Next Article in Journal
The Evolution and Expression Profiles of EC1 Gene Family during Development in Cotton
Next Article in Special Issue
Meta-Analysis of Transcriptome-Wide Association Studies across 13 Brain Tissues Identified Novel Clusters of Genes Associated with Nicotine Addiction
Previous Article in Journal
Polychromatic Flow Cytometric Analysis of Stromal Vascular Fraction from Lipoaspirate and Microfragmented Counterparts Reveals Sex-Related Immunophenotype Differences
Previous Article in Special Issue
Novel In-Frame Deletion in HTRA1 Gene, Responsible for Stroke at a Young Age and Dementia—A Case Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spatial and Temporal Expression of High-Mobility-Group Nucleosome-Binding (HMGN) Genes in Brain Areas Associated with Cognition in Individuals with Down Syndrome

by
Alejandra Rodríguez-Ortiz
1,
Julio César Montoya-Villegas
1,
Felipe García-Vallejo
1 and
Yecid Mina-Paz
2,*
1
Laboratory of Molecular Biology and Pathogenesis, Department of Physiological Sciences, School of Basic Sciences, Faculty of Health, Universidad del Valle, Cali 76001, Colombia
2
Health and Movement Research Group, Faculty of Health, Universidad Santiago de Cali, Cali 76001, Colombia
*
Author to whom correspondence should be addressed.
Genes 2021, 12(12), 2000; https://doi.org/10.3390/genes12122000
Submission received: 30 October 2021 / Revised: 8 December 2021 / Accepted: 9 December 2021 / Published: 17 December 2021
(This article belongs to the Special Issue Gene Expression and Chromatin Modification in the Brain)

Abstract

:
DNA methylation and histone posttranslational modifications are epigenetics processes that contribute to neurophenotype of Down Syndrome (DS). Previous reports present strong evidence that nonhistone high-mobility-group N proteins (HMGN) are epigenetic regulators. They play important functions in various process to maintain homeostasis in the brain. We aimed to analyze the differential expression of five human HMGN genes in some brain structures and age ranks from DS postmortem brain samples. Methodology: We performed a computational analysis of the expression of human HMGN from the data of a DNA microarray experiment (GEO database ID GSE59630). Using the transformed log2 data, we analyzed the differential expression of five HMGN genes in several brain areas associated with cognition in patients with DS. Moreover, using information from different genome databases, we explored the co-expression and protein interactions of HMNGs with the histones of nucleosome core particle and linker H1 histone. Results: We registered that HMGN1 and HMGN5 were significantly overexpressed in the hippocampus and areas of prefrontal cortex including DFC, OFC, and VFC of DS patients. Age-rank comparisons between euploid control and DS individuals showed that HMGN2 and HMGN4 were overexpressed in the DS brain at 16 to 22 gestation weeks. From the BioGRID database, we registered high interaction scores of HMGN2 and HMGN4 with Hist1H1A and Hist1H3A. Conclusions: Overall, our results give strong evidence to propose that DS would be an epigenetics-based aneuploidy. Remodeling brain chromatin by HMGN1 and HMGN5 would be an essential pathway in the modification of brain homeostasis in DS.

1. Introduction

The continuous chromatin modification and the binding of tissue-specific transcription factors to their specific targets in chromatin maintain the epigenetic landscape necessary to regulate the cell-type-specific transcription [1,2]. However, additional chromatin modifiers, including the H1 linker histones [2] and the high mobility group N (HMGN) proteins, can remodel the chromatin organization and transcription regulation, playing important functions in several process to maintain the general homeostasis [3]. Down Syndrome (DS) is a chromosomal aneuploidy caused by a total or partial triplication of chromosome 21, but in rare cases it can be associated with a process of chromosome translocation [4]. In people with DS, the gene dose imbalance by triplication of genes on HSA21 is mostly associated with a wide spectrum of pathologies that include neurological and systemic diseases [4]. The incidence of trisomy 21 is influenced by maternal age and differs throughout the population [5,6]. In developed countries, the average life span for DS population is 55 years [7]. Although DS is the result of the increased copy number of a single 21 chromosome, the regulation of gene expression is affected at a genome-wide level [8,9,10,11].
There is a growing line of evidence proposing that beyond HSA21 trisomy, DS is an epigenetics-based syndrome [12]. For instance, a previous study with fetal skin fibroblasts from a set of monozygotic twins revealed regions that were predominantly hypermethylated in DS in genes involved in embryonic organ morphogenesis. Reprogramming of the DS fibroblasts to induced pluripotent stem cells (iPSCs) showed that these regions were maintained in the pluripotent state and correlated with differential gene expression and increased expression of the DNA methyltransferases DNMT3B and DNMT3L [13]. Thus, the genome-wide differences seen in DS tissues are correlated with epigenetic modifications that would be responsible, in part, for the establishment and/or maintenance of differential expression of genes in and outside of HSA21 in DS.
HMGN proteins are a nonhistone protein family that includes five members encoded by five specific genes with a similar intron-exon organization, localized along human genomes in different chromosome loci (Table 1) [13,14,15,16,17,18]. Previous reports show that HMGN proteins are the only nuclear proteins known to specifically recognize the generic structural features of the 147 base pair nucleosome core particles. In vitro analyses showed that at low ionic strength, nucleosome core particles can bind to HMGN proteins with high affinity [18,19,20,21]. The interaction between HMGN proteins and nucleosomes is dynamic, and the proteins compete among themselves and with the linker histone H1 for chromatin binding sites [22,23,24]; in fact, all HMGN proteins have similar affinities when binding to chromatin [22]. This interaction has been shown to affect the compaction level of the chromatin, modulating epigenetic events, and defining transcription profiles [21].
Since HMGN proteins play an important role as molecules that reshape the organization of chromatin and transcription levels, we performed a bioinformatic simulation analysis of their spatial and temporal expression in several areas of the brain with DS. To perform these analyses, we obtained log2 data from a free-access microarray previously consigned in the GEO DataSets of NCBI (https://www.ncbi.nlm.nih.gov/ (accessed on 1 December 2018). To explore the differential brain expression of HMGN genes, we calculated the Z-ratio from DS postmortem brain samples, specifically from those brain areas associated with cognitive processes previously described by Olmos-Serrano et al., 2016 [25].

2. Methodology

2.1. Data Mining

Raw gene expression data of DS samples and normal samples were downloaded from the Gene Expression Omnibus (GEO) (http://www.ncbi.nlm.nih.gov/geo/ (accessed on 1 December 2018)) of the National Center for Biotechnology Information (NCBI). For the analyses performed in the present study, we selected the human HMGN genes previously consigned in the Gene Entrez of the NCBI database (https://www.ncbi.nlm.nih.gov/gene (accessed on 17 September 2020)) (Table 1). Moreover, for all calculations, we used the log2 transformed expression values of free-access DNA microarray experiment whose registration code in the GEO database is GSE59630 (http://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE59630 (accessed on 1 December 2018)), previously deposited by Olmos-Serrano et al. 2016. [25].
According to the information consigned in the GEO database, the selected microarray experiment included gene expression data of more than 17000 probes from 58 post-mortem brain samples of DS individuals (25 from females and 33 from males) and 58 euploid samples as euploid controls (25 from females and 33 from males) that were classified by sex, age, and also by some brain areas including the hippocampus (HIP), cerebellar cortex (CBC), dorsolateral prefrontal cortex (DFC), orbital prefrontal cortex (OFC), ventrolateral prefrontal cortex (VFC), medial prefrontal cortex (MFC), primary somatosensory cortex (S1C), inferior parietal cortex (IPC), primary visual cortex (V1C), superior temporal cortex (STC), and inferior temporal cortex (ITC). Nevertheless, for the present study, we decided to analyze, not only the brain as a whole, but also OFC, MFC, HIP, and CBC brain regions that are highly associated with neurophenotypes of DS.

2.2. Data Preprocessing

Using the Partek Genomics Suite version 6.7 (Partek Incorporated, St. Louis, MO, USA), the robust multiarray analysis (RMA) algorithm [26] in Affymetrix Power Tools (APT; http://www.affymetrix.com/, accessed on 1 December 2018) was applied by Olmos-Serrano et al. 2016. [25], combined with an R-script to perform background correction and standardization for all raw data, aiming to filter false-positive data. The applied criterion was as follows: at least half the samples had PLIER signal intensity values greater than 100 [27].

2.3. Quantification of the Differential HMGN Genes Expression

Raw intensity log2 data from each experiment were used for the calculation of Z-score [28]. Z-scores of the protein coding genes analyzed were calculated according to Equation (1):
Z s c o r e = ( L o g   i n t e n s i t y   o f   G m e a n log i n t e n s i t y   G G n ) S t a n d a r d   D e v i a t i o n log G G n
Equation (1). Z-score formula
All Z-score values were normalized on a linear scale −3.0 ≤ 0 ≥ +3.0 (two-tailed p value < 0.001). From Z-score data, we calculated the mean values per gene and per structure in brain samples of DS and euploid controls. These data were used to calculate the Z-ratio (Equation (2)), a measure to estimate differential gene expression, where genes with values over 1.96 are considered over-expressed [28].
Z r a t i o = [ ( Z s c o r e G 1 a v e ) D S ( Z s c o r e G 1 a v e ) C o n ] S D   o f   Z s c o r e   d i f f e r e n c e s G 1 G n
Equation (2). Z-ratio formula

2.4. Gene-Dosage Imbalanced Quantification

To find out the gene dosage imbalance of the five HMGN genes in the structures of DS brain samples, first we calculated the M values according to Equation (3), and then we used the M value to calculate the ratio of the dosage imbalance R (DS Control ratio) as shown in Equations (3) and (4) [29].
Equation (3) M-value formula
M = [ M e a n   L o g 2 ( D S ) M e a n   L o g 2 ( C o n t r o l )
Equation (4) R (DS/Control ratio) formula
R = 2 M
R values ranging from 0.80 to 1.30 were considered as normal balanced (two copies of the gene); on the contrary, if R values were in the range 1.4 ≤ 1.5 ≥ 1.7, genes were dosage-imbalanced by triplication (three copies per gene), but if R ratio was greater than 1.8, genes were amplified (more than three copies).

2.5. Construction of HMGN Genes Network Using GeneMania

To build the gene interaction network of HMGN genes, we used the free-access platform GeneMANIA (http://www.genemania.org (accessed on 17 September 2020)), a real-time multiple association network actively developed at the University of Toronto, in the Donnelly Centre for Cellular and Biomolecular Research that uses a massive set of functional association data [30]. All calculations carried out in the present study were processed using the updated 2018 version [31].

2.6. Protein–Protein Interaction Analysis

To simulate the interaction between each HMGN with several histones of the core particle and H1, we obtained data from BioGRID (Database of Protein, Chemical, and Genetic Interactions), a free-access database (https://thebiogrid.org/ (accessed on 17 September 2020)) [32,33]. BioGRID is an interaction repository with data compiled through comprehensive curation efforts. The current index is version 3.5 and all data are freely provided via their search index and available for download in standardized formats. The different searches performed in the present study, were from data updated by January of 2019 [32,33].

2.7. Statistical Analysis

To compare mean values of Z-ratio of DS brain, we performed multivariate statistical analyses among the different brain cortex structures between DS patients and euploid controls. The Z test/Two-tailed was used to calculate differences in HMGN differential expression. The p-values were calculated using the web tool p-value from Z-score Calculator (https://www.socscistatistics.com/pvalues/normaldistribution.aspx (accessed on 17 September 2020)). In all cases, we used an alpha of 0.05 to test the significance of H0. All analyses were run in SPSS program version 25.0 (https://spss.softonic.com/ (accessed on 17 September 2020)) and Cytoscape 3.6 (https://cytoscape.org/release_notes_3_6_0.html (accessed on 17 September 2020)).

3. Results

3.1. Expression of HMGN Genes in Brain Areas from Individuals with DS

In general, we observed that the expression of five HMGN genes was variable along all structures under analysis. Moreover, we recorded significant differences in their overexpression values depending of the brain area under analysis. In this sense, the genes encoding for HMGN1 and HMGN5 were overexpressed not only in HIP, CBC, and V1C but also in some areas of prefrontal cortex including DFC, OFC, and VFC (values of Z-ratio > 1.96); in contrast, HMGN2 and HMGN3 genes had not significant overexpression. Only in ITC did the HMGN5 gene register a significant Z-ratio (Z-ratio = 2.0) (Table 2).
Since the HMGN1 gene is localized at the 21q22.2 band, we calculated the level of dose imbalance in those brain structures including in the present study. Our results showed that HMGN1 was dosage imbalanced, in OFC, VFC, and CBC by triplication (R > 1.4), but in HIP, DFC, and ITC, it was dysbalanced by amplification (R > 1.8).

3.2. Age Dependent Expression of HMGN Genes in the Brain of Individuals with DS

HMGN4 (Z-ratio = 4.72) and HMGN2 (Z-ratio = 2.13) were significantly overexpressed in prenatal samples of DS brain (16 to 22 weeks of gestation) in comparison to other age ranks (Table 3). Z-ratio data for HMGN3 showed significant overexpression values in the brain of DS during the first year (0-12 months), childhood (2 to 10 years), 12- to 22-year-old samples, and adulthood (32 to 42 years old), but not in brain samples of prenatal brain samples (16 to 22 weeks of gestation; Z-ratio = 1.08) (Table 3). In contrast, Z-ratio values for HMGN1 and HMGN5 along the different age ranks were non-significant, except for HMGN1 that was overexpressed in rank of 12 to 22 years old (Z-ratio = 2.0) (Table 3).

3.3. Protein–Protein Interaction Network and GO Categories

The Protein–Protein Interaction (PPI) network made with the five HMGN genes accounted for a total of 73 nodes, two connected components, one multi-edge node pair, an average number of neighbors of 2.374, and a heterogeneity of 2.012 (Figure 1). The node with the highest number of interactions was HMGN1 with 30, followed closely by HMGN2 with 28. Most relevant GO categories of biological processes obtained from the network included indispensable epigenetic processes for chromatin activation or inactivation such as histone deacetylation (p-value 2.46 × 10−8) and histone H3-K4 methylation (p-value 1.88 × 10−7).

3.4. HMGN Protein Interaction with Histones of Nucleosome Core and Linker H1

Data from several experimental methods reported in BioGRID database showed differential high interaction scores of HMGN proteins with histones of the nucleosome core HIST1H2AG, HIST1H2BA, HIST1H3A, and HIST1H4A (Table 4). Only HMGN2 had a significant high score of interaction with the linker histone HIST1H1A and the histones of nucleosome core HIST1H3A and HIST1H2AG. HMGN1 and HMGN5 showed significant interaction scores only with HIST1H4A. Finally, HMGN3 interacts with the HIST1H4A.

4. Discussion

Previously, some studies presented strong evidence that in DS individuals the genome-wide epigenomic alterations occur not only in chromosome 21 but also in some other chromosomes [34,35,36]. These include changes in gene expression, RNA content, and epigenetic histone modifications, nucleosome spacing, and DNA methylation process, which are dependent on health status and age [12]. Specifically, there is strong evidence that HMGN proteins play a role in epigenetic regulation of gene expression and play important functions in several biological processes to maintain normal homeostasis and altered gene expression in disease [12,37,38]. Thus, we aimed to analyze, using a bioinformatics approach, the gene expression of human HMGN genes in different human brain structures and age ranks, comparing DS brain samples and euploid controls.
Most of the literature reports about the expression of HMGNs in the brain as as whole and also in neuronal derived cells come from experiments carried out in mice [39]. In this sense, its interpretation could be cautiously extended to the brain of individuals with DS. In this scenario, our results are the first to analyze, in a representative sample of euploid individuals and DS individuals, the differential expression of the five human HMGNs in several area of the brain that are involved in learning and memory and also its age rank variation. Our bioinformatics approach allowed us to obtain strong statistical evidence of the differential regulation of HMGNs in the disruption of the normal brain homeostasis in some areas associated with the DS neurophenotype.
HMGN1 is in a region of human chromosome 21, and it is frequently found triplicated in DS samples. Our results not only confirmed the previous reports but extended the data to the brains of individuals with DS. The HMGN1 gene was dose-dysbalanced by triplication in the whole brain and brain cortex and dysbalanced by amplification in the hippocampus. The hippocampus is a brain structure that plays a major role in neural plasticity and cognition [40], which is known to the dysregulated in individuals with DS. Our results showed that methyl CpG-binding protein 2 (MeCP2) is underexpressed in several structures of the brain of DS, which can be linked to the dysregulation of the HMGN1 gene, given that this latter gene can affect the expression of MeCP2 by changing the chromatin structure and histone modifications in the MeCP2 promoter [40].
According to the PPI network, the nodes with the highest number of interactions were HMGN1 and HMGN2. Moreover, the GO categories showed a global implication of these genes in chromatin remodeling processes such as acetylation, methylation of histones, and dendritic spine morphogenesis. The dysregulation of gene expression recorded in these genes would most certainly affect the interactions with others and would possibly lead to the epigenomic changes found in individuals with DS.
Some analysis suggests that HMGNs could differentially modulate the global gene transcription in not only some brain structures but also in other tissues [40]. Therefore, the contribution of HMGN1 and HMGN5 to the transcriptional dysregulation of DS neurophenotype needs to be studied separately in specific developmental scenarios [41]. In mice, HMGN1 is a negative regulator of the brain expression of MeCP2, which promotes HMGN1 overexpression associated with some effects not only in general behavioral activities but also in anxiety and social deficits [38]. On the other hand, HMGN5, is thought to reduce the compaction of the chromatin fiber nucleosomes, thereby enhancing transcription from chromatin templates; nevertheless, it has not been related to DS.
Excluding the gestational period (12–16 WG), only HMGN3 had a statistically significant expression across different age-ranks. It contrasts with the expression of HMGN2 and HMGN4, which had significant differential expression in the gestational period but not in the rest of the age-ranks we evaluated. In this sense, our results support some data found in the literature that HMGN3 control part some epigenetic mechanisms during the neuronal development [42]. HMGN2 expression has been widely associated with embryogenesis [43]; anti-sense manipulation of HMGN2 gene leads to early embryonic abnormalities [44,45]. Our results suggest that HMGN2 regulates active and bivalent genes by promoting an epigenetic landscape of active histone modifications at promoters and enhancers, stabilizing the epigenetic landscape necessary to maintain the pluripotent identity of pluripotent stem cells [45].

5. Conclusions

Our results gave strong evidence to support the hypothesis of the crucial role of non-histones HMGN1 and HMGN5 proteins as important spatial and temporal remodelers that would change, by epigenetic process, the brain proteostasis in patients with DS. It is important to highlight that even though not all HMGN genes are located in chromosome 21, they presented a distinctive dysregulation, showing that the complexity seen in DS goes beyond chromosome 21. We also report the differential interaction of HMGN family proteins with histones of the nucleosome core HIST1H2AG, HIST1H4A, HIST1H2AG, HIST1H3A, HIST2H2AB, and HIST1H4A and also with the linker HIST1H1A. In this context, we propose that HMGN proteins play an important role in the topological process of remodeling the chromatin in several brain areas of individuals with DS that are associated with memory and learning processes. The global effect of this epigenetic deregulation would be the alteration of the brain homeostasis that potentially conditions the DS brain’s epigenetics mode.

Author Contributions

A.R.-O., F.G.-V. and Y.M.-P. searched data from a DNA microarray experiment in SD and found microarray experiment with ID GSE59630 previously deposited in the GEO DataSet of NCBI database. A.R.-O. and Y.M.-P. using the transformed log2 data, analyzed the differential expression of five HMGN genes in several brain areas associated with cognition in patients with DS. A.R.-O., Y.M.-P. and J.C.M.-V. explored the co-expression and protein interactions with the histones of nucleosome core particle and linker H1 histone. All authors worked on the results and conclusions. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been funded by the Directorate General of Research (Dirección General de Investigaciones) of the Universidad Santiago de Cali under call N° 01-2021.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

DNA microarray experiment whose registration code in the GEO database is GSE59630 (http://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE59630 (accessed on 1 December 2018)).

Acknowledgments

To the Universidad Santiago de Cali and the Universidad del Valle.

Conflicts of Interest

None of the authors presents a conflict of interest.

References

  1. Zhu, F.; Farnung, L.; Kaasinen, E.; Sahu, B.; Yin, Y.; Wei, B.; Dodonova, S.O.; Nitta, K.R.; Morgunova, E.; Taipale, M.; et al. The interaction landscape between transcription factors and the nucleosome. Nature 2018, 562, 76–81. [Google Scholar] [CrossRef]
  2. Hochedlinger, K.; Jaenisch, R. Induced pluripotency and epigenetic reprogramming. Cold Spring Harb. Perspect. Biol. 2015, 7, a019448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kugler, J.E.; Deng, T.; Bustin, M. The HMGN family of chromatin-binding proteins: Dynamic modulators of epigenetic processes. Biochimica et Biophysica acta 2012, 1819, 652–656. [Google Scholar] [CrossRef] [Green Version]
  4. Patterson, D. Genetic mechanisms involved in the phenotype of Down Syndrome. Ment. Retard Dev. Disabil. Res. Rev. 2007, 13, 199–206. [Google Scholar] [CrossRef] [PubMed]
  5. Carothers, A.D.; Hecht, C.A.; Hook, E.B. International variation in reported live birth prevalence rates of Down Syndrome, adjusted for maternal age. J. Med. Genet. 1999, 36, 386–393. [Google Scholar]
  6. Canfield, M.A.; Honein, M.A.; Yuskiv, N.; Xing, J.; Mai, C.T.; Collins, J.S.; Devine, O.; Petrini, J.; Ramadhani, T.A.; Hobbs, C.A.; et al. National estimates and race/ethnic-specific variation of selected birth defects in the United States, 1999–2001. Birth Defects Res. A Clin. Mol. Teratol. 2006, 76, 747–756. [Google Scholar] [CrossRef]
  7. Glasson, E.J.; Sullivan, S.G.; Hussain, R.; Petterson, B.A.; Montgomery, P.D.; Bittles, A.H. The changing survival profile of people with Down’s syndrome: Implications for genetic counselling. Clin. Genet. 2002, 62, 390–393. [Google Scholar] [CrossRef]
  8. Antonarakis, S.E. Down Syndrome and the complexity of genome dosage imbalance. Nat. Rev. Genet. 2016, 18, 147–163. [Google Scholar] [CrossRef]
  9. Do, C.; Xing, Z.; Yu, Y.E.; Tycko, B. Trans-acting epigenetic effects of chromosomal aneuploidies: Lessons from Down Syndrome and mouse models. Epigenomics 2017, 9, 189–207. [Google Scholar] [CrossRef] [Green Version]
  10. Mao, R.; Zielke, C.L.; Zielke, H.R.; Pevsner, J. Global up-regulation of chromosome 21 gene expression in the developing Down Syndrome brain. Genomics 2003, 81, 457–467. [Google Scholar] [CrossRef]
  11. Antonarakis, S.E.; Lyle, R.; Dermitzakis, E.T.; Reymond, A.; Deutsch, S. Chromosome 21 and down Syndrome: From genomics to pathophysiology. Nat Rev Genet 2004, 5, 725–738. [Google Scholar] [CrossRef]
  12. Gensous, N.; Franceschi, C.; Salvioli, S.; Garagnani, P.; Bacalini, M.G. Down Syndrome, Ageing and Epigenetics. Subcell Biochem. 2019, 91, 161–193. [Google Scholar] [CrossRef] [PubMed]
  13. Sailani, M.R.; Santoni, F.A.; Letourneau, A.; Borel, C.; Makrythanasis, P.; Hibaoui, Y.; Antonarakis, S.E. DNA-methylation patterns in trisomy 21 using cells from monozygotic twins. PLoS ONE 2015, 10, e013555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Rochman, M.; Postnikov, Y.; Correll, S.; Malicet, C.; Wincovitch, S.; Karpova, T.S.; McNally, J.G.; Wu, X.; Bubunenko, N.A.; Grigoryev, S.; et al. The Interaction of NSBP1 with Nucleosomes in Euchromatin Counteracts Linker Histone-Mediated Chromatin Compaction and Modulates The Fidelity Of The Cellular Transcription Profile. Mol. Cell. 2009, 35, 642–656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Shirakawa, H.; Landsman, D.; Postnikov, Y.V.; Bustin, M. NBP-45, a novel nucleosomal binding protein with a tissue-specific and developmentally regulated expression. J. Biol. Chem. 2000, 275, 6368–6374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Ueda, T.; Catez, F.; Gerlitz, G.; Bustin, M. Delineation of the protein module that anchors HMGN proteins to nucleosomes in the chromatin of living cells. Mol. Cell Biol. 2008, 28, 2872–2883. [Google Scholar] [CrossRef] [Green Version]
  17. Mardian, J.K.; Paton, A.E.; Bunick, G.J.; Olins, D.E. Nucleosome cores have two specific binding sites for nonhistone chromosomal proteins HMG 14 and HMG 17. Science 1980, 209, 1534–1536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Sandeen, G.; Wood, W.I.; Felsenfeld, G. The interaction of high mobility proteins HMG14 and 17 with nucleosomes. Nucleic Acids Res. 1980, 8, 3757–3778. [Google Scholar] [CrossRef] [Green Version]
  19. Rice, J.C.; Allis, C.D. Histone methylation versus histone acetylation: New insights into epigenetic regulation. Curr. Opin Cell Biol. 2001, 13, 263–273. [Google Scholar] [CrossRef]
  20. Belova, G.I.; Postnikov, Y.V.; Furusawa, T.; Birger, Y.; Bustin, M. Chromosomal protein HMGN1 enhances the heat shock-induced remodeling of Hsp70 chromatin. J. Biol. Chem. 2008, 283, 8080–8088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Tao, D.; Yuri, P.; Zhang, S.; Lillian, G.; Lore, B.; Ildikó, R.; Sabine, M.; Hölter, W.W.; Helmut, F.; Valerie, G.-D.; et al. Interplay between H1 and HMGN epigenetically regulates OLIG1&2 expression and oligodendrocyte differentiation. Nucleic Acids Res. 2017, 45, 3031–3045. [Google Scholar] [CrossRef]
  22. Singh, G.P.; Ganapathi, M.; Dash, D. Role of intrinsic disorder in transient interactions of hub proteins. Proteins 2007, 66, 761–765. [Google Scholar] [CrossRef]
  23. Uversky, V.N.; Oldfield, C.J.; Midic, U.; Xie, H.; Xue, B.; Vucetic, S.; Iakoucheva, L.M.; Obradovic, Z.; Dunker, A.K. Unfoldomics of human diseases: Linking protein intrinsic disorder with diseases. BMC Genom. 2009, 10, S7. [Google Scholar] [CrossRef] [Green Version]
  24. Zhang, S.; Deng, T.; Tang, W.; He, B.; Furusawa, T.; Ambs, S.; Bustin, M. Epigenetic regulation of REX1 expression and chromatin binding specificity by HMGNs. Nucleic Acids Res. 2019, 47, 4449–4461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Olmos-Serrano, J.L.; Kang, H.J.; Tyler, W.A.; Silbereis, J.C.; Cheng, F.; Zhu, Y.; Pletikos, M.; Jankovic-Rapan, L.; Cramer, N.P.; Galdzicki, Z.; et al. Down Syndrome Developmental Brain Transcriptome Reveals Defective Oligodendrocyte Differentiation and Myelination. Neuron 2016, 89, 1208–1222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Irizarry, R.A.; Hobbs, B.; Collin, F.; Beazer-Barclay, Y.D.; Antonellis, K.J.; Scherf, U.; Speed, T.P. Exploration, normalization, and summaries of high density oligonucleotide array probe level data. Biostatistics 2003, 4, 249–264. [Google Scholar] [CrossRef] [Green Version]
  27. Therneau, T.M.; Ballman, K.V. What does PLIER really do? Cancer Inform. 2008, 6, 423–431. [Google Scholar] [CrossRef] [PubMed]
  28. Cheadle, C.; Vawter, M.P.; Freed, W.J.; Becker, K.G. Analysis of microarray data using Z score transformation. J. Mol. Diagn. JMD 2003, 5, 73–81. [Google Scholar] [CrossRef] [Green Version]
  29. Aït Yahya-Graison, E.; Aubert, J.; Dauphinot, L.; Rivals, I.; Prieur, M.; Golfier, G.; Rossier, J.; Personnaz, L.; Creau, N.; Bléhaut, H.; et al. Classification of human chromosome 21 gene-expression variations in Down Syndrome: Impact on disease phenotypes. Am. J. Hum. Genet. 2007, 81, 475–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Mostafavi, S.; Ray, D.; Warde-Farley, D.; Grouios, C.; Morris, Q. GeneMANIA: A real-time multiple association network integration algorithm for predicting gene function. Genome Biol. 2008, 9 (Suppl. 1), S4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Franz, M.; Rodriguez, H.; Lopes, C.; Zuberi, K.; Montojo, J.; Bader, G.D.; Morris, Q. GeneMANIA update 2018. Nucleic Acids Res 2018, 46, W60–W64. [Google Scholar] [CrossRef] [Green Version]
  32. Breitkreutz, B.J.; Stark, C.; Tyers, M. The GRID: The General Repository for Interaction Datasets. Genome Biol. 2003, 4, R23. [Google Scholar] [CrossRef] [PubMed]
  33. Oughtred, R.; Stark, C.; Breitkreutz, B.J.; Rust, J.; Boucher, L.; Chang, C.; Kolas, N.; O’Donnell, L.; Leung, G.; McAdam, R.; et al. The BioGRID interaction database: 2019 update. Nucleic Acids Res. 2019, 47, D529–D541. [Google Scholar] [CrossRef] [Green Version]
  34. Costa, V.; Angelini, C.; D’Apice, L.; Mutarelli, M.; Casamassimi, A.; Sommese, L.; Gallo, M.A.; Aprile, M.; Esposito, R.; Leone, L.; et al. Massive-scale RNA-Seq analysis of non ribosomal transcriptome in human trisomy 21. PLoS ONE 2011, 6, e18493. [Google Scholar] [CrossRef]
  35. Lane, A.A.; Chapuy, B.; Lin, C.Y.; Tivey, T.; Li, H.; Townsend, E.C.; van Bodegom, D.; Day, T.A.; Wu, S.C.; Liu, H.; et al. Triplication of a 21q22 region contributes to B cell transformation through HMGN1 overexpression and loss of histone H3 Lys27 trimethylation. Nat. Genet. 2014, 46, 618–623. [Google Scholar] [CrossRef] [Green Version]
  36. Kahmann, N.H.; Rake, A.V. Altered nucleosome spacing associated with Down Syndrome. Biochem. Genet. 1993, 31, 207–214. [Google Scholar] [CrossRef] [PubMed]
  37. Hock, R.; Furusawa, T.; Ueda, T.; Bustin, M. HMG chromosomal proteins in development and disease. Trends Cell Biol. 2007, 17, 72–79. [Google Scholar] [CrossRef] [Green Version]
  38. Nanduri, R.; Furusawa, T.; Bustin, M. Biological Functions of HMGN Chromosomal Proteins. Int. J. Mol. Sci. 2020, 21, 449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Rubin, R.D.; Watson, P.D.; Duff, M.C.; Cohen, N.J. The role of the hippocampus in flexible cognition and social behavior. Front. Hum. Neurosci. 2014, 8, 742. [Google Scholar] [CrossRef] [Green Version]
  40. Abuhatzira, L.; Shamir, A.; Schones, D.E.; Schäffer, A.A.; Bustin, M. The chromatin-binding protein HMGN1 regulates the expression of methyl CpG-binding protein 2 (MECP2) and affects the behavior of mice. J. Biol. Chem. 2011, 286, 42051–44262. [Google Scholar] [CrossRef] [Green Version]
  41. Kugler, J.E.; Horsch, M.; Huang, D.; Furusawa, T.; Rochman, M.; Garrett, L.; Becker, L.; Bohla, A.; Hölter, S.M.; Prehn, C.; et al. High mobility group N proteins modulate the fidelity of the cellular transcriptional profile in a tissue- and variant-specific manner. J. Biol. Chem. 2013, 288, 16690–16703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Ito, Y.; Bustin, M. Immunohistochemical localization of the nucleosome-binding protein HMGN3 in mouse brain. J. Histochem. Cytochem. 2002, 50, 1273–1275. [Google Scholar] [CrossRef] [Green Version]
  43. Mohamed, O.A.; Bustin, M.; Clarke, H.J. High-mobility group proteins 14 and 17 maintain the timing of early embryonic development in the mouse. Dev. Biol. 2001, 229, 237–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Korner, U.; Bustin, M.; Scheer, U.; Hock, R. Developmental role of HMGN proteins in Xenopus laevis. Mech. Dev. 2003, 120, 1177–1192. [Google Scholar] [CrossRef] [PubMed]
  45. Garza-Manero, S.; Sindi, A.A.A.; Mohan, G.; Rehbini, O.; Jeantet, V.H.M.; Bailo, M.; Latif, F.A.; West, M.P.; Gurden, R.; Finlayson, L.; et al. Maintenance of active chromatin states by HMGN2 is required for stem cell identity in a pluripotent stem cell model. Epigenetics Chromatin. 2019, 12, 73. [Google Scholar] [CrossRef]
Figure 1. Protein-to-protein interaction network between the five human High-Mobility-Group Nucleosome-Binding (HMGN) proteins and several cellular human genes. (A) Topological structure of the network using Cytoscape program. (B) The statistically significant cellular genes that interact with some of one HMGN. Data were obtained from GeneMania database.
Figure 1. Protein-to-protein interaction network between the five human High-Mobility-Group Nucleosome-Binding (HMGN) proteins and several cellular human genes. (A) Topological structure of the network using Cytoscape program. (B) The statistically significant cellular genes that interact with some of one HMGN. Data were obtained from GeneMania database.
Genes 12 02000 g001
Table 1. Description of gene data of five encoding human high mobility binding nucleosome group proteins (HMGN). [Source of NCBI https://www.ncbi.nlm.nih.gov/gene/ (accessed on 17 September 2020)].
Table 1. Description of gene data of five encoding human high mobility binding nucleosome group proteins (HMGN). [Source of NCBI https://www.ncbi.nlm.nih.gov/gene/ (accessed on 17 September 2020)].
GeneID *LocusGene Name
HMGN1315021q22.2High mobility group nucleosome binding domain 1
HMGN231511p36.11High mobility group nucleosome binding domain 2
HMGN393246q14.1High mobility group nucleosome binding domain 3
HMGN4104736p22.2High mobility group nucleosome binding domain 4
HMGN579366Xq21.1High mobility group nucleosome binding domain 5
(*). According to the classification of the National Center for Biotechnology Information (NCBI).
Table 2. Mean Z-ratio values of expression for the five human High-Mobility-Group Nucleosome-Binding (HMGN) genes in several brain structures of individuals with Down Syndrome (DS).
Table 2. Mean Z-ratio values of expression for the five human High-Mobility-Group Nucleosome-Binding (HMGN) genes in several brain structures of individuals with Down Syndrome (DS).
GeneGene ID BrainHIPCBCDFCOFCV1CVFCITC
HMGN131502.813.002.352.632.073.001.862.45
HMGN231510.060.230.130.540.05−0.620.680.99
HMGN393242.481.291.601.781.572.822.311.86
HMGN4104730.370.850.770.270.840.410.19−0.27
HMGN5793662.652.181.602.491.131.492.532.88
Z-ratio value > 1.96 means significant gene overexpression in the brain of DS individuals. Gene ID and data source were obtained from the information consigned in NCBI GeoDataset of a microarray experiment with registration code of GSE59630. (HIP). Hippocampus; (CBC). Cerebellar brain cortex; (DFC). Dorsolateral prefrontal cortex; (OFC). Orbital prefrontal cortex; (V1C). Primary visual cortex (VFC). Ventrolateral prefrontal cortex and (ITC). Inferior temporal cortex.
Table 3. Mean values of Z-ratio for the five human High-Mobility-Group Nucleosome-Binding (HMGN) genes expressed in different age ranks of the brain of human Down Syndrome (DS) individuals.
Table 3. Mean values of Z-ratio for the five human High-Mobility-Group Nucleosome-Binding (HMGN) genes expressed in different age ranks of the brain of human Down Syndrome (DS) individuals.
Gene16–22 WG0–12 M2–10 Y12–22 Y30–39 Y40–42 Y
HMGN10.661.451.941.972.181.55
HMGN22.130.220.540.540.310.49
HMGN31.083.612.712.552.592.94
HMGN44.711.121.511.651.841.67
HMGN50.161.890.781.150.760.82
WG, Weeks of Gestation. M, Months. Y, Years. Data sources were previously consigned in the NCBI GeoDataset of a DNA microarray experiment under the registration code of GSE59630.
Table 4. BioGRID data of the five human HMGN interactions with the nucleosome core histones H2BA, H2AG, H2AB, H3A, and H4A and with the linker histone H1A.
Table 4. BioGRID data of the five human HMGN interactions with the nucleosome core histones H2BA, H2AG, H2AB, H3A, and H4A and with the linker histone H1A.
InteractorInteractionExperimental EvidenceThroughputScore *
HMGN1HIST1H4AAffinity Capture-MS (§)High>0.75
HMGN2HIST1H2BAAffinity Capture-MSHigh0.99
HIST1H3AAffinity Capture-MSHigh0.90
HIST1H2AGAffinity Capture-MSHigh0.77
HIST1H1AProximity Label-MS (§§)High>0.75
HMGN3HIST1H4AAffinity Capture-MSHigh>0.75
HMGN4HIST1H2AGAffinity Capture-MSHigh0.92
HIST1H3AProximity Label-MSHigh>0.75
HIST2H2ABAffinity Capture-MSHigh0.88
HMGN5HIST1H4AAffinity Capture-MS High>0.75
* The cut-off threshold is >0.75. Data from BioGRID (https://thebiogrid.org/ (accessed on 17 September 2020)). (§). Affinity Capture–MS interaction is inferred when a bait protein is affinity captured from cell extracts by either polyclonal antibody or epitope tag and the associated interaction partner is identified by mass spectrometric methods. (§§). Proximity Label–MS interaction is inferred when a bait-enzyme fusion protein selectively modifies a vicinal protein with a diffusible reactive product, followed by affinity capture of the modified protein and identification by mass spectrometric methods, such as the BioID system.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rodríguez-Ortiz, A.; Montoya-Villegas, J.C.; García-Vallejo, F.; Mina-Paz, Y. Spatial and Temporal Expression of High-Mobility-Group Nucleosome-Binding (HMGN) Genes in Brain Areas Associated with Cognition in Individuals with Down Syndrome. Genes 2021, 12, 2000. https://doi.org/10.3390/genes12122000

AMA Style

Rodríguez-Ortiz A, Montoya-Villegas JC, García-Vallejo F, Mina-Paz Y. Spatial and Temporal Expression of High-Mobility-Group Nucleosome-Binding (HMGN) Genes in Brain Areas Associated with Cognition in Individuals with Down Syndrome. Genes. 2021; 12(12):2000. https://doi.org/10.3390/genes12122000

Chicago/Turabian Style

Rodríguez-Ortiz, Alejandra, Julio César Montoya-Villegas, Felipe García-Vallejo, and Yecid Mina-Paz. 2021. "Spatial and Temporal Expression of High-Mobility-Group Nucleosome-Binding (HMGN) Genes in Brain Areas Associated with Cognition in Individuals with Down Syndrome" Genes 12, no. 12: 2000. https://doi.org/10.3390/genes12122000

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