Next Article in Journal
GBA Variants and Parkinson Disease: Mechanisms and Treatments
Next Article in Special Issue
Hippocampal Excitatory Synaptic Transmission and Plasticity Are Differentially Altered during Postnatal Development by Loss of the X-Linked Intellectual Disability Protein Oligophrenin-1
Previous Article in Journal
NUP133 Controls Nuclear Pore Assembly, Transcriptome Composition, and Cytoskeleton Regulation in Podocytes
Previous Article in Special Issue
Functions of CNKSR2 and Its Association with Neurodevelopmental Disorders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pathophysiological Heterogeneity of the BBSOA Neurodevelopmental Syndrome

by
Michele Bertacchi
1,*,†,
Chiara Tocco
1,†,
Christian P. Schaaf
2 and
Michèle Studer
1,*
1
Institute of Biology Valrose (IBV), University Côte d’Azur, 06108 Nice, France
2
Institute of Human Genetics, Heidelberg University, 69120 Heidelberg, Germany
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2022, 11(8), 1260; https://doi.org/10.3390/cells11081260
Submission received: 3 February 2022 / Revised: 17 March 2022 / Accepted: 29 March 2022 / Published: 8 April 2022
(This article belongs to the Special Issue Pathophysiological Mechanism of Neurodevelopmental Disorders)

Abstract

:
The formation and maturation of the human brain is regulated by highly coordinated developmental events, such as neural cell proliferation, migration and differentiation. Any impairment of these interconnected multi-factorial processes can affect brain structure and function and lead to distinctive neurodevelopmental disorders. Here, we review the pathophysiology of the Bosch–Boonstra–Schaaf Optic Atrophy Syndrome (BBSOAS; OMIM 615722; ORPHA 401777), a recently described monogenic neurodevelopmental syndrome caused by the haploinsufficiency of NR2F1 gene, a key transcriptional regulator of brain development. Although intellectual disability, developmental delay and visual impairment are arguably the most common symptoms affecting BBSOAS patients, multiple additional features are often reported, including epilepsy, autistic traits and hypotonia. The presence of specific symptoms and their variable level of severity might depend on still poorly characterized genotype–phenotype correlations. We begin with an overview of the several mutations of NR2F1 identified to date, then further focuses on the main pathological features of BBSOAS patients, providing evidence—whenever possible—for the existing genotype–phenotype correlations. On the clinical side, we lay out an up-to-date list of clinical examinations and therapeutic interventions recommended for children with BBSOAS. On the experimental side, we describe state-of-the-art in vivo and in vitro studies aiming at deciphering the role of mouse Nr2f1, in physiological conditions and in pathological contexts, underlying the BBSOAS features. Furthermore, by modeling distinct NR2F1 genetic alterations in terms of dimer formation and nuclear receptor binding efficiencies, we attempt to estimate the total amounts of functional NR2F1 acting in developing brain cells in normal and pathological conditions. Finally, using the NR2F1 gene and BBSOAS as a paradigm of monogenic rare neurodevelopmental disorder, we aim to set the path for future explorations of causative links between impaired brain development and the appearance of symptoms in human neurological syndromes.

1. NR2F1 as a Master Regulator of Brain Development and Function

Neurodevelopmental disorders (NDDs) of genetic origin are a highly heterogeneous group of pathological syndromic conditions caused by defects in the basic mechanisms of brain development, such as cell proliferation, migration and differentiation, as well as function and connectivity. The imbalance of these early developmental events can lead to structural and functional brain defects affecting cognitive functions, such as visuospatial processing, motor execution, learning and memory, attention and social skills. They can also represent an underlying risk factor for complex psychiatric conditions, including anxiety and depression.
A rare autosomal dominant neurodevelopmental disorder called Bosch–Boonstra–Schaaf Optic Atrophy Syndrome (BBSOAS), was first reported in 2014 on six patients [1] and then expanded over the past few years by clinicians all over the world (OMIM #615722). With an estimated prevalence between 1 in 100,000 to 250,000 people worldwide, BBSOAS has been so far diagnosed in more than 100 patients. However, new patients are reported every year, suggesting that this proportion could be an underestimation. BBSOAS symptoms are very heterogeneous and include optic nerve atrophy (OA) or optic nerve hypoplasia (ONH), cortical visual impairment (CVI), moderate to severe intellectual disability (ID), developmental delay, hypotonia, seizures, speech difficulties, motor dysfunctions, autism spectrum disorder (ASD) and others. It is the peculiar combination of these diverse symptoms, and particularly the CVI and OA, that differentiate BBSOAS patients from those affected by other NDDs with similar features. After the first description of BBSOAS, patients with NR2F1 mutations from distinct cohorts could finally be grouped into a single clinical category under the name of a novel syndrome.
All BBSOAS patients identified to date present haploinsufficiency of the nuclear factor NR2F1, due mainly to de novo missense mutations or whole-gene deletion of only one of the two alleles. Nevertheless, some rare cases of inherited variants have been described. Complete NR2F1 absence has never been reported to date, suggesting that this condition might be incompatible with life, as reported in mouse mutant models showing perinatal lethality upon loss of both Nr2f1 copies [2]. NR2F1 gene codes for a transcriptional regulator functioning in the form of a dimer and able to either activate or repress target gene expression, depending on the cellular and biological context [3,4,5]. Interestingly, pathogenic BBSOAS point mutations are principally located in the two most conserved functional domains of the protein: the DNA-binding domain (DBD), responsible for the interaction with target gene regulatory sequences, and the ligand-binding domain (LBD), necessary for dimerization and co-factor binding. Even when the same symptoms are shared by multiple BBSOAS patients, their severity can be variable, possibly depending on the specific type of NR2F1 gene perturbation, suggesting the existence of a genotype-phenotype correlation.
The pivotal role of NR2F1 for correct development is supported by its high degree of evolutionary conservation in the animal kingdom, with high protein homology between mice and humans. For this reason, mouse gain-of-function and loss-of-function models have been extensively used to explore the functions of this gene in brain development (reviewed in [5]). Mouse Nr2f1 orchestrates different aspects of brain morphogenesis and maturation: acquisition of neocortical areal identity [6,7], control of proliferation and neurogenesis for laminar and cell-type specification in the cortex [6,8,9,10], proper hippocampal formation and synaptic plasticity [11,12,13], long-range migration of cortical interneuron subtypes [14,15], patterning of distinct domains in the developing eye and optic nerve [16,17], temporal control of neural progenitor competence [18,19] and activity-dependent shaping of pyramidal neuron morphology [20], among many other processes. The highly similar expression profile of mouse Nr2f1 and human NR2F1 suggests that the human gene could play similar roles [8,21,22,23,24]. However, whether the same functions have been preserved in human development is still unknown. Linking Nr2f1 roles in mouse brain assembly with the clinical features described so far in BBSOAS patients could help in further converging diagnostic aspects on present and new patients, as well as unravelling the multi-faceted functions of this gene during brain development in both physiological and pathological conditions.
In this review, we have summarized the available clinical knowledge on individuals affected by BBSOAS, focusing on their predominant symptoms. We supply an updated list of both congenital and acquired BBSOAS features and propose a unified approach for early therapeutic intervention. By linking the clinical knowledge to the available experimental approaches, we revise the state of the art of cellular and animal models used to explore, both in vitro and in vivo, Nr2f1 contribution to early brain development. Furthermore, focusing on the genotype–phenotype correlation that is starting to appear as new patients are reported, we describe the molecular functioning of NR2F1 dimerization and speculate on the possible pathogenic mechanisms originating upon NR2F1 mutation, which could explain a dominant-negative effect during dimer formation. By highlighting what have been so far described in terms of cellular and molecular mechanisms and what is still missing to achieve a full comprehension of BBSOAS pathogenesis, we propose future experimental paths that could translate discoveries from the bench to the clinic.

2. From Distinct Cohorts to a Unified List of NR2F1-Haploinsufficient BBSOAS Patients

Due to the heterogeneity of symptoms, both in terms of presence and severity, a clear diagnosis of BBSOAS has been elusive until recently. In fact, long before NR2F1 haploinsufficient patients were grouped in this syndrome, BBSOAS patients were included in autism, epilepsy or other clinical cohorts, depending on their most prominent pathological features. As an example, a patient with an NR2F1 mutation in the DBD was initially identified in an epilepsy study on infantile spasms [25]. Similarly, a patient with prominent autism-like behaviors was reported after an unbiased search for de novo gene variants associated with autism [26]. The causative link between NR2F1 haploinsufficiency and complex syndromic conditions remained unclear in the first clinical reports, as some of these patients were carrying large deletions also involving additional genes [27,28,29]. The first BBSOAS study, with patients selected from a cohort of individuals with CVI, finally reported loss-of-function point variants, proving once and for all the status of NR2F1 as a disease gene and its causative link with BBSOAS symptoms [1,30]. Located on chromosome 5, the NR2F1 gene is expressed in several tissues and organs, including the eye and the brain (Figure 1). Despite being initially described as an extremely rare syndrome, BBSOAS has now been diagnosed in more than 100 patients (Figure 1; complete list in Table 1). Novel cases are regularly reported and NR2F1 is also recognized as an ASD candidate gene according to the Simons Foundation Autism Research Initiative (SFARI) database [31]. So far, a total of 112 NR2F1 variants has been reported [32], and 92 BBSOAS patients have been clinically described, comprising 43 males and 46 females (Table 1). Most of the identified genetic variants have been diagnosed as de novo (73.9%), but some familial cases are also present (7.6%). Patients harboring other genetic variants or large deletions involving other genes are nonetheless included in Table 1. While 15 patients (16.3%) have small to large deletions involving NR2F1 alone or together with adjacent genes, most of them (83.7%) harbor NR2F1 point variants or small in-frame deletions. Among the latter group, 32 variants (34.8% of all patients) fall in the DBD, 17 (18.5%) in the LBD and 9 (9.8%) in the starting codon, with a single duplication variant falling out of the main functional domains. Additionally, eleven truncation and seven frameshift/truncation variants have been reported (12% and 7.6% of patients, respectively).
Each of the listed BBSOAS patients displays a peculiar combination of visual deficits with ID, DD, ASD and other neurological features that make them stand out as unique when compared to the clinical picture of other distinct neurodevelopmental syndromes. We describe such clinical phenotypes in detail in the next section.

3. Towards a Consensus List of NR2F1-Related Symptoms and Standardized Therapeutic Interventions for BBSOAS Patients

Initially identified for their optic nerve anomalies—notably optic atrophy— BBSOAS patients differentiate themselves from visually deficient patients with a combination of additional neurological symptoms creating a much more complex clinical picture (Figure 2), the heterogeneity of which represents a general hallmark of NDDs. The increasing number of BBSOAS patients reported to date allows the depiction of a general profile of common clinical features resulting from NR2F1 loss-of-function. While the condition is mainly characterized by vision impairment (OA and other optic nerve abnormalities), CVI, developmental delay (defined as a delay in milestone acquisition in distinct domains, such as walking and first words) and ID [1], recent reports have revealed additional clinical characteristics, including hypotonia, infantile spasms and other forms of epilepsy, ASD or autistic traits, feeding difficulties due to oromotor dysfunction and hearing defects, among others [30,33,34,35,36,37,38,39,40]. Together with OA and CVI, the vision/eye phenotype additionally comprises ONH, manifest latent nystagmus/fusional maldevelopment and alacrima [38]. Going into the details of developmental delay and ID, most common features are low intelligence quotient (IQ), speech delay, behavioral abnormalities and fine motor skill impairment [38]. There are few exceptions to the speech delay, as some patients have been reported showing normal or even superior verbal abilities [1,33], suggesting that some cases of BBSOAS speech delay could depend more on impaired motor coordination than on intellectual disability per se.
Further, the more recent and comprehensive reports have also delineated additional features, such as high pain tolerance, remarkable long-term memory, prominent love for music, sleep difficulties and touch sensitivity [38], and novel brain morphological malformations, such as neocortical dysgyria/polymicrogyria in areas involved in language and number processing [8]. Moreover, a recent detailed evaluation of the visual deficits in BBSOAS patients suggested a stable, non-progressive reduction in visual acuity [41]. Finally, new, rarer symptoms were reported in isolated patients, such as mitochondrial deficit, psychosis, transient ataxia and protein-losing enteropathy [42,43,44]
Since the most common BBSOAS symptoms, either appearing at birth or developing during the first years of life, have now been described (Table 2), a panel of recommended clinical tests has been compiled by neuropediatricians for the follow-up of newly diagnosed individuals (Table 3), together with therapeutic approaches to be considered for early intervention (Table 4). However, as our knowledge on the phenotypic spectrum continues to grow, thanks to studies on larger groups of patients with NR2F1 pathogenic variants [30,38], these lists are constantly updated and refined. Hence, families of recently diagnosed BBSOAS children should refer to up-to-date information available online at https://nr2f1.org (USA NR2F1 Foundation; URL accessed date: 17 March 2022) and https://nr2f1france.wordpress.com (French NR2F1 Association; URL accessed date: 17 March 2022).
In clinical practice, physicians should consider a diagnostic evaluation for BBSOAS in every person with ONH or OA in combination with developmental delay and ID. This implies that some of the key features of BBSOAS will only become evident over the course of the first years of life. During infancy, the symptoms of BBSOAS are less specific. Here, the diagnosis may be suggested in children with hypotonia, feeding difficulties, epilepsy and signs of visual impairment, such as poor eye tracking. Given that there is no pathognomonic feature of BBSOAS and considering the great genetic heterogeneity of NDDs such as BBSOAS, it goes without saying that the vast majority of individuals will be diagnosed based on genome-wide technologies: chromosome microarray analysis for deletions and whole-exome sequencing or large next-generation sequencing panels for single nucleotide changes and small indels (see Discussion).
Each of the main BBSOAS pathological features will be discussed in the following sections, focusing on their variability in both presence and severity, as well as cellular and molecular insights coming from in vitro and in vivo experimental models, which have advanced our understanding of NR2F1 function during brain development. Further, in a conclusive chapter, we will examine what is known so far about the correlation between genotype and phenotype in patients carrying different pathogenic NR2F1 variants and speculate on the molecular mechanisms potentially impacting NR2F1 dimer formation and being responsible of this phenomenon.

4. Optic Atrophy (OA), Optic Nerve Hypoplasia (ONH) and Non-Progressive Reduction in Visual Acuity: BBSOAS as a Non-Canonical Optic Neuropathy

Inherited optic neuropathies are an important cause of visual impairment in young children with an estimated prevalence of 1 in 10,000 [45]. Although genetically heterogeneous with both nuclear and mitochondrial genes being implicated, the pathological hallmark is a pronounced vulnerability of retinal ganglion cells (RGCs), ultimately leading to ON degeneration and irreversible vision loss [45,46,47]. This phenomenon can be monitored by high-resolution optical coherence tomography (OCT). BBSOAS can be defined as a genetic optic neuropathy in light of its several and variable clinical features affecting the visual system, such as OA or optic nerve pallor, ONH or small optic discs, CVI, nystagmus (uncontrolled eye movements) and alacrima (decreased tear reflex) [1,30,38]. These features can be present alone or as comorbidities, with a severity degree that presumably varies depending on the type of NR2F1 genetic perturbation.
One of the main and first-reported visual impairments in BBSOAS children is OA, as the name of the syndrome suggests. It can be defined as ON damage anywhere from the retina to the lateral geniculate nucleus of the thalamus, usually caused by RGC death and retraction of RGC axons, with a resulting pale ON in fundoscopy. Notably, OA is particularly common in patients bearing DBD variants (78%), translation initiation variants (78%) and frameshift/truncations (72%), while it is less represented in milder cases caused by LBD variants (47% of patients). However, since the first description of BBSOAS patients, OA and ONH have both been reported [1]. As OA, ONH is also characterized by a deficiency of RGCs and their axons, leading to ganglion cell layer disorganization and a small optic disc with a thin ON. ONH is often associated with poor fixation, abnormal eye movements, nystagmus, strabismus, hyperopia [48] and vision ranging from no light perception to good functional vision [49,50]. Various theories have been proposed to explain the etiology of ONH, including a developmental failure of RGCs [51,52,53,54]. The description of ONH may sometimes be confused with OA, and their overlap may lead to diagnostic challenges [55,56,57].
A main difference between ONH and OA is that while ONH is a congenital, non-progressive disease characterized by underdevelopment of the ON, OA is instead degenerative with a normal early development of the ON that then deteriorates over time. As an example, in autosomal dominant or recessive OA caused by pathogenic OPA1 (OMIM 605290) and WFS1 (OMIM 606201) variants, progressive RGC loss starts in early childhood, and most patients are registered as legally blind by the fifth decade of life [58]. Furthermore, ONH is often syndromic, and occurs in conjunction with other neurodevelopmental abnormalities, such as brain malformations, developmental delay, ID and/or ASD [59,60,61]. The most common neuroanatomical malformation found in patients with ONH is hypoplasia of the corpus callosum, associated with developmental delay, neurological deficits and seizures [60,62], all clinical features also characteristic of BBSOAS children [3,38]. Genetically, syndromic ONH can be caused by perturbations of key transcription factors orchestrating brain, eye and ON development, such as SOX2, PAX2 and PAX6 [63,64,65]. In summary, while ONH has a developmental origin and is often syndromic, stable and non-progressive, OA is instead a progressive, i.e., degenerative, RGC axonal loss due to both genetic and environmental reasons.
Experimental studies in Nr2f1-deficient mouse models can help in filling the lack of clinical data due to the rarity of BBSOAS patients (Figure 3) and contribute to discerning OA from ONH in a BBSOAS-like context. Thanks to a high degree of evolutionary conservation [66], mouse Nr2f1 and human NR2F1 are similarly expressed in different retinal cell types, RGCs included [16], allowing the use of constitutive mouse models with Nr2f1 haploinsufficiency as a model to study BBSOAS-related visual impairments [3,16]. Mouse Nr2f1 retinal expression controls the density of RGCs, which is reduced in mutants due to a differentiation delay during embryogenesis and excessive apoptosis around the time of birth [16,41]. The early decrease in RGCs ultimately affects the number of axonal fibers in the ON, resulting in OA. Besides being atrophic, mutant ONs also show low levels of myelination due to a delay in oligodendrocyte migration and maturation at both embryonic and post-natal stages [16,67]. Interestingly, the defect appears to be restored upon treatment with Miconazole, a chemical drug promoting oligodendrocyte proliferation and maturation [16,68,69,70,71,72], thus proposing a possible therapeutic approach. Mouse mutant ONs are further impacted by an inflammatory process involving ON astrocytes at embryonic stage and in early post-natal life [16]. This early ON inflammation is of particular interest, as it could impact RGC survival and/or worsen myelination defects, as it happens in inflammatory forms of human optic neuritis [73,74]. In a reinforcing feedback loop, hypomyelination could in turn increase the loss of ON axonal fibers and retinal RGC apoptosis [75]. Likely due to these multiple concurrent and early developmental causes (inflammation, hypomyelination, RGC apoptosis and loss of RGC axons), a significant delay in the conduction velocity of visual stimuli along the ONs of mutant animals was revealed by electrophysiological recordings [16]. RGC cells apart, the development of other retinal cell types, such as cone photoreceptors [76], amacrine and bipolar cells [77], seems also to be Nr2f1-dependent, suggesting that its dysfunction could affect color vision and visual acuity by impacting on distinct retinal cell types.
In summary, according to experimental data in Nr2f1-deficient mouse models, BBSOAS-like visual impairments, such as retinal and ON anomalies, result from primary RGC patterning, myelination and inflammatory defects originating during early development (Figure 3) [16,41], and leading to visual system defects that would remain stable during postnatal life [41]. Consistently with early appearance of BBSOAS, ONH could be the major clinical feature accounting for the non-progressive visual dysfunction observed in patients [41]. We propose that since BBSOAS differs from classical optic neuropathies, such as OA, in which vision loss is progressive and visual acuity decreases in time, this syndrome should be defined as a non-progressive ONH with early (embryological) developmental origin. However, more studies in a higher number of patients and during a longer follow-up period will be necessary to confirm the stationary nature of visual clinical features, as current reports cannot exclude a progressive worsening later in a patient’s lifespan.

5. A Genetic Control of Optic Disc Abnormalities in BBSOAS

Several BBSOAS patients show small malformations of the optic disc (OD) at the edge between the neural retina and the ON. Such malformations manifest in excavated, pale or small ODs. Often associated with other congenital eye malformations, OD lesions can negatively affect visual acuity. Interestingly, OD malformations are reported more frequently in patients with NR2F1 deletions (33%) than in patients with other variants (12–22%). A structural change in the form of pale and/or excavated ODs is diagnosed by ophthalmologists as an indirect sign of underlying ON diseases, including OA or ONH. In fact, OD pallor and abnormal shape can result from RGC death (causing axonal fibers loss) and degeneration of pial capillaries entering the optic nerve head [57,78,79]. However, some OD malformations—such as optic disc coloboma—can be the direct consequence of genetic or environmental insults affecting the development of this region, rather than a consequence of ON disease [80,81,82]. The OD pit, for example, consisting in a round or oval localized depression within the OD, is caused by defective occlusion of the embryonic ventral fissure of the optic vesicle [83,84]. By homology, data from Nr2f1 mouse models showed delayed ventral fissure closure in null embryos [16] or severe coloboma in double Nr2f1 and Nr2f2 conditional mutants [17], suggesting that impaired ventral fissure fusion could be responsible for OD malformations in a Nr2f1-deficient context. This would also fit with reports of BBSOAS patients showing cases of coloboma-like malformations [29,30,85].
Furthermore, data in mice revealed a molecular network controlled by Nr2f1 and driving the establishment of the border between the Pax2-expressing optic stalk (i.e., the presumptive ON) and the Pax6-expressing retinal vesicle (i.e., the presumptive neural retina), ultimately allowing the development of the OD between these two regions (Figure 3) [4,17]. One key function of the OD is to produce signaling molecules for RGC axonal guidance, such as Netrin-1 [86,87]. Nr2f1 constitutive mutants demonstrated how the OD genetic network is disrupted by loss of Nr2f1 alone, resulting in Pax6 overexpression at the expense of the Pax2 domain [16]. This results in a shift of the border between NR and OS, with heavy consequences for Netrin-1 expression, ultimately impacting axonal guidance of RGC fibers exiting the eyeball [41]. However, defects in Nr2f1 haploinsufficient animals appear quite subtle compared to OD malformations in BBSOAS patients, suggesting that additional factors could impact OD development in humans and/or that species-specific differences are present.
The involvement of Pax2 in an Nr2f1-regulated network is particularly important, as human PAX2 mutations lead to coloboma-like OD malformations in the context of syndromic conditions [81,88,89], consistent with a key role of the Nr2f1-Pax2-Pax6 genetic network for the correct establishment of the OD region. Human NR2F1 could indeed control OD development by modulating PAX2 and PAX6, as it does in mouse, but further experiments on human cells are necessary to verify the evolutionary conservation of such a network.

6. When Intellectual Disability Meets Visual Disfunction: Cortical Visual Impairment (CVI) in BBSOAS Patients

The condition of CVI (affecting around 42% of BBSOAS patients, and particularly common in the ones bearing DBD variants) is a bilateral visual impairment due to a non-ocular cause in the presence of normal pupil reactivity and characterized by abnormal perception, elaboration and interpretation of visual stimuli [90,91]. Despite the presence of nystagmus, ON atrophy and other structural eye anomalies, the degree of visual impairments in BBSOAS patients with CVI exceeds what would be expected from eye abnormalities alone, implying that higher-order visual centers in the brain (such as the retro-chiasmatic visual pathways, the thalamus, the primary cortex and/or the secondary associative visual cortices) might also be affected [38,91,92,93,94,95,96,97,98]. In BBSOAS patients, poor visual acuity and visual field abnormalities support the diagnosis of CVI. For instance, patients have narrowed visual fields, low visual acuity, and difficulties with distance viewing, following fast moving scenes or recognizing objects in crowded environments (Table 2). Among the possible factors contributing to CVI in BBSOAS, impaired thalamic connection with parietal and occipital cortices has been hypothesized to impact visuospatial ability [33], together with aberrant connections of major fasciculi relaying the occipital lobe to adjacent associative areas [41].
Nr2f1 mouse mutant models challenged the involvement of central thalamic and neocortical structures at the anatomical, electrophysiological and behavioral levels (Figure 3). The size of visual thalamic nuclei is affected upon Nr2f1 loss, in turn acting on the maturation of primary and secondary visual areas as a result of impaired afferent thalamocortical innervation [2,7,99]. At cortical level, the Nr2f1 high caudolateral to low rostromedial expression gradient is key for a dose-dependent establishment of distinct neocortical areal identities, a process termed arealization [100,101]. Upon mouse Nr2f1 loss, the rostral motor area expands at the expense of caudal sensory ones, so that the primary visual area (V1) tends to be compressed caudally in Nr2f1 mutant neocortex, with a striking effect in null (homozygous) brains and a less pronounced area shift in heterozygous HET animals [2,7,66]. Electrophysiological recordings demonstrated a delay in transmitting peripheral visual stimuli along the visual pathway, together with a decrease in amplitude in the visual thalamic nucleus, the superior colliculi and the V1 superficial cortical layers of HET mice [16,41]. Suboptimal conduction velocity of visual stimuli along the visual pathway could depend on ONH associated with hypomyelination and gliosis [16]. Furthermore, visual acuity is decreased in Nr2f1 HET mice compared to wild-type littermates, in line with low visual acuity and nondegenerative vision loss in BBSOAS patients [1,30,38,41]. Finally, CVI could also result from impaired elaboration of visual information in secondary associative cortices. Interestingly, while Nr2f1 haploinsufficient mice are still able to learn and execute complex tasks [12,16], they nevertheless fail to associate the visual stimulus with a specific operating task when challenged with a light-dependent operating procedure, somehow recapitulating a deficit in the interpretation of visual stimuli reminiscent of patients’ CVI [16]. This suggests a specific impairment in the perception and elaboration of visual stimuli in high-order cortices, rather than a generalized defect in learning and cognitive function. Altogether, mouse models point to a scenario in which multiple structures along the visual pathway are impacted by Nr2f1-deficiency, from the retina and ON up to the thalamus, the V1 and visual associative cortices together with their connecting tracts.
Interestingly, human NR2F1 shows similar graded expression along neocortical axes [2,7,21,23], but it is still unknown whether BBSOAS patients display analogous defects in the positioning, size and function of primary and secondary visual areas; functional magnetic resonance imaging (fMRI) analysis and brain tractography could help to tackle this issue in future clinical studies. Undoubtedly, impaired visual learning and aberrant visual processing due to reduced NR2F1 dosage constitute one of the main features of BBSOAS intellectual deficits, and there is an urgent need of specific early therapeutic interventions to help BBSOAS patients in reaching visual developmental milestones.

7. The Many Converging Roads of Intellectual Disability in BBSOAS Patients: From Corpus-Callosum Thinning to Hippocampal and Neocortical Malformations

The second-most prevalent condition in BBSOAS is a moderate to severe ID [38], found in 87% of patients and particularly common upon DBD mutations (94% of patients). Interestingly, even the milder BBSOAS genetic categories, such as patients bearing LBD variants or deletions, show high incidence of ID and speech delay (70% and 80% of described patients, respectively). Despite a high frequency of cognitive symptoms, brain morphological malformations underlying ID in BBSOAS patients are still poorly characterized. Due to the high prevalence of visual deficits in BBSOAS, clinicians tend to focus more on optical features when analyzing MRI brain scans [1,30]. Moreover, the difficulty in obtaining high-resolution MRI scans from young patients also hampers the acquisition of morphological data. Nevertheless, morphological brain malformations have started to emerge, such as a thin corpus callosum (CC) with general presentation or restricted to posterior regions, hippocampal malrotation or dysmorphia, rare white matter loss (demyelination in 14% patients) and cases of localized dysgyria or megalencephaly.
The CC is the largest commissure connecting brain hemispheres, allowing integration between the two cerebral halves. Notably, CC thinning or agenesis could directly contribute to ID, visual problems, motor impairment, speech delay and seizures described in patients [102,103,104]. Consistently, mouse data have demonstrated the role of Nr2f1 during CC formation as a regulator of the differentiation and migration of late-born cortical neurons [105]. Callosal Nr2f1-deficient neurons, found in reduced numbers at postnatal ages, also fail to elongate their axons and topographically innervate the contralateral hemisphere [105,106], ultimately leading to a thinning of the CC, as in human patients. Furthermore, upon fiber elongation via axonal growth, long-range tracts, such as the CC, must be properly myelinated for optimal signal conduction; impaired myelination is associated with human clinical conditions, such as demyelinating diseases or white matter disorders leading to ID, seizures, lack of coordination and other neurocognitive consequences [107]. In mutant mouse models, Nr2f1 regulates myelination levels in the brain [16,67], similarly to what is observed in BBSOAS patients, suggesting a key role for Nr2f1 in the maturation of oligodendrocytes and resulting myelination process.
Alongside CC thinning, specific morphological hippocampal defects have been reported in some BBSOAS patients. Since the hippocampus is a key structure for learning, memory and other cognitive processes [108,109,110] and defects in its development can negatively impact such cognitive performances [111,112,113,114], it is reasonable to hypothesize a hippocampal component underlying ID in BBSOAS patients. Mouse experiments support this hypothesis, as Nr2f1 mutants show defective hippocampal morphogenesis and function [11,12,13]. By regulating both hippocampal progenitor proliferation and neuronal migration during embryonic and early post-natal development [13], Nr2f1 mainly acts on the development of the dorsal-most hippocampal regions [12], which are involved in spatial navigation, learning and memory [115]. In addition, electrophysiological data revealed impairment of two cellular correlates of learning and memory, long-term potentiation and long-term depression in a Nr2f1 HET mouse model [11], suggesting that altered synaptic plasticity may also contribute to BBSOAS intellectual impairment. However, mutant mice also manifest a prolonged fear-memory retention [11], which is in line with unusually good long-term memory reported in up to 75% of BBSOAS patients [38]. This suggests that NR2F1 might play an important role in the retention of memory in both mouse and human hippocampus. The exact impact of hippocampal abnormalities on BBSOAS patients is still unknown to date, and more studies will be necessary to further assess the contribution of other cortical or subcortical regions.
Recently, a novel cohort of six BBSOAS children showed specific cortical morphological defects, such as polymicrogyria- or dysgyria-like malformations in the supramarginal and angular gyri [8], regions specifically involved in language, vision, spatial cognition, memory retrieval, attention and number processing [116,117,118]. Such malformations were sometimes associated with elongated occipital convolutions, reminiscent of a mild caudal megalencephaly [8]. In mice, Nr2f1 fine-tunes neural proliferative potential by controlling cell-cycle progression and balancing progenitor pool amplification and local neurogenesis [8,9]. Upon Nr2f1 loss, a delay in neurogenesis results in amplification of distinct progenitor classes, which ultimately causes caudal cortical expansion [8] and increased overall neocortical volume [11]. The occipital expansion reproduced in mice is somehow reminiscent of the elongated occipital convolutions and megalencephaly reported in some BBSOAS patients, suggesting that common cellular and molecular mechanisms might be shared between mouse and human.
Both the presence of cortical malformations in BBSOAS patients and data coming from mouse models link the NR2F1 gene to a heterogeneous group of NDDs, called malformations of cortical development (MCDs). MCDs comprise a variable class of morphological abnormalities, such as polymicrogyria and macrocephaly, often associated with ID, autism, speech and motor difficulties and/or epilepsy [119,120,121,122]. Hence, these malformations could underlie some of the key BBSOAS cognitive disorders. However, whether they are a common feature of BBSOAS is still unknown, and a more detailed examination of MRI scans in larger cohorts, with a special focus on occipital and parietal regions, will be needed in future analysis. Furthermore, mouse models could only partially recapitulate these morphological defects [8], as they lack specific human-like features, such as cortical convolutions and abundance of a specific class of neural progenitor termed the basal radial glia, highly present in gyrencephalic species [123,124,125,126,127]. Studies in gyrencephalic mammals, such as the ferret [128], or in human-like systems such as the induced pluripotent stem cell (iPSC)-derived brain organoids [129,130,131], could help unravel NR2F1 function in mammalian neural proliferation and neocortex gyrification.

8. NR2F1 as an Autism Spectrum Disorder Gene

Autism spectrum disorder (ASD) or autistic traits constitute another major clinical feature of BBSOAS [38]. ASD is a complex group of NDDs characterized by qualitative impairments in social interaction and communication, with repetitive/stereotyped patterns of behavior, interests and activities. Several genomic studies have unveiled a link between disruptive variants in NR2F1 gene and high susceptibility to ASD phenotypes [26,132,133,134]. As a result, NR2F1 is now classified as an ASD gene with “suggestive evidence” in the SFARI (Simons Foundation Autism Research Initiative) database [26,31,135]. However, while some BBSOAS patients officially meet the requirements for an ASD diagnosis (38% diagnosed with ASD), some of them have at least autistic features (14.1% patients with autistic features but no official ASD diagnosis) [38], such as repetitive/stereotyped movements in the form of head-banging and hand-flapping, repetitive language, circumscribed interests and self-injurious behaviors, among others [136]. The highly variable behavioral phenotype of BBSOAS also includes attention deficit hyperactivity disorder (ADHD), obsessive compulsive disorder (OCD), pervasive developmental disorder-not otherwise specified (PDD-NOS) and, less commonly, psychosis in the form of auditory hallucinations [42].
Due to the limited number of reported patients, together with the variable expressivity of the BBSOAS behavioral spectrum, it is challenging to distinguish behavioral abnormalities due to ASD or to other clinical features, such as visual impairment [137], epilepsy [138] or ID [139,140,141,142]. As an example, children with visual impairments might display head-banging behaviors [137]. Furthermore, some stereotyped movements in patients with ID and concurrent epilepsy could be the result of unrecognized seizure activity [138], and in general, higher rates of disruptive and self-injurious behavior can be found in patients with combined ASD and epilepsy [143]. Additionally, ID is often associated with stereotyped and aggressive behaviors [143,144]. Finally, obsessive-compulsive behaviors typical of ASD and present in some BBSOAS patients partially overlap with typical symptoms associated with OCD [145]. The most probable scenario for BBSOAS is that multiple cognitive impairments—ASD, ID and epilepsy, among others—could converge on a common (although variable) behavioral spectrum, and overlapping pathologies could synergistically lead to shared phenotypes [146]. Hence, to better discern the different aspects of cognitive impairments characteristic of the BBSOAS behavioral spectrum, further neuropsychological assessments are needed.
As for other BBSOAS features, the use of mouse models has helped to unravel possible etiological mechanisms underlying patients’ autistic features. For instance, some cortical morphological changes reported in Nr2f1 mutant mice are the result of a regional-specific control of neurogenesis, which, when impaired, might lead to local megalencephaly and/or abnormal gyrification [8,9]. Moreover, morphological changes are also associated with altered neocortical identity upon Nr2f1 loss, with posterior sensory areas being reduced in size, and a large cortical surface acquiring anterior motor-like properties [7,10]. Interestingly, arealization defects—especially concerning the frontal lobe—regionally altered gyrification of cortical areas and imbalanced neurogenesis leading to early expansion of the brain have all been linked with ASD in patients [147,148,149]. Hence, BBSOAS autistic features and other cognitive impairments could be influenced by impaired arealization upon human NR2F1 perturbations, but further investigation by fMRI is required to investigate potential areal functional impairments.
Another interesting hypothesis is that ASD and other neurodevelopmental impairments described in BBSOAS patients could derive from an excitatory/inhibitory (E/I) dysregulation due to an imbalance between dorsally generated excitatory glutamatergic neurons and ventrally generated inhibitory interneurons [150,151,152,153,154]. In this respect, a recent report introduced a patient-specific DBD variant in mouse Nr2f1, in an attempt to reproduce a genuine point-mutation mouse model of BBSOAS [155]. Mutated Nr2f1 promoted differentiation of inhibitory neurons, while concomitantly reducing the rate of production of glutamatergic ones, resulting in autistic-like behavioral deficits, such as impaired social interaction. Interestingly, the observed behavioral deficits could be partially alleviated by antagonizing the excessive inhibitory synaptic transmission [155]. Consistently, mutations affecting other ASD risk genes have been shown to trigger a similar E/I imbalance through distinct molecular pathways ultimately converging into shared neurodevelopmental abnormalities [156]. In summary, while the exact molecular and cellular mechanisms leading to autistic phenotypes are still under investigation, we expect NR2F1 to be soon considered a bona fide ASD gene.

9. NR2F1 as a Susceptibility Gene for Infantile Epileptic Disorders

Although not always present (46% of BBSOAS patients), another noteworthy BBSOAS feature is the insurgence of various epileptiform pathologies, comprising infantile and febrile spasms, West syndrome and other forms of epilepsy in children [1,30,38]. Particularly common in DBD patients (53%), epileptic features are more rarely reported in patients with LBD variants or deletions (29% and 27%, respectively), suggesting that epilepsy could represent a BBSOAS feature specifically associated with more severe genetic variants. In some cases, altered electroencephalogram (EEG) patterns—often restricted to the occipital hemispheres—were recorded even in the absence of epileptic seizures [85]. The presence of early-onset epileptic forms in infancy (such as infantile spasms) is of particular importance, as these so-called “catastrophic epilepsies” are associated with poor neurodevelopmental outcome and could inflict additional damage to the developing brain, contributing to ID, memory impairment, attention deficits, developmental delay and autistic features [157,158,159,160,161,162]. The discrimination between brain injuries leading to or caused by seizures has always constituted a great challenge for clinicians [163,164].
The disruption of distinct developmental processes has been hypothesized to underlie the onset of epilepsy [165]. As an example, an imbalance of excitatory pyramidal neurons and inhibitory interneuron subtypes has been proposed as common potential cause of epilepsy and autism [166,167]. Alternatively, interneurons could promote seizures in the initial phase of epileptic activity [168], suggesting that epileptiform events could originate from GABAergic dysfunction per se, regardless of their excitatory glutamatergic counterpart. Interestingly, mouse Nr2f1 is known to regulate the number and type of GABAergic interneurons produced in the ventral telencephalon and reaching the cortex [14,15]. Further, epileptic discharges can also be caused by improper electrophysiological activity of neural circuits due to intrinsic electrical properties of glutamatergic neurons and astrocytes [169,170]. Recent studies show that mouse Nr2f1 controls the intrinsic electric properties of pyramidal neurons by directly regulating the expression of distinct voltage-gated ion channels as well as the axon initial segment length and diameter [20,171]. By allowing a fine-tuning of action potentials and proper modulation of spontaneous network activity during circuit maturation, mouse Nr2f1 can ultimately sculpt the emergence of electrical activity in cortical networks [20].
Finally, drug-resistant forms of epilepsy can also be caused by nodular periventricular heterotopia [172,173,174,175], ectopic clusters of grey matter caused by impairment of neuronal migration and connectivity. Defective migration of neuronal cells has been linked to Nr2f1 loss in mouse [105], possibly due to altered expression of specific cytoskeletal proteins [106]. Consistently, periventricular heterotopia has been reported in some BBSOAS patients alongside seizures [8,29], suggesting a cause–effect link between these two phenomena, whereby clusters of misplaced neurons could act as triggering foci.
The NR2F1 gene is starting to emerge from unbiased exome-sequencing approaches in individuals with epilepsy [25,176,177,178], and is associated with West syndrome [34], a severe form of infancy epilepsy characterized by clusters of spasms. We propose the inclusion of NR2F1 in the diagnostic NGS gene panels for epilepsy, which could help families in their diagnostic odyssey to efficient clinical assignment.

10. NR2F1 on the Move: Motor Dysfunction in BBSOAS Patients

The development of the motor system is critical for an individual to experience the environment and engage in social interactions. In NDDs, motor skill impairments are particularly prevalent, to the extent that they are considered by clinicians as one of the first signs of atypical development [179,180,181]. BBSOAS patients have frequently been reported as presenting several motor defects, including delayed motor development, stereotyped and repetitive movements and reduced sensorimotor precision and speed, resulting in defective execution of fine visual motor behavior [33,38]. In children, the typical milestones of motor development are reached at 7 months (sitting), at 10 months (crawling) and around 13 months (walking) [182]. BBSOAS children instead achieve these milestones much later, at an average age of 14, 16 and 33 months, respectively [38]. Interestingly, the incidence of motor delay is higher in patients with DBD variants than those with mutations in other parts of the protein (53% of patients vs. 0% of children bearing all other variants) [38]. Notably, the high severity and penetrance of motor impairments in DBD-variant patients could also be the indirect consequence of other concurring clinical features, such as hypotonia and brain damage following epileptic seizures. Moreover, stereotyped and repetitive movements, such as head-banging and self-injurious behaviors, have been reported in several BBSOAS patients [1,30,38]. The etiological origin of stereotypies is still uncertain, but given the high prevalence of ASD among BBSOAS patients and its association with repetitive behaviors [136], it is possible that these behaviors represent typical ASD features, rather than separate motor symptoms. Another possibility is that the fronto-striatal circuit, responsible for the inhibition of stereotyped repetitive movements, might also be affected in BBSOAS patients, similarly to what is reported for some ASD patients [183].
Abnormal motor behavior can also be observed in the specific context of visual function and eye reflexes. Many BBSOAS patients show nystagmus (repetitive and uncontrolled involuntary movements of the eyeballs), poor object tracking, and saccadic eye movements (rapid, uncontrolled movements of the eyes that abruptly change the point of fixation) [30,38]. In a first thorough characterization of oculomotor skills of a BBSOAS case, the patient showed reduced accuracy when performing specific visual oculomotor tests, compared to age-matched control or ASD individuals [33]. Additionally, Bojanek and colleagues performed a series of tests on general manual skills; the BBSOAS patient showed impaired stopping accuracy and reduced reaction time when compared with both ASD or control age-matched individuals, consistent with a global sensorimotor impairment [33]. This might suggest weakened modulation of the cerebellum on pontine-brainstem burst cells, whereas the slower reaction times could implicate a dysfunction of either the descending cortico-ponto-cerebellar and/or the ascending cerebellar-thalamo-frontal circuits [33]. However, to define the prevalence of these motor dysfunctions among BBSOAS patients and the possible neuroanatomical correlates, a systematic assessment of motor features coupled with MRI examination on bigger cohorts is needed.
In humans, abnormalities of corticopontine and corticospinal descending tracts are usually associated with more broad brain malformations, impacting the execution of voluntary movements and hand dexterity [184,185,186,187,188]. Previous studies in mice helped in describing the development of cortical descending tracts, and noted the existence of several critical steps, each controlled by specific assets of molecular players (revised in [188]). The use of different mouse Nr2f1 mutants suggested altered wiring in the descending corticopontine and corticospinal tracts, and pointed particularly to the topographic organization of corticopontine projections, an essential process for the control of fine voluntary movements in rodents and humans [189]. Furthermore, the impairment observed in both tangential (areal) and radial (laminar) organization of the neocortical layers upon cortical specific loss of Nr2f1 function ultimately results in abnormal connectivity between the cortex and its subcerebral targets [7,10,66,190]. This defective connectivity has a deleterious effect on the accuracy of voluntary motor execution and object reaching behavior: although retaining general normal motor capabilities, fine-skilled paw movements show a remarkable impairment in Nr2f1 mutant animals [10], somehow recapitulating fine motor impairments of BBSOAS patients. Finally, it is reasonable to hypothesize that reduced dexterity could be also impacted by hyperactive features, which have been reported in patients [30,38] and described in Nr2f1 mouse mutants [191].
Further clinical reports will be necessary to fully characterize motor dysfunctions in BBSOAS patients and to distinguish abnormalities due to ASD from those primarily dependent on other motor deficits. However, independently of their clinical origin, motor dysfunctions might represent a first sign of atypical development in BBSOAS. In fact, early-onset motor impairments often emerge before social and communicative deficits [181,192,193,194]. Hence, they may serve as an early clinical indicator for BBSOAS, as happens for ASD patients [181,192,195].

11. Facial Dysmorphia

BBSOAS patients were initially reported not to have a clear facies typica associated with their syndromic conditions [1], as dysmorphic facial features could be present but seemed extremely variable and nonspecific. However, with increasing numbers of patient reports, some characteristic features started to emerge, such as ear abnormalities, including prominent ears and bilateral cupped ear helices [1,27,30,35,37], sometimes treated with cosmetic otoplasty. Other facial features include epicanthal folds, small/high nasal bridge, up-slanting palpebral fissures, tall forehead and thin upper lip [8,38]. Finally, dysmorphic facial features additionally included micrognathia and retrognathia [34,37,38], treated in one case with jaw expansion surgery [33]. Data obtained in zebrafish proved Nr2f factors to be essential for cartilage development of the lower jaw [196,197]. Importantly, NR2F1 has been reported to regulate neural crest gene expression and craniofacial morphogenesis by binding to a subset of neural crest enhancers [198], and Nr2f1-mediated BMP2 regulation is key for the differentiation and ossification of human bone marrow stromal cells [199]. These reports point to a role for NR2F1 in the regulation of neural crest development, face cartilage, and more broadly, bone ossification, which could explain not only facial dysmorphic features, but also bone age abnormalities, reported in a BBSOAS patient [35], and short stature [38]. We believe that NR2F1 functions in other tissues, still poorly characterized due to the major focus being on brain development, could have a great impact on non-neural tissue, and explain specific BBSOAS clinical features involving cartilage and bone development.

12. Feeding Problems and Mouth Stuffing, a Common but Poorly Characterized BBSOAS Feature

A common congenital feature of BBSOAS patients is oromotor dysfunction, consisting in poor swallowing, sucking and chewing, ultimately leading to feeding abnormalities and in some cases requiring gastric tube feeding. Notably, poor feeding could converge on other clinical features by contributing to the global developmental delay and short stature observed in young patients [200,201]. BBSOAS feeding deficits could be caused and/or worsened by hypotonia [202,203,204], one of the most common symptoms in BBSOAS patients. However, studies on mice open an interesting new interpretation of BBSOAS feeding problems. In fact, Nr2f1 null mouse mutants die perinatally of dehydration and starvation, due to multiple defects in cranial glossopharyngeal ganglion (IX) and nerve formation [205]. The ninth nerve supplies both sensory and motor innervation to the pharynx and root of the tongue, innervates the soft palate and induces salivary secretion essential for swallowing by innervating the parotid gland. Hence, even if BBSOAS feeding deficits may be caused indirectly by reduced muscle tone early in infancy, it is tempting to speculate that they could directly result from morphological defects of the IX cranial nerve as found in mouse models.

13. The Emerging BBSOAS Genotype–Phenotype Correlation from a Molecular Point of View: NR2F1 Dimers and the Dominant-Negative Effect

As described in earlier sections, BBSOAS patients are characterized by multiple clinical features. Such high clinical heterogeneity could depend—at least in part—on an equally high diversity in terms of NR2F1 genetic variants. While experimental models are helping to elucidate possible cellular alterations leading to some symptoms, a careful evaluation of NR2F1 function at a molecular level is also necessary to uncover potential genotype—phenotype correlations and obtain a deeper comprehension of BBSOAS features.
The first major genetic difference among BBSOAS patients consists of whether the function of one NR2F1 allele is lost by whole-gene deletion or by point mutations. On one side, BBSOAS patients with chromosome deletions physically lose one copy of the NR2F1 gene, resulting in reduced protein dosage and functional haploinsufficiency. Conversely, patients carrying deleterious (loss-of-function) missense mutations in one allele still express mutated NR2F1 forms. Missense point variants are mainly reported in the coding sequence of two functional domains highly conserved across members of the nuclear receptor family: the DNA-binding domain (DBD) consisting of two zinc-finger domains, and the ligand-binding domain (LBD) (Figure 1). The structure of the NR2F1 LBD and the identity of candidate chemical agonists fitting the ligand-binding pocket start to be characterized [206,207]; however, as the physiological ligands remain unknown, NR2F1 is still ascribed as an orphan nuclear receptor. The LDB is also responsible for protein dimerization, as NR2F monomers combine among them or with other members of the same family (forming homo- or heterodimers, respectively) to regulate gene expression [3,66]. Similarly to other nuclear receptors, NR2Fs also display two activating function domains (AF-1 and AF-2), necessary for binding co-factors that can be either activated by conformational changes induced by ligand binding [208,209,210], or in some cases, be constitutively active [211].
Since NR2F1 binds the DNA in the form of homodimers or heterodimers [4,212], point-mutation variants could result in complex dominant-negative effects, in which the mutated form competes with the wild-type protein (or with other nuclear receptors) for dimer formation. Indeed, a dominant-negative effect of DBD mutations was proven by in vitro luciferase assays [30,155]. Such a molecular mechanism would explain why BBSOAS patients with point mutations falling in the DBD often have a high prevalence and overall increased clinical severity, such as greater severity of speech and motor disabilities, compared to patients with other variants [38]. Additionally, epileptic features are more often reported in DBD-mutated patients (74% vs. 37% in patients with other variants, as previously reported) [38]. However, due to the low number of BBSOAS patients described to date, little is yet known about the emerging genotype–phenotype correlation and the molecular mechanisms underlying such correlations.
The 92 patients reported by different studies (Table 1) show a diverse array of deletions or point variants, completely ablating one NR2F1 allele or affecting its sequence, respectively. We provide here a classification of the NR2F1 perturbations observed so far and speculate on possible genetic/molecular mechanisms that could impact dimer formation and NR2F1 functioning (Figure 4), possibly explaining the molecular origins of the genotype–phenotype correlation. For this purpose, we correlated each patient group to a severity index based on the presence or absence of the main BBSOAS clinical symptoms (abnormal MRI, DD, ID, visual system defects, epilepsy, ASD or autistic-like traits and hypotonia) and ranging from 1 (one symptom only) to 7 (when all symptoms are present). The severity index and the prevalence of each clinical category in different genetic groups are reported in Table 5.
BBSOAS genetic categories and NR2F1 function:
  • Whole-gene deletions or small indels. (Severity index: 4.33). A few BBSOAS patients so far have been reported with chromosome deletions spanning from 582 Kb to 5 Mb in size, all resulting in whole-gene ablation and complete loss of one NR2F1 allele [1,27,28,29,30,38]. The likely consequence of such deletion is a halved NR2F1 protein production, i.e., haploinsufficiency, as was proven by protein quantification in skin fibroblast extracts [30]. However, it must be noted that large deletions and complex chromosomal aberrations can involve additional genes located next to the NR2F1 locus, such as FAM172A, POU5F2, MIR2277 and lnc-NR2F1, adding to the complexity of the condition, and possibly leading to supplementary congenital abnormalities, such as periventricular heterotopia and deafness, among others [27,28,29]. Interestingly, specific deletions only affecting the region adjacent to the NR2F1 gene, where the long non-coding (lnc)-NR2F1 RNA is located, were also shown to cause neurodevelopmental conditions with developmental and speech delay [213], probably due to the ability of lnc-NR2F1 to control the expression of autism-associated neural genes in a similar way to its corresponding protein-coding gene. These data suggest that (i) multiple genes could synergistically converge to cause similar neurological conditions, and that (ii) the loss of other genes together with NR2F1 in BBSOAS patients with large deletions could exacerbate the clinical features and further increase their heterogeneity.
  • Translation initiation variants (TIVs). (Severity index: 5.33). BBSOAS patients with missense variants falling in the translation initiation codon (ATG) show decreased NR2F1 protein resulting from reduced efficiency of both translational and transcriptional processes [30]. Notably, the third codon of the NR2F1 gene sequence is also an ATG, raising the possibility that an alternative initiation site could be present and potentially compensate for the loss or mutation of the main one. However, diminished NR2F1 levels measured in cells from these patients suggest that the second ATG site is not able to efficiently serve as an alternative start codon [30]. The production of a half dosage of the functional NR2F1 protein, without any mutated form competing for dimer formation, could make these patients more similar to BBSOAS individuals with whole-gene deletions than to patients carrying missense DBD/LBD variants, hence leading to milder phenotypes. However, this has been recently questioned by the report of a severe BBSOAS case owing to an NR2F1 start codon variant [85]. While the severity index shows an intermediate severity compared to the two groups, statistical analysis on larger cohorts will be necessary to better detail the phenotypes associated with TIVs.
  • DBD missense variants (or DBD in-frame deletions). (Severity index: 5.62). The effect of missense mutations is tightly linked to the structural and functional importance of the affected amino acids and the protein region in which they are located. Based upon evolutionary conservation of distinct aminoacidic positions, bioinformatic prediction software can be used to identify highly conserved regions, and to evaluate whether they are intolerant towards variation and therefore potentially pathogenic [214,215,216]. BBSOAS pathogenic point mutations falling in the DBD and impacting the 3D structure of evolutionarily conserved sites in the zinc-finger domains have strong effects on NR2F1 structure and function, and consequently, on the clinical phenotype. As an example, substitution of a highly conserved zinc-finger motif of the DBD of NR2F1 leads to heavy changes in molecular structure and stability, as predicted in silico [36]. By contrast, missense variants located in less conserved regions adjacent to the DBD showed only a reduced transcriptional activity in the luciferase assay, in contrast to the almost abolished activity of DBD variants falling in crucial zinc-finger or structural sites [30]. As discussed above, the high penetrance of BBSOAS features following DBD missense mutations, compared to the complete loss of one allele by whole-gene deletion, could result from a possible dominant-negative effect of mutated NR2F1. Indeed, by forming non-functional dimers with the wild-type NR2F1 (produced by the normal allele), mutated NR2F1 proteins could ultimately affect 75% of the total pool of NR2F1 dimers in the cell (Figure 4). Heterodimer formation with other nuclear receptors (NRs) could be affected too, presumably resulting in a 50% loss of functional NR2F1-NRs heterodimers (Figure 4). The high number of reported DBD missense mutations, associated with the highest severity index among BBSOAS genetic groups, confirms the functional relevance of this region.
  • LBD missense variants. (Severity index: 3.76). Patients with variants in the LBD manifest milder developmental delay, often lacking hypotonia, speech defects, seizures and repetitive behaviors [35], consistent with the lowest severity index among BBSOAS genetic categories. As the LBD is necessary for NR2F1 dimerization, it is reasonable to think that mutations falling on interacting surfaces could hamper dimer formation. While this could decrease NR2F1 activity in general, such a situation could also affect the formation of dimers between normal and mutated NR2F1, hence resembling whole-gene deletions, with no (or lower) dominant-negative effect. The LBD also contains the Activating Function domain 2 (AF-2), which is necessary for co-factor binding. While the exact nature of NR2F1 protein partners in neural cells remains elusive, the loss of such interactions following LBD mutations could be responsible for some BBSOAS features. Further studies will be necessary to dissect the specific function of distinct domains in the LBD and the impact of distinct mutations on dimer formation, as well as the identity of NR2F1 physiological co-factors. This will lead to further insights into NR2F1 molecular functions and how they correlate with patients’ symptoms.
  • Truncation or frameshift followed by truncation variants. (Severity index: 5.18 and 5.29, respectively). Nonsense mutations result in premature termination of the protein (i.e., truncation of the peptide sequence) and can have profound impacts on gene function [217]. Obviously, the portion of the protein that is lost depends on the precise location of the aberrant termination site. The low number of described patients carrying NR2F1 truncations makes it challenging to gain a clear picture of the genotype–phenotype correlation for those pathological cases. However, it is reasonable to think that early truncations would trigger nonsense-mediated decay of the mRNA, resulting in loss-of-function, whereas late truncation variants that escape nonsense-mediated decay may create an abnormal, but stable, truncated protein, causing dominant-negative or neomorphic effects [218]. Hence, early variants leading to haploinsufficiency would be more similar to whole-gene deletions and may also protect against more detrimental phenotypic effects (i.e., dominant-negative effect), presumably leading to milder phenotypes [218,219]. Consistently, a recently reported patient presenting NR2F1 early truncation showed a mild phenotype [40], whereby one case has high-functioning ASD with superior verbal abilities [33]. By contrast, late variants that do not affect the dimerization region of the LBD could show a dominant-negative effect, similarly to what is reported for point variants, and ultimately leading to highly penetrant symptoms. Taking all truncation variants together, their collective severity index is more severe than that of deletions, similarly to variants affecting the starting codon.
A further complication is represented by frameshift variants, which can determine large-scale deleterious changes to the overall polypeptide length and chemical composition. Moreover, they are often followed by truncation at a termination site located at variable distance. By dramatically altering the protein sequence, truncating variants with frameshift can result in heavy conformational changes and can have much stronger effects on disease risk [220]. The acquisition of toxic properties can constitute a deleterious event following frameshift [218], further impacting cellular physiology and biochemical processes, thus being more deleterious than the loss of the gene itself. Consistently, a patient with frameshift variant and truncation in the NR2F1 LBD has been reported to show a severe phenotype, including epilepsy [39]. Their severity index shows that these variants could be slightly more severe compared to simple truncations. In vitro assays and in silico predictions of 3D protein structure, together with accurate description of a higher number of patients, could help unravel the exact effect of NR2F1 truncations falling in different protein regions and the possible pathogenic impact due to changes in the aminoacidic sequence following frameshift.
In summary, the severity of BBSOAS symptoms could be the result of both NR2F1 haploinsufficiency and dominant-negative effects affecting protein availability on one side and dimer formation on the other. Consistent with previously reported data [30,38], the severity index presented here points to a scenario where a dominant-negative effect of DBD genetic variants results in the most severe clinical outcomes. Indeed, DBD variants lead to high percentages of main BBSOAS symptoms (Table 5), such as DD (90.6% of patients), ID and/or speech delay (93.8%), CVI (53.1%), OA (78.1%), seizures (62.5%) and hypotonia (75%). On the contrary, the milder phenotypes would be associated with LBD pathogenic variants, showing a decreased prevalence of distinct symptoms when compared to other genetic groups, notably concerning DD (70.6%), ID (70.6%), OA (47.1%), hypotonia (35.3%) and epilepsy (29.4%). It is interesting to note that deletion variants, with respective prevalence of DD (93.3%), hypotonia (60%) and optic disc malformations (33.3%), manifest a higher severity index than LBD point mutations (4.33 vs. 3.82). Similarly, translation initiation variants (TIVs), previously compared to gene deletion in terms of symptom severity, display a quite high severity index of 5.22, due to high prevalence of DD, ID, optic atrophy and very high incidence of hypotonia (88.9%). We expect that the continuous report of novel patients will gradually refine the calculations of clinical prevalence, allowing a better understanding of symptom severity in the distinct genetic groups. It is intriguing to notice that truncation and frameshift/truncation variants show a particularly high prevalence of CC malformations (63.6% and 42.9%, respectively), OA (72.7% and 71.4%, respectively) and DD (90.9% and 100%, respectively). Furthermore, frameshift/truncation variants have a high prevalence of ONH (57.1%) and high risk of epilepsy (57.1%), ASD (71.4%) and hypotonia (57.1%). Altogether, possible pathogenic effects due to frameshift variants cannot be excluded at this point. However, the number of reported patients carrying these types of variants is very limited and could affect the calculation of symptom prevalence.
Overall, the high clinical heterogeneity of BBSOAS patients could be a direct consequence of a yet not well characterized genotype–phenotype correlation, in which specific NR2F1 variants trigger respective clinical outcomes. Careful and consistent clinical investigation of newly reported BBSOAS patients would be a key factor to improve the comprehension of BBSOAS clinical features and genotype–phenotype correlation. In parallel, a more comprehensive characterization of NR2F1 molecular function in both physiological and pathological conditions will help predicting the full spectrum of effects of distinct genetic conditions at a cellular and molecular level and will allow the inference of possible consequences in terms of BBSOAS symptoms.

14. Discussion

The continuous reporting of newly identified BBSOAS patients, made possible by the effort of their families and clinicians, is rapidly expanding our knowledge of this rare autosomal dominant syndrome. However, the variability of clinical descriptions, depending on the scope and depth of the respective study, make it challenging to compare clinical BBSOAS reports from different sources. Before the discovery and naming of BBSOAS as a unique syndrome, individuals with NR2F1 mutations were included in autistic, epileptic or other clinical groups, depending on their most prevalent features, leading to poor description of other, apparently unrelated, symptoms. Studies on larger cohorts of patients with NR2F1 pathogenic variants would further define BBSOAS as a syndromic condition and would lead to a more complete characterization of clinical features [30,38]. Despite this, it remains challenging to compare clinical descriptions from different studies. As an example, ophthalmological features often include OCT scans of the retina and eye fundoscopy, but fewer visual acuity measurements, and an ophthalmological analysis was performed on a large group of patients only recently [41]. Furthermore, MRI analysis of some BBSOAS patients reported morphological malformations of the neocortex [8], but such analysis is still lacking in the vast majority of patients. To help future investigations, it would be necessary to compile a standardized list of clinical exams to regularly perform on newly diagnosed BBSOAS patients. Focus should be on both the clinical challenges BBSOAS-diagnosed children could face (see Table 2) and the clinical evaluations to be implemented (see Table 3).
In light of the rarity of this syndrome, the best approach to date would be to use uniform criteria and to perform in-depth clinical research studies on a higher number of patients, taking advantage of family conferences bringing BBSOAS patients together, as recently undertaken during a meeting organized by the USA parent association [38]. The increase in the number of diagnosed patients would allow exploration of the impact of distinct NR2F1 genetic perturbations at a cellular and molecular level, and of the ways this contributes to their pathophysiological heterogeneity. Furthermore, a unified registry for BBSOAS patients would facilitate the clinical research on both already reported and newly identified patients. The USA NR2F1 Foundation is moving in this direction (BBSOAS Patient Registry, under preparation via Across Healthcare Matrix Platform). Further help could also come from public databases reporting published NR2F1 variants, using standard nomenclature to describe both molecular and phenotypic anomalies. Tools such as the LOVD database will help to further refine the BBSOAS clinical synopsis and allow for a better comparison of patient symptoms [32].
A further challenge in the analysis of BBSOAS patients comes from the complex syndromic nature of their condition. NDDs often show comorbidity of several clinical features, such as ASD, developmental delay, ID and epilepsy [166,221], and BBSOAS makes no exception. Such recurrent comorbidity sustains the idea that the impairment of common mechanisms orchestrating brain development and function could converge on similar pathophysiological phenotypes via shared cellular and molecular pathways [222]. In such a scenario, the crosstalk between different disorders makes it extremely difficult to understand the temporal order of appearance of distinct clinical features and the causative links among them. In the specific case of BBSOAS, for example, ID could result from multiple factors: visual impairments, seizures, cortical or hippocampal malformations, among others. Visual impairments, on the other side, could be further impacted by mitochondrial dysfunction, reported in two patients [42,43], as ON axonal damage easily results from energetic failure secondary to impairments of the mitochondrial respiratory chain [45]. Furthermore, we cannot exclude that both ID and CVI result from secondary damage caused by severe epileptic infantile spasms, as this catastrophic infant epilepsy has been shown to heavily damage visual system integrity and brain development in general [161,162,223,224,225,226].
An example of complex comorbidity in a BBSOAS-like context comes from an Nr2f1-haploinsufficient mouse model recapitulating visual syndromic features, which showed concomitant RGC decreased survival, oligodendrocyte hypomyelination and astrocyte inflammation in the ON of developing mutants [16,41]. In such a complex scenario, understanding the specific contribution of each distinct player in disrupting the ON physiology is most probably unfeasible. A possible solution would be to employ conditional mouse models, in which the use of cell- and time-specific Cre-recombinases allows the removal of Nr2f1 in a very targeted way, thus allowing dissection of its role in specific contexts [5]. As an example, conditional Nr2f1 knock-out in the oligodendrocyte lineage could help in discerning the impact of oligodendrocyte dysfunction and/or astrocyte inflammation in the ONH patient phenotype. In summary, while Nr2f1 constitutive mouse models better recapitulate the BBSOAS condition in toto, Nr2f1 conditional ones help dissect the causative and temporal link between distinct pathological features (reviewed in [5]).
The presence of comorbidities and the phenotypic clinical overlap with other complex syndromic conditions has led to cases of misdiagnosis in BBSOAS patients [44]. This can be avoided by applying an unbiased sequencing approach, such as whole-genome sequencing (WES), to patients with undiagnosed NDDs. WES successfully identified BBSOAS patients among individuals with previously unsolved/unspecific forms of ID [227,228] or epilepsy [229], and the comparison of NGS data helped in compiling the annotation of numerous additional NR2F1 variants in online repositories [230], comprising VUS (variants of unknown significance), likely benign variants and likely pathogenic variants. In this sense, WES or other whole-genome-sequencing approaches should be used as a first choice in the clinical armamentarium, rather than as a last-resource genetic test [231,232,233]. This approach would not only favor an early and accurate molecular diagnosis, but also constitute the foundation of rational and effective therapeutic choices, leading to significant changes in the clinical management of previously unsolved clinical cases [234,235]. However, some NDDs can also be caused by de novo somatic mutations, which occur post-zygotically and are thus present in only a subset of the cells/organs of an affected individual [236]. Therefore, additional BBSOAS patients with somatic mutations in brain tissue only could remain unreported after WES sequencing on blood samples.
Another open question is whether BBSOAS syndromic traits are completely dependent on genetic factors (i.e., NR2F1 perturbations) or can also be influenced by the environment, a plausible mechanism for NDDs. As an example, interactions between genetic and environmental factors have been proposed to explain the complex physiopathology of ASD [237,238,239,240]. The concomitant presence of environmental and genetic influences could further explain the high variability in terms of presence/absence and severity degree of BBSOAS features. A recent report described highly similar, almost overlapping symptoms in two BBSOAS monozygotic twins [36], suggesting a strong genetic component for BBSOAS. However, it would be interesting in the future to compare the clinical descriptions of patients with different genetic backgrounds (i.e., non-consanguineous) and from different familial environments but bearing identical missense mutations. This would allow the gathering of new insights into a putative interplay between genetic and environmental components. However, due to discrepancies in patient descriptions among distinct cohorts discussed above, this will be a challenging task.
Regarding cellular and molecular levels, most of the data coming from classical studies on NR2F1/Nr2f1 molecular function and dimer formation have been obtained in non-physiological conditions or in non-neural cells. Moreover, an in silico prediction of the NR2F1 3D protein structure remains partially elusive to date, with the DBD being the only domain to have been resolved as yet (PDB entry: 2EBL). To overcome the lack of NR2F1 LBD structure, and thanks to the high similarity with its homolog NR2F2 [241], the 3D structure of the latter is currently used as an alternative (PDB entry: 3CJW); however, to model the LBD in its active conformational state, the RXR LBD structure is preferred (PDB entry: 1FM6) [206] even though it only shows a 40% homology to the actual NR2F1 sequence. Further, the exact 3D structure of the linking hinge between LBD and DBD is still missing, as it is too variable to be easily predicted. Further studies unraveling, on one side, the 3D structure of NR2F1, and on the other side, its function in a more reliable and physiological context (i.e., living human neural cells), are of extreme importance, as they would allow the prediction of changes in protein–protein and protein–DNA interactions following specific NR2F1 pathogenic mutations.
Finally, a tight link between clinical reports and experimental research is fundamental for advancing our understanding of BBSOAS and for planning future treatments. One of the key questions to be addressed through research is to what degree BBSOAS is due to altered neurodevelopment vs. altered neuronal homeostasis in the postnatal individual. The greater the contribution of ongoing processes, the greater the potential for therapeutic intervention. As more BBSOAS patients are being reported, and as clinical features are better described, the time will come to test possible therapeutic approaches aiming at improving some patients’ symptoms. As an example, studies in mice showed the feasibility of chemical drug treatments to restore correct myelination levels in the context of several acute or syndromic pathologies, as well as in the ONs of a BBSOAS-like model [16,68,69,70,71,72,75,242,243]. It could be interesting to screen for efficient remyelinating drugs on patient-derived 3D brain organoids containing oligodendrocytes or 2D-differentiated oligodendrocyte precursors [244,245,246,247,248], to identify potential treatments for BBSOAS myelination defects, and more generally, to employ an in vitro BBSOAS-like set-up as a tool for the screening of novel myelinating chemical compounds. Finally, for individuals with point mutations in the DBD, one could envision a therapeutic approach with antisense oligonucleotides targeting the mutated version of NR2F1. Although not being expected to cure BBSOAS, it might lead to an amelioration of the phenotype, shifting it into the milder range, such as what is seen among individuals with heterozygous deletions.
In conclusion, the great breadth and variability of BBSOAS symptoms, due to distinct forms of NR2F1 deficiency, could result from the disruption of converging pathways regulating brain development. The clinical description of newly reported BBSOAS patients, in combination with experimental research in vitro and in vivo, will ultimately allow for a better comprehension of this recently described syndrome and the roles played by NR2F1 during early development.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11081260/s1, Table S1: Updated list of NR2F1 variants and clinical description of BBSOAS reported patients, extended version.

Author Contributions

Conceptualization, M.B.; resources, M.B. and M.S.; data curation, M.B. and C.T.; writing—original draft preparation, M.B., C.T. and M.S.; writing—review and editing, M.B., C.T., C.P.S. and M.S.; visualization, M.B. and C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the “Fondation Recherche Médicale; Équipe FRM 2020” (#EQU202003010222) and by the Fondation de France (#00123416/WB-2021-38751) to MS; by a Ph.D. contract from Région PACA/Inserm and FRM 4th year Ph.D. to CT.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ABA, applied behavioral analysis; ADHD, attention deficit hyperactivity disorder; ADI-R, Autism Diagnostic Interview, Revised; ADOS, Autism Diagnostic Observation Schedule; AF-1 and AF-2, activating function domains 1 and 2, respectively; AIS; axon initial segment; ASD, autism spectrum disorder; BBSOAS, Bosch–Boonstra–Schaaf optic atrophy syndrome; CB, cerebellum; CC, corpus callosum; CS, corticospinal tract; CVI, cortical visual impairment; D, deletion; DBD, DNA binding domain; DD, developmental delay; dLGN, dorso-lateral geniculate nucleus; DM, delayed myelination; DMD, delayed motor development/poor coordination; DQ, developmental quotient; EEG, electroencephalogram; EOE/S, early-onset epilepsy/seizures; FI/NFI, frameshifting indel/non frameshifting indel; FS, febrile seizures; GCL, ganglionic cell layer; GVI, general visual impairments; HET, heterozygous; HP, hippocampus; HPM, hippocampal malrotation; ID, intellectual disability; IQ, intelligence quotient; IS, infantile spasms; KO, knock-out; LBD, ligand-binding domain; LV, lateral ventricle; LVA, low visual acuity; MCDs, malformations of cortical development; MCP, macrocephaly; MD, microdeletion; MM, missense mutation; MRI, magnetic resonance imaging; NDDs, neurodevelopmental disorders; NR, neural retina; OA, optic atrophy; OC, optic chiasm; OCB/RB, obsessive-compulsive/repetitive behaviors; OCD, obsessive compulsive disorder; OCT, optical coherence tomography; OD, optic disc; ON, optic nerve; ONH, optic nerve hypoplasia; P/SOD, pale/small optic disc; PDD-NOS, pervasive developmental disorder-not otherwise specified; PH, bilateral periventricular heterotopia; RB, repetitive behavior; RGCs, retinal ganglion cells; RNFL, retinal nerve fiber layer; TIV, translation initiation variants; V1, primary visual area; VUS, variants of unknown significance; WES, whole-Genome Sequencing; WGD, whole-gene deletion; WM, white matter; WT, wild-type.

References

  1. Bosch, D.G.M.; Boonstra, F.N.; Gonzaga-Jauregui, C.; Xu, M.; de Ligt, J.; Jhangiani, S.; Wiszniewski, W.; Muzny, D.M.; Yntema, H.G.; Pfundt, R.; et al. NR2F1 Mutations Cause Optic Atrophy with Intellectual Disability. Am. J. Hum. Genet. 2014, 94, 303–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Zhou, C. COUP-TFI: An Intrinsic Factor for Early Regionalization of the Neocortex. Genes Dev. 2001, 15, 2054–2059. [Google Scholar] [CrossRef] [Green Version]
  3. Bertacchi, M.; Parisot, J.; Studer, M. The Pleiotropic Transcriptional Regulator COUP-TFI Plays Multiple Roles in Neural Development and Disease. Brain Res. 2018, 1705, 75–94. [Google Scholar] [CrossRef]
  4. Tang, K.; Tsai, S.Y.; Tsai, M.-J. COUP-TFs and Eye Development. Biochim. Biophys. Acta 2015, 1849, 201–209. [Google Scholar] [CrossRef] [Green Version]
  5. Tocco, C.; Bertacchi, M.; Studer, M. Structural and Functional Aspects of the Neurodevelopmental Gene NR2F1: From Animal Models to Human Pathology. Front. Mol. Neurosci. 2021, 14, 767965. [Google Scholar] [CrossRef]
  6. Alfano, C.; Magrinelli, E.; Harb, K.; Hevner, R.F.; Studer, M. Postmitotic Control of Sensory Area Specification during Neocortical Development. Nat. Commun. 2014, 5, 5632. [Google Scholar] [CrossRef] [Green Version]
  7. Armentano, M.; Chou, S.-J.; Srubek Tomassy, G.; Leingärtner, A.; O’Leary, D.D.M.; Studer, M. COUP-TFI Regulates the Balance of Cortical Patterning between Frontal/Motor and Sensory Areas. Nat. Neurosci. 2007, 10, 1277–1286. [Google Scholar] [CrossRef] [PubMed]
  8. Bertacchi, M.; Romano, A.L.; Loubat, A.; Tran Mau-Them, F.; Willems, M.; Faivre, L.; Khau van Kien, P.; Perrin, L.; Devillard, F.; Sorlin, A.; et al. NR2F1 Regulates Regional Progenitor Dynamics in the Mouse Neocortex and Cortical Gyrification in BBSOAS Patients. EMBO J. 2020, 39, e104163. [Google Scholar] [CrossRef]
  9. Faedo, A.; Tomassy, G.S.; Ruan, Y.; Teichmann, H.; Krauss, S.; Pleasure, S.J.; Tsai, S.Y.; Tsai, M.-J.; Studer, M.; Rubenstein, J.L.R. COUP-TFI Coordinates Cortical Patterning, Neurogenesis, and Laminar Fate and Modulates MAPK/ERK, AKT, and ß-Catenin Signaling. Cereb. Cortex 2008, 18, 2117–2131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Tomassy, G.S.; De Leonibus, E.; Jabaudon, D.; Lodato, S.; Alfano, C.; Mele, A.; Macklis, J.D.; Studer, M. Area-Specific Temporal Control of Corticospinal Motor Neuron Differentiation by COUP-TFI. Proc. Natl. Acad. Sci. USA 2010, 107, 3576–3581. [Google Scholar] [CrossRef] [Green Version]
  11. Chen, C.-A.; Wang, W.; Pedersen, S.E.; Raman, A.; Seymour, M.L.; Ruiz, F.R.; Xia, A.; van der Heijden, M.E.; Wang, L.; Yin, J.; et al. Nr2f1 Heterozygous Knockout Mice Recapitulate Neurological Phenotypes of Bosch-Boonstra-Schaaf Optic Atrophy Syndrome and Show Impaired Hippocampal Synaptic Plasticity. Hum. Mol. Genet. 2020, 29, 705–715. [Google Scholar] [CrossRef]
  12. Flore, G.; Di Ruberto, G.; Parisot, J.; Sannino, S.; Russo, F.; Illingworth, E.A.; Studer, M.; De Leonibus, E. Gradient COUP-TFI Expression Is Required for Functional Organization of the Hippocampal Septo-Temporal Longitudinal Axis. Cereb. Cortex 2016, 27, bhv336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Parisot, J.; Flore, G.; Bertacchi, M.; Studer, M. COUP-TFI Mitotically Regulates Production and Migration of Dentate Granule Cells and Modulates Hippocampal Cxcr4 Expression. Dev. Camb. Engl. 2017, 144, 2045–2058. [Google Scholar] [CrossRef] [Green Version]
  14. Lodato, S.; Tomassy, G.S.; De Leonibus, E.; Uzcategui, Y.G.; Andolfi, G.; Armentano, M.; Touzot, A.; Gaztelu, J.M.; Arlotta, P.; Menendez de la Prida, L.; et al. Loss of COUP-TFI Alters the Balance between Caudal Ganglionic Eminence- and Medial Ganglionic Eminence-Derived Cortical Interneurons and Results in Resistance to Epilepsy. J. Neurosci. 2011, 31, 4650–4662. [Google Scholar] [CrossRef]
  15. Touzot, A.; Ruiz-Reig, N.; Vitalis, T.; Studer, M. Molecular Control of Two Novel Migratory Paths for CGE-Derived Interneurons in the Developing Mouse Brain. Development 2016, 143, 1753–1765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bertacchi, M.; Gruart, A.; Kaimakis, P.; Allet, C.; Serra, L.; Giacobini, P.; Delgado-García, J.M.; Bovolenta, P.; Studer, M. Mouse Nr2f1 Haploinsufficiency Unveils New Pathological Mechanisms of a Human Optic Atrophy Syndrome. EMBO Mol. Med. 2019, 11, e10291. [Google Scholar] [CrossRef] [PubMed]
  17. Tang, K.; Xie, X.; Park, J.-I.; Jamrich, M.; Tsai, S.; Tsai, M.-J. COUP-TFs Regulate Eye Development by Controlling Factors Essential for Optic Vesicle Morphogenesis. Development 2010, 137, 725–734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Naka, H.; Nakamura, S.; Shimazaki, T.; Okano, H. Requirement for COUP-TFI and II in the Temporal Specification of Neural Stem Cells in CNS Development. Nat. Neurosci. 2008, 11, 1014–1023. [Google Scholar] [CrossRef]
  19. Okano, H.; Temple, S. Cell Types to Order: Temporal Specification of CNS Stem Cells. Curr. Opin. Neurobiol. 2009, 19, 112–119. [Google Scholar] [CrossRef]
  20. del Pino, I.; Tocco, C.; Magrinelli, E.; Marcantoni, A.; Ferraguto, C.; Tomagra, G.; Bertacchi, M.; Alfano, C.; Leinekugel, X.; Frick, A.; et al. COUP-TFI/Nr2f1 Orchestrates Intrinsic Neuronal Activity during Cortical Area Patterning. bioRxiv 2019. [CrossRef]
  21. Alzu’bi, A.; Lindsay, S.; Kerwin, J.; Looi, S.J.; Khalil, F.; Clowry, G.J. Distinct Cortical and Sub-Cortical Neurogenic Domains for GABAergic Interneuron Precursor Transcription Factors NKX2.1, OLIG2 and COUP-TFII in Early Fetal Human Telencephalon. Brain Struct. Funct. 2017, 222, 2309–2328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Alzu’bi, A.; Lindsay, S.J.; Harkin, L.F.; McIntyre, J.; Lisgo, S.N.; Clowry, G.J. The Transcription Factors COUP-TFI and COUP-TFII Have Distinct Roles in Arealisation and GABAergic Interneuron Specification in the Early Human Fetal Telencephalon. Cereb. Cortex 2017, 27, 4971–4987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Foglio, B.; Rossini, L.; Garbelli, R.; Regondi, M.C.; Mercurio, S.; Bertacchi, M.; Avagliano, L.; Bulfamante, G.; Coras, R.; Maiorana, A.; et al. Dynamic Expression of NR2F1 and SOX2 in Developing and Adult Human Cortex: Comparison with Cortical Malformations. Brain Struct. Funct. 2021, 226, 1303–1322. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, X.; Feng, S.; Tang, K. COUP—TF Genes, Human Diseases, and the Development of the Central Nervous System in Murine Models. In Current Topics in Developmental Biology; Elsevier: Amsterdam, The Netherlands, 2017; Volume 125, pp. 275–301. ISBN 978-0-12-802172-9. [Google Scholar]
  25. Michaud, J.L.; Lachance, M.; Hamdan, F.F.; Carmant, L.; Lortie, A.; Diadori, P.; Major, P.; Meijer, I.A.; Lemyre, E.; Cossette, P.; et al. The Genetic Landscape of Infantile Spasms. Hum. Mol. Genet. 2014, 23, 4846–4858. [Google Scholar] [CrossRef] [Green Version]
  26. Sanders, S.J.; Murtha, M.T.; Gupta, A.R.; Murdoch, J.D.; Raubeson, M.J.; Willsey, A.J.; Ercan-Sencicek, A.G.; DiLullo, N.M.; Parikshak, N.N.; Stein, J.L.; et al. De Novo Mutations Revealed by Whole-Exome Sequencing Are Strongly Associated with Autism. Nature 2012, 485, 237–241. [Google Scholar] [CrossRef]
  27. Al-Kateb, H.; Shimony, J.S.; Vineyard, M.; Manwaring, L.; Kulkarni, S.; Shinawi, M. NR2F1 Haploinsufficiency Is Associated with Optic Atrophy, Dysmorphism and Global Developmental Delay. Am. J. Med. Genet. A 2013, 161A, 377–381. [Google Scholar] [CrossRef]
  28. Brown, K.K.; Alkuraya, F.S.; Matos, M.; Robertson, R.L.; Kimonis, V.E.; Morton, C.C. NR2F1 Deletion in a Patient with a de Novo Paracentric Inversion, Inv(5)(Q15q33.2), and Syndromic Deafness. Am. J. Med. Genet. A 2009, 149A, 931–938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Cardoso, C.; Boys, A.; Parrini, E.; Mignon-Ravix, C.; McMahon, J.M.; Khantane, S.; Bertini, E.; Pallesi, E.; Missirian, C.; Zuffardi, O.; et al. Periventricular Heterotopia, Mental Retardation, and Epilepsy Associated with 5q14.3-Q15 Deletion. Neurology 2009, 72, 784–792. [Google Scholar] [CrossRef]
  30. Chen, C.-A.; Bosch, D.G.M.; Cho, M.T.; Rosenfeld, J.A.; Shinawi, M.; Lewis, R.A.; Mann, J.; Jayakar, P.; Payne, K.; Walsh, L.; et al. The Expanding Clinical Phenotype of Bosch-Boonstra-Schaaf Optic Atrophy Syndrome: 20 New Cases and Possible Genotype-Phenotype Correlations. Genet. Med. Off. J. Am. Coll. Med. Genet. 2016, 18, 1143–1150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Abrahams, B.S.; Arking, D.E.; Campbell, D.B.; Mefford, H.C.; Morrow, E.M.; Weiss, L.A.; Menashe, I.; Wadkins, T.; Banerjee-Basu, S.; Packer, A. SFARI Gene 2.0: A Community-Driven Knowledgebase for the Autism Spectrum Disorders (ASDs). Mol. Autism 2013, 4, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Billiet, B.; Amati-Bonneau, P.; Desquiret-Dumas, V.; Guehlouz, K.; Milea, D.; Gohier, P.; Lenaers, G.; Mirebeau-Prunier, D.; den Dunnen, J.T.; Reynier, P.; et al. NR2F1 Database: 112 Variants and 84 Patients Support Refining the Clinical Synopsis of Bosch–Boonstra–Schaaf Optic Atrophy Syndrome. Hum. Mutat. 2021, 43, 128–142. [Google Scholar] [CrossRef] [PubMed]
  33. Bojanek, E.K.; Mosconi, M.W.; Guter, S.; Betancur, C.; Macmillan, C.; Cook, E.H. Clinical and Neurocognitive Issues Associated with Bosch-Boonstra-Schaaf Optic Atrophy Syndrome: A Case Study. Am. J. Med. Genet. Part A 2019, 182, 213–218. [Google Scholar] [CrossRef] [PubMed]
  34. Hino-Fukuyo, N.; Kikuchi, A.; Yokoyama, H.; Iinuma, K.; Hirose, M.; Haginoya, K.; Niihori, T.; Nakayama, K.; Aoki, Y.; Kure, S. Long-Term Outcome of a 26-Year-Old Woman with West Syndrome and an Nuclear Receptor Subfamily 2 Group F Member 1 Gene (NR2F1) Mutation. Seizure 2017, 50, 144–146. [Google Scholar] [CrossRef] [PubMed]
  35. Kaiwar, C.; Zimmermann, M.T.; Ferber, M.J.; Niu, Z.; Urrutia, R.A.; Klee, E.W.; Babovic-Vuksanovic, D. Novel NR2F1 Variants Likely Disrupt DNA Binding: Molecular Modeling in Two Cases, Review of Published Cases, Genotype-Phenotype Correlation, and Phenotypic Expansion of the Bosch-Boonstra-Schaaf Optic Atrophy Syndrome. Cold Spring Harb. Mol. Case Stud. 2017, 3, a002162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Mio, C.; Fogolari, F.; Pezzoli, L.; D’Elia, A.V.; Iascone, M.; Damante, G. Missense NR2F1 Variant in Monozygotic Twins Affected with the Bosch–Boonstra–Schaaf Optic Atrophy Syndrome. Mol. Genet. Genom. Med. 2020, 8, e1278. [Google Scholar] [CrossRef] [PubMed]
  37. Park, S.E.; Lee, J.S.; Lee, S.-T.; Kim, H.Y.; Han, S.-H.; Han, J. Targeted Panel Sequencing Identifies a Novel NR2F1 Mutations in a Patient with Bosch–Boonstra–Schaaf Optic Atrophy Syndrome. Ophthalmic Genet. 2019, 40, 359–361. [Google Scholar] [CrossRef] [PubMed]
  38. Rech, M.E.; McCarthy, J.M.; Chen, C.-A.; Edmond, J.C.; Shah, V.S.; Bosch, D.G.M.; Berry, G.T.; Williams, L.; Madan-Khetarpal, S.; Niyazov, D.; et al. Phenotypic Expansion of Bosch-Boonstra-Schaaf Optic Atrophy Syndrome and Further Evidence for Genotype-Phenotype Correlations. Am. J. Med. Genet. Part A 2020, 182, 1426–1437. [Google Scholar] [CrossRef] [Green Version]
  39. Walsh, S.; Gösswein, S.S.; Rump, A.; von der Hagen, M.; Hackmann, K.; Schröck, E.; Di Donato, N.; Kahlert, A.-K. Novel Dominant-Negative NR2F1 Frameshift Mutation and a Phenotypic Expansion of the Bosch-Boonstra-Schaaf Optic Atrophy Syndrome. Eur. J. Med. Genet. 2020, 63, 104019. [Google Scholar] [CrossRef]
  40. Zou, W.; Cheng, L.; Lu, S.; Wu, Z. A de Novo Nonsense Mutation in the N-Terminal of Ligand-Binding Domain of NR2F1 Gene Provoked a Milder Phenotype of BBSOAS. Ophthalmic Genet. 2020, 41, 88–89. [Google Scholar] [CrossRef]
  41. Jurkute, N.; Bertacchi, M.; Arno, G.; Tocco, C.; Kim, U.S.; Kruszewski, A.M.; Avery, R.A.; Bedoukian, E.C.; Han, J.; Ahn, S.J.; et al. Pathogenic NR2F1 Variants Cause a Developmental Ocular Phenotype Recapitulated in a Mutant Mouse Model. Brain Commun. 2021, 3, fcab162. [Google Scholar] [CrossRef]
  42. Hobbs, M.M.; Wolters, W.C.; Rayapati, A.O. Bosch-Boonstra-Schaaf Optic Atrophy Syndrome Presenting as New-Onset Psychosis in a 32-Year-Old Man: A Case Report and Literature Review. J. Psychiatr. Pract. 2020, 26, 58–62. [Google Scholar] [CrossRef] [PubMed]
  43. Martín-Hernández, E.; Rodríguez-García, M.E.; Chen, C.-A.; Cotrina-Vinagre, F.J.; Carnicero-Rodríguez, P.; Bellusci, M.; Schaaf, C.P.; Martínez-Azorín, F. Mitochondrial Involvement in a Bosch-Boonstra-Schaaf Optic Atrophy Syndrome Patient with a Novel de Novo NR2F1 Gene Mutation. J. Hum. Genet. 2018, 63, 525–528. [Google Scholar] [CrossRef] [PubMed]
  44. Starosta, R.T.; Tarnowski, J.; e Vairo, F.P.; Raymond, K.; Preston, G.; Morava, E. Bosch-Boonstra-Schaaf Optic Atrophy Syndrome (BBSOAS) Initially Diagnosed as ALG6-CDG: Functional Evidence for Benignity of the ALG6 c.391T>C (p.Tyr131His) Variant and Further Expanding the BBSOAS Phenotype. Eur. J. Med. Genet. 2020, 63, 103941. [Google Scholar] [CrossRef] [PubMed]
  45. Yu-Wai-Man, P.; Griffiths, P.G.; Chinnery, P.F. Mitochondrial Optic Neuropathies—Disease Mechanisms and Therapeutic Strategies. Prog. Retin. Eye Res. 2011, 30, 81–114. [Google Scholar] [CrossRef] [Green Version]
  46. Jurkute, N.; Majander, A.; Bowman, R.; Votruba, M.; Abbs, S.; Acheson, J.; Lenaers, G.; Amati-Bonneau, P.; Moosajee, M.; Arno, G.; et al. Clinical Utility Gene Card for: Inherited Optic Neuropathies Including next-Generation Sequencing-Based Approaches. Eur. J. Hum. Genet. 2018, 27, 494–502. [Google Scholar] [CrossRef]
  47. Yu-Wai-Man, P.; Bailie, M.; Atawan, A.; Chinnery, P.F.; Griffiths, P.G. Pattern of Retinal Ganglion Cell Loss in Dominant Optic Atrophy Due to OPA1 Mutations. Eye 2011, 25, 596–602. [Google Scholar] [CrossRef] [Green Version]
  48. Pang, Y.; Frantz, K.A.; Roberts, D.K. Association of Refractive Error with Optic Nerve Hypoplasia. Ophthalmic Physiol. Opt. J. Br. Coll. Ophthalmic Opt. Optom. 2015, 35, 570–576. [Google Scholar] [CrossRef]
  49. Dahl, S.; Pettersson, M.; Eisfeldt, J.; Schröder, A.K.; Wickström, R.; Fahnehjelm, K.T.; Anderlid, B.M.; Lindstrand, A. Whole Genome Sequencing Unveils Genetic Heterogeneity in Optic Nerve Hypoplasia. PLoS ONE 2020, 15, 1–16. [Google Scholar] [CrossRef]
  50. Karahan, E.; Tulin Berk, A. Ocular, Neurologic and Systemic Findings of the Cases with Optic Nerve Hypoplasia. Open Ophthalmol. J. 2016, 10, 5–11. [Google Scholar] [CrossRef] [Green Version]
  51. Mosier, M.A.; Lieberman, M.F.; Green, W.R.; Knox, D.L. Hypoplasia of the Optic Nerve. Arch. Ophthalmol. Chic. Ill 1960 1978, 96, 1437–1442. [Google Scholar] [CrossRef]
  52. Taylor, D. Developmental Abnormalities of the Optic Nerve and Chiasm. Eye Lond. Engl. 2007, 21, 1271–1284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kaur, S.; Jain, S.; Sodhi, H.B.S.; Rastogi, A. Kamlesh Optic Nerve Hypoplasia. Oman J. Ophthalmol. 2013, 6, 77–82. [Google Scholar] [CrossRef] [PubMed]
  54. Pilat, A.; Sibley, D.; McLean, R.J.; Proudlock, F.A.; Gottlob, I. High-Resolution Imaging of the Optic Nerve and Retina in Optic Nerve Hypoplasia. Ophthalmology 2015, 122, 1330–1339. [Google Scholar] [CrossRef] [Green Version]
  55. Hoyt, C.S.; Good, W.V. Do We Really Understand the Difference between Optic Nerve Hypoplasia and Atrophy? Eye 1992, 6, 201–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Jones, R.; Al-Hayouti, H.; Oladiwura, D.; Karim, R.; Sawczenko, A.; Dahlmann-Noor, A. Optic Atrophy in Children: Current Causes and Diagnostic Approach. Eur. J. Ophthalmol. 2020, 30, 1499–1505. [Google Scholar] [CrossRef] [PubMed]
  57. McClard, C.K.; Shah, V. Pediatric Optic Disc Pallor. Int. Ophthalmol. Clin. 2018, 58, 125–145. [Google Scholar] [CrossRef] [PubMed]
  58. Yu-Wai-Man, P.; Votruba, M.; Burté, F.; La Morgia, C.; Barboni, P.; Carelli, V.; Burte, F.; La Morgia, C.; Barboni, P.; Carelli, V.; et al. A Neurodegenerative Perspective on Mitochondrial Optic Neuropathies. Acta Neuropathol. 2016, 132, 789–806. [Google Scholar] [CrossRef] [Green Version]
  59. Ludwig, P.E.; Lopez, M.J.; Czyz, C.N. Embryology, Eye Malformations. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  60. Ryabets-Lienhard, A.; Stewart, C.; Borchert, M.; Geffner, M.E. The Optic Nerve Hypoplasia Spectrum: Review of the Literature and Clinical Guidelines. Adv. Pediatr. 2016, 63, 127–146. [Google Scholar] [CrossRef] [PubMed]
  61. Garcia-Filion, P.; Borchert, M. Optic Nerve Hypoplasia Syndrome: A Review of the Epidemiology and Clinical Associations. Curr. Treat. Options Neurol. 2013, 15, 78–89. [Google Scholar] [CrossRef] [Green Version]
  62. Garcia-Filion, P.; Epport, K.; Nelson, M.; Azen, C.; Geffner, M.E.; Fink, C.; Borchert, M. Neuroradiographic, Endocrinologic, and Ophthalmic Correlates of Adverse Developmental Outcomes in Children with Optic Nerve Hypoplasia: A Prospective Study. Pediatrics 2008, 121, e653–e659. [Google Scholar] [CrossRef] [Green Version]
  63. Azuma, N.; Yamaguchi, Y.; Handa, H.; Tadokoro, K.; Asaka, A.; Kawase, E.; Yamada, M. Mutations of the PAX6 Gene Detected in Patients with a Variety of Optic-Nerve Malformations. Am. J. Hum. Genet. 2003, 72, 1565–1570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Chen, C.-A.; Yin, J.; Lewis, R.A.; Schaaf, C.P. Genetic Causes of Optic Nerve Hypoplasia. J. Med. Genet. 2017, 54, 441–449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Kelberman, D. Mutations within Sox2/SOX2 Are Associated with Abnormalities in the Hypothalamo-Pituitary-Gonadal Axis in Mice and Humans. J. Clin. Investig. 2006, 116, 2442–2455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Alfano, C. The Nuclear Receptors COUP-TF: A Long-Lasting Experience in Forebrain Assembly. Cell. Mol. Life Sci. 2014, 71, 43–62. [Google Scholar] [CrossRef]
  67. Yamaguchi, H.; Zhou, C.; Lin, S.-C.; Durand, B.; Tsai, S.Y.; Tsai, M.-J. The Nuclear Orphan Receptor COUP-TFI Is Important for Differentiation of Oligodendrocytes. Dev. Biol. 2004, 266, 238–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Eleuteri, C.; Olla, S.; Veroni, C.; Umeton, R.; Mechelli, R.; Romano, S.; Buscarinu, M.; Ferrari, F.; Calò, G.; Ristori, G.; et al. A Staged Screening of Registered Drugs Highlights Remyelinating Drug Candidates for Clinical Trials. Sci. Rep. 2017, 7, 45780. [Google Scholar] [CrossRef] [Green Version]
  69. Harlow, D.E.; Honce, J.M.; Miravalle, A.A. Remyelination Therapy in Multiple Sclerosis. Front. Neurol. 2015, 6, 257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Najm, F.J.; Madhavan, M.; Zaremba, A.; Shick, E.; Karl, R.T.; Factor, D.C.; Miller, T.E.; Nevin, Z.S.; Kantor, C.; Sargent, A.; et al. Drug-Based Modulation of Endogenous Stem Cells Promotes Functional Remyelination In Vivo. Nature 2015, 522, 216–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Porcu, G.; Serone, E.; De Nardis, V.; Di Giandomenico, D.; Lucisano, G.; Scardapane, M.; Poma, A.; Ragnini-Wilson, A. Clobetasol and Halcinonide Act as Smoothened Agonists to Promote Myelin Gene Expression and RxRγ Receptor Activation. PLoS ONE 2015, 10, e0144550. [Google Scholar] [CrossRef] [PubMed]
  72. Su, X.; Tang, W.; Luan, Z.; Yang, Y.; Wang, Z.; Zhang, Y.; Wang, Q.; Suo, L.; Huang, Z.; Wang, X.; et al. Protective Effect of Miconazole on Rat Myelin Sheaths Following Premature Infant Cerebral White Matter Injury. Exp. Ther. Med. 2018, 15, 2443–2449. [Google Scholar] [CrossRef] [PubMed]
  73. Abel, A.; McClelland, C.; Lee, M.S. Critical Review: Typical and Atypical Optic Neuritis. Surv. Ophthalmol. 2019, 64, 770–779. [Google Scholar] [CrossRef] [PubMed]
  74. Redler, Y.; Levy, M. Rodent Models of Optic Neuritis. Front. Neurol. 2020, 11, 580951. [Google Scholar] [CrossRef]
  75. Teixeira, L.B.C.; Hoeve, J.N.V.; Mayer, J.A.; Dubielzig, R.R.; Smith, C.M.; Radcliff, A.B.; Duncan, I.D. Modeling the Chronic Loss of Optic Nerve Axons and the Effects on the Retinal Nerve Fiber Layer Structure in Primary Disorder of Myelin. Investig. Opthalmol. Vis. Sci. 2016, 57, 4859. [Google Scholar] [CrossRef] [PubMed]
  76. Satoh, S.; Tang, K.; Iida, A.; Inoue, M.; Kodama, T.; Tsai, S.Y.; Tsai, M.-J.; Furuta, Y.; Watanabe, S. The Spatial Patterning of Mouse Cone Opsin Expression Is Regulated by Bone Morphogenetic Protein Signaling through Downstream Effector COUP-TF Nuclear Receptors. J. Neurosci. 2009, 29, 12401–12411. [Google Scholar] [CrossRef] [PubMed]
  77. Inoue, M.; Iida, A.; Satoh, S.; Kodama, T.; Watanabe, S. COUP-TFI and -TFII Nuclear Receptors Are Expressed in Amacrine Cells and Play Roles in Regulating the Differentiation of Retinal Progenitor Cells. Exp. Eye Res. 2010, 90, 49–56. [Google Scholar] [CrossRef] [PubMed]
  78. Jonas, J.B.; Budde, W.M.; Panda-Jonas, S. Ophthalmoscopic Evaluation of the Optic Nerve Head. Surv. Ophthalmol. 1999, 43, 293–320. [Google Scholar] [CrossRef]
  79. Rodenbeck, S.J.; Mackay, D.D. Examining the Ocular Fundus in Neurology. Curr. Opin. Neurol. 2019, 32, 105–110. [Google Scholar] [CrossRef] [Green Version]
  80. Cekić, S.; Stanković-Babić, G.; Višnjić, Z.; Jovanović, I.; Risimić, D. Optic Disc Abnormalities—Diagnosis, Evolution and Influence on Visual Acuity. Bosn. J. Basic Med. Sci. 2010, 10, 125–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Dutton, G.N. Congenital Disorders of the Optic Nerve: Excavations and Hypoplasia. Eye 2004, 18, 1038–1048. [Google Scholar] [CrossRef] [PubMed]
  82. Jeng-Miller, K.W.; Cestari, D.M.; Gaier, E.D. Congenital Anomalies of the Optic Disc: Insights from Optical Coherence Tomography Imaging. Curr. Opin. Ophthalmol. 2017, 28, 579–586. [Google Scholar] [CrossRef]
  83. Anglo, M. Topical Review: Optic Disc Pits and Associated Maculopathy. Optom. Vis. Sci. 2020, 97, 531–535. [Google Scholar] [CrossRef] [PubMed]
  84. Uzel, M.M.; Karacorlu, M. Optic Disk Pits and Optic Disk Pit Maculopathy: A Review. Surv. Ophthalmol. 2019, 64, 595–607. [Google Scholar] [CrossRef]
  85. Gazdagh, G.; Mawby, R.; Self, J.E.; Baralle, D. Deciphering Developmental Disorders Study A Severe Case of Bosch– Boonstra–Schaaf Optic Atrophy Syndrome with a Novel Description of Coloboma and Septo-optic Dysplasia, Owing to a Start Codon Variant in the NR2F1 Gene. Am. J. Med. Genet. A 2022, 188, 900–906. [Google Scholar] [CrossRef] [PubMed]
  86. Deiner, M.S.; Kennedy, T.E.; Fazeli, A.; Serafini, T.; Tessier-Lavigne, M.; Sretavan, D.W. Netrin-1 and DCC Mediate Axon Guidance Locally at the Optic Disc: Loss of Function Leads to Optic Nerve Hypoplasia. Neuron 1997, 19, 575–589. [Google Scholar] [CrossRef] [Green Version]
  87. Morcillo, J. Proper Patterning of the Optic Fissure Requires the Sequential Activity of BMP7 and SHH. Development 2006, 133, 3179–3190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Chung, G.W.; Edwards, A.O.; SchimentiI, L.A.; Manligas, G.S.; Zhang, Y.; Ritter, R. Renal-Coloboma Syndrome: Report of a Novel PAX2 Gene Mutation. Am. J. Ophthalmol. 2001, 132, 910–914. [Google Scholar] [CrossRef]
  89. Eccles, M.R.; Schimmenti, L.A. Renal-Coloboma Syndrome: A Multi-System Developmental Disorder Caused by PAX2 Mutations: PAX2 Mutation in Renal-Coloboma Syndrome. Clin. Genet. 1999, 56, 1–9. [Google Scholar] [CrossRef] [PubMed]
  90. Good, V.; Jan, J.E.; Barrovich, A.J.; Hoyt, S. Cortical Visual Impairment in Children. 14. Surv. Ophthalmol. 1994, 38, 351–364. [Google Scholar] [CrossRef]
  91. Good, W.V.; Jan, J.E.; Burden, S.K.; Skoczenski, A.; Candy, R. Recent Advances in Cortical Visual Impairment. Dev. Med. Child Neurol. 2007, 43, 56–60. [Google Scholar] [CrossRef]
  92. Hoyt, C.S. Visual Function in the Brain-Damaged Child. Eye 2003, 17, 369–384. [Google Scholar] [CrossRef] [PubMed]
  93. Huo, R.; Burden, S.K.; Hoyt, C.S.; Good, W.V. Chronic Cortical Visual Impairment in Children: Aetiology, Prognosis, and Associated Neurological Deficits. Br. J. Ophthalmol. 1999, 83, 670–675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Jan, J.E.; Groenveld, M.; Sykanda, A.M.; Hoyt, C.S. Behavioural Characteristics Of Children With Permanent Cortical Visual Impairment. Dev. Med. Child Neurol. 2008, 29, 571–576. [Google Scholar] [CrossRef] [PubMed]
  95. Malkowicz, D.E.; Myers, G.; Leisman, G. Rehabilitation of Cortical Visual Impairment in Children. Int. J. Neurosci. 2006, 116, 1015–1033. [Google Scholar] [CrossRef] [PubMed]
  96. McConnell, E.L.; Saunders, K.J.; Little, J. What Assessments Are Currently Used to Investigate and Diagnose Cerebral Visual Impairment (CVI) in Children? A Systematic Review. Ophthalmic Physiol. Opt. 2021, 41, 224–244. [Google Scholar] [CrossRef] [PubMed]
  97. Philip, S.S.; Dutton, G.N. Identifying and Characterising Cerebral Visual Impairment in Children: A Review: Cerebral Visual Impairment in Children: A Review. Clin. Exp. Optom. 2014, 97, 196–208. [Google Scholar] [CrossRef]
  98. Sakki, H.E.A.; Dale, N.J.; Sargent, J.; Perez-Roche, T.; Bowman, R. Is There Consensus in Defining Childhood Cerebral Visual Impairment? A Systematic Review of Terminology and Definitions. Br. J. Ophthalmol. 2018, 102, 424–432. [Google Scholar] [CrossRef] [Green Version]
  99. Chou, S.-J.; Babot, Z.; Leingartner, A.; Studer, M.; Nakagawa, Y.; O’Leary, D.D.M. Geniculocortical Input Drives Genetic Distinctions Between Primary and Higher-Order Visual Areas. Science 2013, 340, 1239–1242. [Google Scholar] [CrossRef] [Green Version]
  100. Cadwell, C.R.; Bhaduri, A.; Mostajo-Radji, M.A.; Keefe, M.G.; Nowakowski, T.J. Development and Arealization of the Cerebral Cortex. Neuron 2019, 103, 980–1004. [Google Scholar] [CrossRef]
  101. Alfano, C.; Studer, M. Neocortical Arealization: Evolution, Mechanisms, and Open Questions. Dev. Neurobiol. 2013, 73, 411–447. [Google Scholar] [CrossRef]
  102. Hofman, J.; Hutny, M.; Sztuba, K.; Paprocka, J. Corpus Callosum Agenesis: An Insight into the Etiology and Spectrum of Symptoms. Brain Sci. 2020, 10, 625. [Google Scholar] [CrossRef]
  103. Paul, L.K.; Brown, W.S.; Adolphs, R.; Tyszka, J.M.; Richards, L.J.; Mukherjee, P.; Sherr, E.H. Agenesis of the Corpus Callosum: Genetic, Developmental and Functional Aspects of Connectivity. Nat. Rev. Neurosci. 2007, 8, 287–299. [Google Scholar] [CrossRef] [PubMed]
  104. Schell-Apacik, C.C.; Wagner, K.; Bihler, M.; Ertl-Wagner, B.; Heinrich, U.; Klopocki, E.; Kalscheuer, V.M.; Muenke, M.; von Voss, H. Agenesis and Dysgenesis of the Corpus Callosum: Clinical, Genetic and Neuroimaging Findings in a Series of 41 Patients. Am. J. Med. Genet. A 2008, 146A, 2501–2511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Alfano, C.; Viola, L.; Heng, J.I.-T.; Pirozzi, M.; Clarkson, M.; Flore, G.; De Maio, A.; Schedl, A.; Guillemot, F.; Studer, M. COUP-TFI Promotes Radial Migration and Proper Morphology of Callosal Projection Neurons by Repressing Rnd2 Expression. Development 2011, 138, 4685–4697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Armentano, M.; Filosa, A.; Andolfi, G.; Studer, M. COUP-TFI Is Required for the Formation of Commissural Projections in the Forebrain by Regulating Axonal Growth. Development 2006, 133, 4151–4162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Tenembaum, S.N. Acute Disseminated Encephalomyelitis. In Handbook of Clinical Neurology; Elsevier: Amsterdam, The Netherlands, 2013; Volume 112, pp. 1253–1262. ISBN 978-0-444-52910-7. [Google Scholar]
  108. Broadbent, N.J.; Squire, L.R.; Clark, R.E. Spatial Memory, Recognition Memory, and the Hippocampus. Proc. Natl. Acad. Sci. USA 2004, 101, 14515–14520. [Google Scholar] [CrossRef] [Green Version]
  109. Hannula, D.E.; Libby, L.A.; Yonelinas, A.P.; Ranganath, C. Medial Temporal Lobe Contributions to Cued Retrieval of Items and Contexts. Neuropsychologia 2013, 51, 2322–2332. [Google Scholar] [CrossRef] [Green Version]
  110. Kumaran, D.; Maguire, E.A. The Human Hippocampus: Cognitive Maps or Relational Memory? J. Neurosci. 2005, 25, 7254–7259. [Google Scholar] [CrossRef]
  111. Mátéffyová, A.; Otáhal, J.; Tsenov, G.; Mareš, P.; Kubová, H. Intrahippocampal Injection of Endothelin-1 in Immature Rats Results in Neuronal Death, Development of Epilepsy and Behavioral Abnormalities Later in Life. Eur. J. Neurosci. 2006, 24, 351–360. [Google Scholar] [CrossRef]
  112. Ramos, J.M.J. Hippocampal Damage Impairs Long-Term Spatial Memory in Rats: Comparison between Electrolytic and Neurotoxic Lesions. Physiol. Behav. 2008, 93, 1078–1085. [Google Scholar] [CrossRef]
  113. Wan, J.; Shen, C.M.; Wang, Y.; Wu, Q.Z.; Wang, Y.L.; Liu, Q.; Sun, Y.M.; Cao, J.P.; Wu, Y.Q. Repeated Exposure to Propofol in the Neonatal Period Impairs Hippocampal Synaptic Plasticity and the Recognition Function of Rats in Adulthood. Brain Res. Bull. 2021, 169, 63–72. [Google Scholar] [CrossRef]
  114. Zhao, A.; Fang, F.; Li, B.; Chen, Y.; Qiu, Y.; Wu, Y.; Xu, W.; Deng, Y. Visual Abnormalities Associate With Hippocampus in Mild Cognitive Impairment and Early Alzheimer’s Disease. Front. Aging Neurosci. 2021, 12, 522. [Google Scholar] [CrossRef] [PubMed]
  115. Moser, M.B.; Moser, E.I.; Forrest, E.; Andersen, P.; Morris, R.G.M. Spatial Learning with a Minislab in the Dorsal Hippocampus. Proc. Natl. Acad. Sci. USA 1995, 92, 9697–9701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Oberhuber, M.; Hope, T.M.H.; Seghier, M.L.; Parker Jones, O.; Prejawa, S.; Green, D.W.; Price, C.J. Four Functionally Distinct Regions in the Left Supramarginal Gyrus Support Word Processing. Cereb. Cortex 2016, 26, 4212–4226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Seghier, M.L. The Angular Gyrus: Multiple Functions and Multiple Subdivisions. Neuroscientist 2013, 19, 43–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Stoeckel, C.; Gough, P.M.; Watkins, K.E.; Devlin, J.T. Supramarginal Gyrus Involvement in Visual Word Recognition. Cortex 2009, 45, 1091–1096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Barkovich, A.J.; Guerrini, R.; Kuzniecky, R.I.; Jackson, G.D.; Dobyns, W.B. A Developmental and Genetic Classification for Malformations of Cortical Development: Update 2012. Brain 2012, 135, 1348–1369. [Google Scholar] [CrossRef] [PubMed]
  120. Guerrini, R.; Dobyns, W.B. Malformations of Cortical Development: Clinical Features and Genetic Causes. Lancet Neurol. 2014, 13, 710–726. [Google Scholar] [CrossRef] [Green Version]
  121. Juric-Sekhar, G.; Hevner, R.F. Malformations of Cerebral Cortex Development: Molecules and Mechanisms. Annu. Rev. Pathol. Mech. Dis. 2019, 14, 293–318. [Google Scholar] [CrossRef]
  122. Parrini, E.; Conti, V.; Dobyns, W.B.; Guerrini, R. Genetic Basis of Brain Malformations. Mol. Syndromol. 2016, 7, 220–233. [Google Scholar] [CrossRef] [Green Version]
  123. Hansen, D.V.; Lui, J.H.; Parker, P.R.L.; Kriegstein, A.R. Neurogenic Radial Glia in the Outer Subventricular Zone of Human Neocortex. Nature 2010, 464, 554–561. [Google Scholar] [CrossRef]
  124. Shitamukai, A.; Konno, D.; Matsuzaki, F. Oblique Radial Glial Divisions in the Developing Mouse Neocortex Induce Self-Renewing Progenitors Outside the Germinal Zone That Resemble Primate Outer Subventricular Zone Progenitors. J. Neurosci. 2011, 31, 3683–3695. [Google Scholar] [CrossRef] [PubMed]
  125. Vaid, S.; Camp, J.G.; Hersemann, L.; Eugster Oegema, C.; Heninger, A.-K.; Winkler, S.; Brandl, H.; Sarov, M.; Treutlein, B.; Huttner, W.B.; et al. A Novel Population of Hopx-Dependent Basal Radial Glial Cells in the Developing Mouse Neocortex. Development 2018, 145, dev169276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Wang, X.; Tsai, J.-W.; LaMonica, B.; Kriegstein, A.R. A New Subtype of Progenitor Cell in the Mouse Embryonic Neocortex. Nat. Neurosci. 2011, 14, 555–561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Wong, F.K.; Fei, J.-F.; Mora-Bermúdez, F.; Taverna, E.; Haffner, C.; Fu, J.; Anastassiadis, K.; Stewart, A.F.; Huttner, W.B. Sustained Pax6 Expression Generates Primate-like Basal Radial Glia in Developing Mouse Neocortex. PLoS Biol. 2015, 13, e1002217. [Google Scholar] [CrossRef] [Green Version]
  128. Fietz, S.A.; Kelava, I.; Vogt, J.; Wilsch-Bräuninger, M.; Stenzel, D.; Fish, J.L.; Corbeil, D.; Riehn, A.; Distler, W.; Nitsch, R.; et al. OSVZ Progenitors of Human and Ferret Neocortex Are Epithelial-like and Expand by Integrin Signaling. Nat. Neurosci. 2010, 13, 690–699. [Google Scholar] [CrossRef]
  129. Bershteyn, M.; Nowakowski, T.J.; Pollen, A.A.; Di Lullo, E.; Nene, A.; Wynshaw-Boris, A.; Kriegstein, A.R. Human IPSC-Derived Cerebral Organoids Model Cellular Features of Lissencephaly and Reveal Prolonged Mitosis of Outer Radial Glia. Cell Stem Cell 2017, 20, 435–449.e4. [Google Scholar] [CrossRef] [Green Version]
  130. Li, Y.; Muffat, J.; Omer, A.; Bosch, I.; Lancaster, M.A.; Sur, M.; Gehrke, L.; Knoblich, J.A.; Jaenisch, R. Induction of Expansion and Folding in Human Cerebral Organoids. Cell Stem Cell 2017, 20, 385–396.e3. [Google Scholar] [CrossRef] [Green Version]
  131. Qian, X.; Jacob, F.; Song, M.M.; Nguyen, H.N.; Song, H.; Ming, G. Generation of Human Brain Region–Specific Organoids Using a Miniaturized Spinning Bioreactor. Nat. Protoc. 2018, 13, 565–580. [Google Scholar] [CrossRef]
  132. Autism Sequencing Consortium; Lim, E.T.; Uddin, M.; De Rubeis, S.; Chan, Y.; Kamumbu, A.S.; Zhang, X.; D’Gama, A.M.; Kim, S.N.; Hill, R.S.; et al. Rates, Distribution and Implications of Postzygotic Mosaic Mutations in Autism Spectrum Disorder. Nat. Neurosci. 2017, 20, 1217–1224. [Google Scholar] [CrossRef] [Green Version]
  133. Parikshak, N.N.; Luo, R.; Zhang, A.; Won, H.; Lowe, J.K.; Chandran, V.; Horvath, S.; Geschwind, D.H. Integrative Functional Genomic Analyses Implicate Specific Molecular Pathways and Circuits in Autism. Cell 2013, 155, 1008–1021. [Google Scholar] [CrossRef] [Green Version]
  134. The DDD Study; Homozygosity Mapping Collaborative for Autism; UK10K Consortium; The Autism Sequencing Consortium; De Rubeis, S.; He, X.; Goldberg, A.P.; Poultney, C.S.; Samocha, K.; Ercument Cicek, A.; et al. Synaptic, Transcriptional and Chromatin Genes Disrupted in Autism. Nature 2014, 515, 209–215. [Google Scholar] [CrossRef] [PubMed]
  135. Cotney, J.; Muhle, R.A.; Sanders, S.J.; Liu, L.; Willsey, A.J.; Niu, W.; Liu, W.; Klei, L.; Lei, J.; Yin, J.; et al. The Autism-Associated Chromatin Modifier CHD8 Regulates Other Autism Risk Genes during Human Neurodevelopment. Nat. Commun. 2015, 6, 6404. [Google Scholar] [CrossRef] [PubMed]
  136. Watt, N.; Wetherby, A.M.; Barber, A.; Morgan, L. Repetitive and Stereotyped Behaviors in Children with Autism Spectrum Disorders in the Second Year of Life. J. Autism Dev. Disord. 2008, 38, 1518–1533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Molloy, A.; Rowe, F.J. Manneristic Behaviors of Visually Impaired Children. Strabismus 2011, 19, 77–84. [Google Scholar] [CrossRef] [PubMed]
  138. Kerr, M.; Gil-Nagel, A.; Glynn, M.; Mula, M.; Thompson, R.; Zuberi, S.M. Treatment of Behavioral Problems in Intellectually Disabled Adult Patients with Epilepsy. Epilepsia 2013, 54, 34–40. [Google Scholar] [CrossRef] [PubMed]
  139. Furniss, F.; Biswas, A.B. Recent Research on Aetiology, Development and Phenomenology of Self-Injurious Behaviour in People with Intellectual Disabilities: A Systematic Review and Implications for Treatment: Aetiology & Phenomenology of SIB in ID: A Systematic Review. J. Intellect. Disabil. Res. 2012, 56, 453–475. [Google Scholar] [CrossRef]
  140. Inoue, M. Assessments and Interventions to Address Challenging Behavior in Individuals with Intellectual Disability and Autism Spectrum Disorder in Japan: A Consolidated Review. Yonago Acta Med. 2019, 62, 169–181. [Google Scholar] [CrossRef] [Green Version]
  141. Matson, J.L.; Dempsey, T.; Rivet, T.T. The Interrelationships of Psychopathology Symptoms on Social Skills in Adults with Autism or PDD-NOS and Intellectual Disability. J. Dev. Phys. Disabil. 2009, 21, 39–55. [Google Scholar] [CrossRef]
  142. Oliver, C.; Licence, L.; Richards, C. Self-Injurious Behaviour in People with Intellectual Disability and Autism Spectrum Disorder. Curr. Opin. Psychiatry 2017, 30, 97–101. [Google Scholar] [CrossRef]
  143. Smith, K.R.M.; Matson, J.L. Social Skills: Differences among Adults with Intellectual Disabilities, Co-Morbid Autism Spectrum Disorders and Epilepsy. Res. Dev. Disabil. 2010, 31, 1366–1372. [Google Scholar] [CrossRef]
  144. Lowe, K.; Allen, D.; Jones, E.; Brophy, S.; Moore, K.; James, W. Challenging Behaviours: Prevalence and Topographies. J. Intellect. Disabil. Res. 2007, 51, 625–636. [Google Scholar] [CrossRef]
  145. Jacob, S.; Landeros-Weisenberger, A.; Leckman, J.F. Autism Spectrum and Obsessive–Compulsive Disorders: OC Behaviors, Phenotypes and Genetics. Autism Res. 2009, 2, 293–311. [Google Scholar] [CrossRef] [PubMed]
  146. Muehlmann, A.M.; Lewis, M.H. Abnormal Repetitive Behaviours: Shared Phenomenology and Pathophysiology: Abnormal Repetitive Behaviours. J. Intellect. Disabil. Res. 2012, 56, 427–440. [Google Scholar] [CrossRef] [PubMed]
  147. Amaral, D.G.; Schumann, C.M.; Nordahl, C.W. Neuroanatomy of Autism. Trends Neurosci. 2008, 31, 137–145. [Google Scholar] [CrossRef] [PubMed]
  148. Ha, S.; Sohn, I.-J.; Kim, N.; Sim, H.J.; Cheon, K.-A. Characteristics of Brains in Autism Spectrum Disorder: Structure, Function and Connectivity across the Lifespan. Exp. Neurobiol. 2015, 24, 273–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Lee, J.K.; Andrews, D.S.; Ozonoff, S.; Solomon, M.; Rogers, S.; Amaral, D.G.; Nordahl, C.W. Longitudinal Evaluation of Cerebral Growth Across Childhood in Boys and Girls With Autism Spectrum Disorder. Biol. Psychiatry 2020, 90, 286–294. [Google Scholar] [CrossRef]
  150. Adhya, D.; Swarup, V.; Nagy, R.; Dutan, L.; Shum, C.; Valencia-Alarcón, E.P.; Jozwik, K.M.; Mendez, M.A.; Horder, J.; Loth, E.; et al. Atypical Neurogenesis in Induced Pluripotent Stem Cells From Autistic Individuals. Biol. Psychiatry 2021, 89, 486–496. [Google Scholar] [CrossRef] [PubMed]
  151. Mariani, J.; Coppola, G.; Zhang, P.; Abyzov, A.; Provini, L.; Tomasini, L.; Amenduni, M.; Szekely, A.; Palejev, D.; Wilson, M.; et al. FOXG1-Dependent Dysregulation of GABA/Glutamate Neuron Differentiation in Autism Spectrum Disorders. Cell 2015, 162, 375–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Port, R.G.; Oberman, L.M.; Roberts, T.P. Revisiting the Excitation/Inhibition Imbalance Hypothesis of ASD through a Clinical Lens. Br. J. Radiol. 2019, 92, 20180944. [Google Scholar] [CrossRef]
  153. Siegel-Ramsay, J.E.; Romaniuk, L.; Whalley, H.C.; Roberts, N.; Branigan, H.; Stanfield, A.C.; Lawrie, S.M.; Dauvermann, M.R. Glutamate and Functional Connectivity—Support for the Excitatory-Inhibitory Imbalance Hypothesis in Autism Spectrum Disorders. Psychiatry Res. Neuroimaging 2021, 313, 111302. [Google Scholar] [CrossRef]
  154. Willsey, A.J.; Sanders, S.J.; Li, M.; Dong, S.; Tebbenkamp, A.T.; Muhle, R.A.; Reilly, S.K.; Lin, L.; Fertuzinhos, S.; Miller, J.A.; et al. Coexpression Networks Implicate Human Midfetal Deep Cortical Projection Neurons in the Pathogenesis of Autism. Cell 2013, 155, 997–1007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Zhang, K.; Yu, F.; Zhu, J.; Han, S.; Chen, J.; Wu, X.; Chen, Y.; Shen, T.; Liao, J.; Guo, W.; et al. Imbalance of Excitatory/Inhibitory Neuron Differentiation in Neurodevelopmental Disorders with an NR2F1 Point Mutation. Cell Rep. 2020, 31, 107521. [Google Scholar] [CrossRef] [PubMed]
  156. Paulsen, B.; Velasco, S.; Kedaigle, A.J.; Pigoni, M.; Quadrato, G.; Deo, A.J.; Adiconis, X.; Uzquiano, A.; Sartore, R.; Yang, S.M.; et al. Autism Genes Converge on Asynchronous Development of Shared Neuron Classes. Nature 2022, 602, 268–273. [Google Scholar] [CrossRef] [PubMed]
  157. Bjørnæs, H.; Stabell, K.; Henriksen, O.; Løyning, Y. The Effects of Refractory Epilepsy on Intellectual Functioning in Children and Adults. A Longitudinal Study. Seizure 2001, 10, 250–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Duncan, J.S. MRI Studies. Do Seizures Damage the Brain? In Progress in Brain Research; Elsevier: Amsterdam, The Netherlands, 2002; Volume 135, pp. 253–261. ISBN 978-0-444-50814-0. [Google Scholar]
  159. Mameniškienė, R.; Puteikis, K.; Jasionis, A.; Jatužis, D. A Review of Accelerated Long-Term Forgetting in Epilepsy. Brain Sci. 2020, 10, 945. [Google Scholar] [CrossRef]
  160. van Rijckevorsel, K. Cognitive Problems Related to Epilepsy Syndromes, Especially Malignant Epilepsies. Seizure 2006, 15, 227–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Shields, W.D. Infantile Spasms: Little Seizures, BIG Consequences. Epilepsy Curr. 2006, 6, 63–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Sutula, T.P.; Hagen, J.; Pitkänen, A. Do Epileptic Seizures Damage the Brain? Curr. Opin. Neurol. 2003, 16, 189–195. [Google Scholar] [CrossRef]
  163. Pisani, F.; Fusco, C.; Spagnoli, C. Linking Acute Symptomatic Neonatal Seizures, Brain Injury and Outcome in Preterm Infants. Epilepsy Behav. 2020, 112, 107406. [Google Scholar] [CrossRef] [PubMed]
  164. Sutula, T.P.; Pitkanen, A. More Evidence for Seizure-Induced Neuron Loss: Is Hippocampal Sclerosis Both Cause and Effect of Epilepsy? Neurology 2001, 57, 169–170. [Google Scholar] [CrossRef]
  165. Scharfman, H.E. The Neurobiology of Epilepsy. Curr. Neurol. Neurosci. Rep. 2007, 7, 348–354. [Google Scholar] [CrossRef] [PubMed]
  166. Bozzi, Y.; Provenzano, G.; Casarosa, S. Neurobiological Bases of Autism-Epilepsy Comorbidity: A Focus on Excitation/Inhibition Imbalance. Eur. J. Neurosci. 2018, 47, 534–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Powell, E.M.; Campbell, D.B.; Stanwood, G.D.; Davis, C.; Noebels, J.L.; Levitt, P. Genetic Disruption of Cortical Interneuron Development Causes Region- and GABA Cell Type-Specific Deficits, Epilepsy, and Behavioral Dysfunction. J. Neurosci. 2003, 23, 622–631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Lévesque, M.; Biagini, G.; Avoli, M. Neurosteroids and Focal Epileptic Disorders. Int. J. Mol. Sci. 2020, 21, 9391. [Google Scholar] [CrossRef] [PubMed]
  169. Verhoog, Q.P.; Holtman, L.; Aronica, E.; van Vliet, E.A. Astrocytes as Guardians of Neuronal Excitability: Mechanisms Underlying Epileptogenesis. Front. Neurol. 2020, 11. [Google Scholar] [CrossRef]
  170. Anwar, H.; Khan, Q.U.; Nadeem, N.; Pervaiz, I.; Ali, M.; Cheema, F.F. Epileptic Seizures. Discov. Craiova Rom. 2020, 8, e110. [Google Scholar] [CrossRef] [PubMed]
  171. Wu, X.; Li, H.; Huang, J.; Xu, M.; Xiao, C.; He, S. Regulation of Axon Initial Segment Diameter by COUP-TFI Fine-Tunes Action Potential Generation. Neurosci. Bull. 2021, 1–14. [Google Scholar] [CrossRef] [PubMed]
  172. Aghakhani, Y. The Role of Periventricular Nodular Heterotopia in Epileptogenesis. Brain 2005, 128, 641–651. [Google Scholar] [CrossRef] [Green Version]
  173. Khoo, H.M.; Gotman, J.; Hall, J.A.; Dubeau, F. Treatment of Epilepsy Associated with Periventricular Nodular Heterotopia. Curr. Neurol. Neurosci. Rep. 2020, 20, 1–11. [Google Scholar] [CrossRef] [PubMed]
  174. Meroni, A.; Galli, C.; Bramerio, M.; Tassi, L.; Colombo, N.; Cossu, M.; Lo Russo, G.; Garbelli, R.; Spreafico, R. Nodular Heterotopia: A Neuropathological Study of 24 Patients Undergoing Surgery for Drug-Resistant Epilepsy. Epilepsia 2009, 50, 116–124. [Google Scholar] [CrossRef] [PubMed]
  175. Valton, L.; Guye, M.; McGonigal, A.; Marquis, P.; Wendling, F.; Régis, J.; Chauvel, P.; Bartolomei, F. Functional Interactions in Brain Networks Underlying Epileptic Seizures in Bilateral Diffuse Periventricular Heterotopia. Clin. Neurophysiol. 2008, 119, 212–223. [Google Scholar] [CrossRef]
  176. Balciuniene, J.; DeChene, E.T.; Akgumus, G.; Romasko, E.J.; Cao, K.; Dubbs, H.A.; Mulchandani, S.; Spinner, N.B.; Conlin, L.K.; Marsh, E.D.; et al. Use of a Dynamic Genetic Testing Approach for Childhood-Onset Epilepsy. JAMA Netw. Open 2019, 2, e192129. [Google Scholar] [CrossRef]
  177. Dimassi, S.; Labalme, A.; Ville, D.; Calender, A.; Mignot, C.; Boutry-Kryza, N.; de Bellescize, J.; Rivier-Ringenbach, C.; Bourel-Ponchel, E.; Cheillan, D.; et al. Whole-Exome Sequencing Improves the Diagnosis Yield in Sporadic Infantile Spasm Syndrome. Clin. Genet. 2016, 89, 198–204. [Google Scholar] [CrossRef] [PubMed]
  178. Hino-Fukuyo, N.; Kikuchi, A.; Arai-Ichinoi, N.; Niihori, T.; Sato, R.; Suzuki, T.; Kudo, H.; Sato, Y.; Nakayama, T.; Kakisaka, Y.; et al. Genomic Analysis Identifies Candidate Pathogenic Variants in 9 of 18 Patients with Unexplained West Syndrome. Hum. Genet. 2015, 134, 649–658. [Google Scholar] [CrossRef] [PubMed]
  179. Fournier, K.A.; Hass, C.J.; Naik, S.K.; Lodha, N.; Cauraugh, J.H. Motor Coordination in Autism Spectrum Disorders: A Synthesis and Meta-Analysis. J. Autism Dev. Disord. 2010, 40, 1227–1240. [Google Scholar] [CrossRef] [PubMed]
  180. Jeste, S.S.; Geschwind, D.H. Clinical Trials for Neurodevelopmental Disorders: At a Therapeutic Frontier. Sci. Transl. Med. 2016, 8, 321fs1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Wilson, R.B.; Enticott, P.G.; Rinehart, N.J. Motor Development and Delay: Advances in Assessment of Motor Skills in Autism Spectrum Disorders. Curr. Opin. Neurol. 2018, 31, 134–139. [Google Scholar] [CrossRef] [PubMed]
  182. Gerber, R.J.; Wilks, T.; Erdie-Lalena, C. Developmental Milestones: Motor Development. Pediatr. Rev. 2010, 31, 267–277. [Google Scholar] [CrossRef]
  183. Mosconi, M.W.; Kay, M.; D’Cruz, A.M.; Seidenfeld, A.; Guter, S.; Stanford, L.D.; Sweeney, J.A. Impaired Inhibitory Control Is Associated with Higher-Order Repetitive Behaviors in Autism Spectrum Disorders. Psychol. Med. 2009, 39, 1559–1566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Donkelaar, H.J.; Lammens, M.; Wesseling, P.; Hori, A.; Keyser, A.; Rotteveel, J. Development and Malformations of the Human Pyramidal Tract. J. Neurol. 2004, 251, 1429–1442. [Google Scholar] [CrossRef]
  185. Roessmann, U.; Hori, A. Agyria (Lissencephaly) with Anomalous Pyramidal Crossing. Case Report and Review of Literature. J. Neurol. Sci. 1985, 69, 357–364. [Google Scholar] [CrossRef]
  186. Chow, C.W.; Halliday, J.L.; Anderson, R.M.D.; Danks, D.M.; Fortune, D.W. Congenital Absence of Pyramids and Its Significance in Genetic Diseases. Acta Neuropathol. 1985, 65, 313–317. [Google Scholar] [CrossRef] [PubMed]
  187. Marcorelles, P.; Laquerriere, A. Neuropathology of Holoprosencephaly. Am. J. Med. Genet. C Semin. Med. Genet. 2010, 154C, 109–119. [Google Scholar] [CrossRef] [PubMed]
  188. Welniarz, Q.; Dusart, I.; Roze, E. The Corticospinal Tract: Evolution, Development, and Human Disorders. Dev. Neurobiol. 2017, 810–829. [Google Scholar] [CrossRef]
  189. Tocco, C.; Øvsthus, M.; Bjaalie, J.G.; Leergaard, T.B.; Studer, M. The Topography of Corticopontine Projections Is Controlled by Postmitotic Expression of the Area-Mapping Gene Nr2f1. Development 2022, 149, dev200026. [Google Scholar] [CrossRef]
  190. Greig, L.C.; Woodworth, M.B.; Galazo, M.J.; Padmanabhan, H.; Macklis, J.D. Molecular Logic of Neocortical Projection Neuron Specification, Development and Diversity. Nat. Rev. Neurosci. 2013, 14, 755–769. [Google Scholar] [CrossRef] [Green Version]
  191. Contesse, T.; Ayrault, M.; Mantegazza, M.; Studer, M.; Deschaux, O. Hyperactive and Anxiolytic-like Behaviors Result from Loss of COUP-TFI/Nr2f1 in the Mouse Cortex. Genes Brain Behav. 2019, 18, e12556. [Google Scholar] [CrossRef]
  192. Iverson, J.M. Developing Language in a Developing Body: The Relationship between Motor Development and Language Development. J. Child Lang. 2010, 37, 229–261. [Google Scholar] [CrossRef]
  193. Esposito, G.; Venuti, P.; Maestro, S.; Muratori, F. An Exploration of Symmetry in Early Autism Spectrum Disorders: Analysis of Lying. Brain Dev. 2009, 31, 131–138. [Google Scholar] [CrossRef]
  194. Bhat, A.N.; Landa, R.J.; Galloway, J.C. Current Perspectives on Motor Functioning in Infants, Children, and Adults with Autism Spectrum Disorders. Phys. Ther. 2011, 91, 1116–1129. [Google Scholar] [CrossRef] [Green Version]
  195. Karasik, L.B.; Tamis-Lemonda, C.S.; Adolph, K.E. Transition from Crawling to Walking and Infants’ Actions with Objects and People. Child Dev. 2011, 82, 1199–1209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Barske, L.; Rataud, P.; Behizad, K.; Del Rio, L.; Cox, S.G.; Crump, J.G. Essential Role of Nr2f Nuclear Receptors in Patterning the Vertebrate Upper Jaw. Dev. Cell 2018, 44, 337–347.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. DeLaurier, A. Evolution and Development of the Fish Jaw Skeleton. Wiley Interdiscip. Rev. Dev. Biol. 2019, 8, e337. [Google Scholar] [CrossRef] [PubMed]
  198. Rada-Iglesias, A.; Bajpai, R.; Prescott, S.; Brugmann, S.A.; Swigut, T.; Wysocka, J. Epigenomic Annotation of Enhancers Predicts Transcriptional Regulators of Human Neural Crest. Cell Stem Cell 2012, 11, 633–648. [Google Scholar] [CrossRef] [Green Version]
  199. Manikandan, M.; Abuelreich, S.; Elsafadi, M.; Alsalman, H.; Almalak, H.; Siyal, A.; Hashmi, J.A.; Aldahmash, A.; Kassem, M.; Alfayez, M.; et al. NR2F1 Mediated Down-Regulation of Osteoblast Differentiation Was Rescued by Bone Morphogenetic Protein-2 (BMP-2) in Human MSC. Differentiation 2018, 104, 36–41. [Google Scholar] [CrossRef]
  200. Ingulli, E.G.; Mak, R.H. Growth in Children with Chronic Kidney Disease: Role of Nutrition, Growth Hormone, Dialysis, and Steroids. Curr. Opin. Pediatr. 2014, 26, 187–192. [Google Scholar] [CrossRef]
  201. Putnick, D.L.; Bell, E.M.; Ghassabian, A.; Robinson, S.L.; Sundaram, R.; Yeung, E. Feeding Problems as an Indicator of Developmental Delay in Early Childhood. J. Pediatr. 2021, S0022347621010751. [Google Scholar] [CrossRef]
  202. Boyle, K.B.; Anderson, J.M. Hypotonia and Poor Feeding in a Newborn. NeoReviews 2008, 9, e602–e604. [Google Scholar] [CrossRef]
  203. Martin, K.; Inman, J.; Kirschner, A.; Deming, K.; Gumbel, R.; Voelker, L. Characteristics of Hypotonia in Children: A Consensus Opinion of Pediatric Occupational and Physical Therapists. Pediatr. Phys. Ther. 2005, 17, 275–282. [Google Scholar] [CrossRef]
  204. Prasad, A.N.; Prasad, C. The Floppy Infant: Contribution of Genetic and Metabolic Disorders. Brain Dev. 2003, 25, 457–476. [Google Scholar] [CrossRef]
  205. Qiu, Y.; Pereira, F.A.; DeMayo, F.J.; Lydon, J.P.; Tsai, S.Y.; Tsai, M.-J. Null Mutation of MCOUP-TFI Results in Defects in Morphogenesis of the Glossopharyngeal Ganglion, Axonal Projection, and Arborization. Genes Dev. 1997, 11, 1925–1937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Khalil, B.D.; Sanchez, R.; Rahman, T.; Rodriguez-Tirado, C.; Moritsch, S.; Martinez, A.R.; Miles, B.; Farias, E.; Mezei, M.; Nobre, A.R.; et al. An NR2F1-Specific Agonist Suppresses Metastasis by Inducing Cancer Cell Dormancy. J. Exp. Med. 2022, 219, e20210836. [Google Scholar] [CrossRef] [PubMed]
  207. Yoon, K.; Chen, C.-C.; Orr, A.; Barreto, P.; Tamamis, P.; Safe, S. Activation of COUP-TFI by a Novel Diindolylmethane Derivative. Cells 2019, 8, 220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Bourguet, W.; Ruff, M.; Chambon, P.; Gronemeyer, H. Crystal Structure of the Ligand-Binding Domain of the Human Nuclear Receptor RXR-α. Nature 1995, 375, 377–382. [Google Scholar] [CrossRef]
  209. Egea, P.F.; Mitschler, A.; Rochel, N.; Ruff, M.; Chambon, P.; Moras, D. Crystal Structure of the Human RXRa Ligand-Binding Domain Bound to Its Natural Ligand: 9-Cis Retinoic Acid. 10. EMBO J. 2000, 19, 2592–2601. [Google Scholar] [CrossRef]
  210. Germain, P.; Staels, B.; Dacquet, C.; Spedding, M.; Laudet, V. Overview of Nomenclature of Nuclear Receptors. Pharmacol. Rev. 2006, 58, 685–704. [Google Scholar] [CrossRef]
  211. Wang, Z.; Benoit, G.; Liu, J.; Prasad, S.; Aarnisalo, P.; Liu, X.; Xu, H.; Walker, N.P.C.; Perlmann, T. Structure and Function of Nurr1 Identifies a Class of Ligand-Independent Nuclear Receptors. Nature 2003, 423, 555–560. [Google Scholar] [CrossRef]
  212. Tsai, S.Y.; Tsai, M.-J. Chick Ovalbumin Upstream Promoter-Transcription Factors (COUP-TFs): Coming of Age. Endocr. Rev. 1997, 18, 229–240. [Google Scholar]
  213. Ang, C.E.; Ma, Q.; Wapinski, O.L.; Fan, S.; Flynn, R.A.; Lee, Q.Y.; Coe, B.; Onoguchi, M.; Olmos, V.H.; Do, B.T.; et al. The Novel LncRNA Lnc-NR2F1 Is pro-Neurogenic and Mutated in Human Neurodevelopmental Disorders. eLife 2019, 8, 41770. [Google Scholar] [CrossRef]
  214. Cooper, G.M.; Goode, D.L.; Ng, S.B.; Sidow, A.; Bamshad, M.J.; Shendure, J.; Nickerson, D.A. Single-Nucleotide Evolutionary Constraint Scores Highlight Disease-Causing Mutations. Nat. Methods 2010, 7, 250–251. [Google Scholar] [CrossRef] [Green Version]
  215. Gussow, A.B.; Petrovski, S.; Wang, Q.; Allen, A.S.; Goldstein, D.B. The Intolerance to Functional Genetic Variation of Protein Domains Predicts the Localization of Pathogenic Mutations within Genes. Genome Biol. 2016, 17, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Kircher, M.; Witten, D.M.; Jain, P.; O’Roak, B.J.; Cooper, G.M.; Shendure, J. A General Framework for Estimating the Relative Pathogenicity of Human Genetic Variants. Nat. Genet. 2014, 46, 310–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Rivas, M.A.; Pirinen, M.; Conrad, D.F.; Lek, M.; Tsang, E.K.; Karczewski, K.J.; Maller, J.B.; Kukurba, K.R.; DeLuca, D.S.; Fromer, M.; et al. Effect of Predicted Protein-Truncating Genetic Variants on the Human Transcriptome. Science 2015, 348, 666–669. [Google Scholar] [CrossRef] [Green Version]
  218. Holbrook, J.A.; Neu-Yilik, G.; Hentze, M.W.; Kulozik, A.E. Nonsense-Mediated Decay Approaches the Clinic. Nat. Genet. 2004, 36, 801–808. [Google Scholar] [CrossRef] [PubMed]
  219. Miller, J.N.; Pearce, D.A. Nonsense-Mediated Decay in Genetic Disease: Friend or Foe? Mutat. Res. Mutat. Res. 2014, 762, 52–64. [Google Scholar] [CrossRef] [Green Version]
  220. DeBoever, C.; Tanigawa, Y.; Lindholm, M.E.; McInnes, G.; Lavertu, A.; Ingelsson, E.; Chang, C.; Ashley, E.A.; Bustamante, C.D.; Daly, M.J.; et al. Medical Relevance of Protein-Truncating Variants across 337,205 Individuals in the UK Biobank Study. Nat. Commun. 2018, 9, 1612. [Google Scholar] [CrossRef] [Green Version]
  221. Chow, J.; Jensen, M.; Amini, H.; Hormozdiari, F.; Penn, O.; Shifman, S.; Girirajan, S.; Hormozdiari, F. Dissecting the Genetic Basis of Comorbid Epilepsy Phenotypes in Neurodevelopmental Disorders. Genome Med. 2019, 11, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Parenti, I.; Rabaneda, L.G.; Schoen, H.; Novarino, G. Neurodevelopmental Disorders: From Genetics to Functional Pathways. Trends Neurosci. 2020, 43, 608–621. [Google Scholar] [CrossRef] [PubMed]
  223. Hammoudi, D.S.; Lee, S.S.F.; Madison, A.; Mirabella, G.; Buncic, J.R.; Logan, W.J.; Snead, O.C.; Westall, C.A. Reduced Visual Function Associated with Infantile Spasms in Children on Vigabatrin Therapy. Investig. Opthalmol. Vis. Sci. 2005, 46, 514. [Google Scholar] [CrossRef] [PubMed]
  224. Holmes, G.L. Effect of Seizures on the Developing Brain and Cognition. Semin. Pediatr. Neurol. 2016, 23, 120–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Huang, C.-C.; Chang, Y.-C. The Long-Term Effects of Febrile Seizures on the Hippocampal Neuronal Plasticity—Clinical and Experimental Evidence. Brain Dev. 2009, 31, 383–387. [Google Scholar] [CrossRef] [PubMed]
  226. Shields, W.D. Catastrophic Epilepsy in Childhood. Epilepsia 2000, 41, S2–S6. [Google Scholar] [CrossRef] [PubMed]
  227. Eldomery, M.K.; Coban-Akdemir, Z.; Harel, T.; Rosenfeld, J.A.; Gambin, T.; Stray-Pedersen, A.; Küry, S.; Mercier, S.; Lessel, D.; Denecke, J.; et al. Lessons Learned from Additional Research Analyses of Unsolved Clinical Exome Cases. Genome Med. 2017, 9, 26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Jezela-Stanek, A.; Ciara, E.; Jurkiewicz, D.; Kucharczyk, M.; Jędrzejowska, M.; Chrzanowska, K.H.; Krajewska-Walasek, M.; Żemojtel, T. The Phenotype-driven Computational Analysis Yields Clinical Diagnosis for Patients with Atypical Manifestations of Known Intellectual Disability Syndromes. Mol. Genet. Genom. Med. 2020, 8, e1263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Rochtus, A.; Olson, H.E.; Smith, L.; Keith, L.G.; El Achkar, C.; Taylor, A.; Mahida, S.; Park, M.; Kelly, M.; Shain, C.; et al. Genetic Diagnoses in Epilepsy: The Impact of Dynamic Exome Analysis in a Pediatric Cohort. Epilepsia 2020, 61, 249–258. [Google Scholar] [CrossRef]
  230. Fokkema, I.F.A.C.; Velde, K.J.; Slofstra, M.K.; Ruivenkamp, C.A.L.; Vogel, M.J.; Pfundt, R.; Blok, M.J.; Lekanne Deprez, R.H.; Waisfisz, Q.; Abbott, K.M.; et al. Dutch Genome Diagnostic Laboratories Accelerated and Improved Variant Interpretation and Increased Accuracy by Sharing Data. Hum. Mutat. 2019, 40, 2230–2238. [Google Scholar] [CrossRef] [PubMed]
  231. McInerney-Leo, A.M.; Duncan, E.L. Massively Parallel Sequencing for Rare Genetic Disorders: Potential and Pitfalls. Front. Endocrinol. 2021, 11, 628946. [Google Scholar] [CrossRef]
  232. Sun, H.; Shen, X.-R.; Fang, Z.-B.; Jiang, Z.-Z.; Wei, X.-J.; Wang, Z.-Y.; Yu, X.-F. Next-Generation Sequencing Technologies and Neurogenetic Diseases. Life 2021, 11, 361. [Google Scholar] [CrossRef] [PubMed]
  233. Vissers, L.E.L.M.; van Nimwegen, K.J.M.; Schieving, J.H.; Kamsteeg, E.-J.; Kleefstra, T.; Yntema, H.G.; Pfundt, R.; van der Wilt, G.J.; Krabbenborg, L.; Brunner, H.G.; et al. A Clinical Utility Study of Exome Sequencing versus Conventional Genetic Testing in Pediatric Neurology. Genet. Med. 2017, 19, 1055–1063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Niguidula, N.; Alamillo, C.; Shahmirzadi Mowlavi, L.; Powis, Z.; Cohen, J.S.; Farwell Hagman, K.D. Clinical Whole-Exome Sequencing Results Impact Medical Management. Mol. Genet. Genom. Med. 2018, 6, 1068–1078. [Google Scholar] [CrossRef] [PubMed]
  235. Tan, T.Y.; Dillon, O.J.; Stark, Z.; Schofield, D.; Alam, K.; Shrestha, R.; Chong, B.; Phelan, D.; Brett, G.R.; Creed, E.; et al. Diagnostic Impact and Cost-Effectiveness of Whole-Exome Sequencing for Ambulant Children With Suspected Monogenic Conditions. JAMA Pediatr. 2017, 171, 855. [Google Scholar] [CrossRef] [PubMed]
  236. D’Gama, A.M.; Walsh, C.A. Somatic Mosaicism and Neurodevelopmental Disease. Nat. Neurosci. 2018, 21, 1504–1514. [Google Scholar] [CrossRef] [PubMed]
  237. Cheroni, C.; Caporale, N.; Testa, G. Autism Spectrum Disorder at the Crossroad between Genes and Environment: Contributions, Convergences, and Interactions in ASD Developmental Pathophysiology. Mol. Autism 2020, 11, 69. [Google Scholar] [CrossRef] [PubMed]
  238. Hertz-Picciotto, I.; Schmidt, R.J.; Krakowiak, P. Understanding Environmental Contributions to Autism: Causal Concepts and the State of Science. Autism Res. 2018, 11, 554–586. [Google Scholar] [CrossRef] [PubMed]
  239. Imbriani, G.; Panico, A.; Grassi, T.; Idolo, A.; Serio, F.; Bagordo, F.; De Filippis, G.; De Giorgi, D.; Antonucci, G.; Piscitelli, P.; et al. Early-Life Exposure to Environmental Air Pollution and Autism Spectrum Disorder: A Review of Available Evidence. Int. J. Environ. Res. Public. Health 2021, 18, 1204. [Google Scholar] [CrossRef] [PubMed]
  240. Masini, E.; Loi, E.; Vega-Benedetti, A.F.; Carta, M.; Doneddu, G.; Fadda, R.; Zavattari, P. An Overview of the Main Genetic, Epigenetic and Environmental Factors Involved in Autism Spectrum Disorder Focusing on Synaptic Activity. Int. J. Mol. Sci. 2020, 21, 8290. [Google Scholar] [CrossRef] [PubMed]
  241. Kruse, S.W.; Suino-Powell, K.; Zhou, X.E.; Kretschman, J.E.; Reynolds, R.; Vonrhein, C.; Xu, Y.; Wang, L.; Tsai, S.Y.; Tsai, M.-J.; et al. Identification of COUP-TFII Orphan Nuclear Receptor as a Retinoic Acid–Activated Receptor. PLoS Biol. 2008, 6, e227. [Google Scholar] [CrossRef] [PubMed]
  242. Rodgers, J.M. Strategies for Protecting Oligodendrocytes and Enhancing Remyelination in Multiple Sclerosis. Discov. Med. 2013, 16, 53–63. [Google Scholar]
  243. Winters, J.J.; Ferguson, C.J.; Lenk, G.M.; Giger-Mateeva, V.I.; Shrager, P.; Meisler, M.H.; Giger, R.J. Congenital CNS Hypomyelination in the Fig4 Null Mouse Is Rescued by Neuronal Expression of the PI(3,5)P2 Phosphatase Fig4. J. Neurosci. 2011, 31, 17736–17751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Barateiro, A.; Fernandes, A. Temporal Oligodendrocyte Lineage Progression: In Vitro Models of Proliferation, Differentiation and Myelination. Biochim. Biophys. Acta BBA Mol. Cell Res. 2014, 1843, 1917–1929. [Google Scholar] [CrossRef] [Green Version]
  245. Chanoumidou, K.; Mozafari, S.; Baron-Van Evercooren, A.; Kuhlmann, T. Stem Cell Derived Oligodendrocytes to Study Myelin Diseases. Glia 2020, 68, 705–720. [Google Scholar] [CrossRef]
  246. Czepiel, M.; Boddeke, E.; Copray, S. Human Oligodendrocytes in Remyelination Research: Human Oligodendrocytes in Remyelination. Glia 2015, 63, 513–530. [Google Scholar] [CrossRef]
  247. Madhavan, M.; Nevin, Z.S.; Shick, H.E.; Garrison, E.; Clarkson-Paredes, C.; Karl, M.; Clayton, B.L.L.; Factor, D.C.; Allan, K.C.; Barbar, L.; et al. Induction of Myelinating Oligodendrocytes in Human Cortical Spheroids. Nat. Methods 2018, 15, 700–706. [Google Scholar] [CrossRef] [PubMed]
  248. Marton, R.M.; Miura, Y.; Sloan, S.A.; Li, Q.; Revah, O.; Levy, R.J.; Huguenard, J.R.; Pașca, S.P. Differentiation and Maturation of Oligodendrocytes in Human Three-Dimensional Neural Cultures. Nat. Neurosci. 2019, 22, 484–491. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representations of NR2F1 gene localization, expression profile and pathogenic point variants. The human NR2F1 gene, located on chromosome 5 (region 5q14–q15), codes for a 3.824 base pairs (bp)-long transcript containing three distinct exons, translated into a 423 amino acid (aa)-long protein (Source: Human hg38 chr5:93583222-93594611 UCSC Genome Browser v427). The NR2F1 expression profile in different tissues and organs is shown as normalized transcript per million (nTPM). Source: human transcriptome dataset at Human Protein Atlas (HPA) (Query: ENSG00000175745-NR2F1). NR2F1 haploinsufficiency in BBOSAS patients is caused by gene deletion or by loss-of-function mutations affecting one allele. Small indels and point variants (black asterisks) tend to fall in the ATG starting codon (1), in the DBD (2) or in the LBD (3). Protein truncations (or frameshift variants followed by truncation at variable distance) are listed with a grey asterisk (4). All variants indicated by asterisks are also listed in the boxes, grouped by gene region or variant type. Whole-gene deletions, not shown here, are listed together with point variants in Table 1.
Figure 1. Schematic representations of NR2F1 gene localization, expression profile and pathogenic point variants. The human NR2F1 gene, located on chromosome 5 (region 5q14–q15), codes for a 3.824 base pairs (bp)-long transcript containing three distinct exons, translated into a 423 amino acid (aa)-long protein (Source: Human hg38 chr5:93583222-93594611 UCSC Genome Browser v427). The NR2F1 expression profile in different tissues and organs is shown as normalized transcript per million (nTPM). Source: human transcriptome dataset at Human Protein Atlas (HPA) (Query: ENSG00000175745-NR2F1). NR2F1 haploinsufficiency in BBOSAS patients is caused by gene deletion or by loss-of-function mutations affecting one allele. Small indels and point variants (black asterisks) tend to fall in the ATG starting codon (1), in the DBD (2) or in the LBD (3). Protein truncations (or frameshift variants followed by truncation at variable distance) are listed with a grey asterisk (4). All variants indicated by asterisks are also listed in the boxes, grouped by gene region or variant type. Whole-gene deletions, not shown here, are listed together with point variants in Table 1.
Cells 11 01260 g001
Figure 2. BBSOAS clinical spectrum. List of main BBSOAS symptoms and features, subdivided by affected systems. While developmental delay, intellectual disability and optic atrophy are the most common features (88%, 85.9% and 66.3% of patients, respectively), other symptoms are less common, such as CVI (44.6%), epilepsy (46.7%), ASD or autistic traits (39.1% and 14.1%, respectively), hearing impairment (11%) and hypotonia (62%). Abbreviations: ADHD, attention deficit hyperactivity disorder; ASD, autism spectrum disorder; CVI, cortical visual impairment; IQ, intelligence quotient.
Figure 2. BBSOAS clinical spectrum. List of main BBSOAS symptoms and features, subdivided by affected systems. While developmental delay, intellectual disability and optic atrophy are the most common features (88%, 85.9% and 66.3% of patients, respectively), other symptoms are less common, such as CVI (44.6%), epilepsy (46.7%), ASD or autistic traits (39.1% and 14.1%, respectively), hearing impairment (11%) and hypotonia (62%). Abbreviations: ADHD, attention deficit hyperactivity disorder; ASD, autism spectrum disorder; CVI, cortical visual impairment; IQ, intelligence quotient.
Cells 11 01260 g002
Figure 3. Overview of structural and functional defects along the visual pathway in BBSOAS patients and corresponding Nr2f1 mutant mouse models. CVI and other visual impairments reported in BBSOAS patients (blue boxes) might build upon structural impairment affecting several structures in the visual system, such as the neural retina (NR), the optic disc (OD), the optic nerve (ON), the primary visual area of the neocortex (V1) and its connections to secondary associative areas. The use of Nr2f1 mutant mouse models (green boxes) have helped in elucidating the molecular, cellular and functional mechanisms that could potentially cause the defects observed in patients.
Figure 3. Overview of structural and functional defects along the visual pathway in BBSOAS patients and corresponding Nr2f1 mutant mouse models. CVI and other visual impairments reported in BBSOAS patients (blue boxes) might build upon structural impairment affecting several structures in the visual system, such as the neural retina (NR), the optic disc (OD), the optic nerve (ON), the primary visual area of the neocortex (V1) and its connections to secondary associative areas. The use of Nr2f1 mutant mouse models (green boxes) have helped in elucidating the molecular, cellular and functional mechanisms that could potentially cause the defects observed in patients.
Cells 11 01260 g003
Figure 4. NR2F1 molecular functioning and dimer formation hypothetical predictions upon distinct genetic perturbations. Distinct deletions and mutations affecting NR2F1 genomic sequence (left column) could result in specific scenarios of impaired quantity and/or quality of NR2F1 homo- and heterodimers (central and right columns), possibly explaining a genotype–phenotype correlation. While decreased NR2F1 expression has been proven for some ATG point variants and for gene deletions, other scenarios—such as the dominant-negative effect of DBD- or LBD-mutated variants—are poorly understood to date. In a similar way, the possible consequences of the production of truncated NR2F1 forms are purely theoretical and will need further experimental assays to be confirmed and characterized. The exact 3D structure of nuclear receptor dimers and their formation rate, the identity of NR2F1 cofactors and the possible toxic effect of frameshift/truncated variants will require further studies. The asterisks indicate point mutations (color-coded following their genetic category), whereas stop symbols (in red and white) represent truncation sites.
Figure 4. NR2F1 molecular functioning and dimer formation hypothetical predictions upon distinct genetic perturbations. Distinct deletions and mutations affecting NR2F1 genomic sequence (left column) could result in specific scenarios of impaired quantity and/or quality of NR2F1 homo- and heterodimers (central and right columns), possibly explaining a genotype–phenotype correlation. While decreased NR2F1 expression has been proven for some ATG point variants and for gene deletions, other scenarios—such as the dominant-negative effect of DBD- or LBD-mutated variants—are poorly understood to date. In a similar way, the possible consequences of the production of truncated NR2F1 forms are purely theoretical and will need further experimental assays to be confirmed and characterized. The exact 3D structure of nuclear receptor dimers and their formation rate, the identity of NR2F1 cofactors and the possible toxic effect of frameshift/truncated variants will require further studies. The asterisks indicate point mutations (color-coded following their genetic category), whereas stop symbols (in red and white) represent truncation sites.
Cells 11 01260 g004
Table 1. Updated list of NR2F1 variants and clinical description of BBSOAS reported patients. BBSOAS patients, identified by their protein variant and—when available—by their LOVD identifier, are listed following the chronological order of reports and publications describing their cases. Main clinical signs include altered brain morphology as observed by MRI, developmental delay (DD), intellectual disability (ID), visual system deficits, early-onset epilepsy and seizures (EOE/S), autism spectrum disorder (ASD) and behavioral abnormalities and hypotonia. The severity index of each patient has been calculated based on the presence/absence of the main clinical signs, ranging from 1 (one symptom category only found in the patient) to a maximum of 7 (all listed clinical signs are present, to different extent, in the same patient). An extended version of these data, with additional columns describing other less common clinical features, is available in Supplementary Table S1. List of references: AK13, Al-Kateb et al., 2013; BA19, Balciuniene et al., 2019; BE18, Bertacchi et al., 2018; BE20, Bertacchi et al., 2020; BR09, Brown et al., 2009; CA09, Cardoso et al., 2009; BO14, Bosch et al., 2014; BO20, Bojanek et al., 2020; CH16, Chen et al., 2016; DI16, Dimassi et al., 2016; EL17, Eldomery et al., 2017; GA21, Gazdagh et al., 2021; HF15, Hino-Fukuyo et al., 2015; HF17, Hino-Fukuyo et al., 2017; HO20, Hobbs et al., 2020; JS20, Jezela-Stanek et al., 2020; JU21, Jurkute, Bertacchi et al., 2021; KA17, Kaiwar et al., 2017; MH18, Martín-Hernández et al., 2018; MI14, Michaud et al., 2014; MI20, Mio et al., 2020; PA19, Park et al., 2019; SA13, Sanders et al., 2013; ST20, Starosta et al., 2020; RE20, Rech et al., 2020; RO20, Rochtus et al., 2020; VI17, Vissers et al., 2017; WA20 Walsh et al., 2020; ZO20, Zou et al., 2020. Abbreviations: CB, cerebellum; CC, corpus callosum; CS, corticospinal tract; D, deletion; DBD, DNA binding domain; DD, developmental delay; DM, delayed myelination; DMD, delayed motor development/poor coordination; DQ, developmental quotient; EOE/S, early-onset epilepsy/seizures; FI/NFI, frameshifting indel/non frameshifting indel; FS, febrile seizures; GCL, ganglionic cell layer; GVI, general visual impairments; HP, hippocampus; HPM, hippocampal malrotation; ID, intellectual disability; IQ, intelligence quotient; IS, infantile spasms; LBD, ligand binding domain; LV, lateral ventricle; LVA, low visual acuity; MCP, macrocephaly; MD, microdeletion; MM, missense mutation; OA, optic atrophy; OC, optic chiasm; OCB/RB, obsessive-compulsive/repetitive behaviors; OD, optic disc; ON, optic nerve; ONH, ON hypoplasia; P/SOD, pale/small optic disc; PDD-NOS, pervasive developmental disorder—not otherwise specified; PH, bilateral periventricular heterotopia; RB, repetitive behavior; RNFL, retinal nerve fiber layer; TIV, translation initiation variants; WGD, whole-gene deletion; WM, white matter. For the extended report, please refer to Supplementary Table S1.
Table 1. Updated list of NR2F1 variants and clinical description of BBSOAS reported patients. BBSOAS patients, identified by their protein variant and—when available—by their LOVD identifier, are listed following the chronological order of reports and publications describing their cases. Main clinical signs include altered brain morphology as observed by MRI, developmental delay (DD), intellectual disability (ID), visual system deficits, early-onset epilepsy and seizures (EOE/S), autism spectrum disorder (ASD) and behavioral abnormalities and hypotonia. The severity index of each patient has been calculated based on the presence/absence of the main clinical signs, ranging from 1 (one symptom category only found in the patient) to a maximum of 7 (all listed clinical signs are present, to different extent, in the same patient). An extended version of these data, with additional columns describing other less common clinical features, is available in Supplementary Table S1. List of references: AK13, Al-Kateb et al., 2013; BA19, Balciuniene et al., 2019; BE18, Bertacchi et al., 2018; BE20, Bertacchi et al., 2020; BR09, Brown et al., 2009; CA09, Cardoso et al., 2009; BO14, Bosch et al., 2014; BO20, Bojanek et al., 2020; CH16, Chen et al., 2016; DI16, Dimassi et al., 2016; EL17, Eldomery et al., 2017; GA21, Gazdagh et al., 2021; HF15, Hino-Fukuyo et al., 2015; HF17, Hino-Fukuyo et al., 2017; HO20, Hobbs et al., 2020; JS20, Jezela-Stanek et al., 2020; JU21, Jurkute, Bertacchi et al., 2021; KA17, Kaiwar et al., 2017; MH18, Martín-Hernández et al., 2018; MI14, Michaud et al., 2014; MI20, Mio et al., 2020; PA19, Park et al., 2019; SA13, Sanders et al., 2013; ST20, Starosta et al., 2020; RE20, Rech et al., 2020; RO20, Rochtus et al., 2020; VI17, Vissers et al., 2017; WA20 Walsh et al., 2020; ZO20, Zou et al., 2020. Abbreviations: CB, cerebellum; CC, corpus callosum; CS, corticospinal tract; D, deletion; DBD, DNA binding domain; DD, developmental delay; DM, delayed myelination; DMD, delayed motor development/poor coordination; DQ, developmental quotient; EOE/S, early-onset epilepsy/seizures; FI/NFI, frameshifting indel/non frameshifting indel; FS, febrile seizures; GCL, ganglionic cell layer; GVI, general visual impairments; HP, hippocampus; HPM, hippocampal malrotation; ID, intellectual disability; IQ, intelligence quotient; IS, infantile spasms; LBD, ligand binding domain; LV, lateral ventricle; LVA, low visual acuity; MCP, macrocephaly; MD, microdeletion; MM, missense mutation; OA, optic atrophy; OC, optic chiasm; OCB/RB, obsessive-compulsive/repetitive behaviors; OD, optic disc; ON, optic nerve; ONH, ON hypoplasia; P/SOD, pale/small optic disc; PDD-NOS, pervasive developmental disorder—not otherwise specified; PH, bilateral periventricular heterotopia; RB, repetitive behavior; RNFL, retinal nerve fiber layer; TIV, translation initiation variants; WGD, whole-gene deletion; WM, white matter. For the extended report, please refer to Supplementary Table S1.
ReferencesLOVD Database ID; Patient IDVariant TypeVariant (Protein)MRI (General; Optic Nerve and Cortical Morphology)DDIDVisual System Defect(s) and Visual DeficitEOE/SASD Behavioral AbnormalitiesHypotoniaSeverity Index
BR09BE18, #1De novo deletion (400–500 Kb MD at breakpoints following paracentric inversion)DeletedCranial nerve abnormalitiesYesNDNDNDNDYes3
CA09BE18, #2De novo deletionDeletedPHYesYes; speech delayColobomaFSNDYes6
CA09BE18, #3De novo deletionDeletedPHYesYes; speech delayNDISNDND4
CA09BE18, #4De novo deletionDeletedPH; HPM; MCP; polymicrogyriaYesYes; speech delayNDYesNDYes5
AK13BE18, #5; RE20, #27De novo deletion (582 Kb)Deleted; (del. includes FLJ42709, FAM172A, POU5F2, and MIR2277)OA (small OC)Yes, DMDNo but speech delayOA; GVINoADHDYes6
SA13BE18, #6; RE20, #53De novo MM in LBDp.Arg404HisNDNDNDNDNDASDND1
BO14LOVD: NR2F1_000001; BO14, #2; BE18, #8; RE20, #13De novo MM in DBDp.Ser113ArgOA (small OD and OC)YesNo or NDOA; P/SOD; CVI; GVINDNDYes4
BO14LOVD: NR2F1_000002; BO14, #1; BE18, #7; RE20, #14De novo MM in DBDp.Arg115PronormalNoYes (IQ 48)OA; P/SOD; small ON; CVI; GVINDNDND2
BO14LOVD: NR2F1_000003; BO14, #3; BE18, #9; RE20, #48De novo MM in LBDp.Leu252ProNDYesYes (IQ 55–65)P/SOD; CVI; GVINDNDYes4
BO14BO14, #4; BE18, #10; RE20, #28Deletion (0.83 Mb)Deleted; (del. includes FAM172A, KIAA0825)NDNoMild (IQ 61–74)P/SOD; CVI; GVINDNDND2
BO14BO14, #5; BE18, #11; RE20, #24De novo deletion (2.83 Mb)Deleted; (del. includes FAM172A, KIAA0825, ANKRD31)Normal or NDYesNo (IQ ND)P/SOD; small ON; CVI; GVINDNDND2
BO14LOVD: NR2F1_000004; BO14, #6; CH16, #12; RE20, #12De novo MM in DBDp.Arg112LysNormal or NDYesYes (IQ 52)OA; P/SOD; mild GVINDASD; OCDND4
HF15; HF17LOVD: NR2F1_000018; CH16, #14; RE20, #18De novo MM in DBDp.Arg135CysNormal or NDYes, DMDYes (DQ < 20); speech delay; non-verbalBilateral OAWest Syndrome; IS and FS; generalized tonic seizuresASD traitsND5
DI16LOVD: NR2F1_000057; CH16, #15; RE20, #11De novo in-frame deletion in DBDp.Phe110delCC thinning; LV asymmetry; septum pellucidum agenesis; abnormal gyrationYesSevere; non-verbalNo or NDWest Syndrome; spasms at 3 mo; IS; hypsarrythmic EEG; electroclinical spasmsASDGlobal6
CH16LOVD: NR2F1_000041; CH16, #1; BE18, #16; RE20, #16De novo MM in DBDp.Cys128ArgCC thinning; WM reduction; leukodystrophyYes, DMDYes (IQ ND); speech delay; non-verbalOA; GVIEpilepsy with staring spells and generalized tonic-clonic seizuresASD; self-injurious behavior; head-bangingYes7
MI14; CH16LOVD: NR2F1_000075; CH16, #2; BE18, #17; RE20, #17De novo MM in DBDp.Arg135SerCC thinning; WM reduction; ON malformation and OC bilateral hypoplasiaYes, DMDYes (IQ ND); speech delayMild OA; ONH; CVI; GVIIS; Occipital lobe epilepsyASD; head-bangingProfound, axial and appendicular7
CH16; EL17LOVD: NR2F1_000007; CH16, #3; BE18, #18; RE20, #19De novo MM in DBDp.Cys138TyrWM reductionGlobalYesOA; GVIFS; Abnormal EEG during sleepASD; RB (persistent head-banging)No6
CH16LOVD: NR2F1_000058; CH16, #4; BE18, #19; RE20, #21De novo MM in DBDp.Arg142LeuCC thinning at 7 moYes, DMDYes (IQ ND); speech delay; non-verbalOA; P/SOD; small ON; CVI; GVIIS; atonic seizures with markedly abnormal EEGNDYes6
CH16LOVD: NR2F1_000010; CH16, #5; BE18 #20; RE20 #22De novo MM in DBDp.Cys146ArgCC thinning and septo-optic dysplasiaYes, DMDYes (IQ ND); speech delay; non-verbalOA; GVINoASD traits; RB (self-stimulatory behaviors)Yes6
CH16LOVD: NR2F1_000024; CH16, #6; BE18, #21; RE20, #47;De novo MM in DBDp.Ala155ThrNormal or NDNoMild speech delay (pronunciation; dysarthria)No or NDNoNO or NDYes2
CH16LOVD: NR2F1_000078; CH16, #7; BE18, #22; RE20, #51De novo MM in LBDp.Gly368AspNormal or NDYesYes (IQ ND); speech delayNo or NDFirst generalized seizure at 18 yoASD; RB; aggressive behaviorNo4
CH16LOVD: NR2F1_000057; CH16, #8; BE18, #23; RE20, #10De novo in-frame deletion in DBDp.Phe110delCC thinningYes, DMDYes (IQ ND); speech delay; non-verbalOA; ONH; CVI; GVIISNo or NDYes6
CH16LOVD: NR2F1_000053; CH16, #9; CH16, #24; RE20, #40De novo frameshift truncationp.Gly35Argfs*361Normal or NDYes, DMDYes (IQ 55–69; verbal IQ 35–40 at 16 yo); speech delayOA; P/SOD; CVI; GVIFew seizures at 3–4 yoASD; RB including PDD-NOS at 22 yo; ADHDYes6
CH16LOVD: NR2F1_000055; CH16, #10; BE18, #25; RE20, #41De novo frameshift truncationp.His97Hisfs*22NDYesYes (IQ ND); speech delayOA; ONH; GVINoASD; OCB/RB; ADHDYes5
CH16LOVD: NR2F1_000052; CH16, #11; BE18, #26; RE20, #36De novo TIVp?Normal or NDYes, DMDYes (IQ ND); speech delay; non-verbalOA; ONH; GVINoASD; RB (head banging)Yes5
CH16LOVD: NR2F1_000052; CH16, #12; BE18, #27; #37 in RE20, #37De novo TIVp?CC thinning; ONHYes, DMDYes (IQ ND); speech delay; non-verbalOA; ONH; P/SOD; CVI; GVIYesASD traitsYes7
CH16LOVD: NR2F1_000048; CH16, #13; BE18, #28; RE20, #35De novo TIVp.M1?NDYes, DMDYes (FSIQ in the 40 s); speech delayOA; ONH; CVI; GVINoASD; head bangingSubstantial, central5
CH16LOVD: NR2F1_000048; CH16, #14; BE18, #29; RE20, #38De novo TIVp.M1?CC and CS thinning; pyramidal decussation agenesis; right vs. left fiber directionality asymmetryYes, DMDYes (IQ ND); speech delay; non-verbalOA; coloboma; ONH; GVISeizure at 3 yo, complex partial, left parietalOCB (hand stereotypes); ADHDYes7
CH16LOVD: NR2F1_000042; CH16, #15; BE18, #30; RE20, #39De novo TIVp?Cerebral malformations; bilateral HPMYes, DMDYes (IQ ND); speech delay; non-verbalOA; GVITonic-clonic seizures at 13 and 18 yoOCBYes7
CH16CH16, #16; BE18, #31; RE20, #29De novo deletion (0.2 Mb)Deleted; (del. includes FAM172A, partial)NDYes, DMDYes (IQ ND); speech delayOA; P/SOD; pigmented maculae; GVINoADHDYes5
CH16CH16, #17; BE18, #32; RE20, #31Deletion (0.9 Mb)Deleted; (del. includes FAM172A; KIAA0825, partial)NDYesYes (IQ ND); speech delayOA; P/SOD; GVINoASD; ADHDNo4
CH16CH16, #18; BE18, #33; RE20, #32Parental (son of CH16, #17), deletion (0.9 Mb)DeletedCC agenesis; DM of the EC and IC anterior limb; focal abnormality of the right CBYes, DMDYes (IQ ND); speech delayOA; P/SOD; GVINoASD traits; ADHDAxial6
CH16CH16, #19; BE18, #34; RE20, #30Deletion (1.2 Mb)Deleted; (del. Includes FAM172A, KIAA0825, ANKRD31)NDYesYes (verbal IQ 96; non-verbal IQ 70)OA; GVINoASD; PDD-NOSNo4
CH16CH16, #20; BE18, #35; RE20, #23Deletion (5.0 Mb)Deleted; (del. Includes FAM172A, KIAA0825, ANKRD32, MCTP30)MCPYesYes (IQ ND)No or NDNoNDLow muscle tone, normal mass and strength4
KA17LOVD: NR2F1_000039; BE18, #36; RE20, #1De novo MM in DBDp.Cys86PheCC thinning; WM reduction; MCPYes, DMDYes (DQ < 25 at 14 yo); speech delay; non-verbalOA; mild bilateral ONH; CVI; GVIOne episode of IS; left occipital onset seizure in EEG; FSSevere ASD; RB (self-stimulating, self-injurious behavior); limited social interactionYes7
KA17LOVD: NR2F1_000079; BE18, #37; RE20, #52De novo MM in LBDp.Leu372ProNDYes, DMDYes (IQ ND); speech delayOA; GVINDRB; ADHDYes5
VI17LOVD: NR2F1_000017MM in DBDp.Gly105AspCerebral malformationsNDYes (IQ ND)No or NDNDNDND2
MH18LOVD: NR2F1_000040; BE18, #38; RE20, #5De novo MM in DBDp.Lys96GluCC thinningYesYes (IQ ca. 30–50); speech delayMild OA; CVI; GVINoNDYes5
PA19LOVD: NR2F1_000038; RE20, #45; JU21, #10Truncationp.Tyr171*CC thinningYesYes; mild (IQ ca. 77–80)OA; GVINoBehavioral disorders; ADHDND5
BO20LOVD: NR2F1_000037; RE20, #43De novo truncationp.Gln28*ONHNoNo (verbal IQ 141; nonverbal IQ 63)OA; ONH; CVI; mild GVINo; EEG showed rare isolated sharp waves from central regionsASD (hand flapping and toe walking at 24 mo); behavioral disorders; ADHDYes5
BE20LOVD: NR2F1_000059; BE20, #1De novo MM in DBDp.Arg142HisCC thinning; OA (OC and nerve thinning); abnormal gyrationYesYesOA; amblyopiaIS at 8 moASD and ADHD traitsYes7
BE20LOVD: NR2F1_000060; BE20, #2De novo truncationp.Gln244*CC thinning; ventricular asymmetry and enlargement; abnormal gyration; polymicrogyriaYesYesNo or NDNDBehavioral disordersYes5
BE20LOVD: NR2F1_000054; BE20, #3; JU21, #4De novo truncationp.Glu39*CC and OC thinning; CB malformation; ectopic nodular heterotopy; abnormal gyrationYesYes (speech difficulties)Severe bilateral OA; LVA3–4 ES/yStereotypical movements; RB; ADHDYes7
BE20LOVD: NR2F1_000048; BE20, #4De novo TIVp?CC thinning; cortical malformation; abnormal gyrationYesYes (speech difficulties)OANDASD and ADHD traits; behavioral disordersNo5
BE20LOVD: NR2F1_000056; BE20, #5De novo MM in DBDp.Tyr98HisCC thinning; OC hypoplasia; abnormal gyrationYesYesOANDASD traits; behavioral disorders; stereotypical movementsYes6
BE20LOVD: NR2F1_000061; BE20, #6De novo frameshift truncationp.Lys323Serfs*73Short CC; ON and chiasm thinning; hypoplastic olfactory lobes; abnormal gyrationYes, DMDYes (speech difficulties)OA; LVANDASD traitsNo5
ZO20LOVD: NR2F1_000085De novo truncationp.Ser201*NDMild/moderateMild/moderateBilateral P/SOD; LVANDNDND2
HO20LOVD: NR2F1_000084; RE20, #44De novo truncationp.Glu85*Normal or NDYesYes (IQ 69)OA; GVISpells of behavioral arrest and non-responsivenessASD; auditory hallucinations and delusions; crying episodesYes6
WA20LOVD: NR2F1_000051Frameshift truncationp.Asn362fs*33CC; ON and OC hypoplasia; mild MCPApparent at 8 moSpeech delaySevere GVIMyoclonic astatic seizures at 2½ yoASDND6
MI20LOVD: NR2F1_000034; MI20, #1De novo MM in DBDp.Gly105SerBenign enlargement of the subarachnoid spaces (BESS)Yes, DMDSpeech delay; non-verbal until 2 yoBilateral OA; GVIMyoclonic epilepsy diagnosed at 3 yoRBND6
MI20LOVD: NR2F1_000034; MI20, #2De novo MM in DBDp.Gly105SerLV enlargement; intraventricular arachnoid cystYes, DMDSpeech delay; non-verbal until 2 yoBilateral OA; GVIMyoclonic epilepsy diagnosed at 4 yoRBND6
ST20LOVD: NR2F1_000035MM in DBDp.Lys107GluCC; ON; OC and optic tracts atrophy; complex pituitary cyst.Marked and global; DMDNDDeclining visual acuity; legally blind by 10 yo1–3 yo + 30 episodes of FS; occasionally with myoclonus and generalized seizuresAggressive behavior; depression; hallucinationsYes6
RE20LOVD: NR2F1_000065; RE20, #33De novo TIVp.M1?NDNoNDNo or NDNoNDYes1
RE20LOVD: NR2F1_000048; RE20, #34De novo TIVp.M1?NDYes, DMDSpeech delay; non-verbalOA; small ON; CVI; GVIFSASDYes5
RE20LOVD: NR2F1_000067; RE20, #2De novo MM in DBDp.Cys86ArgCC thinningYesSpeech delay; non-verbalOA; CVI; GVIISASD traitsYes7
RE20LOVD: NR2F1_000068; RE20, #3; RO20, #170De novo MM in DBDp.Val88MetNormal or NDYesSpeech delay; non-verbalOA; CVI; GVIOnset at 9 wo; IS; focal and partial seizures; myoclonic jerksASD; RB (head banging)Yes6
RE20LOVD: NR2F1_000009; RE20, #4MM in DBDp.Gly95ValNDYes, DMDYes (IQ 56); speech delayP/SOD; small ON; CVI; GVIIS and absence seizuresASD traitsYes6
RE20LOVD: NR2F1_000069; RE20, #6MM in DBDp.Hys97ProSlightly decreased brain volumeYesYes (IQ ND); speech delay; non-verbalOA; CVI; GVIMyoclonic seizuresASD (severe)Yes7
RE20LOVD: NR2F1_000070; RE20, #7De novo MM in DBDp.Tyr98CysAbnormalYesSpeech delayP/SOD; ONH; CVI; GVIMyoclonic; absence seizuresASD; RB (head banging); ADHDYes7
RE20LOVD: NR2F1_000071; RE20, #8De novo MM in DBDp.Glu104GlyNDYesSpeech delay; non-verbalOA; CVI; GVINoASD traitsYes5
RE20LOVD: NR2F1_000072; RE20, #9MM in DBDp.Ser108IleON thinning and small OCYesYes (IQ ND); speech delay; non-verbalOA; small ON; CVI; GVINoASD traitsYes6
RE20LOVD: NR2F1_000073; RE20, #15MM in DBDp.Cys122SerNDYes, DMDSpeech delay; non-verbalOA; GVIISASD traitsYes5
RE20LOVD: NR2F1_000074; RE20, #42frameshift truncationp.Asn127Lysfs*270NDYes, DMDSpeech delayONH; CVI; GVIYesASD; auditory hallucinationsYes6
RE20LOVD: NR2F1_000076; RE20, #20De novo MM in DBDp.Gln139HisCC thinning; DM; ON thinning and small OCYes, DMDSpeech delayOA; CVI; GVINoASDYes6
RE20LOVD: NR2F1_000077; RE20, #49MM in LBDp.Ala311ProNormal or NDYesMild (FSIQ 80 below average); speech delayP/SOD; mild GVIGeneralized Myoclonic and absence seizuresASDYes6
RE20 De novo MM in LBDp.Glu318AspAbnormalNo but mild DMDNo (IQ 94; performance IQ 54)OA; CVI; GVIAtonic; Rolandic epilepsyASDNo4
RE20LOVD: NR2F1_000021; RE20, #46De novo truncationp.Arg373*CC; ON and OC thinningYes, DMDMild (DQ ca. 60–70); speech delayP/SOD; ONH; CVI; GVINoASDYes6
RE20LOVD: NR2F1_000036; RE20, #54De novo MM in LBDp.Met406ThrSmall ON; Abnormal MRIYes, DMDYes (IQ ND); speech delayCVI; GVINoASDYes6
RE20LOVD: NR2F1_000063; RE20, #25Maternal, deletion (2.5 Mb)Deleted; (del. includes FLJ42709, FAM172A, MIR2277, POU5F2, KIAA0825, MIR1974, ANKRD32, MCTP1, FAM81B, TTC37)NDYesYes (IQ ND); speech delayOA; CVI; GVIAbsence and tonic seizuresASD; PDD-NOS; OCD; pacing and hittingYes5
RE20LOVD: NR2F1_000064; RE20, #26Deletion (0.97 Mb)DeletedNDYesSpeech delayOA; small ON; CVI; GVINoASDMild5
JU21JU21, #1Frameshift truncationp.Ala2Glnfs*3CC thinning; abnormal gyrationYes; mildNDOA; ONH; LVAOccasional epileptic-like state during light sleep; FS at 4 yo; convulsions at 8 yoASD; ADHDHypotonia and hyperlaxity6
JU21JU21, #2Frameshift truncationp.Asn24Glyfs*379Normal or NDYesYes (IQ ND); speech difficulties; learning disabilityOA; ONH; CVI; LVANoNDND3
JU21JU21, #3AA duplicationp.Arg31dupNormal or NDYesYes (IQ ND); learning disabilityOA; CVI; LVANoNDND3
JU21LOVD: NR2F1_000069; JU21, #5De novo MM in DBDp.Hys97ProCC; ON and OC thinning; periventricular leukomalacia; MCPYes; global (delayed visual maturation)Yes (IQ ND); learning disabilityOA; LVAOne episode of FSASDModerate/severe 7
JU21JU21, #6De novo truncationp.Leu118*CC mild foreshorteningYes; globalYes (IQ ND); learning disabilityOA; LVANoNDYes5
JU21JU21, #7(Likely) de novo truncationp.Tyr120*WM abnormalitiesYes; pervasive; global apraxiaYes (IQ ND); speech delay; learning disabilityOA; LVAMyoclonic epilepsy; focal impaired awareness seizuresASD; ADHDND6
BA19; JU21LOVD: NR2F1_000086; JU21, #8De novo MM in DBDp.Cys122TrpWM reduction; CC thinningYes; globalYes (IQ ND); learning disabilityOA; CVI; LVAIS; Myoclonic epilepsyASD; ADHDYes7
JU21JU21, #9De novo MM in DBDp.Ala155ThrWM reduction; ON thinningYesYes (IQ ND); speech delay; learning disabilityCVI; LVANoASD; anxiety; limited attention spanND5
JU21JU21, #11De novo MM in LBDp.Thr200ArgLateral and third ventricles enlargement; MCPGlobalYes (IQ ND); learning disabilityCentral, steady, maintainedNoNDYes5
JU21JU21, #12De novo truncationp.Trp233*CC; ON and OC thinning; WM delayed maturation; brain abnormalitiesYesYes (IQ ND); speech difficultiesOA; microphthalmia; small ON head; CVINoLimited concentration and short attention spanND5
JU21LOVD: NR2F1_000082; JU21, #13De novo MM in LBDp.Glu342LysNormal CC e ON; OC atrophy and defective rotation; Normal gyrationNoNoOA; ONH; LVANoNDND2
JU21JU21, #14(Likely) de novo deletion in LBDp.Glu346_Gln349delNDYes; walking delayYes (IQ ND); speech delay; dyslexia; learning disabilityONH; CVI; LVANoNDND3
JU21LOVD: NR2F1_000079; JU21, #15Familial MM in LBDp.Leu372ProNDYes; walking delayYes (IQ ND); speech delaySmall ON head; CVI; LVANoNDND3
JU21LOVD: NR2F1_000079; JU21, #16Familial MM in LBDp.Leu372ProNormal or NDYes; walking delayYes (IQ ND); speech delay;OA; ONH; CVI; LVAOne episode of FSNDND4
JU21LOVD: NR2F1_000079; JU21, #17Familial MM in LBDp.Leu372ProNormal or NDYes; walking delayYes (IQ ND); speech delay;OA; CVI; LVANoNDND3
JU21JU21, #18Familial deletion in LBDp.Arg373_Leu374delNDNoNDOA; ONH; LVANoNDND1
JU21JU21, #19Familial deletion in LBDp.Arg373_Leu374delON atrophyNoNDOA; ONH; LVANoNDND2
JU21JU21, #20De novo MM in LBDp.Gly395SerON atrophy; WM lossYesYes (IQ ND); learning disabilityCVI; LVANoNDGeneralized 5
JU21LOVD: NR2F1_000083; JU21, #21De novo truncationp.Glu400*CC thinning; ON and OC atrophy; abnormal gyrationYesYes (IQ ND); learning disabilityOA; ONH; LVANoASD; behavioral disordersND5
JU21JU21, #22De novo whole-gene deletion (599 Kb)deleted; (del. includes FAM172A; NR2F1-AS1, partial; KIAA0825, last exon)CC thickening; ON atrophy; cerebral vascular system abnormalitiesYesYes (IQ ND); mild speech delay; learning disabilityOA; LVANoNDND4
JS20LOVD: NR2F1_000036De novo MM in LBDp.Met406ThrDMYesSevere (IQ ND); speech delay; non-verbalOA; suspected ON dysplasia; GVISeizures from 4 moShort attention spanND6
GA21LOVD: NR2F1_000048De novo TIVp.M1?Brain abnormalities; ON; OC and optic tract hypoplasiaYesSevere (IQ ND)Right iris and chorioretinal coloboma; small ON; bilateral P/SOD; LVAEEG at 12 yo showed possible occipital seizuresNDYes6
Table 2. List of BBSOAS main features. Table resuming the main clinical features of BBSOAS patients with NR2F1 haploinsufficiency; some of them are often present at birth (congenital), including hypotonia, nystagmus and oromotor dysfunction. The clinical features of BBSOAS are variable, and not every individual necessarily manifests all features. Further, the severity of the condition varies from one individual to the next. Abbreviations: ASD, autism spectrum disorder; ID, intellectual disability; IQ, intelligence quotient; MRI, magnetic resonance imaging.
Table 2. List of BBSOAS main features. Table resuming the main clinical features of BBSOAS patients with NR2F1 haploinsufficiency; some of them are often present at birth (congenital), including hypotonia, nystagmus and oromotor dysfunction. The clinical features of BBSOAS are variable, and not every individual necessarily manifests all features. Further, the severity of the condition varies from one individual to the next. Abbreviations: ASD, autism spectrum disorder; ID, intellectual disability; IQ, intelligence quotient; MRI, magnetic resonance imaging.
BBSOAS Main Feature(s)Clinical Description(s)
Developmental delay (DD)Delay in reaching language, social or motor skills milestones
Intellectual disability (ID)Significantly reduced ability to understand new or complex information and to learn and apply new skills (impaired intelligence).
IQ ranging from profound ID with IQ < 20, to moderate (35 < IQ < 49) or mild ID (50 < IQ < 69)
Visual impairmentOptic nerve abnormalities and/or brain-based vision impairment:
Optic nerve atrophy or pallor
Optic nerve hypoplasia
Cortical visual impairment (difficulty locating objects in a crowded field and following rapidly moving images and scenes).
Alacrima (abnormal amount of reflex tearing)
Manifest latent nystagmus and poor tracking; congenital
Significant refractive errors
Amblyopia
HypotoniaLow muscle tone; congenital
Oromotor dysfunctionSwallowing, sucking and chewing problems; congenital
Repetitive behaviorHand flapping, head banging and more
Autism spectrum disorder (ASD)ASD or autistic traits
SeizuresInfantile and/or febrile; occipital seizures
Attention-deficit hyperactivity disorder (ADHD)Inattention, impulsivity and hyperactivity
Hearing impairmentAbnormal hearing
Dysmorphic facial featuresMild and inconsistent
Thin corpus callosum and neocortical dysgyriaHypoplasia of the corpus callosum and abnormal pattern of cortical convolutions and sulci (dysgyria in temporal and parietal areas) on brain MRI
Table 3. Clinical exams recommended for children with BBSOAS. Table showing the main clinical tests for patients with NR2F1 haploinsufficiency. Abbreviations: MRI, magnetic resonance imaging; OCT, optical coherence tomography; RNFL, retinal nerve fiber layer.
Table 3. Clinical exams recommended for children with BBSOAS. Table showing the main clinical tests for patients with NR2F1 haploinsufficiency. Abbreviations: MRI, magnetic resonance imaging; OCT, optical coherence tomography; RNFL, retinal nerve fiber layer.
Recommended Clinical Exam(s)Exam Description
Developmental assessmentIdentify areas of impairment and allow for early intervention.
Psychological evaluation for autismADI-R (Autism Diagnostic Interview, Revised) and ADOS (Autism Diagnostic Observation Schedule) testing performed by a certified clinical psychologist.
Brain MRIRecommended at age three years or older.
Full, dilated eye examinationFundus photography and OCT scan of RNFL to document optic nerve health performed by an ophthalmologist every two years.
Visual acuity testsAs appropriate for patient’s age and understanding
Full hearing evaluationEvery two years
Table 4. Recommended therapeutic approaches. Table resuming the recommended therapeutic approaches for BBSOAS patients. Abbreviations: ASD, autism spectrum disorder; CVI, cortical visual impairment.
Table 4. Recommended therapeutic approaches. Table resuming the recommended therapeutic approaches for BBSOAS patients. Abbreviations: ASD, autism spectrum disorder; CVI, cortical visual impairment.
Recommended Therapeutic Approach(es)Therapy Description
Visual therapyFocused on CVI
Physical therapyAiming to increase strength and to improve gross motor skills.
Occupational therapyAiming to improve fine motor skills and coordination.
Speech therapyConsideration of sign language and alternative communication devices
ABA (Applied behavioral analysis) therapyIf ASD is diagnosed
Anti-convulsive treatmentIf epilepsy is present
Table 5. Symptom prevalence: overall and by variant type. Table shows the prevalence of specific morphological and pathological features (lines) in BBSOAS patients, as a whole (first column), or in distinct BBSOAS genetic categories (second to seventh columns), as indicated. The prevalence of each symptom is calculated as a percentage of positive cases above the total number of patients for that category (numbers of patients are indicated in parenthesis). The color code refers to the severity of prevalence, from higher incidence (red) to lower prevalence (blue). Abbreviations: DBD, DNA binding domain, LBD, ligand binding domain; TIVs, translation initiation variants; CC, corpus callosum; DD, developmental delay; DMD, delayed motor development; ID, intellectual disability; ASD, autism spectrum disorder; ADHD, attention deficit hyperactivity disorder; CVI, cortical visual impairment; OA, optic atrophy; ONH, optic nerve hypoplasia; P/SOD, pallid or small optic disc.
Table 5. Symptom prevalence: overall and by variant type. Table shows the prevalence of specific morphological and pathological features (lines) in BBSOAS patients, as a whole (first column), or in distinct BBSOAS genetic categories (second to seventh columns), as indicated. The prevalence of each symptom is calculated as a percentage of positive cases above the total number of patients for that category (numbers of patients are indicated in parenthesis). The color code refers to the severity of prevalence, from higher incidence (red) to lower prevalence (blue). Abbreviations: DBD, DNA binding domain, LBD, ligand binding domain; TIVs, translation initiation variants; CC, corpus callosum; DD, developmental delay; DMD, delayed motor development; ID, intellectual disability; ASD, autism spectrum disorder; ADHD, attention deficit hyperactivity disorder; CVI, cortical visual impairment; OA, optic atrophy; ONH, optic nerve hypoplasia; P/SOD, pallid or small optic disc.
Overall prevalence
(N = 92)
Variants in the DBD
(N = 32)
Variants in the LBD
(N = 17)
Deletions
(N = 15)
TIV
(N = 9)
Truncations
(N = 11)
Frameshift
(N = 7)
Average severity index4.945.623.764.335.335.185.29
Phenotypic
feature
Morphology
Myelin defects14.13% (13/92)25.00% (8/32)11.76% (2/17)6.67% (1/15)0.00% (0/9)18.18% (2/11)0.00% (0/7)
CC malformations32.61% (30/92)46.88% (15/32)0.00% (0/17)13.33% (2/15)33.33% (3/9)63.64% (7/11)42.86% (3/7)
Development and
behavior
DD88.04% (81/92)90.62% (29/32)70.59% (12/17)93.33% (14/15)88.89% (8/9)90.91% (10/11)100% (7/7)
DMD30.43% (28/92)40.63% (13/32)11.76% (2/17)20.00% (3/15)66.67% 6/9)9.09% (1/11)42.86% (3/7)
ID/speech delay86.95% (80/92)93.75% (30/32)70.59% (12/17)86.67% (13/15)88.89% (8/9)90.91% (10/11)85.71% (6/7)
ASD38.04% (32/92)40.63% (13/32)29.41% (5/17)26.67% (4/15)33.33% (3/9)45.45% (5/11)71.43% (5/7)
ASD-like traits14.13% (13/92)28.13% (9/32)0.00% (0/17)6.67% (1/15)22.22% (2/9)0.00% (0/11)14.29% (1/7)
ADHD18.48% (17/92)9.38% (3/32)5.88% (1/17)26.67% (4/15)22.22% (2/9)36.36% (4/11)42.86% (3/7)
Visual system
CVI42.39% (39/92)53.13% (17/32)47.06% (8/17)26.67% (4/15)33.33% (3/9)27.27% (3/11)42.86% (3/7)
OA67.39% (62/92)78.13% (25/32)47.06% (8/17)53.33% (8/15)77.78% (7/15)72.73% (8/11)71.43% (5/7)
ONH21.74% (20/92)12.50% (4/32)29.41% (5/17)0.00% (0/15)44.44% (4/9)27.27% (3/11)57.14% (4/7)
P/SOD19.56% (18/92)18.75% (6/32)11.76% (2/17)33.33% (5/15)22.22% (2/9)18.18% (2/11)14.29% (1/7)
Others
Epilepsy45.65% (42/92)62.50% (20/32)29.41% (5/17)26.67% (4/15)55.56% (5/9)45.45% (5/11)42.86% (3/7)
Hypotonia61.96% (57/92)75.00% (24/32)35.29% (6/17)60.00% (9/15)88.89% (8/9)54.55% (6/11)57.14% (4/7)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bertacchi, M.; Tocco, C.; Schaaf, C.P.; Studer, M. Pathophysiological Heterogeneity of the BBSOA Neurodevelopmental Syndrome. Cells 2022, 11, 1260. https://doi.org/10.3390/cells11081260

AMA Style

Bertacchi M, Tocco C, Schaaf CP, Studer M. Pathophysiological Heterogeneity of the BBSOA Neurodevelopmental Syndrome. Cells. 2022; 11(8):1260. https://doi.org/10.3390/cells11081260

Chicago/Turabian Style

Bertacchi, Michele, Chiara Tocco, Christian P. Schaaf, and Michèle Studer. 2022. "Pathophysiological Heterogeneity of the BBSOA Neurodevelopmental Syndrome" Cells 11, no. 8: 1260. https://doi.org/10.3390/cells11081260

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop