Next Article in Journal
Single-Cell RNA Sequencing: Opportunities and Challenges for Studies on Corneal Biology in Health and Disease
Previous Article in Journal
Soluble Collectin 11 (CL-11) Acts as an Immunosuppressive Molecule Potentially Used by Stem Cell-Derived Retinal Epithelial Cells to Modulate T Cell Response
Previous Article in Special Issue
Congenital Microcephaly: A Debate on Diagnostic Challenges and Etiological Paradigm of the Shift from Isolated/Non-Syndromic to Syndromic Microcephaly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genetic Primary Microcephalies: When Centrosome Dysfunction Dictates Brain and Body Size

1
UMR144, Institut Curie, 75005 Paris, France
2
Inserm UMR-S 1163, Institut Imagine, 75015 Paris, France
3
Service de Neurologie Pédiatrique, DMU INOV-RDB, APHP, Hôpital Robert Debré, 75019 Paris, France
4
Service de Neurologie Pédiatrique, DMU MICADO, APHP, Hôpital Necker Enfants Malades, 75015 Paris, France
5
Université Paris Cité, Inserm UMR-S 1163, Institut Imagine, 75015 Paris, France
6
Université Paris Cité, Inserm UMR 1141, NeuroDiderot, 75019 Paris, France
*
Author to whom correspondence should be addressed.
Cells 2023, 12(13), 1807; https://doi.org/10.3390/cells12131807
Submission received: 6 April 2023 / Revised: 4 June 2023 / Accepted: 13 June 2023 / Published: 7 July 2023
(This article belongs to the Special Issue Cellular Mechanisms of Microcephaly)

Abstract

:
Primary microcephalies (PMs) are defects in brain growth that are detectable at or before birth and are responsible for neurodevelopmental disorders. Most are caused by biallelic or, more rarely, dominant mutations in one of the likely hundreds of genes encoding PM proteins, i.e., ubiquitous centrosome or microtubule-associated proteins required for the division of neural progenitor cells in the embryonic brain. Here, we provide an overview of the different types of PMs, i.e., isolated PMs with or without malformations of cortical development and PMs associated with short stature (microcephalic dwarfism) or sensorineural disorders. We present an overview of the genetic, developmental, neurological, and cognitive aspects characterizing the most representative PMs. The analysis of phenotypic similarities and differences among patients has led scientists to elucidate the roles of these PM proteins in humans. Phenotypic similarities indicate possible redundant functions of a few of these proteins, such as ASPM and WDR62, which play roles only in determining brain size and structure. However, the protein pericentrin (PCNT) is equally required for determining brain and body size. Other PM proteins perform both functions, albeit to different degrees. Finally, by comparing phenotypes, we considered the interrelationships among these proteins.

1. Introduction

Microcephaly is a brain growth defect responsible for a spectrum of neurodevelopmental disorders [1]. Microcephaly identified before birth is congenital and called primary microcephaly (PM), whereas secondary microcephaly is postnatal. This old group of diseases was known only to child neurologists and neuropathologists who had no other etiology to propose at the other time than in utero maternofetal infections to explain this disorder; however, this topic attracted the attention of researchers in the late 1990s–early 2000s, when it was found that this disorder has a genetic origin as well.
The first mutations identified in “microcephaly genes” are responsible for causing congenital, isolated, or primary microcephaly with recessive inheritance, and are termed microcephaly primary hereditary, or MCPH (MCPH1, OMIM # 251200 (MCPH1 [2]), MCPH2, OMIM # 604317 (WDR62 [3]), MCPH3, OMIM # 604804 (CDK5RAP2 [4,5]), MCPH4, OMIM # 604321 (CASC5 [5]) and MCPH5, OMIM #608716 (ASPM [6])). The original OMIM classification continues to reference newly identified PM genes included under the term “MCPH” (30 “MCPHs genes”) until now. However, the phrase “genetic PMs” gained precedence over “MCPH” as new-generation sequencing approaches demonstrated that variants in the same gene could cause several diseases, such as: (i) MCPH and microcephalic primordial dwarfism (MPD) [4,7,8,9], (ii) MCPH and malformations of cortical development (MCDs) [10,11], and (iii) MCPH and neurosensory disorders [4,12]. Considering genotypes and phenotypes together, all these defects, owing mainly to their recessive inheritance, were grouped under the term “genetic PMs”.
Information related to these microcephaly-causing genes and their functions has opened new research avenues in the genetics of species evolution, neuroscience, and developmental biology. This information has helped elucidate a critical step in brain growth, i.e., embryonic neurogenesis [13,14,15]. Significant advances in cell biology approaches have made it possible to identify the mechanisms underlying the origin of PMs. Over the last 20 years, many teams have contributed to deciphering the proteins and organelles involved in cell division and proliferation (see recent reviews [16,17,18,19,20,21,22]). Excellent reviews have summarized the impact of the defects of one of the thousands of proteins that constitute the mitotic spindle on cell division and survival [23,24,25,26,27]).
The biology of PM is a field that involves crosstalk among four fields of expertise, i.e., genetics, neurology, neurosciences, and cell biology. Therefore, the study of PM biology and treatment has been approached from one of those angles. Here, we provide an overview of the current knowledge of neurological and genetic approaches in the microcephaly field to understand this fascinating issue. Although they belong to the same system, that is the mitotic spindle pole and furrow, why does the absence or presence of an abnormal centrosome or spindle protein have different phenotypic consequences in vivo? By reviewing recent studies on the cellular and molecular mechanisms of the most iconic PMs, we will discuss those that only affect brain growth or brain and body growth in humans and those that most impact the cognition of affected individuals.

2. Primary Microcephaly: Small Brain Size or Small Brain and Body Sizes?

2.1. Primary Microcephaly: An Early Defect in Brain Growth with or without Cortical Malformations

Microcephaly is a frequent defect in brain growth that affects 2–3% of the population worldwide [28]. It is defined by physicians based on occipitofrontal head circumference (OFC) below the normal range (<−2 standard deviation (SD)) based on age and sex. Congenital microcephaly or PM is detected before or at least at birth and is considered severe when the deviation from the mean normal OFC is below –3 SD. The prevalence of severe PM is 0.5–1 per 1000 live births. The defect is often detected in the second trimester of pregnancy using high-performance ultrasound, indicating that the slowdown of brain growth occurs early in the first trimester at the peak of neurogenesis in humans. In addition, intrauterine growth retardation may be associated with microcephaly at birth. However, length and weight are usually compensated within the first 24 months of life. Unlike weight and height, the kinetics of embryonic and postnatal brain growth in individuals with PM typically worsen with age until adulthood (Figure 1), suggesting that the later stages of brain development, i.e., astrocytogenesis, myelination, and synaptogenesis, may also be affected (see [29,30]).
Cortical malformations such as pachygyria (thick cerebral cortex), polymicrogyria (small, irregular, and shallow cerebral gyri), schizencephaly (cortical mantle interruption), and periventricular nodular/subcortical neuronal heterotopia or lissencephaly (smooth brain) may be associated with PMs [31]. Examples of MCDs often associated with PMs are shown in Figure 1. In addition to microcephaly, these MCDs attest to critical-staged impairments in neurogenesis and neuronal migration. These MCDs severely affect the motor and intellectual prognoses in patients with PMs and cause epilepsy.

2.2. Primary Microcephaly Is an Early Defect in Brain Growth with or without Defects in Body Growth (Microcephalic Primordial Dwarfism)

In the early 1960s, Helmut P.G. Seckel, Prof. of Pediatrics at the University of Chicago, described individuals with a rare phenotype of proportionate short stature or primordial dwarfism, associated with an extreme smallness of the head, with a peculiar aspect resembling “bird-headed” (Figure 2). Prof. Seckel afforded his name to this rare defect, which is diagnosed based on the following five criteria: (1) extreme statural dwarfism (adult standing height of ~120 cm) and in children, a deviation of −6 to −8 SD from the mean normal standing height or length; (2) a head circumference of 39–42 cm in adults (normal: >52 cm) and ~27 cm in newborns; (3) a proportionate smallness of skull and face; (4) a degree of intellectual disability; and (5) a frequent presence of congenital malformations, in addition to the abnormal bird-like facies [32]. In the 1990s, the term “Seckel syndrome” (OMIM # 210600) was expanded to include other clinical types of MPDs, such as Majeski type 2 syndrome (or osteodysplastic primordial dwarfism type 2 (MOPD2) [33]) and Meier–Gorlin syndrome [34]. The progressive identification of the molecular cause of each syndrome has led physicians to propose a new clinicopathological classification for MPDs that considers their phenotypes and genotypes, which also facilitates the description of new entities (see Box 1). These effects are inherited in a recessive manner.
Box 1. Microcephalic primordial dwarfism (MPD).
Definition:
Severe pre- and postnatal growth failure is associated with proportionate or disproportionate microcephaly, with an autosomal recessive inheritance.
Included syndromes:
Seckel syndrome: Intrauterine and postnatal growth retardation, proportionate microcephaly with intellectual disability, and a characteristic “bird-headed” facial appearance [32]. Seckel syndrome PMDs are caused by biallelic variants in ATR, RBBP8, CENPJ, CEP152, CEP63, NIN, DNA2, TRAIP, and NSMCE2 genes.
Osteodysplastic primordial dwarfism type 2 (MOPD2): Severe intrauterine growth retardation with proportionate microcephaly, mesomelia (shortness of the middle portion of a limb), skeletal dysplasia, abnormal dentition, insulin resistance, and cerebral vascular disease with progressive stenosis and occlusion of the cerebral arteries and moyamoya disease. MOPD2 is caused by biallelic variants in the PCNT gene [35,36].
Meier–Gorlin syndrome: Severe intrauterine and postnatal growth retardation, microcephaly, bilateral microtia (underdevelopment of the external ear), and aplasia or hypoplasia of the patellae. Meier–Gorlin syndrome MPDs are caused by biallelic variants in ORC1, ORC4, ORC6, CDT1, CDC6, GMNN, CDC45L, and MCM5 genes [37].
Bloom syndrome: Severe intrauterine and postnatal growth retardation, severe microcephaly, immunodeficiency, sensitivity to sunlight, insulin resistance, and a high risk of cancers with multiple types and sites at an early age. Cognitive abilities seem to be more preserved than in other MPDs. Bloom syndrome is caused by biallelic variants in RECQL3/BLM gene [38].
Ligase IV deficiency: Severe dwarfism with severe and disproportionate microcephaly, combined immunodeficiency, sensitivity to ionizing radiation, and predisposition to cancer. A neurodevelopmental delay seems expected in this syndrome. Ligase IV deficiency is caused by biallelic variants in the LIG4 gene [39].
XRCC4 deficiency: Severe intrauterine and postnatal growth retardation similar to LIG4 deficiency, extreme microcephaly, sensitivity to ionizing radiation without tumor or immunodeficiency in affected children is reported. XRCC4 deficiency is caused by biallelic variants in the XRCC4 gene [40].

2.3. Primary Microcephaly: An Early Defect in Brain Growth with or without Neurosensory Disorders

In addition to microcephaly with or without primordial dwarfism, individuals with PM may exhibit neurosensory disorders (Box 2). These include visual and hearing impairments rarely present at birth and usually occurring during the first few years of life. It is hard to affirm that this neurosensory impairment is caused by variants in PM genes and not due to an undercurrent environmental cause, mainly infectious embryo-fetopathy. Consequently, clinicians have to refer patients to experienced ophthalmologists and ENT specialists who can discriminate genetic from environmental causes.
Visual impairment associated with PM may be caused by anomalies in the development of: (i) the anterior segment of the eye, i.e., microphthalmia, microcornea, and cataracts and (ii) posterior segment, i.e., the abnormal pigmentation of the retina, chorioretinopathies with chorioretinal lacunae and retinal folds, leading to retinal detachment or cone-rod retinal dystrophy. Optic nerve hypoplasia has also been previously reported [41].
Sensorineural hearing loss may be associated with PM. However, as this congenital sensory disorder is common and affects one in 500 newborns, environmental etiologies that account for half of all etiologies need to be ruled out first [42]. The remaining 50% of neurosensory hearing loss cases are of genetic origin and classified into either syndromic causes or isolated, i.e., non-syndromic, causes. Sensorineural hearing loss associated with PMs is syndromic, with inner-ear malformations observed in 40% of patients [42,43,44].
Box 2. Primary microcephaly and neurosensory disorders.
Definition:
Primary microcephaly is associated with chorioretinopathy or sensorineural hearing loss, or both.
Included syndromes:
KIF11 PM: Autosomal dominant microcephaly due to heterozygous KIF11 variant characterized by (i) developmental ocular abnormalities, e.g., chorioretinopathy (choroidal atrophy and non-progressive dysplasia of the retina), retinal folds and detachment, microphthalmia (a developmental disorder of eyes abnormally small at birth), and myopic and hypermetropic astigmatism [45] and (ii) feet congenital lymphedema.
TUBGCP4/6 PM: Autosomal recessive microcephaly characterized by (i) chorioretinopathy similar to cone-rod retinal dystrophy due to biallelic variants in TUBGCP6 [46] or (ii) chorioretinal dysplasia, with multiple punched-out retinal lesions, and other anomalies, e.g., microphthalmia, retinal folding, retinal detachment, optic nerve hypoplasia, the absence of retinal vessels, and round areas of chorioretinal atrophy due to biallelic TUBGCP4 variants [47].
PLK4 PM: Autosomal recessive microcephaly due to biallelic PLK4 variants characterized by (i) chorioretinopathy with pale optic discs, thin retinal vessels, bilateral macular atrophy, and severe generalized retinopathy, but also microphthalmia, microcornea, or cataract, and (ii) dwarfism [46] similar to Seckel syndrome [48].
CDK5RAP2 PM: Autosomal recessive microcephaly due to biallelic CDK5RAP2 variants [4] characterized by (i) progressive hearing loss due to a specific cochlear malformation (small cochlea and a simplification of the cochlear spiral), (ii) ocular defects, including microphthalmia and retinal pigmentation defects, and (iii) interhypothalamic adhesion [12].

2.4. Primary Microcephaly: An Early Defect in Brain Growth with or without Intellectual Disability

The functional neurodevelopmental consequences of PMs include the intellectual disability (ID) of variable severity, behavioral disorders, epilepsy, neurosensory impairments, and cerebral palsy [12,30,49,50,51,52,53]. Identifying the impact of microcephaly on patients’ intellect is a central issue, as it is necessary to ascertain the degree of autonomy and future social insertion for these individuals. However, an accurate assessment of the intellectual abilities of these individuals is performed in rare cases, and motor and intellectual prognoses are delineated only for the most frequent forms [12,51,53]. There is a real benefit in precisely identifying the intellectual abilities of these patients to ascertain their autonomy. Moreover, identifying the PM type with the best or worst prognosis is also necessary as this informs the quality of neuronal and neuronoglial networks in these patients.

3. ASPM, WDR62, and Dynein: Three Emblematic PM Genes Implicated in PMs with or without MCDs

3.1. ASPM: Phenotype–Gene Relationships

3.1.1. Genetics

The ASPM (NM_018136.5) gene encodes the abnormal spindle-like microcephaly-associated protein, a minus-end microtubule-associated protein localized to spindle poles, necessary for spindle pole organization and orientation as it modulates microtubule dynamics at the centrosome and cytokinesis [54,55,56] (see Figure 3). ASPM is the most frequently mutated gene in autosomal recessive PMs. Since the identification of the first patients [6], 861 individuals from 390 families carrying 210 different biallelic variants spread over the gene have been reported (for synthesis, see [50,51] and more recently [57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75]). Approximately 55% of published cases were of Pakistani origin. Of the 210 variants, 207 were loss-of-function mutations or large deletions encompassing several exons of the ASPM gene, predicted to lead to the absence of the protein or a nonfunctional truncated protein. Two specific loss-of-function ASPM variants (c.9754del; pArg3252Glufs*10 [76], and c.9984+1G>T, predicting the removal of the intron 25 splice donor site [54]) have been shown to generate a truncated but stable ASPM protein with reduced expression at spindle poles [54].

3.1.2. Growth

Retrospectively, in Europe, microcephaly in children carrying ASPM mutations was detected during pregnancy in 53% of cases, mainly during the third trimester [51]. Although intrauterine growth retardation was often observed, followed by a short stature within the first two years of life, body growth normalized as feeding difficulties disappeared and the children aged. The heights of these individuals in adulthood are usually comparable to those of healthy individuals [51], except when kyphoscoliosis occurs [60], which suggests that ASPM is not essential for body growth.

3.1.3. Brain Development and Cognition

The brain growth of these individuals is below the normal range before birth and typically slows postnatally (Figure 1). The brains of these children have long been considered “small-scaled” brains, with a characteristic simplification gyration. However, an analysis using structural brain imaging has highlighted that ASPM-PM is not a homothetic reduction in brain volume. The volume of the basal ganglia, cerebellum, and brain stem are better preserved than that of the cerebral cortex, and regional differences in volume also exist within the cerebral cortex. The neocortex is reduced by 50% in volume (and more in surface area, which is insufficiently compensated for by an increase in thickness). However, the hippocampal volume is nearly comparable to that of healthy individuals [77]. Cognitive assessment of these children showed retention of their mnesic abilities, concordant with their normal hippocampal volume. In parallel, a decrease was recorded in their intellectual abilities (mean full-scale intellectual quotient (FSIQ): 57.5 ± 10 SD, range 40–82, n = 40), resulting from a marked reduction in neocortical volume [77]. However, MCDs are rare under this condition, and polymicrogyria (an excessive number of abnormally small cerebral gyri with cortical overfolding) is the only reported MCD in ASPM-PM and has been identified in only four of ninety-seven patients who underwent brain MRI [51,78,79]. Epilepsy is much more frequent and affects 20% of patients with ASPM mutations in Europe [49,51]; thus, it is not necessarily caused by an MCD associated with ASPM-PM. A high proportion of this group of patients comes from consanguineous families. It remains unclear whether ASPM mutations alone cause epilepsy without MCD or whether additional variants in other PM- or epilepsy-causing genes exacerbate the phenotype, as shown by Duerincks et al. and Makhdoom [62,66], underlying variabilities between patients or siblings. Finally, although this question remains unaddressed, the loss of normal excitatory–inhibitory neuronal balance in the cerebral cortex of patients carrying ASPM mutations may also explain the high rate of epilepsy in patients without MCD.

3.2. WDR62: Phenotype–Gene Relationships

3.2.1. Genetics

The WDR62 (NM_001083961.2) gene encodes the WDR repeat-containing protein 62, a minus-end microtubule-associated protein localized to spindle poles with roles in spindle assembly and orientation, centriole duplication, cilium disassembly, and thus in mitotic progression and neuronal migration [15,80,81,82,83] (see Figure 3). The WDR62 gene is the second most frequent mutated gene in autosomal recessive PMs. Since the identification of the first patients [3,10,11], 156 individuals from 74 families, along with 66 different biallelic variants spread over the gene, have been reported (see synthesis in [53,84] and more recently [66,85]). Unlike ASPM, the missense and nonsense/frameshift variants of WDR62 exhibited equal representation. However, the effects of these variants on protein function have not been studied.

3.2.2. Growth

Weight and height are normal in patients with WDR62 mutations, suggesting that WDR62 is not essential for body growth.

3.2.3. Brain Development and Cognition

Microcephaly is usually detected at birth, but a few newborns have normal OFC, which may explain the low rate of in utero diagnosis of this PM type despite careful ultrasound scan monitoring. The OFC declines with age and reaches a mean deviation of −6.5 SD from the mean normal OFC at adolescence [53,84]. In addition to the reduction in brain volume characteristic of PM, WDR62-PM is recognizable in routine brain MRI because of its association with MCDs (see Figure 1). These MCDs include pachygyria in 70% of cases resulting from gyral simplification with broad gyri and an abnormally thick cortex. However, relatively more severe MCDs, such as lissencephaly, bilateral schizencephaly (cortical mantel interruption), polymicrogyria, or neuronal heterotopia grouped in an abnormal location (along the ventricle or within the white matter) occur in 30% of affected individuals [3,10,11,53,86,87,88,89,90,91,92]. Fewer severe cortical malformations have been identified in patients with variants inside the WD-domain [53]. These severe MCDs are evidence of neuronal migration disorders and are almost invariably responsible for epilepsy [3,10,11,53,64,87,89,91,92,93,94,95,96,97,98]. It is difficult to understand the cognitive consequences of WDR62 variants. Considering the associated cortical malformations, a poor functional prognosis can be expected. Although only a few individuals could have been subjected to cognitive evaluation, we were able to assess eleven children out of seventeen on international Wechsler scales, identifying that three of them had mild ID, four of them had moderate ID, and four had severe ID (mean FSIQ: 51.8 ± 12.6 SD, range 40–70) [53]. Remarkably, despite their mild-to-moderate ID, these patients have significant autonomy in daily life and adequate social interactions. It is important to note that one patient out of seventeen had signs of motor decline in the second decade of life, with the occurrence of ataxia and tremors raising fears of neuronal degeneration [53].

3.3. Dynein: Phenotype–Gene Relationships

3.3.1. Genetics

DYNC1H1 (NM_001376.5) encodes a part of the cytoplasmic dynein complex essential for retrograde cargo transport in axons and dendrites and is thus involved in neuronal development, morphology, and survival (see Figure 3). Since the first mutations were identified in patients with cortical malformations [99], 130 individuals have been reported to carry the majority of dominant missense variants in evolutionarily well-conserved domains with functional roles in processive and power-stroke movements [99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121].

3.3.2. Growth

Weight and height are normal in DYNC1H1-related disorders or dyneinopathy, suggesting that DYNC1H1 is not essential for body growth.

3.3.3. Brain Development and Cognition

DYNC1H1-related disorders or dyneinopathy encompasses a spectrum of overlapping disorders, ranging from exclusive neuromuscular phenotype or DYNC1H1–NMD, or peripheral dyneinopathy, to combined neuromuscular and central nervous system dyneinopathy, on either side of the spectrum [121]. Nearly half of the reported patients show exclusive peripheral dyneinopathy predominantly involving the lower limbs, termed spinal muscular atrophy with lower-end predominance [113,122], exhibiting delayed motor milestones, muscle weakness, atrophy hyporeflexia, and skeletal limb abnormalities. These patients do not present the disrupted brain structure or function as is observed in microcephaly, intellectual disability, or cortical malformations. DYNC1H1 mutations causing this phenotype are usually located in the tail domain of DYNC1H1 (53AA–1867AA), predominantly within the dimerization domain (300AA–1140AA). Previous studies have demonstrated that these tail domain mutations do not disrupt the retrograde movement of dynein along microtubules, in contrast to motor domain mutations. Instead, they shorten the run lengths of the processive dynein–dynactin–BICD2N complexes, possibly disrupting neuronal cargo delivery [123].
A second group of patients with central and peripheral dyneinopathy demonstrated a complex phenotype combining predominant lower extremity muscle atrophy and a variable degree of intellectual and global developmental delay, brain malformations in MRI, or both. Patients with central dyneinopathy or central and peripheral dyneinopathy show different degrees of ID, with more severe phenotypes causing epilepsy or spastic paraplegia in affected individuals. Microcephaly usually is detected at birth in these patients; however, a few newborns exhibit normal OFC. Microcephaly is associated with severe cortical malformations and correlates with the severity of MCD. Contrary to other causes of PMs, OFC rarely deviates from the mean normal OFC below −4 SD (personal communication). In MRIs, these patients show a spectrum of MCDs, including pachygyria, dysgyria, and polymicrogyria (see Figure 1), frequently with ventricular anomalies, dysmorphic basal ganglia and corpus callosum, and cerebellar hypoplasia [102,119]. DYNC1H1 mutations causing this phenotype are usually located in the motor domain, MTBD, or linker region, resulting in a disturbed motor activity, possibly secondarily to the severe disruption of neuronal migration and myelination.

4. PCNT: The Major Microcephalic Primordial Dwarfism-Causing Gene

4.1. PCNT: Phenotype–Gene Relationships

4.1.1. Genetics

The PCNT (NM_006031.56) gene encodes pericentrin, a central component of the pericentriolar material (PCM) arranged around a pair of centrioles that constitute the centrosome [124,125,126,127,128,129] (see Figure 3). PCNT is the most frequently mutated gene in patients with recessive MPDs. Since the first patients were identified [35,36], 139 individuals from 116 families carrying 115 different biallelic variants spread over the gene have been reported [66,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153]. All variants had loss-of-function mutations (nonsense, frameshift, and splicing). The effects of a few variants on protein function have also been studied. Homozygous (1887delA; S629fs), (3568_3569insT; C1190fs), (c.6329G>A; p.W2110X), and (c.6711delG; p.E2237fsX2244) variants resulted in the absence of the PCNT protein in the lymphoblastoid cell lines of patients [35,36], whereas the (c.658G>T; E220X) variant products were truncated proteins [36].

4.1.2. Growth

Intrauterine growth retardation is a hallmark of this disorder, with deviations of −5.2 ± 1.9 SD and −4.3 ± 1.6 SD from the mean normal length at birth and the mean normal OFC, respectively. PCNT mutations cause MPD, specifically, MOPD2 or Seckel syndrome [35,36]. Based on this definition, growth retardation in both syndromes is proportional and affects the body and brain growth equally [154]. For a mean age of 8.3 years, the reported patients showed deviations of −8.4 ± 2.2 SD and −8.3 ± 2.4 SD from the mean normal height and the mean normal OFC, respectively (n = 89 and 81 measures for height and OFC, respectively). No significant difference was observed between height and OFC parameters (t-test) [35,36,130,132,135,136,137,138,140,141,142,143,144,145,146,149,150,151,155]. The weight of these patients was less affected than height and OFC (−6.4 ± 3 SD, p < 0.0001, one-way analysis of variance). Morphological features, including skeletal dysplasia in MOPD2 and Seckel syndrome, are described in Box 1 and depicted in Figure 2.

4.1.3. Brain Development and Cognition

Microcephaly is severe (deviation of −8.3 ± 2.4 SD from the mean normal OFC, range: −4.6 to −15 SD for a mean age of 8.3 years) without a real disruption of cortical architecture, despite reports of pachygyria and gyral simplification in a few patients [130,140,143,144,145]. Polymicrogyria is rare (one in fifty-five patients subjected to brain MRI) [140]. Stroke is the leading neurological complication associated with MOPD2 syndrome. Strokes may be ischemic, hemorrhagic, or both. They are caused by progressive occlusive cerebral arteriopathy, which leads to stenosis and intracranial arterial aneurysms. This progressive arteriopathy is associated with developing compensatory capillary collaterals (smoke-like vessels) named moyamoya disease. Strokes or aneurysmal subarachnoid hemorrhages affected 51% of the reported patients who underwent brain MRI (28 out of 55) [35,130,131,135,135,138,139,140,141,142,143,144,148,151] and occurred early (median: 4.2 years old, range: 0.5–26 years). Moyamoya disease was reported in 13 out of 55 patients who underwent brain MRI [35,130,135,138,140,144,151,152]. These cerebral neurovascular anomalies worsen the functional prognosis and are a significant cause of death in patients. The consequences of both microcephaly and stroke syndrome are difficult to assess, as only 17 patients have been evaluated using neuropsychological tests. Neuropsychological assessment of these patients showed a mean full-scale intelligence quotient (FSIQ) of 65 ± 17.8 SD (range 38 to 88). No correlation could be established between intellectual abilities and the severity of microcephaly or stroke, owing to the small number of patients evaluated.

4.1.4. Associated Features or Comorbidities

Heart attack with myocardial infarction caused by coronary artery disease has been reported in 3.2% of patients (4 of 123, diagnosed at 1, 15, 20, and 23 years of age) [141,151]. Insulin-resistant diabetes and dyslipidemia have been identified in almost all adult patients [150]. Lorenzo-Betancor et al. and Huang et al. reported that heterozygous carriers of PCNT variants could develop intra-cranial aneurysms or subarachnoid hemorrhages and must be followed up closely [131,156].

5. CDK5RAP2, CEP152, and PLK4: Three Emblematic PM Genes Associated with Short Stature or Chorioretinopathy or Both

5.1. CDK5RAP2: Phenotype–Gene Relationships

5.1.1. Genetics

The CDK5RAP2 (NM_018249.6) gene encodes CDK5 regulatory subunit associated protein 2 (CEP215), a major protein involved in PCM organization [157,158,159,160,161] (see Figure 3). As a core PCM protein, CDK5RAP2 participates in the nucleation of microtubules and the formation of mitotic spindles. Mutations in CDK5RAP2 are a rare cause of autosomal recessive PMs. Since the identification of the first patients [4], 45 individuals from 23 families, along with 26 different biallelic variants which spread over the gene, have been reported (see synthesis in [12] and more recently [64,72]). Approximately 55% of published cases were of Pakistani origin. Of the 26 variants, 24 were loss-of-function mutations predicted to lead to the absence of a protein or a nonfunctional truncated protein. However, the effects of these variants on protein function have not been studied.

5.1.2. Growth

Short stature and MPD are associated with CDK5RAP2-PM in consanguineous families living on the Asian continent [64,162,163,164]. However, European individuals carrying compound heterozygous variants have normal heights. The position of the mutation does not seem to interfere with the phenotype. Many children with short stature carry mutations within the third, twenty-fifth, or thirtieth exon out of 38, and no variants are located in the domain of interaction with PCNT. In contrast, mutations at identical locations cause short stature in a few children but not others. The average height of the 24 individuals for whom the data are available showed a deviation of −2.8 ± 2.2 SD from the mean normal height. The following hypotheses may explain this inter-individual and geographic variability in height growth: (i) additional variants in MPD genes in consanguineous families, (ii) age of children (spontaneous correction of height post-infancy), and (iii) feeding- or nutrition-related difficulties, or both due to developmental disease or geographic or social context.

5.1.3. Brain Development and Cognition

Typically, microcephaly is detected before or at birth. As for other PMs, the kinetics of brain growth decreases with age (deviation from the mean normal OFC for the 33 patients for whom data are available was −6.6 ± 3.5 SD) in only a proportion of patients. However, a remarkable improvement in brain growth was observed after 2 years of age in others [12], suggesting that mature neurons manage to develop an efficient neuron–glial network, as evident from the normal structure of their corpus callosum and the absence of MCD. This observation is consistent with the preserved intellectual abilities of these patients, which were higher than those of patients with ASPM mutations. The mean value of the FSIQ of 17 assessed patients was 64.6 ± 15.3 SD, ranging from borderline intellectual functioning to mild ID [12,165,166,167,168,169,170]. Along with a simplified gyration typical of PMs, CDK5RAP2-PM has a particularity, i.e., a defect in early diencephalon development that corresponds to a “forme fruste” of holoprosencephaly characterized by the non-separation of hypothalamic nuclei along the midline [12]. This malformation has also been described in STIL-PM or MCPH7 [171,172].

5.1.4. Neurosensory Impairment

Progressive sensorineural hearing loss was identified in four of seven patients from a French series after 6 years of age [12]. This loss of hearing, never observed in PMs, revealed cochlear dysplasia in six of these seven patients characterized by a small and incomplete or simplified cochlea, with only one and a half turns instead of the normal two and a half turns. This cochlear simplification, called Mondini dysplasia, is one of the hallmarks of CDK5RAP2-PM and suggests that CDK5RAP2 is crucial for ear development, as shown by its expression in the fetal cochlea. It is associated with an enlarged vestibular aqueduct and cochlear nerve hypoplasia.
Abnormal eye development is also associated with microphthalmia (a developmental disorder of one or both eyes that are abnormally small at birth) and CDK5RAP2-PM mutations. Specific chorioretinopathy, characterized by the hypo-and hyperpigmentation of the retina and reminiscent of lipofuscin deposits/accumulation in the retinal pigment epithelium, has also been observed. This feature did not impair visual acuity in the examined children [12], unlike other PMs with chorioretinopathies cited below and in Box 2.

5.2. CEP152: Phenotype–Gene Relationships

5.2.1. Genetics

The CEP152 gene (NM_001194998.2) encodes a centrosome protein of 152 kDa, located at the proximal end of the parent centriole, that recruits PLK4 to build a new procentriole in the S phase [173,174,175,176,177,178,179] (see Figure 3). Mutations in CEP152 are a rare cause of autosomal recessive PMs. Since the first patients were identified [8,9], 14 individuals from 9 families with 12 biallelic variants have been reported [8,9,93,180]. Five of the nine variants were loss-of-function mutations predicted to lead to the absence of a protein or a non-functional truncated protein. A founder haplotype was identified in seven Turkish individuals via a recurrent homozygous splice donor-site mutation in intron 4 (c.261+1G>C). This variant led to the formation of four different aberrant transcripts likely to cause a loss of protein function; however, the partial functional activity of one mutant protein, Val86_Asn87del, cannot be excluded [8].

5.2.2. Growth

CEP152 mutations may cause two distinct PM phenotypes, i.e., the MCPH phenotype or the MPD phenotype resembling Seckel syndrome. The MCPH phenotype was observed in three Canadian individuals from different families of Acadian descent, with normal stature carrying the p.Q265P protein variant. The MPD phenotype resembling Seckel syndrome was observed in eight individuals from four distinct Turkish, South African, or Chinese families, with a deviation of −5.4 ± 1.9 SD from the mean normal height for a mean age of 9.3 ± 5.4 years [8,180]. Height was not indicated for the two Pakistani individuals [93]. These individuals with CEP152-MPD also exhibited dysmorphic features with a high nasal bridge and beaked nose, i.e., “bird-head”, a fifth finger clinodactyly (curved finger deviated in a radioulnar or mediolateral direction may overlap other fingers), tooth agenesis, and retrognathia (abnormal positioning of the mandible). However, unlike individuals with PCNT-related phenotype, the subgroup with CEP152-MPD did not exhibit proportionate dwarfism (height and OFC were not equally reduced), as expected in patients with Seckel syndrome.

5.2.3. Brain Development and Cognition

In these 13 individuals, the reduction in brain size was relatively more pronounced than in body size and reached a deviation of −7.7 ± 2.9 SD from the mean normal brain size at an average age of 9.38 ± 6.4 years. Gyral simplification was associated with microcephaly in six cases. Three individuals with MCPH-like phenotype seemed to have preserved cognitive functions as they could read and attend regular classes with or without modifications until 11 years of age. One individual benefited from a neuropsychological assessment showing borderline intellectual functioning evident in visual motor skills. Behavioral disorders, such as tantrums, tics, and obsessive/compulsive traits, were reported in three patients [9]. Intellectual disability appeared to be higher in patients with MPD, although there have been no real neuropsychological assessments of these individuals [8,93].

5.2.4. Neurosensory Impairment

No neurosensory impairment has been reported in individuals bearing mutations in the CEP152 gene.

5.3. PLK4: Phenotype–Gene Relationships

5.3.1. Genetics

The PLK4 gene (NM_001190799) encodes the Polo-like Kinase 4, the master regulator of centriole duplication, which is activated upon phosphorylation and recruits STIL and SAS6 proteins at the proximal end of the parent centriole, thereby initiating the assembly of the procentriole [21,177,179,180,181,182] (see Figure 3). Mutations in PLK4 are an exceedingly rare cause of autosomal recessive PMs. Since the identification of the first patients [46], 15 individuals from seven families carrying five different biallelic variants have been reported [46,48,181,182,183]. A recurrent homozygous mutation (c.1299_1303delAAAG; p. Phe433Leufs*6) has been reported in four families from diverse geographical regions (Madagascar, Iran, Pakistan, and Equatorial Guinea). Levels of functional PLK4 transcripts were reduced to 25% of healthy control levels in individuals with such mutations [46]. The c.2811–5G>C variant created a new splice acceptor site that led to the retention of 4 bp from the intron 15 sequence in PLK4 mRNA, resulting in the premature truncation of the protein at its C terminus and disruption of the terminal Polo-box domain [46]. The c.31-3 A4G substitution disrupts the splicing of the first intron and leads to the transfer of 63 nucleotides from the acceptor site of intron 1 to the mature mRNA. This results in a frameshift and premature translation termination (=, Asp11Profs*14). Four mildly affected individuals have been reported to carry a novel missense variant, c.881 T>G, with no effect on OFC and cognitive functions in the homozygous state, and a hemizygous deletion spanning complete PLK4 and MFSD8 genes and exon 1 of ABHD18 [184].

5.3.2. Growth

All 15 patients exhibited an MPD phenotype with a deviation of −6.5 ± 1.3 SD from the mean normal height at an average age of 6.6 ± 6 years. Unlike individuals with PCNT-related phenotype, individuals carrying PLK4 mutations did not exhibit proportionate dwarfism (height and OFC not reduced equally) as expected in Seckel syndrome.

5.3.3. Brain Development and Cognition

Similar to mutations in CEP152, mutations in the PLK4 gene alter brain growth more than body growth, as the reduction in OFC expression is more pronounced (−11.7 ± 2.3 SD) than in height. The gyral pattern of the brain is extremely simplified [182] and represents what is called microlissencephaly. Interhemispheric arachnoid cysts have been reported in two cases in unrelated patients. Neuronal heterotopia was observed in three of seven patients who underwent brain imaging [46,181,182,183]. No cases of epilepsy have been reported to date. The development of these patients was severely delayed (developmental quotient 21 [181], normal range for the general population: 80–120). Most patients were unable to sit unaided [46] and to speak, which indicates a severe-to-profound intellectual disability.

5.3.4. Neurosensory Impairment

In the brain, the development of the optic vesicle was insufficient based on age (microphthalmia, microcornea, and optic nerve hypoplasia) regardless of the mutation [46,48,181,182], abnormal with a persistent hyperplastic primary vitreous at the origin of retinal detachment and blindness in one patient [182]. Deafness has been reported in three unrelated patients carrying different mutations [46,182].

6. Emblematic PM Genes Encode Centrosome or Spindle Pole Proteins

The centrosome is an organelle without a membrane, composed of two centrioles and a surrounding PCM, which considerably increases its size and capacity to nucleate microtubules during mitosis (see reviews detailing recent advances in centrosome structure and function [24,185]). Thus, the centrosome and minus end of microtubules, mainly nucleated from centrioles during mitosis, form the mitotic spindle pole [24,186,187,188,189]. Deficits in one or the other centrosome or spindle pole protein could severely affect brain development, particularly during the development of the cerebral cortex, mainly affecting mitotic spindle organization, stability, and orientation (see excellent and recent reviews on these topics [17,20,21,190,191]).
During mitosis, ASPM and WDR62 are located at the minus end of microtubules at the centrosome, PCNT and CDK5RAP2 at the PCM, whereas CEP152 and PLK4 are centriolar proteins (Figure 3). Despite most PM genes encoding mitotic apparatus proteins, exceptions exist, in particular, ZNF335 and PHC1 related to transcription/chromatin remodeling processes, NCAPD2/3 and NCAPH involved in chromosome condensation, or a few others involved in DNA damage response (see Appendix A Table A1).

7. Unresolved Issues and Clinical Pitfalls

7.1. Mutations in Emblematic PM Genes Playing a Role in Spindle Pole Structure and Function Are Responsible for Different Diseases with Partial Overlap

Researchers and physicians have long wondered whether the absence of one or more of these spindle pole proteins invariably causes a single disease, i.e., “PM”. The preceding paragraphs show that it is not the case. The absence of specific PM proteins will not lead to similar consequences on brain and neurosensory development or the size of individuals. Mutations in ASPM, WDR62, and DYNC1H1 genes only affect brain size and structure, whereas mutations in CDK5RAP2, CEP152, and PLK4 affect brain size more than body size. Only the mutations in PCNT equally reduced brain and body size. In addition, mutations in CDK5RAP2 and PLK4 cause neurosensory impairments (see Figure S1 and Appendix A Table A1). The effects of mutations in one or another PM gene on brain structure and function, i.e., on the intellectual abilities of patients, will differ.
The functional impairment of microtubule minus-end-targeting proteins ASPM and WDR62, and the main minus-end-directed motors, DYNC1H1, affects brain growth and structure, whereas that of centriolar or PCM proteins PLK4, PCNT, and CDK5RAP2 affects both brain and body size, but to different degrees. Together, these ubiquitously expressed proteins are a part of the microtubule-organizing center, the centrosome, or the spindle pole. Many questions remain unanswered. First, a loss of one of these PM proteins out of a thousand results in a collapse of the centrosome structure. Is there no redundancy? Does the absence of this specific PM protein explain cellular and clinical phenotypes? Alternatively, does the loss of interactions between these PM proteins and their partners explain these defects? Second, how are a few of these mitotic apparatus proteins essential for the growth of all organs and others only for brain growth?
In summary, although they participate in the same function, organization, and stability of the mitotic spindle, their role is not redundant at the organ or organismal scale. The absence or presence of a truncated form of one PM protein cannot be compensated for by the existence or the over-representation of others.

7.2. Why Is the Brain Relatively More Vulnerable than Other Organs?

This question has fascinated researchers for many years and remains unsolved. For example, centrosome proteins are expressed ubiquitously. How to explain then that mutations in PM genes regulating the function of centrosomes will have relatively more consequences on the development of the brain than that of other organs? Three alternative hypotheses can be proposed. The first hypothesis would be related to the short time window of neurogenesis. Indeed, neural progenitors do not proliferate throughout life, unlike progenitors from other organs. Neuron production is restricted to a short period via neurogenesis, which occurs between 6 weeks post-conception (WPC) and 22–24 WPC in humans. The final number of neurons generated during neurogenesis is fixed by the end of the second trimester of pregnancy. No compensation is possible later during brain development. The second hypothesis could be that different isoforms are expressed in a tissue- or organ-specific manner in distinct tissues/organs. It has been shown that a few PM genes (ASPM, MCPH1, Cdk5rap2, and Nin) encode different protein isoforms in humans and mice. These isoforms are relatively more expressed in fetal than in adult tissues/cells and may have different localization patterns and functions throughout life [76,192,193,194]. For example, alternative splicing is responsible for a change in the localization of ninein from centrosome in neural progenitors to non-centrosome sites in neurons [21,194]. The third hypothesis could be related to tissue-specific characteristics. For example, it is to note that the embryonic brain, which is a neuroepithelium and all epithelia (skin, kidney, retina, hair cells of the organ of Corti, and secretory cells from the pituitary gland or pancreas) are polarized tissues and also the most affected by mutations in PM genes (see Figure S1 and Appendix A Table A1). Among polarized tissues, the neuroepithelium and the neurosensory epithelium do not proliferate throughout life. The question remains whether such vulnerability is due to the morphological characteristics of these epithelial cells or to different levels of polarity, among other possibilities [195].

7.3. Do Single Nucleotide Polymorphisms, Not Pathogenic, or Heterozygous Pathogenic Variants in PM Genes Affect Cognitive Functions, Brain Size, or Both?

The genes coding for PM proteins evolved rapidly between ancestral primates and humans (for a review, see [196]. New haplotypes (benign sequence variations or single nucleotide polymorphisms (SNPs)) in PM genes emerged during human evolution as a result of a positive selection that may contribute to brain enlargement [197,198,199]. Researchers investigated possible correlations between common PM-related SNPs and variation in brain volume or intelligence. No correlation has been identified between ancestor SNPs in ASPM and MCPH1 and (i) IQ scores, (ii) head circumference, or (iii) brain volume assessed by MRI [200,201,202]. Sex-specific associations have been found between common non-exonic SNPs in ASPM, MCPH1, and CDK5RAP2 genes and brain volume or cortical surface area in two different series, a Norwegian and a North American series of healthy controls and patients with mental illness or dementia, respectively [203].
The consequences on brain size or cognition of pathogenic variants carried in a homozygous state by the parents of patients remain unknown. Intriguingly, it may influence brain size, as inferred from the OFC measurements of parents, relatively more frequently below the normal range than in the general population (personal observation). Nevertheless, it does not seem to affect the cognitive abilities of these individuals, as indicated by their academic course and economic and social position.

7.4. Do Mutations in One or More PM Genes Allow the Establishment of Prognosis?

Although it is and will always be extremely difficult for physicians to determine long-term prognosis in these patients, this review shows that there are specificities for each of these PMs and that a few PM types are relatively more severe than others. In addition to brain volume reduction, PM severity is based on the following factors: (i) risk of a life-threatening situation (stroke), (ii) associated comorbidities (neurosensory impairment and epilepsy), or (iii) degree of intellectual disability.
Molecular diagnosis enables prognosis to be specified. PCNT mutations are the cause of premature death when aneurysms rupture. Although they represent one of the rarest forms of PM, mutations in PLK4 appear to affect intellectual abilities more severely than mutations in other PM genes. Efforts must continue to identify PMs that have the severest impact on the cognitive functions of patients, which will require more patients to be assessed intellectually and additional anatomical-functional and genotype-functional correlations to be examined.
However, identifying causal mutations is insufficient for establishing a prognosis. Other factors come into play and modulate the effect of mutations on brain growth and cognitive functions. Variability exists among affected patients and siblings carrying mutations in the same gene. These factors may be genetic, and the possibility of digenism in primitive microcephaly is beginning to emerge [49,66]. However, they may also be epigenetic. Such mutations or the presence of other likely benign mutations might together modulate regulatory components or chromatin modifiers, thus increasing the risk of neurodevelopmental disorders that worsen patient prognosis. Finally, environmental factors (infectious or toxic) may affect brain development during neurogenesis. None of these additional hypothetical factors will be easy to identify without carefully following the underlying factors in each pregnant woman, sequencing the whole genome, and exploring the epigenetic pathways for each patient.

7.5. Understanding What Occurred during Neurogenesis in the Brain of Each Individual Affected by PM

Several knockdown animal models have facilitated deciphering mechanisms underlying these emblematic PMs. These mechanisms are based mainly on premature neuronal differentiation and the death of neural progenitors [81,204,205,206,207,208]. However, the absence of a PM protein may not produce the same alterations as a mutated/truncated form of this protein in developing organs, which needs to be considered to elucidate what occurs in the brains of patients during development. Brain organoids are obtained by reprogramming patient cells into induced pluripotent stem cells (IPSCs) and then differentiating them into neural progenitors used to study defects caused by PM mutations during brain development in the genomic context of each patient. Such approaches were used initially to investigate the mechanisms underlying PM and associated defects. Brain organoids have become a powerful tool to model human microcephaly [163]. Generating a mutation or deletion in a PM gene in the IPSCs of healthy controls and differentiating them into brain organoids is the most common method to dissect the role of a gene in human brain development [81], but not to identify altered pathways in patients’ brains. Another approach, centered on their progenitors and neurons, will be necessary to develop personalized medicine for patients.

7.6. Identifying the Temporal Window during Which Defects Occur and during Which Intervention Will Be Possible to Improve the Conditions of Patients

The specificity of the brain, compared with other organs, results from the fact that neural progenitors do not proliferate throughout life. Neurons are produced between 6 weeks post-conception (WPC) and 22–24 WPC in humans. Ultrasound detects neurogenesis impairments with a delay of several weeks, and despite a quick molecular diagnosis, within days or a few weeks, it is generally late in pregnancy or postnatally, and it is also too late to act. However, no compensatory mechanism has been considered. Owing to the non-feasibility of restarting neuronal production after the end of neurogenesis, and hence, enlarging the brain of individuals affected by PM, the only medical option is early intervention in childhood to improve the skills of affected children. This program includes physical and occupational therapies, speech therapy, and psychological support. Antiepileptic drugs are often required in children with seizures.
Physicians, patients, and their families have different expectations. To accept that everything is written before or at birth for the remainder of life is hard. After identifying the underlying mechanisms involved, the question arises of when and how to influence the trajectory of neurons in these individuals to make them relatively more efficient. Physicians and researchers must work together to reflect on this issue, find a way to specifically answer the needs of each patient, and propose a personalized intervention.

8. Conclusions

Since the discovery of PM genes in the early 2000s, our knowledge of neurodevelopmental diseases has increased considerably. Genetics uncovered genes encoding components of the mitotic spindle implicated in primary microcephaly. Animal models have revealed the underlying mechanisms and consequences of the absence of one of the PM proteins in brain development. However, the impact of mutated/truncated PM proteins on brain growth, brain structure, and cognitive functioning in affected patients remains under investigation, making long-term prognosis and early intervention difficult. Nevertheless, this review, combining clinical, cognitive, and imaging data of reported patients with PM, highlights that not all primary microcephalies are similar. Dysfunction in microtubule minus-end proteins at the spindle pole (ASPM and WDR62) or minus-end motor proteins (DYNC1H1) only affects brain growth and structure, whereas dysfunction in centriolar proteins (PLK4 and CEP152) or pericentriolar matrix proteins (PCNT and CDK5RAP2) affects both brain and body growth. Therefore, in addition to conducting further research on impaired pathways in the developing brains of patients using brain organoids, physicians need to explore how the brains of these patients function by assessing their cognitive abilities and thus propose innovative interventions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells12131807/s1, Figure S1: Affected organs related to mutations in PM and MPD genes (created with BioRender.com, accessed on 5 April 2023).

Author Contributions

Conceptualization, S.P.; methodology, formal analysis, drawing of figures, S.F. and H.H.; writing—original draft, review, and editing, S.P. and N.B.-B. All authors have read and agreed to the published version of the manuscript.

Funding

Sandrine Passemard’s work was funded by DGOS (PHRC—NCT01565005), ANR (EuroMicro, ANR-13-RARE-0007-01; MiCMac, ANR 22-CE16-0008), and Université Paris Diderot (DBDD 2014-2018). Nadia Bahi-Buisson’s work was funded by ANR—Dyneinopathies ANR-2016-CE16-0011-01 and ANR-ATOMy–ANR-2019-CE16-0002-01.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to acknowledge Prof. Charles Joris Roux and Dr. Monique Elmaleh-Berges from pediatric radiology departments at Necker Enfants Malades and Robert Debré hospitals in Paris, respectively, for their expertise in brain MRIs and their help with the figures.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Main genes (proteins) mutated in Primary Microcephaly (PM) and Microcephalic Primary Dwarfism (MPD), their functions, functional partners and affected organs. We do not aim in this table to provide an exhaustive list of the genes mutated in PM, MPD and PM with neurosensory disorders, which were recently documented elsewhere [15,21,191,209]. Note that we report here only the partners that are involved in the same functional pathway than that of the PM/MPD proteins.
Table A1. Main genes (proteins) mutated in Primary Microcephaly (PM) and Microcephalic Primary Dwarfism (MPD), their functions, functional partners and affected organs. We do not aim in this table to provide an exhaustive list of the genes mutated in PM, MPD and PM with neurosensory disorders, which were recently documented elsewhere [15,21,191,209]. Note that we report here only the partners that are involved in the same functional pathway than that of the PM/MPD proteins.
LocusGene/ProteinSubcellular LocationMain Molecular FunctionsFunctional Partners Involved in the Same PathwaySyndrome Related to MutationsAffected OrgansReferences
MCPH1MCPH1/
MICROCEPHALIN
Nucleus, mitochondriaChromosome condensation, DNA replication, DNA damage responseCDC27, H2AX upon DNA damagePM/MPD (AR)Brain, skeleton[2,210,211,212,213,214,215]
MCPH2WD62/WDR62Minus end of microtubulesNeuronal proliferation and migrationASPM, AURORA A, CEP152, CEP63, CDK5RAP2PM + MCD (AR)Brain[3,10,11,216,217]
MCPH3CDK5RAP2/
CDK5RAP2
PCMMicrotubule nucleation, centriole disengagement after duplicationPCNT, GAMMA TUBULIN, WDR62, CEP152, CEP63PM (AR)Brain, eye, ear[4,162,216,218,219,220,221]
MCPH4CASC5/KLN1KinetochoreAttachment of centromeres to spindle microtubules, spindle-assembly checkpoint signalingBUB1, BUB1BPM (AR)Brain[222,223,224,225,226]
MCPH5ASPM/ASPMMinus end of microtubulesSpindle pole organization, microtubule dynamics at spindle poleKATNA1, KATNB1, WDR62, CITPM (AR)Brain[6,56,76,217]
MCPH6 & SCKL4CENPJ (CPAP)/
CENPJ (CPAP)
CentrioleCentriole duplication/biogenesisCEP152, STIL, SAS6, GAMMA TUBULIN, CEP135, microtubulesPM/MPD (AR)Brain, skeleton[4,221,227,228,229,230,231]
MCPH7STIL/STILCentriole cartwheelCentriole duplication/
biogenesis, primary cilium
PLK4, CENPJ/CPAP, SAS6PM/MPD (AR)Brain[229,232,233,234,235,236]
MCPH8CEP135/CEP135CentrioleCentriole biogenesis & cohesionCEP152, CENPJ/CPAP, SAS6, GAMMA TUBULIN, MicrotubulesPM/MPD (AR)Brain[231,237,238,239,240]
MCPH9 & SCKL5CEP152/CEP152CentrioleCentriole duplication/
Biogenesis
PLK4, CENPJ/CPAP, WDR62, CEP192, CEP63, CEP135, CDK5RAP2PM/MPD (AR)Brain, skeleton, teeth[8,173,174,216,221,228,231,241]
MCPH10ZNF332/ZNF332NucleusTranscription, neural progenitor proliferationCCAR2, EMSY, ASCL2PM (AR)Brain, skeleton, eye, ear[242,243]
MCPH11PHC1/PHC1NucleusChromatin remodelingPRC1 complexPM/MPD (AR)Brain[244]
MCPH12CDK6/CDK6Nucleus, centrosomeCell cycle control (G1/S transition)CYCLIN DPM (AR)Brain[245,246]
MCPH13CENPE/CENPEPlus-ends of microtubules, kinesin-like motor proteinMicrotubule attachment at the kinetochore, chromosome congressionCENPF, BUB1BPM/MPD (AR)Brain, skeleton, heart[247,248,249,250]
MCPH14SAS6/SAS6Centriole cartwheelCentriole duplication/
Biogenesis
CENPJ/CPAP, STIL, CEP135, CEP152PM (AR)Brain[174,229,231,251,252,253]
MCPH15MSFD2A/
MSFD2A
Cell membraneBlood-brain barrier formation PM (AR)Brain[254,255,256]
MCPH16ANKLE2/ANKLE2Endoplasmic reticulumNuclear envelope reassemblyBAF/BANF1PM/MPD (AR)Brain, eye, skin, bone marrow, skeleton, heart[148,257,258]
MCPH17CIT/CITMidbodyCytokinesisASPM, KIF14PM/MPD (AR)Brain[259,260,261,262,263]
MCPH18WDFY3/WDFY3NucleusAutophagyATG5PM (AD)Brain[264,265]
MCPH19COPB2/COPB2Golgi, cytoplasmic vesicleER-to-Golgi and Golgi-to-ER transport. Coatomer complex proteinalpha, beta, gamma, delta, epsilon and zeta COP-subunitsPM (AR)Brain, skeleton[266]
MCPH20KIF14/KIF14Mitotic spindle, midbodyCell cycle progression, cytokinesisCIT (AR)PM/MPD (AR)Brain, +/− kidney and eye[262,267,268,269]
MCPH21NCAPD2/NCAPD2Nucleus, chromosomesMitotic chromosome condensation/integrityComponent of the condensing complex NCAPD3/H/GMPD (AR)Brain[270,271]
MCPH22NCAPD3/NCAPD3Nucleus, chromosomesMitotic chromosome condensation/integrityComponent of the condensing complex NCAPD2/H/GPM/MPD (AR)Brain[270,271]
MCPH23NCAPH/NCAPHNucleus, chromosomesMitotic chromosome condensation/integrityComponent of the condensing complex NCAPD2/D3/GPM (AR)Brain[270,271]
MCPH24NUP37/NUP37Nucleus, nuclear pore complexKinetochore microtubule attachment, mitotic progressionComponent of the nuclear pore complex (NUP)PM (AR)Brain[272,273]
MCPH25MAP11/MAP11Mitotic spindle, midbodyMitotic spindle dynamics, cytokinesisTUBA1APM (AR)Brain[274]
MCPH26LMNB1/LMNB1Nucleus, inner nuclear membraneNuclear shape, spindle assemblyLamin-associated polypeptidesPM (AD)Brain[275,276,277,278]
MCPH27LMNB2/LMNB2Nucleus, inner nuclear membraneNuclear shape, spindle assemblyLamin-associated polypeptidesPM (AD)Brain[275,276,277,278]
MCPH28RRP7A/RRP7ANucleus, cilium, microtubulesRibosome biogenesis, cilium resorptionSmall subunits of the processosomePM (AR)Brain[279,280]
MCPH29PDCD6IP/
PDCD6IP
Cytoplasmic vesicle membraneCytokinesis, apoptosisESCRT-III componentsPM (AR)Brain[281]
MCPH30BUB1/BUB1Chromosomes, kinetochorespindle-assembly checkpoint signaling, chromosome segregationMAD1L1, BUB3, CASC5/KNL1, CENPEPM/MPD (AR)Brain, skeleton, heart[282,283,284,285]
MOPD2PCNT/PCNTPCMPCM formation and centrosome cohesionGAMMA TUBULIN, CDK5RAP2MPD, MOPD2 and Seckel syndromes (AR)Brain, cerebral arteries, heart, kidney, bone marrow, skin, pituitary gland, pancreas, skeleton, teeth[35,36,286,287,288,289]
SCKL1ATR/ATRNucleus, chromosomesDNA damage responseATRIP, RAD17MPD (AR)Brain, skeleton, bone marrow, teeth[36,290,291,292]
SCKL2RBBP8 (CTIP)/RBBP8Nucleus, chromosomesDouble-stand break repair, homologous recombinationBRCA1, MRN complex, RAD50MPD (AR)Brain, skeleton, eye, teeth[293,294,295,296]
SCKL6CEP63/CEP63Mother CentrioleCentriole integrity/assemblyCEP152, CDK5RAP2, WDR62MPD (AR)Brain[216,297,298]
SCKL7NIN/NINEINSub-distal appendage of mother centrioleMicrotubule anchoring at mother centrioleCCDC120MPD (AR)Brain, pituitary gland, skeleton[299,300,301]
SCKL8DNA2/DNA2Nucleus, mitochondriaDNA replication/repairBLMMPD (AR)Brain[302,303,304,305]
SCKL9TRAIP/TRAIPNucleusReplication stress responsePCNAMPD (AR)Brain, gonads, skeleton, lung, kidney[306,307,308,309]
SCKL10NSMCE2/
NSMCE2
NucleusDNA double-strand break repair–homologous recombinationComponent of the SMC5-SMC6 complexMPD ‘AR)Brain, eye, pancreas, gonads[310,311,312]
MCCRP1TUBGCP6/
TUBGCP6
PCM/MTOCMicrotubule nucleationTUGCP2/3/4/5, GAMMA TUBULINPM + neurosensory disorder (AR)Brain, eye[46,313]
MCCRP2PLK4/PLK4CentrioleCentriole duplication/
Biogenesis
CEP152, CEP192, STILMPD + neurosensory disorder (AR)Brain, eye, ear, skin[46,174,209,237,314,315]
MCCRP3TUBGCP4/
TUBGCP4
PCM/MTOCMicrotubule nucleationTUGCP2/3/4/5, GAMMA TUBULINPM + neurosensory disorder (AR)Brain, eye[47,316]
MCLRMKIF11/KIF11Plus end of microtubulesBipolar spindle establishmentMicrotubulesPM + neurosensory disorder (AD)Brain, eye[317,318]
CDCBM/SMELEDDYNC1H1MicrotubulesTransporter of cargo towards the minus ends of microtubulesDYNC1LI1/2, DYNACTIN, LIS1, NDEL1PM + MCD (AD)Brain, muscle[100,319,320,321]
Bloom (BLM) syndromeRECQL3/RECQL3NucleusDNA replication/repairRAD51, FANCD2, DNA2MPD (AR)Brain, skin, pituitary gland, bone marrow, immune system, gonads, pancreas[322,323,324,325]
LIG4 syndromeLIG4/LIG4NucleusDNA repairXRCC4/5/6, NHEJ1/XLFMPD (AR)Brain, immune system, skin[39,326,327,328]
SSMEDXRCC4/XRCC4NucleusDNA repairXRCC5/6, LIG4,MPD (AR)Brain, eye, ear, pituitary gland, gonads, pancreas, immune system, kidney, +/− heart[40,328,329]
Meyer Gorlin syndrome 1ORC1/ORC1NucleusDNA replicationORC2/3/4/5/6, CDC6MPD (AR)Brain, skeleton[37,330,331]
Meyer Gorlin syndrome 2ORC4/ORC4NucleusDNA replicationORC1/2/3/5/6, CDC6MPD (AR)Brain, skeleton[241]
Meyer Gorlin syndrome 3ORC6/ORC6NucleusDNA replicationORC1/2/3/4/5, CDC6MPD (AR)Brain, skeleton[37]
Meyer Gorlin syndrome 4CDT1/CDT1Nucleus, chromosomes, kinetochoreDNA replication, mitosisCDC6, PCNA, GMMNMPD (AR)Brain, skeleton[37,241,332,333]
Meyer Gorlin syndrome 5CDC6/CDC6NucleusDNA replicationORC1, CDT1, PCNAMPD (AR)Brain, skeleton, teeth[37,334,335]
Meyer Gorlin syndrome 6GMNN/GMNNNucleusDNA replicationCDT1MPD (AD)Brain, skeleton, ear[336,337]
CDCBM: Cortical dysplasia, complex brain malformation; ER: Endoplasmic reticulum; ESCRT III: endosomal sorting required for transport complex III; MCCRP: MicroCephaly and ChorioRetinoPathy; MCD: malformation of cortical development; MCLRM: microcephaly with or without chorioretinopathy, lymphedema or impaired intellectual development; MOPD2: microcephalic osteodysplastic primordial dwarfism; PCM: pericentriolar material; MPD: microcephalic primordial dwarfism; SMALED: Spinal muscular atrophy, lower extremity predominant; SSMED: short stature, microcephaly, endocrine dysfunction.

References

  1. Wright, C.M.; Emond, A. Head Growth and Neurocognitive Outcomes. Pediatrics 2015, 135, e1393–e1398. [Google Scholar] [CrossRef] [Green Version]
  2. Jackson, A.P.; Eastwood, H.; Bell, S.M.; Adu, J.; Toomes, C.; Carr, I.M.; Roberts, E.; Hampshire, D.J.; Crow, Y.J.; Mighell, A.J.; et al. Identification of Microcephalin, a Protein Implicated in Determining the Size of the Human Brain. Am. J. Hum. Genet. 2002, 71, 136–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Nicholas, A.K.; Khurshid, M.; Desir, J.; Carvalho, O.P.; Cox, J.J.; Thornton, G.; Kausar, R.; Ansar, M.; Ahmad, W.; Verloes, A.; et al. WDR62 Is Associated with the Spindle Pole and Is Mutated in Human Microcephaly. Nat. Genet. 2010, 42, 1010–1014. [Google Scholar] [CrossRef] [PubMed]
  4. Bond, J.; Roberts, E.; Springell, K.; Lizarraga, S.B.; Scott, S.; Higgins, J.; Hampshire, D.J.; Morrison, E.E.; Leal, G.F.; Silva, E.O.; et al. A Centrosomal Mechanism Involving CDK5RAP2 and CENPJ Controls Brain Size. Nat. Genet. 2005, 37, 353–355. [Google Scholar] [CrossRef] [PubMed]
  5. Genin, A.; Desir, J.; Lambert, N.; Biervliet, M.; Van Der Aa, N.; Pierquin, G.; Killian, A.; Tosi, M.; Urbina, M.; Lefort, A.; et al. Kinetochore KMN Network Gene CASC5 Mutated in Primary Microcephaly. Hum. Mol. Genet. 2012, 21, 5306–5317. [Google Scholar] [CrossRef] [Green Version]
  6. Bond, J.; Roberts, E.; Mochida, G.H.; Hampshire, D.J.; Scott, S.; Askham, J.M.; Springell, K.; Mahadevan, M.; Crow, Y.J.; Markham, A.F.; et al. ASPM Is a Major Determinant of Cerebral Cortical Size. Nat. Genet. 2002, 32, 316–320. [Google Scholar] [CrossRef]
  7. Al-Dosari, M.S.; Shaheen, R.; Colak, D.; Alkuraya, F.S. Novel CENPJ Mutation Causes Seckel Syndrome. J. Med. Genet. 2010, 47, 411–414. [Google Scholar] [CrossRef]
  8. Kalay, E.; Yigit, G.; Aslan, Y.; Brown, K.E.; Pohl, E.; Bicknell, L.S.; Kayserili, H.; Li, Y.; Tuysuz, B.; Nurnberg, G.; et al. CEP152 Is a Genome Maintenance Protein Disrupted in Seckel Syndrome. Nat. Genet. 2011, 43, 23–26. [Google Scholar] [CrossRef] [Green Version]
  9. Guernsey, D.L.; Jiang, H.; Hussin, J.; Arnold, M.; Bouyakdan, K.; Perry, S.; Babineau-Sturk, T.; Beis, J.; Dumas, N.; Evans, S.C.; et al. Mutations in Centrosomal Protein CEP152 in Primary Microcephaly Families Linked to MCPH4. Am. J. Hum. Genet. 2010, 87, 40–51. [Google Scholar] [CrossRef] [Green Version]
  10. Bilgüvar, K.; Oztürk, A.K.; Louvi, A.; Kwan, K.Y.; Choi, M.; Tatli, B.; Yalnizoğlu, D.; Tüysüz, B.; Cağlayan, A.O.; Gökben, S.; et al. Whole-Exome Sequencing Identifies Recessive WDR62 Mutations in Severe Brain Malformations. Nature 2010, 467, 207–210. [Google Scholar] [CrossRef] [Green Version]
  11. Yu, T.W.; Mochida, G.H.; Tischfield, D.J.; Sgaier, S.K.; Flores-Sarnat, L.; Sergi, C.M.; Topcu, M.; McDonald, M.T.; Barry, B.J.; Felie, J.M.; et al. Mutations in WDR62, Encoding a Centrosome-Associated Protein, Cause Microcephaly with Simplified Gyri and Abnormal Cortical Architecture. Nat. Genet. 2010, 42, 1015–1020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Nasser, H.; Vera, L.; Elmaleh-Bergès, M.; Steindl, K.; Letard, P.; Teissier, N.; Ernault, A.; Guimiot, F.; Afenjar, A.; Moutard, M.L.; et al. CDK5RAP2 Primary Microcephaly Is Associated with Hypothalamic, Retinal and Cochlear Developmental Defects. J. Med. Genet. 2020, 57, 389–399. [Google Scholar] [CrossRef]
  13. Gotz, M.; Huttner, W.B. The Cell Biology of Neurogenesis. Nat. Rev. Mol. Cell Biol. 2005, 6, 777–788. [Google Scholar] [CrossRef] [PubMed]
  14. Lui, J.H.; Hansen, D.V.; Kriegstein, A.R. Development and Evolution of the Human Neocortex. Cell 2011, 146, 18–36. [Google Scholar] [CrossRef] [Green Version]
  15. Jayaraman, D.; Bae, B.I.; Walsh, C.A. The Genetics of Primary Microcephaly. Annu. Rev. Genom. Hum. Genet. 2018, 19, 177–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Conduit, P.T.; Wainman, A.; Raff, J.W. Centrosome Function and Assembly in Animal Cells. Nat. Rev. Mol. Cell Biol. 2015, 16, 611–624. [Google Scholar] [CrossRef] [PubMed]
  17. Nigg, E.A.; Holland, A.J. Once and Only Once: Mechanisms of Centriole Duplication and Their Deregulation in Disease. Nat. Rev. Mol. Cell Biol. 2018, 19, 297–312. [Google Scholar] [CrossRef]
  18. Iwata, R.; Vanderhaeghen, P. Regulatory Roles of Mitochondria and Metabolism in Neurogenesis. Curr. Opin. Neurobiol. 2021, 69, 231–240. [Google Scholar] [CrossRef]
  19. Saade, M.; Blanco-Ameijeiras, J.; Gonzalez-Gobartt, E.; Martí, E. A Centrosomal View of CNS Growth. Development 2018, 145, dev170613. [Google Scholar] [CrossRef] [Green Version]
  20. Chavali, P.L.; Putz, M.; Gergely, F. Small Organelle, Big Responsibility: The Role of Centrosomes in Development and Disease. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130468. [Google Scholar] [CrossRef] [Green Version]
  21. Marthiens, V.; Basto, R. Centrosomes: The Good and the Bad for Brain Development. Biol. Cell 2020, 112, 153–172. [Google Scholar] [CrossRef]
  22. Yang, J.; Hu, X.; Ma, J.; Shi, S.-H. Centrosome Regulation and Function in Mammalian Cortical Neurogenesis. Curr. Opin. Neurobiol. 2021, 69, 256–266. [Google Scholar] [CrossRef]
  23. Anjur-Dietrich, M.I.; Kelleher, C.P.; Needleman, D.J. Mechanical Mechanisms of Chromosome Segregation. Cells 2021, 10, 465. [Google Scholar] [CrossRef] [PubMed]
  24. Vasquez-Limeta, A.; Loncarek, J. Human Centrosome Organization and Function in Interphase and Mitosis. Semin. Cell Dev. Biol. 2021, 117, 30–41. [Google Scholar] [CrossRef] [PubMed]
  25. Barisic, M.; Rajendraprasad, G.; Steblyanko, Y. The Metaphase Spindle at Steady State–Mechanism and Functions of Microtubule Poleward Flux. Semin. Cell Dev. Biol. 2021, 117, 99–117. [Google Scholar] [CrossRef] [PubMed]
  26. Prosser, S.L.; Pelletier, L. Mitotic Spindle Assembly in Animal Cells: A Fine Balancing Act. Nat. Rev. Mol. Cell Biol. 2017, 18, 187–201. [Google Scholar] [CrossRef]
  27. Storchova, Z. Consequences of Mitotic Failure–The Penalties and the Rewards. Semin. Cell Dev. Biol. 2021, 117, 149–158. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, S.; Pan, Y.; Auger, N.; Sun, W.; Dai, L.; Li, S.; Xie, S.; Wen, S.W.; Chen, D. Small Head Circumference at Birth: An 8-Year Retrospective Cohort Study in China. BMJ Paediatr. Open 2019, 3, e000470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Passemard, S.; Perez, F.; Colin-Lemesre, E.; Rasika, S.; Gressens, P.; El Ghouzzi, V. Golgi Trafficking Defects in Postnatal Microcephaly: The Evidence for “Golgipathies”. Prog. Neurobiol. 2017, 153, 46–63. [Google Scholar] [CrossRef] [Green Version]
  30. Gilmore, E.C.; Walsh, C.A. Genetic Causes of Microcephaly and Lessons for Neuronal Development. Wiley Interdiscip. Rev. Dev. Biol. 2013, 2, 461–478. [Google Scholar] [CrossRef]
  31. Oegema, R.; Barakat, T.S.; Wilke, M.; Stouffs, K.; Amrom, D.; Aronica, E.; Bahi-Buisson, N.; Conti, V.; Fry, A.E.; Geis, T.; et al. International Consensus Recommendations on the Diagnostic Work-up for Malformations of Cortical Development. Nat. Rev. Neurol. 2020, 16, 618–635. [Google Scholar] [CrossRef]
  32. Helmut, P.G. Seckel Bird-Headed Dwarfs: Studies in Developemental Anthropology Including Human Proportions; S. Karger AG: Basel, Switzerland, 1960. [Google Scholar]
  33. Majewski, F.; Ranke, M.; Schinzel, A. Studies of Microcephalic Primordial Dwarfism II: The Osteodysplastic Type II of Primordial Dwarfism. Am. J. Med. Genet. 1982, 12, 23–35. [Google Scholar] [CrossRef] [PubMed]
  34. Boles, R.G.; Teebi, A.S.; Schwartz, D.; Harper, J.F. Further Delineation of the Ear, Patella, Short Stature Syndrome (Meier-Gorlin Syndrome). Clin. Dysmorphol. 1994, 3, 207–214. [Google Scholar] [CrossRef] [PubMed]
  35. Rauch, A.; Thiel, C.T.; Schindler, D.; Wick, U.; Crow, Y.J.; Ekici, A.B.; van Essen, A.J.; Goecke, T.O.; Al-Gazali, L.; Chrzanowska, K.H.; et al. Mutations in the Pericentrin (PCNT) Gene Cause Primordial Dwarfism. Science 2008, 319, 816–819. [Google Scholar] [CrossRef] [PubMed]
  36. Griffith, E.; Walker, S.; Martin, C.A.; Vagnarelli, P.; Stiff, T.; Vernay, B.; Al Sanna, N.; Saggar, A.; Hamel, B.; Earnshaw, W.C.; et al. Mutations in Pericentrin Cause Seckel Syndrome with Defective ATR-Dependent DNA Damage Signaling. Nat. Genet. 2008, 40, 232–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Bicknell, L.S.; Bongers, E.M.H.F.; Leitch, A.; Brown, S.; Schoots, J.; Harley, M.E.; Aftimos, S.; Al-Aama, J.Y.; Bober, M.; Brown, P.A.J.; et al. Mutations in the Pre-Replication Complex Cause Meier-Gorlin Syndrome. Nat. Genet. 2011, 43, 356–359. [Google Scholar] [CrossRef] [Green Version]
  38. Ellis, N.A.; German, J. Molecular Genetics of Bloom’s Syndrome. Hum. Mol. Genet. 1996, 5, 1457–1463. [Google Scholar] [CrossRef] [Green Version]
  39. O’Driscoll, M.; Cerosaletti, K.M.; Girard, P.M.; Dai, Y.; Stumm, M.; Kysela, B.; Hirsch, B.; Gennery, A.; Palmer, S.E.; Seidel, J.; et al. DNA Ligase IV Mutations Identified in Patients Exhibiting Developmental Delay and Immunodeficiency. Mol. Cell 2001, 8, 1175–1185. [Google Scholar] [CrossRef]
  40. Murray, J.E.; van der Burg, M.; IJspeert, H.; Carroll, P.; Wu, Q.; Ochi, T.; Leitch, A.; Miller, E.S.; Kysela, B.; Jawad, A.; et al. Mutations in the NHEJ Component XRCC4 Cause Primordial Dwarfism. Am. J. Hum. Genet. 2015, 96, 412–424. [Google Scholar] [CrossRef] [Green Version]
  41. Shurygina, M.F.; Simonett, J.M.; Parker, M.A.; Mitchell, A.; Grigorian, F.; Lifton, J.; Nagiel, A.; Shpak, A.A.; Dadali, E.L.; Mishina, I.A.; et al. Genotype Phenotype Correlation and Variability in Microcephaly Associated With Chorioretinopathy or Familial Exudative Vitreoretinopathy. Investig. Ophthalmol. Vis. Sci. 2020, 61, 2. [Google Scholar] [CrossRef]
  42. Morton, C.C.; Nance, W.E. Newborn Hearing Screening—A Silent Revolution. N. Engl. J. Med. 2006, 354, 2151–2164. [Google Scholar] [CrossRef] [PubMed]
  43. Wentland, C.J.; Ronner, E.A.; Basonbul, R.A.; Pinnapureddy, S.; Mankarious, L.; Keamy, D.; Lee, D.J.; Cohen, M.S. Utilization of Diagnostic Testing for Pediatric Sensorineural Hearing Loss. Int. J. Pediatr. Otorhinolaryngol. 2018, 111, 26–31. [Google Scholar] [CrossRef] [PubMed]
  44. Teissier, N.; Van Den Abbeele, T.; Sebag, G.; Elmaleh-Berges, M. Computed Tomography Measurements of the Normal and the Pathologic Cochlea in Children. Pediatr. Radiol. 2010, 40, 275–283. [Google Scholar] [CrossRef]
  45. Birtel, J.; Gliem, M.; Mangold, E.; Tebbe, L.; Spier, I.; Müller, P.L.; Holz, F.G.; Neuhaus, C.; Wolfrum, U.; Bolz, H.J.; et al. Novel Insights Into the Phenotypical Spectrum of KIF11-Associated Retinopathy, Including a New Form of Retinal Ciliopathy. Investig. Ophthalmol. Vis. Sci. 2017, 58, 3950–3959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Martin, C.A.; Ahmad, I.; Klingseisen, A.; Hussain, M.S.; Bicknell, L.S.; Leitch, A.; Nurnberg, G.; Toliat, M.R.; Murray, J.E.; Hunt, D.; et al. Mutations in PLK4, Encoding a Master Regulator of Centriole Biogenesis, Cause Microcephaly, Growth Failure and Retinopathy. Nat. Genet. 2014, 46, 1283–1292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Scheidecker, S.; Etard, C.; Haren, L.; Stoetzel, C.; Hull, S.; Arno, G.; Plagnol, V.; Drunat, S.; Passemard, S.; Toutain, A.; et al. Mutations in TUBGCP4 Alter Microtubule Organization via the Gamma-Tubulin Ring Complex in Autosomal-Recessive Microcephaly with Chorioretinopathy. Am. J. Hum. Genet. 2015, 96, 666–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Shaheen, R.; Al Tala, S.; Almoisheer, A.; Alkuraya, F.S. Mutation in PLK4, Encoding a Master Regulator of Centriole Formation, Defines a Novel Locus for Primordial Dwarfism. J. Med. Genet. 2014, 51, 814–816. [Google Scholar] [CrossRef]
  49. Duerinckx, S.; Désir, J.; Perazzolo, C.; Badoer, C.; Jacquemin, V.; Soblet, J.; Maystadt, I.; Tunca, Y.; Blaumeiser, B.; Ceulemans, B.; et al. Phenotypes and Genotypes in Non-Consanguineous and Consanguineous Primary Microcephaly: High Incidence of Epilepsy. Mol. Genet. Genom. Med. 2021, 9, e1768. [Google Scholar] [CrossRef]
  50. Verloes, A.; Drunat, S.; Passemard, S. ASPM Primary Microcephaly. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J., Mirzaa, G., Amemiya, A., Eds.; University of Washington, Seattle: Seattle, WA, USA, 1993. [Google Scholar]
  51. Letard, P.; Drunat, S.; Vial, Y.; Duerinckx, S.; Ernault, A.; Amram, D.; Arpin, S.; Bertoli, M.; Busa, T.; Ceulemans, B.; et al. Autosomal Recessive Primary Microcephaly Due to ASPM Mutations: An Update. Hum. Mutat. 2018, 39, 319–332. [Google Scholar] [CrossRef]
  52. Woods, C.G.; Parker, A. Investigating Microcephaly. Arch. Dis. Child. 2013, 98, 707–713. [Google Scholar] [CrossRef]
  53. Ruaud, L.; Drunat, S.; Elmaleh-Bergès, M.; Ernault, A.; Guilmin Crepon, S.; The MCPH Consortium; El Ghouzzi, V.; Auvin, S.; Verloes, A.; Passemard, S.; et al. Neurological Outcome in WDR62 Primary Microcephaly. Dev. Med. Child Neurol. 2022, 64, 509–517. [Google Scholar] [CrossRef]
  54. Higgins, J.; Midgley, C.; Bergh, A.M.; Bell, S.M.; Askham, J.M.; Roberts, E.; Binns, R.K.; Sharif, S.M.; Bennett, C.; Glover, D.M.; et al. Human ASPM Participates in Spindle Organisation, Spindle Orientation and Cytokinesis. BMC Cell Biol. 2010, 11, 85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Wakefield, J.G.; Bonaccorsi, S.; Gatti, M. The Drosophila Protein Asp Is Involved in Microtubule Organization during Spindle Formation and Cytokinesis. J. Cell Biol. 2001, 153, 637–648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Jiang, K.; Rezabkova, L.; Hua, S.; Liu, Q.; Capitani, G.; Altelaar, A.F.M.; Heck, A.J.R.; Kammerer, R.A.; Steinmetz, M.O.; Akhmanova, A. Microtubule Minus-End Regulation at Spindle Poles by an ASPM-Katanin Complex. Nat. Cell Biol. 2017, 19, 480–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bhargav, D.S.; Sreedevi, N.; Swapna, N.; Vivek, S.; Kovvali, S. Whole Exome Sequencing Identifies a Novel Homozygous Frameshift Mutation in the ASPM Gene, Which Causes Microcephaly 5, Primary, Autosomal Recessive. F1000Res 2017, 6, 2163. [Google Scholar] [CrossRef] [Green Version]
  58. Weitensteiner, V.; Zhang, R.; Bungenberg, J.; Marks, M.; Gehlen, J.; Ralser, D.J.; Hilger, A.C.; Sharma, A.; Schumacher, J.; Gembruch, U.; et al. Exome Sequencing in Syndromic Brain Malformations Identifies Novel Mutations in ACTB, and SLC9A6, and Suggests BAZ1A as a New Candidate Gene. Birth Defects Res. 2018, 110, 587–597. [Google Scholar] [CrossRef] [Green Version]
  59. Khan, A.; Wang, R.; Han, S.; Ahmad, W.; Zhang, X. Identification of a Novel Nonsense ASPM Mutation in a Large Consanguineous Pakistani Family Using Targeted Next-Generation Sequencing. Genet. Test. Mol. Biomark. 2018, 22, 159–164. [Google Scholar] [CrossRef]
  60. Marakhonov, A.V.; Konovalov, F.A.; Makaov, A.K.; Vasilyeva, T.A.; Kadyshev, V.V.; Galkina, V.A.; Dadali, E.L.; Kutsev, S.I.; Zinchenko, R.A. Primary Microcephaly Case from the Karachay-Cherkess Republic Poses an Additional Support for Microcephaly and Seckel Syndrome Spectrum Disorders. BMC Med. Genom. 2018, 11, 8. [Google Scholar] [CrossRef] [Green Version]
  61. Okamoto, N.; Kohmoto, T.; Naruto, T.; Masuda, K.; Imoto, I. Primary Microcephaly Caused by Novel Compound Heterozygous Mutations in ASPM. Hum. Genome Var. 2018, 5, 18015. [Google Scholar] [CrossRef] [Green Version]
  62. Duerinckx, S.; Jacquemin, V.; Drunat, S.; Vial, Y.; Passemard, S.; Perazzolo, C.; Massart, A.; Soblet, J.; Racapé, J.; Desmyter, L.; et al. Digenic Inheritance of Human Primary Microcephaly Delineates Centrosomal and Non-Centrosomal Pathways. Hum. Mutat. 2019, 41, 512–524. [Google Scholar] [CrossRef] [Green Version]
  63. Bazgir, A.; Agha Gholizadeh, M.; Sarvar, F.; Pakzad, Z. A Novel Frameshift Mutation in Abnormal Spindle-Like Microcephaly (ASPM) Gene in an Iranian Patient with Primary Microcephaly: A Case Report. Iran. J. Public Health 2019, 48, 2074–2078. [Google Scholar] [CrossRef] [PubMed]
  64. Rasool, S.; Baig, J.M.; Moawia, A.; Ahmad, I.; Iqbal, M.; Waseem, S.S.; Asif, M.; Abdullah, U.; Makhdoom, E.U.H.; Kaygusuz, E.; et al. An Update of Pathogenic Variants in ASPM, WDR62, CDK5RAP2, STIL, CENPJ, and CEP135 Underlying Autosomal Recessive Primary Microcephaly in 32 Consanguineous Families from Pakistan. Mol. Genet. Genom. Med. 2020, 8, e1408. [Google Scholar] [CrossRef]
  65. Naseer, M.I.; Abdulkareem, A.A.; Muthaffar, O.Y.; Sogaty, S.; Alkhatabi, H.; Almaghrabi, S.; Chaudhary, A.G. Whole Exome Sequencing Identifies Three Novel Mutations in the ASPM Gene From Saudi Families Leading to Primary Microcephaly. Front. Pediatr. 2020, 8, 627122. [Google Scholar] [CrossRef] [PubMed]
  66. Makhdoom, E.U.H.; Waseem, S.S.; Iqbal, M.; Abdullah, U.; Hussain, G.; Asif, M.; Budde, B.; Höhne, W.; Tinschert, S.; Saadi, S.M.; et al. Modifier Genes in Microcephaly: A Report on WDR62, CEP63, RAD50 and PCNT Variants Exacerbating Disease Caused by Biallelic Mutations of ASPM and CENPJ. Genes 2021, 12, 731. [Google Scholar] [CrossRef] [PubMed]
  67. Khan, N.M.; Hussain, B.; Zheng, C.; Khan, A.; Masoud, M.S.; Gu, Q.; Qiu, L.; Malik, N.A.; Qasim, M.; Tariq, M.; et al. Updates on Clinical and Genetic Heterogeneity of ASPM in 12 Autosomal Recessive Primary Microcephaly Families in Pakistani Population. Front. Pediatr. 2021, 9, 695133. [Google Scholar] [CrossRef]
  68. Batool, T.; Irshad, S.; Mahmood, K. Novel Pathogenic Mutation Mapping of ASPM Gene in Consanguineous Pakistani Families with Primary Microcephaly. Braz. J. Biol. 2021, 83, e246040. [Google Scholar] [CrossRef] [PubMed]
  69. Tran, T.H.; Diep, Q.M.; Cao, M.H.; Luong, L.H.; Pham, V.A.; Lan Dinh, O.T.; Bui, T.-H.; Van Ta, T.; Tran, V.K. Microcephaly Primary Hereditary (MCPH): Report of Novel ASPM Variants and Prenatal Diagnosis in a Vietnamese Family. Taiwan J. Obstet. Gynecol. 2021, 60, 907–910. [Google Scholar] [CrossRef] [PubMed]
  70. Xu, S.; Zhang, W.; Zhou, R.; Huang, H.; Chen, W.; Xiang, W.; Liu, L.; Song, J. Two Novel Truncating Variants of the ASPM Gene Identified in a Nonconsanguineous Chinese Family Associated with Primary Microcephaly. Clin. Dysmorphol. 2022, 31, 1–5. [Google Scholar] [CrossRef]
  71. Hussain, S.; Nawaz, A.; Hamid, M.; Ullah, W.; Khan, I.N.; Afshan, M.; Rehman, A.; Nawaz, H.; Halswick, J.; Rehman, S.-U.; et al. Mutation Screening of Multiple Pakistani MCPH Families Revealed Novel and Recurrent Protein-Truncating Mutations of ASPM. Biotechnol. Appl. Biochem. 2022, 69, 2296–2303. [Google Scholar] [CrossRef]
  72. Makhdoom, E.U.H.; Anwar, H.; Baig, S.M.; Hussain, G. Whole Exome Sequencing Identifies a Novel Mutation in ASPM and Ultra-Rare Mutation in CDK5RAP2 Causing Primary Microcephaly in Consanguineous Pakistani Families. Pak. J. Med. Sci. 2022, 38, 84–89. [Google Scholar] [CrossRef]
  73. Naqvi, S.F.; Shabbir, R.M.K.; Tolun, A.; Basit, S.; Malik, S. A Two-Base Pair Deletion in IQ Repeats in ASPM Underlies Microcephaly in a Pakistani Family. Genet. Test. Mol. Biomark. 2022, 26, 37–42. [Google Scholar] [CrossRef] [PubMed]
  74. Türkyılmaz, A.; Sager, S.G. Two New Cases of Primary Microcephaly with Neuronal Migration Defect Caused by Truncating Mutations in the ASPM Gene. Mol. Syndromol. 2022, 13, 56–63. [Google Scholar] [CrossRef]
  75. Li, M.; Luo, J.; Yang, Q.; Chen, F.; Chen, J.; Qin, J.; He, W.; Chen, J.; Yi, S.; Qin, Z.; et al. Novel and Recurrent ASPM Mutations of Founder Effect in Chinese Population. Brain Dev. 2022, 44, 540–545. [Google Scholar] [CrossRef] [PubMed]
  76. Kouprina, N.; Pavlicek, A.; Collins, N.K.; Nakano, M.; Noskov, V.N.; Ohzeki, J.; Mochida, G.H.; Risinger, J.I.; Goldsmith, P.; Gunsior, M.; et al. The Microcephaly ASPM Gene Is Expressed in Proliferating Tissues and Encodes for a Mitotic Spindle Protein. Hum. Mol. Genet. 2005, 14, 2155–2165. [Google Scholar] [CrossRef] [PubMed]
  77. Passemard, S.; Verloes, A.; Billette de Villemeur, T.; Boespflug-Tanguy, O.; Hernandez, K.; Laurent, M.; Isidor, B.; Alberti, C.; Pouvreau, N.; Drunat, S.; et al. Abnormal Spindle-like Microcephaly-Associated (ASPM) Mutations Strongly Disrupt Neocortical Structure but Spare the Hippocampus and Long-Term Memory. Cortex 2016, 74, 158–176. [Google Scholar] [CrossRef]
  78. Passemard, S.; Titomanlio, L.; Elmaleh, M.; Afenjar, A.; Alessandri, J.L.; Andria, G.; de Villemeur, T.B.; Boespflug-Tanguy, O.; Burglen, L.; Del Giudice, E.; et al. Expanding the Clinical and Neuroradiologic Phenotype of Primary Microcephaly Due to ASPM Mutations. Neurology 2009, 73, 962–969. [Google Scholar] [CrossRef]
  79. Hu, H.; Suckow, V.; Musante, L.; Roggenkamp, V.; Kraemer, N.; Ropers, H.H.; Hubner, C.; Wienker, T.F.; Kaindl, A.M. Previously Reported New Type of Autosomal Recessive Primary Microcephaly Is Caused by Compound Heterozygous ASPM Gene Mutations. Cell Cycle 2014, 13, 1650–1651. [Google Scholar] [CrossRef] [Green Version]
  80. Huang, J.; Liang, Z.; Guan, C.; Hua, S.; Jiang, K. WDR62 Regulates Spindle Dynamics as an Adaptor Protein between TPX2/Aurora A and Katanin. J. Cell Biol. 2021, 220, e202007167. [Google Scholar] [CrossRef]
  81. Zhang, W.; Yang, S.-L.; Yang, M.; Herrlinger, S.; Shao, Q.; Collar, J.L.; Fierro, E.; Shi, Y.; Liu, A.; Lu, H.; et al. Modeling Microcephaly with Cerebral Organoids Reveals a WDR62-CEP170-KIF2A Pathway Promoting Cilium Disassembly in Neural Progenitors. Nat. Commun. 2019, 10, 2612. [Google Scholar] [CrossRef] [Green Version]
  82. Chen, J.-F.; Zhang, Y.; Wilde, J.; Hansen, K.C.; Lai, F.; Niswander, L. Microcephaly Disease Gene Wdr62 Regulates Mitotic Progression of Embryonic Neural Stem Cells and Brain Size. Nat. Commun. 2014, 5, 3885. [Google Scholar] [CrossRef] [Green Version]
  83. Sgourdou, P.; Mishra-Gorur, K.; Saotome, I.; Henagariu, O.; Tuysuz, B.; Campos, C.; Ishigame, K.; Giannikou, K.; Quon, J.L.; Sestan, N.; et al. Disruptions in Asymmetric Centrosome Inheritance and WDR62-Aurora Kinase B Interactions in Primary Microcephaly. Sci. Rep. 2017, 7, 43708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Verloes, A.; Ruaud, L.; Drunat, S.; Passemard, S. WDR62 Primary Microcephaly. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J., Gripp, K.W., Mirzaa, G.M., Amemiya, A., Eds.; University of Washington, Seattle: Seattle WA, USA, 1993. [Google Scholar]
  85. Aryan, H.; Zokaei, S.; Farhud, D.; Keykhaei, M.; Ashrafi, M.R.; Rasulinezhad, M.; Hosseini, S.M.M.; Razmara, E.; Tavasoli, A.R. Novel Phenotype and Genotype Spectrum of WDR62 in Two Patients with Associated Primary Autosomal Recessive Microcephaly. Ir. J. Med. Sci. 2022, 191, 2733–2741. [Google Scholar] [CrossRef]
  86. Bhat, V.; Girimaji, S.C.; Mohan, G.; Arvinda, H.R.; Singhmar, P.; Duvvari, M.R.; Kumar, A. Mutations in WDR62, Encoding a Centrosomal and Nuclear Protein, in Indian Primary Microcephaly Families with Cortical Malformations. Clin. Genet. 2011, 80, 532–540. [Google Scholar] [CrossRef] [PubMed]
  87. Murdock, D.R.; Clark, G.D.; Bainbridge, M.N.; Newsham, I.; Wu, Y.-Q.; Muzny, D.M.; Cheung, S.W.; Gibbs, R.A.; Ramocki, M.B. Whole-Exome Sequencing Identifies Compound Heterozygous Mutations in WDR62 in Siblings with Recurrent Polymicrogyria. Am. J. Med. Genet. A 2011, 155A, 2071–2077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Kousar, R.; Hassan, M.J.; Khan, B.; Basit, S.; Mahmood, S.; Mir, A.; Ahmad, W.; Ansar, M. Mutations in WDR62 Gene in Pakistani Families with Autosomal Recessive Primary Microcephaly. BMC Neurol. 2011, 11, 119. [Google Scholar] [CrossRef] [Green Version]
  89. Poulton, C.J.; Schot, R.; Seufert, K.; Lequin, M.H.; Accogli, A.; Annunzio, G.D.; Villard, L.; Philip, N.; de Coo, R.; Catsman-Berrevoets, C.; et al. Severe Presentation of WDR62 Mutation: Is There a Role for Modifying Genetic Factors? Am. J. Med. Genet. A 2014, 164A, 2161–2171. [Google Scholar] [CrossRef]
  90. Bastaki, F.; Mohamed, M.; Nair, P.; Saif, F.; Tawfiq, N.; Aithala, G.; El-Halik, M.; Al-Ali, M.; Hamzeh, A.R. Novel Splice-Site Mutation in WDR62 Revealed by Whole-Exome Sequencing in a Sudanese Family with Primary Microcephaly. Congenit. Anom. 2016, 56, 135–137. [Google Scholar] [CrossRef] [Green Version]
  91. Kvarnung, M.; Taylan, F.; Nilsson, D.; Anderlid, B.-M.; Malmgren, H.; Lagerstedt-Robinson, K.; Holmberg, E.; Burstedt, M.; Nordenskjöld, M.; Nordgren, A.; et al. Genomic Screening in Rare Disorders: New Mutations and Phenotypes, Highlighting ALG14 as a Novel Cause of Severe Intellectual Disability. Clin. Genet. 2018, 94, 528–537. [Google Scholar] [CrossRef]
  92. Zombor, M.; Kalmár, T.; Nagy, N.; Berényi, M.; Telcs, B.; Maróti, Z.; Brandau, O.; Sztriha, L. A Novel WDR62 Missense Mutation in Microcephaly with Abnormal Cortical Architecture and Review of the Literature. J. Appl. Genet. 2019, 60, 151–162. [Google Scholar] [CrossRef] [Green Version]
  93. Sajid Hussain, M.; Marriam Bakhtiar, S.; Farooq, M.; Anjum, I.; Janzen, E.; Reza Toliat, M.; Eiberg, H.; Kjaer, K.W.; Tommerup, N.; Noegel, A.A.; et al. Genetic Heterogeneity in Pakistani Microcephaly Families. Clin. Genet. 2013, 83, 446–451. [Google Scholar] [CrossRef]
  94. Memon, M.M.; Raza, S.I.; Basit, S.; Kousar, R.; Ahmad, W.; Ansar, M. A Novel WDR62 Mutation Causes Primary Microcephaly in a Pakistani Family. Mol. Biol. Rep. 2013, 40, 591–595. [Google Scholar] [CrossRef] [PubMed]
  95. Farag, H.G.; Froehler, S.; Oexle, K.; Ravindran, E.; Schindler, D.; Staab, T.; Huebner, A.; Kraemer, N.; Chen, W.; Kaindl, A.M. Abnormal Centrosome and Spindle Morphology in a Patient with Autosomal Recessive Primary Microcephaly Type 2 Due to Compound Heterozygous WDR62 Gene Mutation. Orphanet J. Rare Dis. 2013, 8, 178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. McDonell, L.M.; Warman Chardon, J.; Schwartzentruber, J.; Foster, D.; Beaulieu, C.L.; FORGE Canada Consortium; Majewski, J.; Bulman, D.E.; Boycott, K.M. The Utility of Exome Sequencing for Genetic Diagnosis in a Familial Microcephaly Epilepsy Syndrome. BMC Neurol. 2014, 14, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Wang, R.; Khan, A.; Han, S.; Zhang, X. Molecular Analysis of 23 Pakistani Families with Autosomal Recessive Primary Microcephaly Using Targeted Next-Generation Sequencing. J. Hum. Genet. 2017, 62, 299–304. [Google Scholar] [CrossRef] [PubMed]
  98. Nardello, R.; Fontana, A.; Antona, V.; Beninati, A.; Mangano, G.D.; Stallone, M.C.; Mangano, S. A Novel Mutation of WDR62 Gene Associated with Severe Phenotype Including Infantile Spasm, Microcephaly, and Intellectual Disability. Brain Dev. 2018, 40, 58–64. [Google Scholar] [CrossRef]
  99. Poirier, K.; Lebrun, N.; Broix, L.; Tian, G.; Saillour, Y.; Boscheron, C.; Parrini, E.; Valence, S.; Pierre, B.S.; Oger, M.; et al. Mutations in TUBG1, DYNC1H1, KIF5C and KIF2A Cause Malformations of Cortical Development and Microcephaly. Nat. Genet. 2013, 45, 639–647. [Google Scholar] [CrossRef]
  100. Strickland, A.V.; Schabhüttl, M.; Offenbacher, H.; Synofzik, M.; Hauser, N.S.; Brunner-Krainz, M.; Gruber-Sedlmayr, U.; Moore, S.A.; Windhager, R.; Bender, B.; et al. Mutation Screen Reveals Novel Variants and Expands the Phenotypes Associated with DYNC1H1. J. Neurol. 2015, 262, 2124–2134. [Google Scholar] [CrossRef] [Green Version]
  101. Weedon, M.N.; Hastings, R.; Caswell, R.; Xie, W.; Paszkiewicz, K.; Antoniadi, T.; Williams, M.; King, C.; Greenhalgh, L.; Newbury-Ecob, R.; et al. Exome Sequencing Identifies a DYNC1H1 Mutation in a Large Pedigree with Dominant Axonal Charcot-Marie-Tooth Disease. Am. J. Hum. Genet. 2011, 89, 308–312. [Google Scholar] [CrossRef] [Green Version]
  102. Willemsen, M.H.; Vissers, L.E.L.; Willemsen, M.A.A.P.; van Bon, B.W.M.; Kroes, T.; de Ligt, J.; de Vries, B.B.; Schoots, J.; Lugtenberg, D.; Hamel, B.C.J.; et al. Mutations in DYNC1H1 Cause Severe Intellectual Disability with Neuronal Migration Defects. J. Med. Genet. 2012, 49, 179–183. [Google Scholar] [CrossRef]
  103. Beecroft, S.J.; McLean, C.A.; Delatycki, M.B.; Koshy, K.; Yiu, E.; Haliloglu, G.; Orhan, D.; Lamont, P.J.; Davis, M.R.; Laing, N.G.; et al. Expanding the Phenotypic Spectrum Associated with Mutations of DYNC1H1. Neuromuscul. Disord. 2017, 27, 607–615. [Google Scholar] [CrossRef]
  104. Niu, Q.; Wang, X.; Shi, M.; Jin, Q. A Novel DYNC1H1 Mutation Causing Spinal Muscular Atrophy with Lower Extremity Predominance. Neurol. Genet. 2015, 1, e20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ding, D.; Chen, Z.; Li, K.; Long, Z.; Ye, W.; Tang, Z.; Xia, K.; Qiu, R.; Tang, B.; Jiang, H. Identification of a de Novo DYNC1H1 Mutation via WES According to Published Guidelines. Sci. Rep. 2016, 6, 20423. [Google Scholar] [CrossRef] [Green Version]
  106. Lin, Z.; Liu, Z.; Li, X.; Li, F.; Hu, Y.; Chen, B.; Wang, Z.; Liu, Y. Whole-Exome Sequencing Identifies a Novel de Novo Mutation in DYNC1H1 in Epileptic Encephalopathies. Sci. Rep. 2017, 7, 258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Chan, S.H.S.; van Alfen, N.; Thuestad, I.J.; Ip, J.; Chan, A.O.-K.; Mak, C.; Chung, B.H.-Y.; Verrips, A.; Kamsteeg, E.-J. A Recurrent de Novo DYNC1H1 Tail Domain Mutation Causes Spinal Muscular Atrophy with Lower Extremity Predominance, Learning Difficulties and Mild Brain Abnormality. Neuromuscul. Disord. 2018, 28, 750–756. [Google Scholar] [CrossRef] [PubMed]
  108. Gelineau-Morel, R.; Lukacs, M.; Weaver, K.N.; Hufnagel, R.B.; Gilbert, D.L.; Stottmann, R.W. Congenital Cataracts and Gut Dysmotility in a DYNC1H1 Dyneinopathy Patient. Genes 2016, 7, 85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Chen, Y.; Xu, Y.; Li, G.; Li, N.; Yu, T.; Yao, R.-E.; Wang, X.; Shen, Y.; Wang, J. Exome Sequencing Identifies De Novo DYNC1H1 Mutations Associated With Distal Spinal Muscular Atrophy and Malformations of Cortical Development. J. Child Neurol. 2017, 32, 379–386. [Google Scholar] [CrossRef]
  110. Laquerriere, A.; Maillard, C.; Cavallin, M.; Chapon, F.; Marguet, F.; Molin, A.; Sigaudy, S.; Blouet, M.; Benoist, G.; Fernandez, C.; et al. Neuropathological Hallmarks of Brain Malformations in Extreme Phenotypes Related to DYNC1H1 Mutations. J. Neuropathol. Exp. Neurol. 2017, 76, 195–205. [Google Scholar] [CrossRef]
  111. Hertecant, J.; Komara, M.; Nagi, A.; Suleiman, J.; Al-Gazali, L.; Ali, B.R. A Novel de Novo Mutation in DYNC1H1 Gene Underlying Malformation of Cortical Development and Cataract. Meta Gene 2016, 9, 124–127. [Google Scholar] [CrossRef]
  112. Peeters, K.; Bervoets, S.; Chamova, T.; Litvinenko, I.; De Vriendt, E.; Bichev, S.; Kancheva, D.; Mitev, V.; Kennerson, M.; Timmerman, V.; et al. Novel Mutations in the DYNC1H1 Tail Domain Refine the Genetic and Clinical Spectrum of Dyneinopathies. Hum. Mutat. 2015, 36, 287–291. [Google Scholar] [CrossRef]
  113. Tsurusaki, Y.; Saitoh, S.; Tomizawa, K.; Sudo, A.; Asahina, N.; Shiraishi, H.; Ito, J.-I.; Tanaka, H.; Doi, H.; Saitsu, H.; et al. A DYNC1H1 Mutation Causes a Dominant Spinal Muscular Atrophy with Lower Extremity Predominance. Neurogenetics 2012, 13, 327–332. [Google Scholar] [CrossRef]
  114. Das, J.; Lilleker, J.B.; Jabbal, K.; Ealing, J. A Missense Mutation in DYNC1H1 Gene Causing Spinal Muscular Atrophy–Lower Extremity, Dominant. Neurol. Neurochir. Pol. 2018, 52, 293–297. [Google Scholar] [CrossRef] [PubMed]
  115. Punetha, J.; Monges, S.; Franchi, M.E.; Hoffman, E.P.; Cirak, S.; Tesi-Rocha, C. Exome Sequencing Identifies DYNC1H1 Variant Associated With Vertebral Abnormality and Spinal Muscular Atrophy With Lower Extremity Predominance. Pediatr. Neurol. 2015, 52, 239–244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Harms, M.B.; Ori-McKenney, K.M.; Scoto, M.; Tuck, E.P.; Bell, S.; Ma, D.; Masi, S.; Allred, P.; Al-Lozi, M.; Reilly, M.M.; et al. Mutations in the Tail Domain of DYNC1H1 Cause Dominant Spinal Muscular Atrophy. Neurology 2012, 78, 1714–1720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Zillhardt, J.L.; Poirier, K.; Broix, L.; Lebrun, N.; Elmorjani, A.; Martinovic, J.; Saillour, Y.; Muraca, G.; Nectoux, J.; Bessieres, B.; et al. Mosaic Parental Germline Mutations Causing Recurrent Forms of Malformations of Cortical Development. Eur. J. Hum. Genet. 2016, 24, 611–614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Fiorillo, C.; Moro, F.; Yi, J.; Weil, S.; Brisca, G.; Astrea, G.; Severino, M.; Romano, A.; Battini, R.; Rossi, A.; et al. Novel Dynein DYNC1H1 Neck and Motor Domain Mutations Link Distal Spinal Muscular Atrophy and Abnormal Cortical Development. Hum. Mutat. 2014, 35, 298–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Scoto, M.; Rossor, A.M.; Harms, M.B.; Cirak, S.; Calissano, M.; Robb, S.; Manzur, A.Y.; Martínez Arroyo, A.; Rodriguez Sanz, A.; Mansour, S.; et al. Novel Mutations Expand the Clinical Spectrum of DYNC1H1-Associated Spinal Muscular Atrophy. Neurology 2015, 84, 668–679. [Google Scholar] [CrossRef] [Green Version]
  120. Jamuar, S.S.; Lam, A.-T.N.; Kircher, M.; D’Gama, A.M.; Wang, J.; Barry, B.J.; Zhang, X.; Hill, R.S.; Partlow, J.N.; Rozzo, A.; et al. Somatic Mutations in Cerebral Cortical Malformations. N. Engl. J. Med. 2014, 371, 733–743. [Google Scholar] [CrossRef] [Green Version]
  121. Becker, L.-L.; Dafsari, H.S.; Schallner, J.; Abdin, D.; Seifert, M.; Petit, F.; Smol, T.; Bok, L.; Rodan, L.; Krapels, I.; et al. The Clinical-Phenotype Continuum in DYNC1H1-Related Disorders-Genomic Profiling and Proposal for a Novel Classification. J. Hum. Genet. 2020, 65, 1003–1017. [Google Scholar] [CrossRef]
  122. Harms, M.B.; Allred, P.; Gardner, R.; Fernandes Filho, J.A.; Florence, J.; Pestronk, A.; Al-Lozi, M.; Baloh, R.H. Dominant Spinal Muscular Atrophy with Lower Extremity Predominance: Linkage to 14q32. Neurology 2010, 75, 539–546. [Google Scholar] [CrossRef]
  123. Hoang, H.T.; Schlager, M.A.; Carter, A.P.; Bullock, S.L. DYNC1H1 Mutations Associated with Neurological Diseases Compromise Processivity of Dynein-Dynactin-Cargo Adaptor Complexes. Proc. Natl. Acad. Sci. USA 2017, 114, E1597–E1606. [Google Scholar] [CrossRef]
  124. Lawo, S.; Hasegan, M.; Gupta, G.D.; Pelletier, L. Subdiffraction Imaging of Centrosomes Reveals Higher-Order Organizational Features of Pericentriolar Material. Nat. Cell Biol. 2012, 14, 1148–1158. [Google Scholar] [CrossRef] [PubMed]
  125. Woodruff, J.B.; Wueseke, O.; Hyman, A.A. Pericentriolar Material Structure and Dynamics. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Luo, Y.; Pelletier, L. Pericentrin: Critical for Spindle Orientation. Curr. Biol. 2014, 24, R962–R964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Chen, C.-T.; Hehnly, H.; Yu, Q.; Farkas, D.; Zheng, G.; Redick, S.D.; Hung, H.-F.; Samtani, R.; Jurczyk, A.; Akbarian, S.; et al. A Unique Set of Centrosome Proteins Requires Pericentrin for Spindle-Pole Localization and Spindle Orientation. Curr. Biol. 2014, 24, 2327–2334. [Google Scholar] [CrossRef] [Green Version]
  128. Gavilan, M.P.; Gandolfo, P.; Balestra, F.R.; Arias, F.; Bornens, M.; Rios, R.M. The Dual Role of the Centrosome in Organizing the Microtubule Network in Interphase. EMBO Rep. 2018, 19, e45942. [Google Scholar] [CrossRef]
  129. Pimenta-Marques, A.; Bettencourt-Dias, M. Pericentriolar Material. Curr. Biol. 2020, 30, R687–R689. [Google Scholar] [CrossRef]
  130. Waich, S.; Janecke, A.R.; Parson, W.; Greber-Platzer, S.; Müller, T.; Huber, L.A.; Valovka, T.; Vodopiutz, J. Novel PCNT Variants in MOPDII with Attenuated Growth Restriction and Pachygyria. Clin. Genet. 2020, 98, 282–287. [Google Scholar] [CrossRef]
  131. Lorenzo-Betancor, O.; Blackburn, P.R.; Edwards, E.; Vázquez-do-Campo, R.; Klee, E.W.; Labbé, C.; Hodges, K.; Glover, P.; Sigafoos, A.N.; Soto, A.I.; et al. PCNT Point Mutations and Familial Intracranial Aneurysms. Neurology 2018, 91, e2170–e2181. [Google Scholar] [CrossRef]
  132. Liu, H.; Tao, N.; Wang, Y.; Yang, Y.; He, X.; Zhang, Y.; Zhou, Y.; Liu, X.; Feng, X.; Sun, M.; et al. A Novel Homozygous Mutation of the PCNT Gene in a Chinese Patient with Microcephalic Osteodysplastic Primordial Dwarfism Type II. Mol. Genet. Genom. Med. 2021, 9, e1761. [Google Scholar] [CrossRef]
  133. Kantaputra, P.; Tanpaiboon, P.; Porntaveetus, T.; Ohazama, A.; Sharpe, P.; Rauch, A.; Hussadaloy, A.; Thiel, C.T. The Smallest Teeth in the World Are Caused by Mutations in the PCNT Gene. Am. J. Med. Genet. A 2011, 155A, 1398–1403. [Google Scholar] [CrossRef]
  134. Dehghan Tezerjani, M.; Vahidi Mehrjardi, M.Y.; Hozhabri, H.; Rahmanian, M. A Novel PCNT Frame Shift Variant (c.7511delA) Causing Osteodysplastic Primordial Dwarfism of Majewski Type 2 (MOPD II). Front. Pediatr. 2020, 8, 340. [Google Scholar] [CrossRef] [PubMed]
  135. Willems, M.; Geneviève, D.; Borck, G.; Baumann, C.; Baujat, G.; Bieth, E.; Edery, P.; Farra, C.; Gerard, M.; Héron, D.; et al. Molecular Analysis of Pericentrin Gene (PCNT) in a Series of 24 Seckel/Microcephalic Osteodysplastic Primordial Dwarfism Type II (MOPD II) Families. J. Med. Genet. 2010, 47, 797–802. [Google Scholar] [CrossRef] [Green Version]
  136. Piane, M.; Della Monica, M.; Piatelli, G.; Lulli, P.; Lonardo, F.; Chessa, L.; Scarano, G. Majewski Osteodysplastic Primordial Dwarfism Type II (MOPD II) Syndrome Previously Diagnosed as Seckel Syndrome: Report of a Novel Mutation of the PCNT Gene. Am. J. Med. Genet. A 2009, 149A, 2452–2456. [Google Scholar] [CrossRef] [PubMed]
  137. Pachajoa, H.; Ruiz-Botero, F.; Isaza, C. A New Mutation of the PCNT Gene in a Colombian Patient with Microcephalic Osteodysplastic Primordial Dwarfism Type II: A Case Report. J. Med. Case Rep. 2014, 8, 191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Müller, E.; Dunstheimer, D.; Klammt, J.; Friebe, D.; Kiess, W.; Kratzsch, J.; Kruis, T.; Laue, S.; Pfäffle, R.; Wallborn, T.; et al. Clinical and Functional Characterization of a Patient Carrying a Compound Heterozygous Pericentrin Mutation and a Heterozygous IGF1 Receptor Mutation. PLoS ONE 2012, 7, e38220. [Google Scholar] [CrossRef]
  139. Bober, M.B.; Niiler, T.; Duker, A.L.; Murray, J.E.; Ketterer, T.; Harley, M.E.; Alvi, S.; Flora, C.; Rustad, C.; Bongers, E.M.H.F.; et al. Growth in Individuals with Majewski Osteodysplastic Primordial Dwarfism Type II Caused by Pericentrin Mutations. Am. J. Med. Genet. A 2012, 158A, 2719–2725. [Google Scholar] [CrossRef]
  140. Unal, S.; Alanay, Y.; Cetin, M.; Boduroglu, K.; Utine, E.; Cormier-Daire, V.; Huber, C.; Ozsurekci, Y.; Kilic, E.; Simsek Kiper, O.P.; et al. Striking Hematological Abnormalities in Patients with Microcephalic Osteodysplastic Primordial Dwarfism Type II (MOPD II): A Potential Role of Pericentrin in Hematopoiesis. Pediatr. Blood Cancer 2014, 61, 302–305. [Google Scholar] [CrossRef]
  141. Dieks, J.-K.; Baumer, A.; Wilichowski, E.; Rauch, A.; Sigler, M. Microcephalic Osteodysplastic Primordial Dwarfism Type II (MOPD II) with Multiple Vascular Complications Misdiagnosed as Dubowitz Syndrome. Eur. J. Pediatr. 2014, 173, 1253–1256. [Google Scholar] [CrossRef]
  142. Li, F.-F.; Wang, X.-D.; Zhu, M.-W.; Lou, Z.-H.; Zhang, Q.; Zhu, C.-Y.; Feng, H.-L.; Lin, Z.-G.; Liu, S.-L. Identification of Two Novel Critical Mutations in PCNT Gene Resulting in Microcephalic Osteodysplastic Primordial Dwarfism Type II Associated with Multiple Intracranial Aneurysms. Metab. Brain Dis. 2015, 30, 1387–1394. [Google Scholar] [CrossRef]
  143. Weiss, K.; Ekhilevitch, N.; Cohen, L.; Bratman-Morag, S.; Bello, R.; Martinez, A.F.; Hadid, Y.; Shlush, L.I.; Kurolap, A.; Paperna, T.; et al. Identification of a Novel PCNT Founder Pathogenic Variant in the Israeli Druze Population. Eur. J. Med. Genet. 2020, 63, 103643. [Google Scholar] [CrossRef]
  144. Abdel-Salam, G.M.H.; Sayed, I.S.M.; Afifi, H.H.; Abdel-Ghafar, S.F.; Abouzaid, M.R.; Ismail, S.I.; Aglan, M.S.; Issa, M.Y.; El-Bassyouni, H.T.; El-Kamah, G.; et al. Microcephalic Osteodysplastic Primordial Dwarfism Type II: Additional Nine Patients with Implications on Phenotype and Genotype Correlation. Am. J. Med. Genet. A 2020, 182, 1407–1420. [Google Scholar] [CrossRef] [PubMed]
  145. Rossi-Espagnet, M.C.; Dentici, M.L.; Pasquini, L.; Carducci, C.; Lucignani, M.; Longo, D.; Agolini, E.; Novelli, A.; Gonfiantini, M.V.; Digilio, M.C.; et al. Microcephalic Osteodysplastic Primordial Dwarfism Type II and Pachygyria: Morphometric Analysis in a 2-Year-Old Girl. Am. J. Med. Genet. A 2020, 182, 2372–2376. [Google Scholar] [CrossRef] [PubMed]
  146. Ma, Y.; Xu, Z.; Zhao, J.; Shen, H. Novel Compound Heterozygous Mutations of PCNT Gene in MOPD Type II with Central Precocious Puberty. Gynecol. Endocrinol. 2021, 37, 190–192. [Google Scholar] [CrossRef]
  147. Shaheen, R.; Maddirevula, S.; Ewida, N.; Alsahli, S.; Abdel-Salam, G.M.H.; Zaki, M.S.; Tala, S.A.; Alhashem, A.; Softah, A.; Al-Owain, M.; et al. Genomic and Phenotypic Delineation of Congenital Microcephaly. Genet. Med. 2019, 21, 545–552. [Google Scholar] [CrossRef] [PubMed]
  148. Alrajhi, H.; Alallah, J.; Shawli, A.; Alghamdi, K.; Hakami, F. Majewski Dwarfism Type II: An Atypical Neuroradiological Presentation with a Novel Variant in the PCNT Gene. BMJ Case Rep. 2019, 12, e224197. [Google Scholar] [CrossRef] [PubMed]
  149. Meng, L.; Tu, C.; Lu, G.; Lin, G.; Tan, Y. Novel Biallelic PCNT Deletion Causing Microcephalic Osteodysplastic Primordial Dwarfism Type II with Congenital Heart Defect. Sci. China Life Sci. 2019, 62, 144–147. [Google Scholar] [CrossRef] [PubMed]
  150. Huang-Doran, I.; Bicknell, L.S.; Finucane, F.M.; Rocha, N.; Porter, K.M.; Tung, Y.C.L.; Szekeres, F.; Krook, A.; Nolan, J.J.; O’Driscoll, M.; et al. Genetic Defects in Human Pericentrin Are Associated with Severe Insulin Resistance and Diabetes. Diabetes 2011, 60, 925–935. [Google Scholar] [CrossRef] [Green Version]
  151. Chung, H.; Kim, S.Y.; Kang, J.; Phi, J.H.; Kim, W.-H.; Yang, S.W.; Kwon, H.W.; Lee, S.Y.; Kim, G.B.; Bae, E.J.; et al. Siblings With Familial Dwarfism Presenting With Acute Myocardial Infarction at Adolescence. JACC Case Rep. 2021, 3, 795–800. [Google Scholar] [CrossRef]
  152. Eslava, A.; Garcia-Puig, M.; Corripio, R. A 10-Year-Old Boy with Short Stature and Microcephaly, Diagnosed with Moyamoya Syndrome and Microcephalic Osteodysplastic Primordial Dwarfism Type II (MOPD II). Am. J. Case Rep. 2021, 22, e933919. [Google Scholar] [CrossRef]
  153. Hettiarachchi, D.; Subasinghe, S.M.V.; Anandagoda, G.G.; Panchal, H.; Lai, P.S.; Dissanayake, V.H.W. Novel Frameshift Variant in the PCNT Gene Associated with Microcephalic Osteodysplastic Primordial Dwarfism (MOPD) Type II and Small Kidneys. BMC Med. Genom. 2022, 15, 82. [Google Scholar] [CrossRef]
  154. Rauch, A. The Shortest of the Short: Pericentrin Mutations and Beyond. Best Pract. Res. Clin. Endocrinol. Metab. 2011, 25, 125–130. [Google Scholar] [CrossRef] [PubMed]
  155. Nguyen, T.H.; Nguyen, N.-L.; Vu, C.D.; Ngoc, C.T.B.; Nguyen, N.K.; Nguyen, H.H. Identification of Three Novel Mutations in PCNT in Vietnamese Patients with Microcephalic Osteodysplastic Primordial Dwarfism Type II. Genes Genom. 2021, 43, 115–121. [Google Scholar] [CrossRef] [PubMed]
  156. Huang, J.; Zhou, D.; Dong, N.; Ding, C.; Liu, Y.; Li, F. Clinical and Genetic Analysis of a Patient With Coexisting 17a-Hydroxylase/17,20-Lyase Deficiency and Moyamoya Disease. Front. Genet. 2022, 13, 845016. [Google Scholar] [CrossRef] [PubMed]
  157. Graser, S.; Stierhof, Y.D.; Nigg, E.A. Cep68 and Cep215 (Cdk5rap2) Are Required for Centrosome Cohesion. J. Cell Sci. 2007, 120, 4321–4331. [Google Scholar] [CrossRef] [Green Version]
  158. Fong, K.W.; Choi, Y.K.; Rattner, J.B.; Qi, R.Z. CDK5RAP2 Is a Pericentriolar Protein That Functions in Centrosomal Attachment of the {gamma}-Tubulin Ring Complex. Mol. Biol. Cell 2008, 19, 115–125. [Google Scholar] [CrossRef] [Green Version]
  159. Barr, A.R.; Kilmartin, J.V.; Gergely, F. CDK5RAP2 Functions in Centrosome to Spindle Pole Attachment and DNA Damage Response. J. Cell Biol. 2010, 189, 23–39. [Google Scholar] [CrossRef] [Green Version]
  160. Lee, K.S.; Park, J.-E.; Ahn, J.I.; Zeng, Y. Constructing PCM with Architecturally Distinct Higher-Order Assemblies. Curr. Opin. Struct. Biol. 2021, 66, 66–73. [Google Scholar] [CrossRef]
  161. Watanabe, S.; Meitinger, F.; Shiau, A.K.; Oegema, K.; Desai, A. Centriole-Independent Mitotic Spindle Assembly Relies on the PCNT-CDK5RAP2 Pericentriolar Matrix. J. Cell Biol. 2020, 219, e202006010. [Google Scholar] [CrossRef]
  162. Yigit, G.; Brown, K.E.; Kayserili, H.; Pohl, E.; Caliebe, A.; Zahnleiter, D.; Rosser, E.; Bogershausen, N.; Uyguner, Z.O.; Altunoglu, U.; et al. Mutations in CDK5RAP2 Cause Seckel Syndrome. Mol. Genet. Genom. Med. 2015, 3, 467–480. [Google Scholar] [CrossRef]
  163. Lancaster, M.A.; Renner, M.; Martin, C.A.; Wenzel, D.; Bicknell, L.S.; Hurles, M.E.; Homfray, T.; Penninger, J.M.; Jackson, A.P.; Knoblich, J.A. Cerebral Organoids Model Human Brain Development and Microcephaly. Nature 2013, 501, 373–379. [Google Scholar] [CrossRef] [Green Version]
  164. Alfares, A.; Alhufayti, I.; Alsubaie, L.; Alowain, M.; Almass, R.; Alfadhel, M.; Kaya, N.; Eyaid, W. A New Association between CDK5RAP2 Microcephaly and Congenital Cataracts. Ann. Hum. Genet. 2018, 82, 165–170. [Google Scholar] [CrossRef]
  165. Moynihan, L.; Jackson, A.P.; Roberts, E.; Karbani, G.; Lewis, I.; Corry, P.; Turner, G.; Mueller, R.F.; Lench, N.J.; Woods, C.G. A Third Novel Locus for Primary Autosomal Recessive Microcephaly Maps to Chromosome 9q34. Am. J. Hum. Genet. 2000, 66, 724–727. [Google Scholar] [CrossRef] [Green Version]
  166. Hassan, M.J.; Khurshid, M.; Azeem, Z.; John, P.; Ali, G.; Chishti, M.S.; Ahmad, W. Previously Described Sequence Variant in CDK5RAP2 Gene in a Pakistani Family with Autosomal Recessive Primary Microcephaly. BMC Med. Genet. 2007, 8, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Issa, L.; Mueller, K.; Seufert, K.; Kraemer, N.; Rosenkotter, H.; Ninnemann, O.; Buob, M.; Kaindl, A.M.; Morris-Rosendahl, D.J. Clinical and Cellular Features in Patients with Primary Autosomal Recessive Microcephaly and a Novel CDK5RAP2 Mutation. Orphanet J. Rare Dis. 2013, 8, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Jouan, L.; Ouled Amar Bencheikh, B.; Daoud, H.; Dionne-Laporte, A.; Dobrzeniecka, S.; Spiegelman, D.; Rochefort, D.; Hince, P.; Szuto, A.; Lassonde, M.; et al. Exome Sequencing Identifies Recessive CDK5RAP2 Variants in Patients with Isolated Agenesis of Corpus Callosum. Eur. J. Hum. Genet. 2016, 24, 607–610. [Google Scholar] [CrossRef] [PubMed]
  169. Pagnamenta, A.T.; Murray, J.E.; Yoon, G.; Sadighi Akha, E.; Harrison, V.; Bicknell, L.S.; Ajilogba, K.; Stewart, H.; Kini, U.; Taylor, J.C.; et al. A Novel Nonsense CDK5RAP2 Mutation in a Somali Child with Primary Microcephaly and Sensorineural Hearing Loss. Am. J. Med. Genet. A 2012, 158A, 2577–2582. [Google Scholar] [CrossRef] [Green Version]
  170. Tan, C.A.; Topper, S.; Ward Melver, C.; Stein, J.; Reeder, A.; Arndt, K.; Das, S. The First Case of CDK5RAP2-Related Primary Microcephaly in a Non-Consanguineous Patient Identified by next Generation Sequencing. Brain Dev. 2014, 36, 351–355. [Google Scholar] [CrossRef] [PubMed]
  171. Kakar, N.; Ahmad, J.; Morris-Rosendahl, D.J.; Altmüller, J.; Friedrich, K.; Barbi, G.; Nürnberg, P.; Kubisch, C.; Dobyns, W.B.; Borck, G. STIL Mutation Causes Autosomal Recessive Microcephalic Lobar Holoprosencephaly. Hum. Genet. 2015, 134, 45–51. [Google Scholar] [CrossRef] [PubMed]
  172. Mouden, C.; de Tayrac, M.; Dubourg, C.; Rose, S.; Carré, W.; Hamdi-Rozé, H.; Babron, M.-C.; Akloul, L.; Héron-Longe, B.; Odent, S.; et al. Homozygous STIL Mutation Causes Holoprosencephaly and Microcephaly in Two Siblings. PLoS ONE 2015, 10, e0117418. [Google Scholar] [CrossRef] [PubMed]
  173. Hatch, E.M.; Kulukian, A.; Holland, A.J.; Cleveland, D.W.; Stearns, T. Cep152 Interacts with Plk4 and Is Required for Centriole Duplication. J. Cell Biol. 2010, 191, 721–729. [Google Scholar] [CrossRef] [Green Version]
  174. Cizmecioglu, O.; Arnold, M.; Bahtz, R.; Settele, F.; Ehret, L.; Haselmann-Weiss, U.; Antony, C.; Hoffmann, I. Cep152 Acts as a Scaffold for Recruitment of Plk4 and CPAP to the Centrosome. J. Cell Biol. 2010, 191, 731–739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Gönczy, P.; Hatzopoulos, G.N. Centriole Assembly at a Glance. J. Cell Sci. 2019, 132, jcs228833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Sonnen, K.F.; Gabryjonczyk, A.-M.; Anselm, E.; Stierhof, Y.-D.; Nigg, E.A. Human Cep192 and Cep152 Cooperate in Plk4 Recruitment and Centriole Duplication. J. Cell Sci. 2013, 126, 3223–3233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Avidor-Reiss, T.; Gopalakrishnan, J. Building a Centriole. Curr. Opin. Cell Biol. 2013, 25, 72–77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Loncarek, J.; Bettencourt-Dias, M. Building the Right Centriole for Each Cell Type. J. Cell Biol. 2018, 217, 823–835. [Google Scholar] [CrossRef]
  179. Breslow, D.K.; Holland, A.J. Mechanism and Regulation of Centriole and Cilium Biogenesis. Annu. Rev. Biochem. 2019, 88, 691–724. [Google Scholar] [CrossRef]
  180. Zhang, L.; Teng, Y.; Hu, H.; Zhu, H.; Wen, J.; Liang, D.; Li, Z.; Wu, L. Two Novel Variants in CEP152 Caused Seckel Syndrome 5 in a Chinese Family. Front. Genet. 2022, 13, 1052915. [Google Scholar] [CrossRef]
  181. Tsutsumi, M.; Yokoi, S.; Miya, F.; Miyata, M.; Kato, M.; Okamoto, N.; Tsunoda, T.; Yamasaki, M.; Kanemura, Y.; Kosaki, K.; et al. Novel Compound Heterozygous Variants in PLK4 Identified in a Patient with Autosomal Recessive Microcephaly and Chorioretinopathy. Eur. J. Hum. Genet. 2016, 24, 1702–1706. [Google Scholar] [CrossRef] [Green Version]
  182. Martín-Rivada, Á.; Pozo-Román, J.; Güemes, M.; Ortiz-Cabrera, N.V.; Pérez-Jurado, L.A.; Argente, J. Primary Dwarfism, Microcephaly, and Chorioretinopathy Due to a PLK4 Mutation in Two Siblings. Horm. Res. Paediatr. 2020, 93, 567–572. [Google Scholar] [CrossRef]
  183. Dinçer, T.; Yorgancıoğlu-Budak, G.; Ölmez, A.; Er, İ.; Dodurga, Y.; Özdemir, Ö.M.; Toraman, B.; Yıldırım, A.; Sabir, N.; Akarsu, N.A.; et al. Analysis of Centrosome and DNA Damage Response in PLK4 Associated Seckel Syndrome. Eur. J. Hum. Genet. 2017, 25, 1118–1125. [Google Scholar] [CrossRef]
  184. Neitzel, H.; Varon, R.; Chughtai, S.; Dartsch, J.; Dutrannoy-Tönsing, V.; Nürnberg, P.; Nürnberg, G.; Schweiger, M.; Digweed, M.; Hildebrand, G.; et al. Transmission Ratio Distortion of Mutations in the Master Regulator of Centriole Biogenesis PLK4. Hum. Genet. 2022, 141, 1785–1794. [Google Scholar] [CrossRef]
  185. Bornens, M. Centrosome Organization and Functions. Curr. Opin. Struct. Biol. 2021, 66, 199–206. [Google Scholar] [CrossRef] [PubMed]
  186. Akhmanova, A.; Steinmetz, M.O. Microtubule Minus-End Regulation at a Glance. J. Cell Sci. 2019, 132, jcs227850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Wu, J.; Akhmanova, A. Microtubule-Organizing Centers. Annu. Rev. Cell Dev. Biol. 2017, 33, 51–75. [Google Scholar] [CrossRef]
  188. Borgal, L.; Wakefield, J.G. Context-Dependent Spindle Pole Focusing. Essays Biochem. 2018, 62, 803–813. [Google Scholar] [CrossRef] [PubMed]
  189. Fraschini, R. Factors That Control Mitotic Spindle Dynamics. Adv. Exp. Med. Biol. 2017, 925, 89–101. [Google Scholar] [CrossRef]
  190. Goundiam, O.; Basto, R. Centrosomes in Disease: How the Same Music Can Sound so Different? Curr. Opin. Struct. Biol. 2021, 66, 74–82. [Google Scholar] [CrossRef] [PubMed]
  191. Phan, T.P.; Holland, A.J. Time Is of the Essence: The Molecular Mechanisms of Primary Microcephaly. Genes Dev. 2021, 35, 1551–1578. [Google Scholar] [CrossRef]
  192. Gavvovidis, I.; Rost, I.; Trimborn, M.; Kaiser, F.J.; Purps, J.; Wiek, C.; Hanenberg, H.; Neitzel, H.; Schindler, D. A Novel MCPH1 Isoform Complements the Defective Chromosome Condensation of Human MCPH1-Deficient Cells. PLoS ONE 2012, 7, e40387. [Google Scholar] [CrossRef] [Green Version]
  193. Kraemer, N.; Issa-Jahns, L.; Neubert, G.; Ravindran, E.; Mani, S.; Ninnemann, O.; Kaindl, A.M. Novel Alternative Splice Variants of Mouse Cdk5rap2. PLoS ONE 2015, 10, e0136684. [Google Scholar] [CrossRef] [Green Version]
  194. Zhang, X.; Chen, M.H.; Wu, X.; Kodani, A.; Fan, J.; Doan, R.; Ozawa, M.; Ma, J.; Yoshida, N.; Reiter, J.F.; et al. Cell-Type-Specific Alternative Splicing Governs Cell Fate in the Developing Cerebral Cortex. Cell 2016, 166, 1147–1162.e15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Knouse, K.A.; Lopez, K.E.; Bachofner, M.; Amon, A. Chromosome Segregation Fidelity in Epithelia Requires Tissue Architecture. Cell 2018, 175, 200–211.e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Gilbert, S.L.; Dobyns, W.B.; Lahn, B.T. Genetic Links between Brain Development and Brain Evolution. Nat. Rev. Genet. 2005, 6, 581–590. [Google Scholar] [CrossRef]
  197. Mekel-Bobrov, N.; Gilbert, S.L.; Evans, P.D.; Vallender, E.J.; Anderson, J.R.; Hudson, R.R.; Tishkoff, S.A.; Lahn, B.T. Ongoing Adaptive Evolution of ASPM, a Brain Size Determinant in Homo Sapiens. Science 2005, 309, 1720–1722. [Google Scholar] [CrossRef] [PubMed]
  198. Evans, P.D.; Anderson, J.R.; Vallender, E.J.; Gilbert, S.L.; Malcom, C.M.; Dorus, S.; Lahn, B.T. Adaptive Evolution of ASPM, a Major Determinant of Cerebral Cortical Size in Humans. Hum. Mol. Genet. 2004, 13, 489–494. [Google Scholar] [CrossRef] [Green Version]
  199. Evans, P.D.; Gilbert, S.L.; Mekel-Bobrov, N.; Vallender, E.J.; Anderson, J.R.; Vaez-Azizi, L.M.; Tishkoff, S.A.; Hudson, R.R.; Lahn, B.T. Microcephalin, a Gene Regulating Brain Size, Continues to Evolve Adaptively in Humans. Science 2005, 309, 1717–1720. [Google Scholar] [CrossRef]
  200. Timpson, N.; Heron, J.; Smith, G.D.; Enard, W. Comment on Papers by Evans et al. and Mekel-Bobrov et al. on Evidence for Positive Selection of MCPH1 and ASPM. Science 2007, 317, 1036, author reply 1036. [Google Scholar] [CrossRef] [Green Version]
  201. Mekel-Bobrov, N.; Posthuma, D.; Gilbert, S.L.; Lind, P.; Gosso, M.F.; Luciano, M.; Harris, S.E.; Bates, T.C.; Polderman, T.J.C.; Whalley, L.J.; et al. The Ongoing Adaptive Evolution of ASPM and Microcephalin Is Not Explained by Increased Intelligence. Hum. Mol. Genet. 2007, 16, 600–608. [Google Scholar] [CrossRef] [Green Version]
  202. Wang, J.K.; Li, Y.; Su, B. A Common SNP of MCPH1 Is Associated with Cranial Volume Variation in Chinese Population. Hum. Mol. Genet. 2008, 17, 1329–1335. [Google Scholar] [CrossRef] [Green Version]
  203. Dobson-Stone, C.; Gatt, J.M.; Kuan, S.A.; Grieve, S.M.; Gordon, E.; Williams, L.M.; Schofield, P.R. Investigation of MCPH1 G37995C and ASPM A44871G Polymorphisms and Brain Size in a Healthy Cohort. NeuroImage 2007, 37, 394–400. [Google Scholar] [CrossRef]
  204. Johnson, M.B.; Sun, X.; Kodani, A.; Borges-Monroy, R.; Girskis, K.M.; Ryu, S.C.; Wang, P.P.; Patel, K.; Gonzalez, D.M.; Woo, Y.M.; et al. Aspm Knockout Ferret Reveals an Evolutionary Mechanism Governing Cerebral Cortical Size. Nature 2018, 556, 370–375. [Google Scholar] [CrossRef] [PubMed]
  205. Fish, J.L.; Kosodo, Y.; Enard, W.; Paabo, S.; Huttner, W.B. Aspm Specifically Maintains Symmetric Proliferative Divisions of Neuroepithelial Cells. Proc. Natl. Acad. Sci. USA 2006, 103, 10438–10443. [Google Scholar] [CrossRef] [PubMed]
  206. Pulvers, J.N.; Bryk, J.; Fish, J.L.; Wilsch-Brauninger, M.; Arai, Y.; Schreier, D.; Naumann, R.; Helppi, J.; Habermann, B.; Vogt, J.; et al. Mutations in Mouse Aspm (Abnormal Spindle-like Microcephaly Associated) Cause Not Only Microcephaly but Also Major Defects in the Germline. Proc. Natl. Acad. Sci. USA 2010, 107, 16595–16600. [Google Scholar] [CrossRef] [PubMed]
  207. Lizarraga, S.B.; Margossian, S.P.; Harris, M.H.; Campagna, D.R.; Han, A.P.; Blevins, S.; Mudbhary, R.; Barker, J.E.; Walsh, C.A.; Fleming, M.D. Cdk5rap2 Regulates Centrosome Function and Chromosome Segregation in Neuronal Progenitors. Development 2010, 137, 1907–1917. [Google Scholar] [CrossRef]
  208. Marthiens, V.; Rujano, M.A.; Pennetier, C.; Tessier, S.; Paul-Gilloteaux, P.; Basto, R. Centrosome Amplification Causes Microcephaly. Nat. Cell Biol. 2013, 15, 731–740. [Google Scholar] [CrossRef]
  209. Nano, M.; Basto, R. Consequences of Centrosome Dysfunction during Brain Development. Adv. Exp. Med. Biol. 2017, 1002, 19–45. [Google Scholar] [CrossRef]
  210. Lin, S.Y.; Rai, R.; Li, K.; Xu, Z.X.; Elledge, S.J. BRIT1/MCPH1 Is a DNA Damage Responsive Protein That Regulates the Brca1-Chk1 Pathway, Implicating Checkpoint Dysfunction in Microcephaly. Proc. Natl. Acad. Sci. USA 2005, 102, 15105–15109. [Google Scholar] [CrossRef]
  211. Alderton, G.K.; Galbiati, L.; Griffith, E.; Surinya, K.H.; Neitzel, H.; Jackson, A.P.; Jeggo, P.A.; O’Driscoll, M. Regulation of Mitotic Entry by Microcephalin and Its Overlap with ATR Signalling. Nat. Cell Biol. 2006, 8, 725–733. [Google Scholar] [CrossRef]
  212. Gruber, R.; Zhou, Z.; Sukchev, M.; Joerss, T.; Frappart, P.O.; Wang, Z.Q. MCPH1 Regulates the Neuroprogenitor Division Mode by Coupling the Centrosomal Cycle with Mitotic Entry through the Chk1-Cdc25 Pathway. Nat. Cell Biol. 2011, 13, 1325–1334. [Google Scholar] [CrossRef]
  213. Shao, Z.; Li, F.; Sy, S.M.-H.; Yan, W.; Zhang, Z.; Gong, D.; Wen, B.; Huen, M.S.Y.; Gong, Q.; Wu, J.; et al. Specific Recognition of Phosphorylated Tail of H2AX by the Tandem BRCT Domains of MCPH1 Revealed by Complex Structure. J. Struct. Biol. 2012, 177, 459–468. [Google Scholar] [CrossRef]
  214. Singh, N.; Wiltshire, T.D.; Thompson, J.R.; Mer, G.; Couch, F.J. Molecular Basis for the Association of Microcephalin (MCPH1) Protein with the Cell Division Cycle Protein 27 (Cdc27) Subunit of the Anaphase-Promoting Complex. J. Biol. Chem. 2012, 287, 2854–2862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Journiac, N.; Gilabert-Juan, J.; Cipriani, S.; Benit, P.; Liu, X.; Jacquier, S.; Faivre, V.; Delahaye-Duriez, A.; Csaba, Z.; Hourcade, T.; et al. Cell Metabolic Alterations Due to Mcph1 Mutation in Microcephaly. Cell Rep. 2020, 31, 107506. [Google Scholar] [CrossRef] [PubMed]
  216. Kodani, A.; Yu, T.W.; Johnson, J.R.; Jayaraman, D.; Johnson, T.L.; Al-Gazali, L.; Sztriha, L.; Partlow, J.N.; Kim, H.; Krup, A.L.; et al. Centriolar Satellites Assemble Centrosomal Microcephaly Proteins to Recruit CDK2 and Promote Centriole Duplication. eLife 2015, 4, e07519. [Google Scholar] [CrossRef] [PubMed]
  217. Jayaraman, D.; Kodani, A.; Gonzalez, D.M.; Mancias, J.D.; Mochida, G.H.; Vagnoni, C.; Johnson, J.; Krogan, N.; Harper, J.W.; Reiter, J.F.; et al. Microcephaly Proteins Wdr62 and Aspm Define a Mother Centriole Complex Regulating Centriole Biogenesis, Apical Complex, and Cell Fate. Neuron 2016, 92, 813–828. [Google Scholar] [CrossRef] [Green Version]
  218. Choi, Y.-K.; Liu, P.; Sze, S.K.; Dai, C.; Qi, R.Z. CDK5RAP2 Stimulates Microtubule Nucleation by the Gamma-Tubulin Ring Complex. J. Cell Biol. 2010, 191, 1089–1095. [Google Scholar] [CrossRef] [Green Version]
  219. Pagan, J.K.; Marzio, A.; Jones, M.J.K.; Saraf, A.; Jallepalli, P.V.; Florens, L.; Washburn, M.P.; Pagano, M. Degradation of Cep68 and PCNT Cleavage Mediate Cep215 Removal from the PCM to Allow Centriole Separation, Disengagement and Licensing. Nat. Cell Biol. 2015, 17, 31–43. [Google Scholar] [CrossRef] [Green Version]
  220. Barrera, J.A.; Kao, L.R.; Hammer, R.E.; Seemann, J.; Fuchs, J.L.; Megraw, T.L. CDK5RAP2 Regulates Centriole Engagement and Cohesion in Mice. Dev. Cell 2010, 18, 913–926. [Google Scholar] [CrossRef] [Green Version]
  221. Firat-Karalar, E.N.; Rauniyar, N.; Yates, J.R.; Stearns, T. Proximity Interactions among Centrosome Components Identify Regulators of Centriole Duplication. Curr. Biol. 2014, 24, 664–670. [Google Scholar] [CrossRef] [Green Version]
  222. Cheeseman, I.M.; Hori, T.; Fukagawa, T.; Desai, A. KNL1 and the CENP-H/I/K Complex Coordinately Direct Kinetochore Assembly in Vertebrates. Mol. Biol. Cell 2008, 19, 587–594. [Google Scholar] [CrossRef] [Green Version]
  223. Ghongane, P.; Kapanidou, M.; Asghar, A.; Elowe, S.; Bolanos-Garcia, V.M. The Dynamic Protein Knl1–a Kinetochore Rendezvous. J. Cell Sci. 2014, 127, 3415–3423. [Google Scholar] [CrossRef] [Green Version]
  224. Rosenberg, J.S.; Cross, F.R.; Funabiki, H. KNL1/Spc105 Recruits PP1 to Silence the Spindle Assembly Checkpoint. Curr. Biol. 2011, 21, 942–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Kiyomitsu, T.; Obuse, C.; Yanagida, M. Human Blinkin/AF15q14 Is Required for Chromosome Alignment and the Mitotic Checkpoint through Direct Interaction with Bub1 and BubR1. Dev. Cell 2007, 13, 663–676. [Google Scholar] [CrossRef] [Green Version]
  226. Kiyomitsu, T.; Murakami, H.; Yanagida, M. Protein Interaction Domain Mapping of Human Kinetochore Protein Blinkin Reveals a Consensus Motif for Binding of Spindle Assembly Checkpoint Proteins Bub1 and BubR1. Mol. Cell. Biol. 2011, 31, 998–1011. [Google Scholar] [CrossRef] [Green Version]
  227. Tang, C.J.; Fu, R.H.; Wu, K.S.; Hsu, W.B.; Tang, T.K. CPAP Is a Cell-Cycle Regulated Protein That Controls Centriole Length. Nat. Cell Biol. 2009, 11, 825–831. [Google Scholar] [CrossRef]
  228. Dzhindzhev, N.S.; Yu, Q.D.; Weiskopf, K.; Tzolovsky, G.; Cunha-Ferreira, I.; Riparbelli, M.; Rodrigues-Martins, A.; Bettencourt-Dias, M.; Callaini, G.; Glover, D.M. Asterless Is a Scaffold for the Onset of Centriole Assembly. Nature 2010, 467, 714–718. [Google Scholar] [CrossRef] [PubMed]
  229. Tang, C.-J.C.; Lin, S.-Y.; Hsu, W.-B.; Lin, Y.-N.; Wu, C.-T.; Lin, Y.-C.; Chang, C.-W.; Wu, K.-S.; Tang, T.K. The Human Microcephaly Protein STIL Interacts with CPAP and Is Required for Procentriole Formation. EMBO J. 2011, 30, 4790–4804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Gabriel, E.; Wason, A.; Ramani, A.; Gooi, L.M.; Keller, P.; Pozniakovsky, A.; Poser, I.; Noack, F.; Telugu, N.S.; Calegari, F.; et al. CPAP Promotes Timely Cilium Disassembly to Maintain Neural Progenitor Pool. EMBO J. 2016, 35, 803–819. [Google Scholar] [CrossRef]
  231. Lin, Y.-C.; Chang, C.-W.; Hsu, W.-B.; Tang, C.-J.C.; Lin, Y.-N.; Chou, E.-J.; Wu, C.-T.; Tang, T.K. Human Microcephaly Protein CEP135 Binds to HSAS-6 and CPAP, and Is Required for Centriole Assembly. EMBO J. 2013, 32, 1141–1154. [Google Scholar] [CrossRef] [Green Version]
  232. Izraeli, S.; Colaizzo-Anas, T.; Bertness, V.L.; Mani, K.; Aplan, P.D.; Kirsch, I.R. Expression of the SIL Gene Is Correlated with Growth Induction and Cellular Proliferation. Cell Growth Differ. 1997, 8, 1171–1179. [Google Scholar]
  233. Izraeli, S.; Lowe, L.A.; Bertness, V.L.; Good, D.J.; Dorward, D.W.; Kirsch, I.R.; Kuehn, M.R. The SIL Gene Is Required for Mouse Embryonic Axial Development and Left-Right Specification. Nature 1999, 399, 691–694. [Google Scholar] [CrossRef]
  234. Kumar, A.; Girimaji, S.C.; Duvvari, M.R.; Blanton, S.H. Mutations in STIL, Encoding a Pericentriolar and Centrosomal Protein, Cause Primary Microcephaly. Am. J. Hum. Genet. 2009, 84, 286–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Arquint, C.; Sonnen, K.F.; Stierhof, Y.-D.; Nigg, E.A. Cell-Cycle-Regulated Expression of STIL Controls Centriole Number in Human Cells. J. Cell Sci. 2012, 125, 1342–1352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Arquint, C.; Nigg, E.A. STIL Microcephaly Mutations Interfere with APC/C-Mediated Degradation and Cause Centriole Amplification. Curr. Biol. 2014, 24, 351–360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Kleylein-Sohn, J.; Westendorf, J.; Le Clech, M.; Habedanck, R.; Stierhof, Y.-D.; Nigg, E.A. Plk4-Induced Centriole Biogenesis in Human Cells. Dev. Cell 2007, 13, 190–202. [Google Scholar] [CrossRef] [Green Version]
  238. Kim, K.; Lee, S.; Chang, J.; Rhee, K. A Novel Function of CEP135 as a Platform Protein of C-NAP1 for Its Centriolar Localization. Exp. Cell Res. 2008, 314, 3692–3700. [Google Scholar] [CrossRef] [PubMed]
  239. Fu, J.; Lipinszki, Z.; Rangone, H.; Min, M.; Mykura, C.; Chao-Chu, J.; Schneider, S.; Dzhindzhev, N.S.; Gottardo, M.; Riparbelli, M.G.; et al. Conserved Molecular Interactions in Centriole-to-Centrosome Conversion. Nat. Cell Biol. 2016, 18, 87–99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Hussain, M.S.; Baig, S.M.; Neumann, S.; Nürnberg, G.; Farooq, M.; Ahmad, I.; Alef, T.; Hennies, H.C.; Technau, M.; Altmüller, J.; et al. A Truncating Mutation of CEP135 Causes Primary Microcephaly and Disturbed Centrosomal Function. Am. J. Hum. Genet. 2012, 90, 871–878. [Google Scholar] [CrossRef] [Green Version]
  241. Guernsey, D.L.; Matsuoka, M.; Jiang, H.; Evans, S.; Macgillivray, C.; Nightingale, M.; Perry, S.; Ferguson, M.; LeBlanc, M.; Paquette, J.; et al. Mutations in Origin Recognition Complex Gene ORC4 Cause Meier-Gorlin Syndrome. Nat. Genet. 2011, 43, 360–364. [Google Scholar] [CrossRef]
  242. Yang, Y.J.; Baltus, A.E.; Mathew, R.S.; Murphy, E.A.; Evrony, G.D.; Gonzalez, D.M.; Wang, E.P.; Marshall-Walker, C.A.; Barry, B.J.; Murn, J.; et al. Microcephaly Gene Links Trithorax and REST/NRSF to Control Neural Stem Cell Proliferation and Differentiation. Cell 2012, 151, 1097–1112. [Google Scholar] [CrossRef] [Green Version]
  243. Garapaty, S.; Xu, C.-F.; Trojer, P.; Mahajan, M.A.; Neubert, T.A.; Samuels, H.H. Identification and Characterization of a Novel Nuclear Protein Complex Involved in Nuclear Hormone Receptor-Mediated Gene Regulation. J. Biol. Chem. 2009, 284, 7542–7552. [Google Scholar] [CrossRef] [Green Version]
  244. Awad, S.; Al-Dosari, M.S.; Al-Yacoub, N.; Colak, D.; Salih, M.A.; Alkuraya, F.S.; Poizat, C. Mutation in PHC1 Implicates Chromatin Remodeling in Primary Microcephaly Pathogenesis. Hum. Mol. Genet. 2013, 22, 2200–2213. [Google Scholar] [CrossRef] [Green Version]
  245. Meyerson, M.; Harlow, E. Identification of G1 Kinase Activity for Cdk6, a Novel Cyclin D Partner. Mol. Cell. Biol. 1994, 14, 2077–2086. [Google Scholar] [CrossRef] [PubMed]
  246. Hussain, M.S.; Baig, S.M.; Neumann, S.; Peche, V.S.; Szczepanski, S.; Nürnberg, G.; Tariq, M.; Jameel, M.; Khan, T.N.; Fatima, A.; et al. CDK6 Associates with the Centrosome during Mitosis and Is Mutated in a Large Pakistani Family with Primary Microcephaly. Hum. Mol. Genet. 2013, 22, 5199–5214. [Google Scholar] [CrossRef] [PubMed]
  247. Thrower, D.A.; Jordan, M.A.; Schaar, B.T.; Yen, T.J.; Wilson, L. Mitotic HeLa Cells Contain a CENP-E-Associated Minus End-Directed Microtubule Motor. EMBO J. 1995, 14, 918–926. [Google Scholar] [CrossRef] [PubMed]
  248. Bancroft, J.; Auckland, P.; Samora, C.P.; McAinsh, A.D. Chromosome Congression Is Promoted by CENP-Q- and CENP-E-Dependent Pathways. J. Cell Sci. 2015, 128, 171–184. [Google Scholar] [CrossRef] [Green Version]
  249. Chan, G.K.; Schaar, B.T.; Yen, T.J. Characterization of the Kinetochore Binding Domain of CENP-E Reveals Interactions with the Kinetochore Proteins CENP-F and HBUBR1. J. Cell Biol. 1998, 143, 49–63. [Google Scholar] [CrossRef] [Green Version]
  250. Mirzaa, G.M.; Vitre, B.; Carpenter, G.; Abramowicz, I.; Gleeson, J.G.; Paciorkowski, A.R.; Cleveland, D.W.; Dobyns, W.B.; O’Driscoll, M. Mutations in CENPE Define a Novel Kinetochore-Centromeric Mechanism for Microcephalic Primordial Dwarfism. Hum. Genet. 2014, 133, 1023–1039. [Google Scholar] [CrossRef] [Green Version]
  251. Strnad, P.; Leidel, S.; Vinogradova, T.; Euteneuer, U.; Khodjakov, A.; Gönczy, P. Regulated HsSAS-6 Levels Ensure Formation of a Single Procentriole per Centriole during the Centrosome Duplication Cycle. Dev. Cell 2007, 13, 203–213. [Google Scholar] [CrossRef] [Green Version]
  252. Van Breugel, M.; Hirono, M.; Andreeva, A.; Yanagisawa, H.; Yamaguchi, S.; Nakazawa, Y.; Morgner, N.; Petrovich, M.; Ebong, I.-O.; Robinson, C.V.; et al. Structures of SAS-6 Suggest Its Organization in Centrioles. Science 2011, 331, 1196–1199. [Google Scholar] [CrossRef] [Green Version]
  253. Khan, M.A.; Rupp, V.M.; Orpinell, M.; Hussain, M.S.; Altmuller, J.; Steinmetz, M.O.; Enzinger, C.; Thiele, H.; Hohne, W.; Nurnberg, G.; et al. A Missense Mutation in the PISA Domain of HsSAS-6 Causes Autosomal Recessive Primary Microcephaly in a Large Consanguineous Pakistani Family. Hum. Mol. Genet. 2017, 23, 5940–5949. [Google Scholar] [CrossRef]
  254. Nguyen, L.N.; Ma, D.; Shui, G.; Wong, P.; Cazenave-Gassiot, A.; Zhang, X.; Wenk, M.R.; Goh, E.L.K.; Silver, D.L. Mfsd2a Is a Transporter for the Essential Omega-3 Fatty Acid Docosahexaenoic Acid. Nature 2014, 509, 503–506. [Google Scholar] [CrossRef] [PubMed]
  255. Ben-Zvi, A.; Lacoste, B.; Kur, E.; Andreone, B.J.; Mayshar, Y.; Yan, H.; Gu, C. Mfsd2a Is Critical for the Formation and Function of the Blood-Brain Barrier. Nature 2014, 509, 507–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Guemez-Gamboa, A.; Nguyen, L.N.; Yang, H.; Zaki, M.S.; Kara, M.; Ben-Omran, T.; Akizu, N.; Rosti, R.O.; Rosti, B.; Scott, E.; et al. Inactivating Mutations in MFSD2A, Required for Omega-3 Fatty Acid Transport in Brain, Cause a Lethal Microcephaly Syndrome. Nat. Genet. 2015, 47, 809–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Asencio, C.; Davidson, I.F.; Santarella-Mellwig, R.; Ly-Hartig, T.B.N.; Mall, M.; Wallenfang, M.R.; Mattaj, I.W.; Gorjánácz, M. Coordination of Kinase and Phosphatase Activities by Lem4 Enables Nuclear Envelope Reassembly during Mitosis. Cell 2012, 150, 122–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Link, N.; Chung, H.; Jolly, A.; Withers, M.; Tepe, B.; Arenkiel, B.R.; Shah, P.S.; Krogan, N.J.; Aydin, H.; Geckinli, B.B.; et al. Mutations in ANKLE2, a ZIKA Virus Target, Disrupt an Asymmetric Cell Division Pathway in Drosophila Neuroblasts to Cause Microcephaly. Dev. Cell 2019, 51, 713–729.e6. [Google Scholar] [CrossRef]
  259. Di Cunto, F.; Calautti, E.; Hsiao, J.; Ong, L.; Topley, G.; Turco, E.; Dotto, G.P. Citron Rho-Interacting Kinase, a Novel Tissue-Specific Ser/Thr Kinase Encompassing the Rho-Rac-Binding Protein Citron. J. Biol. Chem. 1998, 273, 29706–29711. [Google Scholar] [CrossRef] [Green Version]
  260. Harding, B.N.; Moccia, A.; Drunat, S.; Soukarieh, O.; Tubeuf, H.; Chitty, L.S.; Verloes, A.; Gressens, P.; El Ghouzzi, V.; Joriot, S.; et al. Mutations in Citron Kinase Cause Recessive Microlissencephaly with Multinucleated Neurons. Am. J. Hum. Genet. 2016, 99, 511–520. [Google Scholar] [CrossRef] [Green Version]
  261. Li, H.; Bielas, S.L.; Zaki, M.S.; Ismail, S.; Farfara, D.; Um, K.; Rosti, R.O.; Scott, E.C.; Tu, S.; Chi, N.C.; et al. Biallelic Mutations in Citron Kinase Link Mitotic Cytokinesis to Human Primary Microcephaly. Am. J. Hum. Genet. 2016, 99, 501–510. [Google Scholar] [CrossRef] [Green Version]
  262. Gruneberg, U.; Neef, R.; Li, X.; Chan, E.H.Y.; Chalamalasetty, R.B.; Nigg, E.A.; Barr, F.A. KIF14 and Citron Kinase Act Together to Promote Efficient Cytokinesis. J. Cell Biol. 2006, 172, 363–372. [Google Scholar] [CrossRef] [Green Version]
  263. Gai, M.; Bianchi, F.T.; Vagnoni, C.; Verni, F.; Bonaccorsi, S.; Pasquero, S.; Berto, G.E.; Sgro, F.; Chiotto, A.M.; Annaratone, L.; et al. ASPM and CITK Regulate Spindle Orientation by Affecting the Dynamics of Astral Microtubules. EMBO Rep. 2016, 17, 1396–1409. [Google Scholar] [CrossRef] [Green Version]
  264. Kadir, R.; Harel, T.; Markus, B.; Perez, Y.; Bakhrat, A.; Cohen, I.; Volodarsky, M.; Feintsein-Linial, M.; Chervinski, E.; Zlotogora, J.; et al. ALFY-Controlled DVL3 Autophagy Regulates Wnt Signaling, Determining Human Brain Size. PLoS Genet. 2016, 12, e1005919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Clausen, T.H.; Lamark, T.; Isakson, P.; Finley, K.; Larsen, K.B.; Brech, A.; Øvervatn, A.; Stenmark, H.; Bjørkøy, G.; Simonsen, A.; et al. P62/SQSTM1 and ALFY Interact to Facilitate the Formation of P62 Bodies/ALIS and Their Degradation by Autophagy. Autophagy 2010, 6, 330–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. DiStasio, A.; Driver, A.; Sund, K.; Donlin, M.; Muraleedharan, R.M.; Pooya, S.; Kline-Fath, B.; Kaufman, K.M.; Prows, C.A.; Schorry, E.; et al. Copb2 Is Essential for Embryogenesis and Hypomorphic Mutations Cause Human Microcephaly. Hum. Mol. Genet. 2017, 26, 4836–4848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Carleton, M.; Mao, M.; Biery, M.; Warrener, P.; Kim, S.; Buser, C.; Marshall, C.G.; Fernandes, C.; Annis, J.; Linsley, P.S. RNA Interference-Mediated Silencing of Mitotic Kinesin KIF14 Disrupts Cell Cycle Progression and Induces Cytokinesis Failure. Mol. Cell. Biol. 2006, 26, 3853–3863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Zhu, C.; Zhao, J.; Bibikova, M.; Leverson, J.D.; Bossy-Wetzel, E.; Fan, J.-B.; Abraham, R.T.; Jiang, W. Functional Analysis of Human Microtubule-Based Motor Proteins, the Kinesins and Dyneins, in Mitosis/Cytokinesis Using RNA Interference. Mol. Biol. Cell 2005, 16, 3187–3199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Moawia, A.; Shaheen, R.; Rasool, S.; Waseem, S.S.; Ewida, N.; Budde, B.; Kawalia, A.; Motameny, S.; Khan, K.; Fatima, A.; et al. Mutations of KIF14 Cause Primary Microcephaly by Impairing Cytokinesis. Ann. Neurol. 2017, 82, 562–577. [Google Scholar] [CrossRef]
  270. Martin, C.-A.; Murray, J.E.; Carroll, P.; Leitch, A.; Mackenzie, K.J.; Halachev, M.; Fetit, A.E.; Keith, C.; Bicknell, L.S.; Fluteau, A.; et al. Mutations in Genes Encoding Condensin Complex Proteins Cause Microcephaly through Decatenation Failure at Mitosis. Genes Dev. 2016, 30, 2158–2172. [Google Scholar] [CrossRef] [Green Version]
  271. Kimura, K.; Cuvier, O.; Hirano, T. Chromosome Condensation by a Human Condensin Complex in Xenopus Egg Extracts. J. Biol. Chem. 2001, 276, 5417–5420. [Google Scholar] [CrossRef] [Green Version]
  272. Zuccolo, M.; Alves, A.; Galy, V.; Bolhy, S.; Formstecher, E.; Racine, V.; Sibarita, J.-B.; Fukagawa, T.; Shiekhattar, R.; Yen, T.; et al. The Human Nup107-160 Nuclear Pore Subcomplex Contributes to Proper Kinetochore Functions. EMBO J. 2007, 26, 1853–1864. [Google Scholar] [CrossRef] [Green Version]
  273. Braun, D.A.; Lovric, S.; Schapiro, D.; Schneider, R.; Marquez, J.; Asif, M.; Hussain, M.S.; Daga, A.; Widmeier, E.; Rao, J.; et al. Mutations in Multiple Components of the Nuclear Pore Complex Cause Nephrotic Syndrome. J. Clin. Investig. 2018, 128, 4313–4328. [Google Scholar] [CrossRef] [Green Version]
  274. Perez, Y.; Bar-Yaacov, R.; Kadir, R.; Wormser, O.; Shelef, I.; Birk, O.S.; Flusser, H.; Birnbaum, R.Y. Mutations in the Microtubule-Associated Protein MAP11 (C7orf43) Cause Microcephaly in Humans and Zebrafish. Brain 2019, 142, 574–585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Cristofoli, F.; Moss, T.; Moore, H.W.; Devriendt, K.; Flanagan-Steet, H.; May, M.; Jones, J.; Roelens, F.; Fons, C.; Fernandez, A.; et al. De Novo Variants in LMNB1 Cause Pronounced Syndromic Microcephaly and Disruption of Nuclear Envelope Integrity. Am. J. Hum. Genet. 2020, 107, 753–762. [Google Scholar] [CrossRef]
  276. Parry, D.A.; Martin, C.-A.; Greene, P.; Marsh, J.A.; Genomics England Research Consortium; Blyth, M.; Cox, H.; Donnelly, D.; Greenhalgh, L.; Greville-Heygate, S.; et al. Heterozygous Lamin B1 and Lamin B2 Variants Cause Primary Microcephaly and Define a Novel Laminopathy. Genet. Med. 2021, 23, 408–414. [Google Scholar] [CrossRef] [PubMed]
  277. Goldman, R.D.; Gruenbaum, Y.; Moir, R.D.; Shumaker, D.K.; Spann, T.P. Nuclear Lamins: Building Blocks of Nuclear Architecture. Genes Dev. 2002, 16, 533–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Tsai, M.-Y.; Wang, S.; Heidinger, J.M.; Shumaker, D.K.; Adam, S.A.; Goldman, R.D.; Zheng, Y. A Mitotic Lamin B Matrix Induced by RanGTP Required for Spindle Assembly. Science 2006, 311, 1887–1893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Farooq, M.; Lindbæk, L.; Krogh, N.; Doganli, C.; Keller, C.; Mönnich, M.; Gonçalves, A.B.; Sakthivel, S.; Mang, Y.; Fatima, A.; et al. RRP7A Links Primary Microcephaly to Dysfunction of Ribosome Biogenesis, Resorption of Primary Cilia, and Neurogenesis. Nat. Commun. 2020, 11, 5816. [Google Scholar] [CrossRef]
  280. Singh, S.; Vanden Broeck, A.; Miller, L.; Chaker-Margot, M.; Klinge, S. Nucleolar Maturation of the Human Small Subunit Processome. Science 2021, 373, eabj5338. [Google Scholar] [CrossRef]
  281. Khan, A.; Alaamery, M.; Massadeh, S.; Obaid, A.; Kashgari, A.A.; Walsh, C.A.; Eyaid, W. PDCD6IP, Encoding a Regulator of the ESCRT Complex, Is Mutated in Microcephaly. Clin. Genet. 2020, 98, 80–85. [Google Scholar] [CrossRef]
  282. Klebig, C.; Korinth, D.; Meraldi, P. Bub1 Regulates Chromosome Segregation in a Kinetochore-Independent Manner. J. Cell Biol. 2009, 185, 841–858. [Google Scholar] [CrossRef]
  283. Carvalhal, S.; Bader, I.; Rooimans, M.A.; Oostra, A.B.; Balk, J.A.; Feichtinger, R.G.; Beichler, C.; Speicher, M.R.; van Hagen, J.M.; Waisfisz, Q.; et al. Biallelic BUB1 Mutations Cause Microcephaly, Developmental Delay, and Variable Effects on Cohesion and Chromosome Segregation. Sci. Adv. 2022, 8, eabk0114. [Google Scholar] [CrossRef]
  284. Seeley, T.W.; Wang, L.; Zhen, J.Y. Phosphorylation of Human MAD1 by the BUB1 Kinase in Vitro. Biochem. Biophys. Res. Commun. 1999, 257, 589–595. [Google Scholar] [CrossRef] [PubMed]
  285. Johnson, V.L.; Scott, M.I.F.; Holt, S.V.; Hussein, D.; Taylor, S.S. Bub1 Is Required for Kinetochore Localization of BubR1, Cenp-E, Cenp-F and Mad2, and Chromosome Congression. J. Cell Sci. 2004, 117, 1577–1589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Flory, M.R.; Moser, M.J.; Monnat, R.J.; Davis, T.N. Identification of a Human Centrosomal Calmodulin-Binding Protein That Shares Homology with Pericentrin. Proc. Natl. Acad. Sci. USA 2000, 97, 5919–5923. [Google Scholar] [CrossRef] [PubMed]
  287. Li, Q.; Hansen, D.; Killilea, A.; Joshi, H.C.; Palazzo, R.E.; Balczon, R. Kendrin/Pericentrin-B, a Centrosome Protein with Homology to Pericentrin That Complexes with PCM-1. J. Cell Sci. 2001, 114, 797–809. [Google Scholar] [CrossRef] [PubMed]
  288. Matsuo, K.; Nishimura, T.; Hayakawa, A.; Ono, Y.; Takahashi, M. Involvement of a Centrosomal Protein Kendrin in the Maintenance of Centrosome Cohesion by Modulating Nek2A Kinase Activity. Biochem. Biophys. Res. Commun. 2010, 398, 217–223. [Google Scholar] [CrossRef]
  289. Buchman, J.J.; Tseng, H.-C.; Zhou, Y.; Frank, C.L.; Xie, Z.; Tsai, L.-H. Cdk5rap2 Interacts with Pericentrin to Maintain the Neural Progenitor Pool in the Developing Neocortex. Neuron 2010, 66, 386–402. [Google Scholar] [CrossRef] [Green Version]
  290. Hall-Jackson, C.A.; Cross, D.A.; Morrice, N.; Smythe, C. ATR Is a Caffeine-Sensitive, DNA-Activated Protein Kinase with a Substrate Specificity Distinct from DNA-PK. Oncogene 1999, 18, 6707–6713. [Google Scholar] [CrossRef] [Green Version]
  291. Liu, Q.; Guntuku, S.; Cui, X.S.; Matsuoka, S.; Cortez, D.; Tamai, K.; Luo, G.; Carattini-Rivera, S.; DeMayo, F.; Bradley, A.; et al. Chk1 Is an Essential Kinase That Is Regulated by Atr and Required for the G(2)/M DNA Damage Checkpoint. Genes Dev. 2000, 14, 1448–1459. [Google Scholar] [CrossRef]
  292. Casper, A.M.; Nghiem, P.; Arlt, M.F.; Glover, T.W. ATR Regulates Fragile Site Stability. Cell 2002, 111, 779–789. [Google Scholar] [CrossRef] [Green Version]
  293. Sartori, A.A.; Lukas, C.; Coates, J.; Mistrik, M.; Fu, S.; Bartek, J.; Baer, R.; Lukas, J.; Jackson, S.P. Human CtIP Promotes DNA End Resection. Nature 2007, 450, 509–514. [Google Scholar] [CrossRef] [Green Version]
  294. Qvist, P.; Huertas, P.; Jimeno, S.; Nyegaard, M.; Hassan, M.J.; Jackson, S.P.; Børglum, A.D. CtIP Mutations Cause Seckel and Jawad Syndromes. PLoS Genet. 2011, 7, e1002310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Yuan, J.; Chen, J. N Terminus of CtIP Is Critical for Homologous Recombination-Mediated Double-Strand Break Repair. J. Biol. Chem. 2009, 284, 31746–31752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Yu, X.; Chen, J. DNA Damage-Induced Cell Cycle Checkpoint Control Requires CtIP, a Phosphorylation-Dependent Binding Partner of BRCA1 C-Terminal Domains. Mol. Cell. Biol. 2004, 24, 9478–9486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Loffler, H.; Rebacz, B.; Ho, A.D.; Lukas, J.; Bartek, J.; Kramer, A. Chk1-Dependent Regulation of Cdc25B Functions to Coordinate Mitotic Events. Cell Cycle 2006, 5, 2543–2547. [Google Scholar] [CrossRef] [Green Version]
  298. Sir, J.-H.; Barr, A.R.; Nicholas, A.K.; Carvalho, O.P.; Khurshid, M.; Sossick, A.; Reichelt, S.; D’Santos, C.; Woods, C.G.; Gergely, F. A Primary Microcephaly Protein Complex Forms a Ring around Parental Centrioles. Nat. Genet. 2011, 43, 1147–1153. [Google Scholar] [CrossRef] [Green Version]
  299. Dauber, A.; Lafranchi, S.H.; Maliga, Z.; Lui, J.C.; Moon, J.E.; McDeed, C.; Henke, K.; Zonana, J.; Kingman, G.A.; Pers, T.H.; et al. Novel Microcephalic Primordial Dwarfism Disorder Associated with Variants in the Centrosomal Protein Ninein. J. Clin. Endocrinol. Metab. 2012, 97, E2140–E2151. [Google Scholar] [CrossRef]
  300. Ou, Y.Y.; Mack, G.J.; Zhang, M.; Rattner, J.B. CEP110 and Ninein Are Located in a Specific Domain of the Centrosome Associated with Centrosome Maturation. J. Cell Sci. 2002, 115, 1825–1835. [Google Scholar] [CrossRef]
  301. Huang, N.; Xia, Y.; Zhang, D.; Wang, S.; Bao, Y.; He, R.; Teng, J.; Chen, J. Hierarchical Assembly of Centriole Subdistal Appendages via Centrosome Binding Proteins CCDC120 and CCDC68. Nat. Commun. 2017, 8, 15057. [Google Scholar] [CrossRef] [Green Version]
  302. Duxin, J.P.; Dao, B.; Martinsson, P.; Rajala, N.; Guittat, L.; Campbell, J.L.; Spelbrink, J.N.; Stewart, S.A. Human Dna2 Is a Nuclear and Mitochondrial DNA Maintenance Protein. Mol. Cell. Biol. 2009, 29, 4274–4282. [Google Scholar] [CrossRef] [Green Version]
  303. Shaheen, R.; Faqeih, E.; Ansari, S.; Abdel-Salam, G.; Al-Hassnan, Z.N.; Al-Shidi, T.; Alomar, R.; Sogaty, S.; Alkuraya, F.S. Genomic Analysis of Primordial Dwarfism Reveals Novel Disease Genes. Genome Res. 2014, 24, 291–299. [Google Scholar] [CrossRef] [Green Version]
  304. Zheng, L.; Zhou, M.; Guo, Z.; Lu, H.; Qian, L.; Dai, H.; Qiu, J.; Yakubovskaya, E.; Bogenhagen, D.F.; Demple, B.; et al. Human DNA2 Is a Mitochondrial Nuclease/Helicase for Efficient Processing of DNA Replication and Repair Intermediates. Mol. Cell 2008, 32, 325–336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Nimonkar, A.V.; Genschel, J.; Kinoshita, E.; Polaczek, P.; Campbell, J.L.; Wyman, C.; Modrich, P.; Kowalczykowski, S.C. BLM-DNA2-RPA-MRN and EXO1-BLM-RPA-MRN Constitute Two DNA End Resection Machineries for Human DNA Break Repair. Genes Dev. 2011, 25, 350–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Harley, M.E.; Murina, O.; Leitch, A.; Higgs, M.R.; Bicknell, L.S.; Yigit, G.; Blackford, A.N.; Zlatanou, A.; Mackenzie, K.J.; Reddy, K.; et al. TRAIP Promotes DNA Damage Response during Genome Replication and Is Mutated in Primordial Dwarfism. Nat. Genet. 2016, 48, 36–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Hoffmann, S.; Smedegaard, S.; Nakamura, K.; Mortuza, G.B.; Räschle, M.; Ibañez de Opakua, A.; Oka, Y.; Feng, Y.; Blanco, F.J.; Mann, M.; et al. TRAIP Is a PCNA-Binding Ubiquitin Ligase That Protects Genome Stability after Replication Stress. J. Cell Biol. 2016, 212, 63–75. [Google Scholar] [CrossRef] [PubMed]
  308. Sonneville, R.; Bhowmick, R.; Hoffmann, S.; Mailand, N.; Hickson, I.D.; Labib, K. TRAIP Drives Replisome Disassembly and Mitotic DNA Repair Synthesis at Sites of Incomplete DNA Replication. eLife 2019, 8, e48686. [Google Scholar] [CrossRef]
  309. Chapard, C.; Meraldi, P.; Gleich, T.; Bachmann, D.; Hohl, D.; Huber, M. TRAIP Is a Regulator of the Spindle Assembly Checkpoint. J. Cell Sci. 2014, 127, 5149–5156. [Google Scholar] [CrossRef] [Green Version]
  310. Potts, P.R.; Yu, H. Human MMS21/NSE2 Is a SUMO Ligase Required for DNA Repair. Mol. Cell. Biol. 2005, 25, 7021–7032. [Google Scholar] [CrossRef] [Green Version]
  311. Potts, P.R.; Porteus, M.H.; Yu, H. Human SMC5/6 Complex Promotes Sister Chromatid Homologous Recombination by Recruiting the SMC1/3 Cohesin Complex to Double-Strand Breaks. EMBO J. 2006, 25, 3377–3388. [Google Scholar] [CrossRef] [Green Version]
  312. Payne, F.; Colnaghi, R.; Rocha, N.; Seth, A.; Harris, J.; Carpenter, G.; Bottomley, W.E.; Wheeler, E.; Wong, S.; Saudek, V.; et al. Hypomorphism in Human NSMCE2 Linked to Primordial Dwarfism and Insulin Resistance. J. Clin. Investig. 2014, 124, 4028–4038. [Google Scholar] [CrossRef] [Green Version]
  313. Murphy, S.M.; Preble, A.M.; Patel, U.K.; O’Connell, K.L.; Dias, D.P.; Moritz, M.; Agard, D.; Stults, J.T.; Stearns, T. GCP5 and GCP6: Two New Members of the Human Gamma-Tubulin Complex. Mol. Biol. Cell 2001, 12, 3340–3352. [Google Scholar] [CrossRef]
  314. Bettencourt-Dias, M.; Rodrigues-Martins, A.; Carpenter, L.; Riparbelli, M.; Lehmann, L.; Gatt, M.K.; Carmo, N.; Balloux, F.; Callaini, G.; Glover, D.M. SAK/PLK4 Is Required for Centriole Duplication and Flagella Development. Curr. Biol. 2005, 15, 2199–2207. [Google Scholar] [CrossRef] [Green Version]
  315. Habedanck, R.; Stierhof, Y.-D.; Wilkinson, C.J.; Nigg, E.A. The Polo Kinase Plk4 Functions in Centriole Duplication. Nat. Cell Biol. 2005, 7, 1140–1146. [Google Scholar] [CrossRef]
  316. Fava, F.; Raynaud-Messina, B.; Leung-Tack, J.; Mazzolini, L.; Li, M.; Guillemot, J.C.; Cachot, D.; Tollon, Y.; Ferrara, P.; Wright, M. Human 76p: A New Member of the Gamma-Tubulin-Associated Protein Family. J. Cell Biol. 1999, 147, 857–868. [Google Scholar] [CrossRef] [Green Version]
  317. Rapley, J.; Nicolàs, M.; Groen, A.; Regué, L.; Bertran, M.T.; Caelles, C.; Avruch, J.; Roig, J. The NIMA-Family Kinase Nek6 Phosphorylates the Kinesin Eg5 at a Novel Site Necessary for Mitotic Spindle Formation. J. Cell Sci. 2008, 121, 3912–3921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Ostergaard, P.; Simpson, M.A.; Mendola, A.; Vasudevan, P.; Connell, F.C.; van Impel, A.; Moore, A.T.; Loeys, B.L.; Ghalamkarpour, A.; Onoufriadis, A.; et al. Mutations in KIF11 Cause Autosomal-Dominant Microcephaly Variably Associated with Congenital Lymphedema and Chorioretinopathy. Am. J. Hum. Genet. 2012, 90, 356–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Splinter, D.; Razafsky, D.S.; Schlager, M.A.; Serra-Marques, A.; Grigoriev, I.; Demmers, J.; Keijzer, N.; Jiang, K.; Poser, I.; Hyman, A.A.; et al. BICD2, Dynactin, and LIS1 Cooperate in Regulating Dynein Recruitment to Cellular Structures. Mol. Biol. Cell 2012, 23, 4226–4241. [Google Scholar] [CrossRef] [PubMed]
  320. McKenney, R.J.; Huynh, W.; Tanenbaum, M.E.; Bhabha, G.; Vale, R.D. Activation of Cytoplasmic Dynein Motility by Dynactin-Cargo Adapter Complexes. Science 2014, 345, 337–341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  321. Chaaban, S.; Carter, A.P. Structure of Dynein-Dynactin on Microtubules Shows Tandem Adaptor Binding. Nature 2022, 610, 212–216. [Google Scholar] [CrossRef] [PubMed]
  322. Ellis, N.A.; Lennon, D.J.; Proytcheva, M.; Alhadeff, B.; Henderson, E.E.; German, J. Somatic Intragenic Recombination within the Mutated Locus BLM Can Correct the High Sister-Chromatid Exchange Phenotype of Bloom Syndrome Cells. Am. J. Hum. Genet. 1995, 57, 1019–1027. [Google Scholar]
  323. Ellis, N.A.; Groden, J.; Ye, T.Z.; Straughen, J.; Lennon, D.J.; Ciocci, S.; Proytcheva, M.; German, J. The Bloom’s Syndrome Gene Product Is Homologous to RecQ Helicases. Cell 1995, 83, 655–666. [Google Scholar] [CrossRef] [Green Version]
  324. Karow, J.K.; Chakraverty, R.K.; Hickson, I.D. The Bloom’s Syndrome Gene Product Is a 3′-5′ DNA Helicase. J. Biol. Chem. 1997, 272, 30611–30614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  325. Langland, G.; Elliott, J.; Li, Y.; Creaney, J.; Dixon, K.; Groden, J. The BLM Helicase Is Necessary for Normal DNA Double-Strand Break Repair. Cancer Res. 2002, 62, 2766–2770. [Google Scholar] [PubMed]
  326. Grawunder, U.; Wilm, M.; Wu, X.; Kulesza, P.; Wilson, T.E.; Mann, M.; Lieber, M.R. Activity of DNA Ligase IV Stimulated by Complex Formation with XRCC4 Protein in Mammalian Cells. Nature 1997, 388, 492–495. [Google Scholar] [CrossRef] [PubMed]
  327. Grawunder, U.; Zimmer, D.; Fugmann, S.; Schwarz, K.; Lieber, M.R. DNA Ligase IV Is Essential for V(D)J Recombination and DNA Double-Strand Break Repair in Human Precursor Lymphocytes. Mol. Cell 1998, 2, 477–484. [Google Scholar] [CrossRef] [PubMed]
  328. Gu, J.; Lu, H.; Tippin, B.; Shimazaki, N.; Goodman, M.F.; Lieber, M.R. XRCC4:DNA Ligase IV Can Ligate Incompatible DNA Ends and Can Ligate across Gaps. EMBO J. 2007, 26, 1010–1023. [Google Scholar] [CrossRef] [Green Version]
  329. Rosin, N.; Elcioglu, N.H.; Beleggia, F.; Isguven, P.; Altmuller, J.; Thiele, H.; Steindl, K.; Joset, P.; Rauch, A.; Nurnberg, P.; et al. Mutations in XRCC4 Cause Primary Microcephaly, Short Stature and Increased Genomic Instability. Hum. Mol. Genet. 2015, 24, 3708–3717. [Google Scholar] [CrossRef] [Green Version]
  330. Bicknell, L.S.; Walker, S.; Klingseisen, A.; Stiff, T.; Leitch, A.; Kerzendorfer, C.; Martin, C.-A.; Yeyati, P.; Al Sanna, N.; Bober, M.; et al. Mutations in ORC1, Encoding the Largest Subunit of the Origin Recognition Complex, Cause Microcephalic Primordial Dwarfism Resembling Meier-Gorlin Syndrome. Nat. Genet. 2011, 43, 350–355. [Google Scholar] [CrossRef]
  331. Ohta, S.; Tatsumi, Y.; Fujita, M.; Tsurimoto, T.; Obuse, C. The ORC1 Cycle in Human Cells: II. Dynamic Changes in the Human ORC Complex during the Cell Cycle. J. Biol. Chem. 2003, 278, 41535–41540. [Google Scholar] [CrossRef] [Green Version]
  332. Wohlschlegel, J.A.; Dwyer, B.T.; Dhar, S.K.; Cvetic, C.; Walter, J.C.; Dutta, A. Inhibition of Eukaryotic DNA Replication by Geminin Binding to Cdt1. Science 2000, 290, 2309–2312. [Google Scholar] [CrossRef]
  333. Varma, D.; Chandrasekaran, S.; Sundin, L.J.R.; Reidy, K.T.; Wan, X.; Chasse, D.A.D.; Nevis, K.R.; DeLuca, J.G.; Salmon, E.D.; Cook, J.G. Recruitment of the Human Cdt1 Replication Licensing Protein by the Loop Domain of Hec1 Is Required for Stable Kinetochore-Microtubule Attachment. Nat. Cell Biol. 2012, 14, 593–603. [Google Scholar] [CrossRef] [Green Version]
  334. Walter, D.; Hoffmann, S.; Komseli, E.-S.; Rappsilber, J.; Gorgoulis, V.; Sørensen, C.S. SCF(Cyclin F)-Dependent Degradation of CDC6 Suppresses DNA Re-Replication. Nat. Commun. 2016, 7, 10530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Saha, P.; Chen, J.; Thome, K.C.; Lawlis, S.J.; Hou, Z.H.; Hendricks, M.; Parvin, J.D.; Dutta, A. Human CDC6/Cdc18 Associates with Orc1 and Cyclin-Cdk and Is Selectively Eliminated from the Nucleus at the Onset of S Phase. Mol. Cell. Biol. 1998, 18, 2758–2767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Burrage, L.C.; Charng, W.-L.; Eldomery, M.K.; Willer, J.R.; Davis, E.E.; Lugtenberg, D.; Zhu, W.; Leduc, M.S.; Akdemir, Z.C.; Azamian, M.; et al. De Novo GMNN Mutations Cause Autosomal-Dominant Primordial Dwarfism Associated with Meier-Gorlin Syndrome. Am. J. Hum. Genet. 2015, 97, 904–913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Pefani, D.-E.; Dimaki, M.; Spella, M.; Karantzelis, N.; Mitsiki, E.; Kyrousi, C.; Symeonidou, I.-E.; Perrakis, A.; Taraviras, S.; Lygerou, Z. Idas, a Novel Phylogenetically Conserved Geminin-Related Protein, Binds to Geminin and Is Required for Cell Cycle Progression. J. Biol. Chem. 2011, 286, 23234–23246. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Body size, brain growth evolution, and structural neuroanatomy of ASPM-, WDR62-, and DYNC1H1-related primary microcephalies. (A): Schema illustrating the most emblematic primary microcephalies (PMs) caused by ASPM, WDR62, or DYMC1H1 mutations, characterized by a reduction in brain volume without reduction in body size. (B): Occipitofrontal head circumference (OFC) showing that brain growth in individuals with ASPM, WDR62, or DYMC1H1 PM (in blue) is mainly below the normal range, taking age and sex into account. (CF): Brain magnetic resonance imaging (MRI) of healthy control (C) as compared to patients with primary PM caused by ASPM (D), WDR62 (E), or DYNC1H1 (F) mutations. From top to bottom (1–4): Axial T1, coronal T1 (or axial and coronal T2 FLAIR for (F), coronal T2, and sagittal T1-weighted images illustrating the main cortical malformations associated with PMs caused by mutations in ASPM, WDR62, and DYNC1H1. Reduction in brain volume and gyral simplification is evident in all cases. Bilateral polymicrogyria is observed frequently in patients with WDR62 mutations ((E), white arrow). Corpus callosum agenesis (F3–4, yellow arrows), pachygyria (F1–3), and brainstem hypoplasia (white arrows, F4) are frequently associated with DYNC1H1 mutations.
Figure 1. Body size, brain growth evolution, and structural neuroanatomy of ASPM-, WDR62-, and DYNC1H1-related primary microcephalies. (A): Schema illustrating the most emblematic primary microcephalies (PMs) caused by ASPM, WDR62, or DYMC1H1 mutations, characterized by a reduction in brain volume without reduction in body size. (B): Occipitofrontal head circumference (OFC) showing that brain growth in individuals with ASPM, WDR62, or DYMC1H1 PM (in blue) is mainly below the normal range, taking age and sex into account. (CF): Brain magnetic resonance imaging (MRI) of healthy control (C) as compared to patients with primary PM caused by ASPM (D), WDR62 (E), or DYNC1H1 (F) mutations. From top to bottom (1–4): Axial T1, coronal T1 (or axial and coronal T2 FLAIR for (F), coronal T2, and sagittal T1-weighted images illustrating the main cortical malformations associated with PMs caused by mutations in ASPM, WDR62, and DYNC1H1. Reduction in brain volume and gyral simplification is evident in all cases. Bilateral polymicrogyria is observed frequently in patients with WDR62 mutations ((E), white arrow). Corpus callosum agenesis (F3–4, yellow arrows), pachygyria (F1–3), and brainstem hypoplasia (white arrows, F4) are frequently associated with DYNC1H1 mutations.
Cells 12 01807 g001
Figure 2. Stature, brain growth evolution, and structural neuroanatomy in patients with primordial dwarfism (MPD) due to PCNT mutations. (A): Schema illustrating PCNT primordial dwarfism characterized by a proportionate reduction in both brain size and body size. (B): Height (top) and weight (bottom) growth curve of a girl carrying PCNT mutations with a deviation of −7.5 SD from the mean normal for weight and height. (C): The OFC growth curve of a girl carrying PCNT mutations with a deviation of −8 SD from the mean. (DF): Brain MRI (1–3) and magnetic resonance angiography (MRA, 4) of healthy control (D), as compared to the same 6.5 year-old girl carrying PCNT mutations (E,F). Axial T1 (1 (D,E)), coronal T1 (2 (D,E)), sagittal T1-weighted images (3 (D,E)), coronal T2 (1 (F)), and coronal and sagittal T2 FLAIR-weighted images (2–3, (F)) illustrating brain structure and vascular complications related to PCNT mutations. Note the left occipital cortico-subcortical lesion secondary to sequelae from an ischemic stroke (2–3 (F), white arrow) due to a tight stenosis at the origin of the left posterior cerebral artery. This patient exhibited the typical arteriopathy related to PCNT mutations: (i) a bilateral tight and long stenosis of the extracranial internal carotid arteries from the carotid bifurcation, with neovascularization, and (ii) a narrowing and rigid appearance of intracranial internal carotid arteries, circle of Willis, vertebral and basilar arteries. (G): Skeletal complications in PCNT primordial dwarfism, including bilateral coxa vara, narrow ischia, irregular metaphysis (yellow arrows), bilateral proximal femoral epiphyseolysis (white arrows), and brachymesophalangia of the Vth fingers.
Figure 2. Stature, brain growth evolution, and structural neuroanatomy in patients with primordial dwarfism (MPD) due to PCNT mutations. (A): Schema illustrating PCNT primordial dwarfism characterized by a proportionate reduction in both brain size and body size. (B): Height (top) and weight (bottom) growth curve of a girl carrying PCNT mutations with a deviation of −7.5 SD from the mean normal for weight and height. (C): The OFC growth curve of a girl carrying PCNT mutations with a deviation of −8 SD from the mean. (DF): Brain MRI (1–3) and magnetic resonance angiography (MRA, 4) of healthy control (D), as compared to the same 6.5 year-old girl carrying PCNT mutations (E,F). Axial T1 (1 (D,E)), coronal T1 (2 (D,E)), sagittal T1-weighted images (3 (D,E)), coronal T2 (1 (F)), and coronal and sagittal T2 FLAIR-weighted images (2–3, (F)) illustrating brain structure and vascular complications related to PCNT mutations. Note the left occipital cortico-subcortical lesion secondary to sequelae from an ischemic stroke (2–3 (F), white arrow) due to a tight stenosis at the origin of the left posterior cerebral artery. This patient exhibited the typical arteriopathy related to PCNT mutations: (i) a bilateral tight and long stenosis of the extracranial internal carotid arteries from the carotid bifurcation, with neovascularization, and (ii) a narrowing and rigid appearance of intracranial internal carotid arteries, circle of Willis, vertebral and basilar arteries. (G): Skeletal complications in PCNT primordial dwarfism, including bilateral coxa vara, narrow ischia, irregular metaphysis (yellow arrows), bilateral proximal femoral epiphyseolysis (white arrows), and brachymesophalangia of the Vth fingers.
Cells 12 01807 g002
Figure 3. Localization of the most emblematic primary microcephaly proteins during mitosis. Scheme showing ASPM, WDR62, and DYNC1H1 at the mitotic spindle, PCNT and CDK5RAP2 at the pericentriolar matrix of the centrosome, and CEP152 and PLK4 at the centriole.
Figure 3. Localization of the most emblematic primary microcephaly proteins during mitosis. Scheme showing ASPM, WDR62, and DYNC1H1 at the mitotic spindle, PCNT and CDK5RAP2 at the pericentriolar matrix of the centrosome, and CEP152 and PLK4 at the centriole.
Cells 12 01807 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Farcy, S.; Hachour, H.; Bahi-Buisson, N.; Passemard, S. Genetic Primary Microcephalies: When Centrosome Dysfunction Dictates Brain and Body Size. Cells 2023, 12, 1807. https://doi.org/10.3390/cells12131807

AMA Style

Farcy S, Hachour H, Bahi-Buisson N, Passemard S. Genetic Primary Microcephalies: When Centrosome Dysfunction Dictates Brain and Body Size. Cells. 2023; 12(13):1807. https://doi.org/10.3390/cells12131807

Chicago/Turabian Style

Farcy, Sarah, Hassina Hachour, Nadia Bahi-Buisson, and Sandrine Passemard. 2023. "Genetic Primary Microcephalies: When Centrosome Dysfunction Dictates Brain and Body Size" Cells 12, no. 13: 1807. https://doi.org/10.3390/cells12131807

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