Next Article in Journal
Comprehensive Identification of Mitochondrial Pseudogenes (NUMTs) in the Human Telomere-to-Telomere Reference Genome
Previous Article in Journal
FOS Inhibits the Differentiation of Intramuscular Adipocytes in Goats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Whole-Exome Sequencing of a Cohort of 19 Families with Adolescent Idiopathic Scoliosis (AIS): Candidate Pathways

1
Brain Institute of Paris, 43-87 bd de l’Hôpital, 75013 Paris, France
2
Clinique Maussins Nollet, Ramsay Génerale de Santé, 67 Rue de Romainville, 75019 Paris, France
*
Author to whom correspondence should be addressed.
Submission received: 17 October 2023 / Revised: 9 November 2023 / Accepted: 15 November 2023 / Published: 17 November 2023

Abstract

:
A significant genetic involvement has been known for decades to exist in adolescent idiopathic scoliosis (AIS), a spine deformity affecting 1–3% of the world population. However, though biomechanical and endocrinological theories have emerged, no clear pathophysiological explanation has been found. Data from the whole-exome sequencing performed on 113 individuals in 19 multi-generational families with AIS have been filtered and analyzed via interaction pathways and functional category analysis (Varaft, Bingo and Panther). The subsequent list of 2566 variants has been compared to the variants already described in the literature, with an 18% match rate. The familial analysis in two families reveals mutations in the BICD2 gene, supporting the involvement of the muscular system in AIS etiology. The cellular component analysis revealed significant enrichment in myosin-related and neuronal activity-related categories. All together, these results reinforce the suspected role of the neuronal and muscular systems, highlighting the calmodulin pathway and suggesting a role of DNA-binding activities in AIS physiopathology.

1. Introduction

Adolescent idiopathic scoliosis (AIS) is the most common spinal deformity, affecting 1–3% of the worldwide population [1]. It is defined as a three-dimensional spinal deformity with curves exceeding 10° of magnitude in the coronal plane [2]. The medical management of AIS consists of screening, bracing and surgical correction for curves progressing over 40° despite bracing [3,4]. AIS has a long-term impact on patients’ quality of life and may lead to significant disability [5]. Moreover, up to 1–10% of AIS patients will require corrective surgery with a cost of approximatively USD 50,000, resulting in a heavy financial burden on society [6]. For decades, researchers have tried to understand the etiopathology of scoliosis in order to find treatment options that are not only symptomatic. The role of polygenic risk factors has been supported by the work of Riseborough et al. in 1973, finding a 12% recurrence risk of AIS for first-degree relatives [7]. The two main hypotheses that emerged to explain AIS were the biomechanical/musculoskeletal theory and the hormonal theory (related to estrogen and melatonin) [8]. The biomechanical/musculoskeletal theory hypotheses that the cartilage, bone and/or muscle are subject to disequilibrium during growth, resulting in the three-dimensional deformity of the spine. FBN1 (fibrillin-1, a component of the extracellular matrix), MATN1 (matrilin-1, a cartilage protein) and LBX1 (ladybird homeaobox-1, a regulator of muscle precursor cell migration) have been identified as the most promising candidate genes related to the musculoskeletal theory [9,10,11]. The hormonal theory states that a dysfunction in hormones regulating bone formation (such as estrogen), promoting osteoblast proliferation (such as melatonin) or regulating bone formation (such as leptin) may also induce disequilibrium in spine growth, resulting in AIS. Variants in CALM1 and ESR1 have been found to be associated with AIS in previous studies [10,12]. The recent progress in the field of genetics has allowed family-based studies and GWAS (Genome-Wide Association Study) in scoliosis cohorts. New variants have been identified, and a third hypothesis related to the nervous system has emerged, identifying new candidate genes such as CELSR2, TTL1 or CFAP298 [12,13,14,15]. Those works suggest that impairment in processes involving cerebellum growth, axonal migration, or cilium function result in AIS. However, these variants have failed to replicate in other cohorts and none of them seem to explain AIS as a monogenic disease [7,16,17,18]. Indeed, recent reports suggest that the pathophysiology of AIS seems to combine several phenomena involving the muscular and neural systems, regulated by complex hormonal and epigenetic phenomena. Subsequently, studies have tried to identify the pathways and functional categories that may be useful for explaining this highly complex polygenic disease [7,19,20,21]. These studies revealed enrichment in extra-cellular matrices’ genes and actin-based tubular projections and provide a promising lead for the future understanding of AIS.
The main goal of this study was to find new candidate pathways and functional categories involved in scoliosis by analyzing a large cohort of AIS families.

2. Materials and Methods

Patients: As a prospective data collection cohort, nineteen French families with a suppositively dominant inheritance of AIS were enrolled. The inclusion criteria consisted of at least two cases of idiopathic scoliosis in two successive generations. For any affected individual, the participation of at least one sibling without AIS as a healthy relative and both parents were sought ought. All patients underwent an orthopedic examination to confirm the idiopathic nature of the scoliosis (no neurological or syndromic pathology was associated, no dysraphism, normal abdominal reflexes, normal motor function, normal reflexes, absence of vertebral malformation on the radiographic analysis, concordant age of onset, or concordant rate of progression) [22,23]. The radiographic diagnostic of scoliosis relied on the presence of a curve greater than 10° with vertebral rotation [1]. For all subjects, age, gender and medical history was recorded. Additionally, for all AIS patients, magnitude of the main curve and contra-curve (in °) and gibbosity (in cm) were measured, and clinical data associated with AIS, such as age of onset and treatment history (bracing, surgery) were recorded. The pedigree charts of all families were performed on CeGat Software (Version 2.17). This study received ethical committee approval number 2020.02.05 bis_20.01.23.63418 (Comité de protection des personnes Sud-Méditerranée III). All procedures involving human participants were performed in accordance with the 1964 Declaration of Helsinki and its later amendments.

Methods

AIS patients and genetically relevant healthy relatives were sampled for blood or saliva, and anonymization was obtained by the numerical coding of the identity of the patients. DNA was obtained from 63 affected individuals with AIS and 50 non-affected relatives. DNA was extracted to a final quantity of 100 ng (35 µL at 3 ng·µL−1), with quality controls made by gel electrophoresis. Whole Exome Sequencing (WES) was performed on each sample, using the Nextera Rapid Capture MedExome v1.2 kit with paired-end 150-bp sequencing (Illumina NextSeq 500, San Diego, CA, USA), and realigned to the hg38 reference genome using the Burrows-Wheeler Aligner (BWA). The variants were annotated using VARAFT version 2.17 (Université Aix-Marseille, Marseille, France). The genetic variants of the 63 affected subjects were filtered by VARAFT to keep only variants of potential significance. The filtering criterions were stringent: exonic mutation (non-synonymous SNV, frameshift deletion or insertion, stop-gain), CADD score > 20 and frequency < 0.001. The results of the familial analysis of the cohort (affected patient versus healthy relatives, by family) are still ongoing and will be presented in a following report; functional validation of variants are still in progress, except for two interesting variants in one gene found in two families.
  • Literature comparison: A systematic analysis of the literature was performed on PubMed with the MeSH terms “idiopathic scoliosis” and “variant” to collect a list of genes significantly associated with AIS in clinical cohorts with confirmed genetic testing. Matches between this list and the list of SCOGEN variants were searched using ExcelStat.
  • Interaction pathways analysis: The list of variants obtained was then submitted to GO-enrichment analysis using BiNGO (Biological Networks Gene Ontology tool) Cytoscape plugin. The data were analyzed using an overrepresentation binomial test, and p-values were corrected via Benjamini-Hochberg false discovery rate (FDR). Significance level was set at p < 0.05.
  • Functional categories analysis: In order to identify the functional categories to which the variants present in our patients with AIS belonged, an analysis with Panther Version 18.0, (available at http://pantherdb.org/, accessed on 1 May 2023) was conducted. To retain only the most meaningful variants, the SCOGEN list was reduced to those present in 3 or more individuals. This subsequent list of 490 genes’ variants was then compared using a statistical overrepresentation test (binomial testing) to the GO cellular component list with the reference list on homo sapiens’ genome.

3. Results

3.1. Patients

The cohort comprised 19 families, consisting of 63 AIS patients and 50 healthy relatives, amounting to 113 individuals in total. The families’ pedigrees are displayed in Figure 1.
All 63 patients presented idiopathic scoliosis, with a mean Cobb angle for the cohort of 28° ± 8. Among the 63 patients, 21 (34%) underwent corrective surgery and 30 (48%) underwent bracing. Forty-nine patients with AIS were female (78%), whereas fourteen (22%) were male. All families were constituted of French citizens.

3.2. Variants

The initial analysis of the dataset of the 63 affected individuals resulted in a list of 2566 variants (Supplementary Materials) that passed the filtering criterion of the VARAFT analysis. In total, 756 (67%) of the variant’s genes were unique, 160 (14%) occurred twice and 216 (19%) occurred three times or more. Most variants were missense (Figure 2).
  • Comparison to the literature
The literature search of the terms “idiopathic scoliosis” and “variant” on PubMed retrieved 131 studies, of which 25 were excluded, being off-topic. Out of the 106 remaining studies, a list of 114 genes of interest was extracted. The comparison of the SCOGEN datalist to the literature list found 19 matches (18%) between the two datasets. The results are summarized in Table 1.
  • BINGO analysis
The 409 genes selected in the SCOGEN datalist were analyzed using BINGO to decipher their molecular function. Twenty-three clusters with a corrected p-value < 1 × 10−3 were identified. These results are presented in Table 2. The first nine clusters in terms of p-value were linked to binding activities, notably to nucleosides, nucleotides, proteins and calmodulin.
A visual representation of these results is also displayed in Figure 3.
  • Panther analysis
The goal of this analysis was to reveal the most significant cellular component location of the genes of interest, represented at least three times in the SCOGEN list, according to the GO annotation system. The most significant results (according to the corrected p-value (False Discovery Rate)) are presented in Table 3.
Myosin-related categories were represented in three of the top ten categories sorted by p-value, such as the myosin II complex (15.7-fold enrichment), myosin complex (8.6-fold enrichment) and muscle myosin complex (19.6-fold enrichment). The neuronal network was the second-most represented system with neuron-projection (2.01-fold enrichment), dendrite (2.35-fold enrichment), dendritic tree (6.4-fold enrichment), and somatodendritic compartment (8.79-fold enrichment).

3.3. Familial Analysis—BICD2 Gene

The familial analysis, looking for relevant variants present in at least 75% of the affected individuals in one family and absent in the healthy relatives, highlighted the presence of a variant in BICD2 (BICD cargo adaptor 2) in the family SCOGEN017 (Figure 4). This variant was present in four out of the five scoliosis patients and absent in the healthy relative. This mutation is a non-synonymous SNV (single nucleotide variation), occurring on chromosome 9, position 92722745, on the BICD2 gene, exon3: c.517C>G. This mutation is located in the CC1 zone (coiled-coil 1) of the protein in the dynein-dynactin complex. The predictive scores of this variant are 32 for the CADD score and 0.999 for the DANN score. This mutation is predicted to damage the protein: Provean predictive score: −6.51 (Protein Variation Effect Analyser, available at: https://www.jcvi.org/research/provean, accessed on 1 May 2023). Additionally, another variant of interest located on the BICD2 gene was found in another family: SCOGEN018. This variant, also a non-synonymous SNV, c.1069C>T (rs202119238), located in the CC2 zone (kinesin interaction), was present in 2/2 of the affected individuals (females) but also on one healthy relative (male) (Figure 4). The associated predictive scores were CADD 30, DANN: 0.999 and Provean −4.6.
Those variants were considered to be relevant to AIS, due to both their distribution in the families, especially in SCOGEN017, but also to the physiopathology of BICD2. The absence of phenotype in the male subject in SCOGEN018 family that presented the same variant as his affected mother and sister might be explained by the Carter effect, as described in AIS by Kruse et al. [24]. BICD2 is a motor-adaptor protein implicated in the dynein-dynactin motor complex within the muscular system which also interacts with microtubules [25]. In human pathology, mutations of BICD2 have been found to be responsible for SMA (Spinal Muscular Amyotrophy) in several studies [26,27,28]. The phenotype varies from the pauci-symptomatic form to severe ones in neonatal mortality [26,29]. The patients in SCOGEN017 and SCOGEN018 did not clinically present any sign of muscle- or second motoneuron impairment (no fasciculation, amyotrophia or motor deficit). They did not consent to an additional electromyogram to the study protocol, unfortunately. We may hypothesize that those BICD2 mutations might be responsible for the AIS phenotype either via sub-clinical forms of muscular impairment or via their link to the microtubules, as suggested by Baschal et al. [19].

4. Discussion

This study provides a comprehensive description of the rare genetic variants found in 409 genes in a large cohort of familial AIS. The analysis of their cellular function highlights an important number of genes involved in binding activity, suggesting a highly connected network leading to the pathophysiology of AIS. The analysis of cellular component categories reveals a hyper-representation of several myosin- and neuronal-related categories.
The 18% concordance with genes already reported in the literature is encouraging given the heterogeneity of the cohorts in terms of geographical origin (mostly North America and China) and selection criteria (familial studies and GWAS, with or without functional validation). However, several of the most significant genes with variants found in AIS, such as MATN1 and LBX1, were not observed in this cohort [30]. This rate of replication is an echo of the replication studies conducted in Chinese and Canadian cohorts of the ScoliScore designed in an American population, for which only 0–2 of the 53 SNP have significantly replicated in other cohorts [31,32,33]. This genetic scoring system had shown encouraging early results in predicting curve progression on a geographically limited cohort, but results failed to replicate broadly. The results of this cohort are in favor of a heterogenicity of genetic causation in AIS, which may vary between populations. The multiplication of similar studies in different geographic populations of AIS patients might therefore be helpful for understanding AIS etiology broadly.
Amongst the molecular function categories organized in clusters in the datalist retrieved, some referred to functions that may seem ubiquitous. Indeed, eight out of ten of the most represented categories represent nucleoside or nucleotide binding activities. However, the possible involvement of DNA-binding genes has been scarcely described in the literature, except for CHD7 (a DNA binding activity), a gene associated with AIS [34]. One other interesting finding of this study is the representation of “calmodulin-binding” as a molecular function cluster, with a frequency of 2.5% and a p-value of 1.5 × 10–4. Calmodulin is a calcium-binding receptor protein, involved in muscular contraction and interactions with melatonin—a suspected factor in AIS physiopathology. Calmodulin has been studied in AIS patients and via animal models. Acaroglu et al. have shown that the injection of tamoxifen (a calmodulin antagonist) may influence scoliosis curve regression in pinealectomized chickens, a scoliosis model [35]. The more noticeable clinical studies on AIS patients and calmodulin were conducted by Lowe et al. [36], who discovered that platelet calmodulin levels had then been sampled in AIS patients and correlate closely with curve progression, and by Marcucio et al. and Zhao et al. [37,38], who both found different levels of calmodulin in the paraspinal muscles (convex and concave sides) of AIS patients. CALM-1 (calmodulin 1) variants have been found to be associated with AIS in several cohorts [10,39,40]. In addition, the melatonin-calmodulin interactions involve cytoskeleton, filamentous actin and microtubule assembly, which have been shown by Miller et al. to be key factors in AIS physiopathology [19,41,42].
Panther analysis of the SCOGEN datalist reveals a significant enrichment in several GO cellular components such as myosin-related categories (myosin II complex, myosin complex and muscle myosin complex), plasma membrane, neuronal network (neuron projection, dendrite and dendritic tree) and stereocilium categories. The enrichment in the stereocilium category confirms the findings of Terhune et al., who reported significant cytoskeletal variant enrichment in a comparable cohort [7]. The enrichment in myosin found in the SCOGEN analysis, combined with the results of Baschal et al. in a similar work highlighting actin-based projection enrichment, suggests a clear involvement of the muscular complex, with possible variations amongst series that may be explained by geographical differences between studied populations [19]. Those results altogether may suggest that the binding activities impaired in AIS are predominantly located at the neuro-muscular junction and/or in muscles. The finding of two variants on a candidate gene BICD2 in the familial analysis of this cohort also re-enforces that hypothesis. The BICD2 protein is involved in dynein and dynactin interactions in active motor complexes but also plays a role in neural tissue development. Additionally, a novel mutation in BICD2 has recently been described as causative for a Cohen-like syndrome, with neurological impairment and amyotrophy [43]. BICD2 mutations have been shown to increase microtubules’ stability in motor neurons leading to axonal aberrations [44]. Interestingly, several studies have highlighted the role of microtubules, and especially stereocilium and motile cilium in AIS etiopathology [19,45]. Could AIS be the result of the impairment of several connections at the neuro-muscular junction? These hypotheses are still yet to be confirmed and the answer might result in a more complex model involving genetic variants as described in this study, combined with epigenetic factors [46,47].

5. Conclusions

This analysis of a large cohort of idiopathic scoliosis families confirmed the involvement of myosin, calmodulin and genes related to the neural system in AIS etiology, all sophisticatedly intricated with multiple interactions via protein bindings. Two candidate variants for AIS in the BICD2 gene were uncovered, re-enforcing the role of the muscular system and the microtubules in AIS physiopathology.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/genes14112094/s1.

Author Contributions

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

Funding

The research leading to these results had received funding from the program “Investissements d’avenir” ANR-10-IAIHU-06, from the Cotrel foundation and the SFCR (Société française de chirurgie rachidienne) and support from Ramsay Génerale de Santé.

Institutional Review Board Statement

This study received ethical committee approval number 2020.02.05 bis_20.01.23.63418 (Comité de protection des personnes Sud-Méditerranée III). All procedures involving human participants were performed in accordance with the ethical standards of these institutional review boards, the 1964 Declaration of Helsinki and its later amendments, or comparable ethical standards.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Data are available as Supplementary Materials.

Acknowledgments

We firstly thank all the families that have agreed to participate in this study. We particularly thank Myriam Petot, clinical research assistant, for her valuable help to establish patients’ pedigrees. We thank Raphaël Vialle (Trousseau Hospital, Paris) for his support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Horne, J.P.; Flannery, R.; Usman, S. Adolescent Idiopathic Scoliosis: Diagnosis and Management. Am. Fam. Physician 2014, 89, 193–198. [Google Scholar] [PubMed]
  2. Weinstein, S.L.; Dolan, L.A.; Cheng, J.C.; Danielsson, A.; Morcuende, J.A. Adolescent Idiopathic Scoliosis. Lancet 2008, 371, 1527–1537. [Google Scholar] [CrossRef] [PubMed]
  3. Altaf, F.; Gibson, A.; Dannawi, Z.; Noordeen, H. Adolescent Idiopathic Scoliosis. BMJ 2013, 346, f2508. [Google Scholar] [CrossRef]
  4. Miller, N.H. Cause and Natural History of Adolescent Idiopathic Scoliosis. Orthop. Clin. N. Am. 1999, 30, 353–365. [Google Scholar] [CrossRef] [PubMed]
  5. Ragborg, L.C.; Dragsted, C.; Ohrt-Nissen, S.; Andersen, T.; Gehrchen, M.; Dahl, B. Health-Related Quality of Life in Patients 40 Years after Diagnosis of an Idiopathic Scoliosis. Bone Jt. J. 2023, 105, 166–171. [Google Scholar] [CrossRef]
  6. Theologis, A.A.; Wu, H.-H.; Diab, M. Thoracic Posterior Spinal Instrumented Fusion vs. Thoracic Anterior Spinal Tethering for Adolescent Idiopathic Scoliosis with a Minimum of 2-Year Follow-up: A Cost Comparison of Index and Revision Operations. Spine Deform. 2022, 11, 359–366. [Google Scholar] [CrossRef] [PubMed]
  7. Terhune, E.A.; Wethey, C.I.; Cuevas, M.T.; Monley, A.M.; Baschal, E.E.; Bland, M.R.; Baschal, R.; Trahan, G.D.; Taylor, M.R.G.; Jones, K.L.; et al. Whole Exome Sequencing of 23 Multigeneration Idiopathic Scoliosis Families Reveals Enrichments in Cytoskeletal Variants, Suggests Highly Polygenic Disease. Genes 2021, 12, 922. [Google Scholar] [CrossRef]
  8. Zaydman, A.M.; Strokova, E.L.; Pahomova, N.Y.; Gusev, A.F.; Mikhaylovskiy, M.V.; Shevchenko, A.I.; Zaidman, M.N.; Shilo, A.R.; Subbotin, V.M. Etiopathogenesis of Adolescent Idiopathic Scoliosis: Review of the Literature and New Epigenetic Hypothesis on Altered Neural Crest Cells Migration in Early Embryogenesis as the Key Event. Med. Hypotheses 2021, 151, 110585. [Google Scholar] [CrossRef]
  9. Sheng, F.; Xia, C.; Xu, L.; Qin, X.; Tang, N.L.-S.; Qiu, Y.; Cheng, J.C.-Y.; Zhu, Z. New Evidence Supporting the Role of FBN1 in the Development of Adolescent Idiopathic Scoliosis. Spine 2019, 44, E225–E232. [Google Scholar] [CrossRef]
  10. De Salvatore, S.; Ruzzini, L.; Longo, U.G.; Marino, M.; Greco, A.; Piergentili, I.; Costici, P.F.; Denaro, V. Exploring the Association between Specific Genes and the Onset of Idiopathic Scoliosis: A Systematic Review. BMC Med. Genom. 2022, 15, 115. [Google Scholar] [CrossRef]
  11. Luo, M.; Zhang, Y.; Huang, S.; Song, Y. The Susceptibility and Potential Functions of the LBX1 Gene in Adolescent Idiopathic Scoliosis. Front. Genet. 2020, 11, 614984. [Google Scholar] [CrossRef]
  12. Zhuang, Q.; Wu, Z.; Qiu, G. Is Polymorphism of CALM1 Gene or Growth Hormone Receptor Gene Associated with Susceptibility to Adolescent Idiopathic Scoliosis? Zhonghua Yi Xue Za Zhi 2007, 87, 2198–2202. [Google Scholar] [PubMed]
  13. Einarsdottir, E.; Grauers, A.; Wang, J.; Jiao, H.; Escher, S.A.; Danielsson, A.; Simony, A.; Andersen, M.; Christensen, S.B.; Åkesson, K.; et al. CELSR2 Is a Candidate Susceptibility Gene in Idiopathic Scoliosis. PLoS ONE 2017, 12, e0189591. [Google Scholar] [CrossRef] [PubMed]
  14. Mathieu, H.; Patten, S.A.; Aragon-Martin, J.A.; Ocaka, L.; Simpson, M.; Child, A.; Moldovan, F. Genetic Variant of TTLL11 Gene and Subsequent Ciliary Defects Are Associated with Idiopathic Scoliosis in a 5-Generation UK Family. Sci. Rep. 2021, 11, 11026. [Google Scholar] [CrossRef] [PubMed]
  15. Marie-Hardy, L.; Cantaut-Belarif, Y.; Pietton, R.; Slimani, L.; Pascal-Moussellard, H. The Orthopedic Characterization of Cfap298tm304 Mutants Validate Zebrafish to Faithfully Model Human AIS. Sci. Rep. 2021, 11, 7392. [Google Scholar] [CrossRef] [PubMed]
  16. Xia, C.; Xu, L.; Xue, B.; Sheng, F.; Qiu, Y.; Zhu, Z. Rare Variant of HSPG2 Is Not Involved in the Development of Adolescent Idiopathic Scoliosis: Evidence from a Large-Scale Replication Study. BMC Musculoskelet. Disord. 2019, 20, 24. [Google Scholar] [CrossRef] [PubMed]
  17. Takahashi, Y.; Matsumoto, M.; Karasugi, T.; Watanabe, K.; Chiba, K.; Kawakami, N.; Tsuji, T.; Uno, K.; Suzuki, T.; Ito, M.; et al. Replication Study of the Association between Adolescent Idiopathic Scoliosis and Two Estrogen Receptor Genes. J. Orthop. Res. 2011, 29, 834–837. [Google Scholar] [CrossRef] [PubMed]
  18. Miller, N.H. Idiopathic Scoliosis: Cracking the Genetic Code and What Does It Mean? J. Pediatr. Orthop. 2011, 31, S49–S52. [Google Scholar] [CrossRef] [PubMed]
  19. Baschal, E.E.; Terhune, E.A.; Wethey, C.I.; Baschal, R.M.; Robinson, K.D.; Cuevas, M.T.; Pradhan, S.; Sutphin, B.S.; Taylor, M.R.G.; Gowan, K.; et al. Idiopathic Scoliosis Families Highlight Actin-Based and Microtubule-Based Cellular Projections and Extracellular Matrix in Disease Etiology. G3 Genes Genomes Genet. 2018, 8, 2663–2672. [Google Scholar] [CrossRef]
  20. Otomo, N.; Lu, H.-F.; Koido, M.; Kou, I.; Takeda, K.; Momozawa, Y.; Kubo, M.; Kamatani, Y.; Ogura, Y.; Takahashi, Y.; et al. Polygenic Risk Score of Adolescent Idiopathic Scoliosis for Potential Clinical Use. J. Bone Miner. Res. 2021, 36, 1481–1491. [Google Scholar] [CrossRef]
  21. Haller, G.; Alvarado, D.; Mccall, K.; Yang, P.; Cruchaga, C.; Harms, M.; Goate, A.; Willing, M.; Morcuende, J.A.; Baschal, E.; et al. A Polygenic Burden of Rare Variants across Extracellular Matrix Genes among Individuals with Adolescent Idiopathic Scoliosis. Hum. Mol. Genet. 2016, 25, 202–209. [Google Scholar] [CrossRef] [PubMed]
  22. Diab, M. Physical Examination in Adolescent Idiopathic Scoliosis. Neurosurg. Clin. N. Am. 2007, 18, 229–236. [Google Scholar] [CrossRef]
  23. Burton, M.S. Diagnosis and Treatment of Adolescent Idiopathic Scoliosis. Pediatr. Ann. 2013, 42, 224–228. [Google Scholar] [CrossRef] [PubMed]
  24. Kruse, L.M.; Buchan, J.G.; Gurnett, C.A.; Dobbs, M.B. Polygenic Threshold Model with Sex Dimorphism in Adolescent Idiopathic Scoliosis: The Carter Effect. J. Bone Jt. Surg. 2012, 94, 1485–1491. [Google Scholar] [CrossRef] [PubMed]
  25. Feng, Q.; Gicking, A.M.; Hancock, W.O. High-Resolution Tracking of Dynein-Dynactin-BicD2 Complexes. Methods Mol. Biol. 2023, 2623, 177–186. [Google Scholar] [CrossRef] [PubMed]
  26. Koboldt, D.C.; Waldrop, M.A.; Wilson, R.K.; Flanigan, K.M. The Genotypic and Phenotypic Spectrum of BICD2 Variants in Spinal Muscular Atrophy. Ann. Neurol. 2020, 87, 487–496. [Google Scholar] [CrossRef] [PubMed]
  27. Martinez-Carrera, L.A.; Wirth, B. Dominant Spinal Muscular Atrophy Is Caused by Mutations in BICD2, an Important Golgin Protein. Front. Neurosci. 2015, 9, 401. [Google Scholar] [CrossRef] [PubMed]
  28. Neveling, K.; Martinez-Carrera, L.A.; Hölker, I.; Heister, A.; Verrips, A.; Hosseini-Barkooie, S.M.; Gilissen, C.; Vermeer, S.; Pennings, M.; Meijer, R.; et al. Mutations in BICD2, Which Encodes a Golgin and Important Motor Adaptor, Cause Congenital Autosomal-Dominant Spinal Muscular Atrophy. Am. J. Hum. Genet. 2013, 92, 946–954. [Google Scholar] [CrossRef]
  29. Storbeck, M.; Horsberg Eriksen, B.; Unger, A.; Hölker, I.; Aukrust, I.; Martínez-Carrera, L.A.; Linke, W.A.; Ferbert, A.; Heller, R.; Vorgerd, M.; et al. Phenotypic Extremes of BICD2-Opathies: From Lethal, Congenital Muscular Atrophy with Arthrogryposis to Asymptomatic with Subclinical Features. Eur. J. Hum. Genet. 2017, 25, 1040–1048. [Google Scholar] [CrossRef]
  30. AlMekkawi, A.K.; Caruso, J.P.; El Ahmadieh, T.Y.; Palmisciano, P.; Aljardali, M.W.; Derian, A.G.; Al Tamimi, M.; Bagley, C.A.; Aoun, S.G. Single Nucleotide Polymorphisms and Adolescent Idiopathic Scoliosis: A Systematic Review and Meta-Analysis of the Literature. Spine 2023, 48, 695–701. [Google Scholar] [CrossRef]
  31. Ward, K.; Ogilvie, J.W.; Singleton, M.V.; Chettier, R.; Engler, G.; Nelson, L.M. Validation of DNA-Based Prognostic Testing to Predict Spinal Curve Progression in Adolescent Idiopathic Scoliosis. Spine 2010, 35, E1455–E1464. [Google Scholar] [CrossRef] [PubMed]
  32. Xu, L.; Qin, X.; Sun, W.; Qiao, J.; Qiu, Y.; Zhu, Z. Replication of Association Between 53 Single-Nucleotide Polymorphisms in a DNA-Based Diagnostic Test and AIS Progression in Chinese Han Population. Spine 2016, 41, 306–310. [Google Scholar] [CrossRef] [PubMed]
  33. Tang, Q.L.; Julien, C.; Eveleigh, R.; Bourque, G.; Franco, A.; Labelle, H.; Grimard, G.; Parent, S.; Ouellet, J.; Mac-Thiong, J.-M.; et al. A Replication Study for Association of 53 Single Nucleotide Polymorphisms in ScoliScore Test with Adolescent Idiopathic Scoliosis in French-Canadian Population. Spine 2015, 40, 537–543. [Google Scholar] [CrossRef] [PubMed]
  34. Wu, Z.; Dai, Z.; Yuwen, W.; Liu, Z.; Qiu, Y.; Cheng, J.C.-Y.; Zhu, Z.; Xu, L. Genetic Variants of CHD7 Are Associated with Adolescent Idiopathic Scoliosis. Spine 2021, 46, E618–E624. [Google Scholar] [CrossRef] [PubMed]
  35. Akel, I.; Kocak, O.; Bozkurt, G.; Alanay, A.; Marcucio, R.; Acaroglu, E. The Effect of Calmodulin Antagonists on Experimental Scoliosis: A Pinealectomized Chicken Model. Spine 2009, 34, 533–538. [Google Scholar] [CrossRef]
  36. Lowe, T.; Lawellin, D.; Smith, D.; Price, C.; Haher, T.; Merola, A.; O’Brien, M. Platelet Calmodulin Levels in Adolescent Idiopathic Scoliosis: Do the Levels Correlate with Curve Progression and Severity? Spine 2002, 27, 768–775. [Google Scholar] [CrossRef] [PubMed]
  37. Acaroglu, E.; Akel, I.; Alanay, A.; Yazici, M.; Marcucio, R. Comparison of the Melatonin and Calmodulin in Paravertebral Muscle and Platelets of Patients with or without Adolescent Idiopathic Scoliosis. Spine 2009, 34, E659–E663. [Google Scholar] [CrossRef]
  38. Zhao, Y.; Qiu, G. Expression of Calmodulin and NNOS in the Paraspinal Muscles in Idiopathic Scoliosis. Zhonghua Yi Xue Za Zhi 2004, 84, 1358–1361. [Google Scholar]
  39. Zhao, D.; Qiu, G.; Wang, Y.; Zhang, J.; Shen, J.; Wu, Z.; Wang, H. Association of Calmodulin1 Gene Polymorphisms with Susceptibility to Adolescent Idiopathic Scoliosis. Orthop. Surg. 2009, 1, 58–65. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Gu, Z.; Qiu, G. The Association Study of Calmodulin 1 Gene Polymorphisms with Susceptibility to Adolescent Idiopathic Scoliosis. Biomed. Res. Int. 2014, 2014, 168106. [Google Scholar] [CrossRef]
  41. Huerto-Delgadillo, L.; Antón-Tay, F.; Benítez-King, G. Effects of Melatonin on Microtubule Assembly Depend on Hormone Concentration: Role of Melatonin as a Calmodulin Antagonist. J. Pineal Res. 1994, 17, 55–62. [Google Scholar] [CrossRef] [PubMed]
  42. Benítez-King, G.; Antón-Tay, F. Calmodulin Mediates Melatonin Cytoskeletal Effects. Experientia 1993, 49, 635–641. [Google Scholar] [CrossRef]
  43. Caglayan, A.O.; Tuysuz, B.; Gül, E.; Alkaya, D.U.; Yalcinkaya, C.; Gleeson, J.G.; Bilguvar, K.; Gunel, M. Biallelic BICD2 Variant Is a Novel Candidate for Cohen-like Syndrome. J. Hum. Genet. 2022, 67, 553–556. [Google Scholar] [CrossRef] [PubMed]
  44. Carrera, L.A.M.; Gabriel, E.; Donohoe, C.D.; Hölker, I.; Mariappan, A.; Storbeck, M.; Uhlirova, M.; Gopalakrishnan, J.; Wirth, B. Novel Insights into SMALED2: BICD2 Mutations Increase Microtubule Stability and Cause Defects in Axonal and NMJ Development. Hum. Mol. Genet. 2018, 27, 1772–1784. [Google Scholar] [CrossRef] [PubMed]
  45. Grimes, D.T.; Boswell, C.W.; Morante, N.F.C.; Henkelman, R.M.; Burdine, R.D.; Ciruna, B. Zebrafish Models of Idiopathic Scoliosis Link Cerebrospinal Fluid Flow Defects to Spine Curvature. Science 2016, 352, 1341–1344. [Google Scholar] [CrossRef] [PubMed]
  46. Faldini, C.; Manzetti, M.; Neri, S.; Barile, F.; Viroli, G.; Geraci, G.; Ursini, F.; Ruffilli, A. Epigenetic and Genetic Factors Related to Curve Progression in Adolescent Idiopathic Scoliosis: A Systematic Scoping Review of the Current Literature. Int. J. Mol. Sci. 2022, 23, 5914. [Google Scholar] [CrossRef] [PubMed]
  47. Marya, S.; Tambe, A.D.; Millner, P.A.; Tsirikos, A.I. Adolescent Idiopathic Scoliosis: A Review of Aetiological Theories of a Multifactorial Disease. Bone Jt. J. 2022, 104, 915–921. [Google Scholar] [CrossRef]
Figure 1. Pedigrees of the 19 families of the SCOGEN cohort. The pedigrees were created on CeGat Software. Patients with confirmed AIS are represented in black and healthy relatives in white.
Figure 1. Pedigrees of the 19 families of the SCOGEN cohort. The pedigrees were created on CeGat Software. Patients with confirmed AIS are represented in black and healthy relatives in white.
Genes 14 02094 g001
Figure 2. Pie-chart representing the repartition of category types of the variants found in the SCOGEN list. SNV: Single nucleotide variation. Most of the variants (58%) were missense variants and frameshift insertions (18%).
Figure 2. Pie-chart representing the repartition of category types of the variants found in the SCOGEN list. SNV: Single nucleotide variation. Most of the variants (58%) were missense variants and frameshift insertions (18%).
Genes 14 02094 g002
Figure 3. BiNGO visualization of molecular function GO terms in the SCOGEN list of genes. The most significant GO terms are represented in orange (see legend bar with gradient of color according to significance). The size of the node is proportional to the number of genes identified in the SCOGEN list related to this GO term. For easier visualization, the nodes have been pooled into four categories: binding (which is predominantly represented), receptor and kinase, hydrolase and phosphatase, and channel and transporter.
Figure 3. BiNGO visualization of molecular function GO terms in the SCOGEN list of genes. The most significant GO terms are represented in orange (see legend bar with gradient of color according to significance). The size of the node is proportional to the number of genes identified in the SCOGEN list related to this GO term. For easier visualization, the nodes have been pooled into four categories: binding (which is predominantly represented), receptor and kinase, hydrolase and phosphatase, and channel and transporter.
Genes 14 02094 g003
Figure 4. Pedigrees of the SCOGEN017 and SCOGEN018 families regarding mutation status for the two variants in BICD2. In the SCOGEN017 family, patient I.1 had a 35° curve, treated by brace. Patients II.2 and II.4 underwent surgical correction of the scoliosis. Patient II.1 and II.4 were treated by bracing with curves of 20 and 30°, respectively. In the SCOGEN018 family, patients I.1 and II.2 were treated by bracing for thoracic curves of 25 and 35°. All main curves in both families were right thoracic curves, with an age of onset of the scoliosis between 11–15 years old.
Figure 4. Pedigrees of the SCOGEN017 and SCOGEN018 families regarding mutation status for the two variants in BICD2. In the SCOGEN017 family, patient I.1 had a 35° curve, treated by brace. Patients II.2 and II.4 underwent surgical correction of the scoliosis. Patient II.1 and II.4 were treated by bracing with curves of 20 and 30°, respectively. In the SCOGEN018 family, patients I.1 and II.2 were treated by bracing for thoracic curves of 25 and 35°. All main curves in both families were right thoracic curves, with an age of onset of the scoliosis between 11–15 years old.
Genes 14 02094 g004
Table 1. List of the 114 genes of interests associated with AIS in the literature, sorted into two categories, depending on their occurrence in the SCOGEN list of genes.
Table 1. List of the 114 genes of interests associated with AIS in the literature, sorted into two categories, depending on their occurrence in the SCOGEN list of genes.
Found in SCOGEN datalistABCC1- AJAP1- CHD7- COL11A1- DLG1- DST- EPHA7- EPHB6- LRRK2- MACF1- MAGI1- MCM3- MTNR1A- NTF3- PAX3- PCDHA11- SEMA4C- TNIK- VANGL1
Not found in SCOGEN datalistABO- ADAMTSL2- ADGRG6- ADIPOQ- AKAP2- AKAP9- AQP2- ATP12A- ATRN- BLC2- BMP4- BNC2- BOC- C2CD3- CALM1- CCKBR- CDH13- CELSR2- CEP290- CHI3L1- COL11A2- COMP- CPEB1- CSF3R- DNAAF1- DNAH6- DNAH8- DNHD1- EGR1- EPHA4- EPS8- ESR1- ESR2- FAT3- FBN1- FBN2- FGFR1- FLRT2- FNLB- GJB4- GNGT2- GPR126- GPR50- GRID2- HGF- HHIP- IGF1- IL17RC- IL6- KCNJ2- KIF15- KIF6- LBX1- LEP- LEPR- LRP2- LTBP4- MAPK7- MATN1- MEIS1- MGA- MMP3- MTNR1B- NEDD4- NPHP4- NPY4R- NUCKS1- PAX1- PCDHA4- PITX1- POC5- PRMT5- PTK7- PTPRB- ROBO3- SCNN1A- SCO- SEC16B- SELPLG- SLC22A14- SLC26A7- SLC39A8- SNTG1- SOCS3- SOX6- SOX9- STAB1- SULT1C2- TBX6- TGFB1- TNFRSF10C- TPH1- TTLL11- ZMYND10
Table 2. Corrected p-values (Benjamini-Hochberg false discovery rate) and cluster frequencies of the most represented categories of molecular function (overrepresentation binomial test) according to the GO categories. The analysis was created on BINGo.
Table 2. Corrected p-values (Benjamini-Hochberg false discovery rate) and cluster frequencies of the most represented categories of molecular function (overrepresentation binomial test) according to the GO categories. The analysis was created on BINGo.
GO IDGO DescriptionCorrected p-ValueCluster
Frequency
1882nucleoside binding6.6291 × 10−12179/970 18.4%
1883purine nucleoside binding6.6291 × 10−12178/970 18.3%
17076purine nucleotide binding6.6291 × 10−12203/970 20.9%
30554adenyl nucleotide binding9.4134 × 10−12174/970 17.9%
32553ribonucleotide binding1.7312 × 10−11194/970 20.0%
32555purine ribonucleotide binding1.7312 × 10−11194/970 20.0%
32559adenyl ribonucleotide binding3.0105 × 10−11165/970 17.0%
5524ATP binding1.1851 × 10−10161/970 16.5%
166nucleotide binding2.1163 × 10−9217/970 22.3%
3824catalytic activity9.1279 × 10−9413/970 42.5%
5488binding1.2959 × 10−7846/970 87.2%
5515protein binding1.3637 × 10−7600/970 61.8%
16887ATPase activity1.9803 × 10−544/970 4.5%
17111nucleoside-triphosphatase activity1.9803 × 10−579/970 8.1%
16462pyrophosphatase activity8.5471 × 10−579/970 8.1%
16818hydrolase activity, acting on acid anhydrides, in phosphorus-containing anhydrides9.4250 × 10−579/970 8.1%
16817hydrolase activity, acting on acid anhydrides1.0988 × 10−479/970 8.1%
5516calmodulin binding1.4874 × 10−425/970 2.5%
16772transferase activity, transferring phosphorus-containing groups1.4972 × 10−490/970 9.2%
16301kinase activity1.4972 × 10−479/970 8.1%
16787hydrolase activity2.7345 × 10−4190/970 19.5%
16773phosphotransferase activity, alcohol group as acceptor2.7345 × 10−473/970 7.5%
42623ATPase activity, coupled6.2897 × 10−435/970 3.6%
Table 3. Occurrences of the variants in the reference genome of Panther, and in the SCOGEN datalist of genes according to their GO cellular component categories. The number of expected occurrences and the corresponding fold-enrichments with the p-values are also presented.
Table 3. Occurrences of the variants in the reference genome of Panther, and in the SCOGEN datalist of genes according to their GO cellular component categories. The number of expected occurrences and the corresponding fold-enrichments with the p-values are also presented.
GO Cellular ComponentHomo sapiensSCOGENExpectedFold-EnrichmentCorrected p Value
plasma membrane region13073013.332.259.53 × 10−3
apical part of cell461154.703.191.07 × 10−2
myosin II complex2540.2515.691.08 × 10−2
apical plasma membrane396134.043.221.24 × 10−2
myosin complex5750.588.601.26 × 10−2
neuron projection13662813.932.011.31 × 10−2
muscle myosin complex1530.1519.611.51 × 10−2
specific granule15971.624.321.57 × 10−2
cluster of actin-based cell projections16671.694.131.68 × 10−2
cytosol55157656.251.352.03 × 10−2
myosin filament2430.2412.262.13 × 10−2
dendrite625156.372.352.17 × 10−2
plasma membrane59098060.271.332.24 × 10−2
dendritic tree627156.402.352.29 × 10−2
cell periphery63958565.231.302.31 × 10−2
intracellular anatomical structure14,910170152.081.122.48 × 10−2
stereocilium5640.577.002.59 × 10−2
vesicle tethering complex5840.596.762.86 × 10−2
somatodendritic compartment862188.792.053.13 × 10−2
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

Marie-Hardy, L.; Courtin, T.; Pascal-Moussellard, H.; Zakine, S.; Brice, A. The Whole-Exome Sequencing of a Cohort of 19 Families with Adolescent Idiopathic Scoliosis (AIS): Candidate Pathways. Genes 2023, 14, 2094. https://0-doi-org.brum.beds.ac.uk/10.3390/genes14112094

AMA Style

Marie-Hardy L, Courtin T, Pascal-Moussellard H, Zakine S, Brice A. The Whole-Exome Sequencing of a Cohort of 19 Families with Adolescent Idiopathic Scoliosis (AIS): Candidate Pathways. Genes. 2023; 14(11):2094. https://0-doi-org.brum.beds.ac.uk/10.3390/genes14112094

Chicago/Turabian Style

Marie-Hardy, Laura, Thomas Courtin, Hugues Pascal-Moussellard, Serge Zakine, and Alexis Brice. 2023. "The Whole-Exome Sequencing of a Cohort of 19 Families with Adolescent Idiopathic Scoliosis (AIS): Candidate Pathways" Genes 14, no. 11: 2094. https://0-doi-org.brum.beds.ac.uk/10.3390/genes14112094

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