Next Article in Journal
The Clinical Concept of LTDpathy: Is Dysregulated LTD Responsible for Prodromal Cerebellar Symptoms?
Previous Article in Journal
Interoceptive Attentiveness Induces Significantly More PFC Activation during a Synchronized Linguistic Task Compared to a Motor Task as Revealed by Functional Near-Infrared Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Stroke Genomics: Current Knowledge, Clinical Applications and Future Possibilities

1
Government Medical College, Kozhikode 673008, India
2
Baptist Health South Florida, Miami, FL 33176, USA
3
Department of Internal Medicine and Cardiology, Charité—Universitätsmedizin Berlin, Corporate Member of Freie Universität Berlin and Humboldt Universität zu Berlin, 10117 Berlin, Germany
*
Authors to whom correspondence should be addressed.
Brain Sci. 2022, 12(3), 302; https://doi.org/10.3390/brainsci12030302
Submission received: 28 January 2022 / Revised: 19 February 2022 / Accepted: 22 February 2022 / Published: 23 February 2022

Abstract

:
The pathophysiology of stoke involves many complex pathways and risk factors. Though there are several ongoing studies on stroke, treatment options are limited, and the prevalence of stroke is continuing to increase. Understanding the genomic variants and biological pathways associated with stroke could offer novel therapeutic alternatives in terms of drug targets and receptor modulations for newer treatment methods. It is challenging to identify individual causative mutations in a single gene because many alleles are responsible for minor effects. Therefore, multiple factorial analyses using single nucleotide polymorphisms (SNPs) could be used to gain new insight by identifying potential genetic risk factors. There are many studies, such as Genome-Wide Association Studies (GWAS) and Phenome-Wide Association Studies (PheWAS) which have identified numerous independent loci associated with stroke, which could be instrumental in developing newer drug targets and novel therapies. Additionally, using analytical techniques, such as meta-analysis and Mendelian randomization could help in evaluating stroke risk factors and determining treatment priorities. Combining SNPs into polygenic risk scores and lifestyle risk factors could detect stroke risk at a very young age and help in administering preventive interventions.

1. Introduction

Several risk factors and complex pathways are involved in the pathophysiology of stroke. Stroke is the second leading cause of death worldwide after heart attack [1]. The human genome project has helped in understanding many genetic factors that are associated with stroke [2,3,4]. Several studies have reported genetic predisposition to stroke in both human beings and animal models. However, the definition of genetic risk factors for stroke is not well established. Since no single specific gene has been responsible for stroke, it has been hypothesized to be a multifactorial polygenic disorder [5]. In this study, we used a narrative review method to understand the current advances, clinical applications, and future possibilities of the associations between genetic factors and stroke.

2. Genetic Factor Associated with Stroke (Non-Modifiable Factors in Stroke)

Several studies, such as the classical twin study which consisted of 15,924 twin pairs have been designed to assess the genetic factors associated with stroke [6]. Likewise, another twin study provided evidence for genetic factors that may increase the risk of stroke related events, such as death and hospitalization [7]. This study found greater concordance rates for these associations among monozygotic twins, compared to dizygotic twins [7]. These two studies were designed long before the human genome project. There could be different environmental effects affecting the results of these studies, which was a major limitation [8]. Previous studies have reported that first degree relatives are at an increased risk for stroke [9]. The preponderance of large and small vessel strokes, compared to cardioembolic strokes, is higher among subjects with a family history of stroke [9]. Sex is an important factor to influence stroke outcome indicating the possible role of the sex chromosome and associated genes; however, recently a review reported no association between sex and stroke [10]. Similarly, though ethnicity is not widely considered as an important factor affecting acute stroke outcome; it may influence the long-term outcome [10,11]. A recent study identified that levels of lipoprotein-A were significantly associated with adverse stroke outcomes, and were substantially higher in the Black, compared to the White population [12]. In addition, hematological disorders are responsible for nearly 1.3% of acute stroke. Some of the common hematological disorders associated with stroke include polycythemia vera, sickle-cell disease, Waldenström macroglobulinemia, multiple myeloma, essential thrombocythemia, thrombotic thrombocytopenic purpura, protein C deficiency, Protein S deficiency, antithrombin deficiency, and Factor V Leiden. A substantial number of these disorders have a genetic predisposition. For example, a large proportion of polycythemia vera patients have a mutation in the exon 14 of the JAK2 gene (JAK2V617F), whereas a smaller proportion has mutations in the JAK2 exon 12 [4].

2.1. Heritability Genes in Stroke (Monogenic and Polygenic Inheritance in Stroke Etiology)

Several animal model studies were conducted to identify potential candidate genes associated with stroke outcome. These studies analyzed the association of single nucleotide polymorphisms (SNPs) in targeted genes. The SNP of COX-2 and rs20417 genes were associated with early neurological deterioration [13,14]. However, these studies are not supported with further replicational studies and hence warrant further in-depth research. A study reported that several single-gene disorders might influence stroke, such as sickle cell disease, Fabry’s disease, homocystinuria, mitochondrial myopathy, and encephalopathy [15]. A rare stroke case caused by mutations in the Notch 3 gene (OMIM*600276) showed heritable patterns [16], which was also reported as a single-gene disorder. Cerebral Autosomal Dominant Arteriopathy with Subcortical Infarcts and Leukoencephalopathy (CADASIL) caused by different types of mutations of Notch 3 gene are associated with extensive cerebral small vessel damage, marked by the accumulation of granular osmiophilic material (GOM) [17]. Molecular evaluation of the vascular smooth muscles in CADASIL patients showed increased oxidation of soluble guanyl cyclase associated with decreased cyclic GMP levels, which impaired vasorelaxation of the cerebral vasculature [17]. A number of molecular pathways associated with cell adhesion, extracellular matrix components, misfolding control, autophagy, angiogenesis, and transforming growth factor β (TGFβ) signaling pathway are altered in CADASIL. Metabolic impairment, such as diabetes mellitus further expedites the pathological damage to the cerebral small blood vessels in Notch 3 mutation, resulting in endothelium mitochondrial dysfunction and vascular basement membrane injuries [18]. This suggests that the heritability of Notch 3 mutation increases the risk for ischemic stroke from small vessel diseases, such as CADASIL (Table 1).
Heterozygous mutations in the 3ʹuntranslated region (UTR) of the collagen 4A1 encoding gene may also influence ischemic stroke [19]. A glycine substitution mutation in the triple-helical domains of COL4A1 and COL4A2 may develop neurological and non-neurological manifestations, including hemorrhagic stroke [20]. The genomic data enables accurate analysis of heterozygous mutations. Another study identified heterozygous mutations in High-Temperature Requirement Serine protease A1 (HTRA1) encoding gene that manifest as stroke and cognitive decline in people aged more than 45 years [21]. Other mutations were also identified in the HTRA1 gene that may cause cerebral autosomal recessive arteriopathy in younger people who are between 10 to 30 years of age [22]. Similarly, mutations in adenosine deaminase 2 (ADA2), cathepsin A (CTSA) and forkhead-box C1 (FOXC1) genes were also found to be associated with autosomal dominant small vessel disease [23,24,25]. In addition, there are several other candidate genes under investigation for a possible association with stroke.

2.2. Multifactorial Stroke and SNPs

It is challenging to identify individual causative mutations in a single gene because many alleles are responsible for minor effects. Therefore, multiple factorial analyses using SNPs were used to gain newer insight by identifying potential genetic risk factors. For example, a study by Mola-Caminal et al. identified a locus located within a candidate gene [26], which can help in understanding the genetic mechanisms involved in stroke. Newer variants in the gene pals1-associated tight junction (PATJ) were linked to poor functional outcomes at 3-month post-stroke [26]. rs76221407 was the major SNP variant of the PATJ gene, which was associated with poor outcomes in stroke subjects after 3 months. The locus STRK1 was mapped to identify a susceptible gene for stroke for the first time [27]. Another study identified a strong association between the phosphodiesterase 4D gene (PDE4D; OMIM 600129*) and two major subtypes of stroke, cardiogenic and carotid stroke. Among 260 PDE4D gene SNPs, six were found to be significantly associated with stroke. Some of the SNPs were from UTR; therefore, these SNPs may affect the transcription of PDE4D [28]. The 5-lipoxygenase activating protein gene (ALOX5AP; OMIM 603700*) was also associated with an increased risk of stroke [29]. ALOX5AP SNP haplotypes increase the production of leukotriene B4 in stimulated neutrophils, thereby contributing to vascular inflammation in myocardial infarction and stroke [29]. The main limitation of studying candidate genes for SNPs and their association with stroke is that they are time consuming and require significant resources [30,31], and could be associated with false positive results.

3. Genomic Evaluation in Stroke

Several studies were designed during the 1990s to observe the effect of Mendelian genetics and candidate genes on stroke [32]. Subsequently, the human genome project enabled accurate SNP analysis by using the Genome-Wide Association Study (GWAS) [33].

3.1. Genome-Wide Association Study (GWAS) in Stroke

The first GWAS in stroke, Ischemic Stroke Genetics Study (ISGS) which included 250 patients and controls, was published in 2007 [34]. This study failed to identify any genetic locus, which was explicitly associated with stroke. Subsequently, studies focused on a specific region of chromosome 9 (9p21.3) and found an association with stroke [35]. This region was associated with coronary heart disease [36], and hence it was suggested that heart disease and ischemic stroke share similar polymorphisms. Another research group also studied chromosome 9 and found modest associations between ischemic stroke and variants (rs2383207 and rs10757274) of the 9p21 region [37]. Finally, six SNPs were identified, including rs2383207 in the 9p21 region, which were independently associated with the ischemic stroke (large artery atherosclerotic subtype) [38]. This suggests that chromosome 9p21 is an important risk locus that shares SNP variants that are common for both ischemic stroke and coronary artery disease.
A case-control study found a significant association between the 4q25 region and the cardioembolic subtype of ischemic stroke [39]. This region was also associated with all types of ischemic stroke, though to a lesser degree [39]. This study found that markers of atrial fibrillation, such as rs2200733 and rs10033464, have a strong association with ischemic stroke by increasing the risk for cardioembolic events. Another locus, the 16q22 was also found to be associated with cardioembolic stroke [40]. GWAS also found robust associations between intracranial aneurysms and loci on 2q, 8q, and 9p21 regions [41,42] The first prospective GWAS on stroke was the Heart and Aging Research in Genomic Epidemiology (CHARGE) study, which included 19,600 participants with 1544 strokes incidence [43]. This study identified two SNPs (rs11833579 and rs12425791) in the 12p13 region of chromosome 12 and within 11 kb upstream of the gene NINJ2 (Ninjurin 2), all of which were significantly associated with stroke.
The GWAS projects for ischemic stroke have identified many SNPs that are associated with stroke [44,45,46,47,48,49,50,51,52,53,54]. Among them, one study identified variants associated with different subtypes of stroke. This study showed that variants close to PITX2 (paired like homeodomain 2) and ZFHX3 (zinc finger homeobox 3) were linked to cardioembolic stroke. Variants on chromosome 9p21 locus and a novel variant on chromosome 7p21.1 within the histone deacetylase 9 (HDAC9) gene were associated with large vessel stroke [53]. This study suggested that genetic heterogeneity was associated with different stroke subtypes and would further demand subtype-specific studies for understanding genetic alterations in ischemic stroke.
Several GWAS consortia have been using and analyzing extensive datasets from major national and international projects. For example, SiGN project contains 14,549 cases from 24 genetic research centers located in the United States (n = 13) and Europe (n = 11) [55]. The MEGASTROKE consortium analyzed multi-ancestry GWAS data from more than 67,000 stroke cases and 454,000 controls and identified 32 significant loci to be associated with stroke [56]. Among them, two loci were independently associated with large artery stroke, and one with cardioembolic stroke. However, GWAS data has provided different associations between genes and stroke among different population and ethnic groups. For example, variants of the apelin receptor gene (APLNR, rs9943582) were associated with increased risk of ischemic stroke among the Japanese population, while these variants had no association with stroke among the Chinese Han Population [57]. Similarly, GWAS identified that rs2107595 SNP in the HDAC9 gene was associated with large-vessel ischemic stroke among the European population, while not among the Chinese Han population [58]. Another GWAS identified an SNP locus on region 10q25.3 of chromosome 10 (rs11196288) to be associated with the risk of early-onset ischemic stroke among the European population [59]. However, this SNP locus showed different susceptibility levels among the Chinese Han population. Similarly, in another study, there were differing associations between Caucasians and Chinese Han populations with respect to the relationship between SNPs of rs2200733 and rs6843082 on chromosome 4q25 and stroke [60]. These SNPs were associated with ischemic stroke among Caucasians, but not among the Chinese Han population. These varying results suggested that genetic factors are also modulated by other racial and ethnic factors and could provide unclear results. Therefore, it is recommended that GWAS data should be analyzed after population based sub-grouping.

3.2. GWAS and Comorbidities of Stroke

GWAS not only identified the genetic basis for stroke but also associations between genetic factors and comorbidities for stroke. The most important comorbidity associated with stroke is hypertension and is responsible for 30–40% of population-attributable risk for stroke [61]. Other comorbidities associated with stroke are smoking, diabetes, atrial fibrillation, and coronary heart disease. Several loci have been identified which have an association with stroke and its comorbidities. For example, 4q25 region of chromosome 4 [39] and a variant of the ZFHX3 gene in the 16q22 region of chromosome 16 [40] were associated with ischemic stroke and atrial fibrillation, 9p21 region was associated with stroke and diabetes [62], and serine/threonine kinase gene (STK39) variants were associated with stroke and hypertension [63]. The MEGASTROKE consortium identified a total of 32 loci that were significantly associated with stroke. Among them, five were associated with blood pressure, five with coronary heart disease, two with low-density lipoprotein (LDL) cholesterol, two with atrial fibrillation, two with venous thromboembolism, one with white matter hyperintensities, and one with carotid plaque [56]. This study identified a strong association between coronary heart disease and large artery stroke as well as blood pressure and all stroke subtypes. This study also found that cardioembolic stroke and large artery stroke, though not small vessel stroke, were associated with venous thromboembolism. However, an interesting finding from this study was that high-density lipoprotein (HDL) cholesterol was inversely associated with small vessel stroke [56]. Another study reported that the rs4376531 variant among diabetes predicted the risk for atherothrombotic stroke [64]. Another GWAS study found that C-reactive protein gene polymorphisms increased its synthesis level, which in turn increased the risk for stroke [65].

3.3. Genomic Determinants of Stroke Outcomes

Researchers started using data from GWAS for identifying genetic determinants of stroke outcomes only recently [2]. A genome wide meta-analysis (GWMA) of 12 stroke cohorts identified that the Pals1-associated tight junction (PATJ) variant was significantly associated with adverse functional outcomes after three months of stroke [26]. However, the molecular mechanism of how the variants of the PATJ gene led to these outcomes is still unclear. Mola-Caminal et al. reported that the major variant rs76221407 in the PATJ gene was a key genotypic trait associated with poor functional outcomes after three months of stroke onset [26]. Another GWMA study identified the SNP rs184681 to be significantly associated with functional outcomes of neural plasticity between 60 and 190 days after stroke onset [66]. In another study, the genetic imbalance was associated with unfavorable outcomes after 2–6 months of stroke, after adjusting for age, sex, race, and stroke subtypes [67].

4. Preclinical Studies Supporting Genomic Analysis in Stroke

It is challenging to study preclinical stroke models, although several animal studies have been designed to overcome the effects of host genetic variations. However, even after genetic background restriction, studies using animal models have been questioned for assessing complex polygenetic disorders, such as stroke [68]. Nevertheless, animal models are still valuable to study basic mechanisms and factors, such as environmental and dietary factors related to stroke. In addition, animal studies have been used for identifying potential targets involved in inflammatory signaling of stroke outcomes among humans [69].
Studies have shown that Metastasis-Associated Lung Adenocarcinoma Transcript 1 (MALAT1) expression could be induced in vitro in endothelial cells undergoing oxygen-glucose deprivation (OGD) [70,71]. Transcriptional downregulation of MALAT1 in OGD induced primary mouse brain microvascular endothelial cells led to overexpression of pro-apoptotic factor bim and increased pro-inflammatory cytokines, such as MCP-1, IL-6, and E-selectin [72]. Moreover, in vivo MALAT1 knockout mice showed severe neurological deficits, compared to wild-type controls in response to transient focal ischemia [72]. Additionally, other studies have demonstrated that MALAT1 promotes endothelial cell survival, angiogenesis, and vascular integrity in stroke [73,74,75,76,77]. MALAT1 plays a crucial role in regulating post-stroke pathophysiology; however, further studies are required to understand the contexts and conditions under which MALAT1 mediates beneficial versus deleterious outcomes.
Upregulated maternally expressed gene (MEG3) in the mouse brain and primary neurons were linked to increased cell death in cerebral ischemia [78,79,80]. Long non-coding RNA (lncRNA) MEG3 functions as a competing endogenous RNA (ceRNA) and binds to miR-21 and downregulates the miR-21/PDAC pathway, leading to neuronal death in ischemic neurons [80]. miR-21 overexpression reverses the effect of OGD reperfusion induced neuronal apoptosis in vitro. Another investigation showed that downregulation of MEG3 was associated with increased micro-vessel density in rat neurons [81]. Therefore, MEG3 exhibits differential expression after stroke among different species and cell types while downregulation of MEG3 is strongly associated with post-stroke neuroprotection.
The Small Nucleolar RNA Host Gene 12 (SNHG12) expression after ischemic injury was increased both in vitro and in vivo [82,83,84]. The N2a cell line and mouse primary hippocampal neuronal study has shown higher expressions of lncRNA SNHG12 among neuronal cells undergoing ischemia [83], while downregulation of miR-199a by SNHG12 decreases cell death and inflammation [82]. lncRNA SNHG12 improves neuronal survival following OGD reperfusion induced ischemia through miR-199a downregulation by sirtuin-1 upregulation and activation of adenosine 5′ monophosphate-activated protein kinase (AMPK) pathway [83]. This suggests that increased expression of lncRNA SNHG12 salvages injured ischemic neurons. During transient ischemia, circulating H19 levels become higher in blood and brain among stroke patients. Experimental mouse model studies have shown that knockdown of H19 could decrease edema, infarct volume, and neurological deficits after stroke [85,86]. Though several studies have looked for the role of lncRNA in modulating the post-stroke pathophysiology, only a few have explored the genetic variations of lncRNA and altered expression among stroke patients [85,87,88,89,90,91,92]. Overall, these studies introduce the possibility that the evaluation of lncRNA expression or lncRNA gene loci could be a useful clinical tool for assessing the risk for developing stroke.

5. Extending Genome-Based Evaluation into the Clinical Scenario

The global prevalence of stroke is consistently increasing, and there are limited therapeutic interventions. Therefore, developing advanced treatment strategies to manage stroke and post-stroke brain damage is important. Several studies have already completed preliminary research to integrate genetic data into routine clinical practice and precision medicine. Extending genome-based studies could help in developing therapeutic and predictability capabilities in managing the early stages of stroke.

5.1. Stroke Risk Prediction in Childhood

Several studies were designed for the identification of stroke by gene expression profiling. Studies have predicted ischemic stroke with as high as 80% accuracy through analysis of a panel of 22 genes from peripheral blood mononuclear cells (PBMC) [93,94]. Using the latest technology and developments from genetic data, high-risk individuals could be identified by applying polygenic risk scores for common genetic variants, even during childhood [95,96]. These methods can enable the opportunities for early prevention of stroke. Recently, a study developed a polygenic risk score derived from a panel of 90 SNPs to identify individuals with 35% increased risk for stroke [97]. Risk scores for stroke based on lifestyle factors, such as smoking, diet, body mass index (BMI), and physical activity have shown that lifestyle risks were similar across all polygenic risk score strata.
Recently, studies have applied Mendelian randomization to identify risk for stroke [98]. For example, a mendelian randomization study identified factors, such as BMI and waist-to-hip ratio, in order to identify individuals with greater risk for ischemic stroke [99]. The differential effects of LDL and HDL on cardioembolic stroke, small vessel stroke, and large artery stroke observed in the MEGASTROKE consortium study were also confirmed by a Mendelian randomization study [100]. Similarly, differential effects of type 2 diabetes were observed for different etiological stroke subtypes by the Mendelian randomization studies [101,102]. Mendelian randomization studies could be further applied for identifying novel risk factors for stroke. With the increasing availability of genomic data, Mendelian randomization studies will become more relevant and applicable in clinical practice. Although additional research is required to evaluate and improve genetic risk prediction of stroke, these studies highlight the potential for early risk stratification and prevention of stroke via genetic evaluation.

5.2. Exploration of Potential Therapeutics

Currently, the pharmacological treatment of stroke is mainly based on recombinant tissue plasminogen activator (rtPA). rtPA was developed based on genetic data. Nevertheless, pharmacological treatment strategies for stroke have not significantly progressed over the years. The FDA first approved RNA-targeting antisense oligos therapy for spinal muscular atrophy in 2017. After that, in 2018, FDA approved the first RNAi therapy as a treatment option for peripheral nerve disease. Increasing transthyretin in tissues is the main reason for this disease and is caused by hereditary transthyretin-mediated amyloidosis [103]. The exploration of this genetic target could significantly expand the pharmacological applications of stroke treatment in the future.

5.3. Exploiting Genetics for Potential Drug Discovery

Genomic data offers a great potential for drug development of stroke by identifying causal pathways and drug targets and could determine the safety and efficacy of pharmacological interventions [104,105,106]. These approaches were developed to personalize the dose and minimize the side effects. Mendelian randomization studies [107] and other studies have shown the use of protective variants [108,109,110] and have demonstrated naturally occurring human knockouts for phenotypic effects on stroke outcomes [111]. Currently, phenome-wide association studies (PheWASs) show promise and have analyzed large datasets with detailed genotyping and phenotyping data with multiple traits [112,113]. Therefore, using genetic and phenetic data for potential drug discovery and precision medicine is now an advancing and emerging major research focus.

6. Future Directives

Several genetic and genomic factors have already been identified for stroke and some overlap with comorbidities as previously described (Table 1). Many of these studies, such as those associated with vascular risk, monogenic vasculopathies, the leukotriene pathway, and other GWAS require further detailed investigation. Although vasculopathy of CADASIL was not associated with heart diseases, higher rates of myocardial infarction [114] and unexplained sudden deaths [115] among these patients require additional investigations.
Currently, there are several ongoing GWAS and PheWAS with large sample sizes for identifying newer and undiscovered loci associated with stroke [56]. Biobanks and databases will enormously expand the opportunity for gene discovery [116], and thereby, accelerate the progress in this field. However, data from non-European ancestry is inadequate. Therefore, for studies to be effective across all populations, ancestry-specific genetic data should be developed. The development of prospective drugs has now become much easier after GWAS and PheWAS for several common diseases [104]. However, further, improvement is required to develop novel cell and tissue models to study functional genomics and multilevel omics of stroke. Nevertheless, genetic studies focusing on treatment and recovery after stroke are in their infancy and require much more details for clinical applications [26,117]. Strokes could lead to significant cognitive decline and vascular dementia because of cerebral small vessel diseases. However, there is very little data from studies estimating the heritability of cerebral small vessel diseases. A growing body of evidence from epidemiological and genetic studies suggests that early cerebral small vessel diseases are heritable. It is, therefore, imperative that future studies should address the genetic factors associated with cerebral small vessel disease, as well as the potential clinical outcomes, to assess the genomics of vascular cognitive decline.

7. Limitations

Though we have reviewed several genetic factors associated with stroke, several others that have not been covered in this review require additional exploration. We have primarily focused on genetic factors that are adversely associated with stroke. There are some protective genetic factors as well which need further exploration. Though we have explored genetic factors associated with stroke, there are epigenetic factors that need additional evaluation. In addition, we could not explore in detail how genetic factors could be incorporated in precision medicine and how genetic data could be integrated with other omics data, such as proteomic, metabolomic, and transcriptomic data since they are beyond the scope of this review.

8. Conclusions

The prevalence of stroke and the global burden remains high. Therefore, discovering genetic variants and biological pathways offer has revived hopes for novel therapeutics, drug targets, and effective interventions. Genetic information can be used to improve stroke diagnosis and prognosis. Several GWAS and PheWAS have identified many independent loci associated with stroke which could be instrumental in developing newer drug targets and novel therapies. The application of analytical techniques, such as meta-analysis and Mendelian randomization could also facilitate evaluating risk factors and stroke outcomes and prioritizing potential therapeutic targets. Accumulating SNPs into polygenic risk scores and combining them with lifestyle risk factor scores could enable the possibility of identifying individuals who are at a greater risk for stroke even at a younger age.

Author Contributions

Conceptualization, M.R. and S.A.; methodology, M.R.; validation, P.M. and A.S.; investigation, P.M.; resources, E.V.; writing—original draft preparation, M.R.; writing—review and editing, M.R., S.A., V.R., P.M., and E.V.; supervision, M.R and A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization. The Top 10 Causes of Death; World Health Organization: Geneva, Switzerland, 2020. [Google Scholar]
  2. Meschia, J.F. Effects of Genetic Variants on Stroke Risk. Stroke 2020, 51, 736–741. [Google Scholar] [CrossRef] [PubMed]
  3. Devaraddi, N.; Jayalakshmi, G.; Mutalik, N.R. CARASIL, a rare genetic cause of stroke in the young. Neurol. India 2018, 66, 232–234. [Google Scholar] [CrossRef]
  4. Chauhan, G.; Debette, S. Genetic Risk Factors for Ischemic and Hemorrhagic Stroke. Curr. Cardiol. Rep. 2016, 18, 124. [Google Scholar] [CrossRef] [Green Version]
  5. Orlacchio, A.; Bernardi, G. Research actuality in the genetics of stroke. Clin. Exp. Hypertens. 2006, 28, 191–197. [Google Scholar] [CrossRef] [PubMed]
  6. Braun, M.; Haupt, R.; Caporaso, N. The National Academy of Sciences–National Research Council Veteran Twin Registry. Acta Genet. Med. Gemellol. 1994, 43, 89–94. [Google Scholar] [CrossRef] [PubMed]
  7. Bak, S.; Gaist, D.; Sindrup, S.H.; Skytthe, A.; Christensen, K. Genetic liability in stroke: A long-term follow-up study of Danish twins. Stroke 2002, 33, 769–774. [Google Scholar] [CrossRef] [Green Version]
  8. Kendler, K.; Holm, N. Differential enrollment in twin registries: Its effect on prevalence and concordance rates and estimates of genetic parameters. Acta Genet. Med. Gemellol. 1985, 34, 125–140. [Google Scholar] [CrossRef] [PubMed]
  9. Floßmann, E.; Schulz, U.G.; Rothwell, P.M. Systematic review of methods and results of studies of the genetic epidemiology of ischemic stroke. Stroke 2004, 35, 212–227. [Google Scholar] [CrossRef] [Green Version]
  10. Torres-Aguila, N.P.; Carrera, C.; Muiño, E.; Cullell, N.; Cárcel-Márquez, J.; Gallego-Fabrega, C.; González-Sánchez, J.; Bustamante, A.; Delgado, P.; Ibañez, L.; et al. Clinical Variables and Genetic Risk Factors Associated with the Acute Outcome of Ischemic Stroke: A Systematic Review. J. Stroke 2019, 21, 276–289. [Google Scholar] [CrossRef] [Green Version]
  11. Boehme, A.K.; Siegler, J.E.; Mullen, M.T.; Albright, K.C.; Lyerly, M.J.; Monlezun, D.J.; Jones, E.M.; Tanner, R.; Gonzales, N.R.; Beasley, T.M. Racial and gender differences in stroke severity, outcomes, and treatment in patients with acute ischemic stroke. J. Stroke Cerebrovasc. Dis. 2014, 23, e255–e261. [Google Scholar] [CrossRef] [Green Version]
  12. Kamin Mukaz, D.; Zakai, N.A.; Cruz-Flores, S.; McCullough, L.D.; Cushman, M. Identifying Genetic and Biological Determinants of Race-Ethnic Disparities in Stroke in the United States. Stroke 2020, 51, 3417–3424. [Google Scholar] [CrossRef] [PubMed]
  13. Yi, X.; Ming, B.; Wang, C.; Chen, H.; Ma, C. Variants in COX-2, PTGIS, and TBXAS1 are associated with carotid artery or intracranial arterial stenosis and neurologic deterioration in ischemic stroke patients. J. Stroke Cerebrovasc. Dis. 2017, 26, 1128–1135. [Google Scholar] [CrossRef] [PubMed]
  14. Yi, X.; Wang, C.; Zhou, Q.; Lin, J. Interaction among COX-2, P2Y1 and GPIIIa gene variants is associated with aspirin resistance and early neurological deterioration in Chinese stroke patients. BMC Neurol. 2017, 17, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Munshi, A.; Kaul, S. Genetic basis of stroke: An overview. Neurol. India 2010, 58, 185. [Google Scholar] [CrossRef] [PubMed]
  16. Joutel, A.; Corpechot, C.; Ducros, A.; Vahedi, K.; Chabriat, H.; Mouton, P.; Alamowitch, S.; Domenga, V.; Cécillion, M.; Maréchal, E. Notch3 mutations in CADASIL, a hereditary adult-onset condition causing stroke and dementia. Nature 1996, 383, 707–710. [Google Scholar] [CrossRef]
  17. Neves, K.B.; Morris, H.E.; Alves-Lopes, R.; Muir, K.W.; Moreton, F.; Delles, C.; Montezano, A.C.; Touyz, R.M. Peripheral arteriopathy caused by Notch3 gain-of-function mutation involves ER and oxidative stress and blunting of NO/sGC/cGMP pathway. Clin. Sci. 2021, 135, 753–773. [Google Scholar] [CrossRef]
  18. Felczak, P.; Cudna, A.; Błażejewska-Hyżorek, B.; Buczek, J.; Kurkowska-Jastrzębska, I.; Stępień, T.; Acewicz, A.; Wierzba-Bobrowicz, T. Ultrastructure of mitochondria and damage to small blood vessels in siblings with the same mutation in the NOTCH 3 and coexisting diseases. Pol. J. Pathol. 2021, 72, 148–159. [Google Scholar]
  19. Verdura, E.; Hervé, D.; Bergametti, F.; Jacquet, C.; Morvan, T.; Prieto-Morin, C.; Mackowiak, A.; Manchon, E.; Hosseini, H.; Cordonnier, C. Disruption of a mi R-29 binding site leading to COL4A1 upregulation causes pontine autosomal dominant microangiopathy with leukoencephalopathy. Ann. Neurol. 2016, 80, 741–753. [Google Scholar] [CrossRef]
  20. Jeanne, M.; Gould, D.B. Genotype-phenotype correlations in pathology caused by collagen type IV alpha 1 and 2 mutations. Matrix Biol. 2017, 57, 29–44. [Google Scholar] [CrossRef] [Green Version]
  21. Verdura, E.; Herve, D.; Scharrer, E.; Amador, M.d.M.; Guyant-Marechal, L.; Philippi, A.; Corlobe, A.; Bergametti, F.; Gazal, S.; Prieto-Morin, C. Heterozygous HTRA1 mutations are associated with autosomal dominant cerebral small vessel disease. Brain 2015, 138, 2347–2358. [Google Scholar] [CrossRef] [Green Version]
  22. Beaufort, N.; Scharrer, E.; Kremmer, E.; Lux, V.; Ehrmann, M.; Huber, R.; Houlden, H.; Werring, D.; Haffner, C.; Dichgans, M. Cerebral small vessel disease-related protease HtrA1 processes latent TGF-β binding protein 1 and facilitates TGF-β signaling. Proc. Natl. Acad. Sci. USA 2014, 111, 16496–16501. [Google Scholar] [CrossRef] [Green Version]
  23. Zhou, Q.; Yang, D.; Ombrello, A.K.; Zavialov, A.V.; Toro, C.; Zavialov, A.V.; Stone, D.L.; Chae, J.J.; Rosenzweig, S.D.; Bishop, K. Early-onset stroke and vasculopathy associated with mutations in ADA2. N. Engl. J. Med. 2014, 370, 911–920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bugiani, M.; Kevelam, S.H.; Bakels, H.S.; Waisfisz, Q.; Ceuterick-de Groote, C.; Niessen, H.W.; Abbink, T.E.; Oberstein, S.A.L.; van der Knaap, M.S. Cathepsin A–related arteriopathy with strokes and leukoencephalopathy (CARASAL). Neurology 2016, 87, 1777–1786. [Google Scholar] [CrossRef] [PubMed]
  25. French, C.R.; Seshadri, S.; Destefano, A.L.; Fornage, M.; Arnold, C.R.; Gage, P.J.; Skarie, J.M.; Dobyns, W.B.; Millen, K.J.; Liu, T. Mutation of FOXC1 and PITX2 induces cerebral small-vessel disease. J. Clin. Investig. 2014, 124, 4877–4881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Mola-Caminal, M.; Carrera, C.; Soriano-Tárraga, C.; Giralt-Steinhauer, E.; Díaz-Navarro, R.M.; Tur, S.; Jiménez, C.; Medina-Dols, A.; Cullell, N.; Torres-Aguila, N.P. PATJ low frequency variants are associated with worse ischemic stroke functional outcome: A genome-wide meta-analysis. Circ. Res. 2019, 124, 114–120. [Google Scholar] [CrossRef] [PubMed]
  27. Stover, V.; Stern, B.J.; Sherman, S. Analytic strategies for stroke genetics. J. Stroke Cerebrovasc. Dis. 2002, 11, 272–278. [Google Scholar] [CrossRef]
  28. Gretarsdottir, S.; Thorleifsson, G.; Reynisdottir, S.T.; Manolescu, A.; Jonsdottir, S.; Jonsdottir, T.; Gudmundsdottir, T.; Bjarnadottir, S.M.; Einarsson, O.B.; Gudjonsdottir, H.M. The gene encoding phosphodiesterase 4D confers risk of ischemic stroke. Nat. Genet. 2003, 35, 131–138. [Google Scholar] [CrossRef]
  29. Helgadottir, A.; Manolescu, A.; Thorleifsson, G.; Gretarsdottir, S.; Jonsdottir, H.; Thorsteinsdottir, U.; Samani, N.J.; Gudmundsson, G.; Grant, S.F.; Thorgeirsson, G. The gene encoding 5-lipoxygenase activating protein confers risk of myocardial infarction and stroke. Nat. Genet. 2004, 36, 233–239. [Google Scholar] [CrossRef]
  30. Rosand, J.; Bayley, N.; Rost, N.; de Bakker, P.I. Many hypotheses but no replication for the association between PDE4D and stroke. Nat. Genet. 2006, 38, 1091–1092. [Google Scholar] [CrossRef]
  31. Matarin, M.; Brown, W.M.; Dena, H.; Britton, A.; De Vrieze, F.W.; Brott, T.G.; Brown, R.D., Jr.; Worrall, B.B.; Case, L.D.; Chanock, S.J. Candidate gene polymorphisms for ischemic stroke. Stroke 2009, 40, 3436–3442. [Google Scholar] [CrossRef] [Green Version]
  32. Hunter, D.J.; Altshuler, D.; Rader, D.J. From Darwin’s finches to canaries in the coal mine—Mining the genome for new biology. N. Engl. J. Med. 2008, 358, 2760–2763. [Google Scholar] [CrossRef] [PubMed]
  33. Hardy, J.; Singleton, A. Genomewide association studies and human disease. N. Engl. J. Med. 2009, 360, 1759–1768. [Google Scholar] [CrossRef] [Green Version]
  34. Matarín, M.; Brown, W.M.; Scholz, S.; Simón-Sánchez, J.; Fung, H.-C.; Hernandez, D.; Gibbs, J.R.; De Vrieze, F.W.; Crews, C.; Britton, A. A genome-wide genotyping study in patients with ischaemic stroke: Initial analysis and data release. Lancet Neurol. 2007, 6, 414–420. [Google Scholar] [CrossRef] [Green Version]
  35. Matarin, M.; Brown, W.M.; Singleton, A.; Hardy, J.A.; Meschia, J.F. Whole genome analyses suggest ischemic stroke and heart disease share an association with polymorphisms on chromosome 9p21. Stroke 2008, 39, 1586–1589. [Google Scholar] [CrossRef] [PubMed]
  36. McPherson, R.; Pertsemlidis, A.; Kavaslar, N.; Stewart, A.; Roberts, R.; Cox, D.R.; Hinds, D.A.; Pennacchio, L.A.; Tybjaerg-Hansen, A.; Folsom, A.R. A common allele on chromosome 9 associated with coronary heart disease. Science 2007, 316, 1488–1491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Smith, J.G.; Melander, O.; Lövkvist, H.K.; Hedblad, B.; Engström, G.; Nilsson, P.; Carlson, J.; Berglund, G.R.; Norrving, B.; Lindgren, A. Common genetic variants on chromosome 9p21 confers risk of ischemic stroke: A large-scale genetic association study. Circ. Cardiovasc. Genet. 2009, 2, 159–164. [Google Scholar] [CrossRef] [Green Version]
  38. Gschwendtner, A.; Bevan, S.; Cole, J.W.; Plourde, A.; Matarin, M.; Ross-Adams, H.; Meitinger, T.; Wichmann, E.; Mitchell, B.D.; Furie, K. Sequence variants on chromosome 9p21. 3 confer risk for atherosclerotic stroke. Ann. Neurol. 2009, 65, 531–539. [Google Scholar] [CrossRef] [Green Version]
  39. Gretarsdottir, S.; Thorleifsson, G.; Manolescu, A.; Styrkarsdottir, U.; Helgadottir, A.; Gschwendtner, A.; Kostulas, K.; Kuhlenbäumer, G.; Bevan, S.; Jonsdottir, T. Risk variants for atrial fibrillation on chromosome 4q25 associate with ischemic stroke. Ann. Neurol. 2008, 64, 402–409. [Google Scholar] [CrossRef]
  40. Gudbjartsson, D.F.; Holm, H.; Gretarsdottir, S.; Thorleifsson, G.; Walters, G.B.; Thorgeirsson, G.; Gulcher, J.; Mathiesen, E.B.; Njølstad, I.; Nyrnes, A. A sequence variant in ZFHX3 on 16q22 associates with atrial fibrillation and ischemic stroke. Nat. Genet. 2009, 41, 876–878. [Google Scholar] [CrossRef]
  41. Helgadottir, A.; Thorleifsson, G.; Magnusson, K.P.; Grétarsdottir, S.; Steinthorsdottir, V.; Manolescu, A.; Jones, G.T.; Rinkel, G.J.; Blankensteijn, J.D.; Ronkainen, A. The same sequence variant on 9p21 associates with myocardial infarction, abdominal aortic aneurysm and intracranial aneurysm. Nat. Genet. 2008, 40, 217. [Google Scholar] [CrossRef]
  42. Bilguvar, K.; Yasuno, K.; Niemelä, M.; Ruigrok, Y.M.; Von Und Zu Fraunberg, M.; Van Duijn, C.M.; Van Den Berg, L.H.; Mane, S.; Mason, C.E.; Choi, M. Susceptibility loci for intracranial aneurysm in European and Japanese populations. Nat. Genet. 2008, 40, 1472–1477. [Google Scholar] [CrossRef]
  43. Ikram, M.A.; Seshadri, S.; Bis, J.C.; Fornage, M.; DeStefano, A.L.; Aulchenko, Y.S.; Debette, S.; Lumley, T.; Folsom, A.R.; Van Den Herik, E.G. Genomewide association studies of stroke. N. Engl. J. Med. 2009, 360, 1718–1728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lupski, J.R.; Belmont, J.W.; Boerwinkle, E.; Gibbs, R.A. Clan genomics and the complex architecture of human disease. Cell 2011, 147, 32–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Bevan, S.; Traylor, M.; Adib-Samii, P.; Malik, R.; Paul, N.L.; Jackson, C.; Farrall, M.; Rothwell, P.M.; Sudlow, C.; Dichgans, M. Genetic heritability of ischemic stroke and the contribution of previously reported candidate gene and genomewide associations. Stroke 2012, 43, 3161–3167. [Google Scholar] [CrossRef] [PubMed]
  46. Cepeda, M.R.M.; Sierra, L.E. Genetic heritability of ischemic vascular disease subtypes. Rev. Mex. Neurocienc. 2013, 14, 61–62. [Google Scholar]
  47. Foo, J.-N.; Liu, J.-J.; Tan, E.-K. Whole-genome and whole-exome sequencing in neurological diseases. Nat. Rev. Neurol. 2012, 8, 508–517. [Google Scholar] [CrossRef] [PubMed]
  48. Hacke, W.; Grond-Ginsbach, C. Commentary on a GWAS: HDAC9 and the risk for ischaemic stroke. BMC Med. 2012, 10, 70. [Google Scholar] [CrossRef] [Green Version]
  49. Yamada, Y.; Fuku, N.; Tanaka, M.; Aoyagi, Y.; Sawabe, M.; Metoki, N.; Yoshida, H.; Satoh, K.; Kato, K.; Watanabe, S. Identification of CELSR1 as a susceptibility gene for ischemic stroke in Japanese individuals by a genome-wide association study. Atherosclerosis 2009, 207, 144–149. [Google Scholar] [CrossRef]
  50. Meschia, J.F.; Singleton, A.; Nalls, M.A.; Rich, S.S.; Sharma, P.; Ferrucci, L.; Matarin, M.; Hernandez, D.G.; Pearce, K.; Brott, T.G. Genomic risk profiling of ischemic stroke: Results of an international genome-wide association meta-analysis. PLoS ONE 2011, 6, e23161. [Google Scholar]
  51. Sun, H.; Wu, H.; Zhang, J.; Wang, J.; Lu, Y.; Ding, H.; Xiao, H.; Zhang, J. A tagging SNP in ALOX5AP and risk of stroke: A haplotype-based analysis among eastern Chinese Han population. Mol. Biol. Rep. 2011, 38, 4731–4738. [Google Scholar] [CrossRef]
  52. Holliday, E.G.; Maguire, J.M.; Evans, T.-J.; Koblar, S.A.; Jannes, J.; Sturm, J.W.; Hankey, G.J.; Baker, R.; Golledge, J.; Parsons, M.W. Common variants at 6p21. 1 are associated with large artery atherosclerotic stroke. Nat. Genet. 2012, 44, 1147–1151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Bellenguez, C.; Bevan, S.; Gschwendtner, A.; Spencer, C.C.; Burgess, A.I.; Pirinen, M.; Jackson, C.A.; Traylor, M.; Strange, A.; Su, Z. Genome-wide association study identifies a variant in HDAC9 associated with large vessel ischemic stroke. Nat. Genet. 2012, 44, 328–333. [Google Scholar] [CrossRef] [PubMed]
  54. Arregui, M.; Fisher, E.; Knüppel, S.; Buijsse, B.; di Giuseppe, R.; Fritsche, A.; Corella, D.; Willich, S.N.; Boeing, H.; Weikert, C. Significant associations of the rs2943634 (2q36. 3) genetic polymorphism with adiponectin, high density lipoprotein cholesterol and ischemic stroke. Gene 2012, 494, 190–195. [Google Scholar] [CrossRef] [PubMed]
  55. Meschia, J.F.; Arnett, D.K.; Ay, H.; Brown, R.D., Jr.; Benavente, O.R.; Cole, J.W.; De Bakker, P.I.; Dichgans, M.; Doheny, K.F.; Fornage, M. Stroke Genetics Network (SiGN) study: Design and rationale for a genome-wide association study of ischemic stroke subtypes. Stroke 2013, 44, 2694–2702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Malik, R.; Chauhan, G.; Traylor, M.; Sargurupremraj, M.; Okada, Y.; Mishra, A.; Rutten-Jacobs, L.; Giese, A.-K.; Van Der Laan, S.W.; Gretarsdottir, S. Multiancestry genome-wide association study of 520,000 subjects identifies 32 loci associated with stroke and stroke subtypes. Nat. Genet. 2018, 50, 524–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Zhang, H.; Sun, L.; Wang, H.; Cai, H.; Niu, G.; Bai, Y.; Zhang, Y.; Yang, D.; Gu, M.; Xu, P. A Study of GWAS-supported variants of rs9943582 in a Chinese Han population with ischemic stroke: No associations with disease onset and clinical outcomes. J. Stroke Cerebrovasc. Dis. 2017, 26, 2294–2299. [Google Scholar] [CrossRef]
  58. Su, L.; Shen, T.; Liang, B.; Xie, J.; Tan, J.; Chen, Q.; Wei, Q.; Jiang, H.; Gu, L. Association of GWAS-supported loci rs2107595 in HDAC9 gene with ischemic stroke in southern Han Chinese. Gene 2015, 570, 282–287. [Google Scholar] [CrossRef]
  59. Li, S.-h.; Shi, C.-h.; Li, Y.-s.; Li, F.; Tang, M.-b.; Liu, X.-j.; Zhang, S.; Wang, Z.-l.; Song, B.; Xu, Y.-m. Association of GWAS-reported variant rs11196288 near HABP2 with ischemic stroke in Chinese Han population. J. Mol. Neurosci. 2017, 62, 209–214. [Google Scholar] [CrossRef]
  60. Su, L.; Shen, T.; Xie, J.; Yan, Y.; Chen, Z.; Wu, Y.; Yang, J.; Gu, L. Association of GWAS-supported variants rs2200733 and rs6843082 on chromosome 4q25 with ischemic stroke in the southern Chinese Han population. J. Mol. Neurosci. 2015, 56, 585–592. [Google Scholar] [CrossRef]
  61. Baird, A.E. Genetics and genomics of stroke: Novel approaches. J. Am. Coll. Cardiol. 2010, 56, 245–253. [Google Scholar] [CrossRef] [Green Version]
  62. Saxena, R.; Voight, B.F.; Lyssenko, V.; Burtt, N.P.; de Bakker, P.I.; Chen, H.; Roix, J.J.; Kathiresan, S.; Hirschhorn, J.N.; Daly, M.J. Genome-wide association analysis identifies loci for type 2 diabetes and triglyceride levels. Science 2007, 316, 1331–1336. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, Y.; O’Connell, J.R.; McArdle, P.F.; Wade, J.B.; Dorff, S.E.; Shah, S.J.; Shi, X.; Pan, L.; Rampersaud, E.; Shen, H. Whole-genome association study identifies STK39 as a hypertension susceptibility gene. Proc. Natl. Acad. Sci. USA 2009, 106, 226–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Li, Z.; Sun, B.; Gu, M.; Wang, M.; Cheng, X.; Lv, J.; Cen, S.; Zhang, S.; Dai, Z.; Bai, Y. A GWAS-supported variant interacting with diabetes predicts risk of atherothrombotic stroke in Han Chinese population. Int. J. Neurosci. 2019, 129, 167–172. [Google Scholar] [CrossRef] [PubMed]
  65. Ye, Z.; Zhang, H.; Sun, L.; Cai, H.; Hao, Y.; Xu, Z.; Zhang, Z.; Liu, X. GWAS-supported CRP gene polymorphisms and functional outcome of large artery atherosclerotic stroke in Han Chinese. NeuroMol. Med. 2018, 20, 225–232. [Google Scholar] [CrossRef]
  66. Söderholm, M.; Pedersen, A.; Lorentzen, E.; Stanne, T.M.; Bevan, S.; Olsson, M.; Cole, J.W.; Fernandez-Cadenas, I.; Hankey, G.J.; Jimenez-Conde, J. Genome-wide association meta-analysis of functional outcome after ischemic stroke. Neurology 2019, 92, e1271–e1283. [Google Scholar] [CrossRef] [Green Version]
  67. Pfeiffer, D.; Chen, B.; Schlicht, K.; Ginsbach, P.; Abboud, S.; Bersano, A.; Bevan, S.; Brandt, T.; Caso, V.; Debette, S. Genetic imbalance is associated with functional outcome after ischemic stroke. Stroke 2019, 50, 298–304. [Google Scholar] [CrossRef]
  68. Nabika, T.; Ohara, H.; Kato, N.; Isomura, M. The stroke-prone spontaneously hypertensive rat: Still a useful model for post-GWAS genetic studies? Hypertens. Res. 2012, 35, 477–484. [Google Scholar] [CrossRef]
  69. Verma, R.; Ritzel, R.M.; Harris, N.M.; Lee, J.; Kim, T.; Pandi, G.; Vemuganti, R.; McCullough, L.D. Inhibition of miR-141-3p ameliorates the negative effects of poststroke social isolation in aged mice. Stroke 2018, 49, 1701–1707. [Google Scholar] [CrossRef]
  70. Chen, C.; Chu, S.-F.; Liu, D.-D.; Zhang, Z.; Kong, L.-L.; Zhou, X.; Chen, N.-H. Chemokines play complex roles in cerebral ischemia. Neurochem. Int. 2018, 112, 146–158. [Google Scholar] [CrossRef]
  71. Zhang, J.; Yuan, L.; Zhang, X.; Hamblin, M.; Zhu, T.; Meng, F.; Li, Y.; Chen, Y.; Yin, K. Altered long non-coding RNA transcriptomic profiles in brain microvascular endothelium after cerebral ischemia. Exp. Neurol. 2016, 277, 162–170. [Google Scholar] [CrossRef] [Green Version]
  72. Zhang, X.; Tang, X.; Liu, K.; Hamblin, M.H.; Yin, K.-J. Long noncoding RNA Malat1 regulates cerebrovascular pathologies in ischemic stroke. J. Neurosci. 2017, 37, 1797–1806. [Google Scholar] [CrossRef] [PubMed]
  73. Yang, H.; Xi, X.; Zhao, B.; Su, Z.; Wang, Z. KLF4 protects brain microvascular endothelial cells from ischemic stroke induced apoptosis by transcriptionally activating MALAT1. Biochem. Biophys. Res. Commun. 2018, 495, 2376–2382. [Google Scholar] [CrossRef]
  74. Wang, C.; Qu, Y.; Suo, R.; Zhu, Y. Long non-coding RNA MALAT1 regulates angiogenesis following oxygen-glucose deprivation/reoxygenation. J. Cell. Mol. Med. 2019, 23, 2970–2983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ren, L.; Wei, C.; Li, K.; Lu, Z. LncRNA MALAT1 up-regulates VEGF-A and ANGPT2 to promote angiogenesis in brain microvascular endothelial cells against oxygen–glucose deprivation via targetting miR-145. Biosci. Rep. 2019, 39, BSR20180226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Ruan, W.; Li, J.; Xu, Y.; Wang, Y.; Zhao, F.; Yang, X.; Jiang, H.; Zhang, L.; Saavedra, J.M.; Shi, L. MALAT1 up-regulator polydatin protects brain microvascular integrity and ameliorates stroke through C/EBPβ/MALAT1/CREB/PGC-1α/PPARγ pathway. Cell. Mol. Neurobiol. 2019, 39, 265–286. [Google Scholar] [CrossRef]
  77. Li, Z.; Li, J.; Tang, N. Long noncoding RNA Malat1 is a potent autophagy inducer protecting brain microvascular endothelial cells against oxygen-glucose deprivation/reoxygenation-induced injury by sponging miR-26b and upregulating ULK2 expression. Neuroscience 2017, 354, 1–10. [Google Scholar] [CrossRef] [PubMed]
  78. Liu, X.; Hou, L.; Huang, W.; Gao, Y.; Lv, X.; Tang, J. The mechanism of long non-coding RNA MEG3 for neurons apoptosis caused by hypoxia: Mediated by miR-181b-12/15-LOX signaling pathway. Front. Cell. Neurosci. 2016, 10, 201. [Google Scholar] [CrossRef] [Green Version]
  79. Yan, H.; Yuan, J.; Gao, L.; Rao, J.; Hu, J. Long noncoding RNA MEG3 activation of p53 mediates ischemic neuronal death in stroke. Neuroscience 2016, 337, 191–199. [Google Scholar] [CrossRef]
  80. Yan, H.; Rao, J.; Yuan, J.; Gao, L.; Huang, W.; Zhao, L.; Ren, J. Long non-coding RNA MEG3 functions as a competing endogenous RNA to regulate ischemic neuronal death by targeting miR-21/PDCD4 signaling pathway. Cell Death Dis. 2017, 8, 3211. [Google Scholar] [CrossRef]
  81. Liu, J.; Li, Q.; Zhang, K.-S.; Hu, B.; Niu, X.; Zhou, S.-M.; Li, S.-G.; Luo, Y.-P.; Wang, Y.; Deng, Z.-F. Downregulation of the long non-coding RNA Meg3 promotes angiogenesis after ischemic brain injury by activating notch signaling. Mol. Neurobiol. 2017, 54, 8179–8190. [Google Scholar] [CrossRef] [Green Version]
  82. Long, F.-Q.; Su, Q.-J.; Zhou, J.-X.; Wang, D.-S.; Li, P.-X.; Zeng, C.-S.; Cai, Y. LncRNA SNHG12 ameliorates brain microvascular endothelial cell injury by targeting miR-199a. Neural Regen. Res. 2018, 13, 1919. [Google Scholar] [PubMed]
  83. Yin, W.-L.; Yin, W.-G.; Huang, B.-S.; Wu, L.-X. LncRNA SNHG12 inhibits miR-199a to upregulate SIRT1 to attenuate cerebral ischemia/reperfusion injury through activating AMPK signaling pathway. Neurosci. Lett. 2019, 690, 188–195. [Google Scholar] [CrossRef] [PubMed]
  84. Zhao, M.; Wang, J.; Xi, X.; Tan, N.; Zhang, L. SNHG12 promotes angiogenesis following ischemic stroke via regulating miR-150/VEGF pathway. Neuroscience 2018, 390, 231–240. [Google Scholar] [CrossRef]
  85. Wang, J.; Cao, B.; Han, D.; Sun, M.; Feng, J. Long non-coding RNA H19 induces cerebral ischemia reperfusion injury via activation of autophagy. Aging Dis. 2017, 8, 71. [Google Scholar] [CrossRef] [Green Version]
  86. Wang, J.; Zhao, H.; Fan, Z.; Li, G.; Ma, Q.; Tao, Z.; Wang, R.; Feng, J.; Luo, Y. Long noncoding RNA H19 promotes neuroinflammation in ischemic stroke by driving histone deacetylase 1–dependent M1 microglial polarization. Stroke 2017, 48, 2211–2221. [Google Scholar] [CrossRef] [PubMed]
  87. Dykstra-Aiello, C.; Jickling, G.C.; Ander, B.P.; Shroff, N.; Zhan, X.; Liu, D.; Hull, H.; Orantia, M.; Stamova, B.S.; Sharp, F.R. Altered expression of long noncoding RNAs in blood after ischemic stroke and proximity to putative stroke risk loci. Stroke 2016, 47, 2896–2903. [Google Scholar] [CrossRef] [Green Version]
  88. Guo, X.; Yang, J.; Liang, B.; Shen, T.; Yan, Y.; Huang, S.; Zhou, J.; Huang, J.; Gu, L.; Su, L. Identification of novel LncRNA biomarkers and construction of LncRNA-related networks in Han Chinese patients with ischemic stroke. Cell. Physiol. Biochem. 2018, 50, 2157–2175. [Google Scholar] [CrossRef]
  89. Han, X.; Zheng, Z.; Wang, C.; Wang, L. Association between MEG3/miR-181b polymorphisms and risk of ischemic stroke. Lipids Health Dis. 2018, 17, 292. [Google Scholar] [CrossRef] [Green Version]
  90. Zhang, T.; Wang, H.; Li, Q.; Fu, J.; Huang, J.; Zhao, Y. MALAT1 activates the P53 signaling pathway by regulating MDM2 to promote ischemic stroke. Cell. Physiol. Biochem. 2018, 50, 2216–2228. [Google Scholar] [CrossRef]
  91. Zhu, R.; Liu, X.; He, Z. Long non-coding RNA H19 and MALAT1 gene variants in patients with ischemic stroke in a northern Chinese Han population. Mol. Brain 2018, 11, 58. [Google Scholar] [CrossRef]
  92. Zhu, W.; Tian, L.; Yue, X.; Liu, J.; Fu, Y.; Yan, Y. LncRNA expression profiling of ischemic stroke during the transition from the acute to subacute stage. Front. Neurol. 2019, 10, 36. [Google Scholar] [CrossRef] [PubMed]
  93. Moore, D.F.; Li, H.; Jeffries, N.; Wright, V.; Cooper, R.A., Jr.; Elkahloun, A.; Gelderman, M.P.; Zudaire, E.; Blevins, G.; Yu, H. Using peripheral blood mononuclear cells to determine a gene expression profile of acute ischemic stroke: A pilot investigation. Circulation 2005, 111, 212–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Tang, Y.; Xu, H.; Du, X.L.; Lit, L.; Walker, W.; Lu, A.; Ran, R.; Gregg, J.P.; Reilly, M.; Pancioli, A. Gene expression in blood changes rapidly in neutrophils and monocytes after ischemic stroke in humans: A microarray study. J. Cereb. Blood Flow Metab. 2006, 26, 1089–1102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Malik, R.; Bevan, S.; Nalls, M.A.; Holliday, E.G.; Devan, W.J.; Cheng, Y.-C.; Ibrahim-Verbaas, C.A.; Verhaaren, B.F.; Bis, J.C.; Joon, A.Y. Multilocus genetic risk score associates with ischemic stroke in case–control and prospective cohort studies. Stroke 2014, 45, 394–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Khera, A.V.; Chaffin, M.; Aragam, K.G.; Haas, M.E.; Roselli, C.; Choi, S.H.; Natarajan, P.; Lander, E.S.; Lubitz, S.A.; Ellinor, P.T. Genome-wide polygenic scores for common diseases identify individuals with risk equivalent to monogenic mutations. Nat. Genet. 2018, 50, 1219–1224. [Google Scholar] [CrossRef]
  97. Rutten-Jacobs, L.C.; Larsson, S.C.; Malik, R.; Rannikmäe, K.; Sudlow, C.L.; Dichgans, M.; Markus, H.S.; Traylor, M.; Consortium, I.S.G. Genetic risk, incident stroke, and the benefits of adhering to a healthy lifestyle: Cohort study of 306 473 UK Biobank participants. BMJ 2018, 363, k4168. [Google Scholar] [CrossRef] [Green Version]
  98. Dichgans, M.; Pulit, S.L.; Rosand, J. Stroke genetics: Discovery, biology, and clinical applications. Lancet Neurol. 2019, 18, 587–599. [Google Scholar] [CrossRef]
  99. O’Donnell, M.J.; Chin, S.L.; Rangarajan, S.; Xavier, D.; Liu, L.; Zhang, H.; Rao-Melacini, P.; Zhang, X.; Pais, P.; Agapay, S. Global and regional effects of potentially modifiable risk factors associated with acute stroke in 32 countries (INTERSTROKE): A case-control study. Lancet 2016, 388, 761–775. [Google Scholar] [CrossRef]
  100. Hindy, G.; Engström, G.; Larsson, S.C.; Traylor, M.; Markus, H.S.; Melander, O.; Orho-Melander, M. Role of blood lipids in the development of ischemic stroke and its subtypes: A Mendelian randomization study. Stroke 2018, 49, 820–827. [Google Scholar] [CrossRef]
  101. Larsson, S.C.; Scott, R.A.; Traylor, M.; Langenberg, C.C.; Hindy, G.; Melander, O.; Orho-Melander, M.; Seshadri, S.; Wareham, N.J.; Markus, H.S. Type 2 diabetes, glucose, insulin, BMI, and ischemic stroke subtypes: Mendelian randomization study. Neurology 2017, 89, 454–460. [Google Scholar] [CrossRef] [Green Version]
  102. Liu, J.; Rutten-Jacobs, L.; Liu, M.; Markus, H.S.; Traylor, M. Causal impact of type 2 diabetes mellitus on cerebral small vessel disease: A Mendelian randomization analysis. Stroke 2018, 49, 1325–1331. [Google Scholar] [CrossRef] [PubMed]
  103. Akella, A.; Bhattarai, S.; Dharap, A. Long noncoding RNAs in the pathophysiology of ischemic stroke. NeuroMol. Med. 2019, 21, 474–483. [Google Scholar] [CrossRef] [PubMed]
  104. Khera, A.V.; Kathiresan, S. Genetics of coronary artery disease: Discovery, biology and clinical translation. Nat. Rev. Genet. 2017, 18, 331. [Google Scholar] [CrossRef] [PubMed]
  105. Plenge, R.M.; Scolnick, E.M.; Altshuler, D. Validating therapeutic targets through human genetics. Nat. Rev. Drug Discov. 2013, 12, 581–594. [Google Scholar] [CrossRef]
  106. Meschia, J.F. Pharmacogenetics and stroke. Stroke 2009, 40, 3641–3645. [Google Scholar] [CrossRef] [Green Version]
  107. Holmes, M.V.; Ala-Korpela, M.; Smith, G.D. Mendelian randomization in cardiometabolic disease: Challenges in evaluating causality. Nat. Rev. Cardiol. 2017, 14, 577–590. [Google Scholar] [CrossRef] [Green Version]
  108. Sabatine, M.S.; Giugliano, R.P.; Keech, A.C.; Honarpour, N.; Wiviott, S.D.; Murphy, S.A.; Kuder, J.F.; Wang, H.; Liu, T.; Wasserman, S.M. Evolocumab and clinical outcomes in patients with cardiovascular disease. N. Engl. J. Med. 2017, 376, 1713–1722. [Google Scholar] [CrossRef]
  109. Gaudet, D.; Alexander, V.J.; Baker, B.F.; Brisson, D.; Tremblay, K.; Singleton, W.; Geary, R.S.; Hughes, S.G.; Viney, N.J.; Graham, M.J. Antisense inhibition of apolipoprotein C-III in patients with hypertriglyceridemia. N. Engl. J. Med. 2015, 373, 438–447. [Google Scholar] [CrossRef] [Green Version]
  110. Tsimikas, S.; Viney, N.J.; Hughes, S.G.; Singleton, W.; Graham, M.J.; Baker, B.F.; Burkey, J.L.; Yang, Q.; Marcovina, S.M.; Geary, R.S. Antisense therapy targeting apolipoprotein (a): A randomised, double-blind, placebo-controlled phase 1 study. Lancet 2015, 386, 1472–1483. [Google Scholar] [CrossRef]
  111. Saleheen, D.; Natarajan, P.; Armean, I.M.; Zhao, W.; Rasheed, A.; Khetarpal, S.A.; Won, H.-H.; Karczewski, K.J.; O’Donnell-Luria, A.H.; Samocha, K.E. Human knockouts and phenotypic analysis in a cohort with a high rate of consanguinity. Nature 2017, 544, 235–239. [Google Scholar] [CrossRef] [Green Version]
  112. Bush, W.S.; Oetjens, M.T.; Crawford, D.C. Unravelling the human genome–phenome relationship using phenome-wide association studies. Nat. Rev. Genet. 2016, 17, 129. [Google Scholar] [CrossRef]
  113. Emdin, C.A.; Khera, A.V.; Natarajan, P.; Klarin, D.; Won, H.-H.; Peloso, G.M.; Stitziel, N.O.; Nomura, A.; Zekavat, S.M.; Bick, A.G. Phenotypic characterization of genetically lowered human lipoprotein (a) levels. J. Am. Coll. Cardiol. 2016, 68, 2761–2772. [Google Scholar] [CrossRef] [PubMed]
  114. Oberstein, S.A.L.; Jukema, J.W.; Van Duinen, S.G.; Macfarlane, P.W.; van Houwelingen, H.C.; Breuning, M.H.; Ferrari, M.D.; Haan, J. Myocardial infarction in cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy (CADASIL). Medicine 2003, 82, 251–256. [Google Scholar] [CrossRef]
  115. Rufa, A.; Guideri, F.; Acampa, M.; Cevenini, G.; Bianchi, S.; De Stefano, N.; Stromillo, M.L.; Federico, A.; Dotti, M.T. Cardiac autonomic nervous system and risk of arrhythmias in cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy (CADASIL). Stroke 2007, 38, 276–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Bycroft, C.; Freeman, C.; Petkova, D.; Band, G.; Elliott, L.T.; Sharp, K.; Motyer, A.; Vukcevic, D.; Delaneau, O.; O’Connell, J. The UK Biobank resource with deep phenotyping and genomic data. Nature 2018, 562, 203–209. [Google Scholar] [CrossRef] [Green Version]
  117. Maguire, J.M.; Bevan, S.; Stanne, T.M.; Lorenzen, E.; Fernandez-Cadenas, I.; Hankey, G.J.; Jimenez-Conde, J.; Jood, K.; Lee, J.-M.; Lemmens, R. GISCOME–Genetics of Ischaemic Stroke Functional Outcome network: A protocol for an international multicentre genetic association study. Eur. Stroke J. 2017, 2, 229–237. [Google Scholar] [CrossRef]
Table 1. Studies showing stroke related events and clinical or pathological outcomes.
Table 1. Studies showing stroke related events and clinical or pathological outcomes.
Author, YearStroke Related or Associated EventsOutcome (Clinical or Pathological)
Bak et al., 2002Higher stroke death and hospitalization in MZ compared to DZ twinsPotential role of genetic factors in stroke etiology
Flossmann et al., 2004Large and small vessel stroke in comparison to cardioembolic strokePositive family history enhances the risk of large and small vessel stroke
Neves et al., 2021CADASIL with gain of function mutation of notch 3Enhanced cerebral small vessel disease marked by GOM
Neves et al., 2021CADASIL and impaired cerebral vasorelaxation Augmented soluble guanyl cyclase oxidation and reduced cGMP
Felczak et al., 2021Diabetes mellitus and cerebral small blood vessel injury in notch 3 mutationAssociated with mitochondrial dysfunction in endothelial cells and vascular basement membrane injury
Mola-Caminal et al., 2019 PATJ variantsPoor functional outcome post-stroke
Helgadottir et al., 2004Vascular inflammation triggered by 5-lipoxygenase activating protein gene variantsIncrease the risk for myocardial infarction and stroke
Smith et al., 2009Genetic determinants for ischemic stroke on chromosome 9p21 shared with coronary artery disease2 common variants, rs2383207 and rs10757274 associated with modest increase in ischemic stroke risk
Ikram et al., 2009SNPs (rs11833579 and rs12425791) in chromosome 12p13Increased risk for ischemic stroke
Gretarsdottir et al., 2008Atrial fibrillation associated cardioembolic events Increased risk for ischemic stroke associated with markers rs2200733 and rs10033464 located on chromosome 4q25
Malik et al., 201832 loci associated with ischemic stroke and its subtypes Shared traits with blood pressure, cardiac abnormalities, LDL cholesterol, atrial fibrillation and venous thromboembolism
Mola-Caminal et al., 2018Top variant rs76221407 in PATJ geneAssociated with poor functional outcome in ischemic stroke
Zhang et al., 2017Downregulation of MALAT1 expression in in vitro and in vivo stroke modelEnhanced pro-apoptotic bim and pro-inflammatory cytokines (MCP-1, IL-6, and E-selectin) in vitro. Post-stroke functional deterioration in vivo.
Yan et al., 2017OGD-reperfusion induced neuronal injury in vitroLncRNA MEG3 induced neuronal death by downregulating miR-21/PDAC pathway
Long et al., 2018LncRNA SNHG12 overexpression following neuronal ischemiaSuppress neuronal death by downregulating miR-199a
Yin et al., 2019LncRNA SNHG12 salvages ischemia injured neuronsUpregulated Sirtuin-1 and activated AMPK pathway in vitro
Abbreviations: MZ, monozygotic; DZ, dizygotic; CADASIL, cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy; GOM, granular osmiophilic material; cGMP, cyclic guanosine monophosphate; PATJ, pals1-associated tight junction; SNP, single nucleotide polymorphism; LDL, low density lipoprotein; MALAT1, metastasis-associated lung adenocarcinoma transcript 1; MCP-1, monocyte chemoattractant protein-1; IL-6, interleukin-6; OGD, oxygen-glucose deprivation; LncRNA, long non-coding RNA; MEG3, maternally expressed gene 3; miR-21/PDAC, micro ribonucleic acid-21/pancreatic ductal adenocarcinoma; SNHG12, small nucleolar RNA host gene 12; miR-199a, micro ribonucleic acid-199a; AMPK, adenosine monophosphate-activated protein kinase.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Appunni, S.; Rubens, M.; Ramamoorthy, V.; Saxena, A.; McGranaghan, P.; Veledar, E. Stroke Genomics: Current Knowledge, Clinical Applications and Future Possibilities. Brain Sci. 2022, 12, 302. https://doi.org/10.3390/brainsci12030302

AMA Style

Appunni S, Rubens M, Ramamoorthy V, Saxena A, McGranaghan P, Veledar E. Stroke Genomics: Current Knowledge, Clinical Applications and Future Possibilities. Brain Sciences. 2022; 12(3):302. https://doi.org/10.3390/brainsci12030302

Chicago/Turabian Style

Appunni, Sandeep, Muni Rubens, Venkataraghavan Ramamoorthy, Anshul Saxena, Peter McGranaghan, and Emir Veledar. 2022. "Stroke Genomics: Current Knowledge, Clinical Applications and Future Possibilities" Brain Sciences 12, no. 3: 302. https://doi.org/10.3390/brainsci12030302

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