Next Article in Journal
Dystonia: Still a Mysterious Syndrome
Next Article in Special Issue
MicroRNA: A Linking between Astrocyte Dysfunction, Mild Cognitive Impairment, and Neurodegenerative Diseases
Previous Article in Journal
Emerging SARS-CoV-2 Mutational Variants of Concern Should Not Vary in Susceptibility to Microbicidal Actives
Previous Article in Special Issue
Early Salivary miRNA Expression in Extreme Low Gestational Age Newborns
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NSD1 Mutations in Sotos Syndrome Induce Differential Expression of Long Noncoding RNAs, miR646 and Genes Controlling the G2/M Checkpoint

1
Laboratory of Human Genetics, IRCCS Istituto Giannina Gaslini, 16147 Genoa, Italy
2
Clinical Bioinformatics Unit, IRCCS Istituto Giannina Gaslini, 16147 Genoa, Italy
3
Lung Cancer Unit, IRCCS Ospedale Policlinico San Martino, 16132 Genoa, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Life 2022, 12(7), 988; https://doi.org/10.3390/life12070988
Submission received: 26 May 2022 / Revised: 28 June 2022 / Accepted: 1 July 2022 / Published: 2 July 2022

Abstract

:
An increasing amount of evidence indicates the critical role of the NSD1 gene in Sotos syndrome (SoS), a rare genetic disease, and in tumors. Molecular mechanisms affected by NSD1 mutations are largely uncharacterized. In order to assess the impact of NSD1 haploinsufficiency in the pathogenesis of SoS, we analyzed the gene expression profile of fibroblasts isolated from the skin samples of 15 SoS patients and of 5 healthy parents. We identified seven differentially expressed genes and five differentially expressed noncoding RNAs. The most upregulated mRNA was stratifin (SFN) (fold change, 3.9, Benjamini–Hochberg corrected p < 0.05), and the most downregulated mRNA was goosecoid homeobox (GSC) (fold change, 3.9, Benjamini–Hochberg corrected p < 0.05). The most upregulated lncRNA was lnc-C2orf84-1 (fold change, 4.28, Benjamini–Hochberg corrected p < 0.001), and the most downregulated lncRNA was Inc-C15orf57 (fold change, −0.7, Benjamini–Hochberg corrected p < 0.05). A gene set enrichment analysis reported the enrichment of genes involved in the KRAS and E2F signaling pathways, splicing regulation and cell cycle G2/M checkpoints. Our results suggest that NSD1 is involved in cell cycle regulation and that its mutation can induce the down-expression of genes involved in tumoral and neoplastic differentiation. The results contribute to defining the role of NSD1 in fibroblasts for the prevention, diagnosis and control of SoS.

1. Introduction

The nuclear receptor SET domain protein 1 (NSD1) histone methyltransferase has been identified as a protein that interacts with several nuclear receptors, and it may also act as a bifunctional transcriptional cofactor, playing a dual role in transcription [1,2]. The mono- and di-methylation of histone 3 lysine 36 (H3K36) and lysine 168 (H3K168) have been proposed to be the major cellular NSD1 substrates [3,4]. In addition, NSD1 (NM 022455.4) can act as a tumor suppressor gene [5,6]. The structure of the NSD1 protein is characterized by two nuclear interaction domains (NIDs); two proline–tryptophan–tryptophan–proline (PWWP) domains; five plant homeodomains (PHD); an atypical plant homeodomain (C5HCH) finger; and a catalytic domain (CD), which is composed of pre-SET, subset of SET and post-SET-containing proteins. A so-called ‘associated with SET’ (AWS) domain is sometimes found instead of the pre-SET domain, containing Su(var)3–9, enhancer-of-zeste, Trithorax (SET) and a post-SET domain [7].
Germline mutations in the NSD1 gene cause Sotos syndrome (SoS) (OMIM 117550), a rare genetic disease with a prevalence of 1/14.000 births. It is characterized by overgrowth, macrocephaly, advanced bone age, characteristic facial features (long face) and intellectual delay [8,9,10]. SoS is due to an NSD1 gene haploinsufficiency that can be caused by several different mechanisms (truncating mutations, missense mutations, splice-site mutations, partial gene deletions and 5q35 microdeletions) [11,12]. Intragenic mutations account for 80–85% of SoS cases among European and American populations; 5q35 microdeletions are present in 10–15% of European and American cases and in more than 50% of Japanese cases [11].
In SoS patients bearing germline NSD1 haploinsufficiency, a deregulation of the mitogen-activated protein kinase (MAPK)/ERK signaling pathway has been observed downstream of KRAS activation, and this has been postulated to contribute to accelerated skeletal outgrowth [13]. Even if different functions of NSD1 have been documented, the molecular mechanisms that induce the phenotypic characteristics of SoS remain largely unknown.
The Human Genome Project showed that approximately 76% of the human genome is transcribed, with about 70% of transcripts corresponding to noncoding RNAs [14,15]. Noncoding RNAs can be divided into two groups depending on their lengths. Small RNAs are defined as being shorter than 200 nucleotides (nt), which include microRNA, and RNAs longer than 200 nt are referred to as long noncoding RNAs (lncRNAs). lncRNAs and microRNAs have roles in different mechanisms of gene regulation [16,17,18,19,20,21,22,23].
Moreover, somatically acquired alterations are frequently detected in various human cancers [24,25,26,27,28,29,30,31,32,33,34], and RNAi-mediated knock-down seems to increase the proliferation of tumor cells, suggesting the role of NSD1 as a tumor suppressor [5]. Studies have also described NSD1 as a potential biomarker for drug resistance in tumors.
In this study, we generated genome-wide transcription profiles for fibroblasts taken from SoS patients and their sex-matched healthy parents in order to identify which lncRNAs, microRNAs, mRNAs and downstream pathways are perturbed by NSD1 mutations in SoS syndrome. Our results underline the involvement of NSD1 in the G2/M cell cycle checkpoint, which prevents cells from entering mitosis when DNA is damaged, stopping cell proliferation. We also observed that NSD1 mutations can induce dysregulation in the expression of noncoding RNA.

2. Materials and Methods

2.1. Patients

This study was conducted with the approval of the Ethical Committee of the Liguria Region (Approval #OG01IGG, 12 July 2021). Written informed consent was obtained from all participants. Fifteen individuals with the classic clinical features of SoS and with molecular diagnoses of NSD1 pathogenic variants (point mutation or deletion) were enrolled in this study. As controls, five parents with the same sexes as their co-respective SoS patients, selected on the basis of skin biopsy availability, were also enrolled. The details are presented in Tables S1 and S2. The fibroblast cell lines were deposited at the Gaslini Biobank.

2.2. Cell Culture

Skin fibroblasts were obtained via punch biopsy of the forearm skin. Cells were maintained in the RPMI-1640 medium (Life Technologies, Grand Island, NY, USA), supplemented with 10% fetal calf serum (FCS), 2 mM L-glutamine, 100 U/mL penicillin and 100 μg/mL streptomycin (Euroclone S.p.a, Milan, Italy). Cell lines were tested for mycoplasma with a mycoplasma detection kit (Lonza, Basel, CH, Switzerland). In all experiments, cells with between 2 and 15 passages were used.

2.3. Analysis of NSD1 Pathogenetic Variants

DNA was extracted from the fibroblasts using QIAamp® DNA Blood kit (Qiagen, Milan, Italy) according to the manufacturer’s protocol. PCR amplifications were performed using platinum-Taq DNA polymerase (Thermo Fisher Scientific, Carlsbad, CA, USA) and specific primers for the 23 different NSD1 exons (see Table S3). The PCR conditions were a single denaturation step at 95 °C for 3 min followed by 40 cycles of 95 °C/30 s, 58 °C/30 s and 72 °C (1 min/kb), with a final extension step at 72 °C for 5 min. PCR products were sequenced using the ABI BigDye Terminator Ready Reaction Mix (Thermo Fisher Scientific, Foster City, CA, USA) and analyzed on an ABI 3130XL Genetic Analyzer (Applied Biosystems, Foster City, CA, USA) according to the manufacturer’s instructions. The sequences were aligned with Seqscape analysis software V.2.5 (Thermo Fisher Scientific, Foster City, CA, USA). To verify variants with deletions of the 5q35.3 region, array-CGH was performed with CGH 8 × 60 K (Agilent Technologies, Santa Clara, CA, USA) according to the manufacturer’s protocol. The data were analyzed with the Agilent Cytogenomics 4.0.3.12 software (Agilent Technologies, Santa Clara, CA, USA). All genomic positions were reported according to the human genome assembly (GRCh37/hg19).

2.4. Gene Expression Profiling

The total RNA was extracted using Trizol reagent (Thermo Fisher Scientific, MA, USA). The total RNA from each sample was quantified with the NanoDrop ND-1000, and RNA integrity was assessed by an Agilent 2100 bioanalyzer. Microarray hybridization was performed using the Agilent One Color microarray gene expression kit and the SurePrint G3, 8 × 60 K Human Gene Expression V3 array (Agilent Technologies, Santa Clara, CA, USA) following established protocols. Briefly, cyanine-3-CTP-labeled cRNA was hybridized onto SurePrint G3 microarray chips (Design ID: G4851C), which contained 50,599 probes for 32,776 human mRNAs and 17,438 human lncRNAs; these were derived from authoritative databases, including RefSeq, Ensemble, GenBank and the Broad Institute. After hybridization and washing, processed slides were scanned with a Agilent microarray scanner (Agilent Technologies). Raw data were extracted using Feature Extraction (version 12.0.1.1; Agilent Technologies). Next, data preprocessing, including normalization and filtering, was carried out with the Genespring software (version 14.3; Agilent Technologies). Raw data were normalized by a 75-percentile shift, log2-transformed and shifted to the median of all samples. The microarray data were deposited in GEO with the accession number GSE204775 (http://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE204775, accessed on 24 May 2022). Only samples in which quality control was excellent were used for subsequent analyses in order to reduce potential biases introduced by analyzing low-quality specimens.

2.5. Quantitative Real-Time RT-PCR Validation of Microarray Gene Expression Patterns

To confirm the significant modulation of the differentially expressed genes resulting from microarray data analysis, we performed real-time quantitative PCR using gene-specific primer (Table S3) and ran each experiment in triplicate for analysis robustness. Briefly, cDNA was synthesized from 400 ng of the total RNA using the Advantage RT cDNA Kit (Clontech, Mountain View, CA, USA) following the manufacturer’s instructions. Specifically, samples were incubated at 42 °C for 90 min, followed by 2 min at 90 °C. Quantitative real-time PCR was performed using LightCycler 480 SYBR Green I Master (Roche Diagnostics, Mannheim, Germany) in a 15 µL reaction mixture. β-actin was used as an endogenous reference to normalize gene expression values with the 2−ΔΔCt method [35]. PCR products were also confirmed via sequence analysis using the ABI BigDye Terminator Ready Reaction Mix (Life Technologies, Carlsbad, CA, USA) and analyzed on an ABI 3130XL genetic analyzer (Life Technologies, Carlsbad, CA, USA) according to the manufacturer’s protocol.

2.6. Bioinformatic Analysis

Using a bioinformatic analysis, based on the normalized fluorescence signal values of the lncRNA/mRNA probes, differentially expressed mRNAs and lncRNAs were identified by examining fold changes as well as by p-values calculated via Student’s t-tests between the SoS patients and the healthy controls using the “Statistical Analysis” module from the analysis section, provided by the Genespring software (version 14.3; Agilent Technologies) with default parameters. The PO value was adjusted with the Benjamini–Hochberg method to account for the increased probability of false positive findings derived from multiple comparisons. A fold change of ≥2.0 or ≤0.5 and a p-adjusted value of <0.05 were considered to be significantly up- or downregulated. In silico protein–protein functional interactions among the differentially expressed genes were assessed with the STRING database (http://stringdb.org, accessed on 24 May 2022) [36] using default parameters. An extension of the network was analyzed to assess potential indirect interactions between the differentially expressed genes. Using and Graphpad Prism 8 software, a Bonferroni-adjusted significance level of 0.05 was calculated to account for the increased possibility of obtaining false positive results.
A gene set enrichment analysis (GSEA) [37] was used to assess the enrichment of functionally related gene sets in SoS samples [37]. Chemical and genetic perturbations (C2.CGP), hallmark (H), gene ontology biological processes (C5.GO.BP) or gene set collections retrieved from the Molecular Signature Database (MSigDB) v7.4 [38] were used to perform enrichment analyses. GSEA calculated an enrichment score (ES) and a normalized enrichment score (NES) for each gene set and estimated the statistical significance of the NES with an empirical permutation test using 1000 gene permutations in order to obtain the nominal p-value (NOM p-value). When multiple gene sets are evaluated, GSEA adjusts the estimate of the significance level to account for multiple hypothesis testing. To this end, GSEA computed the false discovery rate q-value (FDR q-value), measuring the estimated probability that the normalized enrichment score represented a false positive finding. We considered gene sets containing between 15 and 250 genes. Gene sets with nominal p-values lower than 0.05 and FDR q-values lower than 0.05 were considered significantly enriched.

3. Results

3.1. Comparison of the Gene Expression Profile of Fibroblast Lines between SoS and Control Individuals

Fifteen patients with SoS and five healthy parents were enrolled in the study. The characteristics of the enrolled individuals are reported in Tables S1 and S2. A flowchart summarizing the main steps of these analyses is shown in Figure 1. RNA obtained from dermal fibroblasts from those SoS patients with confirmed NSD1 alterations and their sex-matched controls were analyzed using Agilent SurePrint G3 microarrays. Gene expression profiles were collected into one dataset for subsequent analysis.
Quality control was carried out on all samples to exclude potential artifacts derived from sample preparation and processing. One sample from the SoS patient group was filtered out because it did not pass the conformity test (Supplementary Figure S1). A differential expression analysis identified five significantly differentially expressed noncoding RNAs (four lncRNA and one microRNA) (Table 1) and seven probe sets relative to the mRNAs (Table 2). Four probe sets relative to the noncoding RNAs were upregulated (MIR646, lnc-C2orf84-1, lnc-C00665 and lnc-C20orf197-3) and one was downregulated (lnc-C15orf57, fold change, −0.7, p < 0.05). The most upregulated lncRNA was associated with lnc-C2orf84-1 (fold change, 4.28, p < 0.001; Table 1). Among the probe sets relative to mRNAs, four were upregulated (NOS3, CD19, SFN and ZNF883) and three were downregulated (NDRG2, GSC and SORBS1). The most upregulated probe set was associated with the stratifin (SFN) gene (fold change, 3.9, p < 0.05), and the most downregulated was associated with the goosecoid homeobox (GSC) gene (fold change, 3.9, p < 0.05; Table 1).
To assess the biological connection between the differentially expressed genes, we performed a network analysis using the STRING-DB software [36]. We extended the network to assess likely indirect interactions between differentially expressed genes. The resulting network, reported in Figure 2, showed that, statistically, the significantly modulated genes did not display any direct interaction with NSD1, but it did report an interaction with a subset of other gene products, including KRAS and pTEN, which are known to interact with NSD1. These findings indicate that NSD1 is an important regulator of different genes involved in cell differentiation and proliferation.

3.2. Validation of Microarray Results by Real-Time Quantitative PCR

All seven genes and six lncRNAs that were found to be significantly modulated in the microarray analysis were subjected to further validation using real-time PCR. Interestingly, the qPCR results confirmed all of the deregulations in the genes and lncRNAs assessed by microarray analysis (Figure 3a,b). In particular, a Student’s t-test analysis, by applying a 0.05 cutoff on the Benjamini–Hochberg-corrected p-values, showed that the expressions of MIR646 (p < 0.05), lnc-C2orf84-1 (p < 0.05), lnc-C00665 (p < 0.001), lnc-C20orf197-3 (p < 0.001), NOS3 (p < 0.05), ZNF883 (p < 0.001), CD19 (p < 0.05) and SFN (p < 0.05) were upregulated and that those of GSC (p < 0.05), lnc-C15orf57 (p < 0.001), NDRG2 (p < 0.001) and SORBS1 (p < 0.05) were downregulated in SoS patients.

3.3. Gene Set Enrichment Analysis

Gene set enrichment analysis (GSEA) is a computational technique that assesses the coordinated gene expression modulation of functionally related genes (gene sets) between two groups [37]. To perform this analysis, we used gene sets included in the chemical and genetic perturbation, hallmark (H) and C2 gene ontology collections from the MSigDB database [38]. The analysis showed the 89 significantly enriched biological processes and pathways in the NSD1-mutated fibroblasts of SoS patients, as compared to the healthy controls; Table S4 reports the complete list of significant processes and pathways. The most statistically significant over-represented gene sets were related to the cell cycle and proliferation, cell differentiation, P53-mediated cell cycle arrest, cellular senescence and cancer (Table 3). In particular, upregulated genes were mainly involved in the regulation of the meiotic cell cycle, the negative regulation of nuclear division (Supplementary Figure S2A,B) and kinetochore organization. The most statistically significant underrepresented gene sets were related to the E2F pathway (Supplementary Figure S2), pediatric cancer markers, neoplastic transformation via the STAT3 pathway, MYC and TFRC targets, the cell cycle G2/M checkpoint, TNFA-signaling via NFKB, epithelial mesenchymal transition, the apoptosis process caused by CDKN1A via TP53 and neoplastic transformation KRAS signaling pathways (Supplementary Figure S3A,B).
Then, an interaction network was generated using the STRING database. Nodes represent gene products, and edges represent protein–protein associations. Only the associations with an evidence score higher than 0.3 are shown; their colors indicate different kinds of interaction evidence (key, bottom right).

4. Discussion

SoS is an autosomal dominant disorder with manifestations characterized by tall stature, facial dysmorphism and mental retardation [1,2,3,4,5].
Currently, the pathogenetic cause of SoS is thought to be due to haploinsufficiency in the NSD1 gene [6]. However, the molecular mechanisms of action are still unknown. At present, an increasing number of studies have indicated that NSD1 is involved in the regulation of a wide variety of biological processes [27,28,29,30,31,32]. The NSD1 gene is expressed in most tissues from different organisms. Elevated NSD1 levels are detected in normal brains, pancreases and male reproductive tracts, in hematopoietic organs, bone marrow and lymphoid tissues [9]. Significant NSD1 mRNA expression in bone marrow polymorphonuclear cells, CD4, CD8 and NK has also been reported by genevisible.com [10].
NSD1 defects induce the alternate methylation of H3K36, blocking cellular differentiation and promoting oncogenesis [24]. Reduced NSD1 activity has been observed in head and neck cell cancers, while abnormal DNA-promoter hypermethylation is associated with renal clear-cell carcinoma [25,27,28,29,30]. Furthermore, NSD1 mutations are present in hematological malignancies, and a recurrent t(5;11)(q35;p15) chromosomal translocation is associated with aggressive pediatric acute myeloid leukemia (AML). The resulting fusion protein with leukemogenic activity is composed of the N-terminus of nucleopore 98 (NUP98) and the C-terminal NSD1 SET domain [32,33,34]. The abnormal expression of NSD1 in different cancers has also been detected, which is associated with tumorigenesis, survival and chemoresistance. In pancreatic ductal adenocarcinoma, a high level of NSD1 is significantly correlated with the advanced clinical stage [39]. Moreover, NSD1 gene-silencing inhibits the proliferative, migratory and invasive abilities of hepatocellular carcinoma cells [40]. Furthermore, upregulation of NSD1 might improve oncogenic initiation through the reinforced methylation of H3K36 [41].
In the present study, the expression of lncRNAs and mRNAs in SoS patients and the healthy control was investigated using a microarray analysis to reveal their potential role in the pathogenesis of SoS. We decided to compare the expression profiles of SoS patients with those of their healthy parents, as they share approximately 50% of their DNA; this allowed us to minimize differences caused by other genes variants with respect to gene expression [42,43].
In total, five noncoding RNAs and seven mRNAs were identified as being differentially expressed. GSEA analyses were used to explore the possible biological functions and potential mechanisms of the mRNAs and noncoding RNAs in SoS.
The results of the present study demonstrated that these biological processes, including G2/M checkpoints, can control transitions between cell cycle phases, cell senescence, and meiotic and mitotic division and are among the most significantly enriched mRNAs in SoS. Most of these functions are involved in neoplastic disease and tumor development, and this is consistent with previous studies [24,25,26,27,28,29,30,31,32,33,34,39,40,41]. Furthermore, we observed that the KRAS signaling pathway, the E2F target and apoptosis by CDKN1A via TP53-signaling were remarkably downregulated in the SoS samples compared to the healthy controls. On the other hand, genes involved in the regulation of nuclear division, meiotic cell cycle and kinetochore organization were remarkably upregulated in SoS patients.
lncRNAs have long been considered the only transcriptional noise; however, several lncRNAs analyzed with microarrays showed greater expression in immortalized lymphocytes [44]. lncRNAs were also found to be upregulated in neuroblastoma cell line LAN-I [45,46]. Different studies have shown that lncRNA alterations contribute to neuronal and neurodegenerative disease development [19,47,48]. lncRNAs may provide basic information that could be used to understand pathways related to the disease course of SoS and to find more effective targeted therapies.
The results of the present study revealed that, among the differentially expressed genes, SFN was the most upregulated mRNA (FC = 3.9) in SoS samples. The SFN protein is ubiquitously expressed and exerts many relevant intracellular functions, such as the control of cell cycle and apoptosis, the regulation of signal transduction pathways, cellular trafficking, cell proliferation and differentiation, and protein folding and processing, among others [49]. Both p53 and SFN prevent DNA errors during mitosis [40,41,42,43,44,45,46,47,48,49,50,51,52]. Therefore, according to the reported evidence, our study suggests that NSD1 may be an important protein for cell cycle and mitotic translation, regulating SFN G2/M checkpoint expression.
In addition, in our study, GSC, SORBS1 and NDRG2 mRNA were down-expressed in Sotos patients compared to the healthy controls. These three genes encode proteins that are associated with the actin cytoskeleton, the formation of actin stress fibers and focal adhesions, and receptor tyrosinase kinase-signaling, and they control different aspects of morphogenesis and cell differentiation [53,54,55]. Some studies have reported that the expression of GSC inhibits the proliferation of PC12 cells and that—as with NSD1—GSC mediates the regulation of erythropoiesis functions [56]. NDRG2 is frequently downregulated in cancer, and it plays an important role in the control of tumor growth and metastasis [57]. SORBS1 expression has been detected in colorectal cancer cell lines, and its overexpression increases the proliferation and migration abilities of tumor cells [58].
We found a downregulation of the SORBS1, NDRG2 and GSC genes in the SoS samples, similar to the RNA expression profile of the mesenchymal cell. The down-expression of these three genes induces the epithelial–mesenchymal transition [59,60].
Furthermore, GSC acts in mesenchyme-derived tissues during craniofacial development, and it has been reported that mice that are homozygous for the deletion of the homeobox gene goosecoid can have multiple craniofacial defects, and various bone and cartilage malformations [61,62]. Therefore, we propose that GSC down-expression may contribute to the advanced bone age and characteristic facial features of Sotos syndrome, including macrocephaly and long face.
Epilepsy is one of the clinical manifestations found in SoS patients. We found an over-expression of ZNF833, a gene that is correlated with the progression of epilepsy in rats [63].
In conclusion, we propose that NSD1 may participate in Sotos syndrome’s mechanism of action by regulating the function of lncRNA, inducing the down-expression of GSC, NDRG2 and SORBS1 and the up-expression of SFN and ZNF883.
These expression signatures may be useful tools for screening and monitoring the disease and for predicting its prognosis.
Additional investigations should be performed to establish and clarify the detailed molecular mechanism of action behind the lncRNA-mediated regulation of potential coding genes in SoS. Further studies with a larger sample size should also be performed to verify the relevant results.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/life12070988/s1, Figure S1: Quality control evaluation sample 19; Figure S2: Enrichment plots of representative over represented biological processes in SoS fibroblasts samples by GSEA; Figure S3: Enrichment plots of representative down represented of C2,CGP and hallmark gene sets in SoS fibroblasts samples by GSEA; Table S1: Sotos patients enrolled; Table S2: Healthy parents enrolled; Table S3: Primers used for validation of microarray results by real-time PCR; Table S4: GSEA analysis.

Author Contributions

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

Funding

This research was funded by the AssiGulliver Italian Association; Fondazione Sardegna; Italian Ministry of Health, Ricerca Corrente 2020 and Ricerca Corrente 2021.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

This study was approved by the Ethical Committee of the Liguria Region (Approval #OG01IGG on 12 July 2021), and written informed consent was provided by patients and/or their parents or legal guardians.

Data Availability Statement

The gene expression profiles of the patients and healthy controls are accessible in the Gene Expression Omnibus repository (GEO accession: GSE204775).

Acknowledgments

We thank the Biobank of the Laboratory of Human Genetics of the Giannina Gaslini Institute, Genoa (Italy) for the cell line provided and the patients for their participation in this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, N.; Vom Baur, J.M.; Garnier, T.; Lerouge, J.; Vonesch, L.; Lutz, Y.; Chambon, P.; Losson, R. Two distinct nuclear receptor interaction domains in NSD1, a novel SET protein that exhibits characteristics of both corepressors and coactivators. EMBO J. 1998, 17, 3398–3412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Wang, X.; Yeh, S.; Wu, G.; Hsu, C.L.; Wang, L.; Chiang, T.; Yang, Y.; Guo, Y.; Chang, C. Identification and characterization of a novel androgen receptor coregulator ARA267-alpha in prostate cancer cells. J. Biol. Chem. 2001, 276, 40417–40423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wagner, E.J.; Carpenter, P.B. Understanding the language of Lys36 methylation at histone H3. Nat. Rev. Mol. Cell Biol. 2012, 13, 115–126. [Google Scholar] [CrossRef] [Green Version]
  4. Kudithipudi, S.; Lungu, C.; Rathert, P.; Happel, N.; Jeltsch, A. Substrate specificity analysis and novel substrates of the protein lysine methyltransferase NSD1. Chem. Biol. 2014, 21, 226–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Berdasco, M.; Ropero, S.; Setien, F.; Fraga, M.; Lapunzina, P.; Losson, R.; Alaminos, M.; Cheung, N.-K.; Rahman, N.; Esteller, M. Epigenetic inactivation of the Sotos overgrowth syndrome gene histone methyltransferase NSD1 in human neuroblastoma and glioma. Proc. Natl. Acad. Sci. USA 2009, 106, 21830–21835. [Google Scholar] [CrossRef] [Green Version]
  6. Bui, N.; Huang, J.K.; Bojorquez-Gomez, A.; Licon, K.; Sanchez, K.S.; Tang, S.N.; Beckett, A.N.; Wang, T.; Zhang, W.; Shen, J.P.; et al. Disruption of NSD1 in Head and Neck Cancer Promotes Favorable Chemotherapeutic Responses Linked to Hypomethylation. Mol. Cancer Ther. 2018, 17, 1585–1594. [Google Scholar] [CrossRef] [Green Version]
  7. Amin, N.; Nietlispach, D.; Qamar, S.; Coyle, J.; Chiarparin, E.; Williams, G. NMR backbone resonance assignment and solution secondary structure determination of human NSD1 and NSD2. Biomol. NMR Assign. 2016, 10, 315–320. [Google Scholar] [CrossRef] [Green Version]
  8. Tatton-Brown, K.; Douglas, J.; Coleman, K.; Baujat, G.; Cole, T.R.; Das, S.; Horn, D.; Hughes, H.E.; Temple, I.K.; Faravelli, F.; et al. Genotype-phenotype associations in Sotos syndrome: An analysis of 266 individuals with NSD1 aberrations. Am. J. Hum. Genet. 2005, 77, 193–204. [Google Scholar] [CrossRef] [Green Version]
  9. The Human Protein Atlas. Available online: https://www.proteinatlas.org/ENSG00000165671-NSD1/tissue (accessed on 13 July 2021).
  10. GeneCards: The Human Gene Database: NSD1. Available online: https://www.genecards.org/cgibin/carddisp.pl?gene=NSD1#protein_expression (accessed on 13 July 2021).
  11. Tatton-Brown, K.; Rahman, N. Sotos syndrome. Eur. J.Hum. Genet. 2007, 15, 264–271. [Google Scholar] [CrossRef]
  12. Tatton-Brown, K.; Cole, T.R.P.; Rahman, N. Sotos syndrome. In GeneReviews® [Internet]; Pagon, R.A., Adam, M.P., Ardinger, H.H., Wallace, S.E., Bean, L.J.H., Mefford, H.C., Gripp, K.W., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 2014; pp. 1993–2017. [Google Scholar]
  13. Visser, R.; Landman, E.B.; Goeman, J.; Wit, J.M.; Karperien, M. Sotos syndrome is associated with deregulation of the MAPK/ERK-signaling pathway. PLoS ONE 2012, 7, e49229. [Google Scholar] [CrossRef]
  14. Lee, H.; Zhang, Z.; Krause, K.M. Long Noncoding RNAs and Repetitive Elements: Junk or Intimate Evolutionary Partners? Trends Genet. 2019, 35, 892–902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kashi, K.; Henderson, L.; Bonetti, A.; Carninci, P. Discovery and functional analysis of lncRNAs: Methodologies toinvestigate an uncharacterized transcriptome. Biochim. Biophys. Acta 2016, 1859, 3–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Hangauer, M.J.; Vaughn, I.W.; McManus, M.T. Pervasive transcription of the human genome produces thousands of previously unidentified long intergenic noncoding RNAs. PLoS. Genet. 2013, 6, e1003569. [Google Scholar] [CrossRef] [PubMed]
  17. Li, C.J.; Hong, T.; Tung, Y.T.; Yen, Y.P.; Hsu, H.C.; Lu, Y.L.; Chang, M.; Nie, Q.; Chen, J.A. MicroRNA filters Hox temporal transcription noise to confer boundary formation in the spinal cord. Nat. Commun. 2017, 8, 14685. [Google Scholar] [CrossRef] [Green Version]
  18. Hoye, M.L.; Koval, E.D.; Wegener, A.J.; Hyman, T.S.; Yang, C.; O’Brien, D.R.; Miller, R.L.; Cole, T.; Schoch, K.M.; Shen, T.; et al. MicroRNA Profiling Reveals Marker of Motor Neuron Disease in ALS Models. J. Neurosci. 2017, 37, 5574–5586. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, C.; Wang, L.; Ding, Y.; Lu, X.; Zhang, G.; Yang, J.; Zheng, H.; Wang, H.; Jiang, Y.; Xu, L. LncRNA Structural Characteristics in Epigenetic Regulation. Int. J. Mol. Sci. 2017, 18, 2659. [Google Scholar] [CrossRef] [Green Version]
  20. Engreitz, J.M.; Haines, J.E.; Perez, E.M.; Munson, G.; Chen, J.; Kane, M.; McDonel, P.E.; Guttman, M.; Lander, E.S. Local regulation of gene expression by lncRNA promoters, transcription and splicing. Nature 2016, 539, 452–455. [Google Scholar] [CrossRef]
  21. Vangoor, V.R.; Gomes-Duarte, A.; Pasterkamp, R.J. Long non-coding RNAs in motor neuron development and disease. J. Neurochem. 2021, 156, 777–801. [Google Scholar] [CrossRef]
  22. Anderson, D.M.; Anderson, K.M.; Chang, C.L.; Makarewich, C.A.; Nelson, B.R.; McAnally, J.R.; Kasaragod, P.; Shelton, J.M.; Liou, J.; Bassel-Duby, R.; et al. A micropeptide encoded by a putative long noncoding RNA regulates muscle performance. Cell 2015, 160, 595–606. [Google Scholar] [CrossRef] [Green Version]
  23. Liu, L.; Fang, F. Long Noncoding RNA Mediated Regulation in Human Embryogenesis, Pluripotency, and Reproduction. Stem Cells Int. 2022, 2022, 8051717. [Google Scholar] [CrossRef]
  24. Cancer Genome Atlas Network. Comprehensive genomic characterization of head and neck squamous cell carcinomas. Nature 2015, 517, 576–582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Bianco-Miotto, T.; Chiam, K.; Buchanan, G.; Jindal, S.; Day, T.K.; Thomas, M.; Pickering, M.A.; O’Loughlin, M.A.; Ryan, N.K.; Raymond, W.A.; et al. Global Levels of Specific Histone Modifications and an Epigenetic Gene Signature Predict Prostate Cancer Progression and Development. Cancer Epidemiol. Biomark. Prev. 2010, 19, 2611–2622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Spurr, L.; Li, M.; Alomran, N.; Zhang, Q.; Restrepo, P.; Movassagh, M.; Trenkov, C.; Tunnessen, N.; Apanasovich, T.; Crandall, K.A.; et al. Systematic pan-cancer analysis of somatic allele frequency. Sci. Rep. 2018, 8, 7735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Su, X.; Zhang, J.; Mouawad, R.; Compérat, E.; Rouprêt, M.; Allanic, F.; Malouf, G.G. NSD1 Inactivation and SETD2 Mutation Drive a Convergence toward Loss of Function of H3K36 Writers in Clear Cell Renal Cell Carcinomas. Cancer Res. 2017, 77, 4835–4845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Mohammad, F.; Helin, K. Oncohistones: Drivers of pediatric cancers. Genes Dev. 2017, 31, 2313–2324. [Google Scholar] [CrossRef] [Green Version]
  29. Schuhmacher, M.; Kusevic, D.; Kudithipudi, S.; Jeltsch, A. Kinetic Analysis of the Inhibition of the NSD1, NSD2 and SETD2 Protein Lysine Methyltransferases by a K36M Oncohistone Peptide. ChemistrySelect 2017, 2, 9532–9536. [Google Scholar] [CrossRef]
  30. Farhangdoost, N.; Horth, C.; Hu, B.; Bareke, E.; Chen, X.; Li, Y.; Coradin, M.; Garcia, B.A.; Lu, C.; Majewski, J. Chromatin dysregulation associated with NSD1 mutation in head and neck squamous cell carcinoma. Cell Rep. 2021, 34, 108769. [Google Scholar] [CrossRef]
  31. Rajagopalan, K.N.; Chen, X.; Weinberg, D.N.; Chen, H.; Majewski, J.; Allis, C.D.; Lu, C. Depletion of H3K36me2 recapitulates epigenomic and phenotypic changes induced by the H3.3K36M oncohistone mutation. Proc. Natl. Acad. Sci. USA 2021, 118, e2021795118. [Google Scholar] [CrossRef]
  32. Jaju, R.J. A novel gene, NSD1, is fused to NUP98 in the t(5;11)(q35;p15.5) in de novo childhood acute myeloid leukemia. Blood 2001, 98, 1264–1267. [Google Scholar] [CrossRef] [Green Version]
  33. Ntziachristos, P.; Abdel-Wahab, O.; Aifantis, I. Emerging concepts of epigenetic dysregulation in hematological malignancies. Nat. Immunol. 2016, 17, 1016–1024. [Google Scholar] [CrossRef] [Green Version]
  34. Tauchmann, S.; Almosailleakh, M.; Schwaller, J. NSD1 in erythroid differentiation and leukemogenesis. Mol. Cell Oncol. 2020, 7, 1809919. [Google Scholar] [CrossRef] [PubMed]
  35. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)). Method 2001, 25, 402–408. [Google Scholar] [CrossRef]
  36. Jensen, L.J.; Kuhn, M.; Stark, M.; Chaffron, S.; Creevey, C.; Muller, J.; Doerks, T.; Julien, P.; Roth, A.; Simonovic, M.; et al. STRING 8—A global view on proteins and their functional interactions in 630 organisms. Nucleic Acids Res. 2009, 37, D412–D416. [Google Scholar] [CrossRef] [PubMed]
  37. Subramanian, A.; Tamayo, P.; Mootha, V.K.; Mukherjee, S.; Ebert, B.L.; Gillette, M.A.; Paulovich, A.; Pomeroy, S.L.; Golub, T.R.; Lander, E.S.; et al. Gene set enrichment analysis: A knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl. Acad. Sci. USA 2005, 102, 15545–15550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Liberzon, A.; Subramanian, A.; Pinchback, R.; Thorvaldsdóttir, H.; Tamayo, P.; Mesirov, J.P. Molecular signatures database (MSigDB) 3.0. Bioinformatics 2011, 27, 1739–1740. [Google Scholar] [CrossRef] [PubMed]
  39. Ettel, M.; Zhao, L.; Schechter, S.; Shi, J. Expression and prognostic value of NSD1 and SETD2 in pancreatic ductal adenocarcinoma and its precursor lesions. Pathology 2019, 51, 392–398. [Google Scholar] [CrossRef] [PubMed]
  40. Zhang, S.; Zhang, F.; Chen, Q.; Wan, C.; Xiong, J.; Xu, J. CRISPR/Cas9- mediated knockout of NSD1 suppresses the hepatocellular carcinoma development via the NSD1/H3/ Wnt10b signaling pathway. J. Exp. Clin. Cancer Res. 2019, 38, 467. [Google Scholar] [CrossRef]
  41. Lucio-Eterovic, A.K.; Singh, M.M.; Gardner, J.E.; Veerappan, C.S.; Rice, J.C.; Carpenter, P.B. Role for the nuclear receptor-binding SET domain protein 1 (NSD1) methyltransferase in coordinating lysine 36 methylation at histone 3 with RNA polymerase II function. Proc. Natl. Acad. Sci. USA. 2010, 107, 16952–16957. [Google Scholar] [CrossRef] [Green Version]
  42. Miko, I. Gregor Mendel and the principles of inheritance. Nat. Educ. 2008, 1, 134. [Google Scholar]
  43. Chial, H. Mendelian genetics: Patterns of inheritance and single-gene disorders. Nat. Educ. 2008, 1, 63. [Google Scholar] [CrossRef] [Green Version]
  44. Ronchetti, D.; Manzoni, M.; Agnelli, L.; Vinci, C.; Fabris, S.; Cutrona, G.; Matis, S.; Colombo, M.; Galletti, S.; Taiana, E.; et al. lncRNA profiling in early-stage chronic lymphocytic leukemia identifies transcriptional fingerprints with relevance in clinical outcome. Blood Cancer J. 2016, 6, e468. [Google Scholar] [CrossRef] [PubMed]
  45. Hu, Y.; Sun, H.; Hu, J.; Zhang, X. LncRNA DLX6-AS1 Promotes the Progression of Neuroblastoma by Activating STAT2 via Targeting miR-506-3p. Cancer Manag. Res. 2020, 12, 7451–7463. [Google Scholar] [CrossRef] [PubMed]
  46. Utnes, P.; Løkke, C.; Flægstad, T.; Einvik, C. Clinically Relevant Biomarker Discovery in High-Risk Recurrent Neuroblastoma. Cancer Inform. 2019, 18, 1176935119832910. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Ang, C.E.; Trevino, A.E.; Chang, H.Y. Diverse lncRNA mechanisms in brain development and disease. Curr. Opin. Genet. Dev. 2020, 65, 42–46. [Google Scholar] [CrossRef] [PubMed]
  48. Deng, Q.; Ma, L.; Chen, T.; Yang, Y.; Ma, Y.; Ma, L. NF-κB 1-induced LINC00665 regulates inflammation and apoptosis of neurons caused by spinal cord injury by targeting miR-34a-5p. Neurol Res. 2021, 43, 418–427. [Google Scholar] [CrossRef]
  49. Khalil, A.M.; Guttman, M.; Huarte, M.; Garber, M.; Raj, A.; Rivea Morales, D.; Thomas, K.; Presser, A.; Bernstein, B.E.; Van Oudenaarden, A.; et al. Many human large intergenic noncoding RNAs associate with chromatin modifying complexes and affect gene expression. Proc. Natl. Acad. Sci. USA 2009, 106, 11667–11672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Mattick, J.S.; Gagen, M.J. The evolution of controlled multitasked gene networks: The role of introns and other noncoding RNAs in the development of complex organisms. Mol. Biol. Evol. 2001, 18, 1611–1630. [Google Scholar] [CrossRef] [Green Version]
  51. Popadin, K.; Gutierrez Arcelus, M.; Dermitzakis, E.T.; Antonarakis, S.E. Genetic and epigenetic regulation of human lincRNA gene expression. Am. J. Hum. Genet. 2013, 93, 1015–1026. [Google Scholar] [CrossRef] [Green Version]
  52. Luk, S.C.; Siu, S.W.; Lai, C.K.; Wu, Y.J.; Pang, S.F. Cell Cycle Arrest by a Natural Product via G2/M Checkpoint. Int. J. Med. Sci. 2005, 2, 64–69. [Google Scholar] [CrossRef] [Green Version]
  53. Sawada, K.; Konishi, Y.; Tominaga, M.; Watanabe, Y.; Hirano, J.; Inoue, S.; Kageyama, R.; Blum, M.; Tominaga, A. Goosecoid suppresses cell growth and enhances neuronal differentiation of PC12 cells. J. Cell Sci. 2000, 113, 2705–2713. [Google Scholar] [CrossRef]
  54. Foletta, V.C.; Prior, M.J.; Stupka, N.; Carey, K.; Segal, D.H.; Jones, S.; Swinton, C.; Martin, S.; Cameron-Smith, D.; Walder, K.R. NDRG2, a novel regulator of myoblast proliferation, is regulated by anabolic and catabolic factors. J. Physiol. 2009, 587 Pt 7, 1619–1634. [Google Scholar] [CrossRef]
  55. Cho, W.C.; Jang, J.E.; Kim, K.H.; Yoo, B.C.; Ku, J.L. SORBS1 serves a metastatic role via suppression of AHNAK in colorectal cancer cell lines. Int. J. Oncol. 2020, 56, 1140–1151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Konishi, Y.; Tominaga, M.; Watanabe, Y.; Imamura, F.; Goldfarb, A.; Maki, R.; Blum, M.; De Robertis, E.M.; Tominaga, A. GOOSECOID inhibits erythrocyte differentiation by competing with Rb for PU.1 binding in murine cells. Oncogene 1999, 18, 6795–6805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Golestan, A.; Ghaderi, A.; Mojtahedi, Z. Effects of NDRG2 Overexpression on Metastatic Behaviors of HCT116 Colorectal Cancer Cell Line. Adv. Pharm. Bull. 2017, 7, 661–664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Xu, Y.; Xin, R.; Sun, H.; Long, D.; Li, Z.; Liao, H.; Xue, T.; Zhang, Z.; Kang, Y.; Mao, G. Long Non-coding RNAs LOC100126784 and POM121L9P Derived from Bone Marrow Mesenchymal Stem Cells Enhance Osteogenic Differentiation via the miR-503-5p/SORBS1 Axis. Front. Cell. Dev. Biol. 2021, 9, 723759. [Google Scholar] [CrossRef] [PubMed]
  59. Moradi Monfared, M.; Alizadeh Zarei, M.; Rafiei Dehbidi, G.; Behzad Behbahani, A.; Arabsolghar, R.; Takhshid, M.A. NDRG2 Regulates the Expression of Genes Involved in Epithelial Mesenchymal Transition of Prostate Cancer Cells. Iran. J. Med. Sci. 2019, 44, 118–126. [Google Scholar]
  60. Kang, K.W.; Lee, M.J.; Song, J.A.; Jeong, J.Y.; Kim, Y.K.; Lee, C.; Kim, T.H.; Kwak, K.B.; Kim, O.J.; An, H.J. Overexpression of goosecoid homeobox is associated with chemoresistance and poor prognosis in ovarian carcinoma. Oncol. Rep. 2014, 32, 189–198. [Google Scholar] [CrossRef] [Green Version]
  61. Ulmer, B.; Tingler, M.; Kurz, S.; Maerker, M.; Andre, P.; Mönch, D.; Campione, M.; Deißler, K.; Lewandoski, M.; Thumberger, T.; et al. A novel role of the organizer gene Goosecoid as an inhibitor of Wnt/PCP-mediated convergent extension in Xenopus and mouse. Sci. Rep. 2017, 7, 43010. [Google Scholar] [CrossRef] [Green Version]
  62. Rivera-Perez, J.A.; Wakamiya, M.; Behringer, R.R. Goosecoid acts cell autonomously in mesenchyme-derived tissues during craniofacial development. Development 1999, 126, 3811–3821. [Google Scholar] [CrossRef]
  63. Gong, L.; Yang, P.; Hu, L.; Zhang, C. MiR-181b suppresses the progression of epilepsy by regulation of lncRNA ZNF883. Am. J. Transl. Res. 2020, 12, 2769–2780. [Google Scholar]
Figure 1. Schematic representation of the whole study strategy used. A GeneSpring differential expression analysis assessed the significant modulation of target genes between the SoS patient group and the healthy control. Differentially expressed genes were validated by real-time PCR. A GSEA pathway analysis identified the most significantly altered biological processes and pathways.
Figure 1. Schematic representation of the whole study strategy used. A GeneSpring differential expression analysis assessed the significant modulation of target genes between the SoS patient group and the healthy control. Differentially expressed genes were validated by real-time PCR. A GSEA pathway analysis identified the most significantly altered biological processes and pathways.
Life 12 00988 g001
Figure 2. Functional interaction network among differentially expressed genes between SoS patients and healthy controls.
Figure 2. Functional interaction network among differentially expressed genes between SoS patients and healthy controls.
Life 12 00988 g002
Figure 3. Real-time PCR validation of differentially expressed genes. (a) Quantitative expression measurements of five noncoding RNAs (i.e., lnc-C15orf57, MIR646, lnc-C2orf84-1, Inc-C00665 and lnc-C20orf197-3) and (b) seven mRNAs (i.e., NOS3, ZNF883, CD19, SFN, GSC, NDRG2 and SORBS1). The fold change represents the average difference in expression level of the respective genes between the SoS fibroblasts and the controls. The black bars depict the fold changes detected with the microarray, whereas the white bars show the average fold change with respect to the triplicate real-time PCR assessments. The fold changes indicating a downregulated expression are represented with negative values. The error bars represent the standard error of the mean. Significant levels were at p < 0.05. (c) Compared to healthy controls, five noncoding RNAs (lin-C005, Inc-C20orf97, MIR646HG, lnc-C15orf57 and Inc-C20orf84-1) and (d) seven mRNAs (NOS3, ZNF883, CD19, SFN, NDRG2, SORBS1 and GSC) were selected. The black bars depict healthy control (Ctr), and the white bars represent the Sotos syndrome patients (SoS pts). The results were consistent with the findings obtained from the microarray analysis. Data are presented as the mean ± standard deviation. * p < 0.05, ** p < 0.01.
Figure 3. Real-time PCR validation of differentially expressed genes. (a) Quantitative expression measurements of five noncoding RNAs (i.e., lnc-C15orf57, MIR646, lnc-C2orf84-1, Inc-C00665 and lnc-C20orf197-3) and (b) seven mRNAs (i.e., NOS3, ZNF883, CD19, SFN, GSC, NDRG2 and SORBS1). The fold change represents the average difference in expression level of the respective genes between the SoS fibroblasts and the controls. The black bars depict the fold changes detected with the microarray, whereas the white bars show the average fold change with respect to the triplicate real-time PCR assessments. The fold changes indicating a downregulated expression are represented with negative values. The error bars represent the standard error of the mean. Significant levels were at p < 0.05. (c) Compared to healthy controls, five noncoding RNAs (lin-C005, Inc-C20orf97, MIR646HG, lnc-C15orf57 and Inc-C20orf84-1) and (d) seven mRNAs (NOS3, ZNF883, CD19, SFN, NDRG2, SORBS1 and GSC) were selected. The black bars depict healthy control (Ctr), and the white bars represent the Sotos syndrome patients (SoS pts). The results were consistent with the findings obtained from the microarray analysis. Data are presented as the mean ± standard deviation. * p < 0.05, ** p < 0.01.
Life 12 00988 g003aLife 12 00988 g003b
Table 1. Differentially expressed probe sets relative to noncoding RNAs (ncRNAs) in SoS samples.
Table 1. Differentially expressed probe sets relative to noncoding RNAs (ncRNAs) in SoS samples.
NrProbe Set IDGene SymbolGene NameSeq. NameLog2 Fold Changep-Valuep-Value Adjusted
1A_22_P00002837Lnc-C20orf197-3long intergenic nonprotein-coding RNAlnc-C20orf197-3:133.24.95 × 10−70.007
2A_22_P00002837Lnc-C2orf84-1long intergenic nonproteinlnc-C2orf84-1:13.24.95 × 10−70.007
3A_22_P00002715MIR646HGMIR646 host geneENSG000002283402.88.64 × 10−60.042
4A_24_P16214LINC00665long intergenic nonprotein-coding RNA 665ENST000004278681.22.57 × 10−60.016
5A_32_P98975C15orf57chromosome 15 open reading frame 57NM_052849−0.71.40 × 10−50.047
Table 2. Significant differentially expressed probe sets relative to mRNAs in SoS samples.
Table 2. Significant differentially expressed probe sets relative to mRNAs in SoS samples.
NrProbe Set IDGene SymbolGene NameSeq. NameLog2 Fold Changep-Valuep-Value Adjusted
1A_33_P3389286SFNstratifinNM_0061423.991.62 × 10−60.015
2A_23_P113572CD19CD19 moleculeNM_0017703.361.28 × 10−60.014
3A_33_P3305790NOS3nitric oxide synthase 3 (endothelial cell)NM_0006032.41.41 × 10−70.0027
4A_22_P00017915ZNF883zinc finger protein 883NM_0011013381.334.95 × 10−60.026
5A_33_P3334515NDRG2NDRG family member 2NM_001282213−1.241.06 × 10−50.047
6A_24_P317907SORBS1sorbin and SH3 domain containing 1NM_001034954−3.11.14 × 10−50.047
7A_23_P76774GSCgoosecoid homeoboxNM_173849−3.981.38 × 10−50.047
Table 3. Selected significantly enriched gene sets in SoS samples found via GSEA.
Table 3. Selected significantly enriched gene sets in SoS samples found via GSEA.
Pathway or Process Description aNumber of Genes bFDR
q-Value c
Type of Regulation dOntology e
Cancer Process
Florio neocortex basal radial glia DN1510.00DOWNCGP
Kong E2F3 targets770.00DOWNCGP
Kobayashi EGFR-signaling 24 h DN2130.00DOWNCGP
Whitfield cell cycle literature390.00DOWNCGP
Whiteford pediatric cancer markers1020.00DOWNCGP
Nakayama soft-tissue tumors PCA2 UP760.00DOWNCGP
Zhan multiple myeloma pr UP350.00DOWNCGP
Pyeon HPV positive tumors UP790.00DOWNCGP
Sotiriou breast cancer grade 1 vs. 3 UP1273.65 × 10−4DOWNCGP
Chiang liver cancer subclass proliferation UP1468.15 × 10−4DOWNCGP
Farmer breast cancer cluster 2290.001DOWNCGP
Villanueva liver cancer KRT19 UP1470.002DOWNCGP
Rodrigues thyroid carcinoma DN660.002DOWNCGP
Finetti breast cancer kinome red150.002DOWNCGP
Riggi Ewing sarcoma progenitor DN1580.002DOWNCGP
Azare neoplastic transformation by STAT3 DN160.002DOWNCGP
West adrenocortical tumor markers UP1030.005DOWNCGP
Li Wilms tumor anaplastic UP150.010DOWNCGP
Li prostate cancer epigenetic270.014DOWNCGP
GAL leukemic stem cell DN1880.014DOWNCGP
Rickman head and neck cancer B370.018DOWNCGP
Poola invasive breast cancer UP2280.019DOWNCGP
Sengupta nasopharyngeal carcinoma UP2400.024DOWNCGP
SMID breast cancer luminal A UP680.025DOWNCGP
Vantveer breast cancer BRCA1 UP280.027DOWNCGP
Winnepenninckx melanoma metastasis UP1340.029DOWNCGP
Ferreira Ewing’s sarcoma unstable vs. stable UP1310.029DOWNCGP
Lopes methylated in colon cancer UP220.047DOWNCGP
Chiaradonna neoplastic transformation KRAS UP1130.047DOWNCGP
Cell Cycle and
Proliferation Process
Fischer G2 M cell cycle1910.00DOWNCGP
Croonquist NRAS signaling DN550.00DOWNCGP
Molenaar targets of CCND1 and CDK4 DN480.00DOWNCGP
Ishida E2F targets390.00DOWNCGP
Reichert mitosis LIN9 targets210.00DOWNCGP
Graham normal quiescent vs. normal dividing DN741.15 × 10−4DOWNCGP
Odonnell TFRC targets DN1102.18 × 10−4DOWNCGP
Graham cml dividing vs. normal quiescent UP1570.002DOWNCGP
Eguchi cell cycle RB1 targets170.002DOWNCGP
Yu MYC targets UP400.003DOWNCGP
Zhou cell cycle genes in IR response 24HR1030.003DOWNCGP
Odonnell targets of MYC and TFRC DN360.006DOWNCGP
Alcalay AML by NPM1 localization DN1580.010DOWNCGP
Zhou cell cycle genes in IR response 6HR700.011DOWNCGP
Plasari TGFB1 targets 10HR DN2050.014DOWNCGP
Chicas RB1 targets growing1960.02DOWNCGP
ULE-splicing via NOVA2350.02DOWNCGP
Benporath proliferation1240.033DOWNCGP
Whitfield cell cycle G2 M1620.034DOWNCGP
Liang silenced by methylation 2500.034DOWNCGP
Nojima SFRP2 targets DN190.034DOWNCGP
Kamminga EZH2 targets370.037DOWNCGP
Graham CML quiescent_VS normal quiescent UP740.040DOWNCGP
GOBP heat generation150.014UPGO BP
GOBP regulation of meiotic cell cycle310.07UPGO BP
GOBP negative regulation of nuclear division420.016UPGO BP
GOBP regulation of feeding behavior200.021UPGO BP
GOBP positive regulation of organic acid transport290.017UPGO BP
GOBP regulation of nuclear division organization1060.046UPGO BP
GOBP kinetochore organization160.041UPGO BP
HALLMARK G2M checkpoint1720.001DOWNH
HALLMARK KRAS-signaling UP1740.0001DOWNH
HALLMARK mitotic spindle1630.008DOWNH
HALLMARK inflammatory response1590.006DOWNH
HALLMARK E2F targets1670.005DOWNH
HALLMARK apoptosis1300.017DOWNH
HALLMARK KRAS-signaling DN1660.032DOWNH
HALLMARK allograft rejection1600.036DOWNH
HALLMARK TNFA-signaling via NFKB1630.040DOWNH
Cell Differentiation Process
Boquest stem cell DN1860.05DOWNCGP
Sarrio epithelial mesenchymal transition UP1490.006DOWNCGP
Le neuronal differentiation DN160.034DOWNCGP
P53-Mediated Cell Cycle Arrest and Cellular Senescence Process
Wu apoptosis by CDKN1A via TP53470.002DOWNCGP
Tang senescence TP53 targets DN480.00DOWNCGP
a. Official name of the biological process. b. Number of genes in the biological process. c. The FDR q-value estimates the significance of the enrichment of a biological process or a pathway. FDR q-values <= 0.05 are considered acceptable. d. Type of regulation of the genes involved in a process or a pathway. e. Name of the ontology defining a biological process or pathway. GO BP: gene ontology biological process; H: hallmark pathways; CGP: chemical and genetic perturbations pathways.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Conteduca, G.; Cangelosi, D.; Coco, S.; Malacarne, M.; Baldo, C.; Arado, A.; Pinto, R.; Testa, B.; Coviello, D.A. NSD1 Mutations in Sotos Syndrome Induce Differential Expression of Long Noncoding RNAs, miR646 and Genes Controlling the G2/M Checkpoint. Life 2022, 12, 988. https://doi.org/10.3390/life12070988

AMA Style

Conteduca G, Cangelosi D, Coco S, Malacarne M, Baldo C, Arado A, Pinto R, Testa B, Coviello DA. NSD1 Mutations in Sotos Syndrome Induce Differential Expression of Long Noncoding RNAs, miR646 and Genes Controlling the G2/M Checkpoint. Life. 2022; 12(7):988. https://doi.org/10.3390/life12070988

Chicago/Turabian Style

Conteduca, Giuseppina, Davide Cangelosi, Simona Coco, Michela Malacarne, Chiara Baldo, Alessia Arado, Rute Pinto, Barbara Testa, and Domenico A. Coviello. 2022. "NSD1 Mutations in Sotos Syndrome Induce Differential Expression of Long Noncoding RNAs, miR646 and Genes Controlling the G2/M Checkpoint" Life 12, no. 7: 988. https://doi.org/10.3390/life12070988

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