首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Considerable evidence supports the idea that odorant recognition depends on specific sequence variations in olfactory receptor (OR) proteins. Much of this emerges from in vitro screens in heterogenous expression systems. However, the ultimate proof should arise from measurements of odorant thresholds in human individuals harboring different OR genetic variants, a research vein that has so far been only scantly explored. The study of McRae et al., published in this issue of Chemical Senses, shows how the recognition of a grassy odorant depends on specific OR interindividual sequence changes. It provides a clear relevant example for the impact of genetics on olfaction and is an excellent portrayal of the power of human genomics to decipher olfactory perception.  相似文献   

2.
Connexons and cell adhesion: a romantic phase   总被引:3,自引:1,他引:2  
Recent evidence indicates, that gap junction forming proteins do not only contribute to intercellular communication (Kanno and Saffitz in Cardiovasc Pathol 10:169-177, 2001; Saez et al. in Physiol Rev 83:1359-1400, 2003), ion homeostasis and volume control (Goldberg et al. in J Biol Chem 277:36725-36730, 2002; Saez et al. in Physiol Rev 83:1359-1400, 2003). They also serve biological functions in a mechanical sense, supporting adherent connections between neighbouring cells of epithelial and non-epithelial tissues (Clair et al. in Exp Cell Res 314:1250-1265, 2008; Shaw et al. in Cell 128:547-560, 2007), where they stabilize migratory pathways in the developing central nervous system (Elias et al. in Nature 448:901-907, 2007; Malatesta et al. in Development 127:5253-5263, 2000; Noctor et al. in Nature 409:714-720, 2001; Rakic in Brain Res 33:471-476, 1971; J Comp Neurol 145:61-83 1972; Science 241:170-176, 1988), or mediate polarized movements and directionality of neural crest cells during organogenesis (Kirby and Waldo in Circ Res 77:211-215, 1995; Xu et al. in Development 133:3629-3639, 2006). Since, most data describing adhesive properties of gap junctions delt with connexin 43 (Cx43) (Beardslee et al. in Circ Res 83:629-635, 1998), we will focus our brief review on this isoform.  相似文献   

3.
Iron regulatory proteins (IRPs), the cytosolic proteins involved in the maintenance of cellular iron homeostasis, bind to stem loop structures found in the mRNA of key proteins involved iron uptake, storage, and metabolism and regulate the expression of these proteins in response to changes in cellular iron needs. We have shown previously that HFE-expressing fWTHFE/tTA HeLa cells have slightly increased transferrin receptor levels and dramatically reduced ferritin levels when compared to the same clonal cell line without HFE (Gross et al., 1998, J Biol Chem 273:22068-22074). While HFE does not alter transferrin receptor trafficking or non-transferrin mediated iron uptake, it does specifically reduce (55)Fe uptake from transferrin (Roy et al., 1999, J Biol Chem 274:9022-9028). In this report, we show that IRP RNA binding activity is increased by up to 5-fold in HFE-expressing cells through the activation of both IRP isoforms. Calcein measurements show a 45% decrease in the intracellular labile iron pool in HFE-expressing cells, which is in keeping with the IRP activation. These results all point to the direct effect of the interaction of HFE with transferrin receptor in lowering the intracellular labile iron pool and establishing a new set point for iron regulation within the cell.  相似文献   

4.
R A Broglia  G Tiana 《Proteins》2001,45(4):421-427
While all the information required for the folding of a protein is contained in its amino acid sequence, one has not yet learned how to extract this information to predict the detailed, biological active, three-dimensional structure of a protein whose sequence is known. Using insight obtained from lattice model simulations of the folding of small proteins (fewer than 100 residues), in particular of the fact that this phenomenon is essentially controlled by conserved contacts (Mirny et al., Proc Natl Acad Sci USA 1995;92:1282) among (few) strongly interacting ("hot") amino acids (Tiana et al., J Chem Phys 1998;108:757-761), which also stabilize local elementary structures formed early in the folding process and leading to the (postcritical) folding core when they assemble together (Broglia et al., Proc Natl Acad Sci USA 1998;95:12930, Broglia & Tiana, J Chem Phys 2001;114:7267), we have worked out a successful strategy for reading the three-dimensional structure of lattice model-designed proteins from the knowledge of only their amino acid sequence and of the contact energies among the amino acids.  相似文献   

5.
A quantitative structure-activity relationship (QSAR) model relating electrotopological state (E-state) indices and mutagenic potency was previously described by Cash [Mutat. Res. 491 (2001) 31-37] using a data set of 95 aromatic amines published by Debnath et al. [Environ. Mol. Mutagen. 19 (1992) 37-52]. Mutagenic potency was expressed as the number of Salmonella typhimurium TA98 revertants per nmol (LogR). Earlier work on the development of QSARs for the prediction of genotoxicity indicated that numerous methods could be effectively employed to model the same aromatic amines data set, namely, Debnath et al.; Maran et al. [Quant. Struct.-Act. Relat. 18 (1999) 3-10]; Basak et al. [J. Chem. Inf. Comput. Sci. 41 (2001) 671-678]; Gramatica et al. [SAR QSAR Environ. Res. 14 (2003) 237-250]. However, results obtained from external validations of those models revealed that the effective predictivity of the QSARs was well below the potential indicated by internal validation statistics (Debnath et al., Gramatica et al.). The purpose of the current research is to externally validate the model published by Cash using a data set of 29 aromatic amines reported by Glende et al. [Mutat. Res. 498 (2001) 19-37; Mutat. Res. 515 (2002) 15-38] and to further explore the potential utility of using E-state sums for the prediction of mutagenic potency of aromatic amines.  相似文献   

6.
Volatile odorous chemicals are detected by around a thousand different G protein-coupled odorant receptors in the mouse. We demonstrated that exposure of the behaving mouse to odorant for a few minutes led to induction of the immediate early gene c-fos for several hours in a fraction of the olfactory sensory neurones in the nasal cavity. Associated with this odorant-specific induction event was activation of extracellular-regulated kinase (ERK)1/2 that preceded increased c-fos expression. The distribution of odorant-activated neurones mimicked the scattered and spatially limited distribution of neurones expressing a single odorant receptor gene. A small change in odorant chemical structure caused a zonal shift in the spatial distribution of activated neurones, suggesting that the gene expression change resulted from specific receptor interaction. Repeated exposure to odorant or use of different concentrations did not change the pattern of c-fos induction. These results indicate that odorant-induced c-fos expression can be used to visualize odorant representations in the olfactory epithelium that reflect late cellular events regulated by adequate odorant receptor stimulation.  相似文献   

7.
基于线粒体控制区序列的猕猴属系统发育研究   总被引:6,自引:1,他引:6  
通过线粒体部分控制区DNA序列数据探讨7种猕猴属物种的分子系统发育关系。结果表明熊猴的核苷酸多样度最高,而藏酋猴核苷酸多样度较低。基于控制区序列数据所构建的最大似然树,不考虑食蟹猴的位置,7种猕猴物种可粗略地分为3个种组,即狮尾猴组(包括北平顶猴)、头巾猴组(包括红面猴、熊猴和藏酋猴)和食蟹猴组(包括恒河猴和台湾猴)。与前人(Fooden&Lanyon,1989;Tosi et al,2003a;Deinard&Smith,2001;Evans et al,1999;Hayasaka et al,1996;Morales&Melnick,1998)的结果不同,我们的结果支持食蟹猴比北平顶猴分化早的假设;东部恒河猴(相对于台湾猴)和东部熊猴(相对于藏酋猴)出现并系。与Y染色体、等位酶、核基因以及部分形态学数据推测的结果(Delson,1980;Fooden&Lanyon。1989;Fooden,1990;Tosi et al,2000,2003a,b;Deinard&Smith,2001)一致,红面猴应归于头巾猴组,但此结论与前人(Hayasaka et al,1996;Morales&Melnick,1998;Tosi et al,2003a)依据线粒体得到的结果有较大分歧。  相似文献   

8.
Template-directed replication is known to obey a parabolic growth law due to product inhibition (Sievers & Von Kiedrowski 1994 Nature 369, 221; Lee et al. 1996 Nature 382, 525; Varga & Szathmáry 1997 Bull. Math. Biol. 59, 1145). We investigate a template-directed replication with a coupled template catalysed lipid aggregate production as a model of a minimal protocell and show analytically that the autocatalytic template-container feedback ensures balanced exponential replication kinetics; both the genes and the container grow exponentially with the same exponent. The parabolic gene replication does not limit the protocellular growth, and a detailed stoichiometric control of the individual protocell components is not necessary to ensure a balanced gene-container growth as conjectured by various authors (Gánti 2004 Chemoton theory). Our analysis also suggests that the exponential growth of most modern biological systems emerges from the inherent spatial quality of the container replication process as we show analytically how the internal gene and metabolic kinetics determine the cell population's generation time and not the growth law (Burdett & Kirkwood 1983 J. Theor. Biol. 103, 11-20; Novak et al. 1998 Biophys. Chem. 72, 185-200; Tyson et al. 2003 Curr. Opin. Cell Biol. 15, 221-231). Previous extensive replication reaction kinetic studies have mainly focused on template replication and have not included a coupling to metabolic container dynamics (Stadler et al. 2000 Bull. Math. Biol. 62, 1061-1086; Stadler & Stadler 2003 Adv. Comp. Syst. 6, 47). The reported results extend these investigations. Finally, the coordinated exponential gene-container growth law stemming from catalysis is an encouraging circumstance for the many experimental groups currently engaged in assembling self-replicating minimal artificial cells (Szostak 2001 et al. Nature 409, 387-390; Pohorille & Deamer 2002 Trends Biotech. 20 123-128; Rasmussen et al. 2004 Science 303, 963-965; Szathma ry 2005 Nature 433, 469-470; Luisi et al. 2006 Naturwissenschaften 93, 1-13).  相似文献   

9.
Recently a putative mammalian neutral-sphingomyelinase was cloned [Tomiuk et al. (1998) Proc. Natl. Acad. Sci. USA 95, 3638-3643; GenBank accession number AJ222801]. We have overexpressed this enzyme in cultured cells and demonstrate, using four different tagged constructs, that it is localized at the endoplasmic reticulum and not at the plasma membrane. This localization precludes a role for enzyme AJ222801 in the sphingomyelin cycle. Furthermore, a recent publication demonstrated that this enzyme has lyso-platelet activating factor (PAF) phospholipase C activity [Sawai et al. (1999) J. Biol. Chem. 274, 38131-38139]. Together, these data suggest a role for enzyme AJ222801 in the regulation of PAF metabolism.  相似文献   

10.
Human sniffing behavior usually involves bouts of short, high flow rate inhalation (>300 ml/s through each nostril) with mostly turbulent airflow. This has often been characterized as a factor enabling higher amounts of odorant to deposit onto olfactory mucosa than for laminar airflow and thereby aid in olfactory detection. Using computational fluid dynamics human nasal cavity models, however, we found essentially no difference in predicted olfactory odorant flux (g/cm2 s) for turbulent versus laminar flow for total nasal flow rates between 300 and 1000 ml/s and for odorants of quite different mucosal solubility. This lack of difference was shown to be due to the much higher resistance to lateral odorant mass transport in the mucosal nasal airway wall than in the air phase. The simulation also revealed that the increase in airflow rate during sniffing can increase odorant uptake flux to the nasal/olfactory mucosa but lower the cumulative total uptake in the olfactory region when the inspired air/odorant volume was held fixed, which is consistent with the observation that sniff duration may be more important than sniff strength for optimizing olfactory detection. In contrast, in rats, sniffing involves high-frequency bouts of both inhalation and exhalation with laminar airflow. In rat nose odorant uptake simulations, it was observed that odorant deposition was highly dependent on solubility and correlated with the locations of different types of receptors.  相似文献   

11.
Molecular properties of odorant compounds essential for activation of the human olfactory receptor hOR17-40 were investigated using a collection of 23 variants of its cognate ligand helional. Coupling receptor activation to an optically detectable intracellular Ca(2+) ion flux allowed dose-dependent screening of different odorant molecules in human embryonic kidney (HEK)293 cells. We found an extended collection of activating ligands and provide first evidence for hOR17-40-specific antagonists. The C-terminal fusion of enhanced green fluorescent protein to the hOR17-40 retained full receptor function and permitted the selection of cells with defined receptor expression levels, which was an essential step for optimizing our screening protocol. Interestingly, cells with a low EGFP fluorescence intensity exhibited efficient hOR17-40 cell surface targeting and odorant-evoked signal transduction; in contrast, highly fluorescent cells displayed mainly incorrectly targeted, intracellular receptors. Fluorescence-activated cell sorting was used to separate hOR17-40-expressing cells on the basis of their endogenous EGFP fluorescence intensity, thereby increasing the fraction of odorant-responsive cells to up to 80% of the total cell number.  相似文献   

12.
Fenton chemistry [Fenton (1894) J. Chem. Soc. 65, 899-910] techniques were employed to identify the residues involved in metal binding located at the active sites of restriction endonucleases. This process uses transition metals to catalytically oxidize the peptide linkage that is in close proximity to the amino acid residues involved in metal ligation. Fe2+ was used as the redox-active transition metal. It was expected that Fe2+ would bind to the endonucleases at the Mg2+-binding site [Liaw et al. (1993) Biochemistry 32, 7999-4003; Ermácora et al. (1992) Proc. Natl. Acad. Sci. U.S.A. 89, 6383-6387; Soundar and Colman (1993) J. Biol. Chem. 268, 5264-5271; Wei et al. (1994) Biochemistry 33, 7931-7936; Ettner et al. (1995) Biochemistry 34, 22-31; Hlavaty and Nowak (1997) Biochemistry 36, 15515-15525). Fe2+-mediated oxidation was successfully performed on TaqI endonulease, suggesting that this approach could be applied to a wide array of endonucleases [Cao and Barany (1998) J. Biol. Chem. 273, 33002-33010]. The restriction endonucleases BamHI, FokI, BglI, BglII, PvuII, SfiI, BssSI, BsoBI, EcoRI, EcoRV, MspI, and HinP1I were subjected to oxidizing conditions in the presence of Fe2+ and ascorbate. All proteins were inactivated upon treatment with Fe2+ and ascorbate. BamHI, FokI, BglI, BglII, PvuII, SfiI, BssSI, and BsoBI were specifically cleaved upon treatment with Fe2+/ascorbate. The site of Fe2+/ascorbate-induced protein cleavage for each enzyme was determined. The Fe2+-mediated oxidative cleavage of BamHI occurs between residues Glu77 and Lys78. Glu77 has been shown by structural and mutational studies to be involved in both metal ligation and catalysis [Newman et al. (1995) Science 269, 656-663; Viadiu and Aggarwal (1998) Nat. Struct. Biol. 5, 910-916; Xu and Schildkraut (1991) J. Biol. Chem. 266, 4425-4429]. The sites of Fe2+/ascorbate-induced cleavage for PvuII, FokI, BglI, and BsoBI agree with the metal-binding sites identified in their corresponding three-dimensional structures or from mutational studies [Cheng et al. (1994) EMBO J. 13, 3297-3935; Wah et al. (1997) Nature 388, 97-100; Newman et al. (1998) EMBO J. 17, 5466-5476; Ruan et al. (1997) Gene 188, 35-39]. The metal-binding residues of BglII, SfiI, and BssSI are proposed based on amino acid sequencing of their Fe2+/ascorbate-generated cleavage fragments. These results suggest that Fenton chemistry may be a useful methodology in identifying amino acids involved in metal binding in endonucleases.  相似文献   

13.
The conformation of orexin-A, an orphan G-protein coupled receptor agonist has been determined when bound to sodium dodecylsulphate-d(25) (SDS) micelles by (1)H and (13)C NMR and molecular modeling. Orexin-A has been implicated in sleep-wakefulness and feeding regulation. The conformational preference of orexin-A consists of a short helical section, involving Asp(5) to Gln(9) that makes up helix I, followed by a bend from Lys(10) to Ser(13). Residues Leu(16) to Gly(22) make up helix II. The conformation of orexin-A can now be used to explain the results of earlier Ala substitution mutagenesis experiments (J. G. Darker et al., Bioorg. Med. Chem. Lett. 11, 737-740 (2001); S. Ammoun, et al., J. Pharmacol. Expt. Ther. 305, 507-514 (2003)). Darker et al., working with orexin-A (15-33) amide, observed a significant drop in functional potency at the OX(1)R receptor when Leu(16), Leu(19), Leu(20), His(26), Gly(29), Ile(30), Leu(31), Thr(32), and Leu(33) were replaced by Ala. Ammoun et al. identified three areas of interest, which were the same for OX(1)R and OX(2)R receptors, as amino acids 15-17, 20 and 25-26 with the most marked reduction in activity being produced by the replacement of Leu(20) by Ala. We suggest that Leu(16), Leu(19), and Leu(20), which are in helix II, are likely responsible for binding orexin-A to the surface of the micelle.  相似文献   

14.
Odorant deposition in the nasal and olfactory mucosas is dependent on a number of factors including local air/odorant flow distribution patterns, odorant mucosal solubility and odorant diffusive transport in the mucosa. Although many of these factors are difficult to measure, mucosal solubility in the bullfrog mucus has been experimentally determined for a few odorants. In the present study an experimental procedure was combined with computational fluid dynamic (CFD) techniques to further describe some of the factors that govern odorant mucosal deposition. The fraction of odorant absorbed by the nasal mucosa (eta) was experimentally determined for a number of odorants by measuring the concentration drop between odorant 'blown' into one nostril and that exiting the contralateral nostril while the subject performed a velopharyngeal closure. Odorant concentrations were measured with a photoionization detector. Odorants were delivered to the nostrils at flow rates of 3.33 and 10 l/min. The velopharyngeal closure nasal air/odorant flows were then simulated using CFD techniques in a 3-D anatomically accurate human nose modeland the mucosal odorant uptake was numerically calculated. The comparison between the numerical simulations and the experimental results lead to an estimation of the human mucosal odorant solubility and the mucosal effective diffusive transport resistance. The results of the study suggest that the increase in diffusive resistance of the mucosal layer over that of a thin layer of water seemed to be general and non-odorant-specific; however, the mucosa solubility was odorant specific and usually followed the trend that odorants with lower water solubility were more soluble in the mucosa than would be predicted from water solubility alone. The ability of this approach to model odorant movement in the nasal cavity was evaluated by comparison of the model output with known values of odorant mucosa solubility.  相似文献   

15.
In past reports we illustrated the importance of Y131, Y322, and T137 within the intracellular (IC) face of the rat bradykinin B2 receptor (rBKB2R) for signal transduction and receptor maintenance (Prado et al. [1997] J. Biol. Chem. 272:14638-14642; Prado et al. [1998] J. Biol. Chem. 273:33548-33555). In this report, we mutate the remaining hydroxyl possessing residues located within the rBKB2R IC region. Exchange of S139A (IC2) or T239V (IC3) did not affect BK activated phosphatidylinositol (PI) turnover or receptor internalization. Chimeric exchange of the last 34 amino acids of BKB2R C-terminus with the corresponding 34 amino acids of the rat angiotensin II AT1a receptor (rAT1aR), both containing an S/T cluster, resulted in a mutant with normal endocytosis and BK activated PI turnover. A more selective chimera of these S/T clusters, with an exchange of BKB2R (333-351) with a rAT1aR fragment (326-342), resulted in a receptor with a retarded internalization but a normal BK activated PI turnover. Subsequent mutation of rBKB2R T344V showed little change in receptor uptake but a pronounced loss of BK activated PI turnover. The mutation of S335A, S341A, S348A, and S350A resulted in very poor receptor internalization and loss of activated PI turnover. Closer examination of this serine cluster illustrated that the replacement of S348A led to poor internalization; whereas the retention of S348 and mutation of S341A resulted in a receptor with a much greater internalization than WT. These and other results suggest that the presence of S348 promotes internalization while the presence of S341 dampens it. Conversely, S341 and S350 proved important for receptor signaling. In sum, our results illustrate that the distal C-terminus including its S/T cluster is important for both rBKB2R internalization and signal transduction. Individual S/T residues within this cluster appear involved in either signal transmission or receptor uptake capacity. However, replacement of the entire distal tail region with the corresponding rAT1aR sequence, also containing an S/T cluster, enables the BKB2R/AT1aR chimera to act in a very similar manner to wild type rBKB2R.  相似文献   

16.
Odorant-binding proteins (OBPs) are small abundant extracellular proteins belonging to the lipocalin superfamily. They are thought to participate in perireceptor events of odor detection by carrying, deactivating, and/or selecting odorant molecules. Putative human OBP genes (hOBP) have recently been described [Lacazette et al. (2000) Hum. Mol. Genet. 9, 289-301], but the presence of the corresponding proteins remained to be established in the human olfactory mucus. This paper reports the first evidence of such expression in the mucus covering the olfactory cleft, where the sensory olfactory epithelium is located. On the contrary, hOBPs were not observed in the nasal mucus covering the septum and the lower turbinate. To demonstrate the odorant binding activity of these proteins, a corresponding recombinant protein variant, hOBP(IIa)(alpha), was secreted by the yeast Pichia pastoris and thoroughly characterized. It appears as a monomer with one disulfide bond located between C59 and C151, a conservative feature of all other vertebrate OBPs. By measuring the displacement of several fluorescent probes, we show that hOBP(IIa)(alpha) is able to bind numerous odorants of diverse chemical structures, with a higher affinity for aldehydes and large fatty acids. A computed 3D model of hOBP(IIa)(alpha) is proposed and reveals that two lysyl residues of the binding pocket may account for the increased affinity for aldehydes. The relatively limited specificity of hOBP(IIa)(alpha) suggests that other human OBPs are expected to take into account the large diversity of odorant molecules.  相似文献   

17.
In this paper we present results on the vacuolar uptake mechanism for two flavone glucuronides present in rye mesophyll vacuoles. In contrast to barley flavone glucosides (Klein et al. (1996) J. Biol. Chem. 271, 29666-29671), the flavones luteolin 7-O-diglucuronyl-4'-O-glucuronide (R1) and luteolin 7-O-diglucuronide (R2) were taken up into vacuoles isolated from rye via a directly energized mechanism. Kinetic studies suggested that the vacuolar glucuronide transport system is constitutively expressed throughout rye primary leaf development. Competition experiments argued for the existence of a plant MRP-like transporter for plant-specific and non-plant glucuronides such as beta-estradiol 17-(beta-D-glucuronide) (E217G). The interaction of ATP-dependent vacuolar glucuronide uptake with glutathione and its conjugates turned out to be complex: R1 transport was stimulated by dinitrobenzene-GS and reduced glutathione but was inhibited by oxidized glutathione in a concentration-dependent manner. In contrast, R2 uptake was not increased in the presence of reduced glutathione. Thus, the transport system for plant-derived glucuronides differed from the characteristic stimulation of vacuolar E217G uptake by glutathione conjugates but not by reduced glutathione (Klein et al. (1998) J. Biol. Chem. 273, 262-270). Using tonoplast vesicles isolated with an artificial K+ gradient, we demonstrate for the first time for plant MRPs that the ATP-dependent uptake of R1 is membrane-potential dependent. We discuss the kinetic capacity of the ABC-type glucuronide transporter to explain net vacuolar flavone glucuronide accumulation in planta during rye primary leaf development and the possibility of an interaction of potential substrates at both the substrate binding and allosteric sites of the MRP transporter regulating the activity towards a certain substrate.  相似文献   

18.
Four days after a single systemic injection of 50 mg/kg of dichlobenil(2,6-dichlorobenzonitrile), the olfactory neuroepithelium ofthe frog is extensively damaged. At the same time, electrophysiologicalresponses to odorant stimulations (2-heptanone, D-limonene,amyl acetate, camphor) are largely reduced. Pretreatment ofthe animals with metyrapone, an inhibitor of cytochrome P-450enzymatic systems, inhibits the histological and physiologicaltoxic effects of the dichlobenil injection. The olfactory tissuerecovered 3months after the dichlobenil injections and responsesto odorant stimulations returned. The same dichlobenil injectionsdid not induce lesions in the vomeronasal neuroepithelium. Chem.Senses 20: 433–440, 1995.  相似文献   

19.
昆虫对外界气味的感受作用是一个庞大而复杂的体系,多种蛋白参与了这一过程。其中包括气味结合蛋白,气味结合蛋白受体,气味降解酶等多种蛋白。昆虫不仅可以通过外界气味分子携带的信息来识别配偶,天敌,还可以通过对外界环境特征的识别来寻找食物来源,产卵等。明确昆虫的化学感受机制不仅可以帮助我们理解昆虫的行为,还有助于深入了解动物的行为机制。文章综述昆虫对气味分子的识别、气味分子在昆虫体内的运输以及电化学信号传导机制等方面的进展。  相似文献   

20.
Linkage of chromosome 11q13 to type 1 diabetes (T1D) was first reported from genome scans (Davies et al. 1994; Hashimoto et al. 1994) resulting in P <2.2 x 10(-5) (Luo et al. 1996) and designated IDDM4 ( insulin dependent diabetes mellitus 4). Association mapping under the linkage peak using 12 polymorphic microsatellite markers suggested some evidence of association with a two-marker haplotype, D11S1917*03-H0570POLYA*02, which was under-transmitted to affected siblings and over-transmitted to unaffected siblings ( P=1.5 x 10(-6)) (Nakagawa et al. 1998). Others have reported evidence for T1D association of the microsatellite marker D11S987, which is approximately 100 kb proximal to D11S1917 (Eckenrode et al. 2000). We have sequenced a 400-kb interval surrounding these loci and identified four genes, including the low-density lipoprotein receptor related protein (LRP5) gene, which has been considered as a functional candidate gene for T1D (Hey et al. 1998; Twells et al. 2001). Consequently, we have developed a comprehensive SNP map of the LRP5 gene region, and identified 95 SNPs encompassing 269 kb of genomic DNA, characterised the LD in the region and haplotypes (Twells et al. 2003). Here, we present our refined linkage curve of the IDDM4 region, comprising 32 microsatellite markers and 12 SNPs, providing a peak MLS=2.58, P=5 x 10(-4), at LRP5 g.17646G>T. The disease association data, largely focused in the LRP5 region with 1,106 T1D families, provided no further evidence for disease association at LRP5 or at D11S987. A second dataset, comprising 1,569 families from Finland, failed to replicate our previous findings at LRP5. The continued search for the variants of the putative IDDM4 locus will greatly benefit from the future development of a haplotype map of the genome.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号