Next Article in Journal
Emerging Mechanisms of Skeletal Muscle Homeostasis and Cachexia: The SUMO Perspective
Next Article in Special Issue
Genetic Primary Microcephalies: When Centrosome Dysfunction Dictates Brain and Body Size
Previous Article in Journal
The RUNX/CBFβ Complex in Breast Cancer: A Conundrum of Context
Previous Article in Special Issue
Cortical Organoids to Model Microcephaly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Congenital Microcephaly: A Debate on Diagnostic Challenges and Etiological Paradigm of the Shift from Isolated/Non-Syndromic to Syndromic Microcephaly

by
Maria Asif
1,2,†,
Uzma Abdullah
3,†,
Peter Nürnberg
1,2,
Sigrid Tinschert
4 and
Muhammad Sajid Hussain
1,2,*
1
Cologne Center for Genomics (CCG), Faculty of Medicine, University Hospital Cologne, University of Cologne, 50931 Cologne, Germany
2
Center for Molecular Medicine Cologne (CMMC), Faculty of Medicine, University Hospital Cologne, University of Cologne, 50931 Cologne, Germany
3
University Institute of Biochemistry and Biotechnology (UIBB), PMAS-Arid Agriculture University, Rawalpindi, Rawalpindi 46300, Pakistan
4
Zentrum Medizinische Genetik, Medizinische Universität, 6020 Innsbruck, Austria
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2023, 12(4), 642; https://doi.org/10.3390/cells12040642
Submission received: 8 January 2023 / Revised: 13 February 2023 / Accepted: 14 February 2023 / Published: 16 February 2023
(This article belongs to the Special Issue Cellular Mechanisms of Microcephaly)

Abstract

:
Congenital microcephaly (CM) exhibits broad clinical and genetic heterogeneity and is thus categorized into several subtypes. However, the recent bloom of disease–gene discoveries has revealed more overlaps than differences in the underlying genetic architecture for these clinical sub-categories, complicating the differential diagnosis. Moreover, the mechanism of the paradigm shift from a brain-restricted to a multi-organ phenotype is only vaguely understood. This review article highlights the critical factors considered while defining CM subtypes. It also presents possible arguments on long-standing questions of the brain-specific nature of CM caused by a dysfunction of the ubiquitously expressed proteins. We argue that brain-specific splicing events and organ-restricted protein expression may contribute in part to disparate clinical manifestations. We also highlight the role of genetic modifiers and de novo variants in the multi-organ phenotype of CM and emphasize their consideration in molecular characterization. This review thus attempts to expand our understanding of the phenotypic and etiological variability in CM and invites the development of more comprehensive guidelines.

1. The Definition of Congenital (Primary) Microcephaly and Its Subtypes

Microcephaly opens an aperture to glimpse into the evolution of the human brain, thus uncovering the veils of its architectural orchestration and immense adaptability. Microcephaly is defined as a reduction in the occipital–frontal head circumference (OFC) of at least 2 SD below the mean of age, gender, and ethnicity-matched populations [1]. It underlies the reduced volume of the cerebral cortex, and thus the brain [2]. “Primary” microcephaly means an abnormal OFC at birth [1] and is also known as “congenital microcephaly” (CM) [3]. CM is the currently favored nomenclature and is broadly divided into three categories: (1) isolated microcephaly, that is, without any further clinical (cerebral or extra-cerebral) abnormalities, including no developmental disability or intellectual disability; (2) non-syndromic microcephaly with neurological or psychiatric features, but without cerebral malformations or extra-cerebral abnormalities; and (3) syndromic microcephaly with malformations of the cerebrum and/or extra-cerebral morphologic or functional anomalies. The term microcephaly primary hereditary (MCPH) refers to the genetic forms of CM [4,5]. Adopting these definitions is often difficult and has therefore not been handled uniformly.

2. The Inconsistent Diagnostic Criterion

Clinically, the first step towards the diagnosis of CM is measuring the OFC for a significant reduction. However, different cutoff values are used to define microcephaly. For instance, the American Academy of Neurology and the Practice Committee of the Child Neurology Society have recommended a cutoff of <−2 SD OFC at the time of birth, while <−3 SD would indicate a severe degree of microcephaly [6]. However, multiple studies followed −3 SD as a baseline [5].
The reference values being used to calculate the deviation also vary considerably. As already highlighted by van der Hagen et al. [7], for most of the non-European patients, the SD of OFC is often calculated with reference to values taken from studies performed on Nellhaus (Caucasian children) [8] or Prader (Swiss children) [9], irrespective of the ethnicity of the patients. It should be noted that the mean OFC of children from industrial countries is larger than that shown by the WHO standard values (www.who.int/childgrowth/en accessed on 13 February 2023) [7]. Therefore, it is of critical importance that population-specific data on OFC are taken into consideration when defining microcephaly; the existing data on head circumferences from different populations of Swedish (265,456 singleton neonates), Norwegian (77,044 families), Finnish (533,666 singletons and 15,033 twins) and European, Chinese and South Asian ancestries (2695 infants) may be taken into account for patients of the respective population [10,11,12]. Similarly, the growth charts published by the Centers for Disease Control and Prevention (CDC; www.cdc.gov/growthcharts/ accessed on 13 February 2023) may also be utilized for post-infantile patients of American ancestry.
Furthermore, measuring parental OFCs should be considered for the diagnosis of familial microcephaly. Additionally, a pattern in SD values of OFC of patients with mutations in the same gene could be taken into consideration for a more precise genotype–phenotype correlation.

Inconsistencies with Using the Terms Primary Microcephaly and MCPH

Hereditary CM (MCPH) is sub-categorized—depending on the absence or presence of co-existing features—into isolated, non-syndromic, and syndromic forms. It is important to emphasize that the term “MCPH” is often used for autosomal recessive non-syndromic microcephaly only (see MIM #251200, MCPH1), which is characterized by microcephaly and intellectual deficiency; in some patients, short stature was present. Non-syndromic microcephaly almost exclusively follows an autosomal recessive mode of inheritance and is known to be associated with 30 different genes [13]. However, mutations in one of the MCPH genes, WDFY3, act in an autosomal dominant manner and are associated with MCPH18, characterized by microcephaly with mild to moderate intellectual disability [14].
With an increasing number of patients being clinically and molecularly characterized, mutations in the “non-syndromic MCPH genes” are being reported with a broad range of morphologic or functional anomalies such as brain malformations, epilepsy, behavioral problems, abnormal motor development, ataxia, hearing loss, pigment anomalies, dysmorphism, skeletal anomalies, ophthalmological anomalies, cardiac anomalies, etc. [15]. Consequently, the term “MCPH syndromes” was established [4].
For instance, a recent study reports patients harboring CDK5RAP2 variants manifesting extra-cerebral features such as retinal and cochlear developmental defects and hypothalamic anomalies. Although CDK5RAP2 is traditionally established as a “non-syndromic MCPH gene”, the presence of further anomalies in these patients is more consistent with the criteria of syndromic microcephaly [16]. It may also be noted that clinical data of CDK5RAP2 patients are too sparse in most of the reports to exclude extra-neuroglial features in those patients [16]. Therefore, variants of CDK5RAP2 may also cause multi-organ phenotypes, as suggested by the reports of patients with CDK5RAP2 variants manifesting CM along with sparse eyebrows [17], congenital cataracts [18], and pigmentary abnormalities [15].
Additionally, a large body of data indicates that the phenotypic consequences of different variants of the same gene can be highly divergent depending on the functional consequences [2,19,20,21,22,23]. For instance, a study has recently grouped CEP135-mutated cases of MCPH and MOPD (microcephalic osteodysplastic primordial dwarfism) under the umbrella term of “CEP135 microcephaly” [2].
These observations have led us to believe that it could be more practical to assign the clinical category of CM (isolated, non-syndromic, or syndromic form of microcephaly) to a specific variant of the relevant gene instead of the gene itself. Furthermore, the causality of a gene with specific phenotypic features should be assigned only when ample and consistent clinical and genetic data are available to draw such a conclusion. Towards this aim, it should be helpful to redefine non-syndromic microcephaly by explicitly specifying the associated phenotypes. Another study reported variants of ZNF335 (MCPH10, #615095) that cause “severe autosomal recessive primary microcephaly” even though the patients exhibit cardiac, eye, limb, and digital defects, thus consistent with the definition of syndromic MCPH [24]. Such studies can be refined based on the availability of accumulated data to demonstrate that ZNF335 variants mostly cause multi-organ phenotypes. Otherwise, the observed multi-organ phenotype may be defined as a ZNF335-variant-specific phenotype. The same can be said for other genes such as CEP135 and PCNT which cause both non-syndromic and syndromic microcephaly. In those cases where the same variant of a gene causes both non-syndromic and syndromic microcephaly, genetic modifiers or de novo events should also be taken into consideration when assessing the etiology of a multi-organ phenotype [19].
Furthermore, “ASPM primary microcephaly” based on biallelic pathogenic variants in ASPM is mostly a typical non-syndromic MCPH without extra-cephalic abnormalities, yet still manifesting as an intellectual disability [25]. On the contrary, a few studies are showing an additional phenotype of short stature in patients with ASPM variants [26,27,28,29,30,31]; thus, short stature may be added to the phenotypic criteria of ASPM microcephaly. As a matter of fact, short stature is frequently observed in “non-syndromic MCPH” patients with mutations in MCPH1, WDR62, PHC1, STIL, CDK5RAP2, and KNL1 (formerly known as CASC5) (Table 1) [32,33,34,35,36,37,38,39,40,41,42,43]. Additionally, the association of short stature is shown in 11/21 (52%) affected members of molecularly and clinically diagnosed “non-syndromic MCPH” patients [2]. Interestingly, patients with variants in CEP135, CENPJ, and CEP152 show short stature only when diagnosed with MOPD [2,44]. Apparently, variants in centrosomal genes seem to be more likely to cause multi-organ phenotypes. Nevertheless, establishing a differential, inclusive, and preferably tier-based diagnostic criterion is necessary at this time.
Notably, developmental delay is frequently associated with CM features [45,46,47,48,49], as is evident from the 935 entries of developmental delays in the clinical synopsis of CM in MIM (Mendelian Inheritance in Man) [last checked on 12.09.2022]. Therefore, its inclusion as one of the diagnostic features of CM could be helpful. Similarly, lissencephaly and pachygyria, as well as other architectonic cerebral defects, are frequently observed in patients harboring variants in ASPM, WDR62, CEP135, MCPH1, CIT, CENPJ, and STIL and have also been proposed to be included as MCPH-defining features [2]. Less commonly, mild intrauterine growth restriction with postnatal appearance, subnormal motor development, articulation impairment, spasticity, and dysmorphic facial features have also been observed [25]. Hence, it would be of great diagnostic aid if a uniform list of clinical features is established that would guide the clinicians/researchers when defining CM subtypes. Such reports, though greatly valuable for expanding the clinical spectrum of the mutated genes, are not in line with the conventional diagnostic criteria of MCPH.

3. Genetic Etiology of Microcephaly Primary Hereditary (MCPH)

Hunting for causal gene variants of microcephaly traces back to 1985, when Pérez Castillo published the first microcephaly-linked locus at cytoband 1q31–1q321 [50]. To date, Online Mendelian Inheritance in Man (OMIM) retrieves 1255 entries (last checked on 6 September 2022) with microcephaly in a clinical synopsis section. According to our knowledge, 32 genes have been reported to cause non-syndromic primary microcephaly (Table 2). Interestingly, the 31st and 32nd genes of MCPH i.e., SNRPE and AKNA, respectively, are yet to be updated by OMIM (last checked on 16 December 2022) [51,52]. Among these, ASPM and WDR62 are the most frequently mutated genes, accounting for 68% and 14%, respectively, of the reported families [53,54].
The genetic causes of microcephaly provide insights into the basic processes of brain development and neurogenic mitosis [55]. The majority of variants reported in these genes are frameshift or nonsense variants, eventually leading to truncated, absent, or nonfunctional proteins [56]. MCPH proteins have particular roles in mitosis and are expressed in the neuroepithelium during embryonic neurogenesis [4,57]. The mitosis in neural progenitor cells is surpassingly reliant on the optimal function of these proteins. Any aberrations can cause them to prematurely switch from symmetric proliferative to asymmetric differentiating cell division, thus reducing the progenitor pool and consequently the number of neurons as a whole, producing the MCPH phenotype [58].
The proteins encoded by MCPH-associated genes (Table 2) are expressed ubiquitously and most of them (microcephalin, CDK5RAP2, CENP-J, STIL, CEP135, CEP152, CDK6, HsSAS-6, PDCD6IP, and AKNA) are constituent of either centrosome, spindle pole (WDR62, ASPM, and CENP-E) or spindle microtubules (MAP11), during at least part of the cell cycle [52,59,60,61]. Some MCPH-associated proteins are solely localized at the centrosome; others have a nuclear or cytoplasmic localization in interphase, and a centrosomal localization only in mitotic cells [60,62,63]. Due to the centrosomal localization of several MCPH-associated proteins and their role in centriole duplication, these disorders have also been coined as “centriolopathies” [64]. In addition to the centrosome, several MCPH proteins localize at the kinetochore (KNL1, CENP-E, p37, and hBUB1) and midbody (CRIK, KIF14, and MAP11), particularly during the cell cycle, indicating that dysfunctions of kinetochore and midbody also play critical roles in the etiology of primary microcephaly [13,59,65]. Two of the MCPH proteins, kinetochore scaffold 1 (KNL1) and centromere-associated protein E (CENP-E), are essential components of the kinetochore, where the former controls kinetochore formation and proper segregation of chromosomes during mitosis through interaction with MIS12 whereas the latter helps to capture spindle microtubules to the kinetochore [40,66,67]. Intriguingly, MCPH1-encoded protein microcephalin is reported to promote telomere replication by interacting with the telomere-binding protein TRF2. This complex facilitates the recruitment of DNA damage response factors to the defective telomeres. Defective telomere replication has thus been implicated as the underlying mechanism of microcephaly [68]. A few MCPH patients have been reported with biallelic loss of function mutations in CIT which encodes “citron Rho-interacting kinase (CRIK)” that localizes on the cleavage furrow and midbody of the dividing cells and mediates cytokinesis [69]. After this finding, additional midbody components such as KIF14 and MAP11 were identified as brain size determinants. The components of chromatin-remodeling complexes encoded by ZNF335 and PHC1, the components of the condensin complex encoded by NCAPD2, NCAPH, and NCAPD3, as well as the pre-mRNA-processing spliceosome component encoded by the SNRPE gene represent regulators of mitotic division and their dysfunction in early development, are associated with microcephaly [33,51,70,71]. Other than altered cell cycle monitors, mutations in genes encoding the non-cell-cycle regulators such as sodium-dependent lysophosphatidylcholine symporter 1 (NLS1) encoded by MFSD2A, ankyrin repeat and LEM domain-containing protein 2 encoded by ANKLE2, a subunit of the Golgi coatomer complex (Coatomer subunit beta, COPB2), components of the nuclear lamina encoded by the LMNB1 and LMNB2 genes, and the nucleolar protein (ribosomal RNA processing 7 homolog A) encoded by RRP7A—which is an essential regulator of ribosomal RNA (rRNA) processing—have also been reported in primary microcephaly patients [53,72,73,74,75].
Taken together, the identified MCPH proteins evidently play critical roles in several cellular processes such as cell division and cell cycle regulation, ciliogenesis, cytokinesis, transcription regulation, kinetochore function, chromosome segregation, chromosome condensation, telomere replication, transmembrane or intracellular transport, lipid metabolism and transportation through the blood–brain barrier, ribosomal RNA processing, pre-mRNA processing, DNA damage response, autophagy, and Wnt signaling (Table 2), implicating a crucial role of all these processes for normal brain development [76,77,78].
Table 2. List of genes involved in the etiology of MCPH.
Table 2. List of genes involved in the etiology of MCPH.
LocusGeneProteinSubcellular LocalizationCellular ProcessReferences
MCPH1MCPH1MicrocephalinNucleus (interphase), centrosome (interphase and mitosis)DNA damage signaling and repair, the regulation of chromosome condensation, cell-cycle progression, telomere replication and repair, and centrosome function[68,79,80,81]
MCPH2WDR62WD-repeat containing protein 62Centrosome (interphase), spindle poles (mitosis)Mitogenic kinase signaling, centrosome function, cytoskeletal organization, and cell cycle progression[63,64]
MCPH3CDK5RAP2Cyclin-dependent kinase 5 regulatory subunit-associated protein 2CentrosomeCentrosome function, DNA damage response, and cell cycle progression [4,22]
MCPH4KNL1 * Kinetochore scaffold 1KinetochoreKinetochore assembly, chromosome congression, and mitotic checkpoint signaling[40,82]
MCPH5ASPMAbnormal spindle-like microcephaly-associated proteinCentrosome (interphase), spindle poles (mitosis)Spindle organization, cytokinesis, centriole biogenesis, and microtubule disassembly[83,84]
MCPH6CENPJCentromere protein JCentrosomeCentriole biogenesis, cilium disassembly, and cell cycle progression[22,85]
MCPH7STILSCL/TAL1-interrupting locus proteinCentrosomeCentriole duplication and cell cycle progression[4,86]
MCPH8CEP135Centrosomal protein of 135 kDaCentrosomeCentriole biogenesis[87,88]
MCPH9CEP152Centrosomal protein of 152 kDaCentrosomeCentriole duplication and DNA damage response through ATR-mediated checkpoint signaling[20,21]
MCPH10ZNF335Zinc finger protein 335NucleusTranscription regulation[4,71]
MCPH11PHC1Polyhomeotic-like protein 1NucleusChromatin remodeling, DNA damage and repair, and cell cycle regulation[4,33]
MCPH12CDK6Cyclin-dependent kinase 6Cytosol and nucleus (interphase), centrosome (mitosis) Organization of microtubules, centrosome integrity, and cell proliferation[60]
MCPH13CENPECentromere-associated protein ECentromere, kinetochore, spindle midzoneChromosome alignment and movement toward microtubule bundles [67,89]
MCPH14SASS6Spindle assembly abnormal protein 6 homologCentrosomeCentriole duplication[4,90]
MCPH15MFSD2ASodium-dependent lysophosphatidylcholine symporter 1Cell membraneLipid metabolism and transportation through the blood–brain barrier[91]
MCPH16ANKLE2Ankyrin repeat- and LEM domain-containing protein 2The endoplasmic reticulum, nuclear envelopeMaintain nuclear envelope morphology, cell division, and proliferation[92,93]
MCPH17CITCitron Rho-interacting kinaseCleavage furrow and midbodyCytokinesis[94]
MCPH18WDFY3WD repeat and FYVE domain-containing protein 3Cytosol and nucleusWnt signaling (autophagy-dependent manner)[14]
MCPH19COPB2Coatomer subunit betaGolgi apparatus membraneMay regulate autophagy and maintain the integrity of cellular organelles and cell homeostasis[72]
MCPH20KIF14Kinesin-like protein KIF14Spindle midzone and the midbodyCytokinesis[65]
MCPH21NCAPD2Condensin complex subunit 1Cytoplasm, nucleus (interphase), chromatin (mitosis)Mitotic chromosome condensation[70]
MCPH22NCAPD3Condensin-2 complex subunit D3Cytoplasm, nucleus (interphase), chromatin (mitosis)Mitotic chromosome condensation[70]
MCPH23NCAPHCondensin complex subunit 2 Cytoplasm, nucleus (interphase), chromatin (mitosis)Mitotic chromosome condensation[70]
MCPH24NUP37Nucleoporin Nup37Nuclear envelop (interphase), kinetochore (mitosis)Cell cycle progression[73]
MCPH25MAP11Trafficking protein particle complex subunit 14Spindle microtubules, cleavage furrow and midbodyCytokinesis and cell abscission[59]
MCPH26LMNB1Lamin-B1Nuclear laminaCell cycle regulation, transcription regulation, and DNA repair[74,95]
MCPH27LMNB2Lamin-B2Nuclear laminaTranscription regulation, mitosis, chromosome segregation, and nucleolar morphology[74,95]
MCPH28RRP7ARibosomal RNA-processing protein 7 homolog ANucleolus, centrosomes and ciliaRibosomal RNA processing, primary cilia resorption, and cell cycle progression[75]
MCPH29PDCD6IPProgrammed cell death 6-interacting proteinNucleusCytokinesis, cell proliferation, and abscission and autophagy[61]
MCPH30BUB1BUB1 mitotic checkpoint serine/threonine kinaseNucleus (interphase), kinetochore (mitosis)Chromosome congression and spindle assembly checkpoint[13]
MCPH31SNRPESmall nuclear ribonucleoprotein ENucleusPre-mRNA processing[51]
MCPH32AKNAMicrotubule organization protein AKNACentrosomeCentrosomal microtubule organization[52]
Note: Gene and respective encoded protein names are verified from HUGO (Human Genome Organization), the Gene Nomenclature Committee (HGNC) and UniProt Knowledgebase (UniProtKB), respectively. * This gene was previously known as CASC5. ATR stands for ataxia-telangiectasia mutated and Rad3-related.

4. Model Systems for Microcephaly and Pathomechanisms

The underlying mechanisms of this condition are inferred from the known functions of MCPH proteins investigated during mouse embryonic brain development as well as in cerebral organoids. Knockout mice models of MCPH causative genes (Mcph1, Wdr62, Cdk5rap2, Aspm, Cenpj, Znf335, Phc1, ankle2 (zebrafish), Wdfy3 and Cit) mimic the phenotype of reduced brain size discerned in the MCPH patients [33,56,96,97]. Defective DNA damage response (DDR) is also proposed as an underlying mechanism of microcephaly. Ablation of Mcph1 in mice resulted in elevated genomic instability due to defective homologous recombination repair and mimic microcephaly phenotype. Interestingly, neural progenitor cells (NPCs) examined from these mice were also reported to be vulnerable to ionizing radiation [81]. Similarly, Knl1-deficient mice also showed microcephaly as a result of enhanced DNA damage and p53 activation in NPCs [98]. Furthermore, microcephaly observed in the Cenpj-deficient mice was the result of elevated levels of DNA damage and apoptosis, thus favoring DDR as a crucial process to control brain size [99]. Kif14-knockout mice show small brain size, primarily caused by elevated apoptosis during late neurogenesis in these mice [100]. Mfsd2a-knockout mice demonstrate a severe neurodevelopmental phenotype, including microcephaly, anxiety, cognitive defects, and ataxia leading to postnatal lethality [101,102]. Furthermore, condensin II (Ncaph2) mutant mice showed significantly reduced brain size, accompanied by chromosome segregation errors, leading to micronucleus formation and increased aneuploidy in daughter cells [70]. In addition, the ablation of Cenpj, Knl1, Stil, Cep135, Cenpe, and Copb2 (the orthologs of MCPH3, MCPH4, MCPH7, MCPH8 MCPH13, and MCPH19 loci, respectively) results in the embryonic lethality of mice [56,72,99,103,104]. Conclusively, the viability of MCPH patients mutated for these genes could possibly be explained by the hypomorphic nature of the identified mutations, as compared to more damaging nullimorphic alleles in mice [72,105]. Other than mouse models, microcephaly has also been demonstrated in the cerebral organoids developed from CDK5RAP2 and ASPM mutated patient-derived induced pluripotent stem (iPS) cells, which indicated premature neuronal differentiation as a pathomechanism for the former gene and impaired neuronal functions (due to defective lamination), as well as decreased neural progenitor pool for the later gene [106,107]. Similarly, premature neuronal progenitor differentiation has been proven by yet another study using cerebral organoids developed from iPS cells of Seckel syndrome patients mutated with CENPJ, where the authors claim delayed cell cycle entry as a consequence of impaired cilium disassembly [85]. These reports strengthen the conserved and indispensable role of MCPH genes in brain development. Based on the knowledge gained from these models, two different underlying mechanisms—centrosomal and non-centrosomal—emerge for primary microcephaly. In the centrosomal model, it is speculated that the defective MCPH proteins affect microtubule organization at the centrosome in neural progenitor cells, resulting in premature switching from both symmetric proliferative [108] and asymmetric self-renewing to symmetric consumptive division [109]. The cells prematurely lose their proliferative capacity, causing early depletion of the pool of progenitors, and consequently a smaller brain [108,110]. In the non-centrosomal model of microcephaly, the spindle orientation is not altered, yet the depletion of neural progenitor cells can be caused by other mechanisms such as chromosome segregation defects, which lead to aneuploid cells, eventually destined to apoptosis. This leaves too few progenitors to constitute a normal cortical volume, resulting in primary microcephaly [70,111]. In addition to these two models, perturbation in the migration of neurons and glial cells has also been suggested as the cause of primary microcephaly in patients carrying mutations in WDR62 [112].

5. The Shift from Non-Syndromic to Syndromic Microcephaly Remains Unsolved

The non-linear genotype–phenotype relationship in microcephaly patients has been hugely debated in recent years. Most of the MCPH genes are globally expressed but their mutations exceedingly compromise only brain development, a phenomenon that remains only vaguely understood [2]. Below, we discuss some arguments that might explain the etiologic dichotomy of non-syndromic and syndromic microcephaly.

5.1. The Fragility of Asymmetric Neuronal Division

During neurogenesis, the formation of a neural tube requires precisely orchestrated temporal and spatial arrangements of cell division [113]. This process is highly sophisticated and demands an organized fashion and proper regulation at every step. As explained above, embryonic neurogenesis conceivably only shows exclusive fragility in this context, and any dysregulation in the balance between symmetric and asymmetric cell division, such as premature switching from symmetric proliferative to asymmetric apical progenitor division [108], and from asymmetric self-renewing to symmetric consumptive apical progenitor division, considerably reduces the number of neural progenitor cells, leaving too few of them to constitute a normal-sized brain [108]. Furthermore, spindle pole formation and centrosome duplication, which serve to maintain polarity to neuronal cells, are also proven to be extremely crucial events in neuronal duplication. Interestingly, centrosomal abnormalities show a less profound influence on extra neurogenic mitosis, as shown by studies of neuroblasts in Drosophila [114]. It is worth mentioning that most of the centrosomal proteins function together; MCPH1 and WDR62 are physically interacting partners and make a complex with Cep63 as well, and the knockdown of CEP63 abolishes MCPH1-WDR62 interaction [64]. In addition, Cep63, ASPM, and WDR62 recruit CENP-J (also known as CPAP) to carry out centriole biogenesis [64]. If we trace expression patterns, it is seen that all of them are highly expressed in the cerebral cortex, localized in close intracellular vicinity of each other (centrosome/spindle pole), and are interconnected, indicating that they mediate each other’s function. Not only do these make complexes but they also exhibit functional dependency, which collectively supports the idea of the sensitivity of asymmetric division towards the dysfunctioning of either of these proteins. Hence, it is apparent that neuronal cells are specifically sensitive to centrosomal defects, leading to depletion of the neural progenitor pool.

5.2. Role of Genetic Modifiers/Multiple Allelism

Microcephaly has long been classified as a pure monogenic yet heterogeneous disorder; however, recent pieces of evidence for an oligogenic model of microcephaly are also emerging. The importance of genetic modifiers in phenotypic variability has attained special attention in recent years. A genetic modifier can add to the phenotypic variability due to functional co-dependencies among the genes of related pathways or genetic interaction by enhancing or alleviating the severity of the disease [115]. Due to the reduced cost of genomic sequencing, it is now increasingly feasible to hunt for additional pathogenic variants. The presence of more than one pathogenic variant can completely shift the phenotypic paradigm, as has been shown in a recent study that elegantly proves this fact on functional grounds. In a multiplex Pakistani family, the severity of phenotypes was noted in patients carrying the heterozygous PCNT variant (modifier), additionally harboring the homozygous variant of CENPJ, as compared to those cases where PCNT mutation was not segregated [19]; one loop manifested in Seckel syndrome caused by a pathogenic CENPJ variant (NM_018451.4:c.3586G > A; p.(Asp1196Asn)) which was also segregating in the second loop. However, the second loop manifested microcephalic osteodysplastic primordial dwarfism type II (MOPDII)—a phenotypically severe condition—due to the presence of an additional PCNT variant (NM_006031.5;c.5767C > T; p.(Arg1923*)). Further, this study also provided evidence that heterozygous missense variants of WDR62, CEP63, and RAD50 worsen the MCPH phenotype in patients harboring a homozygous ASPM mutation [19]. Another study also reported the severity of the MCPH phenotype caused by genetic modifiers. In this report a homozygous stop mutation, c.1605dupT; p.(Glu536*) of WDR62, caused MCPH in one patient, whereas the other affected members presented severe phenotypic manifestations in the presence of a chromosomal duplication of 17q25-qter on one allele and a missense mutation, c.3361T > G; p.(Phe1121Val) of TBCD, on a second non-duplicated allele [116]. Conclusively, the authors proposed that the modifying effect of TBCD resulted in the severity of the clinical manifestations in the patients. Interestingly, digenic tri-allelic inheritance in MCPH patients, with the combination of genes WDR62/CDK5RAP2, ASPM/WDR62, and WDR62/CEP135 has also been reported. However, the authors did not show the modifying effects of additional variants on the clinical manifestations of the patients. It should be noted that they did not observe any modifying effects of wdr62 or aspm ablations in zebrafish upon knockout of knl1. However, the quadri-allelic ablation of wdr62 and aspm demonstrated primary microcephaly in zebrafish [117].
In this context, the role of modifiers in exacerbating the phenotype has also been reported in animal models. A previous study analyzed similar modifier effects for Gpr63 in mice, where Ttc21b was ablated in homozygous form; Ttc21b is already known to control brain size, and mutations in this gene cause ciliopathies, including microcephaly [118]. A CRISPR/Cas9-generated heterozygous mutant Gpr63 allele in Ttc21b null showed an increased incidence of neural tube defects, which was not observed in Ttc21b homozygous mutants. Intriguingly, a further severity of phenotypes (early embryonic lethality, neural tube defects, exencephaly, polydactyly of forelimbs and hindlimbs, and spina bifida aperta) was observed in Gpr63 and Ttc21b double homozygous mutants, which were not shown in the single mutants [118].
Additionally, genes encoding centrosomal proteins such as ASPM and WDR62 when mutated cause centrosomal defects leading primarily to the MCPH phenotype. Interestingly, tri-allelism (knockout of two mutant copies of one gene and only one mutant copy of another gene), Wdr62 (+/−), and Aspm (−/−), in mice embryos show comparatively severe cellular defects, and thus microcephaly, showing predominant effects on late neurogenesis. A similar effect was noticed in the case of Wdr62 (−/−) and Aspm (+/−), suggesting severe mutational effects. However, in the case of quadriallelism, the double-knockout mice of Wdr62 (−/−) and Aspm (−/−) showed early embryonic lethality [64]. A similar concept was supported by another study where multiple allelism was analyzed in zebrafish; a double-knockout of genes encoding centrosomal proteins such as Aspm and Wdr62 and one non-centrosomal Knl1. The latter demonstrated no exacerbation of the phenotype. This shows that oligogenic inheritance may show a variable phenotype when the genetic variants are functionally relevant and/or have genetic interaction in one way or another [117].

5.3. Altered Splicing Events

The concept of tissue-specific splicing factors, and hence tissue-specific splicing, is a very well-established phenomenon and can possibly contribute to explaining brain-specific phenotype [119]. It is indicated in a recent study where MCPH-protein KNL1 has been investigated in cerebroids and neural progenitor cells generated by reprogramming the human embryonic stem cells (hESCs) harboring a CRISPR/Cas9-mediated knock-in missense variant (c.6125G > A; p.Met2041Ile) of KNL1 [82]. They observed that the mutation inhibited an exon splicing enhancer (ESE) site and generated an exon splicing silencer (ESS) site. This newly generated ESS site is recognized by the heterogeneous nuclear ribonucleoprotein A1 (HNRNPA1) which acts as an inhibitory splicing factor and thus lowers the expression of KNL1. Authors revealed that HNRNPA1 is dominantly expressed in the brain and only mildly expressed in human fibroblasts and neural crest cells. Therefore, it affects the expression of KNL1 in the brain only, but not in primary fibroblast and neural crest cells. The absence of KNL1, a crucial mediator of chromosome alignments during cell division, results in aneuploidy, leading to reduced cell growth, apoptosis, and reduced size of the cerebroids. However, the patient-derived cells (primary fibroblasts and lymphoblastoid cells), as well as primary fibroblasts and neural crest cells differentiated from isogenic mutant hESC cells, did not show any aberrant cellular phenotype [40,82]. As the mutational effects were observed only in the neural progenitors, this study concluded that the KNL1 mutation exclusively affects neurogenesis, thus causing a brain-restricted phenotype.
KNL1 is not the only gene that has a splice site variant; MCPH genes such as MCPH1, WDR62, CDK5RAP2, ASPM, CEP135, ZNF335, SASS6, MFSD2A, CIT, KIF14, NCAPD2, NCAPD3, and LMNB1 with splice site variants were also reported in MCPH patients [15,63,69,70,71,86,102,120,121,122,123,124]. These splice variants may act in the same manners as observed in the case of KNL1 splice variants and may explain the phenotype of isolated microcephaly. Furthermore, a recent study shows SNRPE variants causing congenital non-syndromic microcephaly; this gene encodes SmE proteins involved in the assembly of the spliceosome complex. The mutation impaired the assembly of the spliceosome components, which leads to the global dysregulation of mRNA splicing in patient-derived human primary fibroblasts and SmE-depleted zebrafish, as well as HEK293 cell line [51]. The RNA profiling of morpholino-injected zebrafish (head and tail only) showed the clustering of retained introns in several transcripts due to SmE deficiency. Notably, authors observed a distinguishable expression pattern of transcripts of head and tail regions with the majority of the downregulated genes enriched for the development of the central nervous system. These data suggest that neurogenesis/brain development is exclusively vulnerable to the splicing alterations and cause a brain-specific phenotype, as the patient of this study presented a non-syndromic form of microcephaly only [51].

5.4. Hypomorphic and Loss-of-Function Variants

It has been indicated that the mutations causing complete loss of the MCPH protein may have severe, multi-organ manifestations (syndromic microcephaly), whereas mutations partially compromising the function of the protein may cause a brain-restricted phenotype [125]. This argument is elegantly supported by establishing the genotype–phenotype correlation of different components of the nuclear pore complex, p37, NUP107, and NUP133, which showed phenotypic variability in microcephalics; it was intriguing to observe that patients segregating a hypomorphic NUP37 mutation did not show any symptoms of steroid-resistant nephrotic syndrome (SRNS), which was diagnosed only in the individuals segregating the NUP107 loss-of-function variant [73]. Authors modeled both hypomorphic and null mutations of NUP107 in zebrafish by CRISPR/Cas9-mediated genome editing and showed that truncating mutation (c.50_56del7; p.Thr81Argfs*74) causes early lethality and developmental defects including microcephaly, whereas in-frame hypomorphic mutation (c.137_139del; p.Ala46del) was compatible with embryonic survival and did not cause any obvious phenotype [73]. These findings thus reflect the inverse relation of a residual function of the gene with the general severity of the disease, broadening the phenotypic spectra [53].
Additionally, the loss-of-function variant in NUP133—the direct binding partner of NUP107—leads to Galloway–Mowat syndrome, which is characterized by microcephaly associated with early onset nephrotic syndrome and hiatus hernia. The mutation causes an abrogation of interaction between NUP107 and NUP133, leading to a reduction in both proteins, which can be related to the severity of the disease [126]. Thus, it is convincing to deduce that hypomorphic mutations show milder phenotypes, whereas loss-of-function mutations lead to severe medical consequences affecting brain development along with other organs such as kidneys in SRNS. Intriguingly, hypomorphic variants of CENPJ and CEP152 are reported to cause only a brain-restricted phenotype (non-syndromic microcephaly), whereas complete loss-of-function variants of these genes result in Seckel syndrome [21,127].

5.5. De Novo Autosomal Dominant Mutations

De novo variants are also reported to cause primary microcephaly. A study presented seven individuals who were reported with de novo variants in LMNB1, all manifesting primary microcephaly associated with additional features such as relatively short stature, neurogenic scoliosis, and severe feeding difficulties. Two of these individuals segregating splice variants causing exon elongation also demonstrated hypotonia, spastic tetraparesis, and gingival hypertrophy. The head circumferences ranged from −3.6 to −12 SD, showing the severity of the disease [123]. Similarly, de novo mutations in KAT6A, TUBB5, ASXL3, SPOCK1, CTNNB1, and KIF11 were also observed to cause severe syndromic microcephaly [128,129,130,131,132,133]. A recent Korean study presented a cohort of 34 patients with primary microcephaly and 6 patients with postnatal microcephaly. Authors identified de novo mutations in 10 genes (GNB1, GNAO1, TCF4, ASXL1, SMC1A, KMT2A, VPS13B, ACTG1, EP300, and KMT2D) segregating in the microcephalics with a broad phenotypic spectrum, including developmental delay and infantile hypotonia consistent in all, variably associated with a movement disorder (in 7 patients) and other anomalies (congenital heart defect, cleft lip/palate, and skeletal anomaly) in 17 patients [48]. Furthermore, another study showed that a de novo heterozygous synonymous variant of EFTUD2, which encodes a spliceosomal component, caused postnatal microcephaly associated with mandibulofacial dysostosis (hearing loss, epileptic seizures, micro retrognathism, malar hypoplasia, and dental anomalies) [134]. Taken together, these studies indicate that the severity of disease in CM cases is dependent on the nature of the mutation, and thus the syndromic cases should be explored for the de novo variants as the first line of strategy. However, the possibility of the de novo variants could not be ruled out, particularly in those cases carrying homozygous/compound heterozygous variants manifesting severe phenotypes.

6. Conclusions

In this review article, we have highlighted the inconsistencies in the differential diagnosis of congenital microcephaly (CM) and outlined the accumulating evidence that should be critically considered when evaluating and defining CM subtypes. Keeping in mind the limited understanding of the molecular mechanism of the paradigm shift from non-syndromic to syndromic microcephaly, we have documented the molecular causes necessary for the shift from brain-restricted to extra-neurological features in CM cases. In conclusion, we recommend that genetic modifiers, de novo variants, and brain-specific splicing events be considered when evaluating the spectrum from brain-specific to multi-organ phenotypes associated with particular genes.

Author Contributions

Conceptualization, M.A. and M.S.H.; methodology, M.A. and U.A.; validation, M.A. and U.A.; investigation, M.A., U.A. and M.S.H.; resources, P.N. and M.S.H.; data curation, M.A., U.A. and M.S.H.; writing—original draft preparation, M.A., U.A., S.T. and M.S.H.; writing—review and editing, P.N., S.T. and M.S.H.; visualization, P.N. and S.T.; supervision, P.N. and M.S.H.; project administration, M.S.H.; funding acquisition, P.N. and P.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Koeln Fortune Program (Faculty of Medicine, University of Cologne; 381/2020 to M.S.H) and the Center for Molecular Medicine Cologne (CMMC) (Projects 38-RP and C12; 2635/8029/01 and 2635/8326/01 to P.N. and M.S.H.).

Institutional Review Board Statement

This study was conducted in accordance with the Declaration of Helsinki and approved by the Ethics Committee for the Use of Human Subjects of Pir Mehr Ali Shah Arid Agriculture University Rawalpindi (PMAS-AAUR/IEC-21/2019).

Informed Consent Statement

Not valid.

Data Availability Statement

Not valid.

Acknowledgments

We are thankful to the scientific community working on microcephaly and related disorders for generating valuable data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Opitz, J.; Holt, M. Microcephaly: General considerations and aids to nosology. J. Craniofacial Genet. Dev. Biol. 1990, 10, 175–204. [Google Scholar]
  2. Shaheen, R.; Maddirevula, S.; Ewida, N.; Alsahli, S.; Abdel-Salam, G.M.; Zaki, M.S.; Tala, S.A.; Alhashem, A.; Softah, A.; Al-Owain, M. Genomic and phenotypic delineation of congenital microcephaly. Genet. Med. 2019, 21, 545–552. [Google Scholar] [CrossRef]
  3. DeSilva, M.; Munoz, F.M.; Sell, E.; Marshall, H.; Kawai, A.T.; Kachikis, A.; Heath, P.; Klein, N.P.; Oleske, J.M.; Jehan, F. Congenital microcephaly: Case definition & guidelines for data collection, analysis, and presentation of safety data after maternal immunisation. Vaccine 2017, 35, 6472. [Google Scholar]
  4. 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] [Green Version]
  5. Piro, E.; Antona, V.; Consiglio, V.; Ballacchino, A.; Graziano, F. Microcephaly a clinical-genetic and neurologic approach. Acta Med. Mediterr. 2013, 29, 327–331. [Google Scholar]
  6. Ashwal, S.; Michelson, D.; Plawner, L.; Dobyns, W.B. Practice parameter: Evaluation of the child with microcephaly (an evidence-based review): Report of the Quality Standards Subcommittee of the American Academy of Neurology and the Practice Committee of the Child Neurology Society. Neurology 2009, 73, 887–897. [Google Scholar] [CrossRef] [Green Version]
  7. Von der Hagen, M.; Pivarcsi, M.; Liebe, J.; Von Bernuth, H.; Didonato, N.; Hennermann, J.B.; Bührer, C.; Wieczorek, D.; Kaindl, A.M. Diagnostic approach to microcephaly in childhood: A two-center study and review of the literature. Dev. Med. Child Neurol. 2014, 56, 732–741. [Google Scholar] [CrossRef]
  8. Nellhaus, G. Head circumference from birth to eighteen years: Practical composite international and interracial graphs. Pediatrics 1968, 41, 106–114. [Google Scholar] [CrossRef]
  9. Prader, A.; Largo, R.H.; Molinari, L.; Issler, C. Physical growth of Swiss children from birth to 20 years of age. First Zurich longitudinal study of growth and development. Helv. Paediatr. Acta. Suppl. 1989, 52, 1–125. [Google Scholar]
  10. Elvander, C.; Högberg, U.; Ekeus, C. The influence of fetal head circumference on labor outcome: A population-based register study. Acta Obstet. Et Gynecol. Scand. 2012, 91, 470–475. [Google Scholar] [CrossRef]
  11. Lunde, A.; Melve, K.K.; Gjessing, H.K.; Skjærven, R.; Irgens, L.M. Genetic and environmental influences on birth weight, birth length, head circumference, and gestational age by use of population-based parent-offspring data. Am. J. Epidemiol. 2007, 165, 734–741. [Google Scholar] [CrossRef] [Green Version]
  12. Janssen, P.A.; Thiessen, P.; Klein, M.C.; Whitfield, M.F.; MacNab, Y.C.; Cullis-Kuhl, S.C. Standards for the measurement of birth weight, length and head circumference at term in neonates of European, Chinese and South Asian ancestry. Open Med. 2007, 1, e74. [Google Scholar]
  13. 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. Biallelic BUB1 mutations cause microcephaly, developmental delay, and variable effects on cohesion and chromosome segregation. Sci. Adv. 2022, 8, eabk0114. [Google Scholar] [CrossRef]
  14. 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] [Green Version]
  15. Pagnamenta, A.T.; Howard, M.F.; Knight, S.J.; Keays, D.A.; Quaghebeur, G.; Taylor, J.C.; Kini, U. Activation of an exonic splice-donor site in exon 30 of CDK5RAP2 in a patient with severe microcephaly and pigmentary abnormalities. Clin. Case Rep. 2016, 4, 952. [Google Scholar] [CrossRef] [Green Version]
  16. Nasser, H.; Vera, L.; Elmaleh-Bergès, M.; Steindl, K.; Letard, P.; Teissier, N.; Ernault, A.; Guimiot, F.; Afenjar, A.; Moutard, M.L. CDK5RAP2 primary microcephaly is associated with hypothalamic, retinal and cochlear developmental defects. J. Med. Genet. 2020, 57, 389–399. [Google Scholar] [CrossRef]
  17. Abdullah, U.; Farooq, M.; Mang, Y.; Bakhtiar, S.M.; Fatima, A.; Hansen, L.; Kjaer, K.W.; Larsen, L.A.; Faryal, S.; Tommerup, N.; et al. A novel mutation in CDK5RAP2 gene causes primary microcephaly with speech impairment and sparse eyebrows in a consanguineous Pakistani family. Eur. J. Med. Genet. 2017, 60, 627–630. [Google Scholar] [CrossRef]
  18. 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]
  19. Makhdoom, E.U.H.; Waseem, S.S.; Iqbal, M.; Abdullah, U.; Hussain, G.; Asif, M.; Budde, B.; Höhne, W.; Tinschert, S.; Saadi, S.M. 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]
  20. Guernsey, D.L.; Jiang, H.; Hussin, J.; Arnold, M.; Bouyakdan, K.; Perry, S.; Babineau-Sturk, T.; Beis, J.; Dumas, N.; Evans, S.C. 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]
  21. 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] [PubMed] [Green Version]
  22. 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]
  23. 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] [PubMed]
  24. Stouffs, K.; Stergachis, A.B.; Vanderhasselt, T.; Dica, A.; Janssens, S.; Vandervore, L.; Gheldof, A.; Bodamer, O.; Keymolen, K.; Seneca, S. Expanding the clinical spectrum of biallelic ZNF335 variants. Clin. Genet. 2018, 94, 246–251. [Google Scholar] [CrossRef]
  25. Verloes, A.; Drunat, S.; Passemard, S. ASPM Primary Microcephaly; 2 April 2020; GeneReviews® [Internet]; Adam, M.P., Everman, D.B., Mirzaa, G.M., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  26. Abdel-Hamid, M.S.; Ismail, M.F.; Darwish, H.A.; Effat, L.K.; Zaki, M.S.; Abdel-Salam, G.M. Molecular and phenotypic spectrum of ASPM-related primary microcephaly: Identification of eight novel mutations. Am. J. Med. Genet. Part A 2016, 170, 2133–2140. [Google Scholar] [CrossRef]
  27. Papari, E.; Bastami, M.; Farhadi, A.; Abedini, S.; Hosseini, M.; Bahman, I.; Mohseni, M.; Garshasbi, M.; Moheb, L.A.; Behjati, F. Investigation of primary microcephaly in Bushehr province of Iran: Novel STIL and ASPM mutations. Clin. Genet. Int. J. Genet. Med. 2013, 83, 488–490. [Google Scholar]
  28. 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]
  29. Hashmi, J.A.; Al-Harbi, K.M.; Ramzan, K.; Albalawi, A.M.; Mehmood, A.; Samman, M.I.; Basit, S. A novel splice-site mutation in the ASPM gene underlies autosomal recessive primary microcephaly. Ann. Saudi Med. 2016, 36, 391–396. [Google Scholar] [CrossRef] [Green Version]
  30. Darvish, H.; Esmaeeli-Nieh, S.; Monajemi, G.; Mohseni, M.; Ghasemi-Firouzabadi, S.; Abedini, S.; Bahman, I.; Jamali, P.; Azimi, S.; Mojahedi, F. A clinical and molecular genetic study of 112 Iranian families with primary microcephaly. J. Med. Genet. 2010, 47, 823–828. [Google Scholar] [CrossRef] [Green Version]
  31. 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. Expanding the clinical and neuroradiologic phenotype of primary microcephaly due to ASPM mutations. Neurology 2009, 73, 962–969. [Google Scholar] [CrossRef]
  32. Mouden, C.; de Tayrac, M.; Dubourg, C.; Rose, S.; Carre, W.; Hamdi-Roze, H.; Babron, M.-C.; Akloul, L.; Heron-Longe, B.; Odent, S. Homozygous STIL mutation causes holoprosencephaly and microcephaly in two siblings. PLoS ONE 2015, 10, e0117418. [Google Scholar] [CrossRef]
  33. 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]
  34. Garshasbi, M.; Motazacker, M.M.; Kahrizi, K.; Behjati, F.; Abedini, S.S.; Nieh, S.E.; Firouzabadi, S.G.; Becker, C.; Rüschendorf, F.; Nürnberg, P. SNP array-based homozygosity mapping reveals MCPH1 deletion in family with autosomal recessive mental retardation and mild microcephaly. Hum. Genet. 2006, 118, 708–715. [Google Scholar] [CrossRef] [PubMed]
  35. Caraffi, S.G.; Pollazzon, M.; Farooq, M.; Fatima, A.; Larsen, L.A.; Zuntini, R.; Napoli, M.; Garavelli, L. MCPH1: A Novel Case Report and a Review of the Literature. Genes 2022, 13, 634. [Google Scholar] [CrossRef] [PubMed]
  36. 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]
  37. Cherkaoui Jaouad, I.; Zrhidri, A.; Jdioui, W.; Lyahyai, J.; Raymond, L.; Egéa, G.; Taoudi, M.; El Mouatassim, S.; Sefiani, A. A novel non sense mutation in WDR62 causes autosomal recessive primary microcephaly: A case report. BMC Med. Genet. 2018, 19, 118. [Google Scholar] [CrossRef] [PubMed]
  38. 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. (1971-) 2022, 191, 2733–2741. [Google Scholar] [CrossRef] [PubMed]
  39. Saadi, A.; Verny, F.; Siquier-Pernet, K.; Bole-Feysot, C.; Nitschke, P.; Munnich, A.; Abada-Dendib, M.; Chaouch, M.; Abramowicz, M.; Colleaux, L. Refining the phenotype associated with CASC5 mutation. Neurogenetics 2016, 17, 71–78. [Google Scholar] [CrossRef] [PubMed]
  40. 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] [PubMed] [Green Version]
  41. 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] [Green Version]
  42. 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]
  43. Rasool, S.; Baig, J.M.; Moawia, A.; Ahmad, I.; Iqbal, M.; Waseem, S.S.; Asif, M.; Abdullah, U.; Makhdoom, E.U.H.; Kaygusuz, E. 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]
  44. Rauch, A. The shortest of the short: Pericentrin mutations and beyond. Best Pract. Res. Clin. Endocrinol. Metab. 2011, 25, 125–130. [Google Scholar] [CrossRef]
  45. Aggarwal, A.; Mittal, H.; Patil, R.; Debnath, S.; Rai, A. Clinical profile of children with developmental delay and microcephaly. J. Neurosci. Rural Pract. 2013, 4, 288–291. [Google Scholar] [CrossRef]
  46. Derwińska, K.; Smyk, M.; Cooper, M.L.; Bader, P.; Cheung, S.W.; Stankiewicz, P. PTCH1 duplication in a family with microcephaly and mild developmental delay. Eur. J. Hum. Genet. 2009, 17, 267–271. [Google Scholar] [CrossRef] [Green Version]
  47. Yatsuka, Y.; Kishita, Y.; Formosa, L.E.; Shimura, M.; Nozaki, F.; Fujii, T.; Nitta, K.R.; Ohtake, A.; Murayama, K.; Ryan, M.T. A homozygous variant in NDUFA8 is associated with developmental delay, microcephaly, and epilepsy due to mitochondrial complex I deficiency. Clin. Genet. 2020, 98, 155–165. [Google Scholar] [CrossRef]
  48. Lee, J.; Park, J.E.; Lee, C.; Kim, A.R.; Kim, B.J.; Park, W.-Y.; Ki, C.-S.; Lee, J. Genomic analysis of Korean patients with microcephaly. Front. Genet. 2020, 11, 1850. [Google Scholar] [CrossRef]
  49. Naseer, M.I.; Rasool, M.; Abdulkareem, A.A.; Chaudhary, A.G.; Zaidi, S.K.; Al-Qahtani, M.H. Novel compound heterozygous mutations in WDR62 gene leading to developmental delay and Primary Microcephaly in Saudi Family. Pak. J. Med. Sci. 2019, 35, 764. [Google Scholar] [CrossRef] [Green Version]
  50. Pérez-Castillo, A.; Martín-Lucas, M.A.; Abrisqueta, J. Is a gene for microcephaly located on chromosome 1? Hum. Genet. 1984, 67, 230–232. [Google Scholar] [CrossRef]
  51. Chen, T.; Zhang, B.; Ziegenhals, T.; Prusty, A.B.; Fröhler, S.; Grimm, C.; Hu, Y.; Schaefke, B.; Fang, L.; Zhang, M. A missense mutation in SNRPE linked to non-syndromal microcephaly interferes with U snRNP assembly and pre-mRNA splicing. PLoS Genet. 2019, 15, e1008460. [Google Scholar] [CrossRef] [Green Version]
  52. Waseem, S.S.; Moawia, A.; Budde, B.; Tariq, M.; Khan, A.; Ali, Z.; Khan, S.; Iqbal, M.; Malik, N.A.; Haque, S.u. A Homozygous AKNA Frameshift Variant Is Associated with Microcephaly in a Pakistani Family. Genes 2021, 12, 1494. [Google Scholar] [CrossRef] [PubMed]
  53. Duerinckx, S.; Abramowicz, M. The genetics of congenitally small brains. Semin. Cell Dev. Biol. 2018, 76, 76–85. [Google Scholar] [CrossRef] [PubMed]
  54. Zaqout, S.; Morris-Rosendahl, D.; Kaindl, A.M. Autosomal Recessive Primary Microcephaly (MCPH): An Update. Neuropediatrics 2017, 48, 135–142. [Google Scholar] [CrossRef] [PubMed]
  55. 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]
  56. Barbelanne, M.; Tsang, W.Y. Molecular and cellular basis of autosomal recessive primary microcephaly. BioMed Res. Int. 2014, 2014, 547986. [Google Scholar] [CrossRef] [Green Version]
  57. Cox, J.; Jackson, A.P.; Bond, J.; Woods, C.G. What primary microcephaly can tell us about brain growth. Trends Mol. Med. 2006, 12, 358–366. [Google Scholar] [CrossRef]
  58. Wollnik, B. A common mechanism for microcephaly. Nat. Genet. 2010, 42, 923–924. [Google Scholar] [CrossRef]
  59. 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] [Green Version]
  60. Hussain, M.S.; Baig, S.M.; Neumann, S.; Peche, V.S.; Szczepanski, S.; Nurnberg, 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]
  61. 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]
  62. Kaindl, A.M.; Passemard, S.; Kumar, P.; Kraemer, N.; Issa, L.; Zwirner, A.; Gerard, B.; Verloes, A.; Mani, S.; Gressens, P. Many roads lead to primary autosomal recessive microcephaly. Prog. Neurobiol. 2010, 90, 363–383. [Google Scholar] [CrossRef]
  63. 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]
  64. 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]
  65. 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] [PubMed]
  66. Kline, S.L.; Cheeseman, I.M.; Hori, T.; Fukagawa, T.; Desai, A. The human Mis12 complex is required for kinetochore assembly and proper chromosome segregation. J. Cell Biol. 2006, 173, 9–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. 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]
  68. Cicconi, A.; Rai, R.; Xiong, X.; Broton, C.; Al-Hiyasat, A.; Hu, C.; Dong, S.; Sun, W.; Garbarino, J.; Bindra, R.S. Microcephalin 1/BRIT1-TRF2 interaction promotes telomere replication and repair, linking telomere dysfunction to primary microcephaly. Nat. Commun. 2020, 11, 5861. [Google Scholar] [CrossRef] [PubMed]
  69. Basit, S.; Al-Harbi, K.M.; Alhijji, S.A.; Albalawi, A.M.; Alharby, E.; Eldardear, A.; Samman, M.I. CIT, a gene involved in neurogenic cytokinesis, is mutated in human primary microcephaly. Hum. Genet. 2016, 135, 1199–1207. [Google Scholar] [CrossRef]
  70. 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]
  71. Yang, Y.W.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]
  72. 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]
  73. 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] [PubMed] [Green Version]
  74. Parry, D.A.; Martin, C.-A.; Greene, P.; Marsh, J.A.; Blyth, M.; Cox, H.; Donnelly, D.; Greenhalgh, L.; Greville-Heygate, S.; Harrison, V. 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]
  75. Farooq, M.; Lindbæk, L.; Krogh, N.; Doganli, C.; Keller, C.; Mönnich, M.; Gonçalves, A.B.; Sakthivel, S.; Mang, Y.; Fatima, A. RRP7A links primary microcephaly to dysfunction of ribosome biogenesis, resorption of primary cilia, and neurogenesis. Nat. Commun. 2020, 11, 5816. [Google Scholar] [CrossRef] [PubMed]
  76. Chellas-Gery, B.; Wolf, K.; Tisoncik, J.; Hackstadt, T.; Fields, K. Biochemical and localization analyses of putative type III secretion translocator proteins CopB and CopB2 of Chlamydia trachomatis reveal significant distinctions. Infect. Immun. 2011, 79, 3036–3045. [Google Scholar] [CrossRef] [Green Version]
  77. Boonsawat, P.; Joset, P.; Steindl, K.; Oneda, B.; Gogoll, L.; Azzarello-Burri, S.; Sheth, F.; Datar, C.; Verma, I.C.; Puri, R.D. Elucidation of the phenotypic spectrum and genetic landscape in primary and secondary microcephaly. Genet. Med. 2019, 21, 2043–2058. [Google Scholar] [CrossRef] [Green Version]
  78. Jean, F.; Stuart, A.; Tarailo-Graovac, M. Dissecting the genetic and etiological causes of primary microcephaly. Front. Neurol. 2020, 11, 570830. [Google Scholar] [CrossRef]
  79. 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. 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] [Green Version]
  80. Zhong, X.; Pfeifer, G.P.; Xu, X. Microcephalin encodes a centrosomal protein. Cell Cycle 2006, 5, 457–458. [Google Scholar] [CrossRef]
  81. Zhou, Z.-W.; Tapias, A.; Bruhn, C.; Gruber, R.; Sukchev, M.; Wang, Z.-Q. DNA damage response in microcephaly development of MCPH1 mouse model. DNA Repair 2013, 12, 645–655. [Google Scholar] [CrossRef]
  82. Javed, A.O.; Li, Y.; Muffat, J.; Su, K.-C.; Cohen, M.A.; Lungjangwa, T.; Aubourg, P.; Cheeseman, I.M.; Jaenisch, R. Microcephaly modeling of kinetochore mutation reveals a brain-specific phenotype. Cell Rep. 2018, 25, 368–382.e365. [Google Scholar] [CrossRef] [Green Version]
  83. 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. ASPM is a major determinant of cerebral cortical size. Nat. Genet. 2002, 32, 316–320. [Google Scholar] [CrossRef]
  84. Gai, M.; Bianchi, F.T.; Vagnoni, C.; Vernì, F.; Bonaccorsi, S.; Pasquero, S.; Berto, G.E.; Sgrò, F.; Chiotto, A.M.; Annaratone, L. ASPM and CITK regulate spindle orientation by affecting the dynamics of astral microtubules. EMBO Rep. 2016, 17, 1396–1409. [Google Scholar] [CrossRef] [Green Version]
  85. 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] [PubMed]
  86. 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]
  87. Hussain, M.S.; Baig, S.M.; Neumann, S.; Nürnberg, G.; Farooq, M.; Ahmad, I.; Alef, T.; Hennies, H.C.; Technau, M.; Altmüller, J. A truncating mutation of CEP135 causes primary microcephaly and disturbed centrosomal function. Am. J. Hum. Genet. 2012, 90, 871–878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. 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]
  89. Tovini, L.; McClelland, S.E. Impaired CENP-E function renders large chromosomes more vulnerable to congression failure. Biomolecules 2019, 9, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Khan, M.A.; Rupp, V.M.; Orpinell, M.; Hussain, M.S.; Altmüller, J.; Steinmetz, M.O.; Enzinger, C.; Thiele, H.; Höhne, W.; Nürnberg, G. A missense mutation in the PISA domain of HsSAS-6 causes autosomal recessive primary microcephaly in a large consanguineous Pakistani family. Hum. Mol. Genet. 2014, 23, 5940–5949. [Google Scholar] [CrossRef]
  91. Alakbarzade, V.; Hameed, A.; Quek, D.Q.; Chioza, B.A.; Baple, E.L.; Cazenave-Gassiot, A.; Nguyen, L.N.; Wenk, M.R.; Ahmad, A.Q.; Sreekantan-Nair, A. A partially inactivating mutation in the sodium-dependent lysophosphatidylcholine transporter MFSD2A causes a non-lethal microcephaly syndrome. Nat. Genet. 2015, 47, 814. [Google Scholar] [CrossRef]
  92. Yamamoto, S.; Jaiswal, M.; Charng, W.-L.; Gambin, T.; Karaca, E.; Mirzaa, G.; Wiszniewski, W.; Sandoval, H.; Haelterman, N.A.; Xiong, B. A drosophila genetic resource of mutants to study mechanisms underlying human genetic diseases. Cell 2014, 159, 200–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Link, N.; Chung, H.; Jolly, A.; Withers, M.; Tepe, B.; Arenkiel, B.R.; Shah, P.S.; Krogan, N.J.; Aydin, H.; Geckinli, B.B. 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.e716. [Google Scholar] [CrossRef] [PubMed]
  94. 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. Biallelic mutations in citron kinase link mitotic cytokinesis to human primary microcephaly. Am. J. Hum. Genet. 2016, 99, 501–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Ranade, D.; Koul, S.; Thompson, J.; Prasad, K.B.; Sengupta, K. Chromosomal aneuploidies induced upon Lamin B2 depletion are mislocalized in the interphase nucleus. Chromosoma 2017, 126, 223–244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Sebastian, W.A.; Shiraishi, H.; Shimizu, N.; Umeda, R.; Lai, S.; Ikeuchi, M.; Morisaki, I.; Yano, S.; Yoshimura, A.; Hanada, R. Ankle2 deficiency-associated microcephaly and spermatogenesis defects in zebrafish are alleviated by heterozygous deletion of vrk1. Biochem. Biophys. Res. Commun. 2022, 624, 95–101. [Google Scholar] [CrossRef]
  97. Dragich, J.M.; Kuwajima, T.; Hirose-Ikeda, M.; Yoon, M.S.; Eenjes, E.; Bosco, J.R.; Fox, L.M.; Lystad, A.H.; Oo, T.F.; Yarygina, O. Autophagy linked FYVE (Alfy/WDFY3) is required for establishing neuronal connectivity in the mammalian brain. Elife 2016, 5, e14810. [Google Scholar] [CrossRef]
  98. Shi, L.; Qalieh, A.; Lam, M.M.; Keil, J.M.; Kwan, K.Y. Robust elimination of genome-damaged cells safeguards against brain somatic aneuploidy following Knl1 deletion. Nat. Commun. 2019, 10, 2588. [Google Scholar] [CrossRef] [Green Version]
  99. McIntyre, R.E.; Lakshminarasimhan Chavali, P.; Ismail, O.; Carragher, D.M.; Sanchez-Andrade, G.; Forment, J.V.; Fu, B.; Del Castillo Velasco-Herrera, M.; Edwards, A.; Van der Weyden, L. Disruption of mouse Cenpj, a regulator of centriole biogenesis, phenocopies Seckel syndrome. PLoS Genet. 2012, 8, e1003022. [Google Scholar] [CrossRef] [Green Version]
  100. Fujikura, K.; Setsu, T.; Tanigaki, K.; Abe, T.; Kiyonari, H.; Terashima, T.; Sakisaka, T. Kif14 mutation causes severe brain malformation and hypomyelination. PLoS ONE 2013, 8, e53490. [Google Scholar] [CrossRef] [Green Version]
  101. 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. 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] [Green Version]
  102. Scala, M.; Chua, G.L.; Chin, C.F.; Alsaif, H.S.; Borovikov, A.; Riazuddin, S.; Riazuddin, S.; Chiara Manzini, M.; Severino, M.; Kuk, A. Biallelic MFSD2A variants associated with congenital microcephaly, developmental delay, and recognizable neuroimaging features. Eur. J. Hum. Genet. 2020, 28, 1509–1519. [Google Scholar] [CrossRef] [PubMed]
  103. Rao, C.V.; Yamada, H.Y.; Yao, Y.; Dai, W. Enhanced genomic instabilities caused by deregulated microtubule dynamics and chromosome segregation: A perspective from genetic studies in mice. Carcinogenesis 2009, 30, 1469–1474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. González-Martínez, J.; Cwetsch, A.W.; Martínez-Alonso, D.; López-Sainz, L.R.; Almagro, J.; Melati, A.; Gómez, J.; Pérez-Martínez, M.; Megías, D.; Boskovic, J. Deficient adaptation to centrosome duplication defects in neural progenitors causes microcephaly and subcortical heterotopias. JCI Insight 2021, 6, e146364. [Google Scholar] [CrossRef]
  105. Szczepanski, S.; Hussain, M.S.; Sur, I.; Altmuller, J.; Thiele, H.; Abdullah, U.; Waseem, S.S.; Moawia, A.; Nurnberg, G.; Noegel, A.A.; et al. A novel homozygous splicing mutation of CASC5 causes primary microcephaly in a large Pakistani family. Hum. Genet. 2016, 135, 157–170. [Google Scholar] [CrossRef] [PubMed]
  106. 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] [PubMed] [Green Version]
  107. Li, R.; Sun, L.; Fang, A.; Li, P.; Wu, Q.; Wang, X. Recapitulating cortical development with organoid culture in vitro and modeling abnormal spindle-like (ASPM related primary) microcephaly disease. Protein Cell 2017, 8, 823–833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. 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] [Green Version]
  109. Fei, J.F.; Haffner, C.; Huttner, W.B. 3’UTR-dependent, miR-92-mediated restriction of Tis21 expression maintains asymmetric neural stem cell division to ensure proper neocortex size. Cell Rep. 2014, 7, 398–411. [Google Scholar] [CrossRef] [Green Version]
  110. Taverna, E.; Gotz, M.; Huttner, W.B. The cell biology of neurogenesis: Toward an understanding of the development and evolution of the neocortex. Annu. Rev. Cell Dev. Biol. 2014, 30, 465–502. [Google Scholar] [CrossRef]
  111. 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]
  112. 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]
  113. Kriegstein, A.R.; Noctor, S.C. Patterns of neuronal migration in the embryonic cortex. Trends Neurosci. 2004, 27, 392–399. [Google Scholar] [CrossRef] [PubMed]
  114. Basto, R.; Brunk, K.; Vinadogrova, T.; Peel, N.; Franz, A.; Khodjakov, A.; Raff, J.W. Centrosome amplification can initiate tumorigenesis in flies. Cell 2008, 133, 1032–1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Rahit, K.T.H.; Tarailo-Graovac, M. Genetic modifiers and rare mendelian disease. Genes 2020, 11, 239. [Google Scholar] [CrossRef] [Green Version]
  116. 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. Part A 2014, 164A, 2161–2171. [Google Scholar] [CrossRef] [PubMed]
  117. Duerinckx, S.; Jacquemin, V.; Drunat, S.; Vial, Y.; Passemard, S.; Perazzolo, C.; Massart, A.; Soblet, J.; Racapé, J.; Desmyter, L. Digenic inheritance of human primary microcephaly delineates centrosomal and non-centrosomal pathways. Hum. Mutat. 2020, 41, 512–524. [Google Scholar] [CrossRef] [Green Version]
  118. Snedeker, J.; Gibbons Jr, W.J.; Paulding, D.F.; Abdelhamed, Z.; Prows, D.R.; Stottmann, R.W. Gpr63 is a modifier of microcephaly in Ttc21b mouse mutants. PLoS Genet. 2019, 15, e1008467. [Google Scholar] [CrossRef] [Green Version]
  119. Barbosa-Morais, N.L.; Irimia, M.; Pan, Q.; Xiong, H.Y.; Gueroussov, S.; Lee, L.J.; Slobodeniuc, V.; Kutter, C.; Watt, S.; Çolak, R. The evolutionary landscape of alternative splicing in vertebrate species. Science 2012, 338, 1587–1593. [Google Scholar] [CrossRef] [Green Version]
  120. Bond, J.; Scott, S.; Hampshire, D.J.; Springell, K.; Corry, P.; Abramowicz, M.J.; Mochida, G.H.; Hennekam, R.C.; Maher, E.R.; Fryns, J.-P. Protein-truncating mutations in ASPM cause variable reduction in brain size. Am. J. Hum. Genet. 2003, 73, 1170–1177. [Google Scholar] [CrossRef] [Green Version]
  121. Ahmad, I.; Baig, S.; Abdulkareem, A.; Hussain, M.; Sur, I.; Toliat, M.; Nürnberg, G.; Dalibor, N.; Moawia, A.; Waseem, S. Genetic heterogeneity in Pakistani microcephaly families revisited. Clin. Genet. 2017, 92, 62–68. [Google Scholar] [CrossRef]
  122. Farooq, M.; Fatima, A.; Mang, Y.; Hansen, L.; Kjaer, K.W.; Baig, S.M.; Larsen, L.A.; Tommerup, N. A novel splice site mutation in CEP135 is associated with primary microcephaly in a Pakistani family. J. Hum. Genet. 2016, 61, 271. [Google Scholar] [CrossRef] [PubMed]
  123. Cristofoli, F.; Moss, T.; Moore, H.W.; Devriendt, K.; Flanagan-Steet, H.; May, M.; Jones, J.; Roelens, F.; Fons, C.; Fernandez, A. 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]
  124. Zhang, Y.; Li, H.; Pang, J.; Peng, Y.; Shu, L.; Wang, H. Novel SASS6 compound heterozygous mutations in a Chinese family with primary autosomal recessive microcephaly. Clin. Chim. Acta 2019, 491, 15–18. [Google Scholar] [CrossRef] [PubMed]
  125. Bettencourt-Dias, M.; Hildebrandt, F.; Pellman, D.; Woods, G.; Godinho, S.A. Centrosomes and cilia in human disease. Trends Genet. 2011, 27, 307–315. [Google Scholar] [CrossRef] [Green Version]
  126. Fujita, A.; Tsukaguchi, H.; Koshimizu, E.; Nakazato, H.; Itoh, K.; Kuraoka, S.; Komohara, Y.; Shiina, M.; Nakamura, S.; Kitajima, M. Homozygous splicing mutation in NUP133 causes Galloway–Mowat syndrome. Ann. Neurol. 2018, 84, 814–828. [Google Scholar] [CrossRef] [PubMed]
  127. 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] [PubMed]
  128. Arboleda, V.A.; Lee, H.; Dorrani, N.; Zadeh, N.; Willis, M.; Macmurdo, C.F.; Manning, M.A.; Kwan, A.; Hudgins, L.; Barthelemy, F. De novo nonsense mutations in KAT6A, a lysine acetyl-transferase gene, cause a syndrome including microcephaly and global developmental delay. Am. J. Hum. Genet. 2015, 96, 498–506. [Google Scholar] [CrossRef] [Green Version]
  129. Breuss, M.; Heng, J.I.-T.; Poirier, K.; Tian, G.; Jaglin, X.H.; Qu, Z.; Braun, A.; Gstrein, T.; Ngo, L.; Haas, M. Mutations in the β-tubulin gene TUBB5 cause microcephaly with structural brain abnormalities. Cell Rep. 2012, 2, 1554–1562. [Google Scholar] [CrossRef] [Green Version]
  130. Dinwiddie, D.L.; Soden, S.E.; Saunders, C.J.; Miller, N.A.; Farrow, E.G.; Smith, L.D.; Kingsmore, S.F. De novo frameshift mutation in ASXL3 in a patient with global developmental delay, microcephaly, and craniofacial anomalies. BMC Med. Genom. 2013, 6, 32. [Google Scholar] [CrossRef] [Green Version]
  131. Dhamija, R.; Graham Jr, J.M.; Smaoui, N.; Thorland, E.; Kirmani, S. Novel de novo SPOCK1 mutation in a proband with developmental delay, microcephaly and agenesis of corpus callosum. Eur. J. Med. Genet. 2014, 57, 181–184. [Google Scholar] [CrossRef]
  132. Mirzaa, G.M.; Enyedi, L.; Parsons, G.; Collins, S.; Medne, L.; Adams, C.; Ward, T.; Davitt, B.; Bicknese, A.; Zackai, E. Congenital microcephaly and chorioretinopathy due to de novo heterozygous KIF11 mutations: Five novel mutations and review of the literature. Am. J. Med. Genet. Part A 2014, 164, 2879–2886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Li, N.; Xu, Y.; Li, G.; Yu, T.; Yao, R.-E.; Wang, X.; Wang, J. Exome sequencing identifies a de novo mutation of CTNNB1 gene in a patient mainly presented with retinal detachment, lens and vitreous opacities, microcephaly, and developmental delay: Case report and literature review. Medicine 2017, 96, e6914. [Google Scholar] [CrossRef] [PubMed]
  134. Jacob, A.; Pasquier, J.; Carapito, R.; Auradé, F.; Molitor, A.; Froguel, P.; Fakhro, K.; Halabi, N.; Viot, G.; Bahram, S. A de novo synonymous variant in EFTUD2 disrupts normal splicing and causes mandibulofacial dysostosis with microcephaly: Case report. BMC Med. Genet. 2020, 21, 182. [Google Scholar] [CrossRef] [PubMed]
Table 1. MCPH cases reported with short stature.
Table 1. MCPH cases reported with short stature.
GeneVariantNo. of
Patients
Additional PhenotypesDiagnosis/Common FeaturesStudy
ASPMc.688delG/c.3340delA 2SSMCPH/SH, SGP, ID, BP[26]
c.688delG/c.9190C > T1SS
c.3979C > T1SS
c.9541C > T1SS
Not identified-1SS, syndactyly, strabismusMCPH/SH, ID [27]
ASPMc.6854_6855del5SSMCPH/ SH, ID, DD, speech delay[28]
ASPMc.3742-1G > C2SSMCPH/SGP, ID, epilepsy, speech delay[29]
ASPMc.9286C > T3SS, hearing impairmentMCPH, ID[30]
c.3055C > T8SS
c.9319C > T3SS
Not identified-2SSMCPH[30]
2SS, joint deformity
3SS, cataract
3SS, strabismus, ataxia
2SS
ASPMc.3945_3946delAG/c.8191_8194delGAAA1SSMCPH, ID[31]
PHC1c.2974C > T2SSPM/SH[33]
STILc.2150G > A2SS, sleep defectsMCPH, ID[32]
STILc.4849C > T2SS, HoloprosencephalyMCPH, ID[27]
MCPH1150–200 kb deletion of first 6 exons6SSMCPH, ID[34]
MCPH161 kb homozygous deletion of first 8 exon1SS, strabismus, ventriculomegalyCM, ID, SGP, speech delay[35]
WDR62c.668T > C2SS, Cortical malformations: hemispherical asymmetry, diffuse pachygyria, thick gray matter, indistinct gray-white matter junction, corpus callosum and white matter hypoplasiaCM, ID, DD[36]
WDR62c.1027C > T2SS, delayed motor skillsMCPH, ID, DD, speech delay[37]
WDR62c.1598A > G2SS, structural abnormalities and cortical malformation of the brain, motor impairment, increased deep tendon reflexes, flexible joint contractures, sensorineural hearing loss, vertical nystagmus, focal seizuresMCPH, ID, DD, speech delay[38]
WDR62c.2195C > T; p.(Thr732Ile) 5SS, seizuresMCPH, ID, speech delay[43]
CDK5RAP2c. 4441C > T1SS at birth (only), motor delay Simian crease, large map-like hyperpigmentation, tic disorderMCPH, SPG, ID, DD, speech delay[41]
CDK5RAP2c.524_528del/c.4005-1G > A1SS, delayed bone age, asthma, sleep apneaMCPH, ID, DD, speech defects[42]
CDK5RAP2c.448C > T; p.(Arg150*)3SSMCPH, DD, ID, speech delay[43]
KNL1c.6125G > A4SSMCPH, ID, DD, speech delay[40]
KNL1c.6125G > A4SS, cerebellar vermis hypoplasiaMCPH, ID, DD, speech delay[39]
Note: PM: primary microcephaly, SH: small head, SS: short stature, SPG: simplified gyral pattern, ID: intellectual disability, BP: behavioral problems, and DD: developmental delay.
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

Asif, M.; Abdullah, U.; Nürnberg, P.; Tinschert, S.; Hussain, M.S. Congenital Microcephaly: A Debate on Diagnostic Challenges and Etiological Paradigm of the Shift from Isolated/Non-Syndromic to Syndromic Microcephaly. Cells 2023, 12, 642. https://doi.org/10.3390/cells12040642

AMA Style

Asif M, Abdullah U, Nürnberg P, Tinschert S, Hussain MS. Congenital Microcephaly: A Debate on Diagnostic Challenges and Etiological Paradigm of the Shift from Isolated/Non-Syndromic to Syndromic Microcephaly. Cells. 2023; 12(4):642. https://doi.org/10.3390/cells12040642

Chicago/Turabian Style

Asif, Maria, Uzma Abdullah, Peter Nürnberg, Sigrid Tinschert, and Muhammad Sajid Hussain. 2023. "Congenital Microcephaly: A Debate on Diagnostic Challenges and Etiological Paradigm of the Shift from Isolated/Non-Syndromic to Syndromic Microcephaly" Cells 12, no. 4: 642. https://doi.org/10.3390/cells12040642

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