Next Article in Journal
hMSH5 Regulates NHEJ and Averts Excessive Nucleotide Alterations at Repair Joints
Previous Article in Journal
Electrophysiological and Behavioral Evidence for Hyper- and Hyposensitivity in Rare Genetic Syndromes Associated with Autism
Previous Article in Special Issue
Comparative Study of Classification Algorithms for Various DNA Microarray Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Silico Analysis Identified Putative Pathogenic Missense nsSNPs in Human SLITRK1 Gene

by
Muhammad Zeeshan Ali
1,
Arshad Farid
1,
Safeer Ahmad
1,
Muhammad Muzammal
1,
Mohammed Al Mohaini
2,3,
Abdulkhaliq J. Alsalman
4,
Maitham A. Al Hawaj
5,
Yousef N. Alhashem
6,
Abdulmonem A. Alsaleh
6,
Eman M. Almusalami
7,
Mahpara Maryam
8 and
Muzammil Ahmad Khan
1,9,*
1
Gomal Center of Biochemistry and Biotechnology, Gomal University, Dera Ismail Khan 29111, Pakistan
2
Basic Sciences Department, College of Applied Medical Sciences, King Saud Bin Abdulaziz University for Health Sciences, Al Ahsa 31982, Saudi Arabia
3
King Abdullah International Medical Research Center, Al Ahsa 31982, Saudi Arabia
4
Department of Clinical Pharmacy, Faculty of Pharmacy, Northern Border University, Rafha 91911, Saudi Arabia
5
Department of Pharmacy Practice, College of Clinical Pharmacy, King Faisal University, Al Ahsa 31982, Saudi Arabia
6
Clinical Laboratory Sciences Department, Mohammed Al-Mana College for Medical Sciences, Dammam 34222, Saudi Arabia
7
King’s College London, Strand, London WC2R 2LS, UK
8
Department of Zoology, Government College No.1, Dera Ismail Khan 29111, Pakistan
9
Department of Human Genetics, Sidra Medical and Research Centre, Doha 26999, Qatar
*
Author to whom correspondence should be addressed.
Genes 2022, 13(4), 672; https://doi.org/10.3390/genes13040672
Submission received: 6 February 2022 / Revised: 5 April 2022 / Accepted: 7 April 2022 / Published: 11 April 2022

Abstract

:
Human DNA contains several variations, which can affect the structure and normal functioning of a protein. These variations could be single nucleotide polymorphisms (SNPs) or insertion-deletions (InDels). SNPs, as opposed to InDels, are more commonly present in DNA and may cause genetic disorders. In the current study, several bioinformatic tools were used to prioritize the pathogenic variants in the SLITRK1 gene. Out of all of the variants, 16 were commonly predicted to be pathogenic by these tools. All the variants had very low frequency, i.e., <0.0001 in the global population. The secondary structure of all filtered variants was predicted, but no structural change was observed at the site of variation in any variant. Protein stability analysis of these variants was then performed, which determined a decrease in protein stability of 10 of the variants. Amino acid conservation analysis revealed that all the amino acids were highly conserved, indicating their structural and functional importance. Protein 3D structure of wildtype SLITRK1 and all of its variants was predicted using I-TASSER, and the effect of variation on 3D structure of the protein was observed using the Missense3D tool, which presented the probable structural loss in three variants, i.e., Asn529Lys, Leu496Pro and Leu94Phe. The wildtype SLITRK1 protein and these three variants were independently docked with their close interactor protein PTPRD, and remarkable differences were observed in the docking sites of normal and variants, which will ultimately affect the functional activity of the SLITRK1 protein. Previous studies have shown that mutations in SLITRK1 are involved in Tourette syndrome. The present study may assist a molecular geneticist in interpreting the variant pathogenicity in research as well as diagnostic setup.

1. Introduction

Human DNA contains several variations in its sequence including single nucleotide polymorphisms (SNPs) and insertion deletions (InDels). However, SNPs are the most frequently occurring variations in the human genome. These variations in the genome may alter the protein structure and function and can affect the normal character(s) of an organism [1,2,3]. InDels cause substantial genetic variation in the genome of an organism. Many of the InDels occur at the functionally important part of the genome and hence may also play their role in disease onset [4]. Single nucleotide substitution may either cause missense or nonsense effect. The detailed classification of variant is shown in Figure 1. In missense effect, one amino acid is replaced by another amino acid, while the nonsense variants replace the coding codon with a stop codon that eventually leads to the truncation of protein [5]. About 90% of human genome polymorphisms comprises SNPs. Through genome wide prioritization, 0.12% of the variants out of total human genome are predicted to be pathogenic [6]. Several SNPs do not contribute to the causation of disease, but there are certain SNPs that are called missense SNPs or non-synonymous SNPs (nsSNPs) and that are involved in genetic disorders [7]. In about 50% of the total known mutations, nsSNPs are the major contributing factor [8,9].
Most of the disease-causing SNPs are reported at evolutionary-conserved regions of the human genome, which have great importance in the structure and function of proteins. It is very important to identify the pathogenicity of specific SNP for disease prognosis. The identification of SNPs involved in disease is a difficult job, as it requires multiple tests for hundreds to thousands of SNPs in candidate genes. Prioritizing SNPs using bioinformatical tools would be a possible way to overcome this problem. Bioinformatics prediction tools help us to discriminate disease-causing variants from neutral ones.
In the current study, several bioinformatics tools were used to investigate the structural and functional consequences of nsSNPs present in the coding region of the human SLITRK1 gene. We also predicted the 3D structure of the wildtype SLITRK1 protein and its prioritized predicted pathogenic variants. This is the first in silico study of the human SLITRK1 gene, which is helpful in predicting pathogenic nsSNPs in the coding region of the SLITRK1 gene.

2. Materials and Methods

2.1. Variant Recruitment

Variants of the SLITRK1 gene were recruited from the Ensembl genome browser (https://asia.ensembl.org/index.html, accessed on 28 February 2021). Manual variant filtration was performed on an MS Excel supported file enlisting all the variants of the SLITRK1 gene. Only those variants were selected for further analyses, which were nsSNPs (i.e., missense) and which fell within the coding region of the SLITRK1 gene. The nsSNPs were analyzed through various bioinformatics tools to find putative pathogenic variants (Figure 2).

2.2. Predicting Pathogenicity of Missense nsSNPs

Different online bioinformatics tools were used to predict the pathogenicity of filtered nsSNPs. These online tools were Polyphen2 (Polymorphism Phenotyping v2) (http://genetics.bwh.harvard.edu/pph2/, accessed on 20 March 2021), SNPs&Go (https://snps.biofold.org/snps-and-go/snps-and-go.html, accessed on 15 April 2021), Meta-SNP (https://snps.biofold.org/meta-snp/, accessed on 5 May 2021), PROVEAN (Protein Variation Effect Analyzer) (http://provean.jcvi.org/index.php, accessed on 2 June 2021), SIFT (Sorting Intolerant From Tolerant) (https://sift.bii.a-star.edu.sg, accessed on 27 June 2021), MutationAssessor (http://mutationassessor.org/r3/, accessed on 15 July 2021), PANTHER (Protein ANalysisTHrough Evolutionary Relationships) (http://pantherdb.org, accessed on 2 August 2021), PhD-SNP (Predictor of human Deleterious SNP) (https://snps.biofold.org/phd-snp/phd-snp.html, accessed on 17 August 2021), SNAP2 (https://rostlab.org/services/snap/, accessed on 5 September 2021) and PMut (http://mmb.irbbarcelona.org/PMut/, accessed on 28 September 2021).

2.3. Variant Frequency

The frequency of variants that are commonly predicted to be pathogenic by all the bioinformatics pathogenicity predictor tools was checked using dbSNP (https://www.ncbi.nlm.nih.gov/snp/, accessed on 2 October 2021).

2.4. Secondary Structure Prediction

The secondary structures of a normal SLITRK1 protein and common predicted pathogenic variants were analyzed through an online tool PSIPRED (http://bioinf.cs.ucl.ac.uk/psipred/, accessed on 4 October 2021).

2.5. Protein Stability Analysis

Protein stability of variants was checked using I-Mutant (http://gpcr2.biocomp.unibo.it/cgi/predictors/I-Mutant3.0/I-Mutant3.0.cgi, accessed on 8 October 2021) and MUpro (http://mupro.proteomics.ics.uci.edu, accessed on 9 October 2021). These are web-based tools that predict the stability of a mutated protein.

2.6. Conservation Analysis

Conservation analysis was performed using Clustal omega (https://www.ebi.ac.uk/Tools/msa/clustalo/, accessed on 20 October 2021) and ConSurf (http://consurf.tau.ac.il, accessed on 18 January 2022). The amino acid sequence of the SLITRK1 protein of humans and some other species was obtained from the HomoloGene sub-database of NCBI (https://www.ncbi.nlm.nih.gov, accessed on 20 October 2021) and submitted to Clustal Omega (https://www.ebi.ac.uk/Tools/msa/clustalo/, accessed on 20 October 2021). Multiple sequence alignment was performed on Clustal omega to check the conservation of amino acids of the SLITRK1 protein among different species including Xenopus tropicalis (western clawed frog), Gallus (red jungle fowl), Bos Taurus (domestic cow), Canis lupus familiaris (domestic dog), Macaca mulata (rhesus macaque), Pan troglodytes (chimpanzee), Mus musculus (house mouse) and Rattus norvegicus (brown rat).
The ConSurf web server [10] demonstrates the evolutionary pattern of the amino acids and nucleic acids by predicting the structural and functional areas. The results are predicted based on conservation scores that range from 1 to 9, where 1 indicates variable regions, 5 indicates mild conserved regions and 9 indicates highly conserved regions. However, exposed residues with high scores are considered functional residues, whereas buried residues with high scores are considered structural.

2.7. Protein 3D Structure Prediction

I-TASSER (https://zhanglab.ccmb.med.umich.edu/I-TASSER/, accessed on 1 December 2021) tool was used to predict the 3D structure of SLITRK1 protein and its commonly predicted pathogenic variants. Missense3D (http://missense3d.bc.ic.ac.uk/missense3d/, accessed on 10 January 2022) was used to predict structural changes in protein by substitution of an amino acid. UCSF Chimera (candidate version 1.15) was used to visualize the 3D structures of proteins retrieved from I-TASSER. 3D structure of normal and variant proteins was overlapped in UCSF Chimera to observe structural changes.

2.8. Protein–Protein Interactions

The interaction of SLITRK1 protein with other proteins was studied using online tool STRING (https://string-db.org, accessed on 20 January 2022), which predicts the top ten proteins that show interactions with the query gene. STRING predicts the interactors of a protein on the basis of gene fusion, co-expression, function and experimental data. It shows combined scores for each interacting protein, ranging from 0 to 1, where 0 shows the lowest interaction and 1 indicates the highest interaction [11].

2.9. Protein-Protein Docking

The online tool Cluspro was used for docking of the normal and mutant SLITRK1 proteins with its close functional interactor [12].
Figure 2. Flowchart of the methodology.
Figure 2. Flowchart of the methodology.
Genes 13 00672 g002

3. Results

3.1. Variant Recruitment

The total identified variants in SLITRK1 include 07 indels, 321 5′UTR and 624 3′UTR variants, and 5 nonsense variants. A total of 2255 variants were recruited from the Ensembl genome browser, which consisted of synonymous, non-synonymous, intronic, 3′UTR and 5′UTR variants. Out of the nonsynonymous variants, 7 were indels and 447 were SNPs. Of these nsSNPs, 442 variants were missense, i.e., causing change in amino acid, and 5 variants resulted in the formation of a stop codon, causing truncation of the protein. Only missense nsSNPs were selected for further bioinformatical analysis.

3.2. Pathogenicity Prediction of Variants

All the missense nsSNPs were subjected to ten bioinformatical tools to predict their biological pathogenicity. The proportion of pathogenic variants predicted by different software include 146 by Polyphen2, 101 by SNPs&Go, 102 by Meta-SNP, 103 by PROVEAN, 177 by SIFT, 104 by MutationAssessor, 395 by PANTHER, 141 by PhD-SNP, 160 by SNAP2 and 387 by PMut. Among all the variants, 16 were commonly predicted to be pathogenic by all the ten bioinformatical tools (Table 1). These 16 variants were then finally selected for further analysis due to their high susceptibility for being pathogenic.

3.3. Variant Frequency

dbSNP (https://www.ncbi.nlm.nih.gov/snp/, accessed on 2 October 2021) was used to check the frequency of commonly predicted pathogenic variants. The frequency of all the variants was very low (<0.0001) in the global population. This supported the findings produced by pathogenicity prediction tools.

3.4. Secondary Structure Prediction

The secondary structure predictions indicated that all the selected variants lie in the coil of the SLITRK1 protein and only one lies in the helix. There was no structural change observed at the point of change. However, some upstream and downstream changes were observed.

3.5. Protein Stability Analysis

I-Mutant and MUpro tools were used to check the stability of the SLITRK1 protein for selected amino acid substitutions. Among 16 commonly predicted pathogenic variants, 10 of the variants were commonly predicted to show decreases in the stability of the protein, which suggestively may cause greater loss to the SLITRK1 protein (Table 2).

3.6. Amino Acid Conservation

3.6.1. Clustal Omega

The Clustal Omega tool was used for multiple sequence alignment of the SLITRK1 protein of humans with other species. The results revealed that all the amino acids at the point of variations were highly conserved in all other species, indicating the evolutionary and functional importance of selected amino acids (Figure 3).

3.6.2. ConSurf

The ConSurf tool was used to identify the evolutionary conservation of amino acids of the SLITRK1 protein. ConSurf predicts which amino acids play structural or functional roles based on conservation and solvent accessibility. Residues are predicted as being functional when they are highly conserved and exposed and as structural when they are highly conserved and buried. The results indicate that all 10 nsSNPs that are predicted to be damaging are highly conserved i.e., nine variants having a conservation score of 9 and 1 showing a conservation score of 8. Of the above 10 amino acids, half were buried and predicted as structural residues while the rest were exposed and predicted as functional residues (Figure 4).

3.7. 3D Structure Predictions

Three-dimensional models of all SLITRKI variants were designed and superimposed with the 3D structure of wild SLITRK1 (Figure 5). The Missense3D tool was used to detect the structural changes that were caused due to substitution of the amino acids. The Missense3D tool predicted structural damage in variants I, III and VIII, while no structural damage was detected in the other variants. Hence, we selected variants I, III and VIII for further analysis. By manual comparison of all the structures, different changes in the folding pattern of protein were absorbed. The highest similarity index of 71.12% was shown in the 3D model of the mutant Asn529Lys with the wildtype 3D model of the SLITRK1 protein, while the lowest similarity index was shown in the 3D model of mutant Leu94Phe with the wildtype 3D model of the SLITRK1 protein. The similarity index of the 3D model of mutant Leu496Pro with the wildtype SLITRK1 protein was 31.32%.

3.8. Protein–Protein Interactions

The String tool was used to predict the close interactor protein of SLITRK1. The results showed that SLITRK1 has close interactions with the PTPRD, PTPRS, PTPRF, OPCML, PTPRA, PTPRE, DLGAP3, PTPRT, IGHMBP2 and SGCE proteins. However, the SLITRK1 protein determined PTPRD as being the closest functional interactor (Figure 6).

3.9. Protein–Protein Docking

The wildtype SLITRK1 protein and all its variants were docked with the close interactor PTPRD protein and notable differences were observed in the interacting sites of wildtype SLITRK1 and its variants with PTPRD. The results showed that wildtype SLITRK1 has interactions with PTPRD at eight amino acid residues, i.e., Ala680, Lys647, Phe641, Arg584, Thr604, Tyr582, Glu550 and Ser616. These interactions are through 8 interactive forces including 7 H-bonds and 1 unfavorable bond.
Variant Leu94Phe showed the lowest interaction with PTPRD, i.e., interacting at five different residues via four H-bonds and one unfavorable bond, while variant Leu496Pro showed the highest interaction with PTPRD, interacting at 14 different residues via 15 bonds, including 12 H-bonds and 3 unfavorable bonds. The interactions of SLITRK1 (normal and all variants) with close interactor PTPRD are diagrammatically shown in Figure 7.

4. Discussion

SNPs, also known as single nucleotide variants (SNVs), are the most commonly found variants in the human genome. According to an estimate, the human genome contains at least 11 million SNPs (1 per 300 bp on average) [13]. SNPs are found in protein coding as well as in non-protein coding regions [14,15]. Research has shown that variations in non-coding elements may also be the cause of several genetic conditions [16]. Numerous evidence has shown that there are some variants that are found in functional non-coding regions including chromatin marks, DNase hypersensitivity and enhancer elements [17,18]. The role of variations in non-coding regions including intergenic sequence, non-coding RNAs and non-coding elements in protein coding gene, is challenging to determine and needs to be better understood. The non-coding variants involved in causation of different disorders may be found in enhancer regions, promoter sites, or 5′UTR or 3′UTR of the gene [19]. Another important role of the non-coding variant is in the regulation of gene expression, which is a challenging task when identifying the effect of variation in molecular mechanism of gene regulation [20]. Studies have also demonstrated that the role of variations in non-coding regions are associated with the timing of DNA replication [21]. Variations in the non-coding genome are associated with various diseases, but to fully understand their functional effects, much research is still required.
In the human genome, around 24,000 to 60,000 coding SNPs are estimated [22,23]. nsSNPs are more significant because they have the potential to affect the structure and function of expressed proteins and are, therefore, likely to represent modifiers of inherited susceptibility to disease [24]. nsSNPs alter cellular functions in many ways. Indeed, nsSNPs often influence normal protein function through a combination of effects on protein stability, protein–protein interactions and many other features [25]. Numerous studies in the past have shown that nsSNPs are responsible for about 50% of mutations that are involved in various genetic disorders [9]. This information confirms that nsSNPs, especially missense SNPs, are associated with various human diseases. Recent studies on the nsSNPs using computational approaches reveal the potential impact of mutation on understanding the molecular mechanisms of various diseases [26,27,28].
SLITRK1 is a member of the SLITRK family and, similar to other members of the SLITRK family, is an integral membrane protein with the domain ‘2 N-terminal leucine rich repeat (LRR) [29]. SLITRK1 has the LRP1 domain, through which they interact with LAR receptor protein tyrosine phosphatases (PTPs) and control synapse formation [30,31,32].
The SLITRK1 gene was mapped to a region of chromosome 13q31 [29,30,31,32,33]. The SLITRK1 gene is highly expressed in adult and fetal brains; moderately expressed in lungs and pancreas; and has very low expressions in the ovaries, kidneys, heart and liver [33]. Mouse Slitrk1 cDNA was cloned by Aruga and Mikoshiba (2003), who discovered that the protein contains a signal peptide at N-terminus, which was followed by LRR domains and a transmembrane domain at the C-terminus [29]. Aruga and Mikoshiba (2003) performed a Northern blot analysis of various mouse tissues and found a very high expression of Slitrk1 only in the brain [29]. Until now, 14 different types of mutations have been identified in the SLITRK1 gene (HGMD). Mutations in the SLITRK1 gene have been reported to be involved in Tourette Syndrome (TS; OMIM No# 137580) [34]. TS is a neuropsychiatric disorder with an estimated onset in early childhood. It is characterized by vocal and motor tics. It prevails in 1 out of every 100 individuals worldwide. TS patients often have obsessive compulsive disorder (OCD) as well as attention deficit hyperactivity disorder (ADHD) along with some disorders of mood, sleep, depression and anxiety [35,36]. Many genes (such as NTN4, SLC6A4, IMMP2L, CNTNAP2, NLGN4, HDC and SLITRK1) and some chromosomal loci have been known to date to be involved in TS [37]. The dysfunction of serotonin and dopamine neurotransmitters and defects of cortico-striatal-thalamic-cortical pathways are considered to have an association with TS. Despite extensive research in genetics, the pathogenetic mechanism of TS is still largely lacking and the number of variants likely to cause TS is extremely small [38,39].
Here, in the current study, we performed a bioinformatical approach to predict the probably harmful nsSNPs and their possible consequences on the structure and function of SLITRK1 proteins. The total identified variants in SLITRK1 include 07 indels, 321 5′UTR and 624 3′UTR variants, and 05 nonsense variants. The analysis initially identified 442 missense variants out of 2255 total variants. Different pathogenicity prediction tools commonly predicted 16 variants to be presumably harmful for protein structure and function. The frequency of all 16 variants was very low. The secondary structure of these 16 variants did not show any change at the site of variation. A protein stability analysis is necessary to assess the structural and functional activity of a protein [40]. Protein stability governs the conformational structure of the protein and thus determines the function. Any alteration in protein stability may cause miss-folding, degradation or aberrant accumulation of proteins [41]. Out of 16 variants, 10 variants showed decreases in the protein stability based on the I-Mutant and MuPro tools. The conservation of amino acids at the points of substitutions was checked in nine different species including Homo sapiens using the Clustal Omega tool. The results showed that all the amino acids at the site of variations were highly conserved, indicating their structural and functional importance. The Missense 3D tool was used to check the possible loss in the 3D structure of the SLITRK1 protein, which is caused by a substitution of amino acids. The results showed that variants I, III and VIII affect the structural confirmation of SLITRK1 protein. Hence, variants I, III and VIII were selected for further in silico analysis. The wildtype SLITRK1 protein and its three variants i.e., Leu94Phe, Leu496Pro and Asn529Lys were docked with the close interactor protein PTPRD. Molecular docking analysis revealed that the aforementioned variants can possibly affect the functional activity of the SLITRK1 protein.
The limitation of current study is that the analysis was conducted without considering the disease model because a single gene may independently cause different diseases with different segregation patterns. Further to this, the alleles (obtained from genome browser for current analysis) exists in real form; however, until now, it was not associated with any disease onset due to several reasons, such as the allele being associated with some recessive condition where the presence of homozygous genotype would be necessary for disease onset. Nevertheless, the allele floats in the population, which may coincidently or due to extensive consanguinity, unite in a single individual (i.e., homozygous genotype) and cause the disorder. Therefore, in this study, we tried to focus on predictions of the detrimental effect of the allele regardless of disease inheritance pattern or disorder type. Hence, it is speculated that, if this allele segregates in an individual or family, it may have a negative impact. Moreover, the study also did not include the non-coding variants in spite of their significant role in the spatio-temporal gene expression pattern.
Conclusively, the present bioinformatic study would assist a molecular geneticist in interpreting the variant pathogenicity in research as well as diagnostic setup.

Author Contributions

Conceptualization, M.Z.A.; methodology, M.Z.A.; software, M.M. (Muhammad Muzammal), S.A. and M.A.K.; validation, M.M. (Mahpara Maryam), M.M. (Muhammad Muzammal) and M.A.K.; formal analysis, A.F.; investigation, M.A.A.H. and Y.N.A.; resources, A.J.A.; data curation, M.A.M.; writing—original draft preparation, A.F. and M.Z.A.; writing—review and editing, M.Z.A., M.A.A.H. and A.J.A.; visualization, M.A.A.H., E.M.A. and Y.N.A.; supervision, M.A.K.; project administration, A.F., A.J.A. and A.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding; however, MM and MZA are financially sponsored by a higher education commission of Pakistan through NRPU project (No. 5564/KPK/NRPU/R&D/HEC/2016) and indigenous scholarship, respectively.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The computational data are stored in the password-protected personal computers of M.M. and M.A.K., which is available to the editors upon request.

Acknowledgments

We thank the developers and curators of all the freely available online tools.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lantican, D.V.; Cortaga, C.Q.; Manohar, A.N.C.; dela Cueva, F.M.; Sison, M.L.J. Resistance gene analogs of mango: Insights on molecular defenses and evolutionary dynamics. Philipp. J. Sci. 2020, 149, 915–934. [Google Scholar]
  2. Hasan, N.; Choudhary, S.; Naaz, N.; Sharma, N.; Laskar, R.A. Recent advancements in molecular marker-assisted selection and applications in plant breeding programmes. J. Genet. Eng. Biotechnol. 2021, 19, 128. [Google Scholar] [CrossRef] [PubMed]
  3. Cortaga, C.Q.; Latina, R.A.; Habunal, R.R.; Lantican, D.V. Identification and characterization of genome-wide resistance gene analogs (RGAs) of durian (Durio zibethinus L.). J. Genet. Eng. Biotechnol. 2022, 20, 29. [Google Scholar] [CrossRef] [PubMed]
  4. Mullaney, J.M.; Mills, R.E.; Pittard, W.S.; Devine, S.E. Small insertions and deletions (INDELs) in human genomes. Hum. Mol. Genet. 2010, 19, R131–R136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Mills, R.E.; Luttig, C.T.; Larkins, C.E.; Beauchamp, A.; Tsui, C.; Pittard, W.S.; Devine, S.E. An initial map of insertion and deletion (INDEL) variation in the human genome. Genome Res. 2006, 16, 1182–1190. [Google Scholar] [CrossRef] [Green Version]
  6. Rozman, V.; Kunej, T. Harnessing omics big data in nine vertebrate species by genome-wide prioritization of sequence variants with the highest predicted deleterious effect on protein function. Omics A J. Integr. Biol. 2018, 22, 410–421. [Google Scholar] [CrossRef] [Green Version]
  7. Lander, E.S. The new genomics: Global views of biology. Science 1996, 274, 536–539. [Google Scholar] [CrossRef] [Green Version]
  8. Doniger, S.W.; Kim, H.S.; Swain, D.; Corcuera, D.; Williams, M.; Yang, S.P.; Fay, J.C. A catalog of neutral and deleterious polymorphism in yeast. PLoS Genet. 2008, 4, e1000183. [Google Scholar] [CrossRef] [Green Version]
  9. Radivojac, P.; Vacic, V.; Haynes, C.; Cocklin, R.R.; Mohan, A.; Heyen, J.W.; Goebl, M.G.; Iakoucheva, L.M. Identification, analysis, and prediction of protein ubiquitination sites. Proteins Struct. Funct. Bioinform. 2010, 78, 365–380. [Google Scholar] [CrossRef] [Green Version]
  10. Ashkenazy, H.; Abadi, S.; Martz, E.; Chay, O.; Mayrose, I.; Pupko, T.; Ben-Tal, N. ConSurf 2016: An improved methodology to estimate and visualize evolutionary conservation in macromolecules. Nucleic Acids Res. 2016, 44, W344–W350. [Google Scholar] [CrossRef] [Green Version]
  11. Franceschini, A.; Szklarczyk, D.; Frankild, S.; Kuhn, M.; Simonovic, M.; Roth, A.; Lin, J.; Minguez, P.; Bork, P.; Von Mering, C.; et al. STRING v9. 1: Protein-protein interaction networks, with increased coverage and integration. Nucleic Acids Res. 2012, 41, D808–D815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kozakov, D.; Hall, D.R.; Xia, B.; Porter, K.A.; Padhorny, D.; Yueh, C.; Beglov, D.; Vajda, S. The ClusPro web server for protein–protein docking. Nat. Protoc. 2017, 12, 255–278. [Google Scholar] [CrossRef] [PubMed]
  13. Jackson, M.; Marks, L.; May, G.H.; Wilson, J.B. The genetic basis of disease. Essays Biochem. 2018, 62, 643–723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Carninci, P.; Kasukawa, T.; Katayama, S.; Gough, J.; Frith, M.C.; Maeda, N.; Oyama, R.; Ravasi, T.; Lenhard, B.; Wells, C.; et al. The transcriptional landscape of the mammalian genome. Science 2005, 309, 1559–1563. [Google Scholar] [CrossRef] [Green Version]
  15. Liu, J.; Gough, J.; Rost, B. Distinguishing protein-coding from non-coding RNAs through support vector machines. PLoS Genet. 2006, 2, e29. [Google Scholar] [CrossRef] [Green Version]
  16. Ahonen, T.; Saltevo, J.; Laakso, M.; Kautiainen, H.; Kumpusa-lo, E.; Vanhala, M. Gender differences relating to metabolic syndrome and proinflammation in Finnish subjects with elevated blood pressure. Mediators Inflamm. 2009, 2009, 959281. [Google Scholar] [CrossRef] [Green Version]
  17. Degner, J.F.; Pai, A.A.; Pique-Regi, R.; Veyrieras, J.B.; Gaffney, D.J.; Pickrell, J.K.; De Leon, S.; Michelini, K.; Lewellen, N.; Crawford, G.E. DNase I sensitivity QTLs are a major determinant of human expression variation. Nature 2012, 482, 390–394. [Google Scholar] [CrossRef] [Green Version]
  18. Trynka, G.; Sandor, C.; Han, B.; Xu, H.; Stranger, B.E.; Liu, X.S.; Raychaudhuri, S. Chromatin marks identify critic-al cell types for fine mapping complex trait variants. Nat. Genet. 2013, 45, 124–130. [Google Scholar] [CrossRef]
  19. Kapoor, A.; Sekar, R.B.; Hansen, N.F.; Fox-Talbot, K.; Morley, M.; Pihur, V.; Chatterjee, S.; Brandimarto, J.; Moravec, C.S.; Pulit, S.L. An enhancer polymorphism at the car- diomyocyte intercalated disc protein NOS1AP locus is a major regulator of the QT interval. Am. J. Hum. Genet. 2014, 94, 854–869. [Google Scholar] [CrossRef] [Green Version]
  20. Pai, A.A.; Pritchard, J.K.; Gilad, Y. The genetic and mechanistic basis for variation in gene regulation. PLoS Genet. 2015, 11, e1004857. [Google Scholar] [CrossRef] [Green Version]
  21. Rhind, N.; Gilbert, D.M. DNA replication timing. Cold Spring Harb. Perspect. Biol. 2013, 5, a010132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Chakravarti, A. Single base differences between human genomes underlie differences in susceptibility to, or protection from, a host of diseases. Hence the great potential of such information in medicine. Nature 2001, 409, 822–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ng, P.C.; Henikoff, S. Predicting the effects of amino acid substitutions on protein function. Annu. Rev. Genom. Hum. Genet. 2006, 7, 61–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Gonzalez-Castejon, M.; Marin, F.; Soler-Rivas, C.; Reglero, G.; Visioli, F.; Rodriguez-Casado, A. Functional non-synonymous polymorphisms prediction methods: Current approaches and future developments. Curr. Med. Chem. 2011, 18, 5095–5103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Dobson, R.J.; Munroe, P.B.; Caulfield, M.J.; Saqi, M.A. Predicting deleterious nsSNPs: An analysis of sequence and structural attributes. BMC Bioinform. 2006, 7, 217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kumar, A.; Purohit, R. Computational screening and molecular dynamics simulation of disease associated nsSNPs in CENP-E. Mutat. Res. Fundam. Mol. Mech. Mutagenesis 2012, 738, 28–37. [Google Scholar] [CrossRef] [PubMed]
  27. Rajendran, V.; Sethumadhavan, R. Drug resistance mechanism of PncA in Mycobacterium tuberculosis. J. Biomol. Struct. Dyn. 2014, 32, 209–221. [Google Scholar] [CrossRef]
  28. Rajendran, V.; Gopalakrishnan, C.; Sethumadhavan, R. Pathological role of a point mutation (T315I) in BCR-ABL1 protein—A computational insight. J. Cell. Biochem. 2018, 119, 918–925. [Google Scholar] [CrossRef]
  29. Aruga, J.; Yokota, N.; Mikoshiba, K. Human SLITRK family genes: Genomic organization and expression profiling in normal brain and brain tumor tissue. Gene 2003, 315, 87–94. [Google Scholar] [CrossRef]
  30. Takahashi, H.; Katayama, K.I.; Sohya, K.; Miyamoto, H.; Prasad, T.; Matsumoto, Y.; Ota, M.; Yasuda, H.; Tsumoto, T.; Aruga, J.; et al. Selective control of inhibitory synapse development by Slitrk3-PTPδ trans-synaptic interaction. Nat. Neurosci. 2012, 15, 389–398. [Google Scholar] [CrossRef]
  31. Um, J.W.; Kim, K.H.; Park, B.S.; Choi, Y.; Kim, D.; Kim, C.Y.; Kim, S.J.; Kim, M.; Ko, J.S.; Lee, S.G.; et al. Structural basis for LAR-RPTP/Slitrk complex-mediated synaptic adhesion. Nat. Commun. 2014, 5, 5423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Han, K.A.; Jeon, S.; Um, J.W.; Ko, J. Emergent synapse organizers: LAR-RPTPs and their companions. Int. Rev. Cell Mol. Biol. 2016, 324, 39–65. [Google Scholar] [CrossRef] [PubMed]
  33. Nagase, T.; Nakayama, M.; Nakajima, D.; Kikuno, R.; Ohara, O. Prediction of the coding sequences of unidentified human genes. XX. The complete sequences of 100 new cDNA clones from brain which code for large proteins in vitro. DNA Res. 2001, 8, 85–95. [Google Scholar] [CrossRef] [Green Version]
  34. Abelson, J.F.; Kwan, K.Y.; O’Roak, B.J.; Baek, D.Y.; Stillman, A.A.; Morgan, T.M.; Mathews, C.A.; Pauls, D.L.; Rašin, M.R.; Gunel, M.; et al. Sequence variants in SLITRK1 are associated with Tourette’s syndrome. Science 2005, 310, 317–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hoekstra, P.J.; Dietrich, A.; Edwards, M.J.; Elamin, I.; Martino, D. Environmental factors in ourette syndrome. Neurosci. Biobehav. Rev. 2013, 37, 1040–1049. [Google Scholar] [CrossRef]
  36. Mathews, C.A.; Scharf, J.M.; Miller, L.L.; Macdonald-Wallis, C.; Lawlor, D.A.; Ben-Shlomo, Y. Association between pre-and perinatal exposures and Tourette syndrome or chronic tic disorder in the ALSPAC cohort. Br. J. Psychiatry 2014, 204, 40–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Kalanithi, P.S.; Zheng, W.; Kataoka, Y.; DiFiglia, M.; Grantz, H.; Saper, C.B.; Schwartz, M.L.; Leckman, J.F.; Vaccarino, F.M. Altered parvalbumin-positive neuron distribution in basal ganglia of individuals with Tourette syndrome. Proc. Natl. Acad. Sci. USA 2005, 102, 13307–13312. [Google Scholar] [CrossRef] [Green Version]
  38. Peterson, B.S.; Thomas, P.; Kane, M.J.; Scahill, L.; Zhang, H.; Bronen, R.; King, R.A.; Leckman, J.F.; Staib, L. Basal ganglia volumes in patients with Gilles de la Tourette syndrome. Arch. Gen. Psychiatry 2003, 60, 415–424. [Google Scholar] [CrossRef] [Green Version]
  39. Müller-Vahl, K.R.; Meyer, G.J.; Knapp, W.H.; Emrich, H.M.; Gielow, P.; Brücke, T.; Berding, G. Serotonin transporter binding in Tourette Syndrome. Neurosci. Lett. 2005, 385, 120–125. [Google Scholar] [CrossRef]
  40. Deller, M.C.; Kong, L.; Rupp, B. Protein stability: A crystallographer’s perspective. Acta Crystallogr. Sect. F Struct. Biol. Commun. 2016, 72, 72–95. [Google Scholar] [CrossRef] [Green Version]
  41. Witham, S.; Takano, K.; Schwartz, C.; Alexov, E. A missense mutation in CLIC2 associated with intellectual disability is predicted by in silico modeling to affect protein stability and dynamics. Proteins Struct. Funct. Bioinform. 2011, 79, 2444–2454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Classification of variants.
Figure 1. Classification of variants.
Genes 13 00672 g001
Figure 3. Amino acid conservation analysis using Clustal Omega. Note: Conserved regions harboring the SNP of interest are highlighted by the red box.
Figure 3. Amino acid conservation analysis using Clustal Omega. Note: Conserved regions harboring the SNP of interest are highlighted by the red box.
Genes 13 00672 g003
Figure 4. Evolutionary conservation analysis using ConSurf.
Figure 4. Evolutionary conservation analysis using ConSurf.
Genes 13 00672 g004
Figure 5. Three-dimensional structure of the SLITRK1 wild type protein, its variants, and the superimposed structures of the wild type and variants.
Figure 5. Three-dimensional structure of the SLITRK1 wild type protein, its variants, and the superimposed structures of the wild type and variants.
Genes 13 00672 g005
Figure 6. Prediction of protein–protein interaction using STRING.
Figure 6. Prediction of protein–protein interaction using STRING.
Genes 13 00672 g006
Figure 7. Protein–Protein docking of the SLITRK1 wild type and variants with their close interactor PTPRD.
Figure 7. Protein–Protein docking of the SLITRK1 wild type and variants with their close interactor PTPRD.
Genes 13 00672 g007
Table 1. List of variants commonly predicted to be pathogenic by all tools.
Table 1. List of variants commonly predicted to be pathogenic by all tools.
S.NoChr:bpAllelesAAAA CoordPolyphen2SNPs&GoMetaSNPProveanSIFTMutation AssessorPantherPHD SNPSNAP2PMut
PredProbPredProbPredScorePredScorePredScoreF IFI ScorePredPreservation TimePredScorePredScorePredScore
113:83879772G/APro/Leu579ProD0.997D0.736D0.657D−8.14D0M2.995ProD750D8E5D0.6827
213:83879921G/TAsn/Lys529ProD1D0.853D0.81D−5.77D0M3.065ProD456D8E82D0.7961
313:83879932G/ALeu/Phe526ProD0.997D0.673D0.603D−3.85D0M2.74ProD750D7E71D0.7852
413:83879997T/AAsn/Ile504ProD0.999D0.907D0.825D−8.14D0H4.565ProD750D8E85D0.8016
513:83880021A/GLeu/Pro496ProD1D0.834D0.786D−6.15D0H4.75ProD456D6E90D0.8058
613:83880024G/CSer/Trp495ProD0.985D0.616D0.72D−4.2D0.01H4.165ProD456D1E65D0.7634
713:83880354T/CAsn/Ser385ProD0.999D0.812D0.721D−4.37D0M2.995PosD361D8E72D0.5334
813:83880432C/TCys/Tyr359ProD0.999D0.922D0.816D−9.47D0M2.62PosD361D4E82D0.8359
913:83880435T/AAsn/Ile358PosD0.775D0.636D0.761D−6.24D0M2.3PosD361D3E58D0.7522
1013:83880469A/CCys/Gly347ProD1D0.757D0.733D−9.53D0M2.62ProD750D6E86D0.7549
1113:83880930G/APro/Leu193ProD0.949D0.767D0.676D−6.7D0M2.93ProD750D8E17D0.6558
1213:83881017T/CAsn/Ser164ProD1D0.833D0.766D−4.75D0M3.375ProD750D8E66D0.7106
1313:83881162T/AAsn/Tyr116ProD1D0.907D0.865D−7.59D0H4.72ProD750D7E83D0.7989
1413:83881226C/ALeu/Phe94ProD0.98D0.744D0.532D−2.96D0.01M3.18ProD456D7E46D0.8303
1513:83881276A/GPhe/Leu78PosD0.831D0.741D0.511D−4.8D0.02M2.035ProD750D8E52D0.5225
1613:83881305T/CAsn/Ser68ProD0.985D0.753D0.69D−4.72D0H3.555ProD750D1E64D0.8179
Key: Chr: Chromosome, bp: Base pair, AA: Amino Acid, AA coord: Amino Acid Coordinate, Pred: Prediction, Prob: Probability, FI: Functional Impact, Pro D: Probably damaging, Pos D: Possibly damaging, D: Damaging, M: Medium, H: High, E: Effect.
Table 2. List of variants showing decreases in the protein stability.
Table 2. List of variants showing decreases in the protein stability.
Variant No.rs IDAAAA CoordI-MutantMuPro
StabilityRIStabilityScore
Irs1048143268Asn/Lys529Decrease4Decrease−0.86
IIrs1219903976Leu/Phe526Decrease8Decrease−0.99
IIIrs1226852299Leu/Pro496Decrease6Decrease−0.992
IVrs1472728808Asn/Ser385Decrease5Decrease−0.971
Vrs1277399796Cys/Gly347Decrease7Decrease−0.994
VIrs1429907885Pro/Leu193Decrease6Decrease−0.778
VIIrs774612607Asn/Ser164Decrease5Decrease−0.999
VIIIrs1410244448Leu/Phe94Decrease8Decrease−0.999
IXrs954218528Phe/Leu78Decrease7Decrease−0.661
Xrs1228122404Asn/Ser68Decrease2Decrease−0.751
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ali, M.Z.; Farid, A.; Ahmad, S.; Muzammal, M.; Mohaini, M.A.; Alsalman, A.J.; Al Hawaj, M.A.; Alhashem, Y.N.; Alsaleh, A.A.; Almusalami, E.M.; et al. In Silico Analysis Identified Putative Pathogenic Missense nsSNPs in Human SLITRK1 Gene. Genes 2022, 13, 672. https://doi.org/10.3390/genes13040672

AMA Style

Ali MZ, Farid A, Ahmad S, Muzammal M, Mohaini MA, Alsalman AJ, Al Hawaj MA, Alhashem YN, Alsaleh AA, Almusalami EM, et al. In Silico Analysis Identified Putative Pathogenic Missense nsSNPs in Human SLITRK1 Gene. Genes. 2022; 13(4):672. https://doi.org/10.3390/genes13040672

Chicago/Turabian Style

Ali, Muhammad Zeeshan, Arshad Farid, Safeer Ahmad, Muhammad Muzammal, Mohammed Al Mohaini, Abdulkhaliq J. Alsalman, Maitham A. Al Hawaj, Yousef N. Alhashem, Abdulmonem A. Alsaleh, Eman M. Almusalami, and et al. 2022. "In Silico Analysis Identified Putative Pathogenic Missense nsSNPs in Human SLITRK1 Gene" Genes 13, no. 4: 672. https://doi.org/10.3390/genes13040672

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