi C., Matushita H., Mizukawa S., Muro S., Akimoto Y., Uchida D., Tomoda T., Matsumoto K., Horiguchi S., Tsutsumi K., Kato H., Okada H., Utility of serum DNA as a marker for KRAS mutations in pancreatic cancer tissue. Pancreatology 17, 285–290 (2017). [DOI] [PubMed] [Google Scholar]
  • 125.Cheng H., Liu C., Jiang J., Luo G., Lu Y., Jin K., Guo M., Zhang Z., Xu J., Liu L., Ni Q., Yu X., Analysis of ctDNA to predict prognosis and monitor treatment responses in metastatic pancreatic cancer patients. Int. J. Cancer 140, 2344–2350 (2017). [DOI] [PubMed] [Google Scholar]
  • 126.Tidyman W. E., Goodwin A. F., Maeda Y., Klein O. D., Rauen K. A., MEK-inhibitor-mediated rescue of skeletal myopathy caused by activating Hras mutation in a Costello syndrome mouse model. Dis. Model. Mech. 15, dmm049166 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Zhu G., Xie J., Kong W., Xie J., Li Y., Du L., Zheng Q., Sun L., Guan M., Li H., Zhu T., He H., Liu Z., Xia X., Kan C., Tao Y., Shen H. C., Li D., Wang S., Yu Y., Yu Z. H., Zhang Z. Y., Liu C., Zhu J., Phase separation of disease-associated SHP2 mutants underlies MAPK hyperactivation. Cell 183, 490–502.e18 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Dance M., Montagner A., Salles J. P., Yart A., Raynal P., The molecular functions of Shp2 in the Ras/mitogen-activated protein kinase (ERK1/2) pathway. Cell. Signal. 20, 453–459 (2008). [DOI] [PubMed] [Google Scholar]
  • 129.Dhindsa R. S., Zoghbi A. W., Krizay D. K., Vasavda C., Goldstein D. B., A transcriptome-based drug discovery paradigm for neurodevelopmental disorders. Ann. Neurol. 89, 199–211 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Uddin M., Wang Y., Woodbury-Smith M., Artificial intelligence for precision medicine in neurodevelopmental disorders. NPJ Digit. Med. 2, 112 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Zeng X., Zhu S., Liu X., Zhou Y., Nussinov R., Cheng F., deepDR: A network-based deep learning approach to in silico drug repositioning. Bioinformatics 35, 5191–5198 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Peng H., Zeng X., Zhou Y., Zhang D., Nussinov R., Cheng F., A component overlapping attribute clustering (COAC) algorithm for single-cell RNA sequencing data analysis and potential pathobiological implications. PLOS Comput. Biol. 15, e1006772 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Agazie Y. M., Hayman M. J., Molecular mechanism for a role of SHP2 in epidermal growth factor receptor signaling. Mol. Cell. Biol. 23, 7875–7886 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Ostman A., Hellberg C., Bohmer F. D., Protein-tyrosine phosphatases and cancer. Nat. Rev. Cancer 6, 307–320 (2006). [DOI] [PubMed] [Google Scholar]
  • 135.Keppler-Noreuil K. M., Rios J. J., Parker V. E., Semple R. K., Lindhurst M. J., Sapp J. C., Alomari A., Ezaki M., Dobyns W., Biesecker L. G., PIK3CA-related overgrowth spectrum (PROS): Diagnostic and testing eligibility criteria, differential diagnosis, and evaluation. Am. J. Med. Genet. A 167A, 287–295 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES