Skip to main content
Transcription logoLink to Transcription
. 2014 Oct 31;5(4):e964559. doi: 10.4161/21541264.2014.964559

Regulatory SNPs and transcriptional factor binding sites in ADRBK1, AKT3, ATF3, DIO2, TBXA2R and VEGFA

Norman E Buroker 1,*
PMCID: PMC4581348  PMID: 25483406

Abstract

Regulatory single nucleotide polymorphisms (rSNPs) which change the transcriptional factor binding sites (TFBS) for transcriptional factors (TFs) to bind DNA were reviewed for the ADRBK1 (GRK2), AKT3, ATF3, DIO2, TBXA2R and VEGFA genes. Changes in the TFBS where TFs attach to regulate these genes may result in human sickness and disease. The highlights of this previous work were reviewed for these genes.

Keywords: rSNPs, transcriptional factors, transcriptional factor binding site, ADRBK1, AKT3, ATF3, DIO2, TBXA2R and VEGFA genes

Introduction

Genome-wide association studies (GWAS) over the last decade have identified nearly 6,500 disease or trait-predisposing simple nucleotide polymorphisms (SNPs) where only 7% of these are located in protein-coding regions of the genome 1, 2 and the remaining 93% are located within non-coding areas 3, 4 such as regulatory or inter-genic regions. SNPs which occur in the putative regulatory region of a gene where a single base change in the DNA sequence of a potential transcriptional factor binding site (TFBS) may affect the process of gene expression are drawing more attention. 5–7 A SNP in a TFBS can have multiple consequences. Often the SNP does not change the TFBS interaction nor does it alter gene expression since a transcriptional factor (TF) will usually recognize a number of different binding sites in the gene. In some allele-specific gene expression cases, a TF's affinity for binding a response element may be altered where the SNP increases or decreases the TF binding ability. In rare cases, a SNP may eliminate the natural binding site or generate a new binding site. In such cases the gene is no longer regulated by the original TF. Therefore, functional rSNPs in TFBS may result in differences in gene expression, phenotypes and susceptibility to environmental exposure.7 Examples of rSNPs associated with disease susceptibility are numerous and several reviews have been published.7–10 In this review, a rSNP will be considered as a SNP that alters the DNA landscape for TF binding resulting in different TFBS that may affect gene regulation.

In genetics, linkage disequilibrium (LD) is defined as the non-random association in a given population between the alleles of two or more loci.11 LD between SNPs in the regulatory region of a gene can be used as a method of identifying associations of certain haplotypes with sickness or disease in a population.12, 13 This can be achieved when levels of LD between SNPs within haplotypes are seen to change substantially in a disease or sickness group of a population when compared to the normal baseline population. In such cases, the relationship between LD, SNPs and TFBS can be used to identify potential binding changes of TFs responsible for gene regulation which could result in disease or sickness.14–21 In this report, LD is considered to be the non-random association of SNP alleles within the gene. The purpose of this review is to highlight some of these genetic associations for six genes with known human disease and sickness that have been recently studied.14, 16–19, 22, 23

ADRBK1 (GRK2) rSNPs, TFBS and Heart Disease

G protein-coupled receptor kinases (GRKs) are a family of seven serine/threonine protein kinases with important and varied roles in regulating cellular signaling.24–26 The GRK2 gene (ADRBK1), which is located at chromosome 11q13.1 and is about 20 kb in size, is an important regulator of beta-adrenergic signaling and plays a central role in heart failure (HF) pathology. 27–29 Consequently, the inhibition of GRK2 has become an emerging treatment option of HF since the inhibition appears to be a powerful therapeutic approach and appears to provide complementation to β-blockade.28 When the heart is injured or stressed there is activation of the sympathetic nervous system (SNS) and adrenergic overdrive via excessive catecholamines that leads to uncoupling desensitization of β-adrenergic receptors (βARs) in the heart, which is brought about by the up regulation of GRK2.30–32 Lymphocyte GRK2 levels increase during acute myocardial infarction (MI) and are associated with worsening cardiac function 33 enhanced GRK2 activity has been shown to contribute to cardiac decompensation.34, 35 Independent of its uncoupling properties on βARs, it appears to be a pro-death kinase especially in stressed myocytes.36, 37 Consequently, the inhibition of GRK2 or its genetic ablation has been show to improve cardiac function in myocardial injury.38

rSNPs in the promoter region (rs12286664 and -703bp), introns (rs1894111, rs7128315 and rs948988) and 3’ UTR (rs4370946) of the ADRBK1 gene are in LD in the black population.27 These rSNP alleles alter the DNA landscape for potential transcriptional factors (TFs) to attach resulting in changes in transcriptional factor binding sites (TFBS). The alleles of each rSNP were found to produce unique TFBS resulting in potential changes in TF ADRBK1 regulation.18 As an example, the rs7128315 rSNP common ADRBK1-G allele creates TFBS for the SREBF1 & 2 TFs which are involved with the cholesterol synthesis pathway. These TFBS occur only once in the gene and do not occur when the rs7128315 rSNP minor ADRBK1-A allele (0.32) is present.18 SREBF1 & 2 are master regulators of lipid homeostasis 39, 40 and have been associated with coronary heart disease in Han Chinese 41 and heart dysfunction.42

AKT3 rSNPs, TFBS and High Altitude Sickness

The phosphatidylinositol 3-kinase (PI3K)/AKT pathway plays a key role in numerous cellular functions including proliferation, adhesion, migration, invasion, metabolism and survival43 as well as angiogenesis. The PI3K/AKT pathway modulates the expression of angiogenic factors such as nitric oxide and angiopoietins.44 The v-akt murine thymoma viral oncogene homolog 3 (AKT3), which is located at chromosome 1q44 and is about 362.8 kb in size. It is one of three isoforms of the AKTs which are major downstream targets of growth factor receptor tyrosine kinases that signal through PI3K.45 The AKT3 single nucleotide polymorphism (rSNP, rs4590656) from intron one was recently found to be significantly associated with three physiological parameters (hemoglobin, hematocrit and red blood cell count) in chronic mountain sickness patients indicating the strong association of this gene with angiogenesis.46 The rs4590656 rSNP common AKT3-C allele creates TFBS for the ARNT:AHR and HIF1α:ARNT TFs which are involved with xenobiotic metabolism and cellular and systemic responses to hypoxia, respectively; while these TFBS are not present with the rs4590656 rSNP minor AKT3-T allele (0.41).16 Other intron one AKT3 rSNPS (rs10157763, rs10927067 and rs2125230) have been significantly associated with aggressive prostate cancer (PCA). 47 These rSNPs also alter the TFBS for TFs that regulate the AKT3 gene,16 where the rs2125230 rSNP minor AKT3-A allele (0.26) creates a unique TFBS for the IRF1 TF that occurs only once in the gene.16 The IRF1 (interferon regulatory factor 1) TF is a member of the interferon regulatory transcription factor (IRF) family which share several common effects which include antiviral agents that fight tumors. Other intron one rSNPs (rs4132509, rs12031994, rs2345994) have been found to be significantly associated with renal cell carcinoma risk (RCC).48 The rs12031994 rSNP minor AKT3-A allele (0.13) creates a unique TFBS that occurs only once in the gene for the TAL1:GATA1 TF which has been implicated in the genesis of hemopoietic malignancies.49 All seven AKT3 rSNP have been found to alter the DNA landscape resulting in changes in transcriptional factor binding sites (TFBS)16 for potential transcriptional factors (TFs) to attach.

ATF3 rSNPs, TFBS and Hypospadias

The activating transcription factor 3 (ATF3) gene, which is located at chromosome 1q32.3 and is about 55.4 kb in size. It is a member of the activating transcription factor/cAMP responsive element binding (CREB) protein family of transcription factors, which share the basic region-leucine zipper (bZip) DNA binding motif (TGACGTCA). This gene is induced by a variety of signals including many of those encountered by cancer cells, and is involved in the complex process of cellular stress response.50–52 ATF3 has been viewed as a hub of the cellular adaptive response network which allows cells adapt to disturbances in homeostasis.53 This gene has been shown to be up-regulated during sexual differentiation 54 which indicates a potential role in hypospadias. Three unlinked ATF3 rSNPs (rs3125289, rs1877474 and rs11119982) which span a 16 kb region of intron one have been independently found to be significantly associated with the risk of hypospadias.55 These rSNP alleles alter the DNA landscape for potential transcriptional factors (TFs) to attach resulting in changes in transcriptional factor binding sites (TFBS). These TFBS changes have been examined with respect to the human etiology of hypospadias which has been found to be significantly associated with the rSNPs .17

The rs3125289 rSNP AFT3-T allele (0.55) creates a unique TFBS for the SRY TF which is a transcriptional regulator that controls a genetic switch in male development. The TFBS for the SRY TF is only found once in the ATF3 gene and does not occur with the rs3125289 rSNP minor AFT3-C allele.17 The rs11119982 rSNP AFT3-T allele (0.51) creates a unique TFBS for the MYB TF which is plays an important role in the control of proliferation and differentiation of hematopoietic progenitor cells. The TFBS for the MYB TF is only found once in the ATF3 gene and does not occur with the rs11119982 rSNP minor AFT3-C allele.17

DIO2 rSNPs, TFBS and Type 2 Diabetes Mellitus

The type 2 deiodeinase gene (DIO2) encodes a deiodinase that coverts the thyroid prohormone, thyroxine (T4), to the biologically active triiodothyronine (T3) where T3 plays an important role in the regulation of energy balance and glucose metabolism.56–59 DIO2 is found in the thyroid gland, cardiac and skeletal muscle, brown adipose tissue, placenta, pituitary, central nervous system (CNS) and at low levels in kidney and pancreases.60–62 The DIO2 gene maps to human chromosome 14q24.3 and is about 15 kb in size. The coding region consists of two exons separated by an intron of approximately 7.4 kb.63 Several SNPs have been found in the gene which have been studied is association with mental retardation (MR),64 osteoarthritis 65 and early-onset type 2 diabetes mellitus (T2DM).66 Three of the common SNPs in the gene (rs225014, rs225012 and rs225010) have been found to be in strong LD with each other while the rs225012 and rs225010 SNPs have been shown to have a positive association with MR.64 The haplotypes of two SNPs (rs225014 and rs12885300) have been shown to have a significant association with symptomatic osteoarthritis in Dutch women.65 Three SNPs (rs225011, rs225014 and rs225015) which are in LD were found to be modestly associated with early-onset of T2DM in Pima Indians while these SNPs and rs6574549 were found to be nominally associated with hepatic glucose output.66 The rs6574549 SNP was also found to be associated with fasting insulin, insulin action and energy expenditure.66 These studies suggest that some DIO2 SNPs may be affecting the regulatory network for the gene expression in humans. An examination was made between DIO2 SNPs in LD and the transcription factor binding site (TFBS) changes resulting from the SNPs.23 Regulatory SNPs (rSNPs) were found in the promoter region involving a novel SNP (-2035bp), 5’UTR (rs12885300), also in intron one (rs225010, rs225011 and rs225012), exon two [rs225014 (Thr92Ala)] and 3’ UTR (rs6574549 and rs225015) of the DIO2 gene which are all in LD.66 The rs225012 rSNP common DIO2-G allele creates four unique TFBS for the E2F6, ELF1, EGR1 and SPI1 TFs which occur only once in the DIO2 gene and are involved with the control of the cell cycle and action of tumor suppressor proteins, transcription regulation, and gene expression during myeloid and B-lymphoid cell development, respectively.23 The rs225012 rSNP minor DIO2-A allele (0.29) creates two unique TFBS for the HOXA5 and NKX3–2 TFs which occur only once in the DIO2 gene and are involved with the development regulatory system and negative regulation, respectively.23 The rs225014 rSNP common DIO2-T allele (0.82) creates a unique TFBS for the FOXC1 TF which occurs only once in the DIO2 gene and is an important regulator of cell viability and resistance to oxidative stress. 23 The rs6574549 rSNP common DIO2-G allele creates a unique TFBS for the LHX3 TF which occurs only once in the DIO2 gene and is required for pituitary development and motor neuron specification.23 The rs6574549 rSNP rare DIO2-G allele (0.06) creates a unique TFBS for the POU2F2 TF which occurs only once in the DIO2 gene and regulates transcription in a number of tissues in addition to activating immunoglobulin gene expression.23 The rs225015 rSNP common DIO2-G allele creates a unique TFBS for the EBF1 TF which occurs only once in the DIO2 gene and is a transcriptional activator.23 The rs225015 rSNP minor DIO2-A allele (0.40) creates a unique TFBS for the TCF7L2 TF which occurs only once in the DIO2 gene and is implicated in blood glucose homeostasis.23

TBXA2R rSNPs, TFBS and Asthma

The thromboxane A2 receptor (TBXA2R) gene which is located at chromosome 19p13.3 is a member of the seven-transmembrane G-protein-coupled receptor super family, which interacts with intracellular G proteins. It regulates different downstream signaling cascades, and induces many cellular responses including the intracellular calcium influx, cell migration and proliferation, and apoptosis.67 This gene is abundantly expressed in tissues (at the mRNA and protein levels) targeted by the TBXA2R ligand thromboxane A2 (TXA2) that include erythroleukaemia cells, vascular and bronchial smooth muscle, uterus and placental tissue, endothelium, epithelium, trophoblasts, thymus, liver and small intestine.68 The activation of TBXA2R in bronchial smooth muscle cells by its ligand results in intercellular calcium mobilization with subsequent bronchoconstriction. This contributes to bronchial smooth muscle hyperplasia and airway remodeling, which occurs in response to chronic airway inflammation in asthma.69

Four linked TBXA2R rSNPs (rs2238631, rs2238632, rs2238633 and rs2238634) which span a 431bp region of intron one have been found to be in LD with two exon 3 SNPs [rs11318632, (c.795 T > C) and rs4523 (c.924 T > C)],70 which are approximately 8.4 kb downstream from the intron one rSNPs. The rs11318632 and rs4523 rSNPs from exon 3 are synonymous changes and unlikely to influence the characteristics of the receptor protein. The exon 3 rSNPs have been associated with asthma and its related phenotypes in Asian populations. The rs4523 SNP has been found to be associated with adult asthma in a Japanese population71 and childhood atopic asthma in a Chinese population.72 The rs11318632 SNP has been found to be associated with atopic asthma in a Korean population.73 Two haplotypes (H2 & H4) involving the four linked TBXA2R SNPs from intron one where found to influence TBXA2R transcriptional activity and were also associated with asthma-related phenotypes.70

Four rSNPs (rs2238631, rs2238632, rs2238633 and rs2238634) in intron one, two rSNPs (rs1131882 and rs4523) in exon 3 and one rSNP (rs5756) in the 3’UTR of the thromboxane A2 receptor (TBXA2R) gene have been associated with childhood-onset asthma in Asians. These rSNP alleles alter the DNA landscape for potential transcriptional factors (TFs) to attach resulting in changes in transcriptional factor binding sites (TFBS). These TFBS changes were examined with respect to asthma which has been found to be significantly associated with the rSNPs.

The rs2238631 rSNP common TBXA2R-G allele creates a unique TFBS for the FOXC1 TF which occurs only once in the TBXA2R gene and is an important regulator of cell viability and resistance to oxidative stress.19 The rs2238631 rSNP minor TBXA2R-A allele (0.18) creates two unique TFBS for the ELK1 and ELK4 TFs which each occur only once in the TBXA2R gene and are involved with the ras-raf-MAPK signaling cascade, transcriptional activation and repression, respectively.19 The rs2238632 rSNP minor TBXA2R-T allele (0.49) creates two unique TFBS for the CREB1 and HIF1α:ARNT TFs which each occur only once in the gene and are involved with DNA binding, cellular and systemic responses to hypoxia, respectively.19 The rs2238634 rSNP minor TBXA2R-T allele (0.19) creates a unique TFBS for the HLTF TF which occurs only once in the TBXA2R gene and whose protein alters the chromatin structure around certain genes during transcription.19 The rs1131882 rSNP common TBXA2R-T allele (0.57) creates a unique TFBS for the SOX17 TF which occurs only once in the gene and is a transcription regulator that binds target promoter DNA and bends the DNA.19 The rs4523 rSNP minor TBXA2R-C allele (0.21) creates a unique TFBS for the AR TF which occurs only once in the gene and is a steroid hormone receptor that regulate eukaryotic gene expression and affects cellular proliferation and differentiation in target tissues.19

VEGFA rSNPs, TFBS and Hypoxia

The vascular endothelial growth factor (VEGF) is a family of key regulators in critical physiological and pathological angiogenesis74 including tissue growth, wound healing, rheumatoid arthritis, proliferative retinopathies, cardiovascular disease and cancer75 and is a growth factor activator for angiogenesis, vasculogenesis and endothelial cell growth. It has been shown that the VEGF is an important component of the pathogenesis of high altitude adaptation and sickness in most studies76–81 but not all.82 Presently seven VEGF family members and 14 alternative splicing variants have been identified in humans.83–85 Of the 14 splicing variants, 12 are VEGFA isoforms85 with three (VEGFA-121, -165 & -189) being differentially expressed in humans visiting or living in high altitude environments and as well as in chronic mountain sickness (CMS) patients.77, 78 Among all family members, VEGFA, which is located at chromosome 6p12, is the most potent and best know angiogenic protein and exerts its biologic effect through interaction with cell-surface receptors which triggers a cascade of downstream dimerizations and phosphorylations.86

Mountain sickness (MS) occurs among humans visiting or inhabiting high altitude environments. Genetic analyses of seven rSNPs in the promoter region of VEGFA gene were analyzed in lowland (Han) and highland (Tibetan) Chinese 14. The seven VEGFA rSNPs were evaluated in Han and Tibetan patients with acute and chronic mountain sickness, respectively. The rSNPs are rs699947, rs34357231, rs79469752, rs13207351, rs28357093, rs1570360 and rs2010963 which are found in the promoter ranging from -2578 bps to -634bps from the transcriptional start site (TSS), respectively. These rSNPs are found in TFBS and all alter these binding sites.14 Arterial oxygen saturation of hemoglobin (SaO2) has been found to be significantly associated with the rs699947, rs34357231, rs13207351 and rs1570360 rSNPs in Han patients with AMS. 14 MS was found to be significantly associated with these rSNPs when compared to their Han and Tibetan control groups indicating these nucleotide substitutions results in TFBS changes which apparently have a physiological effect in the development of high altitude sickness.14 LD was found between rSNPs (rs13207351 and rs1570360) and between rSNPs (rs79469752 and rs28357093) among the control and sickness groups.14 87 The TFBS [HIF1α:ARNT] of the hypoxia inducible factor-1α and aryl hydrocarbon receptor nuclear translocator TFs is generated by the rs699947 common rSNP VEGFA-C allele but not with the minor VEGFA-A allele (0.23). This rSNP is not in LD with the other rSNPs in the VEGFA promoter.14 The HIF1α and ARNT TFs play an essential role in cellular and systemic responses to hypoxia and its TFBS occurs only in the promoter of the VEGFA gene,14 consequently, the rs699947 rSNP may have an impact on SaO2 in patients with high altitude mountain sickness.

SNPs, TFBS and Ethnic Populations

A given rSNP allele frequency can vary between ethnic or racial groups due to historical population bottlenecks. This would affect the occurrence of TFBS and the TFs regulating genes and should impact the groups that are susceptible to disease or sickness. As an example, the rs225015 rSNP DIO2-A allele (above) creates a TFBS for the TCF7L2 TF, which is implicated in blood glucose homeostasis, and has an allele frequency of 0.40 in most ethnic and racial groups; however, the frequency is 0.81 in Pima Indians .66 This rSNP is modestly associated with early-onset Type 2 Diabetes Mellitus and hepatic glucose output in the Pima Indians of Arizona66 which may in part result from the higher occurrence of DIO2-A allele and TCF7L2 TFBS then in most other ethnic or racial groups.

Concluding Remarks

In summary, SNPs which occur in the non-coding regions of these genes have been found to be associated with human diseases or sicknesses. These non-coding regions host the binding sites for the transcription factors that regulate gene expression. The rs7128315 ADRBK1 rSNP which has been associated with coronary heart disease and dysfunction in Asians alters the SREBF1 & 2 TFBS which could interfere with lipid homeostasis. The rs2125230 AKT3 rSNP which has been associated with aggressive prostate cancer alters the IRF1 TFBS which could interfere with antiviral agents that fight tumors. The rs3125289 AFT3 rSNP which has been associated with the risk of hypospadias alters the SRY TFBS which could affect male development. The rs225015 DIO2 rSNP which is modestly associated with early-onset of T2DM in Pima Indians alters the TCF7L2 TFBS which is implicated in blood glucose homeostasis. The rs4523 TBXA2R rSNP which has been associated with asthma in Asians alters the AR TFBS which could affect cellular proliferation and differentiation. The rs699947 VEGFA rSNP which has been associated with arterial oxygen saturation of hemoglobin alters the HIF1α:ARNT TFBS which could affect the cellular and systemic responses to hypoxia. In conclusion regulatory SNPs underlying the allele-specific changes in transcriptional factor binding sites can alter the transcriptional factors that regulate these genes.

REFERENCES

  • 1.Pennisi E. The Biology of Genomes. Disease risk links to gene regulation. Science; 332:1031. [DOI] [PubMed] [Google Scholar]
  • 2.Kumar V, Wijmenga C, Withoff S. From genome-wide association studies to disease mechanisms: celiac disease as a model for autoimmune diseases. Semin Immunopathol; 34:567–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Hindorff LA, Sethupathy P, Junkins HA, Ramos EM, Mehta JP, Collins FS, et al. . Potential etiologic and functional implications of genome-wide association loci for human diseases and traits. Proc Natl Acad Sci U S A 2009; 106:9362–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kumar V, Westra HJ, Karjalainen J, Zhernakova DV, Esko T, Hrdlickova B, et al. . Human disease-associated genetic variation impacts large intergenic non-coding RNA expression. PLoS Genet; 9:e1003201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Knight JC. Functional implications of genetic variation in non-coding DNA for disease susceptibility and gene regulation. Clin Sci (Lond) 2003; 104:493–501. [DOI] [PubMed] [Google Scholar]
  • 6.Wang X, Tomso DJ, Liu X, Bell DA. Single nucleotide polymorphism in transcriptional regulatory regions and expression of environmentally responsive genes. Toxicol Appl Pharmacol 2005; 207:84–90. [DOI] [PubMed] [Google Scholar]
  • 7.Chorley BN, Wang X, Campbell MR, Pittman GS, Noureddine MA, Bell DA. Discovery and verification of functional single nucleotide polymorphisms in regulatory genomic regions: current and developing technologies. Mutat Res 2008; 659:147–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Prokunina L, Alarcon-Riquelme ME. Regulatory SNPs in complex diseases: their identification and functional validation. Expert Rev Mol Med 2004; 6:1–15. [DOI] [PubMed] [Google Scholar]
  • 9.Buckland PR. The importance and identification of regulatory polymorphisms and their mechanisms of action. Biochim Biophys Acta 2006; 1762:17–28. [DOI] [PubMed] [Google Scholar]
  • 10.Sadee W, Wang D, Papp AC, Pinsonneault JK, Smith RM, Moyer RA, et al. . Pharmacogenomics of the RNA world: structural RNA polymorphisms in drug therapy. Clin Pharmacol Ther; 89:355–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Lewontin RC, Kojima K. The Evolutionary Dynamics of Complex Polymorphisms. Evolution 1960; 14:458–72. [Google Scholar]
  • 12.Buroker NE, Ning XH, Zhou ZN, Li K, Cen WJ, Wu XF, et al. . VEGFA SNPs and transcriptional factor binding sites associated with high altitude sickness in Han and Tibetan Chinese at the Qinghai-Tibetan Plateau. J Physiol Sci. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Buroker NE, Ning XH, Zhou ZN, Li K, Cen WJ, Wu XF, et al. . AKT3, ANGPTL4, eNOS3, and VEGFA associations with high altitude sickness in Han and Tibetan Chinese at the Qinghai-Tibetan Plateau. International journal of hematology 2012. [DOI] [PubMed] [Google Scholar]
  • 14.Buroker NE. VEGFA rSNPs, transcriptional factor binding sites and human disease. J Physiol Sci; 64:73–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Buroker NE, Ning XH, Zhou ZN, Li K, Cen WJ, Wu XF, et al. . VEGFA SNPs and transcriptional factor binding sites associated with high altitude sickness in Han and Tibetan Chinese at the Qinghai-Tibetan Plateau. J Physiol Sci; 63:183–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Buroker NE. AKT3 rSNPs, Transcritional Factor Binding Sites and Human Disease. Open Journal of Blood Diseases 2013; 3:116–29. [Google Scholar]
  • 17.Buroker NE. ATF3 rSNPs, transcriptional factor binding sites and human etiology. Open Journal of Genetics 2013; 3:253–61. [Google Scholar]
  • 18.Buroker NE. ADRBK1 (GRK2) rSNPs, Transcriptional Factor Binding Sites and Cardiovascular Disease in the Black Population. Journal of Cardiovascular Disease 2014; 2. [Google Scholar]
  • 19.Buroker NE. TBXA2R rSNPs, Transcriptional Factor Binding Sites and Asthma in Asians. Open Journal of Pediatrics 2014; 4:148–61. [Google Scholar]
  • 20.Buroker NE. ADRBD1 (GRK2), TBXA2R and VEGFA rSNPs in KLF4 and SP1 TFBS Exhibit Linkage Disequilibrium. Open Journal of Genetics 2014; 4. [Google Scholar]
  • 21.Norman E. Buroker XN, Zhou Zhaonian, Li Kui, Cen Weijun, Wu Xiufeng, Zhu Weizhong, Scott C. Ronald, Chen Shihan. SNPs and TFBS Associated with High Altitude Sickness. Open Journal of Blood Diseases 2013; 3:85–93. [Google Scholar]
  • 22.Buroker NE, Ning XH, Zhou ZN, Li K, Cen WJ, Wu XF, et al. . VEGFA SNPs and transcriptional factor binding sites associated with high altitude sickness in Han and Tibetan Chinese at the Qinghai-Tibetan Plateau. Journal of Physiological Sciences 2013; 63:183–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Buroker NE. DIO2 rSNPs, transcription factor binding sites and disease. Journal of Thyroid Cancer 2014. [Google Scholar]
  • 24.Pitcher JA, Freedman NJ, Lefkowitz RJ. G protein-coupled receptor kinases. Annu Rev Biochem 1998; 67:653–92. [DOI] [PubMed] [Google Scholar]
  • 25.Ribas C, Penela P, Murga C, Salcedo A, Garcia-Hoz C, Jurado-Pueyo M, et al. . The G protein-coupled receptor kinase (GRK) interactome: role of GRKs in GPCR regulation and signaling. Biochim Biophys Acta 2007; 1768:913–22. [DOI] [PubMed] [Google Scholar]
  • 26.So CH, Michal A, Komolov KE, Luo J, Benovic JL. G protein-coupled receptor kinase 2 (GRK2) is localized to centrosomes and mediates epidermal growth factor-promoted centrosomal separation. Mol Biol Cell; 24:2795–806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Lobmeyer MT, Wang L, Zineh I, Turner ST, Gums JG, Chapman AB, et al. . Polymorphisms in genes coding for GRK2 and GRK5 and response differences in antihypertensive-treated patients. Pharmacogenet Genomics; 21:42–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Cannavo A, Liccardo D, Koch WJ. Targeting cardiac beta-adrenergic signaling via GRK2 inhibition for heart failure therapy. Front Physiol; 4:264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Lymperopoulos A. Physiology and pharmacology of the cardiovascular adrenergic system. Front Physiol; 4:240. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 30.Theilade J, Strom C, Christiansen T, Haunso S, Sheikh SP. Differential G protein receptor kinase 2 expression in compensated hypertrophy and heart failure after myocardial infarction in the rat. Basic Res Cardiol 2003; 98:97–103. [DOI] [PubMed] [Google Scholar]
  • 31.Raake PW, Vinge LE, Gao E, Boucher M, Rengo G, Chen X, et al. . G protein-coupled receptor kinase 2 ablation in cardiac myocytes before or after myocardial infarction prevents heart failure. Circ Res 2008; 103:413–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Santulli G. Adrenal Signaling in Heart Failure: Something More Than a Distant Ship's Smoke on the Horizon. Hypertension; 63:215–6. [DOI] [PubMed] [Google Scholar]
  • 33.Santulli G, Campanile A, Spinelli L, Assante di Panzillo E, Ciccarelli M, Trimarco B, et al. . G protein-coupled receptor kinase 2 in patients with acute myocardial infarction. Am J Cardiol; 107:1125–30. [DOI] [PubMed] [Google Scholar]
  • 34.Petrofski JA, Koch WJ. The beta-adrenergic receptor kinase in heart failure. J Mol Cell Cardiol 2003; 35:1167–74. [DOI] [PubMed] [Google Scholar]
  • 35.Huang ZM, Gao E, Fonseca FV, Hayashi H, Shang X, Hoffman NE, et al. . Convergence of G protein-coupled receptor and S-nitrosylation signaling determines the outcome to cardiac ischemic injury. Science signaling; 6:ra95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Brinks H, Boucher M, Gao E, Chuprun JK, Pesant S, Raake PW, et al. . Level of G protein-coupled receptor kinase-2 determines myocardial ischemia/reperfusion injury via pro- and anti-apoptotic mechanisms. Circ Res; 107:1140–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Chen M, Sato PY, Chuprun JK, Peroutka RJ, Otis NJ, Ibetti J, et al. . Prodeath signaling of G protein-coupled receptor kinase 2 in cardiac myocytes after ischemic stress occurs via extracellular signal-regulated kinase-dependent heat shock protein 90-mediated mitochondrial targeting. Circ Res; 112:1121–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Khan M, Mohsin S, Toko H, Alkatib M, Nguyen J, Truffa S, et al. . Cardiac Progenitor Cells Engineered With betaARKct Have Enhanced beta-Adrenergic Tolerance. Molecular therapy : the journal of the American Society of Gene Therapy. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Eberle D, Hegarty B, Bossard P, Ferre P, Foufelle F. SREBP transcription factors: master regulators of lipid homeostasis. Biochimie 2004; 86:839–48. [DOI] [PubMed] [Google Scholar]
  • 40.Raghow R, Yellaturu C, Deng X, Park EA, Elam MB. SREBPs: the crossroads of physiological and pathological lipid homeostasis. Trends Endocrinol Metab 2008; 19:65–73. [DOI] [PubMed] [Google Scholar]
  • 41.Alobeidy BF, Li C, Alzobair AA, Liu T, Zhao J, Fang Y, et al. . The Association Study between Twenty One Polymorphisms in Seven Candidate Genes and Coronary Heart Diseases in Chinese Han Population. PLoS One; 8:e66976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Tang XL, Deng LB, Lin JR, Zhang WL, Liu SM, Wei Y, et al. . [Sterol regulatory element binding protein 1 and its target gene networks]. Yi Chuan; 35:607–15. [DOI] [PubMed] [Google Scholar]
  • 43.Bader AG, Kang S, Zhao L, Vogt PK. Oncogenic PI3K deregulates transcription and translation. Nat Rev Cancer 2005; 5:921–9. [DOI] [PubMed] [Google Scholar]
  • 44.Karar J, Maity A. PI3K/AKT/mTOR Pathway in Angiogenesis. Frontiers in molecular neuroscience 2011; 4:51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Testa JR, Bellacosa A. AKT plays a central role in tumorigenesis. Proc Natl Acad Sci U S A 2001; 98:10983–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Buroker NE, Ning XH, Zhou ZN, Li K, Cen WJ, Wu XF, et al. . AKT3, ANGPTL4, eNOS3, and VEGFA associations with high altitude sickness in Han and Tibetan Chinese at the Qinghai-Tibetan Plateau. International journal of hematology; 96:200–13. [DOI] [PubMed] [Google Scholar]
  • 47.Lavender NA, Rogers EN, Yeyeodu S, Rudd J, Hu T, Zhang J, et al. . Interaction among apoptosis-associated sequence variants and joint effects on aggressive prostate cancer. BMC Med Genomics; 5:11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Shu X, Lin J, Wood CG, Tannir NM, Wu X. Energy balance, polymorphisms in the mTOR pathway, and renal cell carcinoma risk. J Natl Cancer Inst; 105:424–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Kassouf MT, Hughes JR, Taylor S, McGowan SJ, Soneji S, Green AL, et al. . Genome-wide identification of TAL1's functional targets: insights into its mechanisms of action in primary erythroid cells. Genome research; 20:1064–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Chen BP, Liang G, Whelan J, Hai T. ATF3 and ATF3 delta Zip. Transcriptional repression versus activation by alternatively spliced isoforms. J Biol Chem 1994; 269:15819–26. [PubMed] [Google Scholar]
  • 51.Hai T, Hartman MG. The molecular biology and nomenclature of the activating transcription factor/cAMP responsive element binding family of transcription factors: activating transcription factor proteins and homeostasis. Gene 2001; 273:1–11. [DOI] [PubMed] [Google Scholar]
  • 52.Thompson MR, Xu D, Williams BR. ATF3 transcription factor and its emerging roles in immunity and cancer. Journal of molecular medicine 2009; 87:1053–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Hai T, Wolford CC, Chang YS. ATF3, a hub of the cellular adaptive-response network, in the pathogenesis of diseases: is modulation of inflammation a unifying component? Gene Expr; 15:1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Liu B, Agras K, Willingham E, Vilela ML, Baskin LS. Activating transcription factor 3 is estrogen-responsive in utero and upregulated during sexual differentiation. Horm Res 2006; 65:217–22. [DOI] [PubMed] [Google Scholar]
  • 55.Beleza-Meireles A, Tohonen V, Soderhall C, Schwentner C, Radmayr C, Kockum I, et al. . Activating transcription factor 3: a hormone responsive gene in the etiology of hypospadias. Eur J Endocrinol 2008; 158:729–39. [DOI] [PubMed] [Google Scholar]
  • 56.Danforth E., Jr. The role of thyroid hormones and insulin in the regulation of energy metabolism. The American journal of clinical nutrition 1983; 38:1006–17. [DOI] [PubMed] [Google Scholar]
  • 57.Krotkiewski M. Thyroid hormones in the pathogenesis and treatment of obesity. European journal of pharmacology 2002; 440:85–98. [DOI] [PubMed] [Google Scholar]
  • 58.Silva JE. Thyroid hormone control of thermogenesis and energy balance. Thyroid 1995; 5:481–92. [DOI] [PubMed] [Google Scholar]
  • 59.Freake HC, Oppenheimer JH. Thermogenesis and thyroid function. Annu Rev Nutr 1995; 15:263–91. [DOI] [PubMed] [Google Scholar]
  • 60.Bianco AC, Salvatore D, Gereben B, Berry MJ, Larsen PR. Biochemistry, cellular and molecular biology, and physiological roles of the iodothyronine selenodeiodinases. Endocr Rev 2002; 23:38–89. [DOI] [PubMed] [Google Scholar]
  • 61.Bianco AC, Larsen PR. Cellular and structural biology of the deiodinases. Thyroid 2005; 15:777–86. [DOI] [PubMed] [Google Scholar]
  • 62.Bartha T, Kim SW, Salvatore D, Gereben B, Tu HM, Harney JW, et al. . Characterization of the 5’-flanking and 5’-untranslated regions of the cyclic adenosine 3’,5’-monophosphate-responsive human type 2 iodothyronine deiodinase gene. Endocrinology 2000; 141:229–37. [DOI] [PubMed] [Google Scholar]
  • 63.Celi FS, Canettieri G, Yarnall DP, Burns DK, Andreoli M, Shuldiner AR, et al. . Genomic characterization of the coding region of the human type II 5’-deiodinase gene. Mol Cell Endocrinol 1998; 141:49–52. [DOI] [PubMed] [Google Scholar]
  • 64.Guo TW, Zhang FC, Yang MS, Gao XC, Bian L, Duan SW, et al. . Positive association of the DIO2 (deiodinase type 2) gene with mental retardation in the iodine-deficient areas of China. J Med Genet 2004; 41:585–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Meulenbelt I, Min JL, Bos S, Riyazi N, Houwing-Duistermaat JJ, van der Wijk HJ, et al. . Identification of DIO2 as a new susceptibility locus for symptomatic osteoarthritis. Hum Mol Genet 2008; 17:1867–75. [DOI] [PubMed] [Google Scholar]
  • 66.Nair S, Muller YL, Ortega E, Kobes S, Bogardus C, Baier LJ. Association analyses of variants in the DIO2 gene with early-onset type 2 diabetes mellitus in Pima Indians. Thyroid; 22:80–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Huang JS, Ramamurthy SK, Lin X, Le Breton GC. Cell signalling through thromboxane A2 receptors. Cell Signal 2004; 16:521–33. [DOI] [PubMed] [Google Scholar]
  • 68.Miggin SM, Kinsella BT. Expression and tissue distribution of the mRNAs encoding the human thromboxane A2 receptor (TP) alpha and beta isoforms. Biochim Biophys Acta 1998; 1425:543–59. [DOI] [PubMed] [Google Scholar]
  • 69.Rolin S, Masereel B, Dogne JM. Prostanoids as pharmacological targets in COPD and asthma. European journal of pharmacology 2006; 533:89–100. [DOI] [PubMed] [Google Scholar]
  • 70.Takeuchi K, Mashimo Y, Shimojo N, Arima T, Inoue Y, Morita Y, et al. . Functional variants in the thromboxane A2 receptor gene are associated with lung function in childhood-onset asthma. Clin Exp Allergy; 43:413–24. [DOI] [PubMed] [Google Scholar]
  • 71.Unoki M, Furuta S, Onouchi Y, Watanabe O, Doi S, Fujiwara H, et al. . Association studies of 33 single nucleotide polymorphisms (SNPs) in 29 candidate genes for bronchial asthma: positive association a T924C polymorphism in the thromboxane A2 receptor gene. Hum Genet 2000; 106:440–6. [DOI] [PubMed] [Google Scholar]
  • 72.Leung TF, Tang NL, Lam CW, Li AM, Chan IH, Ha G. Thromboxane A2 receptor gene polymorphism is associated with the serum concentration of cat-specific immunoglobulin E as well as the development and severity of asthma in Chinese children. Pediatr Allergy Immunol 2002; 13:10–7. [DOI] [PubMed] [Google Scholar]
  • 73.Shin HD, Park BL, Jung JH, Wang HJ, Park HS, Choi BW, et al. . Association of thromboxane A2 receptor (TBXA2R) with atopy and asthma. J Allergy Clin Immunol 2003; 112:454–7. [DOI] [PubMed] [Google Scholar]
  • 74.Nowak DG, Woolard J, Amin EM, Konopatskaya O, Saleem MA, Churchill AJ, et al. . Expression of pro- and anti-angiogenic isoforms of VEGF is differentially regulated by splicing and growth factors. J Cell Sci 2008; 121:3487–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Woolard J, Bevan HS, Harper SJ, Bates DO. Molecular diversity of VEGF-A as a regulator of its biological activity. Microcirculation 2009; 16:572–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Maloney J, Wang D, Duncan T, Voelkel N, Ruoss S. Plasma vascular endothelial growth factor in acute mountain sickness. Chest 2000; 118:47–52. [DOI] [PubMed] [Google Scholar]
  • 77.Appenzeller O, Minko T, Pozharov V, Bonfichi M, Malcovati L, Gamboa J, et al. . Gene expression in the Andes; relevance to neurology at sea level. Journal of the neurological sciences 2003; 207:37–41. [DOI] [PubMed] [Google Scholar]
  • 78.Appenzeller O, Minko T, Qualls C, Pozharov V, Gamboa J, Gamboa A, et al. . Gene expression, autonomic function and chronic hypoxia:lessons from the Andes. Clinical autonomic research : official journal of the Clinical Autonomic Research Society 2006; 16:217–22. [DOI] [PubMed] [Google Scholar]
  • 79.Patitucci M, Lugrin D, Pages G. Angiogenic/lymphangiogenic factors and adaptation to extreme altitudes during an expedition to Mount Everest. Acta physiologica 2009; 196:259–65. [DOI] [PubMed] [Google Scholar]
  • 80.Walter R, Maggiorini M, Scherrer U, Contesse J, Reinhart WH. Effects of high-altitude exposure on vascular endothelial growth factor levels in man. Eur J Appl Physiol 2001; 85:113–7. [DOI] [PubMed] [Google Scholar]
  • 81.Gao W, Gao Y, Zhang G, Song L, Sun B, Shi J. Hypoxia-induced expression of HIF-1alpha and its target genes in umbilical venous endothelial cells of Tibetans and immigrant Han. Comparative biochemistry and physiology Toxicology & pharmacology : CBP 2005; 141:93–100. [DOI] [PubMed] [Google Scholar]
  • 82.Dorward DA, Thompson AA, Baillie JK, MacDougall M, Hirani N. Change in plasma vascular endothelial growth factor during onset and recovery from acute mountain sickness. Respiratory medicine 2007; 101:587–94. [DOI] [PubMed] [Google Scholar]
  • 83.Dai J, Rabie AB. VEGF: an essential mediator of both angiogenesis and endochondral ossification. Journal of dental research 2007; 86:937–50. [DOI] [PubMed] [Google Scholar]
  • 84.Harper SJ, Bates DO. VEGF-A splicing: the key to anti-angiogenic therapeutics? Nat Rev Cancer 2008; 8:880–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Xu J, Dou T, Liu C, Fu M, Huang Y, Gu S, et al. . The evolution of alternative splicing exons in vascular endothelial growth factor A. Gene 2011; 487:143–50. [DOI] [PubMed] [Google Scholar]
  • 86.Holmes DI, Zachary I. The vascular endothelial growth factor (VEGF) family: angiogenic factors in health and disease. Genome Biol 2005; 6:209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Buroker NE, Ning XH, Zhou ZN, Li K, Cen WJ, Wu XF, et al. . AKT3, ANGPTL4, eNOS3, and VEGFA associations with high altitude sickness in Han and Tibetan Chinese at the Qinghai-Tibetan Plateau. International journal of hematology 2012; 96:200–13. [DOI] [PubMed] [Google Scholar]

Articles from Transcription are provided here courtesy of Taylor & Francis

RESOURCES